10.01.2013 Views

Abstracts S S P - SSPC-15 - California Institute of Technology

Abstracts S S P - SSPC-15 - California Institute of Technology

Abstracts S S P - SSPC-15 - California Institute of Technology

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

S S P C <strong>15</strong><br />

The <strong>15</strong> th International Conference on<br />

Solid State Protonic Conductors<br />

<strong>Abstracts</strong><br />

August <strong>15</strong> - 19, 2010<br />

Santa Barbara, USA


The <strong>15</strong> th International Conference on<br />

Solid State Protonic Conductors<br />

August <strong>15</strong>, 2010 – August 19, 2010<br />

Conference Chairs<br />

Sossina M. Haile, Chair<br />

<strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong><br />

Sangtae Kim, Co-Chair<br />

University <strong>of</strong> <strong>California</strong>, Davis<br />

Stephen J. Paddison, Co-Chair<br />

University <strong>of</strong> Tennessee<br />

Lutgarde De Jonghe, Co-Chair<br />

University <strong>of</strong> <strong>California</strong>, Berkeley<br />

Lawrence Berkeley National Labs<br />

International Advisory Committee and <strong>SSPC</strong> Board<br />

John T.S. Irvine, St. Andrews University<br />

Sossina M. Haile, <strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong><br />

M. Saiful Islam, University <strong>of</strong> Bath<br />

Deborah Jones, University <strong>of</strong> Montpellier<br />

Junichi Kawamura, Tohoku University<br />

Klaus-Dieter Kreuer, Max Planck <strong>Institute</strong> for Solid State Research<br />

Truls Norby, University <strong>of</strong> Oslo<br />

Stephen J. Paddison, University <strong>of</strong> Tennessee<br />

Eivind M. Skou, University <strong>of</strong> Southern Denmark<br />

Robert C.T. Slade, University <strong>of</strong> Surrey<br />

Josh O. Thomas, University <strong>of</strong> Upsala<br />

Shu Yamaguchi, University <strong>of</strong> Tokyo


Local Organizing Committee<br />

Sossina M. Haile<br />

Áron Varga (co-chair)<br />

Yoshihiro Yamazaki (co-chair)<br />

Ayako Ikeda<br />

Mary W. Louie<br />

Chatr Panithipongwut<br />

<strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>


Financial Support:<br />

Acknowledgements<br />

The US National Science Foundation, Division <strong>of</strong> Materials Research<br />

US Army Research Office, Chemical Sciences<br />

The US Department <strong>of</strong> Energy, Office <strong>of</strong> Basic Energy Sciences<br />

Dokiya Symposium Fund, The Electrochemical Society (ECS)<br />

The Electrochemical Society, High Temperature Material Division<br />

Supporting Organizations:<br />

The Electrochemical Society<br />

The Materials Research Society


General Information<br />

Conference Location<br />

Loma Pelona Center<br />

The University <strong>of</strong> <strong>California</strong>, Santa Barbara, CA 93106, USA<br />

Tel: +1 (805)893-6161<br />

Alternate: +1 (805)698-2992<br />

E-mail: sspc@caltech.edu<br />

Meeting Rooms<br />

Oral Presentations: Room 1108<br />

Poster Presentations: Room 1100A (continuous access 8/<strong>15</strong> – 8/17)<br />

Conference Registration-Desk Hours<br />

8/<strong>15</strong> (Sun) <strong>15</strong>:00-20:00<br />

8/16 (Mon) 7:30-16:00<br />

8/17 (Tue) 7:30-10:00<br />

Welcome Reception<br />

16:00-18:00 Sunday, <strong>15</strong> August<br />

Loma Pelona Center 1108 Patio<br />

Meals<br />

Breakfast: 07:<strong>15</strong>-08:<strong>15</strong><br />

Lunch: 12:30-13:45 (except Wednesday: 11:30-13:00)<br />

Dinner 18:00-19:<strong>15</strong> (except Wednesday banquet)<br />

Carillo Dining Commons<br />

Excursion<br />

13:00-18:00 Wednesday, 18 August<br />

Santa Ynez valley wine tour<br />

*pickup/drop-<strong>of</strong>f (1:<strong>15</strong>/5:45pm) Manzanita Village bus loop<br />

Banquet<br />

18:30-21:00 Wednesday, 18 August<br />

Las Encinas Quad, Manzanita Village


Loma Pelona


The <strong>15</strong> th International Conference on<br />

Solid State Protonic Conductors<br />

Program<br />

Sunday, August <strong>15</strong><br />

Registration <strong>15</strong>:00-20:00<br />

Outside Loma Pelona Center, University <strong>of</strong> <strong>California</strong>, Santa Barbara<br />

Welcome Reception 16:00-18:00<br />

Loma<br />

Pelona Center 1108 Patio<br />

Opening Remarks 8:45-8:55 Sossina Haile<br />

Plenary Lecture<br />

Monday, August 16<br />

8:55-9:55 …………………………………………………………………………………………………….. 1<br />

Oxide protonics: Electronic structure and acid base reactions in bulk and surface<br />

Shu Yamaguchi<br />

(C<strong>of</strong>fee Break)<br />

Oral Session (1) Chairperson: Steve Paddison<br />

O 01(K) 10:<strong>15</strong>-11:00 ………………………………………………………………………………………... 2<br />

Ab initio molecular dynamics studies <strong>of</strong> proton structure and transport in solid acids and acid hydrate<br />

crystals<br />

Mark Tuckerman<br />

O 02 (I) 11:00-11:35 ……………………………………………………………………………………….. 3<br />

Polyelectrolyte multilayers for electrochemical energy applications<br />

Avni A. Argun and P. T. Hammond<br />

O 03 11:35-11:55 …………………………………………………………………………………………… 4<br />

About the evolution <strong>of</strong> the microstructures <strong>of</strong> ionomers and polyelectrolytes for PEM fuel cell<br />

applications<br />

Klaus-Dieter Kreuer, C. C. de Araujo, G. Portale, U. Klock and J. Maier<br />

O 04 11:55-12:<strong>15</strong> …………………………………………………………………………………………… 5<br />

Coherent proton transport in nafion<br />

George Reiter<br />

(Group Photo)<br />

(Lunch Break)<br />

‐ i ‐<br />

Page


Oral Session (2) Chairperson: Tetsuya Uda<br />

O 05 (K) 14:00-14:45 ……………………………………………………………………………………….. 6<br />

NMR studies <strong>of</strong> local structure, cation ordering and protonic conduction in perovskites<br />

L. Buannic, F. Blanc, D. Middlemiss, and Clare P. Grey<br />

O 06 (I) 14:45-<strong>15</strong>:20 ……………………………………………………………………………………….. 7<br />

Recent developments in solid acid fuel cells<br />

Calum R.I. Chisholm<br />

O 07 <strong>15</strong>:20-<strong>15</strong>:40 ……………………………………………………………………………………………. 8<br />

Local structure and hydride-ion transport in MgY2H8<br />

Hitoshi Takamura, K. Kurosu, A. Hatakeyama and H. Maekawa<br />

(C<strong>of</strong>fee Break)<br />

O 08 (K) 16:00-16:45 ………………………………………………………………………………………… 9<br />

Improving the proton conductivity <strong>of</strong> yttrium-doped barium zirconate electrolytes towards the<br />

development <strong>of</strong> intermediate temperature solid oxide fuel cells<br />

Enrico Traversa<br />

O 09 16:45-17:05 ……………………………………………………………………………………………. 10<br />

Existence <strong>of</strong> equilibrium mixture between hydrated and unhydrated proton conducting perovskites and<br />

possible implications<br />

John T.S. Irvine, M. C. Verbraeken, H. Viana, A. K. Azad, I. Ahmed and S. Eriksson<br />

O 10 17:05-17:25 …………………………………………………………………………………………….. 11<br />

Local structure and proton dynamics in In-doped BaZrO3<br />

Maths Karlsson, A. Matic, I. Ahmed, C. S. Knee and S. Eriksson<br />

O 11 17:25-17:45 ……………………………………………………………………………………………. 12<br />

Grain boundary conduction in polycrystalline Y-doped BaZrO3: Space charge analysis<br />

Sangtae Kim, C. Chen and C. Danel<br />

Poster Session 20:00-22:30<br />

Loma Pelona Center 1100A<br />

‐ ii ‐


Oral Session (3) Chairperson: John Irvine<br />

Tuesday, August 17<br />

O 12 (K) 8:30-9:<strong>15</strong> ………………………………………………………………………………………….. 13<br />

Atomic-scale insights into proton-conducting oxides: Defects, dopants and transport<br />

M. Saiful Islam<br />

O 13 9:<strong>15</strong>-9:35 ………………………………………………………………………………………………. 14<br />

DFT-MD study <strong>of</strong> local structures and proton transport in disordered brownmillerite-based compounds<br />

Karsten Rasim, F. Boucher, O. Joubert, P. Baranek and M. Marrony<br />

O 14 9:35-9:55 ……………………………………………………………………………………………. <strong>15</strong><br />

Ab initio studies <strong>of</strong> the effect <strong>of</strong> hydrophobic environment on proton transfer in model perfluorsulfonic<br />

acid systems<br />

Bradley F. Habenicht and S. J. Paddison<br />

(C<strong>of</strong>fee Break)<br />

O <strong>15</strong> (K) 10:<strong>15</strong>-11:00 ……………………………………………………………………………………… 16<br />

