Biomass Gasification and Solid Oxide Fuel Cell
Biomass Gasification and Solid Oxide Fuel Cell
Biomass Gasification and Solid Oxide Fuel Cell
17856
A dissertation submitted to
ETH ZURICH
presented by
Florian-Patrice Nagel
Dipl.-Ing. Universität Stuttgart
Born 05.12.1977
citizen of
the Federal Republic of Germany
2008
Für meine Eltern
in Liebe und Dankbarkeit
Danksagung
If we knew what it was we were doing, it would not be called research, would it?
Albert Einstein
Prof. Dr. Alexander Wokaun gilt mein besonderer Dank für die Möglichkeit diese Disser-
tation anzufertigen und die stets vorzügliche Betreuung dieser Arbeit. Ihr Vertrauen, Herr
Prof. Wokaun, in meine Lernfähigkeit und Beharrlichkeit war für mich Verpflichtung und
Ansporn zugleich. Die Quartalsbesprechungen haben meine Arbeit wesentlich beein-
flusst und bereichert.
Die herzliche Aufnahme in die „Thermal Process Engineering“ Gruppe, die Formulierung
des Dissertationsthemas und die Übernahme des Koreferats kann ich Herrn Dr. Serge
Biollaz nicht hoch genug anrechnen. Ich danke Dir, Serge, für die vielen anregenden
Diskussionen die aufgrund Deiner Erfahrung fast immer zur Verbesserung des Ver-
suchsaufbaus geführt haben. Vielen herzlichen Dank für die freundschaftliche Zusam-
menarbeit und für die Unterstützung in Zeiten persönlicher Not und Ungewissheit ob ich
diese Arbeit gesundheitsbedingt abschließen kann.
Für die Übernahme des Koreferats und interessante Diskussionen zum Thema der Mo-
dellierung bedanke ich mich herzlich bei Dr. François Maréchal.
Dr. Tilman Schildhauer könnte ich eine eigene Danksagung widmen und hätte immer
noch nicht genug Platz um meinen Dank ausreichend auszudrücken. Die unzähligen Dis-
kussionen, Fragen und Anregungen in allen Bereichen der Dissertation haben mich nicht
nur akademisch sondern auch menschlich bereichert. Danke für die viele Zeit die Du,
Tilman, in mich und diese Dissertation investiert hast, ohne die diese Arbeit nicht in die-
ser Form möglich gewesen wäre.
Dr. Samuel Stucki danke ich herzlich für die freundliche Aufnahme in das Labor für
Energie- und Stoffkreisläufe und die Einführung in die Welt der Elektrochemie.
Dr. Alexander Schuler von der Hexis AG danke ich für die fachliche Unterstützung und
die Bereitstellung mehrerer Brennstoffzellenmodule ohne die der experimentelle Teil die-
ser Arbeit nicht denkbar gewesen wäre.
Für die entspannenden Gespräche in den Kaffeepausen die mir stets neuen Schwung für
meine Arbeit gaben danke ich Gisela Herlein. Ohne Dich, Gisela, wäre meine Zeit am
LEM nur halb so schön gewesen.
Ein besonderer Dank gilt dem Technikteam des LEM. Für die Lösung unlösbarer Steue-
rungsprobleme beim Aufbau des weltweit ersten Vergaser- Brennstoffzellen Systems
danke ich Peter Hottinger. Peter, der erste HoVer-BZM Versuch, bei dem wir uns vorka-
men wie Spaceshuttlepiloten beim Landeanflug auf den alten Hong Kong Airport werde
ich nie vergessen.
Für sein Talent auch die verrücktesten Ideen aufs Papier und zum schlussendlich sogar
zum funktionieren zu bringen danke ich Thomas Marti. Thomas, ich werde immer mit ei-
nem Lächeln im Gesicht an die Zeit mit dem Schraubenschlüssel in der Hand zurück
denken.
Peter Binkert danke ich post mortem für seine Geduld mit meinen Wünschen und die
Beharrlichkeit an der Dreh- und Fräsmaschine. Spaß hat es gemacht, Peter.
Die perfekten Zyklone und viele andere Bauteile im HoVex-Veruschsaufbau verdanke ich
Marcel Hottiger.
Für die unzähligen Stunden mit dem Schweißapparat in der Hand durch die der HoVex-
Aufbau erst Realität wurde danke ich Markus Schriber.
Die Messungen der abstrusesten Gase, die so manche Überraschung an den Tag brach-
ten, hat Herr Jörg Schneebeli durchgeführt. Danke Jörg, für deine stete Bereitschaft mir
in Diagnostikfragen zu helfen.
Weiterer Dank gebührt Dr. Frédéric Vogel, Christian Pitta, Dr. Sudip Ghosh, Manuel
Damsohn, Markus Jenne, Martin Künstle, Stefan Decker, Mohand Nait-Atmane, Thorsten
Schulz, Jan Kopyscinski und Johannes Judex für die Freundschaft sowie allen anderen
nicht namentlich genannten LEM-Mitarbeiter mit denen ich arbeiten durfte und die immer
noch Zeit für das ein oder andere Späßchen gefunden haben.
Last but not least möchte ich für die außerordentlich entspannte und positive Arbeitsat-
mosphäre am PSI und dem AXPO Naturstrom Fond für die finanzielle Unterstützung
meiner Arbeit danken.
Meine Eltern, Dr. Karl-Heinz Nagel und Chantal Nagel, und meine beiden Schwestern,
Nathalie McCaughey und Sandra Nagel Bolliger, haben mich stets unterstützt und mir
durch eine unfallbedingte Zeit voller dunkler Zweifel hinweggeholfen. Ihre Liebe und ihr
Vertrauen in mich und nicht zuletzt die familiäre Geborgenheit haben mir die notwendige
Kraft gegeben meine Reise in die Welt der Wissenschaft fortzusetzen und letztendlich
abzuschließen.
The availability of energy in general and electricity in particular has become a major con-
cern given the growing global population, numerous fast developing nations and deplet-
ing fossil fuel reserves. Further, the excessive use of fossil fuels with their attributed car-
bon dioxide emissions has lead to changes in the global ecosystem with negative conse-
quences for life on earth. Against this background it comes as no surprise that renewable,
carbon dioxide neutral energy sources have gained increased interest recently. Biomass
has a large potential as renewable and carbon dioxide neutral feedstock for electricity
generation but is comparably expensive. An approach to overcome this economical
drawback is the combination of high-temperature fuel cells with biomass gasification
processes, which is commonly referred to as "Biomass Integrated Gasification Fuel Cell"
systems (B-IGFC). Solid oxide fuel cells (SOFC) are considered as promising candidates
for the application in B-IGFC system due to their less stringent requirements to the fuel
gas quality compared to other fuel cell types and their capability of directly oxidizing car-
bonaceous fuel gases with high efficiencies at small scales. However, the composition of
the fuel gas plays an important role for its conversion through SOFCs which leads to
strong system interactions that must be well understood to allow for the exploitation of
the full potential of the B-IGFC approach. This thesis presents the demonstration of the
B-IGFC technology on kW-scale and a thermo-economic system analysis aiming at the
identification of promising B-IGFC systems with power outputs around 1 MWel.
The PSI B-IGFC system concept comprises an updraft biomass gasification reactor, a
hot gas cyclone, a catalytic partial oxidation (CPO) unit and a 1 kW Hexis SOFC system.
The experiments conducted in this thesis initially concentrated on the characterization of
all processes steps in the PSI B-IGFC system. The first milestone was reached with the
stable operation of the lab-scale updraft gasifier over 165 h non-stop. Experiments with
short stacks showed that tars are to some extent a fuel for SOFCs. A commercial CPO
catalyst was investigated, revealing satisfying conversion performance for not only oxy-
genated tars and aromatics but also organic sulfur compounds. The PSI B-IGFC concept
was subsequently operated for 28 h non-stop employing the above mentioned 1 kW
Hexis SOFC system. Compared to operation with partially oxidized methane, the SOFC
delivered 40 % less current when operated with water and nitrogen diluted producer gas.
Overall, the demonstration unit was operated without problems and valuable experience
for future improvements was gathered. The application of effective means for the re-
moval of micro particles has been identified as very important in this respect.
The system analysis is based on a finite volume SOFC model developed in this thesis
that includes all relevant charge, mass and heat transport processes. Cell internal steam
I
reforming is considered through a Langmuir-Hinshelwood type applied kinetic model. For
the system analysis, the model was applied to anode- and electrolyte-supported planar
cell designs with co- and counter-current flow pattern as well as the standard tubular cell
design promoted by the Siemens AG. The model was validated against experimental and
literature data and proved to behave physically correct with smaller mass and energy
balance errors than comparable models. The robustness of the general trends predicted
for various fuel gas compositions was attested by means of a sensitivity analysis.
Based on measured producer gas compositions originating from downdraft and updraft
gasification processes as well as the fluidized bed steam gasification process, seven
B-IGFC systems with different gas processing strategies were defined. The correspond-
ing gas compositions at the fuel inlet of the SOFCs were computed with ASPEN PLUS.
Compared to pre-reformed methane, the power output of SOFCs decreases by 25 to
70 % when operated with producer gases. The magnitude of the power output decrease
mainly depends on the degree of dilution of the producer gas and on the anode catalyst
activity which both determine the importance of activation losses. The tubular cell design
appears to feature a highly active anode catalyst according to the employed model pa-
rameters. Diffusion limitations in contrast do not gain importance for any of the investi-
gated producer gases and cell designs. The required air-to-fuel ratio to maintain a given
mean cell temperature increases with decreasing internal reforming potential of the fuel
gas. This mechanism is considerably less pronounced for counter-current cell designs
compared to their co-current correspondents. The thermal stress resulting from the op-
eration of SOFCs with producer gases is generally lower than for operation with pre-
reformed methane. This may slow down the corresponding cell degradation processes.
The power outputs predicted by the SOFC model were used as input for overall system
simulations performed with ASPEN PLUS aiming at the investigation of the interactions
between the gasification processes, gas processing technologies and SOFC designs.
The heat integration was conducted by means of a generalized heat exchanger network.
The comparably low operational temperature of the zinc oxide trap beds, generally em-
ployed for the removal of hydrogen sulfide, calls for additional humidification of the pro-
ducer gases to prevent thermodynamic carbon deposition except in the system where a
catalytic partial oxidation unit is employed for tar conversion. The net system efficiency is
preeminently determined by the cold gas efficiency of the biomass gasification process
and the auxiliary power requirements, which directly correlate with the air-to-fuel ratios.
The adiabatic methanation and catalytic partial oxidation are effective means to lower the
required air-to-fuel ratios by increasing the internal reforming potential of the correspond-
ing producer gases, while simultaneously converting undesired organic species.
Consequently with 32.1 %, the highest net AC system efficiency in the analysis is
reached by the combination of the updraft gasification with an adiabatic methanation and
II
the tubular cell design. The updraft gasification yields a producer gas with high cold gas
efficiency and considerable internal reforming potential, which is further increased
through the adiabatic methanation. The high internal reforming potential leads to a very
low air-to-fuel ratio and thus low auxiliary power needs. The steam reforming of the pro-
duced methane can be interpreted as the final gasification step using SOFC waste heat.
With the mass and energy flows determined for of the different B-IGFC systems through
the overall system simulations, all major equipment pieces were sized and pricewise
rated using pertinent cost functions. The fuel cell balance of plant equipment and the cor-
responding heat exchangers in particular are the most important cost drivers amounting
for up to 50 %, while the fuel cell itself accounts for up to 25 % of the total system costs.
This emphasizes the importance of the air-to-fuel ratios which together with the opera-
tional cell temperatures determine the size of the heat exchangers. The total plant costs
were related to the corresponding net system power outputs to yield the specific plant
costs. The downdraft gasification based systems have the highest specific plant costs
due to costly wood drying, large heat exchangers and low power outputs resulting from
high activation losses as a consequence of gas dilution. The fluidized bed steam gasifi-
cation based systems yield lower specific plant costs, mainly because of low activation
losses and thus high power outputs. The updraft gasification based systems feature the
lowest specific plant costs as a result of low total plant costs and auxiliary power needs.
The power production costs (PPC) relate the total plant costs to the corresponding sys-
tem efficiencies. The basis for the calculations was a detailed cost analysis for a given
planning horizon including costs not only for operation and maintenance, feedstock and
utilities but also capital costs resulting from plant costs depreciation and interests. The
annuity method was employed to evenly distribute the net value of all costs throughout
the planning horizon. On average, 40 % of the PPC arise from fuel expenses, the re-
mainder being attributed to capital costs. Hence, reduction of the specific plant costs is
considerably more effective than system efficiency increases with respect to PPC reduc-
tion. With 0.1154 €/kWhel, the updraft gasification based system discussed above yields
the lowest PPCs in the analysis. Increasing the operational voltage of the SOFC from
0.6 V to 0.7 V results in a net system efficiency increase from 32.1 % to 37.1 %. With
0.1912 €/kWhel, the resulting reduced power output yields however considerably higher
PPCs. In contrast, reducing the auxiliary power needs by adjusting the air-to-fuel ratio to
the lowest possible value leads to higher net system efficiencies and lower specific plant
costs, thus considerably lowering the PPCs.
Finally, the revenues from heat sales are crucial for the economical viability of the inves-
tigated B-IGFC systems without bottoming cycles. Future systems analysis should focus
on B-IGFC systems with bottoming cycles. Therefore, the developed SOFC model has to
be integrated in Flowsheeting- Software packages such as e.g. ASPEN PLUS.
III
IV
Zusammenfassung
V
Holzgas aufgrund dessen Verdünnung mit Stickstoff und Wasser 40 % weniger Strom
erzeugt. Insgesamt konnte der Versuchsaufbau ohne nennenswerte Probleme betrieben
und wertvolle Erfahrungen für künftige Verbesserungen gesammelt werden. Insbesonde-
re hat sich gezeigt, dass eine effiziente Kleinstpartikelabscheidung für künftige Langzeit-
experimente von hoher Wichtigkeit ist.
Die in der vorliegenden Arbeit beschriebenen Ergebnisse der Systemanalyse beruht auf
einem eigens dafür entwickelten SOFC Modell. Das Modell wurde nach dem Finite-
Volumen-Ansatz aufgebaut und berücksichtigt alle wichtigen Ladungs-, Masse- und
Wärmetransportprozesse die in SOFCs stattfinden. Mögliche zellinterne Dampfreformie-
rungsreaktionen werden über eine Langmuir-Hinshelwood Kinetik beschrieben. Für die
Systemanalyse wurden planare Anoden- und Elektrolyt-gestützte Zelltypen in Gleich-
und Gegenstromausführung, sowie der röhrenförmige Zelltyp der Siemens AG mit dem
Modell abgebildet und gegen Messdaten und Literaturwerte validiert. Es konnte gezeigt
werden, dass das aufgebaute Model physikalisch sinnvolle Ergebnisse liefert. Die er-
reichte Schließung der Massen- und Energiebilanzen ist deutlich besser als die ver-
gleichbarer Modelle in der einschlägigen Literatur. Um sicherzustellen, dass die vorher-
gesagten Trends für verschiedene Brenngaszusammensetzungen nicht die Folge der
festgelegten Modellparameter ist, wurde eine entsprechende Sensitivitätsanalyse durch-
geführt. Die vom Modell vorhergesagten Trends haben sich als sehr robust erwiesen.
Aufbauend auf gemessenen Zusammensetzungen von Produktgasen aus Gleich- und
Gegenstromholzvergasungsprozessen sowie einem Wirbelschichtdampfvergasungspro-
zess wurden sieben verschiedene B-IGFC Systemkonzepte definiert. Die aus den ver-
schiedenen B-IGFC Systemkonzepte resultierenden Gaszusammensetzungen am Eintritt
der eingesetzten SOFCs wurden mittels ASPEN PLUS berechnet.
Im Vergleich zu vorreformiertem Methan ist die Leistung der holzgasbetriebenen SOFCs
um 25 bis 70 % reduziert. Die Höhe der Leistungsminderung hängt maßgeblich vom
"Verdünnungsgrad" des Holzgases und der Aktivität des Anodenmaterials ab, die beide
zusammen die Aktivierungsspannungsverluste bestimmen. Gemäß den verwendeten
Modellparametern scheint im untersuchten röhrenförmigen Zelltyp ein hochaktives Kata-
lysatormaterial zum Einsatz zu kommen. Nennenswerte Diffusionslimitierung wurde für
keines der untersuchten Holzgase und keinen der untersuchten Zelltypen vorhergesagt.
Die zur Einhaltung der vorgegebenen mittleren Zelltemperatur benötigte Kühlungsluftzahl
nimmt mit sinkendem Potential des Brenngases für zellinterne Reformierung zu. Dieser
Effekt ist für Zelltypen in Gegenstromausführung deutlich schwächer ausgeprägt als für
solche in Gleichstromausführung. Die aus dem Betrieb der verschiedenen Zelltypen mit
Holzgas resultierenden thermischen Belastungen sind grundsätzlich geringer als die die
sich mit vorreformiertem Methan als Brenngas einstellen. Dies könnte sich positiv auf die
entsprechenden Zelldegradationsmechanismen auswirken.
VI
Die vom SOFC Modell vorhergesagten Leistungswerte wurden als Eingabewerte für Ge-
samtsystemrechnungen genutzt. Die Rechnungen wurden mit ASPEN PLUS durchge-
führt um die Wechselwirkungen zwischen den verschiedenen Vergasungsprozessen,
Gasreinigungsvarianten und Zelltypen zu untersuchen. Die Wärmeintegration der be-
trachteten Gesamtsysteme wurde über ein verallgemeinertes Wärmetauschernetzwerk
abgebildet.
Die vergleichsweise niedrige Betriebstemperatur des standardmäßig zur Schwefelwas-
serstoffabscheidung eingesetzten Zinkoxids macht eine zusätzliche Befeuchtung der
verschiedenen Holzgase zwingend um eine thermodynamisch begünstigte Kohlenstoff-
bildung zu vermeiden. Einzige Ausnahme ist dabei des B-IGFC Systemkonzept in dem
eine katalytisch partielle Oxidationsstufe zur teilweisen Teerumwandlung zum Einsatz
kommt. Der Netto-Systemwirkungsgrad wird hauptsächlich vom Kaltgaswirkungsgrad
des Vergasungsprozesses und vom Eigenstrombedarf bestimmt, der direkt von der zur
Zellkühlung benötigten Luftzahl abhängt. Die adiabate Methanierung und die katalytisch
partielle Oxidation sind Prozesschritte, die sich gut für die Senkung der benötigten Luft-
zahl eignen indem sie das Potential der entsprechenden Brenngase für zellinterne Re-
formierung steigern und gleichzeitig den Abbau unerwünschter organischer Verbindun-
gen erlauben.
Folglich wurde der höchste Netto-Systemwirkungsgrad von 32.1 % in dieser Systemana-
lyse von der Kombination der Gegenstromholzvergasung mit einer adiabaten Methanie-
rungsstufe und des röhrenförmigen Zelltyps erreicht. Die Gegenstromvergasung liefert
bereits ein Holzgas mit vergleichsweise hohem Potential für zellinterne Reformierung.
Dieses wird in der adiabaten Methanierungsstufe zusätzlich gesteigert, was zu einer ent-
sprechend niedrigen Kühlungsluftzahl und geringem Eigenstrombedarf dieses B-IGFC
Systemkonzepts führt. Die zellinterne Dampfreformierung des hohen Methananteils im
so aufbereiteten Holzgas entspricht im weitesten Sinne dem letzten Schritt der Biomas-
severgasung unter direkter Nutzung der Hochtemperaturabwärme der SOFC.
Die über Gesamtsystemrechnungen für jedes der untersuchten B-IGFC Systemkonzepte
ermittelten Massen- und Energieströme wurden für die Auslegung der wichtigsten Appa-
rate genutzt. Die Gesamtsystemkosten wurden mit Hilfe entsprechende Kostenfunktio-
nen und Wichtungsfaktoren bestimmt. Die Anlagenperipherie der Brennstoffzellen und
insbesondere die benötigten Wärmetauscher machen bis zu 50 % der Gesamtsystem-
kosten aus, während die Brennstoffzellenstapel selbst nur bis zu 25 % der Gesamtsys-
temkosten verursachen. Dies verdeutlicht nochmals den Stellenwert der Kühlungsluftzahl,
aus der, zusammen mit dem Temperaturniveau des entsprechenden Brennstoffzellen-
stapels, die benötigte Wärmetauscherfläche folgt.
Die Gesamtsystemkosten wurden mit den entsprechenden Gesamtsystemleistungen zu
den spezifischen Anlagenkosten verrechnet. Gleichstromholzvergaser basierte Systeme
VII
weisen die höchsten spezifischen Anlagenkosten in dieser Systemanalyse auf. Die
Gründe dafür sind die benötigte kostenintensive Holztrocknung, große Wärmetauscher-
flächen und relative geringe Gesamtsystemleistungen, die auf hohe Aktivierungsverluste
durch die stark verdünnten Holzgase zurückzuführen sind. Im Vergleich dazu sind die
spezifischen Anlagenkosten der B-IGFC Systemkonzepte mit Wirbelschichtdampfholz-
vergasern etwas geringer. Dies ergibt sich trotz der insgesamt höheren Gesamtsystem-
kosten aus den deutlich kleineren Aktivierungsverlusten. Die auf Gegenstromholzverga-
sern beruhenden B-IGFC Systemkonzepte haben die geringsten spezifischen Anlagen-
kosten dieser Systemanalyse aufgrund der insgesamt niedriger Gesamtsystemkosten
und zusätzlich befriedigenden Gesamtsystemleistungen die sich hauptsächlich aus nied-
rigen Eigenstrombedarfswerten ergeben.
Die Stromgestehungskosten bilden die Synthese aus den Gesamtsystemkosten und den
entsprechenden Netto-Systemwirkungsgraden. Grundlage für deren Berechnung war ei-
ne detaillierte Kostenanalyse für einen festgelegten Planungszeitraum. Die Kostenanaly-
se umfasst Betriebs- und Wartungskosten, Brennstoffkosten und Betriebsmittelkosten
sowie alle Kosten die aus der Anlagenabschreibung und Darlehnszinsen entstehen. Zur
gleichmäßigen Verteilung des Nettobarwerts aller anfallenden Kosten über den Pla-
nungszeitraum wurde die Annuitätenmethode angewandt.
Etwa 40 % der Stromgestehungskosten können dem Brennstoffbedarf zugeordnet wer-
den während der Rest im weitesten Sinne von den spezifischen Anlagenkosten abhängt.
Folglich ist die Senkung der spezifischen Anlagenkosten wirksamer um geringere Strom-
gestehungskosten zu erreichen als die Steigerung des Netto-Systemwirkungsgrades.
Das bereits weiter oben beschriebene gegenstromholzvergasungsbasierte System erzielt
mit 0.1154 €/kWhel die geringsten Stromgestehungskosten in dieser Systemanalyse. Die
Erhöhung der Betriebsspannung des eingesetzten Brennstoffzellenstapels von 0.6 V auf
0.7 V ergibt eine Zunahme des Netto-Systemwirkungsgrads von 32.1 auf 37.1 %. Im Ge-
genzug steigen die Stromgestehungskosten auf 0.1912 €/kWhel, was auf die deutlich re-
duzierte Gesamtsystemleistung zurückzuführen ist. Dagegen birgt die Senkung des Ei-
genstrombedarfs durch optimierte Brennstoffzellenstapelbetriebsparameter zur Senkung
der Kühlungsluftzahl erhebliches Potential zur Senkung der Stromgestehungskosten.
Dies folgt aus der gesteigerten Gesamtsystemleistung und den gleichzeitig reduzierten
Kosten für die Anlagenperipherie der Brennstoffzellen.
Schlussendlich sind die Einkünfte aus dem Verkauf von Nutzwärme entscheidend für die
Wirtschaftlichkeit von B-IGFC Systemen, falls die Wärme nicht anlagenintern zur zusätz-
lichen Stromerzeugung verwendet wird. Der Einfluss entsprechender "Bottoming cycles"
auf die Wirtschaftlichkeit von B-IGFC Systemen ist in weitererführenden Arbeiten zu un-
tersuchen. Dazu muss das aufgebaute SOFC Modell in Flowsheeting- Programme wie
z.B. ASPEN PLUS integriert werden.
VIII
IX
X
Table of Contents
1 Introduction................................................................................................................ 1
1.1 Motivation.............................................................................................................. 2
1.2 Objectives ............................................................................................................. 4
1.3 Methodology ......................................................................................................... 4
1.4 Organization of the thesis ..................................................................................... 5
2 Technology overview ................................................................................................ 9
2.1 Biomass .............................................................................................................. 10
2.1.1 Definition .................................................................................................... 10
2.1.2 Classification.............................................................................................. 10
2.1.3 Construction............................................................................................... 10
2.1.4 Properties................................................................................................... 12
2.2 Gasification ......................................................................................................... 16
2.2.1 Biomass conversion overview.................................................................... 16
2.2.2 Fundamentals of thermochemical conversion ........................................... 17
2.2.3 Wood gasification technology overview ..................................................... 21
2.3 Gas processing................................................................................................... 31
2.3.1 Gas cleaning .............................................................................................. 31
2.3.2 Gas conditioning ........................................................................................ 39
2.4 Solid oxide fuel cell ............................................................................................. 43
2.4.1 Fuel cells in general ................................................................................... 43
2.4.2 Technology features .................................................................................. 46
2.4.3 Working principle........................................................................................ 47
2.4.4 Cell designs................................................................................................ 50
2.4.5 Support designs ......................................................................................... 52
2.4.6 Materials..................................................................................................... 54
2.4.7 Cell failure and performance degradation.................................................. 61
3 Experiments ............................................................................................................ 71
3.1 The PSI B-IGFC system ..................................................................................... 71
3.1.1 System description..................................................................................... 71
3.1.2 Discussion of the chosen approach ........................................................... 72
XI
3.2 Gas analysis ....................................................................................................... 76
3.3 Proof-of-feasibility tests ...................................................................................... 79
3.3.1 Lab scale tests under load conditions........................................................ 79
3.3.2 Lab scale tests under open circuit conditions ............................................ 81
3.4 Characterization of important unit operations ..................................................... 84
3.4.1 Updraft gasification reactor ........................................................................ 84
3.4.2 Hot gas particle removal ............................................................................ 88
3.4.3 Catalytic partial oxidation ........................................................................... 92
3.4.4 Sulfur absorption by zinc oxide .................................................................. 96
3.4.5 Solid oxide fuel cell .................................................................................... 97
3.5 Proof-of-concept of the PSI B-IGFC system..................................................... 103
3.5.1 First demonstration unit tests and derived modifications......................... 104
3.5.2 Present demonstration unit configuration ................................................ 106
3.5.3 Long-term test results .............................................................................. 108
4 Modeling work literature review............................................................................. 117
4.1 Modeling approaches for steady-state investigations....................................... 117
4.1.1 Molecular level models ............................................................................ 118
4.1.2 Electrode level models............................................................................. 118
4.1.3 Cell level models...................................................................................... 120
4.1.4 Stack level models ................................................................................... 123
4.1.5 Plant level models.................................................................................... 124
4.2 Recent studies of biomass fed SOFC cycles.................................................... 124
5 Modeling................................................................................................................ 129
5.1 Overall modeling approach............................................................................... 129
5.2 Flowsheeting models ........................................................................................ 131
5.2.1 Biomass gasifier....................................................................................... 131
5.2.2 Gas humidifier.......................................................................................... 132
5.2.3 Heat integration network .......................................................................... 132
5.2.4 Steam reforming and methanation reactors............................................. 133
5.2.5 Catalytic partial oxidation reactors ........................................................... 134
5.2.6 Sulfur adsorbent beds.............................................................................. 135
5.2.7 SOFC ....................................................................................................... 135
XII
5.3 ATHENA SOFC model ..................................................................................... 136
5.3.1 Model description..................................................................................... 136
5.3.2 Planar model verification.......................................................................... 181
5.3.3 Tubular model validation.......................................................................... 186
5.3.4 Triangular model validation...................................................................... 191
5.3.5 Sensitivity analysis................................................................................... 193
5.4 Economic model ............................................................................................... 219
5.4.1 Equipment purchase cost estimate functions .......................................... 220
5.4.2 Equipment sizing and purchase cost estimation...................................... 222
5.4.3 Total investment, direct and indirect plant costs...................................... 227
5.4.4 Net AC power efficiency........................................................................... 228
5.4.5 Power production costs............................................................................ 229
6 System analysis .................................................................................................... 235
6.1 Analysis constraints .......................................................................................... 235
6.1.1 SOFC simulation assumptions................................................................. 235
6.1.2 Overall system simulation assumptions................................................... 239
6.2 System definition............................................................................................... 240
6.2.1 Natural gas reference system .................................................................. 240
6.2.2 B-IGFC systems....................................................................................... 241
6.3 Technical evaluation ......................................................................................... 249
6.3.1 Simulation results of gas processing sections ......................................... 249
6.3.2 Simulation results of SOFC operation with producer gases .................... 253
6.3.3 System simulation results ........................................................................ 270
6.4 Economic evaluation......................................................................................... 279
6.4.1 Direct plant cost estimates....................................................................... 279
6.4.2 Power production costs............................................................................ 283
6.4.3 Sensitivity analysis................................................................................... 286
7 Concluding remarks .............................................................................................. 293
7.1 Conclusions ...................................................................................................... 293
7.1.1 Experiments ............................................................................................. 293
7.1.2 System analysis ....................................................................................... 294
7.2 Recommendations for future work.................................................................... 297
XIII
7.2.1 Experiments ............................................................................................. 297
7.2.2 Modeling and system analysis ................................................................. 299
Notation.......................................................................................................................... 303
References..................................................................................................................... 311
XIV
XV
XVI
1 Introduction
1 Introduction
Figure 1-1:
Development of
global energy
consumption dur-
ing the last 150
years, [1]
Note:
1 Etajoule = 1018 J
A fairer distribution of the fossils fuels and the foreseeable end of re-
serves is however only one part of the problem arising from the pres-
ently excessive use of fossil fuels, which will only continue to increase
due to the rapidly growing global population and numerous fast develop-
ing nations. In the year 2000 only, mankind consumed an amount of fos-
1
1 Introduction
sil fuels that took approx. one million years to be formed. The emissions
of CO2 attributed to the combustion of fossil fuels have thus reached
levels impacting negatively on the ecosystem, endangering not only the
security of food and water supplies but also irreversibly altering the cli-
mate conditions on earth. Thus, a continuity of the developed nations
currently taken as granted and a fair chance for every human being to
relish a comparable quality of life crucially hinges on finding ways to re-
duce the pace at which global energy consumption increases and to
sustainably satisfy the global energy needs. It comes as no surprise
then that renewable energy sources such as solar (direct), wind, tide,
wave, geothermal and biomass have gained increased interest recently
given their time unlimited abundance and global availability. In the ongo-
ing and often controversial debates between citizens, governments, in-
dustry and scientists regarding the energy problem, the electricity gen-
eration has always played a key role and will continue to gain signifi-
cance in the near future.
1.1 Motivation
Among the renewable energy sources biomass has a large potential for
electricity generation. If the harvest is performed in a sustainable way,
wood is a CO2-neutral energy source. Within the next 20 years, the
worldwide installed capacity of biomass power plants is expected to in-
crease by approx. 65 GW, [3]. This forecasted increase however still
leaves most of the potential untapped. A prerequisite to achieve this in-
crease and to exploit its full potential is to overcome the economical
drawbacks related to biomass as feed stock.
Currently most power plants using wood as feedstock are based on
grate firing and steam cycles, where high electrical efficiencies are only
possible on multi-MW scale. Wood is an expensive feedstock due to the
elaborate harvesting processes. This makes the sale of the produced
heat mandatory for the economical feasibility of power generation using
wood. Consequently the scarcity of major heat consumers and connec-
tion points to large-scale district heating grids limit the number of eco-
nomically feasible power plant locations. Small-scale wood power plants
may overcome the heat sales problem due to comparably small
amounts of heat that can be sold to residential heating. However, the
low efficiencies of small-scale steam cycles waste this advantage to a
major extent.
2
1 Introduction
3
1 Introduction
between the gasification process and the SOFC need to be well under-
stood to allow for the exploitation of the full potential of the B-IGFC ap-
proach.
1.2 Objectives
The present thesis aims at demonstrating the B-IGFC technology on
kW-scale and at identifying promising B-IGFC system concepts with
power outputs around 1 MWel. With respect to the latter the most impor-
tant questions are:
How strongly does the producer gas composition resulting from dif-
ferent wood gasification processes influence the temperature man-
agement and the power output of available SOFC designs and what
is the impact on the corresponding cooling requirements?
Are there producer gases that lead to critical operating conditions in
SOFCs such as critical temperature gradients and cell temperatures?
Do different producer gases yield the same effects in SOFCs of dif-
ferent design?
What role does the eventual treatment of tars and removal of sulfur
play for the overall system efficiency and the specific plant costs?
What are the major cost drivers in B-IGFC systems?
What power production costs are feasible and what are the cost dif-
ferences between the different cell designs?
1.3 Methodology
The chosen approach to fulfill the above aims is based on dedicated ex-
periments and numerical systems analysis:
Investigations of different operational modes of an in-house devel-
oped lab-scale updraft gasification reactor are performed aiming at a
stable long-term operation and producer gas composition.
The removal of impurities such as particles and sulfur species as well
as the partial degradation of tars are experimentally characterized to
explore the performance of state-of-the-art and currently developed
gas processing technologies.
An experimental setup covering the complete chain from wood to
electricity is erected and operated to identify the most important tech-
nical problems, which may arise in B-IGFC systems.
4
1 Introduction
6
7
8
2 Technology overview
2 Technology overview
Gasification
Gasification
PG FC
FCSystem
System
Raw suitable
PG for FC
- -Reactor
Reactortype
type - -FC
FCType
Type
- -Gasification
Gasificationagent
agent
Gas
Gasprocessing
processing - -Catalysts
Catalysts
- -Feedstock
Feedstock - -FC
FCdesign
design
- -Operating
Operating - -Operating
Operating
Conditions
Conditions Conditions
Conditions
Heat Electricity
9
2 Technology overview
2.1 Biomass
2.1.1 Definition
Biomass is a diverse term. In biological and ecological disciplines, the
accumulation of all living matter (plants, bacteria, animals and men) or
specific species in a given area or of a biological community is referred
to as biomass. In contrast, the energy and chemical industry defines
biomass as all living or recently deceased biological matter suitable as
raw material for industrial processing. In this work, the term biomass re-
fers to plant tissue or more specifically to wood.
2.1.2 Classification
The variety of plant matter can be classified based on the origin and the
type of growth of the plants, [5]:
Terrestrial plants
Woody plants, i.e. perennial lignocellulosic crops
2.1.3 Construction
Plants combine water and carbon dioxide to sugar building blocks, Eq. 1.
n ⋅ H 2O + n ⋅ CO2 + Light ⎯Chlorophyl
⎯ ⎯⎯l → (CH 2O) n + n ⋅ O2 Eq. 1
10
2 Technology overview
The sugar building blocks are the starting point for the major fractions
found in all terrestrial plants, [8], lignin, hemicellulose and cellulose.
Lignin
Lignin makes up about 25-35 mass-% (dry) of wood and 10-25 mass-
% (dry) of other plants. It is a complex irregular aromatic polymer, of-
ten linked to cellulose and hemicellulose compounds in the plant cell
walls. The high lignin fraction of wood accounts for its rigidity and
slow growth compared to herbaceous plants, Figure 2-2, [6].
Figure 2-2:
Molecular structure
of lignin, [9]
Hemicellulose
In contrast to cellulose, hemicellulose comprises five sugars, namely
xylose, arabinose, galactose, glucose and mannose. Hemicellulose
consists of 50 to 200 sugar units and is amorphous, see Figure 2-3.
Plants consist to approx. 15-30 mass-% (dry) of hemicellulose.
Figure 2-3:
Molecular structure
of hemicellulose,
[9]
11
2 Technology overview
Cellulose
Plants consist to approx. 40-80 mass-% (dry) of cellulose, being a
linear polysaccaride occasionally made up of over 10000 glucose
units, see Figure 2-4. The elemental formula of polymeric cellulose is
(C6H10O5)n.
Figure 2-4:
Molecular structure
of cellulose, [9]
2.1.4 Properties
This work focuses on woody biomass for power generation, the most
important material properties of which are discussed in the following.
2.1.4.1 Moisture
The intrinsic moisture content comprises the water in the pores and hy-
droscopic bound water in the solid structure of the woody material, [10].
The extrinsic moisture content accounts for complementary moisture
originating from the weather conditions during harvesting and storage.
Regarding wood as feedstock for power generation applications, the ex-
12
2 Technology overview
with Eq. 2
M =water load; ms = moist mass; mH 2O = extrinsic water mass
md = mass of specimen after 24 h drying at 102 °C at 1 atm
with Eq. 5
mstart = original sample mass
mend =sample mass at end of volatile matter test procedure
13
2 Technology overview
terial is considered as ash. The ash fraction is the quotient of ash mass
and sample mass before the test procedure.
mend
X Ash =
mstart
with Eq. 6
mstart = original sample mass
mend =sample mass at end of ash test procedure
The fixed carbon fraction is used to estimate the amount of coke that
can be obtained from a fuel sample. It is calculated according to:
X FC = 100 − X VM − X Ash Eq. 7
The proximate analysis gives a first measure for the suitability of the
analyzed material for thermochemical conversion technologies.