Analysis <strong>of</strong> transport through mixed proton, oxygen ion, and electron (hole) conductors: Fuel cell and<br />

electrolyzer modes<br />

Anil Virkar<br />

O 16 (I) 11:00-11:35 ………………………………………………………………………………………… 17<br />

Hydrogen in nominally anhydrous minerals: analysis and implications<br />

George R. Rossman<br />

O 17 (K) 11:35-12:20 ……………………………………………………………………………………… 18<br />

The dynamical behaviour <strong>of</strong> water and protons in perfluorinated membranes and surfactants<br />

Sandrine Lyonnard, C. Cailleteau, H. Mendil-Jakani, S. Mossa, G. Gebel, A. Guillermo, B. Frick and J. Ollivier<br />

(Lunch Break)<br />

Oral Session (4) Chairperson: Klaus-Dieter Kreuer<br />

O 18 (K) 14:00-14:45 ……………………………………………………………………………………… 19<br />

Tailored tungsten oxide nanoparticles produced with hot wire chemical vapor deposition for protonic<br />

electrochromic and fuel cell applications<br />

Anne C. Dillon, S. Lee, R. Deshpande, K. E. Hurst, S. Kocha, V. R. Anderson and S. M. George<br />

O 19 (I) 14:45-<strong>15</strong>:20 ………………………………………………………………………………………… 20<br />

Uphill gas permeation due to coupled transport in a mixed proton, oxygen vacancy and hole<br />

conducting perovskite membrane<br />

M. Sanders, J. Tong, H. Zhu, R. Kee and Ryan O’Hayre<br />

O 20 <strong>15</strong>:20-<strong>15</strong>:40 ……………………………………………………………………………………….….... 21<br />

Proton migration at Σ5 (310)/[001] tilt grain boundary in Y-doped BaZrO3<br />

Cheonan Kim<br />

(C<strong>of</strong>fee Break)<br />

‐ iii ‐


O 21 (K) 16:00-16:45 ………………………………………………………………………………………. 22<br />

On the path dependence <strong>of</strong> the open-cell voltage <strong>of</strong> a galvanic cell involving a ternary or higher<br />

compound with multiple mobile ionic species under multiple chemical potential gradients<br />

Han-Ill Yoo and M. Martin<br />

O 22 16:45-17:05 ……………………………………………………………………………………………. 23<br />

Development <strong>of</strong> proton conducting SOFCs based on LaNbO4 electrolytes – Status in Norway<br />

Anna. Magrasó, M. L. Fontaine, G. E. Syvertsen, Y. Larring, H. L. Lein, T. Grande, Reidar Haugsrud and T. Norby<br />

O 23 17:05-17:25 ……………………………………………………………………………………………. 24<br />

Charge transfer protonation <strong>of</strong> a manganese-doped strontium zirconate<br />

Hiroshige Matsumoto, T. Sakai, Y. Kawasaki, Y. Sato and T. Ishihara<br />

O 24 17:25-17:45 ……………………………………………………………………………………………. 25<br />

Electronic conduction <strong>of</strong> Sr-doped Ce-La monazite ceramics<br />

Hannah L. Ray and L. C. De Jonghe<br />

Poster Session 20:00-22:30<br />

Loma Pelona Center 1100A<br />

‐ iv ‐


Oral Session (5) Chairperson: Noriko Sata<br />

Wednesday, August 18<br />

O 25 (K) 8:30-9:<strong>15</strong> ………………………………………………………………………………………….. 26<br />

Interfacial proton dynamics in PEM: Theory and molecular modeling<br />

A. Roudgar, S. Vartak and Michael Eikerling<br />

O 26 9:<strong>15</strong>-9:35 ………………………………………………………………………………………………. 27<br />

Proton transport mechanism and pathways in solid acids from experiment and theory<br />

Boris Merinov<br />

O 27 9:35-9:55 ………………………………………………………………………………………………. 28<br />

Adsorption processes related to the surface properties <strong>of</strong> LaNbO4<br />

Kianoosh Hadidi, O.M. Løvvik and T. Norby<br />

(C<strong>of</strong>fee Break)<br />

O 28 (K) 10:<strong>15</strong>-11:00 ……………………………………………………………………………………… 29<br />

High temperature protonic conduction in rare earth phosphates and borates<br />

Koji Amezawa, H. Takahashi, A. Unemoto, H. Kuwabara, N. Kitamura and T. Kawada<br />

O 29 11:00-11:20 ……………………………………………………………………………………………. 30<br />

Thermodynamics and proton conductivity trends in fluorite-pyrochlore structures: A study on<br />

lanthanum cerate<br />

Vasileios Besikiotis, T. S. Bjørheim, S. Ricote, C. Kjølseth, R. Haugsrud and T. Norby<br />

O 30 11:20-11:40 ……………………………………………………………………………………………. 31<br />

Synthesis <strong>of</strong> highly Sr-doped lanthanum orthophosphate and polyphosphate in phosphoric acid<br />

solutions<br />

Naoyuki Hatada, A. Kuramitsu, Y. Nose and T. Uda<br />

(Lunch Break)<br />

Excursion 13:00-18:00<br />

Santa Ynez valley wine tour<br />

*pickup/drop-<strong>of</strong>f (1:<strong>15</strong>/5:45pm) Manzanita Village bus loop<br />

Banquet 18:30-21:00<br />

Las Encinas Quad, Manzanita Village<br />

‐ v ‐


Oral Session (6) Chairperson: Ryan O’Hayre<br />

Thursday, August 19<br />

O 31 (K) 8:30-9:<strong>15</strong> …………………………………...……………………………………………………… 32<br />

Characterization <strong>of</strong> structural change in PLD-fabricated SrZrO3 thin films<br />

Noriko Sata, S. Tamura, Y. Nagao, H. Kageyama, K. Handa, K. Nomura, F. Iguchi and H. Yugami<br />

O 32 9:<strong>15</strong>-9:35 ………………………………………………………………………………………………. 33<br />

Water uptake and conduction property <strong>of</strong> nano-grained yttria-doped zirconia<br />

Y. Akao, Shogo Miyoshi, N. Kuwata, J. Kawamura, Y. Oyama and Shu Yamaguchi<br />

O 33 9:35-9:55 ………………………………………………………………………………………………. 34<br />

Proton conductivity in acceptor doped LaVO4<br />

Morten Huse, T. S. Bjørheim and R. Haugsrud<br />

(C<strong>of</strong>fee Break)<br />

O 34 (I) 10:<strong>15</strong>-10:50 ………………………………………………………………………………………… 35<br />

Quasielastic neutron scattering (QENS) applied to proton dynamics in bulk and confinement – the<br />

case <strong>of</strong> phosphoric acid<br />

Bernhard Frick, A. Thomas, S. Lyonnard, L.Vilciauskas and K.D. Kreuer<br />

O 35 (I) 10:50-11:25 ………………………………………………………………………………………… 36<br />

Ambient and high-pressure synchrotron x-ray diffraction studies <strong>of</strong> heating-induced structural<br />

modifications in phosphate-based solid acids<br />

Cristian E. Botez<br />

O 36 (I) 11:25-12:00 ………………………………………………………………………………………… 37<br />

Proton dynamics <strong>of</strong> CsH2PO4 and related salts containing organic ions<br />

Hideki Maekawa, A. Ishikawa and A. Kudo<br />

O 37 12:00-12:20 ……………………………………………………………………………………………. 38<br />

Oxygen reduction kinetics at Pt | CsHSO4 by conducting atomic force microscopy<br />

Mary W. Louie, A. Hightower, and S. M. Haile<br />

(Lunch Break)<br />

Oral Session (7) Chairperson: Sangtae Kim<br />

O 38 14:00-14:20 ……………………………………………………………………………………………. 39<br />

Impact <strong>of</strong> cation nonstoichiometry on phase behavior, mictrostructure, water incorporation and proton<br />

conductivities in yttrium-doped BaZrO3<br />

Yoshihiro Yamazaki, C. K. Yang, R. Hernandez-Sanchez and S. M. Haile<br />

O 39 14:20-14:40 ……………………………………………………………………………………………. 40<br />

Cost-effective solid-state reactive sintering method for proton-conducting ceramics<br />

Jianhua Tong, D. Clark and R. O’Hayre<br />

O 40 14:40-<strong>15</strong>:00 ……………………………………………………………………………………………. 41<br />

Solid state NMR studies <strong>of</strong> doped BaSnO3 and BaZrO3 protonic conductors: Defect trapping and ionic<br />

mobility<br />

Lucienne Buannic, F. Blanc and C. P. Grey<br />

O 41 <strong>15</strong>:00-<strong>15</strong>:20 ……………………………………………………………………………………………. 42<br />

The thermodynamics and kinetics <strong>of</strong> the dehydration <strong>of</strong> CsH2PO4 studied in the presence <strong>of</strong> SiO2<br />

Ayako Ikeda and S. M. Haile<br />

Closing Remarks <strong>15</strong>:20 Sossina Haile<br />

‐ vi ‐


Poster session (Monday-Tuesday 20:00-22:30)<br />

P 01 ...........................................................................................................................................................................................43<br />

Enhancement <strong>of</strong> proton conductivity in highly oriented poly(aspartic acid) thin film<br />

Yuki Nagao, J. Matsui, T. Abe, H. Yamamoto, T. Miyashita, N. Sata and H.Yugami<br />

P 02 ...........................................................................................................................................................................................44<br />

Crystallinity and morphology <strong>of</strong> PVDF-HFP-based proton-exchange membranes embedding<br />

polytyrenesulfonic acid-grafted silica particles<br />

M. Maréchal, F. Niepceron, J. Bigarré, H. Mendil-Jakani and G. Gebel<br />

P 03 ...........................................................................................................................................................................................45<br />

Proton mobility in tetragonal and monoclinic LaNbO4 through a second-order phase transition<br />

Kazuaki Toyoura, H. Fjeld, R. Haugsrud and T. Norby<br />

P 04 ...........................................................................................................................................................................................46<br />

NMR measurements on proton mobility in nano-crystalline YSZ<br />

Judith Hinterberg, A. Adams, B. Blümich , M. Wilkening, P. Heitjans, S. Kim, Z. A. Munir, R. A. De Souza and M. Martin<br />

P 05 ...........................................................................................................................................................................................47<br />

Determination <strong>of</strong> the amount <strong>of</strong> proton in proton conducting alumina by DC polarization method<br />

Yuji Okuyama, N. Kurita, D. Sato, H. Douhara and N. Fukatsu<br />

P 06 ...........................................................................................................................................................................................48<br />

Dynamic absorption behavior <strong>of</strong> hydrogen within perfluorosulfonic acid polymer electrolyte<br />

membranes during exposure to water vapor<br />

Bun Tsuchiya, S. Nagata, K. Saito, T. Shikama<br />

P 07 ...........................................................................................................................................................................................49<br />

Preparation <strong>of</strong> multilayered thin film fuel cell using titanium oxide as anodic catalyst via layer-by-layer<br />

assembly<br />

Hisatoshi Sakamoto, Y. Daiko, H. Muto, M. Sakai and A. Matsuda<br />

P 08 ...........................................................................................................................................................................................50<br />

Hydration kinetics <strong>of</strong> proton-conducting zirconates upon a change <strong>of</strong> temperature in wet atmosphere<br />

Jung In Yeon and H. Yoo<br />

P 09 ...........................................................................................................................................................................................51<br />

“In-situ” high temperature neutron diffraction study <strong>of</strong> lanthanum tungstate: a proton conductor with a<br />

fluorite-type structure<br />

Anna Magrasó, I. Ahmed, R. Haugsrud<br />

P 10 ...........................................................................................................................................................................................52<br />

Periodic long range proton conduction pathways in pseudo-cubic and orthorhombic perovskites<br />

M. A. Gomez, D. Shepardson, L. T. Nguyen and T. Kehinde<br />

P 11 ...........................................................................................................................................................................................53<br />

Proton conductivity and stability <strong>of</strong> Ba2In2O5 in hydrogen containing atmospheres<br />

Jasna Jankovic, D. P. Wilkinson, R. S. Hui<br />

P 12 ...........................................................................................................................................................................................54<br />

Proton solvation and transport in hydrated nafion<br />

Shulu Feng and G. A. Voth<br />

P 13 ...........................................................................................................................................................................................55<br />

Conductivity study <strong>of</strong> dense BaZr0.9Y0.1O3-δ obtained by spark plasma sintering<br />

Sandrine Ricote , N. Bonanos , H.Wang and B. A. Boukamp<br />

vii


P 14 ...........................................................................................................................................................................................56<br />

Conductivity study <strong>of</strong> A- and B-site (co-)doped LaNbO4<br />

M. Ivanova, Sandrine Ricote, W. A. Meulenberg, R. Haugsrud, T. Norby<br />

P <strong>15</strong> ...........................................................................................................................................................................................57<br />

Preparation <strong>of</strong> Nafion 117�-SnO2 composite membranes using an ion-exchange method<br />

C. F. Nørgaard, U. G. Nielsen and E. M. Skou<br />

P 16 ...........................................................................................................................................................................................58<br />

A study <strong>of</strong> Pt electrodes on proton conducting Ca-doped LaNbO4<br />

Ragnar Strandbakke, T. Norby and R. Haugsrud<br />

P 17 ...........................................................................................................................................................................................59<br />

Proton conductivity in BCY20-Pd ionic hybrid material<br />

Archana Subramaniyan, J. Tong, R. O’Hayre and N. Sammes<br />

P 18 ...........................................................................................................................................................................................60<br />

Defect chemistry and transport properties <strong>of</strong> oxide protonic perovskite materials with transition metals<br />

on B-site: BaZr1-xPrxO3-δ<br />

K. E. J. Eurenius, T. Kikuchi, M. Tamaru, S. Miyoshi and S. Yamaguchi<br />

P 19 ...........................................................................................................................................................................................61<br />

Microstructural analysis <strong>of</strong> Y doped BaZrO3<br />

Yukiko Oyama, T. Tsurui, M. Shogo and S. Yamaguchi<br />

P 20 ...........................................................................................................................................................................................62<br />

Site selectivity <strong>of</strong> Dopants in BaZr1-yMyO3-δ (M = Dy, Eu, Sm) and measurement <strong>of</strong> their water contents<br />

and conductivities<br />

Donglin Han, Y. Nose, K. Shinoda and T. Uda<br />

P 21 ...........................................................................................................................................................................................63<br />

Proton conductors based on the double perovskite Ba2YNbO6<br />

Xiaoxiang Xu, E. Konysheva, S. Tao and J. T.S. Irvine<br />

P 22 ...........................................................................................................................................................................................64<br />

Atmosphere dependent temperature evolution phase transition <strong>of</strong> BaCe0.9Y0.1O2.95 - A neutron powder<br />

diffraction study<br />

Abul K. Azad, A. Kruth and J. T.S. Irvine<br />

P 23 ...........................................................................................................................................................................................65<br />

The effect <strong>of</strong> cation non-stoichiometry <strong>of</strong> proton conducting LaNbO4<br />

Guttorm E. Syvertsen, A. Magrasó, M. Einarsrud and T. Grande<br />

P 24 ...........................................................................................................................................................................................66<br />

Solid state NMR studies <strong>of</strong> CsH2PO4, a protonic conductor for intermediate-temperature solid oxide<br />

fuel cells (IT-SOFCs)<br />

Gunwoo Kim, F. Blanc and C. P. Grey<br />

P 25 ...........................................................................................................................................................................................67<br />

First-principles studies <strong>of</strong> proton - Ba 2+ interactions in LaPO4<br />

Nicole Adelstein, J. Feng, J. A. Reimer, J. B. Neaton and L. C. De Jonghe<br />

P 26 ...........................................................................................................................................................................................68<br />

High temperature protonic conduction properties <strong>of</strong> (La,Sr)MO3 – SrZrO3 (M = Cr, Mn and Fe) solid<br />

solutions<br />

Atsushi Unemoto, K. Amezawa and T. Kawada<br />

viii


P 27 ...........................................................................................................................................................................................69<br />

Electrical conductivity and defect structure <strong>of</strong> Sr-Doped Nd3PO7<br />

Atsushi Unemoto, K. Amezawa and T. Kawada<br />

P 28 ...........................................................................................................................................................................................70<br />

Mechanochemical synthesis <strong>of</strong> proton conductive CsHSO4-azole composites for medium temperature<br />

dry fuel cells<br />

Song-yul Oh, T. Yoshida, G. Kawamura, H. Muto and A. Matsuda<br />

P 29 ...........................................................................................................................................................................................71<br />

Micro solid oxide fuel cells with perovskite type proton conductive thin electrolytes<br />

Fumitada Iguchi, K. Kubota, S. Tanaka, N. Sata, M. Esashi and H. Yugami<br />

P 30 ...........................................................................................................................................................................................72<br />

Room-temperature protonic conduction in nanocrystalline films <strong>of</strong> yttria-stabilized zirconia<br />

Sangtae Kim, H. J. Avila-Paredes and Z. A. Munir<br />

P 31 ...........................................................................................................................................................................................73<br />

Fabrication <strong>of</strong> BaCe0.9Y0.1O3-� thin film on Pd substrate by UV-MOD<br />

Koichi Asano, Y. Kozawa, Y. Mugikura and T. Watanabe<br />

P 32 ...........................................................................................................................................................................................74<br />

First principles calculations <strong>of</strong> defect equilibria in BaZrO3<br />

Akihide Kuwabara, C. A. J. Fisher, H. Moriwake, F. Oba, K. Matsunaga and I. Tanaka<br />

P 33 ...........................................................................................................................................................................................75<br />

On the symmetry <strong>of</strong> defects in perovskite-type protonic conductors: A Sc-45 NMR study<br />

Itaru Oikawa, M. Ando, H. Kiyono, M. Tansho, T. Shimizu and H. Maekawa<br />

P 34 ...........................................................................................................................................................................................76<br />

Performances <strong>of</strong> reversible SOFCs with BaZr0.6Co0.4O3-δ as air electrode<br />

F. He, Ranran Peng and C. Xia<br />

P 35 ...........................................................................................................................................................................................77<br />

Space charge effect and dopant segregation in acceptor-doped BaZrO3 proton conductors<br />

Mona Shirpour, B. Rahmati, W. Sigle, P. A. van Aken, R. Merkle and J. Maier<br />

P 36 ...........................................................................................................................................................................................78<br />

Synthesis and characterization <strong>of</strong> BaCe0.7Zr0.1Y0.1Yb0.1O3-δ proton conducting ceramic<br />

Siwei Wang, F. Zhao, L. Zhang and F. Chen<br />

P 37 ...........................................................................................................................................................................................79<br />

High flux ND-study <strong>of</strong> the structural phase transition in LaNbO4<br />

C. S. Knee, Morten Huse, T. Norby, S. G. Eriksson, R. Haugsrud<br />

P 38 ...........................................................................................................................................................................................80<br />

Performance <strong>of</strong> SOFC with proton-conductor BaCe0.7In0.2Yb0.1O3-δ electrolyte<br />

Fei Zhao, L. Dixon and F. Chen<br />

P 39 ...........................................................................................................................................................................................81<br />