The "ultimate analysis" provides information about the elemental com-
position of solid fuels. Plants mainly consist of carbon, hydrogen, oxy-
gen and nitrogen. Besides, a variety of other chemical elements can be
found in biomass with more or less important concentrations. The impor-
tance of these trace elements depends on the subsequent processes.
For thermochemical conversion, sulfur, chlorine, potassium and sodium
are most important. The rest of the biomass material (e.g. heavy metals)
is usually summed up under the term ash. Table 2-2 gives the proximate
and ultimate analysis of biomass materials in comparison to coal.
The main difference between biomass and solid fossil fuels can be
found in the oxygen to carbon and the hydrogen to carbon ratios, [11].
14
2 Technology overview
/ kg
J /k g
40 MJ/k
Van Krevelen dia- Cellulose
35 MJ
30 M
gram for various 1.6
dry solid fuels, [12],
1.4 Wood
with calculated dry-
Lignin
basis iso-LHV lines
0.4
J/kg
g
0.2 Anthracite
/k
/k
MJ
MJ
25 M
20
15
0.0
0.0 0.2 0.4 0.6 0.8
Atomic O/C Ratio
Figure 2-5 illustrates the significance of these ratios on the lower heating
value of young biomass and solid fossil fuels in a so called Van Kreve-
len diagram. The comparably higher heating value of solid fossil fuels
can be explained with the low number of carbon-oxygen bonds, [6],
which again is a result of the natural diffusion process of biomass bound
oxygen under high pressure and temperature. However, hydrogen also
diffuses away under these conditions. Hence, at a certain stage of the
process, the heating value begins to decrease asymptotically against
the heating value of pure carbon which is 34.8 MJ/kg (daf).
The alkali metal content (Na, K, Ca, etc.) of biomass is of special inter-
est regarding thermochemical conversion processes. Alkali metals can
react with silica present in the biomass or originating from soil material
introduced into the process during harvesting. The resulting sticky, liquid
phase can provoke slagging in furnaces and boilers. Another known
problem which can arise from alkali metals is high temperature corrosion
of e.g. gas turbine blades. The catalysts employed in fuel cells could be
poisoned by alkali metals, [13]. However, alkali metals were also found
to have a positive effect as they reduce the tar formation, while increas-
ing the char formation, in pyrolysis and gasification processes, [14, 15].
15
2 Technology overview
2.2 Gasification
2.2.1 Biomass conversion overview
The type of biomass has a strong impact on the choice of conversion
technology due to varying moisture contents and proportions of the
three main components; cellulose, hemicellulose and lignin. In addition,
the choice of conversion technology depends on the subsequent energy
converting device and its specific requirements.
intermediate prod-
ucts and end-use
Heat Combustible
Combustiblegas Combustible
Combustibleliquid
applications for the Heat gas liquid
combined genera-
tion of power and
Steam
Steamcycle
heat cycle Fuel
Fuelcell
cell Combustion
Combustionengine
engine Gas
Gasturbine
turbine
Heat
Heatand
andElectricity
Electricity
16
2 Technology overview
Oxidation
Heat is produced via exothermal partial or total oxidation of carbon,
Eq. 10 and Eq. 11, at around 1000 to 1600 °C, [19]. The carbon origi-
nates from the devolatilization and reduction sub-processes.
⎛ kJ ⎞
C + 0.5 O2 → CO with ⎜ − ΔH R0 = 110.5 ⎟ Eq. 10
⎝ mol ⎠
⎛ kJ ⎞
C + O2 → CO2 with ⎜ − ΔH R0 = 398.8 ⎟ Eq. 11
⎝ mol ⎠
Reduction
Most of the hydrogen, carbon monoxide, carbon dioxide and methane
found in producer gases from thermochemical biomass conversion is
produced in the reduction sub-process at temperatures between 600
and 1000 °C, [19]. The required heat, steam and carbon dioxide is
usually produced in the oxidation sub-process. The carbon steam re-
action, Eq. 12, is the principal reaction for the gasification of carbon.
Besides, carbon can be gasified via the Boudouard reaction, Eq. 13.
⎛ kJ ⎞
C + H 2O ↔ CO + H 2 with ⎜ − ΔH R0 = −131.3 ⎟ Eq. 12
⎝ mol ⎠
⎛ kJ ⎞
C + CO2 ↔ 2CO with ⎜ − ΔH R0 = −172.5 ⎟ Eq. 13
⎝ mol ⎠
⎛ kJ ⎞
C + 2 H 2 ↔ CH 4 with ⎜ − ΔH R0 = 74.8 ⎟ Eq. 15
⎝ mol ⎠
⎛ kJ ⎞
CO + 3H 2 ↔ CH 4 + H 2O with ⎜ − ΔH R0 = 206.1 ⎟ Eq. 16
⎝ mol ⎠
18
2 Technology overview
Devolatilization
Devolatilization denotes the thermal decomposition of biomass with-
out the participation of an oxidant at temperatures between 200 and
600 °C, [19]. The three main constituents of biomass; cellulose,
hemi-cellulose and lignin, see section 2.1.3, exhibit different thermal
decomposition behaviors which originate from their molecular struc-
ture, [5]. Devolatilization yields gaseous, liquid and solid products.
Gaseous products are mainly hydrogen, carbon monoxide and diox-
ide as well as methane and higher hydrocarbons. Liquid products are
hydrocarboneous species that condense at room temperature, gen-
erally summarized under the term "tars". Non reacted carbon is the
main solid product. Due to transport phenomena in pores and on the
surface, the kinetics of devolatilization depends on the shrinking be-
havior and structure of the feed particles, [21]. Further, the heating
rate and the heat transfer play important roles. The thermal decom-
position reactions are endothermic. A more detailed discussion of the
devolatilization of biomass is given in [20].
Drying
The moisture of biomass is vaporized in the drying sub-process at
temperatures below 200 °C. The amount and nature of moisture de-
termine the required heat quantity. Especially the vaporization of the
water bound in the cell walls involves strengthening of the wood fi-
bers and is accompanied by shrinkage of the biomass, [5]. More de-
tails on recent developments in biomass drying are given in [22].
Table 2-3 summarizes thermochemical conversion technologies with
typical air/steam-to-fuel ratios, process temperatures and products.
19
2 Technology overview
20
2 Technology overview
Figure 2-7:
Impurities
Types of impurities
and their typical
phase of appear-
ance in biomass- Solid Liquid Gaseous
derived producer
gases
Alkali and Nitrogenous Sulfurous
Particles Tars Halides
heavy metals species species
As the gasification is the origin for all impurities found in producer gases,
a short summary is given at this point as a basis for the discussion of
different gasification technologies. The removal of impurities and their
impact on the operation of SOFCs is discussed in the sections 2.3.1 and
2.4.7. As can be seen in Figure 2-7, impurities can be present in the
produced gas in the solid, liquid and gaseous phase. Following types of
impurities can be distinguished:
Particles are solid agglomerations of unreacted carbon and ash.
Alkali metals can be present in the solid state condensed on particles
at temperatures below 600 °C, [26], and in the liquid phase as aero-
sols or even gaseous phase at temperatures above 800 °C, [25].
Tars are a mixture of diverse organic species of oxygenated and
aromatic character formed during the thermal decomposition of the
biomass in the devolatilization sub-process and subsequent reactions
with reactive char. The complex mechanisms of primary and secon-
dary tar formation are discussed in depth in [20, 27]. Depending on
their molecular structure, tars may be in the liquid state as aerosols
21
2 Technology overview
Drying Drying
Devolatilization Devolatilization
Reduction Oxidation
Oxidation
Reduction
Updraft wood gasifiers are fixed bed reactors. The wood is introduced
at the top of the gasifier while the gasification agent, typically air, is in-
troduced at the bottom through a grate. The producer gas is extracted at
the top of the gasifier after passing through the wood material present in
the gasifier. In contrast to the gasification agent and the producer gas,
22
2 Technology overview
the wood particles move downwards in the gasifier, running through the
sub-processes of gasification in the order "drying", "devolatilization",
"reduction" and finally "oxidation". Char is partially combusted in the oxi-
dation zone at temperatures between 1600 and 1000 °C. After the oxi-
dant is spent, the hot exhaust gas reacts with char in the reduction zone
at temperatures between 1000 and 600 °C. Further up in the gasifier,
the temperature of the producer gas becomes to low for reducing reac-
tions. The thermal decomposition of the wood takes place in the subse-
quent devolatilization zone at temperatures between 600 and 200 °C.
Before exiting the gasifier, the producer gas dries the fresh wood at tem-
peratures below 200 °C in the drying zone.
Updraft gasification producer gases are relatively cold with outlet tem-
peratures typically between 200 and 300 °C. In [28], producer gas tem-
peratures of 75 °C are reported. The cold gas efficiency including all
combustible species in the producer gas can reach values in excess of
90 % as a consequence of the low sensible heat content of the producer
gas, [4]. Due to the use of air as gasification agent, the heating value of
updraft gasification producer gases is low in the range of 5 MJ/mn3 on a
dry and tar free basis (dtf), [20]. Since the tars formed in the devolatiliza-
tion zone only pass the colder drying zone before exiting the updraft
gasifier, cracking is not likely to occur resulting in very high tar loads
with values up to 150 g per mn3 (dtf), [4]. The tar mix mainly consists of
oxygenated species such as phenols and acids as well as light aromat-
ics, [29]. Heavier aromatic species might be present in the condensed
phase. The particle load is usually low due to the filtration effect of the
wood material bed which the producer gas passes on its way to the
gasifier exit at its top, [4]. Alkali metals can be assumed to stay almost
completely in the solid phase. For the same reason as for the high tar
load, the presence of a notable amount of organic sulfur species is likely.
The updraft gasification process is very robust towards the size of the
wood particles and their humidity. The large internal drying zone allows
the conversion of wood with up to 50 % humidity. Updraft gasification
reactors are well scalable to a size of 20 MW thermal input, [28], and the
costs are relatively low due to the simple reactor concept.
Downdraft wood gasifiers are the second common type of fixed bed
reactors. The wood is introduced at the top of the gasifier, while the
gasification air is either introduced there as well or right above the
hearth zone in the lower part of the reactor. The producer gas exits the
gasifier at the bottom. Hence, wood particles and gasification air move
23
2 Technology overview
a rated thermal capacity of 4.8 MW and was built during 1993 in Har-
boore, Denmark, by the Babcock & Wilcox Volund A/S. The producer
gas has a temperature of 75 °C and an average tar load of 70 g/mn3 (dtf).
The gas is cleaned by means of a gas scrubber and electrostatic pre-
cipitator. With two gas engines producing 1536 kW electrical power, the
overall power efficiency departing from wood chips with approx. 50 %
humidity is in excess of 30 %, [29]. 3.2 MW of heat at 90 °C are deliv-
ered to a district heating grid. Since commissioning, the updraft gasifica-
tion reactor has been operated several ten thousand hours with 8000
hours per year in average. The gas engines were commissioned in 2003
and operated until 2005 for over ten thousand hours, [25].
The downdraft reactor (right) has a rated thermal input of 0.8 MW. It was
built during 2001 in Spiez, Switzerland, by Pyroforce AG. The almost tar
free producer gas exits the gasifier at approx. 650 °C. Gas cleaning in-
volves the addition of an adsorbent to the producer gas and its subse-
quent removal together with all other particles in a particle filter. A single
gas engine is installed producing a maximum of 200 kW of power and
270 kW of heat. Since commissioning, the gasification reactor has been
operated for 25000 hours and the gas engine for 15000 hours, [31].
Steam
and air Reduction
Ash
Fluidized bed gasifiers are categorized via their fluid dynamics and
heat transfer into stationary and circulating fluidized bed gasifier types.
Generally, the wood is introduced near the bottom of the gasifier where
25
2 Technology overview
of the feed but introduced by other means such as heat pipes, [33]. Sta-
tionary fluidized bed gasification is very well scalable for plant sizes be-
tween 1 to 50 MW thermal input, [4].
Circulating fluidized bed gasifiers feature almost the same properties
as their stationary counterparts, except that the flow velocities are con-
siderably higher. Consequently the mixing of bed material and wood
particles is better which yields fully isothermal conditions. Moreover, bed
material and unreacted char are continuously discharged from the gasi-
fication zone. In standard circulating fluidized bed reactors, the dis-
charged bed material and char are reintroduced to the gasification zone
where the required heat for the gasification reactions is produced via di-
rect partial combustion of the feed. Two-zone circulating fluidized bed
reactors however use the discharge mechanism for the spatial separa-
tion of gasification and combustion reactions. This way, air can be used
as oxidant in the combustion zone and steam as gasification agent in
the gasification zone yielding a non-diluted product gas. The heat for the
gasification reactions is transported from the combustion zone to the
gasification zone via the circulating bed material.
The producer gas from circulating fluidized bed reactors has basically
the same properties as that of stationary fluidized bed reactors. Excep-
tions are the higher heating values of two-zone concepts around 14.2 to
18.1 MJ/mn3 (dtf), [4], and the typically lower tar load in the order of 1 to
20 g/mn3 (dtf), [25]. This can be assigned to the frequently used reactive
bed materials such as olivine, [35]. Reasonable reactor sizes are be-
tween 20 to 200 MW thermal input, [4].
Figure 2-10 shows two fluidized bed gasification reactors employing
steam as gasification agent and air for the combustion of unreacted char
to produce the required heat for the gasification reactions. The station-
ary fluidized bed reactor (left) uses sodium heat pipes for the heat ex-
change between the combustion and the gasification chamber. The
"Biomass heat pipe reformer" (BHPR) was developed and patented by
the Technische Universität München. The first prototype reactor with a
thermal capacity of 150 kW was commissioned in 2002 and operated for
72 hours non-stop. The producer gas had a temperature around 800 °C
and was directly flared. In future, the product gas shall be produced with
a cold gas efficiency around 70 % , [33], and converted to electricity via
micro gas turbines and high-temperature fuel cells.
27
2 Technology overview
Figure 2-10:
Technical outlines Biomass Steam Producer gas Producer gas Exhausts
of the Biomass
heat pipe reformer
Combustion chamber
from TU Munich/D Exhausts
(left), [36], and the
Fast internally cir- Gasification
chamber
culating fluidized
bed reactor in Gasification
Güssing/AT (right), Filtration Biomass
chamber
[37] bed
Bed
Combustion Heatpipes material
chamber
Steam
Combustion
Combustion air air
28
2 Technology overview
Figure 2-11:
Technical outlines
of the Carbo-V
gasifier developed
by CHOREN/D
(left), [39], and the
GSP gasifier de-
veloped by the
Deutsches Brenn-
stoffinstitut/DDR,
[38]
Figure 2-11 shows the technical outlines of two entrained flow biomass
gasifiers. CHOREN Industries GmbH has developed the Carbo-V® gasi-
fier (left) and is currently building a Biomass-to-liquids (BtL) plant in
Freiberg, Germany, based on this gasification technology. The biomass
29
2 Technology overview
is first gasified with very little air, producing a highly tar laden gas and
char. The gas is partially oxidized with oxygen in the entrained flow gasi-
fier. The hot gases are cooled down to approx. 800 °C in the reduction
section of the gasifier by means of quenching with grinded char from the
low temperature gasification. The char is simultaneously gasified, [39].
The GSP gasifier (Gaskombinat Schwarze Pumpe, right) was developed
for the gasification of salt-containing brown coal and therefore suits the
gasification of biomass with high ash content well. In fact, the ash acts
as protective layer for the reactor walls and is therefore required. As the
gasifier requires a liquid feed, the Forschungszentrum Karlsruhe has
developed a pyrolysis process that is suited for the production of slurry
from various types of biomass. With 1200 °C, [38], the producer gas is
hotter than that of the CHOREN gasifier as no reactor internal quench-
ing is conducted. Table 2-4 summarizes the above paragraphs.
30
2 Technology overview
2.3.1.1 Particles
Particles in producer gases can cause erosion and blocking in the
equipment downstream of the gasifier. Further, particles are subject to
emission limits. For these reasons, the removal of particles is an integral
part of every gas cleaning system for producer gases from biomass
gasification. Ordered by their maximum operational temperature, the fol-
lowing particle removal technologies can be distinguished:
Wet scrubbers
Electrostatic precipitators
Cyclones
Barrier filters
Wet scrubbers employ a liquid, often water, for the removal of particles
from gas streams. Their maximum operational temperature is usually
below 100 °C which requires a cool down of the producer gas. The par-
ticles are captured in wash fluid droplets which are then removed from
the gas stream in a demister. The most common wet scrubber type is
the venturi scrubber. It allows removing particles below 1 μm diameter
and achieving particle loads around 10 to 20 mg/mn3 (dtf). However,
venturi scrubbers cause a pressure loss between 30 and 200 mbar.
31
2 Technology overview
Figure 2-12:
Clean gas
Working principle
of cyclones and Pulse gas
rigid barrier filters, Raw gas Valve Injector
[41] Venturi
Clean gas
nozzle
Dip pipe Pulse
control
Cyclone Dust
body cake
Filter
element
Particles Particles
32
2 Technology overview
33
2 Technology overview
2.3.1.3 Tars
Producer gases can contain considerable amounts of tars. At tempera-
tures below their dew point and/or at elevated pressures, tars condense
causing operational problems due to the formation of droplets which ac-
cumulate to sticky films on cold surfaces of e.g. pipes and other equip-
ment. Besides the condensation related issues, tars may cause carbon
deposition problems at very elevated temperatures. Systems employing
gas engines and gas turbines as end use devices require almost com-
34
2 Technology overview
tor. The potential advantage of in situ tar decomposition is that the tars
may be decomposed at the same time that they are formed. However,
experiments showed that in fixed bed gasification reactors the contact
time between catalyst and tars is insufficient for complete tar decompo-
sition and that in fluidized bed reactors the attrition of the catalyst parti-
cles causes a fast deactivation. Further, the means of temperature con-
trol for in situ tar decomposition are limited. The latter is the main advan-
tage of external tar decomposition reactors, which however involve addi-
tional costs. State-of-the-art non-metallic catalyst materials for tar de-
composition are dolomites, zeolites and calcites. These materials are
cheap, insensitive towards most impurities present in producer gases
and allow tar decomposition rates of up to 99 %. Metallic tar decomposi-
tion catalysts employ a wide range of materials such as Ni, Mo, Co, Pt
and Ru in pure form or as mixtures supported on silica alumina, zeolites
etc. The main advantages of metallic catalysts are that they allow almost
complete tar decomposition and simultaneous decomposition of ammo-
nia. However, the long-term application in biomass gasification systems
has not been demonstrated yet and the costs are higher than for non-
metallic catalysts. Sections 2.3.2.1 and 2.3.2.2 give additional informa-
tion for the catalytic decomposition of hydrocarbons and tars.
2.3.1.4 Halides
The average chlorine content of untreated wood is low, which yields low
concentrations of hydrochloric acid in producer gases. If required, hy-
drochloric acid can be removed from producer gases using wet scrub-
bers which are well tested and commercially available. To avoid heat
losses, high temperature capable adsorbent materials can be used such
as e.g. lime, Eq. 17.
CaO + 2 HCl → CaCl2 + H 2O Eq. 17
36
2 Technology overview
38
2 Technology overview
⎛ kJ ⎞
CH 4 + H 2O → CO + 3 ⋅ H 2 with ⎜ − ΔH R0 = −206 ⎟ Eq. 21
⎝ mol ⎠
39
2 Technology overview
Eq. 21 shows the high heat requirement of the STR and explains why
reactor designs are typically limited by heat transfer rather than the re-
action kinetics. STR of most hydrocarbons takes place only over suit-
able catalysts and at high temperatures typically over 800 °C. Group VIII
metals are employed as STR catalysts, [52], of which Ni is the mostly
used material due to cost considerations. SOFCs offer all factors for an
integrated STR process, see also section 2.4.3. The conversion can be
performed cell internally or externally using the SOFC waste heat.
The main problem of the STR process is carbon deposition. High steam
partial pressures help to prevent carbon deposition on the catalyst mate-
rial. Typically, steam-to-carbon ratios (SC) of two and higher, [53], are
required to prevent carbon deposition. The tendency towards carbon
deposition increases with rising molecular weight of the hydrocarbon
species in the educt gas, especially for polycyclic molecules, [54].
Poisoning of the STR catalyst is another issue. Hydrocarbon fuels are
usually contaminated with small amounts of sulfur, which is likely to
chemisorb on any metallic surface. Hence, a desulphurization step has
to be carried out prior to the reformer, see section 2.3.1.6.
Oxygenated organic compounds can be steam reformed using Ni-based
catalysts, [55, 56]. The corresponding reaction can be stated as:
⎛ m ⎞
Cn H m Ok + (n − k ) ⋅ H 2O → n ⋅ CO + ⎜ n + − k ⎟ ⋅ H 2 with (− ΔH < 0) Eq. 22
⎝ 2 ⎠
Although Eq. 22 appears similar to Eq. 20, there are significant differ-
ences between the reaction mechanisms of pure hydrocarbons and
oxygenates. According to [55], the oxygenated organic compounds
found in biomass-derived gases are thermally unstable and decompose
without forming significant amounts of coke at the operating conditions
of STR processes. The formed intermediates are instantly converted to
hydrogen and carbon monoxide. Experiments showed that oxygenate
STR allows lower reforming temperatures and higher space velocities
than the comparable hydrocarbon STR process, [57]. The following sec-
tions provide an overview of state-of-the-art hydrocarbon STR reactors
and integration possibilities of the STR process into SOFC designs.
Figure 2-13:
External reforming
reactor with inte-
grated heat ex-
changer designed
by Haldor Topsoe,
[58]
Figure 2-14:
Indirect internal
reforming concept
by the Siemens
AG, [59]
41
2 Technology overview
42
2 Technology overview
The ATR process requires less steam than the STR process. The re-
quired heat is generated via partial oxidation of the fuel, avoiding the
need for a complex heat management, [58]. Both processes are well
suitable for the conversion of biomass-derived fuel gases. For oxygen-
ated organic compounds, Eq. 25 and Eq. 26 are reformulated as follows:
n−k m
C n H m Ok + ⋅ O2 → n ⋅ CO + ⋅ H 2 with (− ΔH > 0) Eq. 25
2 2
n−k n−k ⎛n−k m⎞
Cn H m Ok + ⋅ O2 + ⋅ H 2O → n ⋅ CO + ⎜ + ⎟ ⋅ H2
4 2 ⎝ 2 2⎠
Eq. 26
with (− ΔH ≥ 0)
CPO and ATR of oxygenated organic compounds have been shown us-
ing pyrolysis gas from cedar wood powder, [64]. It was found, that the
tar content of the pyrolysis gas was almost completely converted to hy-
drogen and carbon monoxide.
43
2 Technology overview
44
2 Technology overview
Figure 2-15:
Comparison of fuel
cell technologies
Pictures: [66]
45
2 Technology overview
Figure 2-16:
Working principle
of an internally
reforming anode-
supported SOFC,
[69]
fuses through the anode and participates in STR, see section 2.3.2.1,
and shifting reactions (SR), Eq. 14.
The required steam for both reactions is generated in the electrochemi-
cal oxidation of hydrogen at the anode triple phase boundary (TPB)
where the fuel gas meets electrode and electrolyte material simultane-
ously. The hydrogen produced in the reforming reactions further diffuses
to the anode TPB, while the major part of the formed carbon monoxide
is converted to hydrogen and carbon dioxide via exothermal shifting re-
actions. Some of the carbon monoxide may be directly converted to car-
bon dioxide at the anode TPB. The heat of the shifting and of electro-
chemical reactions is partially consumed by the reforming reactions. At
the TPB of the anode, hydrogen and carbon monoxide are oxidized by
oxygen ions, at the same time releasing electrons:
H 2 + O 2− → H 2O + 2e − Eq. 27
CO + O 2− → CO2 + 2e − Eq. 28
The free electrons flow through an external circuit to the cathode where
oxygen is converted to oxygen ions at the TPB of the cathode consum-
ing the free electrons:
1
O2 + 2e − → O 2− Eq. 29
2
The oxygen ions pass the electrolyte via vacancies in the crystal struc-
ture, hence closing the electrical circuit. In sum, the electrochemical re-
actions can be written as:
1
H2 + O2 → H 2O Eq. 30
2
1
CO + O2 → CO2 Eq. 31
2
Voltage [V/cell]
0.8 Ohmic or resistance losses
0.4
0.0
Current denisity [mA/cm2]
Current density
49
2 Technology overview
Temperature
Figure 2-17 shows that the activation polarization increases fast with
rising small current density values and then stabilizes on a certain
level.
Concentration losses
The electrochemical reactions (Eq. 27, Eq. 28, Eq. 29) take place at
the TPB. Due to slow diffusion processes, the product partial pres-
sures are higher and the reactant partial pressures are lower at the
TPB compared to the gas flow channels. Consequently, using the
TPB instead of the gas flow channel partial pressures yields a lower
Nernst voltage. The difference between the TPB and the gas flow
channel Nernst voltage is referred to as concentration loss. Concen-
tration losses become important at high current densities, where they
increase against an asymptotic maximum current, see Figure 2-17. A
further increase of the current is impossible because at this point the
concentrations of the reactants at the TPB are equal to zero due to
instant consumption.
50
2 Technology overview
the cathode tube. The fuel flows at the outside of the cell. The depleted
air exits the cell at the open end, where it is combusted with the partially
depleted fuel, 50 to 90 % respectively.
Fuel
Interconnect
Cathode
Oxidant
Electrolyte
Anode
The tubular cell design is at present the most reliable and robust design,
[72]. However, the tubular cell design features only low power density,
[73]. This is due to high ohmic in-plane resistance resulting from long
current paths, [51], and high concentration polarization losses resulting
from long diffusion paths for the oxygen from the bulk gas phase to the
cathode TPB. The employed coating techniques, e.g. electrochemical
vapor deposition (EVD), carry a high cost, [51], leading to high overall
cell costs. Finally, thermal stress can be a concern, [73].
The flattened tubular cell design, commonly referred to as HPD design
(high power density), was also developed by the Siemens AG, see
Figure 2-18. It allows for about 30 to 40 % higher volumetric power den-
sities compared to standard tubular cells, [74], due to better stacking
possibilities, [59]. The seal-less design, similar to the standard tubular
design, avoids gas cross-over problems, [74]. Internal ribs were intro-
duced to reduce the high ohmic resistance of the standard tubular de-
sign by shortening the current paths, [75].
The main difference between planar and tubular cells is that their flow
channels are usually defined by the geometry of the interconnector
plates. Three flow configurations can be distinguished, namely co-flow,
cross-flow and counter-flow, see Figure 2-19.
51
2 Technology overview
Figure 2-19:
Flow configurations
of planar SOFCs Interconnect
Co-Flow
Cathode
Electrolyte
Anode
Ox
id
an
t
Fu
el
Interconnect
Cross-Flow
Cathode
Electrolyte
Anode
t
an
id
Ox
Fu
el
Interconnect
Counter-Flow
Cathode
Electrolyte
Anode
Ox
ida
nt
The planar cell design features higher area and especially volume spe-
cific power densities than standard tubular designs due to more efficient
current collection, [73], and simple in-series connection of the single
cells to stacks, [68]. The interconnectors may consist of ceramic or me-
tallic materials. There exists a high cost reduction potential of the planar
design as cost efficient production processes for high volume manufac-
turing of the flat cell components are available, [51]. The drawback of
these cost efficient manufacturing techniques is that the maximum cell
size is limited, [51]. The most important issue of the planar design is that
sealings are necessary in order to avoid gas cross-over of oxidant and
fuel gases.
Thermal stress is an important problem for planar cells, which are usu-
ally less robust than tubular cells. Thermal stress can result from multi-
ple sources such as material property mismatches, poorly balanced en-
dothermic and exothermic reactions and transients of operational pa-
rameters, [51].
52
2 Technology overview
ca. 50 μm
Anode supported
ca. 50 μm
ca. 10 μm
Passive supported
ca. 50 μm
ca. 10 μm
ca. 10 μm
Cathode
Electrolyte
ca. 40 μm
Anode
ca. 100 μm Passive Support
53
2 Technology overview
The anode-support design has gained a lot of interest recently. The ma-
jor advantage is that the thickness of the electrolyte can be heavily re-
duced as the mechanical strength of the cell is provided by the anode.
Consequently, the ohmic losses caused by the electrolyte are heavily
reduced, [77]. The good electrical conductivity of the supporting anode
material allows for low operation temperatures around 700 to 800 °C,
[75]. This, and the matching thermal expansion rate of the usually em-
ployed Ni-based anode materials makes the application of ferritic chro-
mium interconnects possible, again lowering the internal resistance of
the cells, [75]. Finally, the thick porous anode features a large specific
surface and void volume which improves the conditions for Ni-catalyzed
reactions, such as steam reforming of hydrocarbons and water gas shift.
Although the anode-support design has significant advantages, it bears
a relatively low mechanical strength compared to other support designs.
Consequently, thermal stress and vibrations are the most important is-
sues, [76].
The passive-support design allows very thin anode, electrolyte and cath-
ode layers. These functional layers can be deposited via coating tech-
niques onto either metallic or ceramic support structures.
Metallic support structures offer high mechanical strength, Redox stabil-
ity, [75], and the potential for operation temperatures below 600 °C. The
start-up times can therefore be reduced to values suitable for auxiliary
power units (APUs), [76]. Other advantages of metallic support struc-
tures are the reduced need for expensive ceramic materials, [78], and
the applicability of conventional joining techniques, [76]. The high ther-
mal conductivity of metallic materials reduces spatial thermal gradients.
Ceramic support structures provide reasonable mechanical strength at
relatively low costs but struggle with low gas permeability and high con-
centration polarization losses.
2.4.6 Materials
Material and fabrication process improvements play an important role in
reducing the stack costs. For instance, sulfur tolerant anode materials
are presently an important research field because these materials would
largely simplify the required gas cleaning and hence lower the costs of
SOFC systems. Another important research area is in lowering the op-
eration temperatures, enabling the use of cheaper materials for all cell
elements. The working principle of the SOFC leads to the conductivity
requirements shown in Table 2-5, [68].
54
2 Technology overview
2.4.6.1 Anode
Apart from the properties mentioned above, SOFC anode materials
should feature, [68]:
High surface to volume ratios. The porosity, tortuosity and average
pore radius play an important role for the size of the TPB and the gas
diffusion. These parameters are to a certain extent determined by the
production processes.
High catalytic activity towards electrochemical fuel oxidation, hydro-
carbon reforming and water gas shift reactions.
Chemical compatibility with electrolyte and interconnect materials.
Mechanical strength in case that the cell is anode supported.
Tolerance towards fuel contaminants.
55
2 Technology overview
Figure 2-21:
SEM picture of the
contact area be-
tween anode and
electrolyte, [80]
56
2 Technology overview
Low sintering temperatures around 1000 °C, allowing for cheap pro-
duction processes, [76].
Improved resistance against Ni-NiO cycling, [73].
Strongly improved electrochemical oxidation activity as CGO is a
mixed ionic and electronic conductor (MIEC). MIECs enable the oxi-
dation of the fuel gas in the entire catalyst volume, [68].
However, Ni-CGO anodes have the following disadvantages:
The low sintering temperature limits the operating temperature and
hence the applicability of Ni-CGO to the anode-support design, [82].
If other dopants than nickel are used, e.g. Cu, Au, etc., the STR activ-
ity is comparably low, [68].
Lanthanum strontium chromium perovskite (LSCr) anodes offer good
Redox and sulfur tolerance. However, the electrochemical performance
is lower than that of nickel-based anodes, [73].
Titanate anodes have drawn a lot of interest recently but their properties
are hardly known yet. The material is stable in the reducing atmosphere
and resistant against coking and sulfur poisoning. However, titanate an-
odes show only poor electronic conductivity, thus requiring highly con-
ductive current collectors to keep the ohmic losses small, [82].
2.4.6.2 Electrolyte
Electrolyte materials should satisfy the following requirements besides
those mentioned in section 2.4.6:
Chemical compatibility with the electrode materials.
Mechanical strength in case that the cell is electrolyte-supported.
Gas tightness to prevent gas cross-over phenomena causing a strong
decrease of the cell voltage and power output, [68].
State-of-the-art SOFCs employ yttria stabilized zirconia (YSZ) as elec-
trolyte material. YSZ typically contains 8 mol-% of yttria. The advan-
tages of YSZ electrolytes are, [82]:
Acceptable ionic conductivity between 800 and 1000 °C.
Excellent chemical stability in reducing and oxidizing atmospheres.
Almost no electronic conductivity, [68].
YSZ shows only low ionic conductivity below 800 °C and is therefore not
well suitable for low temperature SOFCs. Being a very refractory mate-
rial, YSZ requires higher sintering temperatures than the other cell com-
57
2 Technology overview
Figure 2-22:
Ionic conductivities
of selected high
temperature con-
ductors, [79]
58
2 Technology overview
2.4.6.3 Cathode
Apart from the properties mentioned in section 2.4.6, SOFC cathode
materials should feature, [68]:
High porosity to allow unrestricted gas diffusion to the cathode TPB.
Catalytic activity towards electrochemical oxygen ionization.
Chemical stability in oxidizing atmospheres.
State-of-the-art SOFCs use La1-xSrxMnO3-δ, (lanthanum strontium man-
ganite or LSM) as cathode material, where x is typically 0.15 to 0.25.
Porosities are in the range of 25 to 40 %. Composite cathodes with YSZ
to LSM ratios of 1:1 recently appeared and may be widely used in the
near future, [79].
The advantages of LSM cathodes are, [82]:
Good phase stability.
Good thermal conductivity match with YSZ electrolytes.
Well characterized in SOFC systems.
However, LSM is an almost pure electronic conductor with low conduc-
tivity. The electrochemically actives sites are therefore limited to the
TPB, leading to a moderate catalytic activity.
Recently investigated mixed ionic-electronic conductors (MIEC) allow
the ionization of oxygen in a larger active volume compared to standard
LSM cathodes, where they are limited to the TPB area. This strongly re-
duces the cathode polarization losses, [82].
Most of the recently investigated cathode materials are mixtures of lan-
thanum and strontium with iron and/or cobalt (LSCF, LSF and LSC). The
main concerns with these materials are low phase stability, yet unknown
interface reactions and poor mechanical properties due to a strong mis-
match of their thermal expansion coefficient with most electrolyte mate-
rials, [73]. For example, LSC cathodes were found to form resistive lay-
ers when combined with YSZ electrolytes. Hence, protective layers or
other electrolyte materials have to be employed with LSC cathodes, [68].
Samarium-based perovskite cobaltites (e.g. Sm5Sr5CoO3-x) are exten-
sively investigated in Japan as they are amongst the best MIE conduc-
tors and feature smaller thermal expansion coefficients than LSCF, thus
fixing the compatibility issue with standard electrolyte materials, [82].
59
2 Technology overview
2.4.6.4 Interconnect
Interconnects, also named bipolar plates in planar stack systems, estab-
lish the mechanical and electrical connection between the anode of one
cell and the cathode of the next cell in SOFC stacks. There are different
interconnect concepts under development which are suitable for differ-
ent stack designs. Metallic, conducting ceramic and non-conducting ce-
ramic interconnects with conducting vias can be distinguished.
Metallic interconnects usually consist of ferritic chromium. The chro-
mium (up to 17 %, [76]) is used to improve the high-temperature oxida-
tion resistance and to realize a sufficient electronic conductivity, [73].
Metallic interconnects feature:
High electronic conductivity at low temperatures, [75].
Production wise almost unlimited size, [75].
Good thermal conductivity, thus reducing temperature gradients, [75].
High mechanical integrity, [83].
Lower production cost than ceramic interconnects, [83].
However, metallic interconnect exhibit a strong mismatch of the thermal
expansion coefficient with all other cell components, [75]. The chromium
can cause evaporation and deposition induced issues leading to severe
cell degradation, [73]. Protective coatings are therefore mandatory, [84].
Conductive ceramic interconnects are usually made of doped lanthanum
chromite perovskites. The specific dopant (typically, Sr, Ca or Mg) and
concentration allow matching the thermal expansion coefficients of the
cathode material and the other materials used for the remaining fuel cell
components. Conducting ceramic interconnects feature negligible corro-
sion and low degradation. However, there are negative aspects to ce-
ramic interconnects, [75]:
The size is limited by available manufacturing processes.
High sintering temperatures are needed.
The thermal expansion coefficient is different in oxidizing and reduc-
ing atmospheres, which may cause thermal stress issues.
The electrical and thermal conductivity is poor inducing thermal stress.
The ceramic materials are difficult to process because chromium
evaporation at high temperatures can lead to poor densification, [79].
A less common approach is the use of non-conducting ceramic inter-
connects with e.g. silver vias, [85]. This approach is well established in
60
2 Technology overview
2.4.6.5 Sealant
In addition to anode, electrolyte, cathode and interconnects, SOFCs re-
quire sealant materials to isolate anode and cathode chambers from
each other in planar configurations. Sealants are typically made of
glasses or glass-ceramics, [79].