Ionic conduction in Mg–Al and Zn-Al layered double hydroxide intercalated with inorganic anions<br />

Kiyoharu Tadanaga, Y. Furukawa, A. Hayashi and M. Tatsumisago<br />

P 40 ...........................................................................................................................................................................................82<br />

Determination <strong>of</strong> the enthalpy <strong>of</strong> hydration <strong>of</strong> oxygen vacancies by TG-DSC<br />

Christian Kjølseth, A. Løken, L.Wang, R. Haugsrud and T. Norby<br />

ix


P 41 ...........................................................................................................................................................................................83<br />

Proton transfer and hydration in 3M ionomers with different protogenic groups<br />

Jeffrey K. Clark and S. J. Paddison<br />

P 42 ...........................................................................................................................................................................................84<br />

Electrical conductivity and defect structure <strong>of</strong> Y-doped BaZrO3<br />

Ho-Il Ji, J. Lee, H. Kim, J. Son, H. Lee and B. Kim<br />

P 43 ...........................................................................................................................................................................................85<br />

Enhanced sinterability <strong>of</strong> Y-doped BaZrO3 powder synthesized with CuO, ZnO addition as a sintering<br />

aid<br />

Jong-Ho Lee, S. Hong, H. Ji, J. Park, H. Kim, J. Son, H. Lee and Byung-Kook Kim<br />

P 44 ...........................................................................................................................................................................................86<br />

Intermediate temperature solid oxide fuel cells based on a thin BaZr1-xYxO3-δ proton conductor<br />

electrolyte<br />

Daniele Pergolesi, E. Fabbri and E. Traversa<br />

P 45 ...........................................................................................................................................................................................87<br />

Does the increase in Y-dopant concentration improve the proton conductivity <strong>of</strong> BaZr1-xYxO3-δ fuel cell<br />

electrolytes?<br />

Fabbri Emiliana, D. Pergolesi, S. Licoccia and E. Traversa<br />

P 46 ...........................................................................................................................................................................................88<br />

Computational modeling <strong>of</strong> transport phenomena in polymer electrolyte membranes, nafion and<br />

hydrocarbon membrane<br />

Yoong-Kee Choe<br />

P 47 ...........................................................................................................................................................................................89<br />

Study <strong>of</strong> proton transport using reactive molecular dynamics<br />

Myvizhi Esai Selvan, D. J. Keffer, S. Cui and S. J. Paddison<br />

P 48 ...........................................................................................................................................................................................90<br />

Hydrogen transport in LaNbO4-LaNb3O9 composites<br />

W. Xing, G. E. Syvertsen, and Reidar Haugsrud<br />

P 49 ...........................................................................................................................................................................................91<br />

Poly(p-phenylene sulfone)s with high ion exchange capacity: Ionomers with unique microstructural and<br />

transport features<br />

C. C. de Araujo, K. D. Kreuer, M. Schuster, G. Portale and J. Maier<br />

P 50 ...........................................................................................................................................................................................92<br />

Rule <strong>of</strong> superprotonic phase transition in CsxRb1-xH2PO4<br />

Yasumitsu Matsuo, J. Hatori, Y. Yoshida and S. Ikehata<br />

P 51 ...........................................................................................................................................................................................93<br />

Hydration and protonic conductivity in LaAsO4 based ceramics<br />

Tor S. Bjørheim, T. Norby and R. Haugsrud<br />

P 52 ...........................................................................................................................................................................................94<br />

Investigation <strong>of</strong> hydrogen permeation in mixed conductor LaWOX (La/W = 5.6)<br />

Skjalg Erdal and R. Haugsrud<br />

P 53 ...........................................................................................................................................................................................95<br />

Fabrication <strong>of</strong> Ca-doped lanthanum niobate electrolyte film by electrophoretic deposition for PCFC<br />

applications<br />

Francesco Bozza and N. Bonanos<br />

x


P 54 ...........................................................................................................................................................................................96<br />

Ab initio molecular dynamics study <strong>of</strong> proton dynamics and mobility in phosphoric acid<br />

Linas Vilciauskas, G. Bester, K. D. Kreuer, M. E. Tuckerman and S. J. Paddison<br />

P 55 ...........................................................................................................................................................................................97<br />

CsH2PO4 nanoparticle synthesis via electrohydrodynamic atomization – aerosol size measurements<br />

Áron Varga, A. J. Downard, H. Wan Do, R. C. Flagan and S. M. Haile<br />

P 56 ...........................................................................................................................................................................................98<br />

Cobalt and yttrium doped barium zirconates as mixed conducting cathodes for SOFC<br />

Björn Björnsson, Juan C. Lucio-Vega, Y. Yamazaki and S. M. Haile<br />

P 57 ...........................................................................................................................................................................................99<br />

High-temperature phase behavior in the Rb3H(SO4)2-RbHSO4 pseudo-binary system<br />

Chatr Panithipongwut and S. M. Haile<br />

P 58 .........................................................................................................................................................................................100<br />

The phase transition and hysteresis behavior <strong>of</strong> Cs1-xKxH2PO4<br />

Daniil Kitchaev, A. Ikeda and S. M. Haile<br />

xi


Oxide Protonics: Electronic Structure and Acid Base Reactions in<br />

Bulk and Surface<br />

Shu Yamaguchi<br />

Plenary<br />

Department <strong>of</strong> Materials Engineering, Graduate School <strong>of</strong> Engineering, The University <strong>of</strong> Tokyo, 7-3-1 Hongo, Bunkyoku,<br />

Tokyo, 113-8656, Japan<br />

The discovery <strong>of</strong> the high temperature oxide protonics materials, defined as oxides in which<br />

proton has a certain transference number, by Pr<strong>of</strong>essor Iwahara opened a novel frontier in late 70’s.<br />

The quest for the proton conducting electrolytes was the hottest topics <strong>of</strong> that period <strong>of</strong> time in solid<br />

state ionics society and many different approaches were carried out for the breakthrough. The oxide<br />

protonics materials with perovskite structure, which is classified generally as a solid base, has a<br />

unique feature other than simple acid or basic characteristics, since the water incorporation reaction<br />

can be expressed by a combination <strong>of</strong> following partial acid/base reactions,<br />

H + �- �<br />

(g) � OO� OHO<br />

OH - �� - �<br />

(g) � VO � OHO<br />

in which H + and OH - gas were supplied from<br />

water by H2O(g) � H . Therefore,<br />

the presence <strong>of</strong> both acid and basic site are<br />

+ (g) � OH - (g)<br />

necessary for the water incorporation.<br />

The electronic band structer in Fermi<br />

energy region <strong>of</strong> the oxide protonics materials is<br />

characterized by a typical wide band gap<br />

semiconductors doped with acceptor dopants as<br />

shown in Figs. 1-(a) through (d), in which O1s<br />

s<strong>of</strong>t X-ray Absorption Spectra (XAS) are shown.<br />

Within the band gap region, there exist deffectinduced<br />

level (DIL) mainly composed <strong>of</strong> d- or forbital<br />

<strong>of</strong> B-site cation and the acceptor level<br />

mainly composed <strong>of</strong> O2p orbital, present with<br />

large unoccupied level in oxydizing stmosphere is<br />

reduced and split into two levels by proton<br />

incorporation. Also it is eviden that holes are created at the top <strong>of</strong> valence band by a thermcal<br />

activation. The H1s orbital cannot be observed by XAS and contributes indirectly to the band gap<br />

region.<br />

Fig. 1 O1s XAS spectra under dry and humidified<br />

conditions for (a) Sc-doped SrTiO3, (b) Y-doped<br />

SrZrO3, (c) Y-doped BaCeO3, (d) In-doped<br />

CaZrO3[1]<br />

In contrast to the fact that well-grown micrograins have a stable surface, nano-sized grains<br />

have unstable surface with dangling bonds or, in other words, strong basic site, which is stabilized<br />

by the formation <strong>of</strong> -OH - group at the surface by the hydration, serving as a strong attracter for water<br />

molecule by the hydrogen bond (H-B) formation. Such surface acid/base reaction site serves as a<br />

platform <strong>of</strong> the surface protonics in addition to the solid acid modified by conjugate ligands such as<br />

sulphates and phosphates. Further discussion on the basic properties <strong>of</strong> surface -OH - , its thermal<br />

stability, and proton conductivity will be presented.<br />

REREFENCES<br />

[1] T. Higuchi et al.: Materials Integration [in Japanese], 18(2005), No. 7, p. 28.<br />

- 1 -


O 01 (K)<br />

Ab initio molecular dynamics studies <strong>of</strong> proton structure and transport in<br />

solid acids and acid hydrate crystals<br />

Mark Tuckerman<br />

Department <strong>of</strong> Chemistry and Courant <strong>Institute</strong> <strong>of</strong> Mathematical Sciences, New York University<br />

We present ab initio molecular dynamics studies <strong>of</strong> two types <strong>of</strong> systems: The superprotonic phase<br />

<strong>of</strong> the solid acid cesium dihydrogen phosphate (CDP), and a series <strong>of</strong> hydrates <strong>of</strong><br />

trifluoromethanesulfonic (triflic) acid, specifically, the mono-, di-, tetra-, and penta-hydrates. The<br />

simulation results for CDP suggest that the cubic, superprotonic phase <strong>of</strong> the system is dynamically<br />

disordered, and two primary mechanisms <strong>of</strong> proton transport are obtained. The first involves<br />

rotation <strong>of</strong> the PO4 groups while the second is a Grotthuss type relay mechanism that can occur<br />

without such PO4 rotational motion. The latter is roughly an order <strong>of</strong> magnitude faster than the<br />

former. Specific time scales associated with these mechanisms are extracted using a new kinetic<br />

framework introduced by us for analyzing the dynamics <strong>of</strong> structural defects in hydrogen-bonded<br />

systems. Studies <strong>of</strong> the hydrates <strong>of</strong> triflic acid suggest preferred locations and structural<br />

characteristics <strong>of</strong> Eigen and Zundel type cations in these systems as a function <strong>of</strong> hydration level.<br />

Because strong nuclear quantum effects are expected in these systems, we have performed ab initio<br />

path integral calculations as well as classical ab initio molecular dynamics calculations, and we will<br />

compare these in order to assess the importance <strong>of</strong> nuclear quantum effects.<br />

- 2 -


Polyelectrolyte multilayers for electrochemical energy applications<br />

Avni A. Argun and Paula T. Hammond<br />

O 02 (I)<br />

Chemical Engineering, Massachusetts <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, 77 Massachusetts Avenue, Cambridge, MA 02139, USA<br />

The increasing global focus on alternative energy sources has<br />

led to a renewed interest in electrochemical energy systems<br />

such as fuel cells, batteries, and photoelectrochemical cells.<br />

At the core <strong>of</strong> these devices is an electrolyte which enables<br />

charge transport between electrodes. Polymeric solid state<br />

electrolytes <strong>of</strong>fer high mechanical strength and great<br />

fabrication flexibility, as well as better physical separation <strong>of</strong><br />

electrodes. Desired properties <strong>of</strong> electrolytes would depend<br />

on the application, but it is essential to utilize approaches that<br />

provide fast proton transport and environmentally friendly<br />

processing. The alternating adsorption <strong>of</strong> oppositely charged<br />

molecular species, known as the electrostatic layer-by-layer<br />

technique, is a simple and elegant method to construct highly<br />

tailored nanocomposites. This approach presents strong<br />

advantages allowing the incorporation <strong>of</strong> many different<br />

functional materials within a single film with exceptional<br />

homogeneity. The process is also flexible; by altering<br />

assembly conditions such as deposition pH, additives (salts or<br />

surfactants), and co-solvent choice, it is possible to greatly<br />

affect the final thin film composition and properties. We have<br />

Fig. 1 A picture <strong>of</strong> a free-standing PEM<br />

(~10 μm) assembled using spray-LbL<br />

method.<br />

utilized this method to engineer a number <strong>of</strong> systems, including the formation <strong>of</strong> proton exchange<br />

membranes in fuel cells, solid-state electrolytes for batteries, and mixed proton/electron conduction<br />

for photoelectrolytic water-splitting. In one example, we have developed a membrane with proton<br />

conductivity value <strong>of</strong> 70.0 mS/cm, the highest conductivity ever obtained from a multilayer<br />

assembled film. These multilayer films also exhibit very low liquid methanol permeability to<br />

improve the power output <strong>of</strong> a direct methanol fuel cell by over 50%. Finally, we have shown the<br />

rapid preparation <strong>of</strong> large area films using an automated spray apparatus, which reduces the<br />

processing<br />

time by over 40 times (Figure 1). Combined with the use <strong>of</strong> inexpensive water-soluble<br />

materials,<br />

this has proven to be a viable alternative to incumbent solution processing technologies.<br />

- 3 -


About the evolution <strong>of</strong> the microstructures <strong>of</strong> ionomers and<br />

polyelectrolytes for PEM fuel cell applications<br />

K. D. Kreuer 1 , C. C. de Araujo 1 , G. Portale 2 , U. Klock 1 , and J. Maier 1<br />

1. Max-Planck-Institut für Festkörperforschung, Heisenbergstraβe 1, D-70590 Stuttgart, Germany<br />

2.DUBBLE, BM26 at ESRF, 6 rue Jules Horowitz, BP220, F- 38043 Grenoble, Cedex, France<br />

kreuer@fkf.mpg.de<br />

O 03<br />

The strong relations between the microstructures <strong>of</strong> polymer electrolyte membranes and their<br />

transport properties (proton conductivity and water transport) are well known for a long time 1-6 , but<br />

the relations to other membrane properties are not investigated yet. Also, knowledge <strong>of</strong> the<br />

interactions driving the microstructural evolution on the relevant scales are still very scarce.<br />

We have therefore investigated the microstructure <strong>of</strong> divers PFSA and poly –(phenylene-sulfone)<br />

membranes as a function <strong>of</strong> RH and T by means <strong>of</strong> Small Angle X-Ray Scattering (SAXS), and the<br />

results have been related to the ones from Dynamical Mechanical (DMA) and Thermo Gravimetric<br />

Analysis (TGA) recorded under the same conditions.<br />

While the structures on the nano-meter scale are found be controlled by electrostatic interactions<br />

(IEC and degree <strong>of</strong> dissociation 7 ), the long range structures seems to be the consequence <strong>of</strong> (e.g.<br />

hydrophobic) backbone / backbone interactions. The first is found to be relevant for the relation<br />

between proton and hydrodynamic water transport (electroosmotic drag and water permeation) and<br />

the second for the viscoelastic properties <strong>of</strong> the membrane. Also long term structural changes are<br />

explained by the interplay <strong>of</strong> these two type <strong>of</strong> driving forces and relaxations.<br />

Finally, the relation between the microstructure <strong>of</strong> diverse poly-(phenylene-sulfone) based<br />

membranes (PBI-blends, block-copolymeres, branched architectures) and their proton conductivity<br />

under minimum hydration conditions is discussed.<br />

1. K. D. Kreuer, Th. Dippel, and J. Maier: Membrane Materials for PEM-Fuel-Cells, A Microstructural Approach; In<br />

Proton Conducting Membrane Fuel Cells I, Sh. Gottesfeld et al., ed., Vol. PV 95-23, The Electrochemical Society,<br />

Pennington, NJ, 1995, pages 241–246.<br />

2. K. D. Kreuer, M. Ise, A. Fuchs, and J. Maier: Proton and Water Transport in Nano-separated Polymer Membranes; J.<br />

Physique IV 10, Pr7–279–Pr7–281 (2000).<br />

3. K. D. Kreuer: On the Development <strong>of</strong> Proton Conducting Polymer Membranes for Hydrogen and Methanol Fuel<br />

Cells; J. Membrane Science 185 (1), 29–39 (2001).<br />

4. K. D. Kreuer, M. Schuster, B. Obliers, O. Diat, U. Traub, A. Fuchs, U. Klock, S. J. Paddison, and J. Maier:<br />

5. Short-side-chain proton conducting perfluorosulfonic acid ionomers: Why they perform better in PEM fuel cells; J.<br />

Power Sources 178(2), 499–509 (2008).<br />

6. C. C. de Araujo, K. D. Kreuer, M. Schuster, G. Portale, H. Mendil-Jakani, G. Gebel, and J. Maier:<br />

Poly(p-phenylene sulfone)s with high ion exchange capacity: ionomers with unique microstructural and transport<br />

features; Phys. Chem. Chem. Phys. 11(17), 3305–3312 (2009).<br />

7. A. Telfah, G. Majer, K. D. Kreuer, M. Schuster, and J. Maier: Formation and mobility <strong>of</strong> protonic charge carriers in<br />

methyl sulfonic acid-water mixtures: A model for sulfonic acid based ionomers at low degree <strong>of</strong> hydration; Solid State<br />

Ionics 181(11-12), 461–465 (2010).<br />

- 4 -


Coherent Proton Transport in Nafion<br />

George Reiter<br />

Physics Department, University <strong>of</strong> Houston<br />

O 04<br />

Measurements <strong>of</strong> the momentum distribution <strong>of</strong> the protons in Nafion and Dow 858 short side chain<br />

polymer show that, at any instant most <strong>of</strong> the protons are coherently delocalized in a double well<br />

potential. The potential differs for the two materials, for the same value <strong>of</strong> lambda(~14), in the<br />

separation <strong>of</strong> the wells and the fraction <strong>of</strong> delocalized protons. The transport in these materials<br />

must differ significantly from that for a localized proton that can be approximated as a classical<br />

particle.<br />

- 5 -


O 05 (K)<br />

NMR studies <strong>of</strong> local structure, cation ordering and protonic conduction in<br />

perovskites<br />

Lucienne Buannic 1 , Frederic Blanc 1 , Derek Middlemiss, and Clare P. Grey 1,2<br />

1 Department <strong>of</strong> Chemistry, Stony Brook University, Stony Brook, NY 11794-3400 USA<br />

2 Department <strong>of</strong> Chemistry, University <strong>of</strong> Cambridge, Lensfield Road, Cambridge, UK CB2 1EW<br />

This talk will illustrate the use <strong>of</strong> high field, multinuclear NMR spectroscopy to investigate the<br />

nature <strong>of</strong> the defects in materials for solid-state electrolytes. In particular, we focus on electrolytes<br />

that operate via protonic conduction in solid oxide fuel cells. For example, BaZrO3 or BaSnO3 can<br />

be doped with Y 3+ or Sc 3+ to create oxygen vacancies. These vacancies can be filled with H2O, the<br />

water molecules dissociating to form mobile ions that contribute to the long-range ionic transport in<br />

these systems. NMR experiments are used to examine the local structure, the locations <strong>of</strong> the<br />

vacancies and how this affects protonic/oxygen ion motion in these systems. NMR studies <strong>of</strong> the<br />

host B cations (Zr, Sn) can be used to quantify (particularly in the case <strong>of</strong> Sn) the ratio <strong>of</strong> 5:6 fold<br />

cations, and thus indirectly, the location <strong>of</strong> the vacancy. NMR studies <strong>of</strong> the dopants (Sc, Y) can be<br />

used to investigate vacancy trapping. The location <strong>of</strong> the dopant ion is <strong>of</strong>ten ambiguous, and we<br />

show that 89 Y can be used to investigate cation doping on the A vs. B site <strong>of</strong> the perovskite, the<br />

former reducing the number <strong>of</strong> oxygen vacancies and thus the number <strong>of</strong> protons, following<br />

hydration. Finally, 17 O NMR experiments are shown to be extremely sensitive to the nature <strong>of</strong> the<br />