2.4.7.1 State-of-the-art
Cell degradation may vary by cell and stack size, fuel composition, fuel
utilization and current load. A standardized degradation test procedure
is yet to be defined by the research community. Consequently, a com-
61
2 Technology overview
Figure 2-23:
SEM pictures of a
Ni-YSZ anode after
1500 h of operation
(a, b) and after 8 h
of operation (c, d);
Width of pictures
equals 100 μm,
[96]
62
2 Technology overview
63
2 Technology overview
Figure 2-24:
Lattice structures
of zirconia, [101]
Figure 2-25:
Forms of perma- Blocked Carbon whisker
Encapsulated
catalyst pore with nickel particle
nent deactivation nickel particle
of nickel catalysts
by deposited car- Nickel particle
bon, [40]
Carbon
marized under the term coke. Methane may contribute to coke deposi-
tion via the reverse hydrogasification reaction, Eq. 15. The decomposi-
tion pathway of higher hydrocarbons is however more complex than that
of carbon monoxide and methane. Hydrocarbons adsorb on the catalytic
surface and split off hydrogen. The remaining carbon structures polym-
erize to form coke, [107]. The coking potential of higher hydrocarbons
increases from paraffins via olefins to acetylenes, [108]. Aromats are
also known for their potential towards coke formation, which is several
times higher than that of carbon monoxide and increases with increasing
molecular weight of the aromatic specie, [54]. More detailed information
about carbon and coke formation can be found in [106, 109, 110].
Once deposited, carbon is difficult to remove at temperatures below
800 °C without the presence of a catalyst due to the slow carbon gasifi-
cation reactions. Depending on the SOFC type, the temperatures may
be significantly higher, which results in reduced carbon deposition risk.
However, buildup of carbon deposits was observed in internally reform-
ing SOFCs, [82], causing reduced STR activity and increased anode ac-
tivation losses. The latter is due to shortened active length of the TPB. A
material-based solution for the coking issues is copper doping of the an-
ode material, [81, 111].
therefore one of the most important limiting factors for the scaling up
of planar cells over 100 cm2 active area, [114].
Spatial and temporal thermal gradients
Residual stresses are aggravated by spatial and temporal thermal
gradients. Important spatial thermal gradients result from poorly bal-
anced exothermic water gas shift and electrochemical reactions with
endothermic steam reforming reactions of methane, higher hydrocar-
bons and tars, [51]. Cracking of the AEC and attributed cell failure
was frequently reported as consequence of high spatial thermal gra-
dients, [53, 113, 115]. Spatial thermal gradients may be reduced by
metallic interconnects with high thermal conductivity, appropriate flow
configurations, increased fuel utilization and operational voltages,
[113], as well as higher degree of pre-reforming of the fuel gas in
question. Temporal thermal gradients are most important during heat-
up and start-up phases. The effects of temporal thermal gradients are
similar to those of spatial thermal gradients.
External mechanical loading
SOFC stacks are usually charged with weights to assure correct ori-
entation and good contact between the single cells. The induced me-
chanical stresses may cause stability issues within the AEC, [113].
Figure 2-26:
Synergetic degra-
dation effects of
phosphorous, ar-
senic and sulfur,
[117]
Arsenic reacts with standard Ni-YSZ anodes and yields Ni5As2, [117].
However, Figure 2-26 shows that the exposure of Ni-YSZ anodes to 1
ppmV AsH3 for a period of 700 h at 800 °C resulted in minor cell degra-
dation. According to [117], this can be explained with strong nickel arse-
nic interaction at the surface of the Ni-YSZ anode. Therefore the arsenic
did not penetrate into the anode and did not provoke a change of the tri-
ple phase boundary structure.
Phosphorous in form of PH3 was shown in [117] to react with nickel,
forming Ni3P which extensively agglomerates inside the anode layer.
Figure 2-26 shows that the operation over 700 hours at 800 °C with 2
ppm of PH3 resulted in considerable degradation. At room temperature,
PH3 can be removed with copper doped zeolites, [122].
68
69
70
3 Experiments
3 Experiments
71
3 Experiments
To stack
HX5 HX4 HX3 HX2
Air
900°C 900°C
HEAT
200°C
POWER C SOFC A
1000°C
1000°C 1000°C
1200°C
Combustor
The producer gas is then heated up again to approx. 700 °C in the gas-
to-gas heat exchanger HX1 and superheated to 900 °C in HX2 before
entering the anode channels of the SOFC. The depleted producer gas is
mixed with the depleted cooling air of the SOFC and completely burned
in a catalytic combustor. The hot flue gas is used to superheat the pro-
ducer gas in HX2 and the cooling air of the SOFC in HX3. Heat at a
temperature level of 200 °C is recuperated in HX4. The remainder of the
flue gas heat is used to pre-heat the SOFC cooling air in HX5 before it is
sent to the stack. The SOFC produces direct current at a given cell volt-
age which is converted to AC power in an inverter.
is the subject of several other investigations, [106, 109, 110, 126]. Car-
bon deposition can either result from CO dissociation following the exo-
thermic Boudouard reaction or from the dehydrogenation of hydrocar-
bons and of aromatics in particular, [127].
The probability of carbon deposition can be roughly estimated with equi-
librium calculations assuming carbon depositions as graphite. This ap-
proach was chosen by several other research teams for similar investi-
gations, [128-131]. Figure 3-2 shows the ternary diagram of the carbon-
hydrogen-oxygen system. The inscribed isotherms divide the diagram
into a region where graphite formation can be expected and one in
which it is unlikely to take place. It can be seen that the probability of
graphite formation decreases with increasing temperatures. Air or oxy-
gen gasification of dry wood shifts the point representing the corre-
sponding producer gas towards the oxygen corner. Increasing wood
humidity and steam gasification shifts it towards the H2O point in the
diagram.
Figure 3-2:
C-H-O ternary dia-
gram for the ther-
modynamic graphite C
0.0
formation at equilib- 1.0
rium conditions with T=1000°C
inscribed fuel gases T= 300°C
as received from 0.2
0.8
different biomass
gasifier types
0.4 Graphite 0.6
formation
CO
region
0.6
0.4
Dry wood
CO2
CH4 Downdraft
0.8
Updraft 0.2
Fluidized bed
1.0
0.0
0.0 0.2 0.4 0.6 0.8 1.0
H H2O O
For fixed bed downdraft gasification, the humidity of the wood is limited
to less than 12 %. The cold gas efficiency of this process is typically
around 80 to 85 %. The corresponding producer gas is close to the con-
73
3 Experiments
nection line between the dry wood point and the oxygen corner of the
diagram as air is employed as gasification agent.
The updraft gasification process also employs air as gasification agent,
but allows for the conversion of wood with humidity up to 50 %. Conse-
quently, the corresponding point in the diagram is close to the connec-
tion line between the dry wood point and the water point. The cold gas
efficiency of this process can be up to 95 %.
For the fluidized bed steam gasification process, the feed humidity
should be kept below 10 % in order to reach satisfying conversion effi-
ciencies in the range of 70 %, [33]. Using steam as gasification agent,
the corresponding producer gas point lies directly on the connection line
between the dry wood point and the water point of the diagram.
Figure 3-2 shows that all raw fuel gases are in the graphite formation
free region at 1000 °C. However at 300 °C, which is considered the low-
est temperature at which carbon formation can occur at a considerable
rate, only the updraft and the fluidized bed producer gases are close to
the graphite formation free zone. The required water addition to prevent
carbon deposition, the low content of hydrocarbons and the conversion
efficiency of 80 % were clear factors against the application of the down-
draft gasification in the PSI B-IGFC system.
The tar species found in producer gases from updraft wood gasifiers are
mainly oxygenates, [29], which decompose in the gas phase at tempera-
tures of 500 °C or higher without forming significant amounts of coke,
[55]. In contrast, the tar load of producer gases originating from fluidized
bed steam gasification typically consists to a major extent of aromatic
species which are prone to coking as discussed above. For this reason
and due to the higher conversion efficiency, the updraft gasification
process was identified as best choice for the application in the PSI B-
IGFC system.
Evaluation of gas processing options
In a next step, the required gas processing was defined based on the
fuel gas requirements imposed by SOFCs and the producer gas proper-
ties of the chosen updraft fixed bed gasification process. Fuel gases for
SOFCs have to be particle free in order to avoid physical blocking of the
gas channels. Another issue are impurities such as e.g. alkali salts. Due
to migration processes of these impurities into the electrode materials,
they can cause serious damage to fuel cells via micro structural
changes of the electrode structure, [13]. The particle load of producer
74
3 Experiments
gases from updraft gasification is usually low and at the typical producer
gas temperatures of updraft gasifiers around 75 to 200 °C, alkali salts
can be assumed to be condensed on particles to a major extent. There-
fore it was concluded, that the particle removal via a cyclone and a par-
ticle filter should resolve both issues.
State-of-the-art gas processing is focused on the application of internal
combustion engines which require fuel gas temperatures below the dew
point of tars. In order to prevent blocking of valves etc. tars are almost
completely removed. SOFCs, in contrast to internal combustion engines,
require high fuel gas temperatures and can, to some extent, use tars as
fuel. The reforming of oxygenated tars, which are mainly present in the
producer gas from the updraft gasification process, was demonstrated in
e.g. [55]. Applying cold gas processing for the complete removal of tars
in a B-IGFC system therefore entails a loss of both fuel and sensible
heat, which ultimately results in lower system efficiencies, [132]. Conse-
quently, hot gas processing at temperatures above the dew point of tars
was considered for the PSI B-IGFC system. This consideration took
place despite the handling of tars being more delicate than their removal,
as especially the heat-up of tar-laden gases can cause serious coking
problems, [133].
Cell internal reforming of hydrocarbons and tars is a possibility for the
chemical cooling of SOFCs, which reduces the cooling air mass flow re-
quired to maintain the operational temperature. Further, internal reform-
ing increases the electrical efficiency of SOFCs as the heat of the elec-
trochemical reactions is used cell internally to increase the heating value
of the fuel gas. However, hydrocarbon and tar reforming reactions are
relatively fast in the presence of nickel at the typical operating tempera-
tures of SOFCs. High hydrocarbon and tar contents can therefore lead
to thermal stress due to cooling of the cell inlet region as a consequence
of the endothermic character of reforming reactions. It was therefore
concluded that the gas processing must allow for a precise partial con-
version of hydrocarbons and tars to yield the best compromise between
thermal stress problems and efficiency benefits.
A further issue which has to be addressed by the gas processing is sul-
fur. Sulfur is known for its tendency to form strong bonds with nickel,
[134-136], and thereby blocking active sites making these unavailable
for reactant molecules. For the poisoning, the type of the sulfur species
is largely irrelevant, [116, 118]. However for the removal of sulfur, the
75
3 Experiments
Figure 3-3:
Outline of used
gas sampling sys-
tem
pane, nitrogen and oxygen. All other species have to be removed from
the analyzed gas prior to the injection to the μ-GC. The employed gas
sampling system is based on the design proposed in [137] and has been
further developed at PSI, Figure 3-3.
In a first step, the raw gas is quenched with a solvent. Condensable
species, such as tars, water and other impurities, are thus dissolved.
The solution is cooled to -25 °C to assure that only traces of condens-
able species remain gaseous. After a gas-liquid separation, the dry and
clean gas is pumped to the μ-GC while the used solvent is either reused
for another quench cycle, as depicted in Figure 3-3, or bottled for later
analysis. 1-methoxy-2-propanol is used as solvent due to its good solu-
bility for both organic carboneous species and water.
At times the employed solvent was analyzed in order to determine the
amount and nature of impurities in the raw gas and its water load. The
employed methods were:
Gas chromatography coupled with mass spectrometry (GC/MS) for
the identification of aromatic species such as benzene, toluene etc.
and quantification of their concentration in the raw gas.
Gravimetric method for the determination of the concentration of hy-
drocarbons with dew point above approx. 140 °C in the raw gas.
Following the gravimetric method, a given amount of used solvent is
placed in a vacuum oven at 40 °C and 30 mbara until the weight loss
of the specimen due to evaporation of volatile compounds subsides.
Based on the remaining mass of the sample, which is attributed to the
mass of higher hydrocarbons, the corresponding concentration in the
raw gas is computed.
Karl-Fischer titration method for the analysis of the water concentra-
tion in the raw gas.
Given that offline analysis is time-consuming, mass balance calculations
were performed to estimate the tar and water concentrations of the raw
gas based on the measured dry and clean gas compositions. The gen-
eral assumptions were:
Carbon (C), hydrogen (H), oxygen (O) and nitrogen (N) are the only
considered elements.
There are no uncontrolled leakages in the balanced system.
Nitrogen appears as molecular nitrogen only.
77
3 Experiments
78
3 Experiments
Figure 3-4:
Technical outline
and photograph of
the experimental
setup used for
proof-of-feasibility
tests in 2004
The gasifier had a maximum feed of 2.7 kg/h of wood pellets. For the
experiments, the mass flow was reduced to 1.3 kg/h or 6 kW of thermal
input, respectively. The non-stop operation period was limited to 16
hours. Typical gasifier air-to-fuel ratios were between 0.2 and 0.4. A pla-
nar lab scale stack provided by the Hexis AG (formerly Sulzer Hexis)
was used for the experiments. The SOFC stack consisted of five single
cells with an active area per cell of 60 cm2. The stack was designed for
a nominal electric power output of 60 W when operated with reformate
gas originating from the catalytic partial oxidation of natural gas from the
Swiss grid. The stack was placed in an oven which kept the operational
temperature of the stack at approx. 950 °C.
Due to the stack and system design aiming at natural gas applications,
means of heating had to be introduced to the normally non-heated feed
gas inlet area. This area as well as the pipe system connecting the gasi-
fier and the SOFC system were heated to 400 °C in order to prevent
condensation of tars and thereby blocking. Further, a particle filter was
79
3 Experiments
900 18
5-cell SOFC stack
with wood gas under 800 16
isothermal conditions
700 14
in an oven
Cell voltages [mV]
100 2
0 0
0 6 12 18 24 30 36 42 48 54 60 66 72 78
Runtime [h]
80
3 Experiments
short periods of time, the pressure loss could be reset to normal levels.
The voltage-current-curve remained almost constant during operation
with real producer gas as well as with synthetic producer gas. The ob-
served minor cell deactivation after the experiments was on normal lev-
els compared to that of natural gas operation of the same duration. The
tar species typically present in updraft gasification producer gases, such
as e.g. phenol and acetic acid, were not found to have a negative impact
on the performance of SOFCs. Unlike most relevant literature, this result
once more supported the hypothesis that SOFCs are, to some degree,
tolerant towards tars.
Humidified hydrogen with a tar load of approx. 240 g/mn3 (dtf) was fed
via a heated transfer pipe to a SOFC lab scale stack. The stack was not
electrically connected to a power unit. This was done in order to reduce
the costs of the experiment in case the stack might get destroyed due to
the high tar load. The stack was mounted in a tinder resistant INCONEL
housing placed in an oven heated to 900 °C. The hydrogen mass flow
was controlled via a mass flow controller (MFC). Further, a nitrogen line
was installed for inertization purposes in emergency cases and for the
conservation of deposits in the SOFC for post-test analysis. The hydro-
gen volume flow was monitored manually using a gas meter prior to en-
tering the water vaporizer. The added steam mass flow was set via the
flow rate of the water pump and monitored manually by means of a bal-
ance. A similar approach was used to add a mixture of tar species rep-
resenting the typical composition of tars found in producer gases from
fluidized bed gasifiers, [138]. This mixture was chosen as it was consid-
81
3 Experiments
ered more difficult to degrade than the mixture of oxygenated tars typi-
cally found in updraft producer gases. The mixture consisted of benzene
(63.0 mass-%), toluene (9.5 mass-%) and naphthalene (27.5 mass-%).
The composition of the gas after passing the SOFC was measured
through gas chromatography using the above described sampling sys-
tem. The gas volume was monitored with a gas meter.
After the experiments, mass balance calculations were conducted yield-
ing tar conversions between 92 and almost 99 % for different steam-to-
carbon ratios, see Table 3-1. In comparison, the tar conversion found for
the empty INCONEL housing was in the order of 10 %. Therefore, the
observed tar conversions were attributed to a major extent to the cata-
lytic activity of the investigated nickel anode catalyst.
2 0.88 12.41 3.23 0.81 7.56 1.04 82.22 12.67 4.97 0.14 93.1
3 0.97 9.06 1.59 0.96 7.78 1.06 80.13 17.44 2.29 0.14 98.9
4 0.95 7.64 0.86 0.88 7.25 0.86 80.14 18.91 0.19 0.76 92.5
Figure 3-7 compares the measured gas composition at the stack outlet
and the equilibrium composition calculated with ASPEN PLUS.
H2 3
H2 [mol-%]
H2O (balanced)
70 20
CO
55 10
CH4 CO2
40 0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
Steam-to-carbon ratio [-]
The equilibrium calculations were conducted for the set oven tempera-
ture and the input gas composition considering H2, CO, CO2, CH4, C2H2,
82
3 Experiments
C2H4, C2H6, C3H6, C6H6, C7H8, C10H8, O2, water and graphite as possible
products. The full lines in Figure 3-7 represent the equilibrium molar
fractions of the most important species in the gas phase for varying
steam-to-carbon ratios. The dotted lines stand for the equilibrium molar
fractions considering solid graphite, which is formed for steam-to-carbon
ratios below one at 900 °C. It can be seen that the tars in the input gas
were degraded to hydrogen and carbon monoxide and that the refor-
mate gas almost reached equilibrium. The assumption for SOFC models
that the water gas shift, [139-141], and the steam reforming reaction,
[142-144], are at equilibrium was therewith confirmed.
In order to provoke carbon depositions the steam-to-carbon ratio of the
last experimental point was chosen below one. In fact, the deposits
caused an increasing pressure loss. For a closer analysis, the deposits
were conserved by flushing the stack with nitrogen during cool down.
This allowed not only the verification of the formation of carbon deposi-
tions but also their location was determined. Figure 3-8 shows the top
view (A) of an interconnector plate with carbon deposition at the gas
inlet. View (B) shows that the carbon depositions grew from the inter-
connector plate in the central gas channel, see view (C). This indicates
that carbon depositions are self-catalyzing. An explanation for the loca-
tion of the carbon depositions at the cell inlet is the local temperature,
which must have been considerably lower than the average cell tem-
perature due to endothermal steam reforming reactions.
Figure 3-8:
Carbon depositions
in anode channel
inlet region and
central fuel gas
channel
83
3 Experiments
In summary, the experiments have shown that the anode catalyst of the
investigated SOFC catalyses the degradation of aromatic tars to hydro-
gen and carbon monoxide. Oxygenated tars should be degraded even
more easily. Water gas shift and steam reforming reactions were found
to be very close to equilibrium while the formation of carbon deposits
was found to be locally limited to the cell inlet and kinetically controlled.
84
3 Experiments
Pressure [mbar]
500 125
sures Temperature in the
400 middle of the 100
gasification reactor
300 75
Gasifier
200 50
pressure
100 25
0 0
0.00 0.25 0.50 0.75 1.00
Runtime [d]
The gasifier produced about 2.2 mn3 of producer gas on a dry and tar
free basis per hour at an air-to-fuel ratio in the range of 0.25 to 0.27. A
85
3 Experiments
mass balance was applied to the gasifier, taking into consideration the
dry gas composition and measured water content. It was found that the
sum of the tar species maintained a C-H-O ratio corresponding to the
stoiciometric formula CH1.2O0.5. This implied abundance of oxygenated
tar species in the producer gas as also reported in [28, 29] for a com-
mercial updraft gasifier. The tar load was calculated to be as high as
160 g/mn3 (dtf). This is an elevated value even for an updraft gasifier,
where most of the volatile tars escape during the pyrolysis occurring at
the upper part of the reactor. Apart from being detrimental to a SOFC
stack at such levels, the high tar load also rendered stable long-term
operations of the gasifier. This called for further improvements.
Thermodynamic calculations indicated that with the injection of steam
into the gasifier, keeping the reaction zone temperature above 650 °C, a
considerable fraction of the tars could be reformed or formation could be
inhibited. Consequently, a steam generator along with a super heater
coil was integrated with the gasifier, capable of injecting up to 800 g/h of
steam at 600 °C into the alumina ball bed.
Figure 3-10 shows 167 h of operation of the gasifier with constant steam
injection at 600 g/h, which corresponds to 460 g per kg of dry wood pel-
lets or a theoretical humidity of the wood pellets of approx. 30 %.
pressure build-up
Temperature [°C]
Pressure [mbar]
500 125
and decline cycles Temperature in the
middle of the
400 gasification reactor 100
300 75
Gasifier
200 pressure 50
100 25
0 0
0 1 2 3 4 5 6 7
Runtime [d]
The temperature in the middle of the gasification reactor and the pro-
ducer gas temperature did not fluctuate and the gasifier pressure
showed no trend for blocking of the gasifier or the downstream piping.
However, the gasifier pressure did not remain at a constant level. It in-
creased up to a maximum of 60 mbar over-pressure, before the deposits
86
3 Experiments
causing this over pressure were removed by the producer gas flow. Be-
sides allowing for a stable operation of the gasifier, the steam injection
had the expected positive impact on the producer gas composition and
tar load. Figure 3-11 shows the improvement of gas composition, start-
ing from no steam injection mode, as the steam rate was gradually in-
creased till 640 g/h. This corresponds to a theoretical wood pellet hu-
midity of approx. 33 %. The air-to-fuel ratio was fixed at 0.26.
Figure 3-11: 20 72
Effect of steam
injection on the
raw producer gas
H2, CO, CO2, CH4, C2H4 [mol-%]
15 54
composition of the
PSI updraft gasi-
fier
N2 [mol-%]
10 36
H2 CH4
CO CO2
C2H4 N2
5 18
0 g/h 150 g/h 400 g/h Water 640 g/h
0 0
0:00 1:00 2:00 3:00 4:00 5:00 6:00
Time [h]
It can be seen that the molar fraction of hydrogen almost doubled with
increasing steam injection from approx. 5 to 9 %. The carbon dioxide
molar fraction increased from approx. 8 to 12 %, while the carbon mon-
oxide molar fraction decreased from 15 to 13 %. This indicates that wa-
ter gas shift reactions were enhanced by the injected excess steam. The
nitrogen fraction decreased from approx. 63 to 56 % with increasing
steam injection. This corresponds to a gain of dry and tar free gas vol-
ume of roughly 13 %.
Mass balance calculations were performed to estimate the tar load of
the producer gas with steam injection, Figure 3-12. It was found that the
steam injection allowed an effective conversion of the tars into prefer-
able permanent gas species, decreasing the tar load by about 50 %.
This also explained the increased producer gas volume. The remaining
tars are not expected to pose problems to a SOFC anode at an operat-
ing temperature around 900 °C due to their oxygenated, thus unstable
nature.
In summary it can be said, that the gasifier built for the demonstration
unit fulfills all the defined requirements, allowing for long-term tests.
87
3 Experiments
100 112
82
50
0
No steam injection 460 g/h steam injection
Figure 3-13:
Updraft wood
gasifier com-
bined with a hot
gas filtration
system and an
oil scrubber Flare Silo Wet scrubber
SP 2
Orifice
Gas pumps
SP 1
Filter housing
Wood gasifier
The time progression of the pressure loss over the investigated HGF
system is a good indicator for its operational condition.
Figure 3-14:
Δpmax
Typical pressure
history of filtra-
Pressure drop Δp
tion systems
with cleaning
cycles, [41, 146] Deep bed filtration
ΔpR
Surface filtration
Filtration
interval
Δp0
Runtime t
Figure 3-14 shows a typical pressure loss time progression of candle fil-
ter element based HGF systems. New filter elements pass through a
commissioning phase prior to reaching stable operating conditions. At
first, the pressure loss over the filter housing Δp0 originates only from
the filter element. Particles deposit in the filter element and induce an
increase of the pressure loss over the filter element. The commissioning
89
3 Experiments
phase is completed after the saturation level of particles in the filter ele-
ment is reached. The particles then deposit only on the surface of the fil-
ter element accumulating to a dense dust cake with lower porosity com-
pared to the fresh filter element. The dust cake is allowed to build up un-
til the pressure loss reaches the design level Δpmax. This initiates a re-
generative cycle, where the dust cake and parts of the particles depos-
ited in the filter element are removed by a reversed gas flow through the
filter element. Due to the flammability of producer gases, pressurized
argon was employed for the regeneration cycles. The filtration interval is
defined as the time period between two consecutive regeneration cycles.
The pressure loss after a regenerative cycle, ΔpR, draws near a given
value for stable operation. If it steadily increases, particles penetrate
deep into the filter element ultimately leading to operational failure.
Two types of filter elements were characterized, namely a non-ceramic
fiber filter candle with a pore diameter less than 1 μm and a metallic fil-
ter candle with a pore diameter less than 5 μm. The gasifier was oper-
ated with a wood pellet mass flow of 1.3 kg/h and an air-to-fuel ratio of
0.36 which translates into a gasification air volume flow of 33 ln/min. The
filtration velocity is defined as the quotient of filtered producer gas vol-
ume flow and filtration area of the investigated filter element. For the
demonstration of a stable operation, the filtration velocity was main-
tained at 1.6 cm/sec. The temperature of the HGF unit was held con-
stant at 400 °C.
27 h operation of
hot and highly tar
laden gas filtration
Pressure drop over HGF unti [mbar]
160 1.6
employing ce-
Filtration velocity [cm/sec]
80 0.8
40 0.4
Gasifier failure
0 0
8:00 10:00 12:00 14:00 16:00 18:00 20:00 22:00 0:00 2:00 4:00 6:00 8:00 10:00
Time [hh:mm]
eration were achieved with this filter element type. This corresponds to a
total filtered gas volume of approx. 57 mN3, thereof approx 27 mN3 non-
stop. The grey area indicates a gasifier failure which made a burnout
cycle necessary. The HGF unit was operated throughout this burnout
cycle, showing no signs of clogging due to the strongly increased par-
ticulate matter load of the gasifier flue gas. The pressure after cleaning
pulses, ΔpR, leveled off at 80 mbar with Δpmax set to 140 mbar. The re-
generation cycles occurred at regular 60 minutes intervals. No particles
were found in the wet scrubber. It was concluded, that stable operation
was reached with the investigated non-ceramic fiber filter element.
In contrast, the metallic filter candles were clogged almost instantly
when operated at 400 °C. The temperature of the HGF unit was there-
fore increased to 500 °C, which allowed a continuous operation. How-
ever, large amounts of soot were found in the wet scrubber fluid. The
soot was either produced in the metallic filter candle or consisted of char
particles originating from the gasifier which passed the filter element.
For both cases it was concluded, that metallic filter candles are highly
problematic for the filtration of tar laden producer gases.
The impact of the HGF system was assessed via the measurement of
the producer gas composition up- and downstream of the filtration unit.
The tar and water load was estimated with mass balance calculations,
see section 3.2. It was found that the HGF system employing the non-
ceramic filter elements influenced the producer gas composition. Due to
the inert character of the filter element material, it is most probable that
the observed activity is due to the soot cake, which is continuously de-
posited on the candles. The molar fractions of H2, CO, H2O and CH4 in-
creased notably passing the HGF system, while the molar fractions of
C2H2, CO2 and tars decreased. This indicates that tar and ethylene de-
composition processes and reverse water gas shift reactions took place
in the HGF system, see Figure 3-16. It has to be pointed out that these
values were calculated based on the measured species concentration
and the balanced tar and water load. As it cannot be excluded, that the
principal tar signature changed, the values should be understood as
rough trend indicators. More detailed studies are required in order to de-
termine the exact reaction scheme. The variation of the filtration velocity
had no measurable impact on the discussed trends.
The metallic filter elements were two to three times more active than the
non-ceramic fiber filter elements towards the decomposition of tars and
91
3 Experiments
Conversion of 1.00
measured gas
CO
species and 0.75
0.25
HGF system CH4 C2H4
H2O
employing a 0.00
C2H2 H2
non-ceramic -0.25
Tars
-1.00
-1.25
92
3 Experiments
Figure 3-17:
Representative tar O
species for CPO
characterization
experiments
S
Toluene Anisole Thiophene
Figure 3-18: Ar Ar
Experimental MFC-Station
setup for catalyst Ar-
H2S
characterization
experiments P-11
CO2
CO
N2
Tar H2O
Saturator Saturator
H2
CH4
Preheater
MS Superheater
93
3 Experiments
The setup was designed to allow for catalyst tests with tar-laden and
humidified synthetic gas mixtures. Six mass flow controllers enabled the
generation of a wide range of dry gas mixtures consisting of H2, CO,
CO2, CH4 and N2. Hydrogen sulfide diluted in argon could be added as
well. The dry gas mixture was then preheated to 180 °C and humidified
with deionized water via a water saturator with argon as carrier gas. The
water load was controlled via the saturator temperature which was set to
70 °C for the experiments. The addition of liquid tars followed the same
scheme. The saturator temperatures were 30 °C for thiophene, 60 °C
for toluene and 90 °C for anisole. The tar-laden and humidified gas mix-
ture was then superheated and piped into a packed-bed reactor housing,
which was placed in an electric furnace. The gas samples for the
Thermo VG ProLab quadrupole mass spectrometer (MS) were sucked
through fused silica capillaries which were placed before and after the
catalyst. A condenser located at the reactor outlet was used for the re-
covery of non converted water or organic compounds.
As mentioned in section 3.1.2, CPO catalysts feature two reaction zones,
the first being the oxidation and the second being the steam reforming
zone. It was assumed, that while oxidation occurs fast, producing water,
carbon dioxide and heat, the steam reforming is a slow process. There-
fore, a diluted "post-oxidation" gas composition was defined as matrix
for the tests, consisting of 9 mol-% H2, 8 mol-% CO, 6 mol-% CO2, 15
mol-% H2O and 62 mol-% N2. The temperature range chosen for the ex-
periments was 700 to 800 °C. The weight hourly space velocity (WHSV)
was varied between 2.75 and 7 gtar/(gcatalyst h).
For WHSVs between 2.75 and 5.5 gtoluene/(gcath), toluene conversions
between 38 and 59 % were observed using 119 mg of catalyst at a tem-
perature of 700 °C, see Figure 3-19. At 800 °C the conversion was con-
siderably higher with values between 72 and 90 %. While at 800 °C
most of the toluene was reformed to hydrogen and carbon monoxide, Eq.
1, at 700 °C an important part of the toluene was degraded to benzene,
Eq. 33.
C7 H 8 + 7 H 2O → 7CO + 11H 2 Eq. 32
94
3 Experiments
Conversion [%]
60
and WHSVs,
[148]
40
20
0
2 3 4 5 6 7 8
WHSV [gtoluene/(gcat h)]
700°C 800°C
Figure 3-20 depicts the conversion of anisole over the investigated CPO
catalyst. The observed conversions of anisole were between 61 and
83 % at 700 °C and between 78 and 87 % at 800 °C. In contrast to the
conversion of toluene, no formation of benzene was observed during the
reforming of anisole. The high conversions and the lack of benzene as
side product support the assumption that oxygenated tars are easier to
degrade than pure aromatic species.
60
and WHSVs,
[148]
40
20
0
2 3 4 5 6 7 8
WHSV [ganisole/ (gcath)]
700°C 800°C
95
3 Experiments
Conversion [%]
tures and WHSV, 60
[148]
40
20
0
2 3 4 5 6 7 8
WHSV [gthiophene/(gcath)]
700°C 800°C
96
3 Experiments
For hydrogen sulfide, the sulfur separation efficiency was close to 100 %,
yielding sulfur concentrations at the adsorbent bed outlet below 1 ppmV.
These values were also reached despite an increase of the water frac-
tion in the producer gas matrix from 5 to 13 mol-%. Tests with thiophene
confirmed, that organic sulfur species are not adsorbed by zinc oxide.
With organic sulfur concentrations up to several ppmVs, this confirms
that a conversion unit for organic sulfur species such as a CPO is most
required in B-IGFC systems. Otherwise sulfur concentrations around 1
ppmV cannot be reached.
The thermal stability of the investigated zinc oxide adsorbent was as-
sessed with thermo-gravimetric measurements. The maximum operation
temperature was found around 450 °C, see Figure 3-22. Starting from
550 °C, zinc oxide is reduced to pure zinc and quickly begins evaporat-
ing. The measured curve compares well with the vapor pressure curve
of pure zinc calculated with ASPEN PLUS.
25 0.25
0 0.00
300 400 500 600 700 800 900
Temperature [°C]
The composition of the producer gas matrix was not altered in the pres-
ence of zinc oxide. In summary, the experiments confirmed that the find-
ings reported in [107] with respect to zinc oxide for the desulphurization
of syngases originating from natural gas or coal gasification can be car-
ried forward to biomass gasification applications.
97
3 Experiments
Figure 3-23:
The Hexis SOFC Air Fuel gas Exhaust gas
system, [150]
Catalytic partial Post-combustion Exhaust channels
oxidation monolith
Insulation
SOFC stack
Air preheater
Figure 3-23 shows the Hexis SOFC system. The Hexis cell design is
planar circular, electrolyte-supported and optimized for co-flow operation.
The stack consists of 63 metallic current collectors and electrode-
98
3 Experiments
electrolyte assemblies. The active area per cell is roughly 100 cm2. The
maximum electrical output of the stack is 1 kW.
Air is introduced to the SOFC stack via an air pre-heater. The pre-
heated air enters the SOFC stack on the cathode side through four air
channels which are formed by the metallic interconnector plates and the
anode-electrolyte-cathode assemblies. After participating in electro-
chemical reactions, the air exits the stack radial and takes part in the
post-combustion of the partially depleted fuel gas. The fuel gas entering
the stack is reformed and heated to a temperature suitable for the SOFC
stack using a catalytic partial oxidation monolith. The processed fuel gas
is introduced to the SOFC stack on the anode side via a central fuel gas
channel. After being partially electrochemically converted to water and
carbon dioxide, the depleted fuel is mixed with the depleted air and
combusted. The exhausts are used for the pre-heating of the fresh air.
Pure methane was used as model fuel for the experiments. A thiophene
saturator, see sections 3.4.3 and 3.4.4, was installed in an ice bed. Ar-
gon was used as carrier gas. The thiophene concentration in the fuel
gas was adjusted via the carrier gas flow. The employed mass flow con-
troller allowed producing thiophene concentrations in the fuel gas up to
400 ppmV. This translates into approx. 80 ppmV at the cell inlet, due to
the volume expansion of the fuel gas resulting from the CPO. To ac-
count for the adsorption of sulfur on the metallic surfaces, the voltage-
current curves were only recorded after several hours of steady-state
operation. Assuming that in steady-state the ad- and desorption of sulfur
on the metallic surfaces of the feed line occur at equal rates, the sulfur
concentration at the cell inlet should correspond to the set value.
Figure 3-24: 65
Voltage-current
curves obtained
from the investi- 60
gated Hexis-stack
for different sulfur
Stack voltage [V]
55
concentrations
ppmV of thiophene in feed gas to
catalytic partial oxidation catalyst
50
45
400 200 90 50 25 12 6 0
40
0 5 10 15 20 25
Stack current [A]
99
3 Experiments
940
920
900
0 50 100 150 200 250 300 350 400
Sulfur concentration in feed gas [ppmV]
100
3 Experiments
and 40 % reform-
ing activity
1.025
Nernst @ 100% kinetics
Nernst @ 40% kinetics
OCV @ 100% kinetics
OCV @ 40% kinetics
1.000
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell length [m]
Figure 3-27 correlates the OCV drop with the reforming activity.
Figure 3-27: 30
Calculated OCV
drop for varying
25
reforming activi-
ties
20
OCV voltage loss [mV]
15
10
0
0 10 20 30 40 50 60 70 80 90 100
Anode reforming activity [% of activity without sulfur poisoning]
Assuming that the first 30 mV of OCV drop in Figure 3-25 result from re-
duced STR activity leads to approx. 90 % deactivation according to
Figure 3-27. This magnitude of deactivation is reasonable for high-
101
3 Experiments
Figure 3-28:
65
Comparison of
voltage- current
curves obtained 60
before and after
sulfur experiments
Stack voltage [V]
55
Before thiophen experiments
45
6% degradation
40
0 5 10 15 20 25
Stack current [A]
102
3 Experiments
ter the sulfur experiments. Given that such degradation values are nor-
mal after Redox and thermo cycle test programs, [84], the role of sulfur
as a limiting factor to the proof-of-concept test can be put in perspective.
In summary, the desulphurization is the key to substantially minimize
sulfur poisoning induced power losses. For the proof-of-concept tests, a
desulphurization was found desirable but not necessary. Indeed sulfur
reduces the power output of the Hexis SOFC but does not inhibit its sta-
ble operation. In addition, omitting the desulphurization neglects the
need for a high temperature heat exchanger, simplifying the experimen-
tal setup for the proof-of-concept test considerably.
103
3 Experiments
104
3 Experiments
105
3 Experiments
Figure 3-29:
Technical outline
and photograph of
the present dem-
onstration unit
configuration
106
3 Experiments
In section (A), a scheme of the PSI wood gasifier is shown. The techni-
cal details were discussed in section 3.4.1. The gasification air is con-
trolled by means of a mass flow controller (MFC). The temperature and
composition of the raw producer gas as well as the pressure loss over
the gasifier are measured approx. 10 cm downstream the gasifier outlet.
Section (B) shows the current configuration of the gas processing setup.
Particles larger than approx. 5 μm are removed in a cyclone and dis-
charged with a screw conveyor. To prevent tar condensation during the
start-up phase of the gasifier, the cyclone and all producer gas flowed
through pipes are electrically heated to 400 °C. The tar load of the pro-
ducer gas is reduced in the first external catalytic partial oxidation unit
(CPO1). The minimum ignition temperature of the CPO catalyst is
approx. 450 °C. CPO1 can hence be electrically heated to over 600 °C.