B cation directly bound to the oxygen site, different resonances being observed for the B-O-B’ sites<br />

on cation doping. Chemical shift calculations, performed by using the CASTEP code, reveal that<br />

oxygen ions in axial and equatorial positions relative to a vacancy (i.e., in a OaxialB(Oeq)-vac local<br />

environment, where vac = oxygen vacancy) can be distinguished. The implications <strong>of</strong> these<br />

results for protonic conduction will be discussed.<br />

- 6 -


Recent Developments in Solid Acid Fuel Cells<br />

Calum R.I. Chisholm<br />

SAFCell, Inc. 36 South Chester Avenue, Pasadena CA 91106, USA<br />

O 06 (I)<br />

Solid acids are compounds whose chemistry and properties are intermediate between those <strong>of</strong> a<br />

normal acid, such as H2SO4 or H3PO4, and a normal salt, such as Cs2SO4. The anhydrous, solidstate<br />

proton conductivity <strong>of</strong> these materials can reach values higher than 10 -2 Ω -1 cm -1 when heated<br />

to moderately elevated temperatures (<strong>15</strong>0-250°C)[1], making solid acids excellent candidates for<br />

fuel cell electrolytes. In particular, fuel cells using cesium dihydrogen phosphate (CsH2PO4), with a<br />

conductivity <strong>of</strong> 2.5 x10 -2 Ω -1 cm -1 at 250 °C, have been demonstrated with thin (10-25 μm) gas tight<br />

electrolyte layers. [2-5]<br />

Recent developments at the cell and stack level show solid acid fuel cells (SAFCs) using<br />

CsH2PO4 to have sufficiently high enough performance/durability and low enough cost for<br />

commercialization in initial markets such as portable and premium power.[6] To date, SAFCs have<br />

measured peak power densities over 0.5 W/cm 2<br />

on hydrogen/air and lifetimes surpassing a<br />

thousand hours. SAFC stacks have also been<br />

fabricated using both <strong>15</strong> and 110 cm 2 active area<br />

cells with demonstrated power outputs over 250<br />

W. Very importantly, SAFCs have also operated<br />

on fuel streams with high levels <strong>of</strong> impurities such<br />

as CO, H2S and NH3, making them a fuel flexible<br />

technology capable <strong>of</strong> running on “dirty”<br />

reformates <strong>of</strong> commercially available fuels such<br />

as methane, propane, methanol, diesel and<br />

kerosene (JP8).[7, 8]<br />

Talk will highlight the recent promising<br />

developments and remaining challenges for<br />

SAFCs to become a broadly accepted commercial<br />

technology.<br />

- 7 -<br />

IR Corrected<br />

Cell Voltage [mV]<br />

1000<br />

900<br />

800<br />

700<br />

600<br />

250 O C<br />

Stoich 1.2/2 H 2 /O 2<br />

pH 2 O = 0.3 bar<br />

Cell Active Area<br />

1.5 cm 2<br />

<strong>15</strong> cm 2<br />

110 cm 2<br />

500<br />

0 100 200 300 400 500 600 700 800<br />

Current Density [mA/cm2]<br />

Fig. 1 Scaling <strong>of</strong> SAFC membrane electrode<br />

assemblies: area normalized performance is<br />

nearly identical over two orders <strong>of</strong><br />

magnitude.<br />

1. A.I. Baranov, L.A. Shuvalov and N.M. Shchagina, JETP<br />

Lett. 36 (1982) (11), p. 459.<br />

2. S.M. Haile, D.A. Boysen, C.R.I. Chisholm and R.B. Merle, Nature 410 (2001), p. 910.<br />

3. D.A. Boysen, T. Uda, C.R.I. Chisholm and S.M. Haile, Science 303 (2004), p. 68.<br />

4. T. Uda and S.M. Haile, Electrochemical and solid-state letters 8 (2005) (5), p. A245.<br />

5. S.M. Haile, R.I.C. Chisholm, K. Sasaki, D.A. Boysen and T. Uda, Faraday Discussions 134 (2007), p. 17.<br />

6. C.R.I. Chisholm, D.A. Boysen, A.B. Papandrew, S. Zecevic, S. Cha, K. Sasaki, A. Varga, K.P. Giapis and S.M.<br />

Haile, Interface Fall (2009), p. 7.<br />

7. T. Uda, D.A. Boysen, C.R.I. Chisholm and S.M. Haile, Electrochemical and solid-state letters 9 (2006) (6), p.<br />

A261.<br />

8. H.H. Duong, Solid Acid Fuel Cell Stack for APU Applications, In: DOE, Editor, DOE Hydrogen Program (2009),<br />

pp. 1177-1182.


Local Structure and Hydride-Ion Transport in MgY2H8<br />

Hitoshi Takamura, Keita Kurosu, Akira Hatakeyama, and Hideki Maekawa<br />

Department <strong>of</strong> Materials Science, Graudate School <strong>of</strong> Engineering, Tohoku University,<br />

6-6-11-301-2 Aramaki Aoba, Sendai 980-8579, Japan<br />

As well as metal oxides, metal hydrides comprising <strong>of</strong><br />

alkaline- and rare-earth elements such as CaH2 and YH3<br />

exhibit ionic-bonding nature. Depending on crystal structure<br />

and defect concentration, hydrogen can migrate as a form <strong>of</strong><br />

hydride ion (H - ) in such metal hydrides[1]. In this study, local<br />

structure and transport property <strong>of</strong> MgY2H8 having a<br />

defective BiF3-type structure are investigated. MgY2H8 is<br />

prepared by using a cubic-anvil-type apparatus at 800 °C<br />

under 5 GPa [2]. To modify hydrogen content in the hydride,<br />

a hydrogen source comprising <strong>of</strong> NaBH4 and Ca(OH)2 is<br />

used. Raman spectra <strong>of</strong> MgY2H8 at room temperature show a<br />

band at around 920 cm -1 corresponding to an F2g mode <strong>of</strong> Tsite<br />

hydrogen, which corresponds to oxide-ion sites in a<br />

stabilized zirconia, and a broad band at around 500 - 1200<br />

cm -1 presumably suggesting the presence <strong>of</strong> O-site hydrogen.<br />

The detailed local structure <strong>of</strong> MgY2H8 will be discussed<br />

based on these spectroscopic analyses. Electrical properties<br />

including hydride-ion transport <strong>of</strong> MgY2H8 are evaluated by<br />

using AC impedance and NMR spectroscopic techniques.<br />

Figure 1 shows Arrhenius plots <strong>of</strong> the total electrical<br />

conductivity <strong>of</strong> MgY2H8 prepared with or without the<br />

O 07<br />

Fig. 1 Arrhenius plots <strong>of</strong> MgY2H8<br />

prepared with or without a hydrogen<br />

source. Atmosphere is either under<br />

evacuation or H2 flow.<br />

hydrogen source. Even though all the sample show electronic conduction as a major carrier, the<br />

conductivity strongly depends on preparation and measurement atmospheres. This implies that<br />

hydrogen content affects the concentration <strong>of</strong> electronic carriers. The conductivity <strong>of</strong> hydride ion in<br />

MgY2H8 is estimated from T1 analysis <strong>of</strong> 1 H NMR, and also plotted in Fig. 1. The hydride-ion<br />

conductivity is found to be smaller than the electronic conductivity by 2 to 3 orders <strong>of</strong> magnitude.<br />

The<br />

transport properties <strong>of</strong> MgY2H8 will be discussed in the context <strong>of</strong> its local and electronic<br />

structures.<br />

1. S. J. van der Molen, M. S. Welling, and R. Griessen, Phys. Rev. Lett. 85 (2000), 3882.<br />

2. H. Takamura, Y. Goto, A. Kamegawa and M. Okada, Mater. Trans. 44 (2003), 583.<br />

- 8 -


O 08 (K)<br />

Improving the proton conductivity <strong>of</strong> yttrium-doped barium zirconate<br />

electrolytes towards the development <strong>of</strong> intermediate temperature solid<br />

oxide fuel cells<br />

Enrico Traversa<br />

International Center for Materials Nanoarchitectonics (MANA), National <strong>Institute</strong> for Materials Science (NIMS),<br />

1-1 Namiki, Tsukuba, Ibaraki 305-0044, Japan<br />

For the widespread deployment <strong>of</strong> solid oxide fuel cells (SOFCs), the high cost is an obstacle,<br />

together with the long-term stability due to the high working temperatures. Reducing the SOFC<br />

operating temperature in the 400-700°C range can reduce fabrication costs and overall improve<br />

performance. This aim can be obtained using high temperature proton conductor (HTPC) oxides as<br />

electrolytes, due to by their lower activation energy for proton conduction (0.3-0.6 eV), with respect<br />

to oxygen-ion conductor electrolytes. Moreover, proton conductor electrolytes <strong>of</strong>fer the advantage<br />

<strong>of</strong> generating water at the cathode, and thus the fuel does not become diluted during cell operation<br />

[1].<br />

Among polycrystalline HTPCs, doped BaCeO3 shows the largest proton conductivity, but it is not<br />

suitable for fuel cell applications since it easily reacts with acidic gases, e.g. CO2, and water vapor.<br />

Doped BaZrO3 <strong>of</strong>fers instead excellent chemical stability against reaction with CO2 and H2O, but<br />

low conductivity values for sintered pellets. The low electrical conductivity is the consequence <strong>of</strong><br />

poor conductive grain boundary regions coupled with the material sintering difficulties.<br />

This work reports the various strategies recently followed in our lab to improve the conductivity <strong>of</strong><br />

Y-doped barium zirconate [2,3], towards the limitation <strong>of</strong> grain boundary surface, the use <strong>of</strong> co-<br />

doping<br />

for improving the sinterability, and the fabrication <strong>of</strong> films by pulsed laser deposition (PLD)<br />

[ 4,5].<br />

1. E. Fabbri, D. Pergolesi and E. Traversa, Chemical Society Reviews, in press.<br />

2. A. D'Epifanio, E. Fabbri, E. Di Bartolomeo, S. Licoccia and E. Traversa, Fuel Cells 8 (2008), 69.<br />

3. E. Fabbri, A. D'Epifanio, E. Di Bartolomeo, S. Licoccia and E. Traversa, Solid State Ionics 179 (2008), 558.<br />

4. E. Fabbri, A. D’Epifanio, S. Sanna, E. Di Bartolomeo, G. Balestrino, S. Licoccia and E. Traversa, Energy &<br />

Environmental Science 3 (2010), 618.<br />

5.<br />

D. Pergolesi, E. Fabbri and E. Traversa, Electrochemistry Communications 12 (2010), 977.<br />

- 9 -


Existence <strong>of</strong> equilibrium mixture between Hydrated and Unhydrated<br />

Proton conducting Perovskites and possible implications<br />

John T.S. Irvine 1 , Maarten C. Verbraeken 1 , Hemangildo Viana 1 , Azul K. Azad 1 , Istaq Ahmed 2 , and<br />

Sten Eriksson 2<br />

1. School <strong>of</strong> Chemistry, University <strong>of</strong> St Andrews, St Andrews, Fife KY16 9ST, UK<br />

2. Department <strong>of</strong> Chemical and Biological Engineering, Chalmers University <strong>of</strong> <strong>Technology</strong>, Sweden<br />

O 09<br />

Neutron powder diffraction <strong>of</strong>fers an excellent technique for the location <strong>of</strong> protons (or deuterons)<br />

in hydrated perovskite proton conductors. It would be expected that the concentration <strong>of</strong> these<br />

protons would increase as temperature is decreased or the humidity <strong>of</strong> equilibrating atmosphere is<br />

increased; however, in recent studies <strong>of</strong> several systems we have observed some quite dissimilar<br />

behaviour. These systems include members <strong>of</strong> the BCN family such as Ba3Ca1.18Ta1.82O8.73 and<br />

Indium doped barium zirconates such as Ba0.5In0.5ZrO3-y. Results will be presented from X-ray and<br />

neutron powder diffraction that demonstrate 2 phases with similar structure are present in<br />

equilibrium, with one being essentially hydrated and one essentially unhydrated. In such a system,<br />

as temperature is decreased or humidity increased, then the dominant change will be an increase in<br />

the hydrated phase proportion. This behavior can be simply interpreted in terms <strong>of</strong> phase equilibria<br />

diagrams and the implications <strong>of</strong> this deviation from assumed solid solution <strong>of</strong> water into a single<br />

phase will be discussed.<br />

- 10 -


Local structure and proton dynamics in In-doped BaZrO3<br />

Maths Karlsson 1 , Aleksandar Matic 2 , Istaq Ahmed 3 , Christopher S. Knee 4 , Sten Eriksson 3<br />

O 10<br />

1 European Spallation Source AB, 221 00 Lund, Sweden, 2 Dept. <strong>of</strong> Applied Physics, Chalmers University <strong>of</strong> <strong>Technology</strong>,<br />

412 96 Göteborg, Sweden, 3 Dept. <strong>of</strong> Chemical and Biological Engineering, Chalmers University <strong>of</strong> <strong>Technology</strong>, 412<br />

96 Göteborg, Sweden, 4 Dept. <strong>of</strong> Chemistry, University <strong>of</strong> Gothenburg, 412 96 Göteborg, Sweden<br />

Hydrated acceptor-doped perovskite-type<br />

oxides, such as In-doped BaZrO3, have been<br />

widely studied in recent years because <strong>of</strong> their<br />

high proton conductivities at elevated<br />

temperatures and concomitant promise for use<br />

as electrolytic membrane in future fuel cells<br />

[1]. The acceptor-doping creates an oxygen<br />

deficient structure, which in protons can be<br />

introduced by exposure to a humid<br />

atmosphere [1]. The protons are generally<br />

believed to migrate into the structure through<br />

jumps between oxygens. From this it is easy<br />

to understand that the local structure plays a<br />

key role for the local proton dynamics, as well<br />

as for the long-range diffusion. In this<br />

context, we have during the last years been<br />

investigating the relationship between local<br />

structure and proton dynamics in protonconducting<br />

perovskites, using neutron<br />

diffraction [2], inelastic [3] and quasielastic<br />

[4] neutron scattering, and optical vibrational<br />

spectroscopy [5-7]. The goal with these<br />

studies has been to increase the understanding<br />

<strong>of</strong> the proton conduction process so that one<br />

can ‘realize’ new compounds with higher<br />

conductivities, which is a critical factor for<br />

practical applications.<br />

As one example <strong>of</strong> our results, neutron spinecho<br />

data <strong>of</strong> BaInxZr1-xO3-x/2Hx (x = 0.10 and<br />

0.50) shows that the diffusional proton selfdynamics<br />

are strongly dependent on the In<br />

concentration and the short-range structure <strong>of</strong><br />

the material [4], see Figure 1. For the highly<br />

doped material (x = 0.50), we observe a<br />

distribution <strong>of</strong> translational diffusional rates<br />

<strong>of</strong> the protons in the structure on the time<br />

scale <strong>of</strong> nanoseconds and with an effective<br />

activation energy <strong>of</strong> about 0.75 eV [4]. The<br />

wide range <strong>of</strong> diffusional rates is found to be<br />

related to dopant-induced short-range<br />

- 11 -<br />

distortions <strong>of</strong> the average cubic structure,<br />

yielding many structurally different local<br />

configurations <strong>of</strong> the protons in the structure,<br />

each configuration with its own energy<br />

barriers for proton migration [4,5]. For<br />

comparison, the data <strong>of</strong> the weakly doped<br />

compound (x = 0.10) reveal dynamics on a<br />

more well defined time scale, ~60 ps at 500<br />

K, as a result <strong>of</strong> a more ordered local structure<br />

[4,5]. In this presentation, I will describe our<br />

results in more detail.<br />

Figure 1. Intermediate scattering function <strong>of</strong> In-doped<br />

BaZrO3 obtained from neutron spin-echo at T = 500 K<br />

and Q = 1.05 Å -1 (Q: momentum transfer during the<br />

neutron-proton scattering process).<br />

1. K. D. Kreuer, Annu. Rev. Mater. R. 33 (2003) 333.<br />

2. I. Ahmed, J. Alloys and Compds 450 (2008) 103.<br />

3. M. Karlsson, Phys. Rev. B 77 (2008) 104302.<br />

4. M. Karlsson, J. Phys. Chem. C 114 (2010) 3292.<br />

5. M. Karlsson, Chem. Mater. 20 (2008) 3480.<br />

6. M. Karlsson, Solid State Ionics 181 (2010) 126.<br />

7. M. Karlsson, J. Phys. Chem. C 114 (2010) 6177.


Grain Boundary Conduction in polycrystalline Y-doped BaZrO3: Space<br />

Charge Analysis<br />

Sangtae Kim, Chien-Ting Chen, Christina Danel<br />

Department <strong>of</strong> Chemical Engineering and Materials Science, University <strong>of</strong> <strong>California</strong>, Davis, CA 95616, USA<br />

O 11<br />

Y-doped BaZrO3 (BZY) is one <strong>of</strong> the most intensively studied high-temperature proton conductors.<br />

The bulk conductivity <strong>of</strong> this chemically stable solid electrolyte (SE) is reported to be higher than<br />

that <strong>of</strong> conventional oxygen-ion conducting solid electrolytes such as doped zirconia and ceria in<br />

the so-called intermediate temperature range (ca 400 – 600 C). On the other hand, the grain<br />

boundaries in the polycrystalline ceramic <strong>of</strong> this material is known to be highly resistive to proton<br />

transport across them, leading to the total conductivity <strong>of</strong> the polycrystalline BZY being<br />

substantially lower than its bulk conductivity. It has been speculated that such high proton<br />

resistance at the grain boundaries<br />

may be due to proton depletion at<br />

the grain boundaries (namely space<br />

charge effects). In fact, some<br />

researchers have recently applied<br />

the space charge model to explain<br />

the conduction behavior at the<br />

grain boundaries in BZY. In this<br />

contribution, we demonstrate the<br />

results <strong>of</strong> our investigation on the<br />

potential space charge effects on<br />

the grain boundary conduction in<br />

BZY. We have applied the space<br />

charge model to the bulk and the<br />

grain boundary conductivity<br />

measured from not only our own<br />

samples but also BZY with<br />

log c H+<br />

4 3 2 1 0 0<br />

x / nm<br />

1 2 3 4<br />

different Y-concentration previous reported. The grain boundary activation energies and widths<br />

measured from BZY will be quantitatively discussed based on the space charge model.<br />