The air volume fed to CPO1 is set using a MFC. The pressure upstream
and the temperature in the middle of CPO1 are monitored. The compo-
sition of the catalytic partially oxidized producer gas is measured directly
at the reactor outlet. The second catalytic partial oxidation unit (CPO2)
is placed downstream valve V2. The methane and air volumes are both
controlled by means of MFCs. In section (C), the SOFC system is out-
lined. The temperature and pressure of either the catalytic partially oxi-
dized methane or the producer gas are measured at the SOFC system
inlet. The cooling air flow for the system is adjusted using a MFC. The
actual condition of the SOFC stack is monitored via thermocouples in
the top, middle and bottom cells. The electrochemical performance is
determined by measuring the voltages of 5 cell clusters, Cl1 to 5. The
net stack voltage is maintained constant using an electrical load genera-
tor which automatically adjusts the stack current.
107
3 Experiments
108
3 Experiments
Figure 3-30:
Time plots of gasi- 1000 1000
Temperatures
fier outlet, CPO1
900 900
mid and SOFC sys- Gasifier out CPO mid SOFC in
tem inlet tempera- 800 800
tures as well as
[°C]
[°C]
measured raw and 700 700
CPO gas composi-
tions plus derived 600 600
volume flows, heat-
ing values, steam- 500 500
to-carbon ratios, tar
and water loads 25 70
Raw gas composition
20 63
H2
[N2 mol.-%]
[mol.-%]
15 CO 56
CO2
10 CH4 49
5
C2 42
N2
0 35
25 70
CPO gas composition
20 63
H2
[N2 mol.-%]
[mol.-%]
15 CO 56
CO2
10 CH4 49
5 C2 42
N2
0 35
10 400
Raw gas properties
[mn3/h] & [-] & [kW]
8 320
Vol. [g/mn3 dtf]
6 LHV 240
4 SC 160
Tar
2 Water 80
0 0
10 400
CPO gas properties
[mn3/h] & [-] & [kW]
8 320
Vol.
[g/mn3 dtf]
6 LHV 240
4 SC 160
Tar
2 Water 80
0 0
0 6 12 18 24 30
Run time [h]
109
3 Experiments
Figure 3-30 shows the raw and CPO gas composition measured during
the first 9 h and the last 7 h of the experiment. Note that the discontinu-
ity in the CPO gas progressions in the 5th hour of the experiment is a re-
sult of a change of the sampling point from the middle to the outlet of the
CPO1 reactor. The characteristic properties of the gases, the tar and
water load, volume flow, LHV power and steam-to-carbon ratio (SC),
were calculated based on the measured dry gas compositions. The SC
was defined as the ratio of the total carbon content of the gases to their
steam content. It can be seen, that the gasifier and the CPO1 operated
steady state. Especially the smooth gas volume and steam-to-carbon ra-
tio progressions are of interest. Table 3-3 summarizes the average and
the standard deviation values of the gas compositions and their charac-
teristic properties.
From Table 3-3 it can be seen that the CPO1 predominantly promoted
the water gas shift reaction. In terms of hydrocarbons, it converted al-
most all C2 species and on average about 40 % of the methane in the
raw gas. The tar conversion was on average around 15 %. The low tar
conversion might be a consequence of diffusion limitations in the wide
channels of the employed CPO monolith with 100 channels per square
inch. A finer meshed monolith could not be used due to blocking issues
caused by the moderate particle removal efficiency of the cyclone. The
LHV loss resulting from the partial combustion of the raw producer gas
was approx. 10 %.
110
3 Experiments
Figure 3-31:
Time plots of stack 5 75
current and voltage, Stack current and voltage Current Voltage
4 60
cluster specific av-
erage single cell 3 45
voltages, stack tem-
[V]
[A]
peratures, pressure 2 30
differences and air
volume flows of the 1 15
main components of
the demonstration 0 0
unit
1.0 1.0
Cluster specific cell voltages
0.8 0.8
0.6 0.6
[V]
[V]
0.4 0.4
0.0 0.0
1000 1000
Stack temperatures
950 950
900 900
[°C]
[°C]
850 850
800 800
Top Mid Bottom
750 750
25 75
Pressure differences Gasifier CPO SOFC
20 60
between in - and outlet
15 45
[mbar]
[mbar]
10 30
5 15
0 0
2000 500
Air volume flows Gasifier CPO SOFC
1600 400
1200 300
[ln/min]
[ln/h]
800 200
400 100
0 0
0 6 12 18 24 30
Run time [h]
111
3 Experiments
112
3 Experiments
Figure 3-32:
C-H-O ternary dia-
gram for the ther-
modynamic graphite C
0.0
Lines
formation with in- 1.0 950°C: SOFC design temperature
scribed C-H-O mo- 800°C: SOFC minimum temperature
lar fractions of dry 650°C: Gasifier outlet temperature
550°C: SOFC inlet temperature
and wet wood and 0.2
0.8 300°C: Min. temperature for
producer gas after noticeable graphite formation
the different process Graphite
steps 0.4 formation
0.6
region Points
CO
0.6 1: Dry wood
0.4 2: Wood humidity
3: Steam addition
1 CO2
CH4 2
4: Gasification air
5: CPO air
0.8
3 0.2
4 5
1.0
0.0
0.0 0.2 0.4 0.6 0.8 1.0
H H2O O
The first SAI at around 9.5 h experimental run time resulted in an instant
stack top temperature increase and system pressure decrease indicat-
ing a successful discharge of deposits from the SOFC system. The fol-
lowing SAI at 12.5 h and at 27 h experimental run time had no impact on
the stack top temperature. The system pressure was lowered during the
SAIs due to an automatic decrease of the cooling air flow resulting from
the cooling effect of the increased gas flow at the anode. However after
the SAIs, the system pressure went back to the initial value as the cool-
ing air flow returned to normal. Post-test examination of the central fuel
gas channel showed that it was clear throughout the experiment. The
assumption of coke deposition was hence shown wrong and the opera-
tional conditions were confirmed as safe in this respect. In contrast, evi-
dence was found for ash deposits in cold regions of the system.
Figure 3-33:
Ash deposits on the
mounting of the
thermocouple at the
SOFC system inlet
113
3 Experiments
114
115
116
4 Modeling work literature review
117
4 Modeling work literature review
118
4 Modeling work literature review
119
4 Modeling work literature review
possesses some flexibility towards cell design details and fuel gas
compositions. Calculation times are still in the range of seconds.
The main problem with this type of lumped model is the gas composi-
tion, which is used in the calculations of the operational voltage for a
given current density. In reality, the gas composition changes along the
fuel channel and therewith the Nernst voltage as well as the voltage
losses and the local current density. If the inlet gas composition is used
for the lumped model calculations, the voltage and power output is sys-
tematically overestimated, while it is systematically underestimated if
the outlet gas composition is used, [161]. To yield realistic results,
most lumped models use a theoretic gas composition, which is calcu-
lated as arithmetic average of the inlet and the outlet gas composition.
The oxygen ion flow, which is transferred to the anode from the cath-
ode side, is determined from the predefined fuel utilization. The outlet
gas composition is calculated based on the inlet gas composition and
the oxygen ion mol flow assuming chemical equilibrium at the pre-
defined operational temperature.
This lumped model type is the most accurate of the three lumped
model types. However, it lacks information about the local conditions in
the SOFC. Critical conditions such as excessive temperature gradients
cannot be identified. [161-168] are examples for investigations con-
ducted with this model type.
120
4 Modeling work literature review
lar to the gas flow direction, Figure 2-19. According to [169], the er-
ror due to this assumption is negligible when the feed properties are
equal for all gas channels. In consequence, all other properties
along the gas channels should be equal and so should the local
temperature be. Hence, no perpendicular heat transfer should occur.
1D cell level models feature short calculation times in the order of
minutes and are therefore suitable for system analysis calculations.
Two dimensional models (2D)
Planar cross-flow cells, Figure 2-19, have to be modeled 2D due to
the perpendicular gas flow channels. The sizeable additional effort
is why this type of flow design remains largely neglected in system
analysis calculations. However, such a model was used in [154] for
the investigation of an SOFC/gas turbine (GT) cycle.
For cell level models the following discretization methods can be distin-
guished:
Finite element method (FEM)
The idea of this method is to discretize the investigated object, e.g.
flow channel, but also solid structure, and to allocate transfer func-
tions to all considered physical properties, which can change within
a discrete element. The solution is found, when the difference of the
boundary values of all directly adjacent elements is within the toler-
ance. FEM based fuel cell models are usually employed for detailed
investigations of flow phenomena in the gas ducts and channels.
The calculation times are rather long. As a result this approach is
seldom employed for systems analysis calculations.
Numerical volume averaging method (NVAM)
The NVA method is extensively discussed in [170]. The major dif-
ference to the FE method is, that transfer functions are not only de-
fined for physical properties, but also for properties such as pres-
sure, flow velocity etc. which are changing due to the internal struc-
ture of the discretized element. The corresponding transfer functions
are computed via detailed FEM models of the discretized elements.
The overall solution is found analogously to the FE method. NVAM
models feature shorter calculation times than FEM models.
Finite volume method (FVM)
The FV method is often used for one and two dimensional SOFC
models. The discrete elements are assumed as continuously stirred
121
4 Modeling work literature review
122
4 Modeling work literature review
123
4 Modeling work literature review
culations conducted with stack level models have hence not been re-
ported as yet.
124
4 Modeling work literature review
sity values. The main finding is that by increasing the current, the carbon
deposition risk decreases. This phenomenon is associated with the in-
creasing oxygen partial pressure at the anode with increasing current.
Furthermore, the threshold current density value, at which no graphite
formation occurs, is found to increase with higher tar loads.
Kivisaari, [156], from the Royal Institute of Sweden explores the rela-
tions between a MCFC and the balance of plant (BoP) components in B-
IGFC systems. The employed model is of the lumped type based on
Faraday's law with pre-defined voltage. The main finding is that increas-
ing fuel utilization results in reduced cooling requirements of the MCFC.
Omosum, [132], from the Imperial College/United Kingdom has per-
formed a cost and efficiency analysis of two B-IGFC systems with hot
and cold gas processing system. The employed model is of the same
type as Kivisaari’s. Omosum concludes that the hot gas processing has
advantages on the overall efficiency side due to improved heat integra-
tion possibilities. Furthermore, process waste heat can be sold leading
to increased revenues. These two aspects are assumed sufficient to
cover the higher investment costs of the hot gas processing. Omosum
also states that an increased rate of internal reforming leads to lower
cooling requirements of the SOFC and hence reduced BoP costs.
Van Herle, [160], from the Swiss Federal Institute of Technology in
Lausanne/Switzerland uses a lumped model to study the impact of sew-
age gas on the operational behavior of a SOFC operated at full and par-
tial load. The model includes empirical correlations for activation, ohmic
and diffusion losses. The carbon deposition risk is also considered
through equilibrium calculations for graphite formation. The system
analysis reveals an electrical system efficiency of 48.5 % by employing
a pre-reformer as compared to 43 % by employing a catalytic partial oxi-
dation, which also causes higher cooling requirements of the stack.
Proell, [163], from the Technical University of Vienna/Austria analyses
different options of heat integration, bottoming cycles and system pres-
sures focused on a B-IGFC system based on the FICFB gasifier in
Güssing/ Austria. The applied lumped model includes detailed electro-
chemical calculations. The operational voltage is calculated based on
the average gas composition between the cell in- and output gas com-
position. The output gas composition is calculated assuming chemical
equilibrium. The overall electrical system efficiency is calculated to be
40 %. The largest losses are found in the gasification section.
125
4 Modeling work literature review
127
128
5 Modeling
5 Modeling
Figure 5-1:
- Raw gas properties
Overall modeling - Temperatures of
approach gas processing units
- Required water addition
In a first step, the producer gas composition at the SOFC inlet is deter-
mined using the commercial flowsheeting software ASPEN PLUS (ver-
sion 12.1). ASPEN PLUS was developed for the simulation of chemical
process applications and solves heat and mass balances in complex
systems. The input data for this first calculation involves the raw gas
composition, temperature and yield, the gas processing unit operations
and their operational temperatures as well as the eventually required
additional amount of water or steam to prevent graphite formation.
The predicted gas composition at the SOFC inlet is transferred to a
SOFC model code implemented in the ATHENA equation solver lan-
guage, [185]. The SOFC model yields the gas outlet temperatures, the
required air volume flow to maintain the design temperature of the
SOFC, the fuel volume flow required for a certain operational voltage at
a given fuel utilization and the direct current (DC) power output. The
SOFC model is discussed in section 5.3.
The results of the SOFC model are used as input for a second flow-
sheeting simulation. Figure 5-2 exemplarily shows the ASPEN PLUS
flowsheet of the B-IGFC system 1, see section 6.2.2.
129
5 Modeling
Figure 5-2:
Flowsheet of
"System 1"
tracted from the flue gas. The remaining heat is used for steam genera-
tion, air pre-heating and biomass drying, see section 5.2.3.
The second flowsheeting simulation yields the overall process DC power
efficiency, the amount of use heat at 200 °C for which all other heat re-
quirements are still satisfied, the areas of all heat exchangers in the heat
integration network and other data relevant to component sizing such as
sulfur adsorbent mass, steam reforming reactor heat duties etc.
The results of the second flowsheeting simulation are passed to an EX-
CEL spread sheet where the net AC power efficiency, the equipment
sizes, the overall system costs and the power production costs are de-
termined, see section 5.4.
131
5 Modeling
suming that all sulfur present in the gasified wood are bound in either
hydrogen sulfide or thiophene. The latter was chosen as representa-
tive of organic sulfur species. The ratio between organic and inorganic
sulfur species was assumed similar to that of carbon bound in tars to
carbon bound in permanent gas species. These are carbon monoxide,
carbon dioxide, methane and ethene, respectively.
Figure 5-2 (A) shows that the dry raw gas composition "GASDTF" is first
humidified in the unit operation block MIXER "H2O-LOAD". The sulfur
species are added to the raw gas in the MIXER "S-LOAD". Acetic acid
and anisole as primary tar species are introduced to the raw gas via the
MIXER "PT-LOAD". And finally the raw gas is loaded with toluene as
tertiary tar species in the MIXER "TT-LOAD". The HEATER "HEATGAS"
is used to adjust the raw gas temperature to the measured value.
133
5 Modeling
Figure 5-3:
ASPEN PLUS AIR
model of catalytic
partial oxidation
CONVER EQUIL CONVER2
PRODGAS PG2CONV CONVPG CPOGAS
PGBYPASS
134
5 Modeling
5.2.7 SOFC
The flowsheeting model shown in Figure 5-2 (D) was implemented to
enable the application of the results of the ATHENA SOFC model, see
section 5.3, to the overall system simulations. The flowsheeting model
reproduces the mass and energy balance results of the more intricate
ATHENA model. The fuel gas mol flow required to achieve a predefined
fuel utilization at a given operational voltage is calculated by the
ATHENA model. In the overall system model, this mol flow is used to
compute the required dry raw gas mol flow "GASDTF" and thus the
mass flow of wood chips. The oxygen ion transfer through the electrolyte
is modeled with the separator block SEP "CATHODE". SEP blocks are
used to divert single species streams to a certain output stream while
the rest of the input stream is guided to given output stream. The oxy-
gen ion mol flow is computed based on the fuel gas mole flow and the
fuel utilization predefined for the ATHENA model calculations. The an-
ode is modeled as equilibrium block RGIBBS "ANODE" with a given
heat duty. The heat duty depends on the heat required to reproduce the
cathode outlet temperature and the power output determined by the
ATHENA model. The depleted anode and cathode off gases are post-
combusted in a catalytic burner for heat recovery reasons. This burner is
modeled with the block RSTOIC "AFTERBUR" as adiabatic combustion.
RSTOIC is used when the reaction stoichiometry and degree of conver-
sion is known and the reaction kinetics are either unknown or irrelevant.
Combustion reactions are predefined in RSTOIC.
135
5 Modeling
136
5 Modeling
the oxidant gas channels due to excessive cooling air flows, [191].
Mass transport in normal direction to the flow direction arises via dif-
fusion only.
A constant Nusselt number with the value 4 is assumed for the calcu-
lation of the convective heat transfer coefficients, [190].
Radiative heat transport between solid structure and gas streams is
ignored as the patch length of the radiation is too short for a signifi-
cant amount of absorbed heat by the gas streams, [191].
Heat is transferred within the solid structure via heat conduction.
All heterogeneous reforming reactions are considered so fast, that no
diffusion into the porous anode takes place.
The heats of reaction of all heterogeneous reactions can be attributed
to the solid structure of the fuel cell.
The geometry- and fuel composition-dependent diffusion of the reac-
tants of the reforming reactions through the gas channel to the cata-
lytic surface is considered.
The water-gas-shift reaction is regarded as non-diffusion limited ho-
mogenous reaction which is always at equilibrium.
The electrochemical oxidation of hydrogen and carbon monoxide are
considered to occur in parallel. Direct electrochemical conversion of
hydrocarbons is disregarded. Instead, hydrocarbons are oxidized
through reforming reactions and subsequent electrochemical conver-
sion of the produced hydrogen and carbon monoxide.
Besides local values such as solid temperature and current density, the
developed model calculates integral values which characterize the op-
erational conditions and the performance of the investigated cell design
and fuel composition. The definitions of the integral values are given in
the following:
The average solid temperature represents the operational tempera-
ture of the cell and is calculate as the arithmetic average of all local
temperatures. It is assumed, that the operation of SOFCs is opti-
mized according to the average solid temperature.
The maximum temperature gradient is the maximum value of the
quotient of the temperature difference between adjacent CVs and
their length. The calculation is performed for all CVs in order to find
the maximum value. In reality this value can hardly be measured.
137
5 Modeling
The hydrogen equivalent mol flow, EqH2, is calculated for the inlet
and outlet fuel composition assuming that all hydrocarbons are com-
pletely reformed to hydrogen and carbon monoxide and that all car-
bon monoxide is shifted to hydrogen.
The air utilization is defined as the ratio of the oxygen mol flows at the
cell inlet and outlet. It is equal to the quotient of the fuel utilization and
the air-to-fuel ratio.
The air-to-fuel ratio is defined as the ratio between the effective and
the stoichiometry oxygen mol flow for total combustion.
The average current density is calculated as the quotient of the cur-
rent produced by the cell or the considered part of the cell and the
overall area of the cell or the considered part of the cell. This defini-
tion accounts for the fact, that in some cases not the entire cell area
participates in the electrochemical reactions e.g. areas which are
covered by the interconnector plates, see section 5.3.1.3. The aver-
age current density is computed for the total current and for the cur-
rents originating from hydrogen or carbon monoxide conversion sepa-
rately.
The direct current (DC) power produced is calculated as the product
of total produced current and the according operational voltage.
The electrochemical efficiency is defined as the ratio between pro-
duced DC power and the change of enthalpy of the fuel gas.
The fuel cell efficiency is defined as the ratio between DC power and
LHV input to the cell or considered part of the cell, Eq. 182.
The developed model consists of three coupled sub-models, namely the
electrochemical performance model, the mass balance model and the
energy balance model. The chosen modular structure of the developed
model is highly flexible with respect to the integration of further reactions
and species.
Figure 5-4 depicts the model structure and the corresponding iterative
solution algorithm. The three sub-models are coupled via the fuel com-
138
5 Modeling
position and the temperatures of the gases and the solid structure yield-
ing a highly non-linear equation system. The solution can therefore only
be found iteratively.
Output:
Thermal conductivities
Pure convective heat streams between solid structure Heat capacities
and gases, solid temperature and gas temperatures
of the actual CV
Solution found
139
5 Modeling
In this work, the modeling package "ATHENA Visual Studio 11.0" was
used to implement the numerical code in the FORTRAN language.
ATHENA includes strong numerical solver algorithms specifically devel-
oped for this kind of numerical problem. Furthermore, ATHENA models
can be easily linked with state-of-the-art flow sheeting software pack-
ages such as e.g. ASPEN PLUS. A more detailed discussion of the
used modeling package can be found elsewhere, [185].
The electrochemical performance model calculates the current density
for a given fuel composition and operational voltage in a control volume.
The core of the model is the voltage balance according to which the
Nernst voltage minus the current density-dependent voltage losses has
to equal the operational voltage.
The mass balance model requires the current density as input for the
calculation of the conversion rates of the electrochemically active spe-
cies. According to the afore mentioned assumptions, these are hydro-
gen, water, carbon monoxide and carbon dioxide. Furthermore, the
mass balance model includes the calculation of reaction rates for all
considered homogenous and heterogeneous reactions based on applied
kinetic models. Besides the calculation of the conversion rates, the
mass balance model also calculates the related heat source terms and
the convective heat transport terms, which are coupled to the mass
transport to and from the catalytic surface.
The energy balance model serves for the calculation of the effective
temperatures in the solid structure and the gas channels based on the
heat source terms and the convective heat transport terms stemming
from the mass balance model. Furthermore, the energy balance model
includes the calculation of the purely convective heat transfer between
the gases and the solid structure as well as the conductive heat trans-
port within the solid structure.
The coefficients (B,C,D,E and F) for the considered pure specie (sub-
script x) in the model were taken from the flow sheeting package AS-
PEN PLUS, Table 5-1. Gas mixture heat capacities are calculated
through molar fraction weighting.
The viscosity and thermal conductivity of the pure gas species consid-
ered in the model are also computed according to DIPPR correlations,
Eq. 38 and Eq. 39, respectively.
K x ⋅ T Lx
η xv =
M N Eq. 38
1 + x + 2x
T T
141
5 Modeling
Gx ⋅ T H x
λvx =
I J Eq. 39
1 + x + x2
T T
The coefficients were taken from ASPEN PLUS and are summarized in
Table 5-2 and Table 5-3.
⎜ ⎜ M j ⎟⎟
⎝ ⎝ ⎠⎠
142
5 Modeling
Eq. 40 involves the molar fractions of the mixture, y, the pure specie vis-
cosity, ηv, and the interaction parameter φij which further involves the
molar weight of the pure species, M.
The calculation of the thermal conductivity of mixtures is done using the
Wassiljewa-Mason-Saxena mixing rule, Eq. 41, [194].
0 .5
⎛ ⎛ η v ⎞ 0.5 ⎛ M ⎞
0.25
⎞
⎜1 + ⎜ i ⎟ ⎜ j ⎟ ⎟
⎜ ⎜η v ⎟ ⎜ M ⎟ ⎟
λvmix = ∑
yi ⋅ λvi
with Aij = ⎝ ⎝ j⎠ ⎝ i ⎠ ⎠ Eq. 41
i ∑ y j ⋅ Aij
j
⎛ ⎛ M ⎞⎞
⎜ 8 ⋅ ⎜1 + i ⎟ ⎟
0.5
⎜ ⎜ M j ⎟⎟
⎝ ⎝ ⎠⎠
found in [194]. The second method is the empirical Fuller method de-
rived from an extensive experimental data base and featuring an aver-
age variance of 4 %. The Fuller equation, Eq. 43, comprises the tem-
perature, T, the pressure, p, the molecular weight relation between the
diffusing specie (subscript x) and the matrix specie (subscript z), Mzx,
and the atomic diffusion volumes, Vz and Vx. The atomic diffusion vol-
umes have to be determined experimentally, which is the major draw-
back of the Fuller method.
−1
0.00143 ⋅ T 1.75 ⎛ 1 1 ⎞
Dm , zx = with M zx = 2 ⋅ ⎜⎜ + ⎟⎟ Eq. 43
(
p ⋅ M zx0.5 ⋅ Vz0.33 + Vx0.33 )
2
⎝ Mz Mx ⎠
Based on the atomic diffusion volumes, which are only available for very
common species such as e.g. hydrogen, carbon monoxide, water etc.,
the Fuller method is less flexible than the kinetic gas theory method.
Approximation correlations are available for the atomic diffusion volume
of more exotic species, [194], but the impact on the accuracy of the re-
sults is unknown. Nevertheless, the Fuller method was chosen as stan-
dard method, the reason being that the partial pressures of the exotic
species considered in the developed model are expected to be small
compared to the common species. The employed atomic diffusion vol-
umes are given in Table 5-4.
1 − yx
Dm,mix , x =
y
∑i D i Eq. 44
m ,i , x
Eq. 44 comprises the molar fraction of the diffusing specie, yx, the molar
fractions of the all species in the gas mixture, yi, and the binary molecu-
lar diffusion coefficients of the diffusing specie with all other species in
the gas mixture, Dm,i,x.
⎛ 10300.0 ⎞
ρ el = 3.34 ⋅104 ⋅ exp⎜ ⎟ Eq. 46
⎝ TsK ⎠
The variable TsK denotes the solid material temperature. The corre-
sponding curves are depicted in Figure 5-5.
It can be seen, that even at very elevated temperatures the electric con-
ductivity of the employed materials for the anode and cathode elec-
trodes as well as for the current collectors (also called interconnectors,
IC) are three orders of magnitude higher than the ionic conductivity of
the electrolyte. Further it can be observed, that small temperature varia-
tions from e.g. 1100 K to 1300 K result in an increase of the ionic con-
ductivity by a factor of five.
145
5 Modeling
[1/Ohm m]
ionic conductivity 20000 10
of electrolyte
(right scale)
10000 5
0 0
300 400 500 600 700 800 900 1000 1100 1200 1300 1400
Temperature [K]
146
5 Modeling
Figure 5-6:
Generalized pla- Cell width
Fu
el
ga
s
Ox Cell length
ida
nt
gas Macro and micro
geometry of the
planar cell
repeating structure
Interconnect
Anode channel
Interconnect
Cathode channel height
height
on anode side
Channel width
Rib width
Anode channel
Thickness of anode
AEC assembly Thickness of electrolyte
Thickness of cathode
Cathode channel
Cathode
channel
Interconnect
height
Anode channel height on
cathode side
147
5 Modeling
148
5 Modeling
assumed. In the case of the tubular geometry, the fuel cell tube itself
was considered as a repeating structure in the complete fuel cell stack.
Similar to the planar cell geometry, the micro geometry of the tubular
cell design is entirely described by the thicknesses of the anode, δt,an,
the electrolyte, δt,el, and the cathode, δt,ca. The complete description of
the tubular cell design macro geometry encompasses the active tube
length, lt, the inner, ri,ADT, and outer radius, ro,ADT, of the air delivery
tube, the inner radius of the fuel cell tube, ri,FC, and the height of the vir-
tual anode channel, ht,an.
Inner radius
of air delivery tube Thickness Virtual anode
Outer radius of cathode channel height
of air delivery tube
Thickness Thickness
Inner radius of cathode layer of electrolyte of anode
149
5 Modeling
150
5 Modeling
Figure 5-8:
Triangular cell
geometry
The outer upper surface of the D8 cell and the outer lower surface of the
next D8 cell form the anode gas channels, Figure 5-8d. Note that the D8
cells are stacked like planar cells, Figure 5-8a. For better contacting and
current distribution, a porous contact layer is placed between the indi-
vidual cells. Due to its high electrical conductivity and porosity, the con-
tact layer was not considered in the developed model. The D8 repeating
structure was defined as one triangular-shaped fuel cell tube.
Similar to the above discussed geometries, the micro geometry of the
triangular cell design is entirely described by the thicknesses of the an-
ode, δD8,an, the electrolyte, δD8,el, and the cathode, δD8,ca, Figure 5-8c.
The description of the triangular cell design macro geometry is more
complex than that of the standard tubular and planar cell design. How-
ever, the complexity can be reduced by introducing geometrical de-
pendencies for the curvature radii. For the model, it was assumed, that
the upper outer curvature radius ro equals 1.5 times the D8 cell tube
wall thickness δ. The lower outer curvature radius rm was assumed
equal to the cell tube wall thickness and the inner curvature radius ri
equal to half the cell tube wall thickness, Figure 5-8b.
Based on the above assumptions and the overall cell width, wD8, the cell
length, lD8, the upper triangle half-angle, α, and the number of triangu-
151
5 Modeling
lar-shaped fuel cell tubes forming the D8 cell, ntri, the following charac-
teristic lengths and areas were derived for the triangular geometry:
Electrochemically and chemically active width, lD8,elact
2δ 90 − α
lD8,elact = lD8,chact = + ⋅π ⋅δ + 2 ⋅ X
ntri ⋅ cos α ntri ⋅ 90
Eq. 69
w 5 δ 90 − α
with X = D8 − ⋅ + ⋅π ⋅δ
ntri 2 tan α 72
The active width of one control volume is the arithmetic middle of the
overall active width of the D8 cell and the number of triangular cell tubes.
Note that the additional electrochemically active areas of the lower cur-
vature of the two outer triangular cells were also considered, Figure 5-8d.
Total active area, AD8,act
AD 8,act = lD 8,elact ⋅ lD 8 Eq. 70
152
5 Modeling
Figure 5-9:
Equivalent circuit Hydrogen electrochemistry
Operational
voltage
of the electro-
Current originating from
chemical per- Nernst
hydrogen conversion
voltage
formance model
Activation Diffusion Oxygen electrochemistry
polarization polarization
Current originating from
hydrogen and CO conversion
153
5 Modeling
⎛ R ⋅ TsK ⎞ ⎛⎜ ⎛ pH ⋅ pO 0.5 ⎞ ⎞
⎜ 2 ⎟⎟
E Nernst ,H 2 = ⎜ ⎟ ⋅ ln Kp + ln 2
Eq. 80
⎝ n⋅F ⎠ ⎝ ⎜ H 2 oxi
⎜ p ⎟⎟
⎝ H 2O ⎠⎠
In Eq. 80 and Eq. 81, R denotes the ideal gas constant, TsK stands for
the solid structure temperature, n represents the number of transferred
electrons (in both cases 2), F is the Faraday constant, Kp is the equilib-
rium constant and the factors pi represent the partial pressures of the
products and educts of the oxidation reactions in the bulk gas phase.
In this work, the equilibrium constant values for the hydrogen and car-
bon monoxide oxidation reaction were taken from [195]. The tabulated
values were compiled to yield two fit correlations, which both take the
form of Eq. 82.
⎛ − TsK ⎞ ⎛ − TsK ⎞ ⎛ − TsK ⎞
X = y0 + A1 exp⎜⎜ ⎟⎟ + A2 exp⎜⎜ ⎟⎟ + A3 exp⎜⎜ ⎟⎟ Eq. 82
⎝ t1 ⎠ ⎝ t2 ⎠ ⎝ t3 ⎠
154
5 Modeling
Eop = E Nernst ,H 2 − η act ,H 2 − η diff ,H 2 − η act ,O2 − η diff ,O2 − η ohm Eq. 83
Eop = E Nernst ,CO − η act ,CO − η diff ,CO − η act ,O2 − η diff ,O2 − η ohm Eq. 84
From Eq. 83 and Eq. 84 or Figure 5-9 it can be seen, that a current
originating from carbon monoxide oxidation can only be produced, when
the corresponding Nernst voltage minus the voltage losses of the oxy-
gen electrochemistry and the ohmic losses, due to an already flowing
current originating from hydrogen oxidation, still yields a positive value.
The same applies vice versa.
CO + O 2− ↔ CO2 + 2e − Eq. 87
O2 + 4e − ↔ 2O 2 − Eq. 88
155
5 Modeling
⎡ ⎛ n ⋅ F ⋅ηact ⎞ ⎛ n ⋅ F ⋅ηact ⎞⎤
I = I 0 ⎢exp⎜ β ⋅ ⎟ − exp⎜ − (1 − β ) ⋅ ⎟ Eq. 89
⎣ ⎝ R ⋅ TsK ⎠ ⎝ R ⋅ TsK ⎠⎥⎦
Other correlations such as the Tafel equation, Eq. 90, the linear current-
potential equation, Eq. 91, and more empirical relations can also be
found in the literature and are mostly derivates of the Butler-Volmer
equation. A detailed discussion of the application of these simplified
equations is given in [141].
R ⋅ TsK ⎛I ⎞
η act = ⋅ ln⎜⎜ ⎟⎟ Eq. 90
n ⋅ F ⋅ β ⎝ I0 ⎠
R ⋅ TsK I
η act = ⋅ Eq. 91
n ⋅ F I0
The Butler-Volmer equation, Eq. 89, which was implemented in the de-
veloped model, comprises two important model parameters. The first
parameter is the transfer coefficient β, which represents the part of the
change in polarization leading to a change in the reaction rate constant.
For fuel cell applications and in the developed model, the value is usu-
ally assumed to be 0.5, [182, 196].
The second parameter is the exchange current density I0, which repre-
sents the forward and reverse electrode reaction rate at the equilibrium
potential. High exchange current density values denote a high electro-
chemical reaction rate and hence low activation losses. More detailed
information can be found in [197].
The literature discusses a vast number of models used to calculate the
exchange current density at the anode and cathode. A detailed discus-
sion of these models is given in [182]. In the present work, the anode
exchange current density is calculated according to Eq. 92 for the hy-
drogen electrochemistry and Eq. 93 for the carbon monoxide electro-
chemistry. The exponent near to the stoichiometry of the anodic electro-
chemical reactions results in a direct proportional dependence of the
anode exchange current density to the educt and an inversely propor-
tional dependence to the product partial pressures. With the inverse de-
pendency of the activation losses to the exchange current density, the
exponents yield lower activation losses with increased educt or de-
creased product partial pressures and vice versa, [182]. The cathode
exchange current density is given by Eq. 94. The positive exponent
yields lower cathode activation losses with rising oxygen partial pres-
sures.
156
5 Modeling
−0.5
⎛ pH ⎞⎛ pH 2 O ⎞ ⎛ − Eact , an , H 2 ⎞
I 0, an , H 2 = γ an , H 2 ⎜⎜ 2 ⎟⎟⎜⎜ ⎟⎟ ⋅ exp⎜⎜ ⎟⎟ Eq. 92
⎝ p ⎠⎝ p ⎠ ⎝ R ⋅ TsK ⎠
−0.5
⎛ p ⎞⎛ pCO2 ⎞ ⎛ − Eact , an ,CO ⎞
I 0, an ,CO = γ an ,CO ⎜⎜ CO ⎟⎟⎜⎜ ⎟⎟ ⋅ exp⎜⎜ ⎟⎟ Eq. 93
⎝ p ⎠⎝ p ⎠ ⎝ R ⋅ TsK ⎠
0.25
⎛ pO ⎞ ⎛ − Eact ,ca ⎞
I 0, ca = γ ca ⎜⎜ 2 ⎟⎟ ⋅ exp⎜⎜ ⎟⎟ Eq. 94
⎝ p ⎠ ⎝ R ⋅ TsK ⎠
157
5 Modeling
⎛ ρ an ρ ca ⎞
⎛ ρ an ρ ca ⎞ ρ elδ el ⎜⎜ + ⎟
δ ca ⎟⎠
1
L ⎜ + ⎟ ⎝ δ an
ρ elδ el ⎜⎝ δ an δ ca ⎟⎠
tanh J
5
in parallel, representing the two halves of the cell tube, and one type 2
cell sub-unit, representing the interconnect area, connected in series.
Figure 5-10:
Cross section of
the standard tubu- Interconnect
lar Siemens AG
design (a) and a Cathode
b
transmission line
model (b)
Anode
Electrolyte
Due to the perfect symmetry of the tubular geometry, the equivalent re-
sistance is only calculated for one half-cell and then divided by 2. The
length of the type 3 cell sub-unit (Le) for one half-cell is calculated ac-
cording to Eq. 96 and for cell sub-unit type 2 (Lic) according to Eq. 97.
Le = π ⋅ rm , AEC ⋅ (1 − xt ,ic ) Eq. 96
The resistance of one half of the cell tube (Re) is given by Eq. 98.
⎛ ⎛ ρ ⎞2 ⎛ ρ ⎞2 ⎞
⎜ ⎜ an ⎟ + ⎜ ca ⎟ ⎟ ⋅ cosh J + ⎛⎜ ρ an ρ ca ⎞⎟ ⋅ (2 + J ⋅ sinh J )
⎜ ⎜⎝ δ an ⎟⎠ ⎜⎝ δ ca ⎟⎠ ⎟ e ⎜δ δ ⎟
⎝ an ca ⎠
e e
Re = ⎝ ⎠
1 ρ an ρca
⋅3 + ⋅ sinh J e
ρelδ el δ an δ ca Eq. 98
1 ⎛ ρ an ρ ca ⎞
with J e = Le ⎜ + ⎟
ρ elδ el ⎜⎝ δ an δ ca ⎟⎠
159
5 Modeling
Figure 5-11:
Cross section of
the triangular
Siemens AG de-
sign (a) and
transmission line
model (b)
The cell sub-unit lengths, Le and Lic, are given by Eq. 101 and Eq. 102.
LD 8, e = 0.5 ⋅ lD 8, elact Eq. 101
wD 8
LD 8,ic = Eq. 102
2 ⋅ ntri
As only one half of the current path is considered, the area specific re-
sistance is obtained by multiplying the equivalent resistance with half of
the electrochemically active width of the repeat element, Eq. 103.
RD 8, equiv = 0.5 ⋅ lD 8,elact ⋅ ( Re + Ric ) Eq. 103
160
5 Modeling
Figure 5-12:
Cross section of a b
the considered Interconnect
planar cell design Cathode channel 6: Type 1
6
(a) and transmis- Interconnect
sion line model (b)
Anode channel 4: Type 1 5: Type 5*
4 5
AEC assembly
3: Type 1
3
Cathode channel
Anode
Electrolyte 1: Type 4 2: Type 5
Anode channel 1 2 Cathode
⎛ ρ an ρca ⎞
ρ elδ el ⎜⎜ + ⎟
⎝ δ an δ ca ⎟⎠ wch 1 ⎛ ρ an ρ ca ⎞ Eq. 105
R2 = with J 2 = ⎜ + ⎟
tanh J 2 2 ρelδ el ⎜⎝ δ an δ ca ⎟⎠
The two cell sub-units are connected in parallel. The interconnect con-
tacts anode and cathode of two superposed AEC assemblies. The resis-
tances of the interconnect ribs on the anode (R3) and cathode (R6) side
are calculated using type 1 cell sub-units, Eq. 106 and Eq. 107.
2 ⋅ ρic hp , an
R3 = Eq. 106
wc nch − wch
2 ⋅ ρic hp , ca
R6 = Eq. 107
wc nch − wch
161
5 Modeling
ρic
R5 =
⎛ ⎛ ⎛ ⎞ ⎞ ⎞⎟
0.41 ⋅ ⎜1 − exp⎜ − 0.6 ⋅ ⎜
wch ⎟⎟ Eq. 109
⎜ ⎜ ⎜ (h + h ) − (h + h ) ⎟ ⎟ ⎟
⎝ ⎝ ⎝ ic , an ic , ca p , an p , ca ⎠⎠⎠
Finally, the equivalent circuit depicted in Figure 5-12 yields Eq. 110.
Similar to the tubular geometry, the area specific resistance is multiplied
with the electrochemically active length of the repeat structure.
⎛ w ⎞ ⎛⎛ 1 ⎞
−1 −1
1 ⎞ ⎛ 1 1 ⎞
R p , equiv = ⎜⎜ ce ⎟⎟ ⋅ ⎜ ⎜⎜ + ⎟⎟ + R3 + ⎜⎜ + ⎟⎟ + R6 ⎟ Eq. 110
⎝ nch ⎠ ⎜⎝ ⎝ R1 R2 ⎠ ⎝ R4 R5 ⎠ ⎟
⎠
In these equations, TaK denotes the temperature of the anode gas flow,
IH2 and ICO represent the currents originating from the hydrogen and
carbon monoxide conversion, δan stands for the thickness of the anode
electrode, τan gives the tortuosity and εan the porosity of the anode elec-
trode material and Deff,H2 and Deff,CO denote the effective diffusion coef-
ficients of hydrogen and carbon monoxide of the anode gas mixture.
The effective diffusion coefficients are calculated based on binary mo-
lecular diffusion coefficients of hydrogen and carbon monoxide with all
other considered species in the gas mixture and the Knudsen diffusion
coefficient. The calculation of binary molecular and Knudsen diffusion
coefficients is discussed in section 5.3.1.2.1. The binary molecular diffu-
sion coefficients are summed up to yield the effective molecular diffusion
coefficient according to Eq. 44. For the calculation of the effective diffu-
sion coefficient it is assumed, that Knudsen and molecular diffusion oc-
cur simultaneously. Therefore, the Bosanquet formula was implemented
in this model, Eq. 112, where x denotes either hydrogen or carbon mon-
oxide.
Dm, x − mix ⋅ DK , x
Deff , x = Eq. 112
Dm , x − mix + DK , x
For known partial pressures of the educts and products in the bulk gas
phase and at the TPB, the diffusion voltage losses are computed ac-
cording to Eq. 113. The temperature used in the calculation is the tem-
perature of the solid structure of the fuel cell and not that of the gas mix-
ture.
⎛ R ⋅ TsK ⎞ ⎛⎜ pi ⋅ p j ⎞⎟
TPB
163
5 Modeling
pOTPB −p
A⋅ = exp(I tot ⋅ TcK ⋅ B )
( )
2
ε ca ⋅ 2 ⋅ n ⋅ F ⋅ Dm,O − N ⋅ DK ,O ⋅ p
2 2 2
⎛ M O2 ⎞
and α = 1 − ⎜ ⎟
⎜ MN ⎟
⎝ 2 ⎠
Eq. 114 was implemented in this model for the implicit calculation of the
partial pressure of oxygen at the cathode TPB, [155]. In contrast to the
diffusion at the anode, the effective diffusion coefficient is calculated for
the non-dilute two-component gas mixture at the cathode according to
[201], which includes the relation between the molecular masses, α, of
the components nitrogen and oxygen. Further, Eq. 114 comprises the
porosity and tortuosity of the cathode electrode, τca and εca, the cathode
electrode thickness, δca, the Knudsen diffusion coefficient of oxygen,
DK,O2, the binary molecular diffusion coefficient of the oxygen-nitrogen
system, Dm,O2-N2, the prevalent total current density, Itot, and the tem-
perature of the cathode gas, TcK. The diffusion voltage loss at the
cathode is calculated analogously to the anode, Eq. 115.
⎛ R ⋅ TsK ⎞ ⎛⎜ pO2 ⎞
⎟
ηdiff ,O = ⎜ ⎟ ⋅ ln TPB Eq. 115
2
⎝ 2 ⋅ n ⋅ F ⎠ ⎜⎝ pO2 ⎟
⎠
164
5 Modeling
⎛ IH ⎞
rH 2 oxi = lt ⋅ ⎜⎜ 2 ⎟⎟ Eq. 116
⎝ 2⋅ F ⎠
p , elact
⎛ I ⎞
rCO oxi = lt ⋅ ⎜ CO ⎟ Eq. 117
⎝ 2⋅ F ⎠
p , elact
⎛ I ⎞
rO2 ion = lt ⋅ ⎜ tot ⎟ Eq. 118
⎝ 4⋅ F ⎠
p , elact
CH 4 + H 2O ↔ CO + 3H 2 Eq. 121
To facilitate the extension of the present model with respect to the con-
sidered chemical reactions and applied kinetic models, a generalized ki-
netic approach was developed, Eq. 126.
A⋅ B ⋅C ⋅ D
rreac = f act ⋅
E⋅F
⎛ − Ea ⎞
with A = k0 ⋅ exp⎜ ⎟
⎝ R ⋅T ⎠
and B = pero1 e1 ⋅ pero2e 2
C = p p1p1 ⋅ p p 2p 2
ro ro
and
ν ν
p p1p1 ⋅ p p p22
D =1−
and
(
K eq ⋅ pνe1e1 ⋅ pνe 2e 2 )
Eq. 126
and E = K eq
and F = (1 + G + H + I ) ro ads
⎛ ΔH ads1 ⎞ ro ads11 ro ads12
with G = K ads
0
1 ⋅ exp⎜ ⎟ ⋅ pads11 ⋅ pads12
⎝ R ⋅T ⎠
⎛ ΔH ads 2 ⎞ roads 21 roads 22
with H = K ads
0
2 ⋅ exp⎜ ⎟ ⋅ pads 21 ⋅ pads 22
⎝ R ⋅T ⎠
⎛ ΔH ads 3 ⎞ roads 31 ro ads 32
with I = K ads
0
3 ⋅ exp⎜ ⎟ ⋅ pads 31 ⋅ pads 32
⎝ R ⋅T ⎠
165
5 Modeling
166
5 Modeling
adapt an available applied kinetic model for another reaction to the reac-
tion in question. The value of fact has to be determined by comparison of
the reaction in question with the reaction for which the applied kinetic
model has been developed. Although this approach lacks a solid data
base, it is nevertheless valuable as it hints towards possible impacts of
hitherto uninvestigated reactions in fuel cells such as e.g. the steam re-
forming of tars.
The water-gas-shift reaction (WGS), Eq. 119, and the reverse water-
gas-shift reaction (REV-WGS), Eq. 120, were assumed to be at equilib-
rium. This assumption is found frequently in the literature, e.g. [141, 202,
203], and was confirmed in experiments conducted by the author, see
section 3.3.2. Besides, the WGS and REV-WGS were considered to be
non-diffusion limited homogenous reactions. The reaction rates were
calculated according to Eq. 127 and Eq. 128.
mol Eq.
rWGS = 10000.0 ⋅ pCO ⋅ pH 2 O
m ⋅ sec⋅ bara 2 127
mol ⎛ pCO ⋅ pH 2O ⎞
rREV −WGS = 10000.0 ⋅ pCO2 ⋅ pH 2 ⋅ ⎜1 − ⎟ Eq.
m ⋅ sec⋅ bara 2 ⎜ K eq ,WGS ⋅ pCO ⋅ pH ⎟ 128
⎝ 2 2 ⎠
Table 5-7 lists the coefficients for the equilibrium constant fit correlation,
Eq. 82, of the WGS and the methane steam reforming reaction (CH4
STR). The equilibrium constant values were taken from [195].
167
5 Modeling
positive reaction order, [53]. Drescher has shown, that all these findings
are correct and a result of the chosen operating conditions of the ex-
periments, more precisely of the chosen steam-to-carbon ratio (SC),
[207]. Small SC yield positive reaction orders, SC in the order of 2 yield
reaction orders of zero while high SC lead to negative reaction orders of
water. Drescher developed a Langmuir-Hinshelwood kinetics model for
temperatures around 650 °C.
To allow for an analysis of the impact of the kinetic models on the model
results, Eq. 129 to Eq. 133 and the equilibrium approach were imple-
mented. The equilibrium constant for the CH4 STR is calculated using
the fit correlation coefficients given in Table 5-7. Note that Dreschers ki-
netic model, Eq. 133, was initially related to the mass of catalyst in the
fuel cell. Consequently the pre-exponential factor had to be recalculated
to an area specific value to allow for the application in the presented
model. The author assumed that only the upper 25 microns of the
nickel-cermet participate in the steam reforming reactions. All other data
for the recalculation were taken from [207] in order to keep the kinetic
model concise.
mol ⎛ pCO ⋅ p H3 2 ⎞
rCH 4 STR ,eq = 10000.0 ⋅ p ⋅ p ⋅
H 2O ⎜
⎜ 1 − ⎟ Eq. 129
m ⋅ sec⋅ bara 2 ⎟
⎝ K eq , STR ⋅ pCH 4 ⋅ p H 2O
CH 4
⎠
⎛ J ⎞
mol ⎜ − 82000 ⎟
rCH 4 STR , Ach = 4274.0 2 ⋅ pCH 4 ⋅ exp⎜ mol ⎟ Eq. 130
m ⋅ sec⋅ bara ⎜ R ⋅ TsK ⎟ [204]
⎜ ⎟
⎝ ⎠
⎛ J ⎞
mol ⎜ − 95000 ⎟
rCH 4 STR , Ahm = 8542.0 2 ⋅ pCH
0.85
⋅ pH− 02.O35 ⋅ exp⎜ mol ⎟ Eq. 131
m ⋅ sec⋅ bara 2 4
⎜ R ⋅ TsK ⎟ [205]
⎜ ⎟
⎝ ⎠
⎛ J ⎞
mol ⎜ − 205000 ⎟
rCH 4 STR , Lei = 30.8e + 10 2 ⋅ pCH 4 ⋅ pH 2 O ⋅ exp⎜ mol ⎟ Eq. 132
m ⋅ sec⋅ bara 2 ⎜ R ⋅ TsK ⎟ [53]
⎜ ⎟
⎝ ⎠
⎛ J ⎞
⎜ − 11000 ⎟
mol
288.52 2 ⋅ p ⋅ p ⋅ exp ⎜ mol ⎟
m ⋅ sec⋅ bara 2 ⎜ R ⋅ TsK ⎟
CH 4 H 2O
⎜ ⎟
rCH 4 STR , Dre = ⎝ ⎠ Eq. 133
⎛ ⎛ J ⎞ ⎞ [207]
⎜ ⎜ 39000 ⎟⎟
⎜1 + 16.0 1 1 mol ⎟ ⎟
⋅ pCH 4 + 0.143 ⋅ pH 2O ⋅ exp⎜
⎜ bara bara ⎜ R ⋅ TsK ⎟⎟
⎜ ⎜ ⎟⎟
⎝ ⎝ ⎠⎠
168
5 Modeling
For the reactions Eq. 122 to Eq. 125, no valid kinetic models regarding
the application in a solid oxide fuel cell model were found in the litera-
ture. Therefore, it was assumed, that the STR of higher hydrocarbons
follows the kinetic model of Achenbach, [204]. Further it was assumed
that the reactions occur faster than the CH4 STR due to the higher reac-
tivity of the considered hydrocarboneous species. Following this as-
sumption, the activity factor, fact, was set to 2.5, 1.5, 2.0 and 1.75 for the
reactions Eq. 122, Eq. 123, Eq. 124 and Eq. 125, respectively. The sub-
sequent expressions for the reaction rates of the STR of ethene, toluene
and anisole and the thermal decomposition of acetic acid were formu-
lated as:
⎛ J ⎞
mol ⎜ − 82000 ⎟
rC 2 H 4 STR , Ach = 10685.0 2 ⋅ pC 2 H 4 ⋅ exp⎜ mol ⎟ Eq. 134
m ⋅ sec⋅ bara ⎜ R ⋅ TsK ⎟
⎜ ⎟
⎝ ⎠
⎛ J ⎞
mol ⎜ − 82000 ⎟
rC7 H 8 STR , Ach = 6411.0 2 ⋅ pC7 H 8 ⋅ exp⎜ mol ⎟ Eq. 135
m ⋅ sec⋅ bara ⎜ R ⋅ TsK ⎟
⎜ ⎟
⎝ ⎠
⎛ J ⎞
mol ⎜ − 82000 ⎟
rC7 H 8O STR , Ach = 7479.5 2 ⋅ pC7 H 8 O ⋅ exp⎜ mol ⎟ Eq. 136
m ⋅ sec⋅ bara ⎜ R ⋅ TsK ⎟
⎜ ⎟
⎝ ⎠
⎛ J ⎞
⎜ − 82000 ⎟
mol
rC2 H 4 O2 Dec , Ach = 8548.0 2 ⋅ pC 2 H 4O2 ⋅ exp⎜ mol ⎟ Eq. 137
m ⋅ sec⋅ bara ⎜ R ⋅ TsK ⎟
⎜ ⎟
⎝ ⎠
169
5 Modeling
The diffusion limited and integration length specific reaction rate, rdl-reac,
is obtained from Eq. 141.
p xro x β x , diff ⋅ β x , reac
rdl − reac = lt p , chact ⋅ β x , diff − reac ⋅ with β x , diff − reac = Eq. 141
R ⋅ TsK β x , diff + β x , reac
In this equation, lt/p,chact denotes the chemically active length of the con-
sidered fuel cell design and βx,diff-reac represents the diffusion limited re-
action conversion coefficient. The partial pressure of the diffusion-limited
specie, px, to the power of its reaction order, rox, is thus reintroduced.
Knowing the reaction rate of all reactions and neglecting axial diffusion
mass transport, the spatial distribution of the species along the anode
channel can be computed according to following differential equations:
170
5 Modeling
dn& H 2
= RH 2
dx
with RH 2 = − rH 2 oxi + 3 ⋅ rdl − CH 4 STR + rWGS − rREV −WGS + 4 ⋅ rdl − C 2 H 4 STR Eq. 142
dn&CO
= RCO
dx
with RCO = − rCO oxi + rdl − CH 4 STR − rWGS + rREV −WGS + 2 ⋅ rdl − C 2 H 4 STR Eq. 143
dn&CO2
= RCO2 with RCO2 = rCO oxi + rWGS − rREV −WGS Eq. 144
dx
dn& H 2O
= RH 2O
dx
with RH 2O = rH 2 oxi − + rdl − CH 4 STR − rWGS + rREV −WGS − 2 ⋅ rdl − C 2 H 4 STR Eq. 145
dn&CH 4
= RCH 4 with RCH 4 = −rdl −CH 4 STR Eq. 146
dx
dn&C 2 H 4
= RC 2 H 4 with RC 2 H 4 = − rdl − C 2 H 4 STR Eq. 147
dx
dn& N 2 , an
= 0 .0 Eq. 148
dx
dn&C7 H 8
= RC 7 H 8 with RC 7 H 8 = − rdl − C7 H 8 STR Eq. 149
dx
dn&C 2 H 4 O2
= RC 2 H 4 O2 with RC 2 H 4 O2 = − rdl − C 2 H 4 O2 Dec Eq. 150
dx
dn&C7 H 8O
= RC 7 H 8O with RC7 H 8O = −rdl − C 7 H 8O STR Eq. 151
dx
171
5 Modeling
dn& N 2 , ca
= 0.0 Eq. 152
dx
dn&O2
= RO2 with RO2 = −rO2 ion (co − flow) or RO2 = rO2 ion (counter − flow) Eq. 153
dx
The total molar flow in the anode and the cathode channel are calcu-
lated according to Eq. 154 and Eq. 155.
n&an = n&H 2 + n&CO + n&CO2 + n& H 2O + n&CH 4 + n&C2 H 4
Eq. 154
+ n& N 2 , an + n&C7 H 8 + n&C2 H 4O2 + n&C7 H 8O
Based on the calculated reaction rates, the mass balance model also
computes the related heat source term according to Eq. 156.
ΔH r = rH 2 oxi ⋅ ΔH R , H 2 oxi + rCO oxi ⋅ ΔH R ,CO oxi + rdl − CH 4 STR ⋅ ΔH R ,CH 4 STR
+ rWGS ⋅ ΔH R ,WGS + rREV −WGS ⋅ ΔH R , REV −WGS + rdl − C 2 H 4 STR ⋅ ΔH R ,C 2 H 4 STR
Eq. 156
+ rdl − C 7 H 8 STR ⋅ ΔH R ,C 7 H 8 STR + rdl − C 2 H 4O2 Dec ⋅ ΔH R ,C 2 H 4O2 Dec
+ rdl − C 7 H 8O STR ⋅ ΔH R ,C7 H 8O STR
The heats of reaction, ΔHR,reaction, are calculated via Eq. 157, where i
denotes each considered specie, νi gives the stoiciometric coefficient of
the specie in the considered reaction, ΔH0f,i stands for the ideal gas en-
thalpy of formation and T0 is the temperature at standard conditions.
TsK
ΔH R , reaction = ΔH R0 , reaction + ∑ν i ⋅ ∫ c p ,i dT
i T0
with ΔH 0
R , reaction = ∑ν i ⋅ ΔH o
f ,i Eq. 157
i
⎛ D ⎞ ⎛ F ⎞
TsK
and ∫c
T0
p ,i dT = B ⋅ TsK + C ⋅ D ⋅ coth⎜ ⎟ − E ⋅ F ⋅ tanh⎜
⎝ TsK ⎠
⎟−O
⎝ TsK ⎠
The coefficient O and the applied ideal gas enthalpy of formation value
for the considered species are given in Table 5-8, whereas the coeffi-
cients B, C, D, E and F can be found in Table 5-1. All values were taken
from ASPEN PLUS.
By definition, the heats of reaction of all heterogeneous reactions are at-
tributed to the solid structure of the fuel cell. The differing heat capaci-
ties of educt and product species as well as the different temperature of
the gas phase and the solid structure result in an enthalpy flux coupled
to the mass flow from the solid structure to the gas phase and vice versa.
172
5 Modeling
This mass transport coupled enthalpy flux was considered in this model.
Eq. 158 to Eq. 160 give the mass transfer coupled enthalpy flux of the
educts, Q& SH ,ed , an , and products, Q& SH , prod , an , at the anode and the enthalpy
flux at the cathode, Q& SH ,ed ,ca .
In these equations, rj stands for the reaction rate of the considered het-
erogeneous reactions, including the electrochemical reactions, occurring
at the anode, Eq. 121 to Eq. 125 as well as Eq. 86 and Eq. 87, and rO2
ion represents the reaction rate of reaction Eq. 88. TaK, TcK and TsK
denote the anode and cathode gas and the solid structure temperature,
respectively. Further, νi represents the stoiciometric coefficient of educt
(subscript ed) and product (subscript prod) species of the according re-
action. Finally, the heat capacity of the educts and products is calculated
at the gas or the solid structure temperature (subscripts TaK, TcK and
TsK).
173
5 Modeling
Figure 5-13:
Outline of the
energy balance
model
The energy balance of the anode gas channel includes the sensible
heat stream, n&an ⋅ c p ,an ⋅ TaK , entering the control volume at the coordi-
nate x and leaving it at x+dx. Furthermore, the enthalpy fluxes due to
the different heat capacity of the educt, Q& SH ,ed , an , and product species,
Q& SH , prod , an , reacting at the anode are accounted for. The value of the en-
thalpy fluxes is calculated according to Eq. 158 and Eq. 159, respec-
tively. Finally the conductive heat stream between solid structure and
anode gas channel is considered, Eq. 161. αan denotes the heat trans-
fer coefficient, TaK stands for the anode gas temperature and TsK
174
5 Modeling
It should be noted, that this model considers the change of the total mo-
lar flow, nan, and the change of the heat capacity, cp,an, due to the
change of the gas composition over the control volume. In some models
found in the literature, only the temperature change is accounted for by
assuming the heat capacity and sometimes also the molar flows are
fixed at the inlet conditions of the corresponding control volume, [140,
182, 208]. For coarse discretization aiming at short calculation times,
this can cause problems with the gas channel energy balance closure.
The energy balance of the solid structure involves the overall heat of re-
action, Eq. 156, and the mass transfer coupled enthalpy fluxes at the
anode and cathode electrode, Eq. 158 to Eq. 160. Besides this, the con-
vective heat streams between the solid structure and anode, Q& conv , an , as
well as cathode gas channel, Q& conv , ca , have to be taken into account. In
the present model, the produced electrical power, Pel, is regarded as a
source term for the solid structure and is computed according to Eq. 163.
Note that the total current density Itot results from the electrochemical
performance model and Eop is the user-defined operational voltage of
the fuel cell. In order to transform this area specific power density into
an integration length specific value, it has to be multiplied with the elec-
trochemically active length of the fuel cell design in question, lt/p,elact.
Pel = −lt p , elact ⋅ I tot ⋅ Eop Eq. 163
The convective heat stream between the cathode gas channel and the
solid structure is determined according to Eq. 164 considering the tem-
perature difference between cathode gas, TcK, and solid structure, TsK,
the effective heat transfer coefficient, αca, and the circumferential length
175
5 Modeling
of the cathode gas channel, lt/p,circ,ca. Analogously to Eq. 161, the heat
transfer coefficient of the cathode gas channel is calculated with the
temperature and gas composition dependent thermal conductivity of the
cathode gas, λca, and the hydraulic diameter of the cathode gas channel,
dt/p,hyd,ca, of the investigated fuel cell design. The Nusselt number, Nu, is
assumed to equal that of the anode gas channel, as in both cases lami-
nar flow is assumed. It is important to emphasize that the convective
heat stream of the cathode gas channel is considered positive, when
heat is transferred from the solid structure to the cathode gas. This is
due to the cooling task of the air flow in the cathode channel.
Nu ⋅ λca
Q& conv , ca = α ca ⋅ lt p , circ , ca ⋅ (TsK − TcK ) with α ca = Eq. 164
dt p , hyd , ca
Finally in this model, the energy balance of the solid structure includes
solid heat conduction. This yields the second order derivate differential
equation, Eq. 165, for the solid structure temperature, TsK, where λs
symbolizes the average solid heat conduction coefficient of the solid
structure and At/p,cross stands for the cross-sectional area which is per-
pendicular to the gas flow channels.
d 2TsK &
λs ⋅ At p , cross ⋅ = Qconv , ca − Q& conv , an − Q& SH ,ed , an − Q& SH , ed ,ca + Q& SH , prod , an + ΔH r + Pel Eq. 165
dx 2
All terms required to compute the cathode gas temperature of the planar
cell design have been previously mentioned and yield Eq. 166.
d (n&ca ⋅ c p , ca ⋅ TcK )
= Q& conv ,ca − Q& SH , ed , an + Q& conv, ADT (co − flow)
dx Eq. 166
= Q& SH , ed , an − Q& conv, ca (counter − flow)
176
5 Modeling
ture TadtK, Eq. 167. In this model, the two processes are accounted for
through an effective heat exchange coefficient, αADT,eff, which was pro-
posed in [209]. The calculation of αADT,eff includes the inner, ri,ADT, and
outer radius, ro,ADT, of the ADT, the heat exchange coefficients of the
cathode gas, αca, and inside the ADT, αADT. Also included is the solid
heat conduction coefficient of the ADT material, λs,ADT, which leads to
either enhanced or reduced convective heat exchange in cold or hot re-
gions of the ADT, respectively. αADT is computed analogously to the
other heat exchange coefficients.
Q& conv, ADT = α ADT , eff ⋅ lt , circ, ADT ⋅ (TadtK − TcK )
1 ⎛ 1 1 ⎛r ⎞ 1 ⎞
with α ADT ,eff = ⋅⎜ + ⋅ ln⎜ o , ADT ⎟+ ⎟
ri , ADT ⎜ ri , ADT ⋅ α ADT 4 ⋅ λs , ADT ⎜r ⎟ r ⋅ α ⎟ Eq. 167
⎝ ⎝ i , ADT ⎠ o , ADT ca ⎠
Nu ⋅ λ ADT
and α ADT =
d t , hyd , ADT
The energy balance of the air flow inside the ADT can then be formu-
lated as Eq. 168. Since the air flow inside the ADT is not partaking in
any chemical or electrochemical reactions, the molar flow is constant.
Accordingly the differential equation for the ADT energy balance was
simplified in the present model.
d (c p , ADT ⋅ TadtK )
n& ADT ⋅ = Q& conv , ADT Eq. 168
dx
177
5 Modeling
I tot , av ⋅ Ap t , act
n&an , 0 =
2 ⋅ F ⋅ UF ⋅ EqH 2
Eq. 169
with EqH 2 = y H 2 + yCO + 4 ⋅ yCH 4 + 6 ⋅ yC 2 H 4
+ 18 ⋅ yC7 H 8 + 4 ⋅ yC 2 H 4 O2 + 17 ⋅ yC 7 H 8 O
In Eq. 169, Ap/t,act denotes the electrochemically active area of the con-
sidered fuel cell design and EqH2 represents the hydrogen equivalent of
the prescribed input gas composition.
Pin
n&an , 0 = with nRE = nch ⋅ ncells or nRE = ntubes Eq. 170
LHVin ⋅ nRE
nRE in Eq. 170 stands for the number of repeating elements in the stack.
For the tubular design, the repeating element is a whole cell tube,
whereas for the planar design only one fuel and air channel is consid-
ered. Hence, for the planar design the number of repeating elements is
the product of the number of cells in the stack, ncells, and the number of
channels per cell, nch. Furthermore, LHVin stands for the lower heating
value of the prescribed fuel gas composition.
In this model, the standard cathode gas input composition is equal to
that of air or more precisely 79 mol-% nitrogen and 21 mol-% oxygen.
However, other compositions such as e.g. pure oxygen can also be de-
fined. The inlet molar flow of the cathode gas, nca,0, is either assigned
directly or by specifying an air-to-fuel ratio, λ.
n&O2 , stoic ⋅ λ
n&ca , 0 =
yO2
Eq. 171
with n&O2 , stoic = n&an , 0 ⋅ (0.5 ⋅ yH 2 + 0.5 ⋅ yCO + 2 ⋅ yCH 4
+ 3 ⋅ yC 2 H 4 + 9 ⋅ yC7 H 8 + 2 ⋅ yC 2 H 4O2 + 8.5 ⋅ yC7 H 8O )
In the latter case Eq. 171 applies, where nO2,stoic gives the oxygen molar
flow required for the stoiciometric combustion of the fuel gas. Depending
on the investigated fuel cell design and the considered flow pattern, the
mass balance boundary values were assigned to different CVs, see
Table 5-9 where (0) denotes the first and (n) the last CV.
178
5 Modeling
The total molar flow values nan, nca and nADT are related to the molar
flows of the single species via the user-defined molar fractions.
Besides the inlet molar flow, the present model requires the definition of
the gas temperatures at the inlet in order to solve the energy balance
equation of the gas channels. The inlet gas temperature of the anode
gas channel, Tan,in, the cathode gas channel, Tca,in and eventually of the
air deliver tube, TADT,in, are user-defined values and assigned as shown
in Table 5-10.
In case of the tubular cell design, the boundary conditions of the cath-
ode gas channel ensure the continuity between the temperature and the
molar flow of the air flow exiting the air delivery tube and entering the
cathode gas channel.
Owing to the considered solid heat conduction, the energy balance of
the solid structure includes a second derivate term. Hence, the solution
of each CV depends on the solutions of the neighboring CVs. Conse-
quently, boundary conditions have to be defined at both ends of the in-
tegration region.
The differential equation applying at the left boundary (first control vol-
ume of the integration region) is given by Eq. 172.
dTsK
λs ⋅ At p,cross ⋅ = Q& hloss Eq. 172
dx
Q& hloss represents the heat loss stream to the surroundings. At the right
boundary, the source terms of the last CV of the integration region have
to be taken into account, Eq. 173.
dTsK
λs ⋅ At p , cross ⋅ =
dx Eq. 173
(Q& conv,ca )
− Q& conv, an − Q& SH , ed , an − Q& SH , ed , ca + Q& SH , prod , an + ΔH r + Pel ⋅ dx − Q& hloss
For the tubular design the cell ends are generally assumed adiabatic. In
contrast, the developed model features the possibility to assume adia-
batic cell ends, radiative or conductive heat losses at the cell ends of
179
5 Modeling
planar cells. Where adiabatic cell ends are assumed for the simulation
of a perfect insulation, Q& hloss is set to zero. The value of the radiative
heat loss streams is calculated according to Eq. 174, [190], where αrad
denotes the radiative heat exchange coefficient, σ stands for the Stefan-
Boltzmann constant and Tsur represents the Temperature of the sur-
roundings.
⎛ Tsur + TsK ⎞
3
Q& hloss =α rad ⋅ At p,cross ⋅ (TsK − Tsur ) with α rad = 4 ⋅ σ ⋅ ⎜ ⎟ Eq. 174
⎝ 2 ⎠
180
5 Modeling
this model to handle fuel gases with almost unlimited hydrocarbon con-
tent. The investigation of different operating conditions implies fewer
convergence difficulties. Guidelines that might promote a faster outcome
can also be formulated. Changes of the model input that strongly influ-
ence the anode gas mass balance, e.g. increase of the fuel mass flow or
change of the fuel composition, have to be undertaken in small incre-
mental steps. On the other hand changes of e.g. the voltage can be per-
formed in comparably large increments. This is due to the absolute
mass flow values remaining virtually constant. Changes of input gas
temperatures have a strong impact on the energy balance and therefore
also require an iterative approach in small steps.
If the employed numerical solver is unable to find a solution, the values
of the state variables of the last solution need to be examined. If the val-
ues of any of the state variables have reached numbers very close to
zero or are disproportionately high the envisaged operational point could
be physically unfeasible. This physical unfeasibility might be the reason
for the perceived non-convergence. If on the other hand physical infea-
sibility was ruled out the solver could be blocked in a semi-stable solu-
tion point. In this case the only operational strategy is to return to a sta-
ble and physically adequate solution and change the input for the next
interim solution so that the semi-stable point is avoided.
Table 5-11: Specie Unit Fuel gas IEA 1 Fuel gas IEA 2 Cathode gas Source
Fuel and cathode
Hydrogen 90.00 26.26 -
gas compositions
of the BMT Carbon monoxide - 2.94 -
Carbon dioxide - 4.36 -
Water vol.-% 10.00 49.34 - [112]
Methane - 17.10 -
Nitrogen - - 79.00
Oxygen - - 21.00
Table 5-12 gives the BMT geometry data of the planar SOFC stack.
Table 5-13 shows the operational conditions defined for the BMT.
182
5 Modeling
The BMT participants agreed to consider only ohmic losses for a better
comparability of the model results. The specific ohmic resistances of the
ceramic components were calculated according to the equations Eq. 45
to Eq. 48. For the planar SOFC, ceramic bipolar plates were assumed.
Further, the applied kinetics data for the steam reforming reaction (STR)
were prescribed, Eq. 130. Table 5-14 gives the BMT parameter used for
the energy balance calculations.
Besides the described model input data, several assumptions were de-
fined for the BMT, Table 5-15.
183
5 Modeling
nition of the sensible heat gradient by holding the molar flow and heat
capacity constant at the value of the antecedent control volume for the
respective control volume. Further information can be found in [112].
Figure 5-14 shows the comparison of results of the developed model for
the co-flow pattern considering both IEA fuel gases, Table 5-11.
Figure 5-14:
Comparison of co- Planar co-flow model comparison for IEA benchmark 1 input
3500
tions IEA 1 and 3000
IEA 2 2500
2000
1500
1000
500
0
[mV] [10E-2 W] [A/m2] [A/m2] [K] [K] [K/m] [K] [K]
Voltage Power Min. Max. Min. Solid Max. Max. Air outlet Fuel outlet
Current Current temp. Solid Temp. temp. temp.
density density temp. gradient
4000
3500
Corresponding units
3000
2500
2000
1500
1000
500
0
[mV] [10E-2 W] [A/m2] [A/m2] [K] [K] [K/m] [K] [K]
Voltage Power Min. Max. Min. Solid Max. Max. Air outlet Fuel outlet
Current Current temp. Solid Temp. temp. temp.
density density temp. gradient
The terms "Minimum value" and "Maximum value" refer to the minimum
and maximum values among all the BMT participants, whereas "Own"
refers to the results obtained from the author's model.
Figure 5-15 shows the comparison of results for the counter-flow pattern.
It can be seen, that the results of the developed model lie mostly be-
tween the minimum and maximum values of the BMT. Exceptions are
found for the maximum current density (Figure 5-14, fuel gas IEA 1) and
the maximum solid temperature gradient (Figure 5-14 and Figure 5-15,
fuel gas IEA 2).
184
5 Modeling
Figure 5-15:
Comparison of Planar counter-flow model comparison for IEA benchmark 1 input
counter-flow 10000
model results with 9000
BMT results for 8000
Corresponding units
fuel compositions 7000
Voltage Power Min. Max. Min. Solid Max. Max. Air outlet Fuel outlet
Current Current temp. Solid Temp. temp. temp.
density density temp. gradient
12000
Corresponding units
10000
8000
6000
4000
2000
0
[mV] [10E-2 W] [A/m2] [A/m2] [K] [K] [K/m] [K] [K]
Voltage Power Min. Max. Min. Solid Max. Max. Air outlet Fuel outlet
Current Current temp. Solid Temp. temp. temp.
density density temp. gradient
The local temperature curves obtained from the author's model were
compared with those of the BMT activity leader, Achenbach, [112], to
find explanations for the discrepancies, Figure 5-16.
pattern computed
by the Achenbach 1250
model, [112], and
the developed
model with 20 and
100 discretization 1200
points
1150
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell length [m]
185
5 Modeling
It was found that different numbers of control volumes were used for the
discretization of the considered planar cell. Figure 5-16 shows the im-
pact of the number of discretization points on the computed temperature
profile. The temperature gradients at the channel ends are determined
by the considered heat losses. Coarse discretization leads to an overes-
timated impact of the cell end heat losses on the temperature profile.
However, the general trends were not affected by the coarse discretiza-
tion which has the advantage of short calculation times around 30 sec-
onds run on a conventional personal computer. In summary it can be
stated that the author's model behaves physically correct and that there
is no systematic discrepancy with the results of other participants in the
BMT. In all the simulations, the error in the mass balance was below
0.5 % and in the energy balance below 2 %. Compared to the state-of-
the-art with 1 % error in the mass balance and up to 6 % error in the en-
ergy balance, [114], the author's model is considered a reliable tool for
the simulation of planar SOFCs.
186
5 Modeling
The aim of validating the developed model considering the tubular cell
geometry is to ensure that the generalized approach, which is based on
characteristic lengths and areas, has been properly adapted to the tubu-
lar cell design. Although the most important parts of the model code
were already confirmed using the BMT for the planar geometry, there is
still a need to also verify the equations involved in the calculation of the
activation and diffusion losses given these were excluded from the BMT.
The experimental voltage-current curves used for the validation are pub-
lished in [70]. The characterized standard tubular Siemens AG cell was
operated with a fuel gas mixture consisting of 89 vol.-% H2 and 11 vol.-
% H2O at three different cell temperatures, 900 °C, 940 °C and 1000 °C.
The latter feature predestines this data set for model validation, as it al-
lows the model to be checked regarding the correct prediction of tem-
perature effects. Model results obtained for other temperatures can be
assumed valid, in case the model reliably predicts the temperature in-
duced effects of the validation case. When model validation is carried
out with data sets measured at a single temperature, the results are only
valid for this temperature. Hence extrapolated results have to be con-
sidered less reliable.
The voltage-current curves were experimentally determined with con-
stant fuel utilization of 85 % and an air-to-fuel ratio of 4. For the simula-
tion, the model further requires as operational input data the gas inlet
temperatures, system pressure and temperature of the surroundings.
The gas inlet temperatures were assumed 100 K lower than the respec-
tive mean cell temperature, [155]. As the author of [70] stated that the
measurements were carried out using a prototypic cell, it was assumed
that the cell was operated in an oven exhibiting a constant wall tempera-
ture. The surroundings temperature required for the radiational heat
transfer calculation was assumed equal to the mean cell temperature,
[155]. The heat transfer calculation input data is given in Table 5-17.
187
5 Modeling
188
5 Modeling
Siemens AG cell
measured at a mean
Voltage [V]
Variance
cell temperature of 0.60 0%
900 °C
0.50 -3%
0.40 -6%
750 1250 1750 2250 2750 3250 3750 4250 4750 5250 5750
2
Current density [A/m ]
Measurement Developed model Variance
Figure 5-17 shows the comparison of simulated (dashed line) and meas-
ured (full line) voltage-current density (UI) characteristics of the tubular
cell operated at 900 °C. The slightly higher slope of the simulated UI
curve indicates an overestimation of the ohmic resistance and an un-
derestimation of the activation polarization at low current density values.
The mean variance of the predicted voltage for 900 °C operational
mean cell temperature compared to the measured values is 1.6 % and
the maximum variance is below +/-3 % (right axis of graph).
Figure 5-18 shows, the simulated and measured U-I curve for 940 °C
mean cell temperature.
Variance
cell temperature of
0.60 0%
940 °C
0.50 -3%
0.40 -6%
750 1250 1750 2250 2750 3250 3750 4250 4750 5250 5750
2
Current density [A/m ]
Measurement Developed model Variance
189
5 Modeling
It can be seen, that the predicted voltage values diverge by less than
1 % for small and medium current density values. At higher current den-
sity values the variance increases up to 5 %, indicating an underestima-
tion of the diffusion polarization losses. However, the activation and oh-
mic losses are predicted well resulting in a good match of the UI curve
slope of the simulated and measured curve. The mean variance be-
tween predicted and measured voltage values for 940 °C mean cell
temperature is 1.3 % and the maximum variance is below +5 %.Figure
5-19 depicts the comparison of simulated and measured voltage-current
curve measured at a mean cell temperature of 1000 °C.
Siemens AG cell
measured at a mean
Voltage [V]
Variance
cell temperature of 0.60 0%
1000 °C
0.50 -3%
0.40 -6%
750 1250 1750 2250 2750 3250 3750 4250 4750 5250 5750
Current density [A/m2]
Measurement Developed model Variance
Similar to the predicted UI curve for 900 °C, the model underestimates
the activation losses at low current densities for 1000 °C mean cell tem-
perature. Further, the slope of the predicted curve is again steeper than
of the measured UI curve indicating a slight overestimation of the ohmic
losses. At high current density values, the model predicts a slight bend
of the UI curve due to increasing diffusion losses. This phenomenon can
be observed more pronounced for the measurement. Nevertheless, the
prediction of this phenomenon clearly shows that the developed tubular
cell model behaves physically correct. The mean variance between pre-
dicted and measured voltage values for 1000 °C mean cell temperature
is 2.3 % and the maximum variance is below +/-4 %.
In summary, the mean variance of the model results from the measure-
ment results was found to be below 2.5 % for all the three considered
mean cell temperatures. The developed tubular cell model behaves
190
5 Modeling
It is important to point out, that the geometry of the air delivery tube was
not taken from a publication but was assumed to yield an adequate
cathode gas channel cross-sectional area. This assumption directly im-
pacts the convective heat transfer between the cathode gas and the air
delivery tube. The convective heat transfer between the cathode gas
and the solid structure however dominates in any case, such that the
overall impact of this assumption is considered less significant.
The experimental voltage-current curve used for the validation of the tri-
angular cell model was measured at the Siemens AG using a D8 cell
segment, [210]. The 1 cm long D8 cell segment was operated under iso-
thermal conditions in an oven at temperatures between 950 and 965 °C
191
5 Modeling
with fuel and air excess. The fuel mixture consisted of 50 vol.-% H2 and
50 vol.-% H2O, being the expected average gas composition of a full-
scale cell operated at 85 % fuel utilization with 89 vol.-% H2 and 11 vol.-
% H2O. Due to the isothermal conditions and the virtually constant reac-
tant partial pressures along the D8 cell segment, the voltage-current
curve was simulated by using the electrochemical performance model
without consideration to the energy and heat balance. With respect to
the model parameters, it was assumed that the anode and cathode ma-
terials used for the D8 cells resemble those used for the standard tubu-
lar cells. Accordingly, the model parameters given in Table 5-18 were
used for the simulation. Exceptions were the electric conductivity of the
interconnect material, which was calculated according to Eq. 48, and the
ionic conductivity of the electrolyte, as D8 cells employ scandia-
stabilized zirconia (ScSZ) instead of the state-of-the-art yttrium-
stabilized zirconia (YSZ) as electrolyte material, [210]. At temperatures
above 900 °C, ScSZ features ionic conductivities about three times
higher than YSZ, [79]. Hence, the conductivity values computed via Eq.
46 were multiplied by three for the D8 cell segment simulation.
of a D8 cell segment
measured at a tem- 0.70
perature between
Voltage [V]
0.50
0.40
0.30
0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000
2
Current density [A/m ]
Figure 5-20 depicts the comparison of the measured and the simulated
voltage-current curve of the D8 cell segment.
The simulation result for 957 °C fits the measured data well. It is con-
cluded, that the author's model was successfully applied to the triangu-
lar geometry thereby validating the discussed assumptions.
192
5 Modeling
193
5 Modeling
The cell geometry is planar and equal to that of the IEA BMT, Table
5-12. The considered flow configuration is co-flow. Operational condi-
tions for the sensitivity analysis are given in Table 5-13. The material
property parameters used in the energy balance calculations are given
in Table 5-14. Table 5-21 summarizes the integral model response val-
ues for the reference case and both considered fuel gases.
Table 5-21: Response value Unit Fuel gas IEA 1 Fuel gas IEA 2
Model response K/mm
Maximum solid temperature gradient 2.2 2.0
values for the
reference case Maximum solid temperature 1327.9 1286.9
for IEA 1 and IEA Minimum solid temperature K 1212.0 1146.3
2 gas
Mean solid temperature 1282.7 1221.9
Operational voltage for 85 % fuel utilization V 0.683 0.624
194
5 Modeling
Figure 5-21 depicts the predicted temperature profiles of the solid struc-
ture for the reference case and both considered fuel gases.
1250 60
[K]
1200 40
1150 20
1100 0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
Temperature IEA 1 Temperature IEA 2 Temperature difference
It can be seen, that at the cell inlet (x-axis: cell length equals zero) the
temperature predicted for the fuel gas IEA 1 is approx. 65 K higher than
for the fuel gas IEA 2. That can be explained with the endothermal
steam reforming reactions taking place due to the methane containing
fuel gas IEA 2. At the cell outlet (x-axis: cell length equals 0.1 m), the
difference between the two temperature distributions is decreased to
approx. 40 K. The difference between the temperature profiles of the
two fuel gases is represented by the curve named "Temperature differ-
ence". This curve is considered an indicator for the thermal prediction
behavior of the model towards changes of the fuel composition.
195
5 Modeling
6000
energy for IEA 1
and IEA 2 gas 5000
4000
3000
2000
1000
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
IEA 1 -50% IEA1 ref. IEA 1 +50% IEA 2 -50% IEA 2 ref. IEA 2 +50%
For the IEA 1 gas, the maximum current is produced at the cell inlet.
This is due to the high hydrogen partial pressure of the IEA 1 gas. From
thereon the current density constantly decreases along the cell length
because of the ongoing fuel conversion and accordingly decreased
educt partial pressures. In contrast, the IEA 2 gas yields the maximum
educt partial pressures in the region where the methane steam reform-
ing reactions are complete. Consequently, the current density is not
constantly decreasing along the cell length, but is somewhat evenly dis-
tributed with a wide and flat peak.
Decreasing the anode activation energy means to increase the activity
towards electrochemical reactions yielding higher current density values
in the cell inlet region for both gases. The increased fuel consumption in
the cell inlet region causes a considerable current density decrease at
the cell outlet for the IEA 1 gas. However, for the IEA 2 gas the current
density at the cell outlet remains constant, as the fuel consumption in
the cell inlet region is dominated by the hydrogen production through
steam reforming reactions. This leaves more hydrogen over for the latter
part of the cell. In short, the less hydrogen is produced cell internally
through steam reforming, the more does a decrease of the activation
energy yield an increased current density at the cell inlet which accom-
panies a current density decrease at the outlet.
196
5 Modeling
Table 5-22 shows that for both fuel gases, the increase of the anode ac-
tivation energy yields an increase of the temperature extremes, the
maximum temperature gradient and the mean cell temperature.
This is the result of the constant fuel utilization, which requires higher
temperatures due to the lower activity towards electrochemical reactions
when the activation energy value is increased.
The operational voltage to meet the defined fuel utilization is strongly
decreased for higher activation energies while it is only little increased
for lower values, Figure 5-23.
energy values
obtained for IEA 1 -40%
and IEA 2 gas
-60%
-80%
-100%
Percentage of reference value
IEA 1 gas IEA 2 gas
197
5 Modeling
It is important to point out that SOFCs are usually not operated at volt-
ages below 0.5 V. The predicted voltage values for increased activation
energies indicate that in reality a fuel cell employing such an anode
catalyst can not reach 85 % fuel utilization. For the lower activation en-
ergy values, the power output increases by a maximum of 15 % and de-
creases by 80 % for the higher values.
80
Difference between temperature
temperature dis-
tributions for vary-
ing anode activa- 60
tion energy values
obtained for IEA 1
and IEA 2 gas 40
20
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
Figure 5-24 depicts the sensitivity of the difference between the tem-
perature profiles of the two fuel gases towards the anode activation en-
ergy. The local temperatures are strongly influenced by the locally pro-
duced power, which depends on the prevalent current density and the
operational voltage, see also Table 5-22. For a fixed operational voltage,
high current densities usually lead to high temperatures. However, en-
dothermal reactions such as steam reforming counteract this, provided
that the reaction rate is considerable.
It can be seen, that higher values of the activation energy yield an in-
crease of the difference between the predicted temperatures near the
cell inlet and a higher reforming peak. This can be traced back to the
fact that the temperature at the cell inlet for the fuel gas IEA 2 is domi-
nated by the steam reforming reaction and does not change significantly
by varying the activation energy. In contrast, the temperature at the cell
inlet increases significantly with increasing activation energy for the IEA
1 gas. The reason for this is that the operational voltage was strongly
reduced in order to meet the required overall fuel utilization. As a con-
198
5 Modeling
sequence, the amount of produced power is low even though the current
density is only slightly decreased with increasing activation energy. As
less electrical power is produced from a still considerable amount of
converted fuel, more sensible heat is released from the electrochemical
reactions resulting in higher local temperatures.
Lowering the anode activation energy causes an increase of the tem-
perature difference towards the cell end. Similar to the above discussed
increase of the activation energy, the cell inlet temperature of the IEA 2
gas remains almost equal to the reference case. However with low acti-
vation energy values, the high amount of produced power plus the heat
consumed by the endothermal steam reforming leads to a lower tem-
perature for the IEA 2 gas at the cell outlet compared to the reference
case. For the IEA 1 gas the missing steam reforming cooling is compen-
sated to a smaller extent by low current densities at the cell outlet due to
fuel depletion. Hence, the temperature difference between the two
gases is increased at the cell end.
Figure 5-25 depicts the sensitivity of the current density distribution to-
wards the cathode activation energy. Note that the cathode activation
energy was not decreased by more than 20 % as the associated high
electrochemical activity at the cell inlet for both fuel gases caused sub-
stantial fuel depletion effects at the cell outlet, which resulted in numeri-
cal instability of the employed solver.
4000
3000
2000
1000
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
IEA 1 -20% IEA1 ref. IEA 1 +50% IEA 2 -20% IEA 2 ref. IEA 2 +50%
For the reference case the anode activation losses are slightly higher
than the cathode activation losses. By decreasing the cathode activation
199
5 Modeling
200
5 Modeling
energy values
obtained for IEA 1 -40%
and IEA 2 gas
-60%
-80%
-100%
Percentage of reference value
IEA 1 gas IEA 2 gas
201
5 Modeling
60
vation energy
values for IEA 1
and IEA 2 gas 40
20
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
Figure 5-27 shows the sensitivity of the difference between the tempera-
ture profiles of the two fuel gases towards the cathode activation energy.
Comparing Figure 5-27 and Figure 5-24 one could pre-maturely con-
clude that the cathode activation energy has a weaker impact than the
anode activation energy values on thermal behavior of the fuel cell for
the two investigated fuel gases. However, the current density distribu-
tions for different cathode activation energy values reveal a different pic-
ture.
The strong reforming peak, which was discussed for higher anode acti-
vation energies, cannot be observed for the increased cathode activa-
tion energy values. Similar to the anode activation energy variation, the
temperature at the cell inlet considering the IEA 2 gas is dominated by
the occurring steam reforming reactions and does barely change with
varying cathode activation values.
Considering the IEA 1 gas with higher cathode activation energy, the
temperature at the cell inlet is slightly lower then in the reference case.
The reason for that is that the produced current in this region is strongly
202
5 Modeling
reduced compared to the reference case due to the low oxygen partial
pressure at the cathode compared to the initial hydrogen partial pres-
sure of the IEA 1 gas. However, a strong temperature decline is inhib-
ited by the low operational voltage, which results in a tangible amount of
released sensible heat sufficient to maintain the inlet temperature on a
level close to the reference case. Hence, the temperature difference
found at the cell inlet for increased cathode activation energy values is
slightly lower than for the reference case, which in turn explains the
missing steam reforming peak.
Towards the cell end, the increased cathode activation causes higher
current density values due to less fuel depletion in the front parts of the
cell. This can be observed for both fuel gases. However, the tempera-
ture increase towards the cell end for the IEA 2 gas is stronger due to
the previously discussed current density peak near the cell end. This
explains the slightly lower temperature difference between the two in-
vestigated fuel gases at the cell end for higher cathode activation energy
values as compared to the reference case.
As discussed for the current density distribution, decreasing the cathode
activation energy puts the anode polarization into a dominant position.
As the anode polarization losses are computed using the reference val-
ues, the temperature difference progression remains almost unchanged
by the lowered cathode activation energy.
In conclusion it can be said, that the cathode and anode activation po-
larization energy values are model parameters which have to be han-
dled carefully. The thermal prediction behavior, the predicted current
density distribution and power output are all strongly influenced by these
model parameters. In particular, the impact of the anode and cathode
activation energy values has in each case a different nature. Thus, it is
important to know if anode or cathode activation polarization is dominant.
Dominating anode activation emphasizes the impact of the educt partial
pressure distribution leading to high fuel conversion where educt partial
pressures are high and vice versa. In contrast, high cathode activation
losses yield a relatively even fuel consumption along the cell due to the
almost constant oxygen partial pressures at the cathode resulting from
the usually low air utilization. Regarding the local temperatures, large
activation losses yield a higher amount of released sensible heat par-
tially compensating the lower amount of converted fuel and vice versa.
Unfortunately, these general rules are distorted in case internal reform-
203
5 Modeling
ing reactions take place. These reactions dominate the local tempera-
ture in the region where they occur.
5%
put for varying
anode pre-
exponential factor
values for IEA 1 0%
and IEA 2 gas 50% 75% 100% 125% 150%
-5%
-10%
Percentage of reference value
For lower anode pre-exponential factor values, the power output de-
creases by a maximum of 8 %, while it increases by a maximum of 4 %
for higher anode pre-exponential factor values. Similar to the variation of
the activation energy values, the IEA 2 gas shows a more pronounced
response. The reasons for which were discussed above.
The minor changes of the current density distribution appear for both
gases to almost equal rates. Further, as shown in Figure 5-28, the an-
204
5 Modeling
80
Difference between temperature
temperature dis-
tributions for vary-
ing anode activa- 60
tion pre-
exponential factor
values for IEA 1 40
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
5%
put for varying
cathode pre-
exponential factor
values for IEA 1 0%
and IEA 2 gas 50% 75% 100% 125% 150%
-5%
-10%
Percentage of reference value
IEA 1 gas IEA 2 gas
205
5 Modeling
In contrast, the opposite is the case for the impact on the predicted
power, Figure 5-30. For lower cathode pre-exponential factor values, the
power output decreases by a maximum of 5 %, while it increases by a
maximum of 2 % for higher cathode pre-exponential factor values. The
reasons for that are the same as already discussed for the cathode acti-
vation energy variation. Higher cathode activation pre-exponential fac-
tors put the anode activation into the dominant role, while lower values
lead to a more homogeneous current density distribution, which equal-
izes fuel depletion losses at the cell ends.
Figure 5-31 depicts the low sensitivity of the temperature distribution
trend towards the cathode pre-exponential factor.
80
Difference between temperature
temperature dis-
tributions for vary-
ing cathode acti- 60
vation pre-
exponential factor
values for IEA 1
40
and IEA 2 gas
20
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
ther depend on the prevalent temperature, the diffusion path length and
prevalent current density. The latter is causing the concentration gradi-
ent between the triple phase boundary (TPB) where the electrochemical
reactions take place and the bulk gas phase. The diffusion path length
has a direct impact on the concentration at the end of the diffusion path.
In particular, the educt partial pressure at the TPB decreases with in-
creasing diffusion path length. Hence, with respect to diffusion losses,
the investigated type of support design is of major importance. Cathode
supported cells, such as are the tubular cells of the Siemens AG, feature
a stronger cathodic diffusion limitation then e.g. electrolyte supported
cells, because of their approx. 20 times longer diffusion path through the
cathode. At low and intermediate current densities, diffusion losses
however are usually small or even negligible. Nevertheless, this sensitiv-
ity study was carried out considering an electrolyte-supported planar cell
in order to investigate whether the according material properties could
lead to a change of these coherences.
The variation of the pore diameter was omitted as it merely re-enforces
the impact of Knudsen for smaller pore diameter values or weakens it
for larger pore diameters. Instead, the material parameters influencing
both diffusion mechanisms, namely the porosity and tortuosity, were var-
ied. For the reference case the quotient of the interdependent porosity
and tortuosity has the value 0.166 assuming a tortuosity of 3 and a po-
rosity of 0.5. It was decreased by 70 % to 0.05, which corresponds to a
porosity of 0.25 and a tortuosity of 5 (typical values of "dense mem-
branes"), and increased by 110 % to 0.35, which corresponds to a po-
rosity of 0.7 and a tortuosity of 2, in order to cover the entire range of
technically possible values.
The current density distribution for varied anode material parameters
remained unchanged. Therefore, no discussion on this aspect is under-
taken. In contrast, Figure 5-32 depicts the sensitivity of the current den-
sity distribution towards the cathode material parameters. Increasing the
parameter quotient does not produce any effect. This can be explained
by the magnitude of the diffusion losses for the investigated operational
conditions, which is about two orders of magnitude lower than ohmic
and activation losses. However, the current density progression com-
puted with the dense membrane values shows a considerable decrease
of the current density at the cell inlet for the IEA 1 gas. This indicates
that in this region, the reached current density is close to the diffusion
limited current density of the cathode electrode. In the latter cell parts
207
5 Modeling
3000
2000
1000
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
IEA 1 -70% IEA 1 ref. IEA 1 +110% IEA 2 -70% IEA 2 ref. IEA 2 +110%
For the increased porosity and decreased tortuosity, the power output
remains constant. In contrast, assuming dense membrane properties for
the anode yields a decreased power output of 0.5 %. Decreasing the
cathode material properties quotient results in a 2.5 % decreased power
output. A cause for this somewhat stronger response can be found in
the smaller effective diffusion coefficient of oxygen as compared to hy-
drogen.
80
Difference between temperature
temperature dis-
tributions for vary-
ing cathode diffu- 60
sion material pa-
rameter for IEA 1
and IEA 2 gas 40
20
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
208
5 Modeling
209
5 Modeling
80
Difference between temperature
temperature dis-
tributions for dif-
ferent Nusselt 60
numbers for IEA 1
and IEA 2 gas
40
20
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
3000
2000
1000
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
IEA 1 -90% IEA 1 ref. IEA 1 +900% IEA 2 -90% IEA 2 ref. IEA 2 +900%
Higher λs values yield a strong increase of the current density at the cell
inlet and corresponding decrease at the cell outlet. In contrast, lower λs
values show a minor impact on the predicted current density distribution,
the power output and the thermal prediction behavior of the model con-
sidering different fuel gases. These results are sensible, given that the
reference case value of λs is already small. Lowering this value further
reduces the already humble conductive heat transfer to values close to
zero. The remainder of this discussion hence focuses on the impact of
increased λs values, which induce stronger changes to the energy bal-
ance of the cell.
211
5 Modeling
Table 5-24 gives the cell temperatures and operational voltage values
for varying λs values. Comparing the cell inlet temperature of both fuel
gases for the reference case and the increased solid heat conductivity it
was found that the IEA 2 gas features a stronger increase by approx.
5 K. In detail, the cell inlet temperature for the IEA 1 gas increased by
72 K versus 77 K for the IEA 2 gas, Figure 5-36.
1200
1150
1100
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
IEA 1 -90% IEA 1 ref. IEA 1 +900% IEA 2 -90% IEA 2 ref. IEA 2 +900%
Similar effects were observable for the cell outlet temperatures which in-
creased by 18 K for the IEA 1 gas and 20 K for the IEA 2 gas. Summed
up these temperature differences result in an increased mean cell tem-
peratures by 13 K for the IEA 2 gas (54 K increase) as compared to the
IEA 1 gas (41 K increase). The differently strong pronounced tempera-
212
5 Modeling
0.0%
10% 100% 1000%
-2.5%
Percentage of reference value
IEA 1 gas IEA 2 gas
80
temperature dis-
Difference between temperature
20
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
Therefore it can be said that the Nusselt number has almost no impact
on the model results which justifies the according assumption. The heat
conductivity coefficient strongly affects the predicted current density dis-
tribution and local temperature distribution. The current density tends to
be less uniform for higher heat conductivity values whereas the opposite
is the case for the temperature distribution. Further for high solid heat
conduction coefficients, the dominating role of the steam reforming reac-
tions occurring at the cell inlet is weakened. Hence, the differences for
the local temperatures between hydrocarbon containing gases and pure
syngases are reduced. A considerable impact on the predicted power
output was found despite the fact that the model parameters used for
the electrochemical performance prediction were held constant at their
reference value. However, the solid heat conduction is a material pa-
rameter which can be reliably determined. Therefore is it not likely to in-
troduce an uncontrolled uncertainty to the model results.
214
5 Modeling
215
5 Modeling
20%
Carbon monoxide
10%
Methane
0%
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
Figure 5-40 shows the predicted current density and Nernst voltage dis-
tribution for the different STR kinetics models. For faster STR kinetics,
the current density at the cell inlet increases yielding slightly decreased
current density values at the cell outlet due to fuel depletion effects.
3000 0.950
2000 0.925
1000 0.900
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
The reason for the higher current density values at the cell inlet ob-
served for faster STR kinetics can be found in the predicted Nernst volt-
age within the STR region. From Eq. 80 and Eq. 81 it can be derived
that higher hydrogen and carbon monoxide partial pressures yield
higher Nernst voltages. Further, the Nernst voltage increases with de-
creasing temperatures due to the endothermal STR reactions.
216
5 Modeling
Figure 5-41 and Table 5-25 show that the cooling effect in the STR re-
gion is more pronounced for faster STR kinetics as the locally higher
heat requirements cannot be fully covered via heat conduction and con-
vection. However, towards the cell outlet, fast STR kinetics yield higher
temperatures than the slow kinetics. Overall, faster STR kinetics yield
higher average cell temperatures.
1200
1150
1100
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
Table 5-25 gives the predicted maximum temperature gradients and op-
erational voltages to maintain 85 % fuel utilization.
217
5 Modeling
to less activation and ohmic voltage losses. However, the gained power
output increase is in the order of 1 % for the equilibrium model as com-
pared to the reference case considering the Achenbach model.
Figure 5-42 depicts the temperature difference obtained for the different
STR kinetics models. The faster the STR reactions occur, the flatter and
the nearer to the cell inlet is the STR peak.
80
Difference between temperature
temperature dis-
tributions for dif-
ferent STR kinetic 60
models obtained
for IEA 1 and IEA
2 gas 40
20
0
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10
Cell lenth [m]
218
5 Modeling
ergy values, the solid heat conduction coefficient and the applied kinet-
ics of the steam reforming showed a considerable impact. However, the
impact of all variations in model parameters on the predicted tempera-
ture distribution trends remained always smaller than the impact of the
different fuel gases themselves. In this respect, the author's model has
shown a fair robustness. The predicted trends for the power output
based on different fuel gases were also hardly affected by the variations
in model parameters. With respect to the absolute value of the power
output, only the activation polarization activation energy values were
found to have a strong impact. All other investigated model parameters,
including the applied kinetics of the steam reforming reactions, had less
then 10 % variance to the reference case within the variation bandwidth.
219
5 Modeling
of different cost
Order of
estimation classes, magnitude
derived from [213] estimate
Definitive
10 estimate
Detailed estimate
0
1 2 5 10 20 50 100 200 500 1000
Cost of estimate [1000 US$]
100000
functions of Garrett
(lines without sym- Shell & tube
bols) and Turton Spiral
(lines with symbols)
10000
Double pipe
1000
1 10 100 1000 10000
Heat transfer area [m2]
220
5 Modeling
The Turton functions yield cost estimates with values on average twice
as high as the Garrett functions. Employing Eq. 176, the renown
"Chemical Engineering magazine Plant Cost Index", CEPCI, was
adopted to adjust the Garrett cost estimates based on data from 1987 to
2001, the year in which the data for the Turton functions was collected.
CEPCI present
C present = ⋅ Creference Eq. 176
CEPCI reference
The CEPCI is a lumped inflation cost indicator, which represents the av-
erage price inflation of all major equipment pieces used in the chemical
industry. Figure 5-45 shows a plot of the CEPCI from 1950 through 2007.
300
200
100
0
1950 1960 1970 1980 1990 2000 2010
Year [-]
It is likely that technology changes and unequal price inflations for dif-
ferent equipment pieces lead to considerable differences between the
estimated costs obtained from functions developed at various times.
However, keeping the general stray area of cost estimates compared to
vendor quotations in mind, the functions of both authors appear to apply
well to the comparisons of different system designs with respect to their
economical performance. The choice of function is thus a matter of taste
rather than of accuracy. The Turton functions were used in e.g. [212],
while Garretts functions were employed in e.g. [215].
As Garrett offers more specified cost functions than Turton, the cost
functions formulated by Garrett were applied in this work. Being formu-
lated in 1987, the reference CEPCI is 320. The cost estimates were up-
dated to the year 2007 with a CEPCI of 530. Further, the estimated
equipment prizes were converted from US$ to €, assuming a currency
rate of 1.4 US$ per € of the year 2007.
221
5 Modeling
223
5 Modeling
224
5 Modeling
The cost of the spray chamber is determined by its water flow rate, see
section 5.2.2. As the steam generator is an integral part of the heat inte-
gration network discussed in section 5.2.3, its area is known from the
flowsheeting simulation.
State-of-the-art producer gas cleaning systems operate at low tempera-
tures and usually consist of a particle removal unit and a wet scrubber.
The wash liquid, mostly water, has to be treated after its use. The
"Spray chamber" function was used for the prize estimation of wet
scrubbers. The considerably higher cost estimates account for the wash
liquid treatment. The wet scrubbers are sized according to [28], where it
is reported that approx. 2 l of water are spent to clean 1 m3n of wet and
tar-laden producer gas.
For the removal of hydrogen sulfide from producer gases, solely zinc ox-
ide adsorbers were considered using the "Activated carbon adsorber"
cost function, [214]. It was assumed that the adsorber should allow for
one month non-stop operation. The molar amount of hydrogen sulfide
which is adsorbed during this period is known from the flowsheeting
simulations, see section 5.2.6. Assuming a maximum sulfur capacity of
zinc oxide of 10 %, [149], the adsorbent mass and the cost is computed.
The adiabatic methanation of producer gas, see section 5.2.4, is carried
out in a fluidized bed reactor. The cost function "Fluid bed" is based on
the reactor volume, [214], which is sized to yield similar operating condi-
tions as applied in [40].
The catalytic partial oxidation of producer gas is conducted using coated
monoliths. Based on the experiments reported in section 3.5, a tolerable
catalyst load of 2.5 kW LHV producer gas power per monolith with 1" di-
ameter, 3" length and 400 cpsi was assumed to estimate the required
number of monoliths and the CPO reactor volume. The CPO reactor is
pricewise rated employing the "Jacketed reactor" cost function, [214]. Eq.
178 is used for the power requirement calculation of the CPO air blower.
The air flow rate is determined for an air-to-fuel ratio of 0.12.
225
5 Modeling
The fuel cell balance of plant module contains all the heat exchangers
required to adjust the producer gas and the cooling air temperature to
the predefined SOFC inlet temperatures, the cooling air blower, the
DC/AC power converter and the SOFC off-gas combustor. The heat ex-
change areas of the air pre- and super heater, the producer gas cooler
(if required) and super heater are known from the flowsheeting simula-
tion, see section 5.2.3. The costs of all heat exchangers are determined
using the "Plate and frame" cost function, except for the air pre-heater
the cost of which is estimated based on the "Shell and tube" function,
[214], see Figure 5-44.
Similar to all other air blowers, Eq. 178 is used for the power require-
ment calculation of the fuel cell cooling air blower. The air flow rate is
determined by the ATHENA SOFC model.
In 2003, the American Department of Energy together with the Depart-
ment of Defense has launched a design competition for DC/AC power
converters tailored to the application with SOFCs. The cost target of less
than 40 $ per kWel, [217], was adopted in this work.
The oxidation of the depleted SOFC fuel gas is performed in a catalytic
combustor to ensure complete conversion. Its heat duty is known from
the flowsheeting calculations, see section 5.2.7, and used for the cost
estimation based on the "Catalytic incinerator" cost function, [214].
The fuel cell module comprises the SOFC stack and all catalyst, which
have to be exchanged in regular intervals. For 2012, the American Solid
State Energy Conversion Alliance (SECA) has defined a cost target for
SOFC systems of 400 2002-$ per kWel output using natural gas as fuel,
[59]. Despite this objective was formulated for SOFC systems, it was
assumed for the stack cost to be more conservative in this work. In sys-
tems using an adiabatic methanation, the catalyst cost are estimated
based on a price of 50000 2007-€ per ton corresponding to the average
nickel price in 2007. The catalyst mass is determined to yield operating
conditions as applied in [40]. For systems using a CPO, 100 € was as-
sumed as piece prize for a 1" CPO catalyst.
The auxiliary equipment module summarizes all equipment units
which are not directly related with the electricity generation. These are
the stack, the emergency flare and the heat exchanger employed to ex-
tract useful heat from the SOFC flue gas.
The stack is assumed by default 20 m height and pricewise rated using
the "Short chimney" cost function, [214].
226
5 Modeling
227
5 Modeling
The gas processing efficiency, ηGP, relates the chemical energy con-
tent of the processed gas at the SOFC inlet (CG) to that of the raw pro-
ducer gas (PG), Eq. 181. It thus describes chemical energy losses oc-
curring in the gas processing system.
V& ⋅ LHV
ηGP = &CG CG
Eq. 181
VPG ⋅ LHVPG
The SOFC DC efficiency, ηSOFC,DC, includes the fuel utilization and the
electrochemical conversion efficiency of the SOFC. It is defined as the
ratio between the produced DC power and the chemical energy content
of the processed gas at the SOFC inlet (CG), Eq. 182
P
η SOFC, DC = & SOFC , DC Eq. 182
VCG ⋅ LHVCG
⎛ yS ⎞ ⎛ yS ⎞
xS ,loss = 0.52695 + 0.20716 ⋅ exp⎜ − ⎟ + 0.26521 ⋅ exp⎜ − ⎟ Eq. 183
⎝ 3.7608 ⎠ ⎝ 27.0748 ⎠
yS represents the molar fraction of the total sulfur content in the fuel gas
in parts per millions. Note that the loss of electrical power due to sulfur is
added to the heat output of the SOFC.
The auxiliary power requirements, PAux, originating from air blowers
for e.g. the SOFC cooling and pumps for e.g. producer gas humidifica-
tion, were calculated according to Eq. 178. As no momentum balance
was performed in the thermodynamical simulations, a counter pressure,
Δpsys, of 375 mbar and a blower or pump efficiency, ηblow, of 75 % was
generally assumed for the auxiliary power requirement estimation.
The inverter efficiency, ηInv, for the conversion of the DC produced by
the SOFC to AC deliverable to the grid was assumed to be 92.5 %.
Figure 5-46:
Total investment Wood price szenario Fixed operation cost
Structure of the
Dept portion Plant factor Insurance
power production
Intrest rate Wood consumption Operation
cost calculation, maintenance
Construction period
derived from [218] Utilities
Depreciation period
Variable cost
229
5 Modeling
The starting point for the calculation is a detailed cost analysis on either
a yearly or monthly basis. Most of the regular and customary recurring
costs are assumed subject to 2 % inflation, IF. This includes costs for
operation and maintenance, feedstock, utilities and replacement of
worn-out equipment. Following cost factors were considered:
Fuel cost
The fuel cost for the operation year i, Cfuel,i, is computed via Eq. 184.
C fuel ,i = m& wood ⋅ 8760 ⋅ PF ⋅ Cwood ⋅ (1 + IF + WI )
(i + SoP − 2007 )
Eq. 184
The wood mass flow per hour, m& wood , is known from the flowsheeting
simulation. The plant factor, PF, describes the availability of the plant
and is assumed to be 85 %. The wood price, Cwood, of 84 2007-€ per
dry ton in the year 2007 is in accordance with the German Coordinat-
ing Office for Renewable Raw Materials, [219]. To account for the in-
creasing future demand of wood, a market-driven wood price in-
crease, WI, of 1 % per annum in addition to regular inflation, IF, was
added.
Insurance cost
For the insurance costs, a fixed rate per annum of 1 % of the total in-
vestment was assumed over the entire planning horizon, Eq. 185.
CIns ,i = 0.01 ⋅ TI Eq. 185
Utility cost
Water for the producer gas humidification was considered the only
utility cost factor with a price of 0.94 2007-€ per m3. The required wa-
ter flow rate is known from the flow sheeting simulation.
Cutility ,i = V&water ⋅ 8760 ⋅ PF ⋅ Cwater ⋅ (1 + IF ) Eq. 187
i
230
5 Modeling
Pth stands for the output of useful heat from the B-IGFC system in
question. The heat output increases with proceeding degradation of
the SOFC stack. However, it was assumed that only the heat amount
produced by a new stack is sold.
Depreciation
Linear depreciation was generally assumed.
Standard equipment
The yearly depreciation of the standard equipment, CDep,SE,i, is
calculated according to Eq. 189. Standard equipment comprises
all equipment pieces except the fuel cell stack and any em-
ployed catalysts, the initial cost of which, C FC and Ccat , are0 0
231
5 Modeling
Catalysts
Catalysts are subject to deactivation processes and have to be
exchanged in due course. It was assumed that both the
methanation and the CPO catalyst materials are operable for a
maximum period of 24 months in which they are depreciated,
DePCat. The yearly depreciation of catalysts, CDep,cat,i, is com-
puted according to Eq. 191. Similar to the fuel cell stack, cata-
lyst costs were considered subject to inflation. Hence, Ccat de-
notes the costs of the catalyst in operation.
12
Ccat
CDep , cat ,i = ∑ Eq. 191
m =1 DePcat
Interest expenses
The yearly interest expenses, CIE,i, are due to the debt owing on total
investment, TI, plus the additional cost resulting from fuel cell stack,
CFC, and catalyst exchanges, Ccat. As interest payments in arrear
were assumed, the first depreciation and amortization payments are
effective only from the second year of operation onwards.
⎛ i
⎞
CIE ,i = ⎜ TI + CFC + CCat − ∑ DAi −1 ⎟ ⋅ NCIR
⎝ 1 ⎠ Eq. 193
with DAi = CDep , SE ,i + CDep , FC , i + CDep, cat ,i + C Am, i
The total yearly capital costs are given by Eq. 195. The net value at SoP
of the operational and capital costs accrued over the planning horizon,
232
5 Modeling
Pel depends on the TRD and the DRFC and is determined through a
yearly averaging of the produced electricity during the entire planning
horizon.
233
234
6 System analysis
6 System analysis
The main objective of this system analysis was to assess the technical
and economical feasibility of different B-IGFC systems with an electrical
power output of around 1 MWel based on already or soon available bio-
mass gasification processes, gas processing technologies and SOFC
designs for stationary power applications. Particular attention was given
to the thermal gradients and temperature extremes caused by different
producer gases during their electrochemical conversion in SOFCs. The
interactions between different biomass gasification processes, gas
processing technologies and SOFC concepts were also investigated
and the overall system efficiencies and power output as well as the re-
quirements for the balance of plant equipment resulting from the SOFC
cooling were determined. By relating the system efficiency to system
cost estimates, the power production costs were approximated and used
as basis for the economic comparison of the various B-IGFC systems.
To cover most of the SOFC designs summarized above, the SOFC cell
designs listed below were investigated in the system analysis.
235
6 System analysis
236
6 System analysis
The heat conductivity coefficients of the ceramic materials used for the
simulation of planar cell designs can be found in Table 5-14. For the
heat conduction in the metallic bipolar plates, 18 W/m2 K was assumed.
Table 6-3 summarizes the parameters applied in the reproduction of the
electrochemical behavior of anode-supported cells. The data used for
the reproduction of planar electrolyte-supported cells is summarized in
Table 5-12, Table 5-14 and Table 5-20. Table 5-16 to Table 5-18 give
the geometry and model input data for the tubular cathode-supported
cell design. The general model settings are compiled in Table 6-4.
237
6 System analysis
238
6 System analysis
239
6 System analysis
Pth
ηth = Eq. 200
m& feed ⋅ LHV feed
Pnet , AC + Pth
ηtot = Eq. 201
m& feed ⋅ LHV feed
In a first step, the odorant of the natural gas, assumed as pure methane,
is removed. The gas is then humidified to yield a steam-to-carbon ratio
of 2, which was found sufficient to prevent carbon deposition in corre-
sponding equilibrium calculations, see section 6.3.1. The humidified
natural gas is introduced to an isothermal convective pre-reformer
where, depending on the reformer temperature, 25 to 30 % of the meth-
ane is converted to hydrogen and carbon monoxide, see Table 6-6. The
heat required for both, the steam generation and the endothermic re-
forming reactions, is taken from the hot flue gas. The temperature of the
pre-reformer was assumed 100 K below the operational temperature of
the various cell designs to ensure that enough heat at the required tem-
240
6 System analysis
perature level is available. After being heated to the cell design specific
temperature, the pre-reformed gas is fed to the SOFC, where its chemi-
cal energy content is partially converted to electricity. For heat recovery,
the depleted anode off gases are post-combusted and then conducted
through the heat integration network (HIN) discussed in section 5.2.3.
The electrical efficiency achieved with a fuel utilization of 85 % at a volt-
age of 0.6 V was 45 to 46 % depending on the pre-reforming degree
achieved at the cell design specific reformer temperature.
For the four investigated cell designs, Table 6-7 lists the number of cells
required to produce 1 MWel DC power and the gas inlet temperatures
needed for the accordance of the mean cell temperature and the opera-
tional temperature at an air-to-fuel ratio of 4.25. The established maxi-
mum temperature gradients and solid temperatures were adopted as
upper limit values for the safe operation of the various cell designs.
241
6 System analysis
The water load of the producer gas originating from the updraft gasifica-
tion process was calculated to satisfy saturation at 75 °C, [28]. The gas
yield, which represents the producer gas volume on dry and tar free ba-
sis (dtf) per kg of dry wood, was computed to yield a cold gas efficiency
of 95 % including the heating value of tars, [4]. The producer gas com-
position as well as the tar load and its composition were reported in [28]
and [29].
242
6 System analysis
For the downdraft gasification process, the water load of the producer
gas was determined by means of mass balance, assuming a char flow
rate in form of carbon of 4 g/mn3 (dtf), [4]. The gas yield was calculated
to achieve a cold gas efficiency including tars of 80 %, [4]. The producer
gas composition was measured in the framework of a measuring cam-
paign conducted by fellow researchers of the author in 2003.
The water load of the fluidized bed steam gasification producer gas was
evaluated by application of mass balance calculations for a steam-to-
fuel ratio of 2.16, [33]. As no tar load and composition data could be
found for the investigated fluidized bed gasification process, values ob-
tained during a measuring campaign at a similar steam gasification plant
were employed. The gas yield was computed with a predefined cold gas
efficiency of 76 %, which corresponds to the development target of the
investigated fluidized bed gasification process, [33], increased by 10 %.
Measured data regarding the sulfur load and composition was not avail-
able for the three gasification processes. Hence, a rough estimation was
performed by hypothesizing that the sulfur present in the gasified wood
is entirely converted to either H2S or thiophene. This assumption yields
a conservative estimate due to the neglect of the amount of sulfur usu-
ally bound in particles, [222]. According to the DIN Plus norm, a sulfur
content of wood of 400 mg/kg on dry basis was assumed. The fractions
of thiophene and hydrogen sulfide were appraised via an analogy. It was
assumed that the ratio between organically bound sulfur and sulfur
bound in hydrogen sulfide equals the ratio between carbon bound in tars
and carbon bound in permanent gas species (CO, CO2, CH4, C2H4).
243
6 System analysis
Figure 6-2 shows the outline of the first downdraft gasification based
B-IGFC system configured for this analysis, named "System 1". In Sys-
tem 1, the humidity of the wood is decreased from 42 to 8 % in a con-
vective exhaust gas dryer. The dried wood is injected to the downdraft
gasifier where it is gasified with air. The gasification air is forced into the
gasifier by an electric air blower. The hot producer gas is de-dusted in a
ceramic particle filter operated at 650 °C and then humidified with wet
steam. This results in a cool down of the producer gas to 400 °C, which
corresponds to the operating temperature of the zinc oxide adsorbent
bed employed for the removal of hydrogen sulfide. Besides yielding the
desired gas cool down, the humidification is necessary to prevent car-
bon deposition at the temperature of the zinc oxide adsorbent bed tem-
perature. The de-dusted and desulphurized gas is heated to the cell de-
sign specific gas inlet temperature, see Table 6-7, and subsequently fed
to the SOFC, where the chemical energy of the producer gas is partially
converted to electricity and heat. The cooling air for the SOFC is pro-
vided by an electric air blower. The depleted electrode off-gases are
then guided through the heat integration network, explained in section
5.2.3, to cover the heat requirements resulting from the air and fuel
heating, steam generation and wood drying. Note that in System 1, or-
ganic sulfur species are not removed from the producer gas.
Figure 6-3 depicts the second downdraft gasification based B-IGFC sys-
tem configured for this analysis, named "System 2". In addition to Sys-
tem 1, an adiabatic methanation unit is placed downstream the ZnO
desulfurizer in System 2. The aim of the methanation step is to increase
the methane content of the producer gas, which should allow for more
SOFC internal reforming reactions and thus lower the required cooling
air flow rate.
244
6 System analysis
The scheme of the first updraft gasification based B-IGFC system de-
fined for this analysis is named "System 3" and depicted in Figure 6-4.
System 3 uses cold gas processing technology, thus representing
state-of-the-art. The wood with a humidity of 42 % is directly introduced
245
6 System analysis
246
6 System analysis
ticles are removed from the hot producer gas in a ceramic particle filter.
After a slight cool down, the hydrogen sulfide content is removed from
the producer gas in a ZnO trap bed at 400 °C. The subsequent heating
of the gas, conversion in the SOFC and heat recovery is carried out as
already mentioned above for the other investigated systems.
Figure 6-6 shows the third updraft gasification based B-IGFC system,
named "System 5". The producer gas originating from the air gasifica-
tion of the un-dried wood is again de-dusted in a multiclone.
Tars and organic sulfur species are degraded in a catalytic partial oxida-
tion reactor (CPO) through combustion and reforming reactions. Thus, a
part of the chemical energy of the producer gas is converted to heat
which results in a considerable increase of the producer gas tempera-
ture. To minimize the loss of chemical energy, the air-to-fuel ratio is kept
as low as possible without risking carbon deposition in the equipment
downstream of the CPO. The catalytic partial oxidation of tar-laden pro-
ducer gases was shown and reported in [153]. The CPO air flow rate is
provided by an electric blower. Ash particles are removed from the par-
tially oxidized producer gas in a ceramic particle filter before it is cooled
to 400 °C for subsequent hydrogen sulfide removal in a ZnO trap bed.
The conversion of the gas to electricity and useful heat through the
SOFC and heat integration network is conducted as discussed above.
247
6 System analysis
Despite the elevated tar load around 16 g/mn3 (dtf), the gas handling
should be rather unproblematic. At the high temperature of the steam
gasification producer gas, all tar species are in the gas phase. In con-
trast, coking problems usually arise during the evaporation of con-
densed tars, [133]. Hence regarding coking issues, it should suffice to
ensure the gas phase state of the tars by restraining a minimum system
temperature of 400 °C, which is well above the dew point of most tars.
According to the assumption in section 6.2.2.1, the sulfur load consists
of hydrogen sulfide to more than 96 %. Hence, it is possible that a re-
moval of organic sulfur species is not necessary.
The only difference between System 7 and System 6 is that a sulfur ad-
sorbent bed containing a metal oxide is placed downstream the particle
filter for coarse desulfurization and cracking of organic sulfur species.
According to [149], the metal oxide material has the potential to degrade
organic sulfur species. Accounting for the found influence of the metal
oxide on the producer gas composition, it was assumed that 10 % of the
producer gas reach equilibrium. The sulfur load in the raw producer gas
can thus be almost completely removed in the subsequent zinc oxide
trap bed, which should have a positive effect on the SOFC performance.
Table 6-9 summarizes the steam-to-carbon ratios (SC) of the raw pro-
ducer gases, the minimum SC at which graphite does not exist at equi-
librium conditions and the SC used in this analysis, which is 33 % higher
than the minimum SC. Note that Table 6-9 gives the minimum and the
safe air-to-fuel ratio for the CPO performed in System 5.
Figure 6-9 shows the C-H-O ternary diagram with isotherm lines for
760 °C, 600 °C and 400 °C, which are the lowest operational tempera-
ture of the four investigated cell designs, the critical temperature for car-
bon deposition from methane and the operational temperature of the
ZnO trap beds in Systems 4 and 5, respectively.
Figure 6-9:
C-H-O molar frac-
tion ternary diagram
for graphite forma-
Raw gases
NG C
0.0
Equlibrium
temperatures
tion under equilib- S4
1.0 T=760°C
rium conditions. S5 T=600°C
T=400°C
The isotherm lines com- Processed gases
puted for different tem- NG 0.2
0.8
peratures divide the S4
diagram in a region S5
where graphite is formed Graphite
at equilibrium and a 0.4
graphite free region. The formation 0.6
inscribed points repre-
sent the raw fuel gases region CO
as received from the 0.6
gasifier or the NG grid 0.4
and the processed fuel
gases. The points are
CO2
only moved when addi-
CH4
tional carbon, hydrogen 0.8
0.2
or oxygen is added to the
gases. Thus, only the
humidification and the air
injection for the CPO 1.0
result in a shift of the 0.0
points representing the 0.0 0.2 0.4 0.6 0.8 1.0
different fuel gases. H H2O O
250
6 System analysis
The raw updraft producer gas used in System 4 and 5 is slightly above
the isotherm line of 400 °C, indicating that carbon deposition is likely to
occur in the ZnO trap bed. The water addition in System 4 shifts the rep-
resenting point out of the graphite formation region following the dashed
line. The air addition for the CPO in System 5 shifts the point represent-
ing the raw updraft producer gas towards the oxygen corner of the dia-
gram following the dotted line. Methane is clearly in the graphite forma-
tion region for all the displayed temperatures. In contrast to the updraft
producer gas, the carbon deposition risk for methane is not highest at
400 °C but at 600 °C. This is because methane is stable at tempera-
tures below 300 °C. The water addition to yield a SC of 2 follows the
dashed-dotted line.
Figure 6-10:
C-H-O molar frac-
tion ternary diagram
for graphite forma-
Raw gases
S1/S2 C
0.0
Equilibrium
temperatures
tion under equilib- S3 1.0 T=760°C
rium conditions. S6/S7 T=400°C
T=300°C
The isotherm lines com- Processed gases
S1/S2 0.2
puted for different tem- 0.8
peratures divide the S3
diagram in a region S6/S7
where graphite is formed
at equilibrium and a 0.4 Graphite 0.6
graphite free region. The
inscribed points repre- formation
sent the raw fuel gases
CO
as received from the
region
0.6
gasifier and the proc- 0.4
essed fuel gases. The CO2
points are only moved CH4
when additional carbon,
hydrogen or oxygen is 0.8
0.2
added to the gases.
Thus, only the humidifi-
cation results in a shift of
the points representing 1.0
the different fuel gases. 0.0
0.0 0.2 0.4 0.6 0.8 1.0
H H2O O
Figure 6-10 shows the C-H-O ternary diagram with isotherm lines for
760 °C, 400 °C and 300 °C, where the latter is the operational tempera-
ture of the ZnO trap bed used in System 3. The lower ZnO temperature
calls for more water addition in System 3 than in System 4. Together
with the removal of almost 75 % of the water content occurring in the
wet scrubber, the amount of additional water needed in System 3 is
three times higher than in System 4, see dashed line in Figure 6-10.
251
6 System analysis
The water load of the raw downdraft producer gas (S1/S2) is low, due to
the limited wood humidity that can be sustained by the downdraft gasifi-
cation process. For the desulfurization at 400 °C, the amount of addi-
tional water is comparable to that required in System 3, see dashed-
dotted line in Figure 6-10.
The raw producer gas originating from the investigated fluidized bed
steam gasification process (S6/S7) exhibits a comparably high water
content despite the low humidity of the gasified wood. This is due to
steam being used as gasification agent. Nevertheless, this producer gas
does not allow for desulfurization at 400 °C without water addition, see
dotted line in Figure 6-10.
The producer gas compositions at the cell inlet, including the above dis-
cussed additional water, obtained from the seven B-IGFC systems de-
fined in section 6.2.2, are compiled in Table 6-10. The models used to
simulate the methanation, catalytic partial oxidation and ZnO trap bed
desulfurization units were explained in the sections 5.2.4 to 5.2.6.
In summary, the use of ZnO trap beds for desulfurization requires an in-
crease of the water content of the producer gas to avoid carbon deposi-
tion in all investigated systems except System 5. One way to overcome
this requirement could be the application of adsorbent materials with
operational temperatures above 700 °C. However, presently such mate-
rials are not available. Regarding the system design, in Systems 1, 2
and 3 the extra water calls for additional equipment such as steam gen-
252
6 System analysis
Figure 6-11: 50
Conversion effi-
ciencies of the
different producer 40
gases and the
Conversion efficiency [%]
reference gas at
85 % fuel utiliza- 30
tion and 0.6 V
operational volt-
age 20
10
0
Reference System 1 System 2 System 3 System 4 System 5 System 6 System 7
253
6 System analysis
1500
1000
500
0
Reference System 1 System 2 System 3 System 4 System 5 System 6 System 7
Hydrogen activation
CO activation
Oxygen activation
Ohmic
Current density
0.0100 0
0.00 0.02 0.04 0.06 0.08 0.10
Cell length [m]
254
6 System analysis
2
Current density
0.0100 0
0.00 0.02 0.04 0.06 0.08 0.10
Cell length [m]
Figure 6-14 shows the voltage loss and current density distributions of
the electrolyte-supported cell operated with the System 2 producer gas.
In contrast to the reference gas, the anode activation losses dominate
the ohmic losses along the entire cell length. Compared to the reference
gas, the activation losses are increased by 50 to 100 %, indicating that
the model parameters used for the electrolyte-supported cell describe
an anode catalyst with low activity in dilute gases.
0.0100 0
0.00 0.25 0.50 0.75 1.00 1.25 1.50
Cell length [m]
255
6 System analysis
The power output of the tubular cell is limited by ohmic losses, which are
at least one order of magnitude higher than the anode activation losses
for all the investigated gases, Figure 6-15. This explains the moderate
power output reduction of the tubular cell by 15 % in average when op-
erated with producer gases and indicates a very active anode catalyst.
distributions of the
anode-supported
2
Voltage loss [V]
0.0100 0
0.00 0.02 0.04 0.06 0.08 0.10
Cell length [m]
Figure 6-16 shows the voltage loss and current density distributions of
the co-current anode-supported cell operated with the reference gas.
The power output is limited by the oxygen activation and to a lesser ex-
tent by the anode activation. The oxygen activation losses are approx.
twice as high as the anode activation losses. The anode diffusion losses
gain significance towards the cell end, which can be attributed to fuel
depletion and the long diffusion path through the thick anode.
distributions of the
anode-supported
Current density [A/m ]
2
co-current cell
Voltage loss [V]
0.0100 0
0.00 0.02 0.04 0.06 0.08 0.10
Cell length [m]
256
6 System analysis
4000 1100
2
anode-supported
co- and counter-
3000 1050
current cells for
the reference gas
2000 1000
1000 950
0 900
0.00 0.02 0.04 0.06 0.08 0.10
Cell length [m]
257
6 System analysis
with a wide and flat peak in the middle part of the cell. This is the result
of the low temperatures at the cell inlet, due to endothermic STR reac-
tions, which yield only low fuel conversion in this cell region leaving over
most of the fuel for the conversion in the latter parts of the cell.
Throughout the entire co-current cell, the anode and oxygen activation
are the dominant voltage losses.
In contrast, the counter-current cell features a plateau of very high cur-
rent densities due to high temperatures at the fuel inlet, despite endo-
thermic STR reactions taking place. The high temperatures are main-
tained by the cooling air, which is heated in the latter parts of the cell.
The fuel conversion at the cell inlet is mainly limited by the oxygen acti-
vation and diffusion. Towards the fuel outlet, the current density is in-
creasingly limited by anode activation losses, which result from the fast
fuel depletion in the front parts of the cell. In sum, the counter-current
cell yields a lower average current density with the reference gas than
the co-current cell due to the diffusion limitations at the cell inlet.
anode-supported
co- and counter-
3000 1050
current cells for
the producer gas
of System 2 2000 1000
1000 950
0 900
0.00 0.02 0.04 0.06 0.08 0.10
Cell length [m]
258
6 System analysis
where the activation losses are relatively high. For the reference gas,
the maximum current density is reached in the second cell half at 1060
K. This is due to the low temperatures at the cell inlet resulting from the
STR inhibiting a fast fuel conversion.
The counter-current cell yields considerably higher current densities at
the fuel inlet than the co-current cell due to 50 K higher solid tempera-
tures. In contrast to the reference gas, with producer gas the counter-
current cell is not limited by oxygen diffusion but by the anode activation
like the co-current cell. Thus, the fast fuel conversion at the cell inlet
(and high temperatures) is beneficial to the current production of the
counter-current cell. In consequence, the total amount of current pro-
duced by the counter-current cell is slightly higher than that produced by
the co-current cell when operated with producer gases.
In summary it can be said that for a given operational voltage the anode
catalyst activity is the dominating factor with respect to the mean current
density obtained from different cell designs operated with producer
gases. Due to their high degree of nitrogen or water dilution, producer
gases require highly active anode catalysts to yield satisfying mean cur-
rent densities. It appears that an initial estimation of the catalyst activity
can be based on the cell design. Designs limited by ohmic losses, such
as the tubular and the planar electrolyte-supported cells, will probably
feature more active anode catalysts to compensate the ohmic losses.
The reason why this does not apply to the electrolyte-supported cell in-
vestigated in this analysis is because it was primarily developed as a
catalytic burner to provide heat to households. Thus, the anode catalyst
is most likely not tuned for high activity with dilute gases. Regarding dif-
fusion limitations resulting from the dilution quality of producer gases it
can be said that only the anode-supported cells showed a slight and
hence negligible response. Finally, it was found that cell designs with
counter-current flow pattern allow for high temperatures at the fuel inlet
and are thus beneficial to the conversion of producer gases.
Figure 6-20: 15
Comparison of the Anode counter-current
predicted air-to- Anode co-current
fuel ratios re- 12 Electrolyte
quired to achieve Cathode
a mean cell tem-
Air-to-fuel ratio [-]
perature equal to 9
the operational
temperature of the
investigated cell 6
designs operated
with the different
producer gases 3
0
Reference System 1 System 2 System 3 System 4 System 5 System 6 System 7
supported cell
operated with the 40 1100
reference gas and
1164 K cell inlet
temperature
20 950
0 800
0.00 0.02 0.04 0.06 0.08 0.10
Cell length [m]
The reason for this can be found in the cell inlet temperatures. Figure
6-21 shows the species molar fractions and the solid temperature distri-
butions predicted for the electrolyte-supported cell operated with the ref-
260
6 System analysis
supported tubular
cell with air deliv-
1200
ery tube operated
with the reference
gas
1100
1000
0.00 0.25 0.50 0.75 1.00 1.25 1.50
Cell length [m]
Once in the cathode gas channel, the temperature of the cooling air is
lowered due to the STR reactions. This results in a considerable in-
261
6 System analysis
crease of the temperature gradient between the solid structure and the
cooling air and accordingly allows a low AF.
When operated with producer gas with low internal reforming potential,
the cell inlet region is not sub-cooled yielding lower temperatures in the
latter parts of the cell. Figure 6-23 shows the temperature distribution
predicted for the tubular cathode-supported cell operated with the pro-
ducer gas of System 1, which has the lowest internal reforming potential
of all investigated gases. Compared to the reference gas, the solid tem-
perature in the later parts of the cell is approx. 50 K lower, while at the
fuel inlet it is almost 200 K higher. The cooling air is gradually heated
along the air delivery tube and the cathode channel and does not act as
heat source for the fuel inlet region. Hence, the AF has to be increased
to preserve a sufficiently high temperature gradient between the solid
structure and the cooling air for the removal of the cell waste heat.
Cathode gas
Gas temperature in
air delivery tube
1000
0.00 0.25 0.50 0.75 1.00 1.25 1.50
Cell length [m]
262
6 System analysis
Figure 6-24 shows that in the anode-supported co-current cell the STR
reactions are distributed over a length roughly three times longer than in
the electrolyte-supported cell, Figure 6-21. Thus, most of the heat re-
quired for the endothermic STR reactions is provided by electrochemical
reactions and heat conduction through the metallic bipolar plates and
needs not to be provided in form of sensible heat of the inflowing fuel
gas and cooling air as is the case for the electrolyte-supported cell.
supported co-
current cell oper- 40 1100
ated with the ref-
erence gas and
914 K cell inlet
20 950
temperature
0 800
0.00 0.02 0.04 0.06 0.08 0.10
Cell length [m]
When operated with producer gas with lower internal reforming potential,
the anode-supported co-current cell requires lower AF than the electro-
lyte-supported cell due to more effective cooling resulting from the lower
cell inlet gas temperatures.
Among all investigated cell designs, the anode-supported counter-
current cell shows the lowest sensitivity of the AF towards the internal
reforming potential of the fuel gas. For gases with high internal reform-
ing potential, the heat amount produced in electrochemical reactions at
the cell inlet is high enough to cover the heat demand of the STR reac-
tions taking place. Further, the cooling air transports heat from the latter
parts of the cell to the cell inlet region where it is also consumed by the
STR reactions. When operated with fuel gas featuring low internal re-
forming potential, the heat amount produced in the latter cell parts is re-
duced due to fuel depletion resulting from the fuel conversion at the cell
inlet. The heat is then primarily transported from the fuel inlet region to
the latter cell parts through solid heat conduction. Thus, the counter-
current cell yields relatively constant AF for different fuel gases.
263
6 System analysis
Overall it was found that the AF increases with decreasing internal re-
forming potential of the fuel gas. The sensitivity of the AF towards the in-
ternal reforming potential depends on the cell design. Co-current cells
were found to be highly sensitive due to the constraint of the cell inlet
gas temperatures to the values determined for the reference case. In
contrast, the AF required by counter-current cells was established to be
almost independent of the internal reforming potential of the fuel gas.
This robustness makes this cell design appealing for the operation with
producer gases, whose temperatures and compositions often fluctuate.
1073
1023
973
Reference System 1 System 2 System 3 System 4 System 5 System 6 System 7
Figure 6-25 compares the maximum solid temperatures predicted for the
investigated cell designs and producer gases. It can be seen that fuel
gases with high internal reforming potential tend to yield high maximum
solid temperatures, see "Reference" and "System 4". This can be attrib-
uted to the low solid temperatures resulting from endothermic STR reac-
tions at the fuel inlet, which are compensated by increased solid tem-
peratures in the latter cell parts to satisfy the accordance of the mean
cell temperature with the operational temperature of the respective cell
design.
264
6 System analysis
-150
0.0 0.2 0.4 0.6 0.8 1.0
Dimensionless cell length [-]
Figure 6-26 gives the comparison of the solid structure temperature dis-
tributions predicted for the electrolyte-supported cell when operated
with the reference gas and the different producer gases. Despite the
producer gas of System 4 having the highest internal reforming potential
of all investigated fuel gases, it does not yield the highest maximum
solid temperature. The reason for this is that the fuel inlet region is less
cooled compared to the reference case. For the System 4 producer gas,
the anode activation losses are dominant. Consequently, the maximum
current density is produced close to the fuel inlet where the reactant par-
tial pressures are comparably high. The released heat of the electro-
chemical reactions is sufficient to hinder a strong cooling of the fuel inlet
area. With the reference gas, the maximum current density is reached in
the second half of the cell as a consequence of the dominating ohmic
losses which only allow for high current densities at elevated tempera-
tures. The cooling of the fuel inlet area is more pronounced because,
compared to the heat required by the STR reactions, less heat is pro-
duced through electrochemical reactions. The more pronounced cooling
results in higher maximum solid temperatures at the cell outlet.
For the cathode-supported cell, Figure 6-27 shows that the maximum
solid temperatures are not reached at the fuel outlet but at around three
quarters of the cell length. This is due to the cooling effect stemming
from the fresh air flowing through the air delivery tube. The producer gas
of System 4 does not yield lower solid temperatures at the fuel inlet than
the reference gas despite its higher internal reforming potential. Never-
theless, the maximum solid temperature attained with the producer gas
of System 4 is higher than that obtained with the reference gas. The
reason for this is discussed based on Figure 6-28.
265
6 System analysis
-300
0.0 0.2 0.4 0.6 0.8 1.0
Dimensionless cell length [-]
The STR reactions taking place in the fuel inlet region with the reference
gas require considerably more heat than is provided by electrochemical
reactions and the pre-heated cooling air. This results in reduced solid
structure temperatures, which are again the reason for the ohmic limita-
tions hindering more electrochemical reactions. The producer gas of
System 4 yields current densities at the fuel inlet comparable to the ref-
erence gas, which in turn allow higher solid structure temperatures at
the cell inlet. Additionally, compared to the reference gas more heat is
provided to the fuel inlet by the pre-heated cooling air.
The STR reactions are complete after roughly one tenth of the cell
length for both gases. For the reference gas, a very fast temperature in-
266
6 System analysis
Reference System 1
solid structure System 2 System 3
80
temperature dis- System 4 System 5
tributions pre- System 6 System 7 Over heating zone
dicted for the 40
anode-supported
co-current cell 0
operated with the
reference gas and
-40
the different pro- Sub cooling zone
ducer gases
-80
-120
0.0 0.2 0.4 0.6 0.8 1.0
Dimensionless cell length [-]
All investigated gases except the producer gas of System 4 yield maxi-
mum temperatures in the narrow range between 1084 and 1098 K. The
undiluted gases of the Systems 6 and 7 and the reference gas achieve
the highest maximum solid temperatures. This is because the current
densities are considerably higher in the fuel inlet region compared to the
other gases due to less activation polarization losses. Compared to the
reference case, the current densities reached with the diluted producer
gas of System 4 are considerably lower due to increased activation
losses. The heat amount released in the electrochemical reactions is
thus not sufficient to cover the heat demand of the STR reactions taking
place close to the fuel inlet. Consequently, the predicted solid tempera-
tures close to the fuel inlet for the System 4 producer gas are lower than
those obtained with all other gases. To satisfy the accordance of the
mean cell temperature and the design temperature, the lower solid tem-
peratures at the fuel inlet are compensated with higher temperatures
towards the fuel outlet region.
The role of tars in the cooling of the cell inlet area was investigated by
comparing the solid temperature distribution predicted for the electro-
lyte-supported cell operated with the tar free producer gas of System 3
with that predicted for the raw updraft producer gas, Table 6-8. Figure
6-31 depicts that the added tar load does contribute to the cooling of the
268
6 System analysis
cell inlet area. With a 6 K lower solid temperature at the cell inlet, the
impact of the tars is however small. Tar removal especially to prevent
excessive maximum solid temperatures or thermal gradients is therefore
not necessary.
1250
electrolyte-
supported cell
1225
operated with the
producer gas of
System 3 and the 1200
producer gas as
obtained from the 1175
updraft gasifier
1150
0.00 0.02 0.04 0.06 0.08 0.10
Cell length [m]
Figure 6-32: 4
Comparison of the Anode counter-current
maximum tem- Anode co-current
Maximum temperature gradient [K/mm]
0
Reference System 1 System 2 System 3 System 4 System 5 System 6 System 7
The reference gas yields the highest temperature gradients for all cell
designs and gases. Similar to the maximum solid temperatures, the
269
6 System analysis
270
6 System analysis
Eq. 203
with EqH 2 ,in = yH 2 ,in + yCO ,in + 4 ⋅ yCH 4 ,in + 6 ⋅ yC 2 H 4 ,in
+ 18 ⋅ yC 7 H 8 ,in + 4 ⋅ yC 2 H 4O2 ,in + 17 ⋅ yC 7 H 8O ,in
In Eq. 203, Itot stands for the mean current densities obtained for the
various producer gases and cell designs, see Figure 5-1, Aact repre-
sents the active area of the corresponding stacks, F is the Faraday con-
stant, UF defines the fuel utilization and EqH2,in is the hydrogen equiva-
lent coefficient of the respective producer gas.
The mass flow rates of wet wood were computed for all the investigated
systems and cell designs based on the raw gas yields, see Table 6-8, to
deliver the required producer gas flow rates at the cell inlet, Figure 6-33.
700
designs
600
500
400
300
200
100
0
System 1 System 2 System 3 System 4 System 5 System 6 System 7
For all investigated systems and cell designs, no heat pinch points were
found, indicating that the sensible heat of the post-combusted SOFC off-
gases is sufficient to cover all heat requirements.
Based on the predefined heat exchange coefficients, see section 5.2.3,
the required heat exchange areas were computed. Figure 6-34 shows
that the heat exchange areas of the systems employing anode-
supported cells are generally smaller than those of systems using elec-
trolyte- or cathode-supported cells. This can be traced down to the con-
siderably lower cell inlet temperatures of the anode-supported cells. De-
spite lower cell inlet temperatures, the electrolyte-supported cells require
larger heat exchange areas than the cathode-supported cells. This is the
consequence of the lower air-to-fuel ratios predicted for the cathode-
supported cells due to the cell internal air pre-heating through the air de-
livery tube. Regarding the different flow patterns of the anode-supported
cells, Figure 6-34 shows that the lower cell inlet temperatures and pre-
dicted air-to-fuel ratios of the counter-current cells result in considerably
smaller heat exchange areas compared to co-current cells.
designs
900
600
300
0
Anode cc
Anode co
Electrolyte
Cathode
Anode cc
Anode co
Electrolyte
Cathode
Anode cc
Anode co
Electrolyte
Cathode
Anode cc
Anode co
Electrolyte
Cathode
Anode cc
Anode co
Electrolyte
Cathode
Anode cc
Anode co
Electrolyte
Cathode
Anode cc
Anode co
Electrolyte
Cathode
1 2 3 4 5 6 7
272
6 System analysis
Figure 6-35:
Efficiency loss Input LHV
Electricity AC
100%
10%
20%
30%
40%
50%
60%
70%
80%
90%
AC electrical effi-
0%
ciencies for the
different B-IGFC
systems and cell
Anode cc
designs Anode co
1
DC/AC losses
Electrolyte
Cathode
Anode cc
Anode co
Auxiliary power
Electrolyte
Cathode
Anode cc
Anode co
Sulfur losses
Electrolyte
Cathode
Anode cc
Anode co
SOFC losses
Electrolyte
Cathode
Anode cc
Fuel utilization
Anode co
5
Electrolyte
Cathode
Anode cc
Gas processing
Anode co
6
Electrolyte
Cathode
Anode cc
Gasification
Anode co
7
Electrolyte
Cathode
273
6 System analysis
274
6 System analysis
for Systems 4 and 5. Similar to System 2, the water produced during the
methanation in System 4 seriously obstructs the attainment of sulfur
concentrations below 2 ppmV. In System 5, total sulfur concentrations
around 2 ppmV are reached despite remaining parts of the organic sul-
fur species in the producer gas. The reason for this is the lower water
content of the producer gas after the CPO which allows achieving H2S
concentrations in the zinc oxide trap bed that are low enough to com-
pensate the organic sulfur species.
The auxiliary power losses resulting from the power consumption of fans
and blowers in System 4 are around 4 to 5 % and therewith the lowest in
this analysis. This is because the corresponding producer gas has the
highest internal reforming potential of all investigated gases and there-
fore requires the lowest air-to-fuel ratios. Due to the comparably high net
DC efficiencies, the DC/AC conversion losses are in the range of 2 to
3 % and thus are slightly higher than in the Systems 1, 2, 6 and 7.
Similar to systems 1 and 2, the anode-supported counter-current cell
yields the highest net AC efficiencies in the Systems 3 and 5 with
24.3 % and 26.5 %, respectively. Regarding System 4, the cathode-
supported cell designs benefits from the cooling air cool down at the fuel
inlet which allows a lower air-to-fuel ratio than that of the anode-
supported counter-current cell. The correspondingly lower auxiliary
power losses lead to a net AC efficiency of 32.1 % compared to 31.7 %
reached by the anode-supported counter-current cell. Overall it can be
stated that System 4 outperforms all other investigated systems be-
cause the combination of updraft gasification and adiabatic methanation
yields a producer gas with high internal reforming potential with a com-
parably high cold gas efficiency of approx. 85 %.
Fluidized bed steam gasification based systems: System 6 and 7
In the fluidized bed steam gasification process, 76 % of the chemical
energy of the feedstock is converted to chemical energy of the producer
gas and the remainder into high-temperature heat.
In System 7, the tar load and the hydrocarbon content are slightly de-
creased over the metal oxide bed used for the degradation of organic
sulfur species. The loss of internal reforming potential is however com-
pensated by a noticeable amount of carbon monoxide that is shifted to
hydrogen and carbon dioxide. If this shift occurs cell internally, the re-
leased heat of reaction adds to the SOFC waste heat thus lowering the
conversion efficiency.
276
6 System analysis
The sulfur induced losses of System 6 are 4.5 % compared with 1.9 %
of System 7. This shows that the neglect of the removal of almost
5 ppmV organic sulfur species in the system design may result in con-
siderable overestimation of the system efficiency. Compared to the Sys-
tems 1 to 5, the excessive water content of the fluidized bed producer
gases hinders the desulfurization in the zinc oxide trap beds more and
yields approx. 0.5 ppmV higher sulfur concentrations in the producer
gas at the cell inlet. A possibility to overcome this would be a reduction
in the amount of added water and thus a decrease of the steam-to-
carbon ratio. This however might cause carbon deposition problems.
The auxiliary power losses for the different cell types are almost similar
for both systems in the range between 4.5 to almost 9 %. The DC/AC
losses amount to roughly 2 %.
System 7 with the anode-supported counter-current cell reaches a net
AC efficiency of 23.7 %, which is slightly higher than the efficiencies
reached by the downdraft gasification based systems and close to the
efficiency of System 3, where 20 % of the thermal input was removed
from the producer gas in form of tars.
In sum it can be stated that the cold gas efficiency of the gasification
process and the auxiliary power requirements of the SOFC resulting
from the SOFC cooling are crucial for the overall system efficiency. In
both cases, the downdraft gasification process is not favorable. The air-
blown updraft gasification in contrast was found very promising for the
application in B-IGFC systems due to the high cold gas efficiency and
typically high hydrocarbon concentration in the producer gas. The latter
allow for the relocation of the final gasification reactions into the SOFC
using waste heat and thus lowering the cooling requirements and attrib-
uted auxiliary power requirements. The fluidized bed steam gasification
yields producer gas with high hydrocarbon content too, however with
lower cold gas efficiencies. The corresponding losses can barely be
compensated by the subsequent producer gas conversion to electricity.
277
6 System analysis
heat below 200 °C of the system flue gases is not used. The heat loss
hence grows with increasing flue gas flow rate, which directly correlates
with the air-to-fuel ratio. The air-to-fuel ratio thus not only considerably
affects the net AC efficiency through the auxiliary power requirements
but also the thermal efficiency.
System 3 yields the smallest amount of useful heat because it was as-
sumed that the entire heating value of the tar load removed in the cold
gas cleaning system is required for the treatment of the wash water.
designs
50%
40%
30%
20%
10%
0%
Cathode
Cathode
Cathode
Cathode
Cathode
Cathode
Cathode
Anode cc
Electrolyte
Anode cc
Electrolyte
Anode cc
Electrolyte
Anode cc
Electrolyte
Anode cc
Electrolyte
Anode cc
Electrolyte
Anode cc
Electrolyte
Anode co
Anode co
Anode co
Anode co
Anode co
Anode co
Anode co
1 2 3 4 5 6 7
With 75.7 %, System 4 with the cathode-supported cell yields the high-
est total system efficiency in this analysis. The corresponding thermal
efficiency is 43.6 %. The useful heat may be converted to additional
electricity through a bottoming cycle, thus covering the auxiliary power
requirements of the cooling air blower. Assuming a moderate electrical
efficiency of 12 % for an organic rankine cycle operated at 200 °C, [163],
the net AC system efficiency of System 4 could be increased to 37.3 %,
while also increasing the system power output by 16 %. This is sufficient
to be superior to state-of-the-art biomass gasification based gas engine
power plants featuring electrical efficiencies around 25 %, [226, 227].
Figure 6-37 depicts the net AC electrical power output of the investi-
gated B-IGFC systems and cell designs. The electrolyte-supported cell
features the lowest power outputs in all investigated systems. This re-
sults from the obtained low current densities when operated with pro-
ducer gas and the comparably small active area, which is a conse-
278
6 System analysis
quence of the reference gas based top-down approach chosen for this
analysis. The auxiliary power requirements are high due to the repeat-
edly discussed high air-to-fuel ratios (AF). The cathode-supported cell in
contrast yields the highest power outputs, due to the high mean current
densities achieved with the producer gases and the moderate AF.
375
250
125
0
System 1 System 2 System 3 System 4 System 5 System 6 System 7
279
6 System analysis
Figure 6-38:
Direct plant cost Direct plant costs [2007-k€]
estimates struc-
1000
1500
2000
2500
3000
3500
500
tured according to
0
the system mod-
ules discussed in Anode cc
section 5.4.2 for
the different B- Anode co
Wood handling
1
Anode cc
Anode co
2
Electrolyte
Gasifier
Cathode
Anode cc
Anode co
3
Electrolyte
Gas processing
Cathode
Anode cc
Anode co
4
Electrolyte
Cathode
FC Balance of Plant
Anode cc
Anode co
5
Electrolyte
Cathode
Anode cc
Anode co
Fuel cell
6
Electrolyte
Cathode
Anode cc
Anode co
Auxiliary
7
Electrolyte
Cathode
280
6 System analysis
281
6 System analysis
is removed in the employed cold gas cleaning system, the steam gen-
erator required for the re-humidification of the producer gas accounts for
almost 80 % of the gas processing module costs. This is because no ce-
ramic particle filter unit and zinc oxide trap bed are employed. The fluid-
ized bed systems have the advantage that the producer gas can be hu-
midified without the need for a steam generator by using the sensible
heat of the producer gas for the evaporation of the additional water.
Consequently, the gas processing module of these systems is compara-
tively inexpensive. The costs of the gas processing could be effectively
reduced in fluidized bed and updraft systems by increasing the wood
humidity at the gasifier inlet. This would not only reduce the cost in-
curred by the gas humidification but also the cost of wood drying as
mentioned above.
The fuel cell balance of plant module (FC-BoP) includes the heat ex-
changers required to adjust the producer gas and the cooling air tem-
perature to the predefined cell inlet values, the cooling air blower, the
DC/AC power converter and the anode off-gas combustor. Except for
the downdraft and fluidized bed steam gasification based systems em-
ploying anode-supported counter-current cells, the FC-BoP module is
the dominant cost factor in all the investigated systems, accounting for
25 to 50 % of the total system costs. In the systems employing electro-
lyte- or cathode-supported cells, the FC-BoP costs are generally higher
than in the corresponding systems based on anode-supported cells.
This is due to the considerably lower cell inlet temperatures and air-to-
fuel ratios required by the anode-supported cells.
The fuel cell module (FC) accounts for one eights to a quarter of the to-
tal system cost. The fuel cell costs are equal for all the systems due to
the top-down approach chosen for this analysis, which considers the FC
size as fixed. Note that the costs entailed by the catalysts employed in
systems 2, 4 and 5 were allocated to the FC module. The absolute
module cost in these systems can thus be increased by up to 25 %
compared to systems 1, 3, 6 and 7.
The auxiliary equipment module accounts for 7 to 12 % of the total
system costs. Half of these costs can be attributed to the heat ex-
changer required to recover useful heat from the system flue gases.
The direct plant costs are related to the net AC power output for a better
comparability of the systems. Figure 6-39 shows the specific direct plant
costs computed for the investigated B-IGFC systems and cell designs.
282
6 System analysis
6,000
4,000
2,000
0
System 1 System 2 System 3 System 4 System 5 System 6 System 7
283
6 System analysis
and maintenance expenses. Further, the revenues from heat sales per
sold kWhel net AC are depicted assuming 100 % heat sales. It can be
seen that for each of the systems and cell designs the sum of the PPCs
resulting from the equipment depreciation (D-costs) equals the capital
costs plus the costs for the operation and maintenance (C-costs).
Figure 6-40:
Power production PP cost and heat revenues [2007-€/kWhel net AC]
Heat revenues
costs structured
-0.15
-0.05
0.05
0.15
0.25
0.35
0.45
0.55
according to the
system modules
discussed in sec-
tion 5.4.2 includ- Anode cc
ing fuel, capital Anode co
Fuel
the different B-
IGFC systems Anode cc
and cell designs
Anode co
2
Electrolyte
Cathode
Gasifier
Anode cc
Anode co
Gas Processing
Electrolyte
Cathode
Anode cc
FC Balance of Plant
Anode co
4
Electrolyte
Cathode
Anode cc
Anode co
Fuel cell
Electrolyte
Cathode
Auxiliary
Anode cc
Anode co
6
Electrolyte
Capital cost
Cathode
Anode cc
Anode co
Operation
Electrolyte
Cathode
284
6 System analysis
In Systems 1 and 2, the fuel costs (F-costs), the D-costs and the
C-costs each contribute to approx. one third to the overall PPCs. In all
other systems, the F-costs account for approx. 35 to 45 % of the PPCs.
This is mainly due to the lower specific plant costs. Regarding the D-
costs, the most important cost drivers are already discussed above.
However, it is worthwhile noting that the fuel cell stack accounts for
merely 5 to 15 % of the overall PPCs.
In the systems 1, 3, 5, 6 and 7, the anode-supported counter-current cell
design yields the lowest PPCs compared to all other cell designs. Sys-
tem 1 is the only system, where the lowest PPCs are reached by a cell
design that does not also yield the lowest specific plant costs. Com-
pared to the cathode-supported cell design, the 8 % (relative) efficiency
advantage of the anode-supported counter-current cell design is just
sufficient to compensate 4 % higher specific plant costs. The counter
example is System 2, where the cathode-supported cell design yields
lower PPCs than the anode-supported counter-current cell design de-
spite the latter featuring an almost 10 % (relative) higher net AC system
efficiency. The reason for this can be found in the 20 % lower specific
plant costs resulting from the higher current density produced by the
cathode-supported cell due to high catalyst activity in dilute gases.
This clearly shows that the PPCs are more sensitive towards the spe-
cific plant costs than towards the system efficiency. The latter is effec-
tive only on the F-cost share, which accounts for moderate 30 to 45 % of
the PPCs even though wood is an expensive feedstock. Hence, effi-
ciency maximization is limited regarding the reduction of the PPCs. In
contrast, power output increase and reduction of specific plant costs
through optimization of the SOFC operational parameters is effective for
55 to 70 % of the PPCs. This emphasizes once more the importance of
the power output of the cell, which mainly depends on the anode cata-
lyst activity especially for the operation with dilute producer gases.
Figure 6-41 depicts the net power production costs assuming that the
entire heat produced is sold. The heat revenues are sufficient to cover
most of the F-costs and are therefore crucial for the market competive-
ness of B-IGFC systems. Given the electricity price currently paid in
Germany for electricity produced from biomass using highly advanced
technologies (gray marked area), [229], the only competitive systems in-
vestigated in this analysis are the Systems 4 and 5. In System 4, the
cathode-supported cell designs yields the lowest PPCs with
0.1154 €/kWhel net AC. In System 5, the anode-supported counter-
285
6 System analysis
0.20
0.15
0.10
0.05
0.00
System 1 System 2 System 3 System 4 System 5 System 6 System 7
75%
50%
-80% -60% -40% -20% 0% 20% 40% 60% 80%
Parameter variation [% of Base case]
286
6 System analysis
Figure 6-42 shows that the plant availability, represented by the plant
factor (PF), has considerable PPC reduction potential. Increasing the PF
by 10 % yields 7 % lower PPCs with values around 0.107 €/kWhel net
AC. Reduced plant availability leads to fast increasing PPCs. With a PF
of 63 %, which is 26 % (relative) lower than the base case value, the
PPCs are already increased by 30 % and reach the lower limiting price
of 0.15 €/kWhel net AC for biomass electricity from highly advanced
technologies in Germany. A further reduction of the PF to approx. 45 %,
which corresponds to 52 % (relative) of the base case, yields an in-
crease of the PPCs by 73 % to 0.2 €/kWhel net AC, which is the upper
limiting electricity price according to [229]. This shows that the plant
availability plays a crucial role for the competitiveness of the B-IGFC
technology due to the high involved capital costs.
Figure 6-42 further depicts that the PPCs increase fast with catalyst life-
times below 50 % of the base case. This indicates that the catalyst de-
velopment should focus on the achievement of catalyst lifetimes around
one to two years. Further improvements regarding the lifetime of the
catalysts employed in the adiabatic methanation and the catalytic partial
oxidation reactors in the investigated systems 2, 4 and 5, have almost
no impact on the PPCs.
The corresponding parameters in the economic model were varied to
evaluate the general importance of the plant cost estimate accuracy, the
wood price and the heat price or demand, respectively. The wood price
and the heat price or demand, respectively, are equally sensitive and
considerably affect the PPCs. However, the PPCs stay below the lower
limiting electricity price of 0.15 €/kWhel net AC despite either a heat price
or demand decrease or a wood price increase by 40 % compared to the
base case. The latter shows that a wood price of 117 € per dry ton of
wood would still allow profitable operation of System 4 with the cathode-
supported cell design. The same applies, when assuming 30 % higher
direct plant costs, which corresponds to the maximum error expected for
the plant cost estimates.
100%
90%
0 1 2 3 4 5 6 7 8 9
Numer of FC exchanges during planning horizon [-]
ciency of 37.2 %. Despite this very high efficiency, the PPCs increase by
almost 70 % compared to the base case yielding 0.1912 €/kWhel net AC,
which renders profitable operation difficult.
25 0.15
20 0.1
0.6 0.625 0.65 0.675 0.7
Voltage [V]
The same trend applies to System 2, which confirms the initial assump-
tion that the operational voltage has to be as low as possible to yield
reasonable power outputs and specific plant costs when SOFCs are op-
erated with dilute fuel gases. For System 7, the PPCs initially decrease
with increasing operational voltage. The minimum PPCs are reached at
0.63 V. Further increasing the operational voltage leads to increasing
PPCs similar to Systems 2 and 4. The reason for the initial PPCs de-
crease is that the un-diluted producer gas of System 7 causes consid-
erably less activation polarization losses. This reduces the power output
decrease resulting from the higher operational voltages.
0.6
0.4
0.6 0.625 0.65 0.675 0.7
Voltage [V]
289
6 System analysis
producer gas of
2
700
0
0.00 0.25 0.50 0.75 1.00 1.25 1.50
Cell length [m]
However, the thermal stress imposed on the SOFCs employed in the in-
vestigated systems increases with higher operational voltages. Figure
6-45 exemplifies that the maximum temperature gradients increase with
increasing operational voltages. The reason for this lies in the current
density distributions, which become increasingly inhomogeneous with
rising operational voltages. Figure 6-46 shows for the producer gas of
System 7 that despite in relative terms more producer gas is converted
at the fuel inlet with higher operational voltages, the absolute amount
decreases. This leads to a drop of the solid temperatures at the fuel inlet
as the heat produced in the electrochemical reactions becomes insuffi-
cient to cover the heat demand of the STR reactions taking place. The
lower solid temperatures at the cell inlet imply higher temperatures in
the latter cell parts, which in turn lead to higher temperature gradients.
Consequently, the PPCs benefit resulting from higher operational volt-
ages for the producer gas of System 7 may be cancelled out by faster
degradation due to higher thermal stress.
In conclusion it can be stated that none of the parameters of the eco-
nomic models and important assumptions of the system analysis inves-
tigated in this sensitivity analysis change the relative differences be-
tween the considered B-IGFC systems. The absolute PPC estimates are
however in parts considerably changed, which could lead to a better or
worse competitiveness of the investigated systems.
290
291
292
7 Concluding remarks
7 Concluding remarks
Biomass has a large potential for electricity generation among the re-
newable energy sources. A promising approach to achieve high electri-
cal efficiencies is the application of fuel cells for the conversion of pro-
ducer gases originating from biomass gasification. This combination is
referred to as "Biomass Integrated Gasification Fuel Cell System"
(B-IGFC). The solid oxide fuel cell in particular has gained much interest
due to its ability of converting hydrocarboneous fuel gases and its com-
parably low fuel gas requirements. This thesis aimed at the demonstra-
tion of the B-IGFC technology on kW-scale and at the identification of
promising B-IGFC system concepts with power outputs around 1 MWel
by means of a systems analysis.
7.1 Conclusions
7.1.1 Experiments
The PSI B-IGFC system is based on the updraft gasification process.
The resulting producer gas is de-dusted and its tar load is reduced
through a catalytic partial oxidation (CPO) before being fed to the SOFC.
The starting point for the experimental part of the thesis was the
achievement of stable long-term operation of the in-house developed
lab-scale updraft gasifier. Diverse modifications of the gasifier and the
addition of a steam generator made a stable operation over 165 h non-
stop possible. Experiments with SOFC stacks showed that tar is to
some extent a fuel for SOFCs and that sulfur reduces the power output
but does not inhibit stable operation of SOFCs. Experiments with a com-
mercial CPO catalyst revealed satisfying conversion performance for
oxygenated tars and aromatics. The CPO catalyst was further found
suitable for the decomposition of organic sulfur compounds.
The PSI B-IGFC concept was first time demonstrated on kW-scale. The
SOFC delivered 40 % less current when operated with producer gas
compared to partially oxidized methane. The gasifier and the CPO were
operated without problems. Ash deposits in the SOFC system were
found. Overall the first-time demonstration of the PSI B-IGFC concept
was successful and possibilities for future improvements of the experi-
mental setup were identified.
293
7 Concluding remarks
294
7 Concluding remarks
dominant role due to the dilute character of producer gases. Thus, high
anode catalyst activity is essential for high power outputs. According to
the model parameters used in the analysis, the tubular cell features a
highly active anode catalyst. The anode-supported counter-current cell
partially balances out the lower catalyst activity with high temperatures
at the cell inlet. Diffusion limitations do not gain importance for any of
the investigated producer gases and cell designs.
The cooling requirements strongly depend on the SOFC flow pattern
and the internal reforming potential of the producer gases. The required
air-to-fuel ratios (AF) generally decrease with increasing internal reform-
ing potential. The co-current cells are considerably more sensitive to-
wards the internal reforming potential than the counter-current cell,
which requires the lowest AFs of all cell designs. Compared to pre-
reformed natural gas, all investigated producer gases cause lower tem-
perature gradients and temperature extremes, which may reduce tem-
perature induced cell degradation.
The interactions between the gasification processes, gas processing
technologies and SOFC designs were investigated with respect to the
net system efficiencies and balance of plant requirements resulting from
the SOFC cooling. A generalized heat integration network was defined.
The corresponding simulations were performed with ASPEN PLUS.
The water concentrations of the raw producer gases are insufficient to
prevent carbon depositions at the operational temperature of the gener-
ally employed zinc oxide sulfur trap beds. In all investigated systems,
except the one employing a CPO, considerable amounts of water have
to be added to the producer gases.
The net system efficiency is preeminently determined by the cold gas ef-
ficiency of the biomass gasification process and the auxiliary power re-
quirements. The latter directly correlate with the required air-to-fuel ra-
tios, which in turn are a consequence of the internal reforming potential
of the corresponding producer gases. The adiabatic methanation and
the low-temperature CPO are effective means to increase the internal
reforming potential of producer gases.
With 32.1 %, the highest system efficiency is reached by the combina-
tion of the updraft gasification with an adiabatic methanation and the tu-
bular cell design. The high system efficiency results from the high cold
gas efficiency of the updraft gasification process which already yields a
producer gas with considerable internal reforming potential. The latter is
295
7 Concluding remarks
296
7 Concluding remarks
fuel cell balance of plant costs. For the fluidized bed steam gasification
based systems, slightly lower specific plant costs are determined mainly
because of the comparably high power outputs as result of the undiluted
character of the corresponding producer gases. The lowest specific
plant costs are established for the updraft gasification based systems.
This is mainly because of the systems simplicity resulting in low direct
plant costs and minor auxiliary power requirements.
The plant costs were related to the corresponding system efficiencies to
yield power production cost (PPC) estimates. The basis for the calcula-
tion of the PPCs was a detailed cost analysis for a given planning hori-
zon including costs not only for operation and maintenance, feedstock
and utilities but also capital costs resulting from plant costs depreciation
and interests. The net value of all costs was evenly distributed through-
out the planning horizon using the annuity method.
The PPCs analysis revealed that the fuel costs amount to on average
40 % of the PPCs, the remainder being capital costs. Thus, specific
plant cost reduction is considerably more effective on reducing the
PPCs than increasing the system efficiency.
An increase of the operational voltage from 0.6 V to 0.7 V yields a sig-
nificant increase of the SOFC conversion efficiency which in turn entails
a gain of the net system efficiency from 32.1 to 37.1 % for the updraft
gasification based system with adiabatic methanation. However, the
PPCs increase from 0.1154 €/kWhel to 0.1912 €/kWhel too, mainly due to
considerably reduced power output of the SOFC.
The reduction of auxiliary power requirements in contrast leads to an in-
crease of the system efficiency and a decrease of the specific plant
costs, thus considerably lowering the PPCs.
Finally, the revenues from heat sales are crucial for the economical vi-
ability of the investigated B-IGFC systems without bottoming cycles.
297
7 Concluding remarks
Figure 7-1:
Hot gas filtration system
To achieve a minimum fly ash load in the filtered producer gas, the fil-
tration temperature must be kept below the ash melting point. Usually
temperatures around 600 to 650 °C are chosen for this purpose.
Means for active cooling may have to be added to the hot gas filtra-
tion unit.
Application of fine meshed CPO monoliths
Measurements of the producer gas composition downstream a CPO
monoliths featuring 400 channels per square inch (cpsi) via a capil-
lary showed that tar conversions over 90 % can be achieved. The
application of such fine meshed monoliths in long-term tests was
however not possible until now due to blocking issues resulting from
the moderate particle removal efficiency of the used cyclone. After
successful commissioning of the hot gas particle filtration unit, the
currently used 100 cpsi CPO monolith should be replaced by a mono-
lith with 400 cpsi.
298
7 Concluding remarks
300
301
302
Notation
Notation
Latin symbols
A Area [m2]
AmP Amortization period [a]
AN Annuity factor [%]
c Carbon mass fraction [-]
C Cost [€]
cp Heat capacity [J/(mol K)]
CP Length of construction period [a]
d Diameter [m]
D Diffusion coefficient [m2/sec]
DA Depreciation/Amortization costs [€]
DeP Depreciation period [a]
DK Knudsen diffusion coefficient [m2/sec]
Dm Molecular diffusion coefficient [m2/sec]
DP Dept portion [%]
DPC Direct plant costs [€]
DR Degradation rate [%/(1000 h)]
E Voltage [V]
Eact Activation energy for exchange current density [J/mol]
EqH2 Hydrogen equivalent molar flow [mol/sec]
F Faraday constant [C/mol]
G Gibbs enthalpy [J/mol]
h Height, Hydrogen mass fraction [m], [-]
H Enthalpy [J/mol]
I Current density [A/m2]
I0 Exchange current density [A/m2]
IF Inflation rate [%]
IPC Indirect plant costs [€]
IR Interest rate [%]
Kp Equilibrium constant [-]
L Length of cell-subunit [m]
303
Notation
l Length [m]
lc Length of planar cell [m]
lD8 Length of Delta8 cell [m]
LHV Lower heating value [J/mol]
lt Length of tubular cell [m]
M Molecular weight, Water load [kg/mol], [%]
m Mass [kg]
n Number of electrons, Nitrogen mass fraction [-]
nch Number of channels of planar cell [-]
NCIR Nominal capital interest rate [%]
ndot Molar flow [mol/sec]
nRE Number of repeating elements [-]
ntri Number of triangular tubes in Delta8 cell [-]
Nu Nusselt number [-]
o Oxygen mass fraction [-]
p Total or partial pressure [N/m2]
P Power [W]
p Price [€]
Pel Electrical Power (DC) [W]
PF Plant factor [%]
PH Planning horizon [a]
PPC Power production costs [€/kWHel]
Qdot Heat flux [W/m]
R Resistance or Ideal gas constant [Ω], [J/(mol K)]
r Radius [m]
rdl-reac Diffusion limited, length specific reaction rate [mol/(sec m)]
ri Inner radius [m]
rj Reaction rate of reaction j [mol/(sec m)]
rm Middle radius [m]
ro Outer radius [m]
ro Reaction order [-]
RoE Return on equity [%]
s Sulfur mass fraction [-]
SoP Start of production [-]
T Temperature [K]
TadtK Gas temperature in air delivery tube [K]
304
Notation
Greek symbols
α Upper triangular half angle [°]
α Convective heat exchange coefficient [W/(m2 K)]
αinsul Heat transfer coefficient through insulation [W/(m2 K)]
β Transfer coefficient [-]
βx,diff Mass transfer coefficient of specie x [m/sec]
βx,diff-reac Diffusion limited reaction conversion coefficient of specie x
βx,reac Reaction conversion coefficient of specie x
δ Thickness of component [m]
ΔH Heat of reaction [J/mol]
ε Porosity of porous media [-]
γ Pre-exponential factor [A/m2]
η Efficiency [%]
ηact Activation polarization voltage loss [V]
ηdiff Diffusion polarization voltage loss [V]
ηohm Ohmic voltage loss [V]
λ Air-to-fuel ratio [-]
λan Thermal conductivity of anode gas [W/(m K)]
λca Thermal conductivity of cathode gas [W/(m K)]
305
Notation
306
Notation
Abbreviations
AC Alternating current
ADT Air delivery tube
AEC Anode-electrolyte-cathode assembly
AF Air-to-fuel ratio
AFC Alkaline fuel cell
APU Auxiliary power unit
ATR Autothermal reforming
BHPR Biomass Heatpipe Reformer
B-IGFC Biomass Integrated Gasification Fuel Cell System
BMT Benchmark test
BoP Balance of plant
BtL Biomass to Liquids
CECPI Chemical Engineering Magazine Plant Cost Index
CFD Computational fluid dynamics
CGO Gadolinium doped ceria
CHP Combined heat and power
CPO Catalytic partial oxidation
cpsi Channels per square inch
CSTR Continuous stirred tank reactor
CV Caloric value, Control volume
D8 Delta8 cell design by the Siemens AG
daf Dry and ash free
DC Direct current
DMFC Direct methanol fuel cell
dtf Dry, tar and ash free
FC Fuel cell
FEM Finite element method
FICFB Fast Internally Circulating Fluidized Bed
FVM Finite volume method
GC Gas chromatography
GT Gas turbine
HDP High power density
HGF Hot gas filtration
308
Notation
309
310
References
References
[1] Novatlantis, Smarter living - Generating a new understanding for natural resources as
the key to sustainable development - the 2000-watt society, Dübendorf, Switzerland,
2005.
[2] A. Wokaun, Erneuerbare Energie, Teubner Verlag, Stuttgart, Leipzig, Germany, 1999.
[3] World Energy Outlook 2004, International Energy Agency, Paris, France, 2004.
[5] M. Gronli, A theoretical and experimental study of the thermal degradation of biomass,
Norwegian University of Science and Technology, Trondheim, Norway, 1996.
[10] N. Arnstein, Experimental investigation of Solid Oxide Fuel Cells using biomass gasi-
fication producer gases, Norwegian University of Science and Technology, Trondheim,
Norway, 2005.
[17] User Documentation for Aspen Plus, Version 12.1, ASPEN Tech, Cambridge, United
States of America, 2002.
311
References
[19] A. Kaupp, J. Goss, Small scale gas producer-engine systems, German Appropriate
Technology Exchange (GATE), Eschborn, Germany, 1984.
[28] B. Teislev, Wood-chips gasifier combined heat and power, Babcock & Wilcox Volund
R&D Centre, Kolding, Denmark, 2001.
[34] D. J. Stevens, Hot Gas Conditioning: Recent Progress with Larger-Scale Biomass
Gasification Systems, National Renewable Energy Laboratory, Golden, United States of
America, 2001.
312
References
[38] E. Heinrich, The status of the FZK concept of biomass gasification, in: Proceedings
of the 2nd European Summer School on Renewable Motor Fuels, Warsaw, Poland, 2007,
29th to 31st August.
[39] O. Schulze, Advanced gas cleaning for biomass gasification, in: Proceedings of the
IEA Workshop 5, Dresden, Germany, 2006, 12th to 14th June.
[40] M. Seemann, Methanation of biosyngas in a fluidized bed reactor, Swiss Federal In-
stitute of Technology, Zurich, Switzerland, 2007.
[47] A. Wojcik, H. Middleton, I. Damopoulos, J. Van herle, J. Power Sources, 118 (2003)
342-348.
[51] T. Sanderson, An updated assesment of the prospects for fuel cells in stationary
power and CHP, DTI, Oxfordshire, England, 2005.
313
References
[54] R. Coll, J. Salvado, X. Farriol, D. Montane, Fuel Process. Technol., 74 (2001) 19-31.
[55] D. Wang, D. Montane, E. Chornet, Appl. Catal. A-Gen., 143 (1996) 245-270.
[58] J. Larminie, A. Dicks, Fuel Cell Systems Explained, John Wiley & Sons, Ltd., Chices-
ter, England, 2000.
[59] S. D. Vora, SECA Program at Siemens Westinghouse, in: Proceedings of the 6th
Annual SECA Workshop, Pacific Grove, United States of America, 2005, 18th to 21st
April.
[61] S. Shaffer, Development Update on Delphi's Solid Oxide Fuel Cell System, in: Pro-
ceedings of the 6th Annual SECA Workshop, Pacific Grove, United States of America,
2005, 18th to 21st April.
[62] M. Klemm, Heissentteerung von Brenngas aus der Vergasung von Biomasse durch
katalytische partielle Oxidation, Fortschritt-Berichte VDI Reihe 6 Nr. 525, Düsseldorf,
Germany, 2005.
[64] T. Miyazawa, T. Abe, K. Kunimori, K. Tomishige, J. Jpn. Pet. Inst, 48 (2005) 162-172.
[65] U. Bossel, The Birth of the Fuel Cell (1835-1845). Complete Correspondence be-
tween Christian Friedrich Schönbein and William Robert Grove, European Fuel Cell Fo-
rum, Lucern, Switzerland, 2000.
[67] R. Bosch, Perspectives on Fuel Cells vs. Incumbent Technologies, in: Proceedings
of the 2005 Fuel Cell Seminar, Palm Springs, United States of America, 2005, 14th to
18th November.
314
References
[71] Fuel Cell Handbook (Sixth Edition), DOE/NETL, Morgantown, West Virginia, United
States of America, 2002.
[74] A. Baker, Fuel Cell Today Market Survey: Large Stationary Applications,
www.fuelcelltoday.com, 2005.
315
References
alloys, in: Proceedings of the 2002 ASM Fall Meeting, Columbus, United States of Amer-
ica, 2002, 7th to 10th October.
[84] C. Voisard, Das planare Sulzer Hexis SOFC Konzept, in: Proceedings of the DGM-
Fortbildungsseminar Werkstofffragen der Hochtemperatur-Brennstoffzelle (SOFC), Jülich,
Germany, 2005, 20th to 22nd April.
[86] L. Jones, Pressurisation of IP-SOFC technology for second generation hybrid appli-
cation, DTI, Oxfordshire, England, 2005.
[88] R. A. George, SECA Project at Siemens Westinghouse, in: Proceedings of the 3rd
Annual Solid State Energy Conversion Alliance (SECA) Workshop, Washington D.C.,
United States of America, 2002, 21st to 22nd March.
[89] N. Bessette, Status of the Acumentrics SOFC Program, in: Proceedings of the 5th
SECA Annual Workshop, Boston, United States of America, 2004, 11th to 13th May.
[90] P. Patel, Thermally Integrated High Power Density SOFC Generator, in: Proceed-
ings of the 6th SECA Annual Workshop, Pacific Grove, United States of America, 2005,
18th to 21st April.
[93] C. Wunderlich, M. Boltze, Auxiliary Power Unit Development - How the System Con-
cept interacts with the Stack Concept, in: Proceedings of the 2005 Fuel Cell Seminar,
Palm Springs, United States of America, 2005, 14th to 18th November.
[94] D. Norrick, Cummins Power Generation SECA Program Phase 1 Results and Ex-
perience, in: Proceedings of the 2005 Fuel Cell Seminar, Palm Springs, United States of
America, 2005, 14th to 18th November.
316
References
[98] W. Z. Zhu, S. C. Deevi, Mater. Sci. Eng. A-Struct. Mater. Prop. Microstruct. Process.,
362 (2003) 228-239.
[100] D. Simwonis, F. Tietz, D. Stover, Solid State Ion., 132 (2000) 241-251.
[101] J. Helbig, U. Schönholzer, Grundzüge der Keramik - Skript zur Vorlesung In-
genieurkeramik I, Swiss Federal Institute of Technology, Zürich, Switzerland, 2001.
[107] M. V. Twigg, Catalyst Handbook, Wolfe Publishing Ltd., Frome, England, 1989.
[112] E. Achenbach, IEA Programm on R,D&D on advanced fuel cells: Final Report of
Activity A2, (1996).
[113] A. Selimovic, M. Kemm, T. Torisson, M. Assadi, J. Power Sources, 145 (2005) 463-
469.
[114] Structural limitations in the scale-up of anode supported SOFCs, Final report to
DOE NETL, Cambridge, United States of America, 2002.
317
References
[116] U. Hennings, M. Brune, R. Reimert, GWF Gas Erdas, 145 (2004) 92-97.
[119] L. Aguilar, S. Zha, Z. Cheng, J. Winnick, M. Liu, J. Power Sources, 135 (2004) 17-
24.
[122] W. C. Li, H. L. Bai, J. N. Hsu, S. N. Li, C. C. Chen, Ind. Eng. Chem. Res., 47 (2008)
1501-1505.
[123] R. Sime, S. Stucki, S. Biollaz, A. Wiasmitinow, Linking wood gasification with SOFC
hybrid processes, in: Proceedings of the 5th European SOFC Forum, Lucerne, Switzer-
land, 2002, 2nd to 5th July.
[125] F. P. Nagel, M. Künstle, S. Biollaz, Link-up of a SOFC to a Wood Gasifier, in: Pro-
ceedings of the 14th European Biomass Conference and Exhibition: Biomass for Energy,
Industry and Climate Protection, Paris, France, 2005, 17th to 21st October.
[128] K. Eguchi, H. Kojo, T. Takeguchi, R. Kikuchi, K. Sasaki, Solid State Ion., 152
(2002) 411-416.
318
References
[133] D. Wang, S. Czernik, D. Montane, M. Mann, E. Chornet, Ind. Eng. Chem. Res., 36
(1997) 1507-1518.
[138] L. de Sousa, Gasification of wood, urban wastewood (Altholz) and other wastes in
a fluidized bed reactor, Swiss Federal Institute of Technology, Zurich, Switzerland, 2001.
[140] C. Stiller, Design, Operation and Control Modelling of SOFC/GT Hybrid Systems,
Norwegian University of Science and Technology, Trondheim, Norway, 2006.
[141] A. Selimovic, Modelling of Solid Oxide Fuel Cells Applied to the Analysis of Inte-
grated Systems with Gas Turbines, Lund University, Lund, Sweden, 2002.
[145] T. Reed, S. Gaur, A Survey of biomass gasification 2001 - Gasifier projects and
manufacturers around the world, Solar Energy Research Institute, Golden, United States
of America, 2001.
319
References
[148] C. Pitta, Investigation of tar reforming employing a catalytic partial oxidation cata-
lyst, Swiss Federal Institute of Technology, Zurich, Switzerland, 2007.
[155] B. Thorud, Dynamic Modelling and Characterisation of a Solid Oxide Fuel Cell Inte-
grated in a Gas Turbine Cycle, Norwegian University of Science and Technology, Trond-
heim, Norway, 2005.
[157] E. Liese, R. S. Gemmen, J. Eng. Gas Turb. Power, 127 (2005) 86-90.
[159] Y. Yi, A. D. Rao, J. Brouwer, G. S. Samuelsen, J. Power Sources, 144 (2005) 67-
76.
320
References
[169] D. Froning, U. Reimer, A. Gubner, Combined Treatment of the Stack and the Sys-
tems Level in SOFC - Modelling, in: Proceedings of the 3rd Fuel Cell Research Sympo-
sium - Modelling and Experimental Validation, Dübendorf, Switzerland, 2006, 16th to
17th March.
[170] M. Roos, E. Batawi, U. Harnisch, T. Hocker, J. Power Sources, 118 (2003) 86-95.
[173] T. Araki, T. Ohba, S. Takezawa, K. Onda, Y. Sakaki, J. Power Sources, 158 (2006)
52-59.
[174] P. Costamagna, A. Selimovic, M. Del Borghi, G. Agnew, Chem. Eng. J., 102 (2004)
61-69.
[176] L. T. Lim, D. Chadwick, L. Kershenbaum, Ind. Eng. Chem. Res., 44 (2005) 9609-
9618.
[179] D. Larrain, J. Van Herle, F. Maréchal, D. Favrat, J. Power Sources, 131 (2004)
304-312.
321
References
[180] N. Autissier, D. Larrain, J. Van Herle, D. Favrat, J. Power Sources, 131 (2004) 313-
319.
[183] S. Cordiner, M. Feola, V. Mulone, F. Romanelli, Appl. Therm. Eng., 27 (2007) 738-
747.
[189] N. Autissier, F. Palazzi, F. Marechal, J. van Herle, D. Favrat, J. Fuel Cell Sci.
Technol., 4 (2007) 123-129.
322
References
[197] N. Q. Minh, T. Takahashi, Science and technology of ceramic fuel cells, Elsevier
Science, Amsterdam, The Netherlands, 1995.
[200] A. Solheim, K. Nisancioglu, Resistance and Current distribution in Fuel Cell Ele-
ments, in: Proceedings of the Second International Symposium on SOFC, Greece, Ath-
ens, 1991, 2nd to 5th July.
[206] A. L. Lee, R. F. Zabransky, W. J. Huber, Ind. Eng. Chem. Res., 29 (1990) 766-773.
[210] F. Lange, M. Ise, W. Kleinlein, B. Schricker, Delta 8 - A new cell design for cost op-
timized generators, in: Proceedings of the Fuel Cell Seminar & Exposition, San Antonio,
USA, 2007, 15th to 19th October.
[211] K. Huang, Development of Delta-Type SOFCs at Siemens Stationary Fuel Cells, in:
Proceedings of the Fuel Cell Seminar & Exposition, San Antonio, United States of Amer-
ica, 2007, 15th to 19th October.
323
References
[214] D. E. Garrett, Chemical Engineering Economics, Van Nostrand Reinhold, New York,
United States of America, 1989.
[220] W. A. Surdoval, The U. S. Department of Energy Fossil Energy Fuel Cell Program -
Solid State Energy Conversion Alliance: Goals and Challenges, in: Proceedings of the
8th SECA Workshop, San Antonio, United States of America, 2007, 7th to 9th August.
[222] S. Jimenez, J. Ballester, Combustion Science And Technology, 178 (2006) 655-
683.
324
References
[228] J. Karl, Biocellus - Highly efficient SOFC systems with indirect gasification, in: Pro-
ceedings of the Bio-energy workshop EU-Russia, Moscow, Russia, 2004, 7th to 8th Oc-
tober.
[229] EEG, "Erneuerbare-Energien-Gesetz vom 21. Juli 2004 (BGBl. I S. 1918), zuletzt
geändert durch Artikel 1 des Gesetzes vom 7. November 2006 (BGBl. I S. 2550)".
325
326
Curriculum vitae
Personal Data
Surname Nagel
Given names Florian-Patrice
Date of birth 5th December 1977
Birthplace Berlin
Nationality German
Post-graduate experience
Education
Under-graduate experience
01/2003 – 03/2003 Intern (LAE Lahmeyer Agua y Energia S.A., Lima, Peru)
327
Publications
Conference pro- F. P. Nagel, M. Künstle, S.Biollaz: Link-up of a solid oxide fuel cell with a wood gasifier;
Proceedings of the 14th European Biomass Conference and Exhibition: Biomass for En-
ceedings th st
ergy, Industry and Climate Protection; Paris, France, 17 -21 October 2005
F. Nagel, M. Jenne, S. Biollaz, S. Stucki: Link-up of a SOFC with an Updraft Wood Gasifier
via Hot Gas Processing; Proceedings of the 2006 Fuel Cell Seminar: Fuel Cells in Para-
th th
dise - 30th Anniversary; Honolulu, HI, United States of America, 13 -17 November 2006
S. Biollaz, F.P. Nagel, K. Dennerlein, H. Landes, W. Straub, J. Karl: Results from long term
testing of tubular SOFC with woodgas; Proceedings of the 16th European Biomass Con-
nd th
ference and Exhibition: From Research to Industry and Markets, Valencia, Spain, 2 -6
June 2008
Oral presentations "Link-up of a SOFC with an updraft-wood gasifier via hot gas processing (B-IGFC)" held at
th
the 2006 Fuel Cell Seminar in Honolulu, HI, USA, the 16 November 2006
"Link-up of a 1kW-SOFC with an updraft-wood gasifier via hot gas processing" held at the
th
15th European Biomass Conference and Exhibition in Berlin, Germany,the 8 May 2007
Scientific papers F.P. Nagel, T.J. Schildhauer, S.M.A. Biollaz, A. Wokaun: Performance comparison of pla-
nar, tubular and Delta8 solid oxide fuel cells using a generalized finite volume model; Jour-
(peer-reviewed) nal of Power Sources 184 (2008) 143–164
F.P. Nagel, T.J. Schildhauer, S.M.A. Biollaz, S. Stucki: Charge, mass and heat transfer
interactions in SOFCs operated with different fuel gases - A Sensitivity analysis; Journal of
Power Sources 184 (2008) 129–142
F.P. Nagel, T.J. Schildhauer, S.M.A. Biollaz, A.Schuler: The impact of sulfur on the per-
formance of a SOFC system operated with hydrocarboneous fuel gas; Journal of Power
Sources (2009), doi:10.1016/j.jpowsour.2008.12.092
F.P. Nagel, S. Ghosh, C. Pitta, T.J. Schildhauer, S.M.A. Biollaz: Biomass integrated gasifi-
cation fuel cell systems - Concept development based on experimental results and long-
term testing; Biomass and Bioenergy, in internal review process
F.P. Nagel, T.J. Schildhauer, S.M.A. Biollaz: Biomass-Integrated gasification fuel cell sys-
tems - Part 1: Definition of systems and technical analysis; International Journal of Hydro-
gen Energy, submitted
F.P. Nagel, T.J. Schildhauer: Improving lumped SOFC models for system simulations by
parameterization of detailed model results; Journal of Power Sources, in preparation
Patent applications F.P. Nagel, T.J. Schildhauer, M. Jenne, S. Biollaz, Verfahren und Anlage zur Verstromung
fester Biomasse, WO 2008/055591 A2
328