H + H + H +<br />

H + H + H +<br />

Fig. 1. Proton concentration pr<strong>of</strong>iles in a space charge zone in BZY at<br />

different core charges estimated at 400�C. The dopant concentration is<br />

1�10 20 cm -3 .<br />

- 12 -<br />

0.1V<br />

0.2V<br />

0.3V<br />

0.4V<br />

0.5V


Atomic-Scale Insights into Proton-Conducting Oxides:<br />

Defects, Dopants and Transport<br />

M. Saiful Islam<br />

Department <strong>of</strong> Chemistry, University <strong>of</strong> Bath, UK<br />

m.s.islam@bath.ac.uk<br />

O 12 (K)<br />

Ion-conducting oxide materials have been attracting considerable attention owing to their potential<br />

applications in solid oxide fuel cells (SOFCs) and other electrochemical technologies. It is clear that<br />

fundamental materials research is a crucial component in the discovery and characterisation <strong>of</strong><br />

materials for such applications. In this context, advanced computational techniques are now<br />

powerful tools for probing the properties <strong>of</strong> energy materials on the atomic- and nano-scale. This<br />

presentation will highlight recent studies on solid-state proton conductors focusing on perovskitetype<br />

materials (such as doped BaZrO3) and novel oxide structures containing tetrahedral units (such<br />

as LaBaGaO4, doped LaNbO4, and Ge-based apatites). Key solid-state properties investigated<br />

include local protonic sites, defect-dopant association and ion conduction mechanisms. In every<br />

case, our results are closely correlated with the available structural and transport studies (e.g.<br />

diffraction, conductivity, NMR).<br />

1. Kendrick E, Kendrick J, Knight KS, Islam MS and Slater PR, Nature Materials, 6, 871 (2007).<br />

2. Islam MS, Slater PR, MRS Bulletin 34, 935 (2009).<br />

3. Panchmatia PM, Orera A, Kendrick E, Smith ME, Slater PR, Islam MS, J. Mater. Chem. 20, 2766 (2010).<br />

- 13 -


DFT-MD study <strong>of</strong> local structures and proton transport in disordered<br />

Brownmillerite-based compounds<br />

Karsten Rasim 1,2 , Florent Boucher 1 , Olivier Joubert 1 , Philippe Baranek 2 , Mathieu Marrony 3<br />

1 Institut des Matériaux Jean Rouxel, 2 rue de la Houssinière, BP32229, 44322 Nantes cedex, France<br />

2 EDF R&D, Dep. MMC, Avenue des Renardières, 77818 Moret sur Loing Cedex, France<br />

3 EIFER, Emmy-Noether-Strasse 11, 76131 Karlsruhe, Germany<br />

O 13<br />

A detailed study <strong>of</strong> heavily cation-substituted, disordered Brownmillerite based compounds is<br />

presented. The focus is on their proton conduction behaviour as well as their structural properties in<br />

the dry state. The study is conducted by DFT calculations (both static and Born-Oppenheimer<br />

molecular dynamics [1]) combined with experimental approaches: XANES and IR-spectra<br />

measurements. Amongst others, by means <strong>of</strong> these techniques, the coordination preference <strong>of</strong><br />

different substituents, the protonic mobility and vibrational spectra can be obtained and compared<br />

to experimental evidence. The focus lies on the Ba2In2(1-x)Ti2xO5+x (BITx)[2] family, proven to be a<br />

well suited electrolyte material for PCFC. Moreover, neighbouring compounds, such as Sr2In2(1x)Ti2xO5+x<br />

(SITx), Ba2In2(1-x)Zr2xO5+x (BIZx) and Ba2In2(1-x)Y2xO5 (BIYx), are being considered as<br />

they prove useful to systematically clarify the influence <strong>of</strong> different chemical environments on<br />

proton diffusivity (e.g. different "proton affinities" <strong>of</strong> the Ti, Zr or Y- substituents, more or less<br />

pronounced hydrogen bonding, distinction between extra- and intra-octahedrally bonded protons,).<br />

All those aspects are obtained in<br />

the molecular dynamics<br />

framework, therefore including<br />

temperature or entropic effects,<br />

both vibrational and<br />

configurational in nature.<br />

A key aspect in this study <strong>of</strong> proton<br />

transport is the detailed analysis <strong>of</strong><br />

each <strong>of</strong> the six unique elementary<br />

displacement steps (shown in<br />

figure 1) that were identified<br />

during the MD-simulations. Each<br />

<strong>of</strong> them can be characterized by an<br />

effective free-energy barrier at a<br />

given temperature and by combining runs at different<br />

temperatures, activation energies and -entropies can be<br />

extracted. It turns out, that the protonic mobility in this class<br />

<strong>of</strong> materials is heavily influenced by the existence <strong>of</strong> the two<br />

distinct proton sites, intra- and extra-octahedral. This is to be<br />

seen in contrast to the case <strong>of</strong> the well studied perovskite-<br />

related compounds that do usually only enable intra-octahedral proton sites.[4]<br />

[1] G. Kresse, J. Hafner, Phys.Rev.B. 49, 20 (1994) 14251.<br />

[2] Jayaraman et al, Solid State Ionics 170 (2004) 17-24.<br />

[3] E. Quarez et al. Journal <strong>of</strong> Power Sources, 195 (<strong>15</strong>), (2010) 4923-4927<br />

[4] K-D. Kreuer, Annu. Rev. Mater. Res. 33, 333-359 (2003)<br />

- 14 -<br />

Fig. 1 The six unique proton<br />

displacement modes in the<br />

Brownmillerite related compounds<br />

studied.


O 14<br />

Ab Initio Studies <strong>of</strong> the Effect <strong>of</strong> Hydrophobic<br />

Environment on Proton Transfer in Model Perfluorsulfonic Acid Systems<br />

Bradley F. Habenicht and Stephen J. Paddison<br />

Department <strong>of</strong> Chemical and Biomolecular Engineering, University <strong>of</strong> Tennessee Knoxville,<strong>15</strong>12 Middle Drive,<br />

Knoxville, TN 37996, USA<br />

Single walled carbon nanotubes (CNT) were functionalized<br />

with perfluorosulfonic acid (-CF2SO3H) groups and hydrated<br />

with water molecules (1 and 3 H2O/SO3H) to investigate<br />

proton dissociation and transport in simplified proton<br />

exchange membranes under conditions <strong>of</strong> minimal hydration<br />

[1]. The systems were studied using ab initio molecular<br />

dynamics (AIMD), which does not require a priori<br />

assumptions about proton dissociation and hydration. The<br />

influence <strong>of</strong> the hydrophobic environment was studied by<br />

comparing proton mobilities with and without fluorine atoms<br />

attached to the CNT walls (Figure 1). The AIMD trajectories<br />

showed that dissociation <strong>of</strong> the acidic proton was increased<br />

as sulfonic acid density increased, however, greater densities<br />

also increased trapping <strong>of</strong> the dissociated proton [2]. The<br />

fluorine atoms accepted hydrogen bonds from the water<br />

molecules, stabilized hydrogen bonding, and enhanced<br />

proton dissociation. The CNT systems with bare walls<br />

exhibited a propensity for Zundel cation (H5O2 + ) formation,<br />

hile the fluorinated systems favoured the hydronium cation<br />

H3O + w<br />

( ) [3].<br />

Fig. 1 Schematic <strong>of</strong> functionalized CNTs<br />

used<br />

in this study. Atoms have been<br />

removed<br />

from CNT walls for clarity.<br />

1. K.D. Kreuer, S.J. Paddison, E. Spohr and M. Schuster, Chem. Rev. 104 (2004), p. 4637.<br />

2. B.F. Habenicht, S.J. Paddison and M.E. Tuckerman, J. Mater. Chem. In press (2010).<br />

3.<br />

B.F. Habenict, S.J. Paddison and M.E. Tuckerman, Phys. Chem. Chem. Phys. In press (2010).<br />

- <strong>15</strong> -


- 16 -<br />

O <strong>15</strong>(K)<br />

Analysis <strong>of</strong> Transport through Mixed Proton, Oxygen Ion, and Electron<br />

(Hole) Conductors: Fuel Cell and Electrolyzer Modes<br />

Anil Virkar<br />

Utah University<br />

This talk is on the analysis <strong>of</strong> transport through mixed proton, oxygen ion and electron (hole)<br />

conducting membranes. The predominant transport is by protons and oxygen ions with electronic<br />

conductivity being small. The only permeation <strong>of</strong> H2, O2, and H2O is assumed to occur via a<br />

coupled transport <strong>of</strong> H + , O 2- , and e (and/or h). Transport is analyzed for three cases: (1) Fuel cell<br />

under normal operating conditions, (2) A driven fuel cell, and (3) Electrolyzer. Transport equations<br />

are presented in the Onsager format. All coefficients are given in terms <strong>of</strong> cell parameters and the<br />

operating conditions. Spatial variations <strong>of</strong> chemical potentials <strong>of</strong> all neutral species through the<br />

membrane are determined as a function <strong>of</strong> operating conditions. Implications <strong>of</strong> the analysis<br />

performed and the results obtained will be discussed. A possible experimental procedure for the<br />

measurement <strong>of</strong> transport coefficients in the various modes <strong>of</strong> operation will be described.


O 16 (I)<br />

Hydrogen in nominally anhydrous minerals: analysis and implications<br />

George R. Rossman<br />

Geological and Planetary Sciences, <strong>California</strong> Inst. <strong>Technology</strong>, 1200 E <strong>California</strong> Blvd., Pasadena, CA 91125, USA<br />

Many naturally occuring minerals and their synthetic<br />

counterparts contain minor amounts <strong>of</strong> hydrous<br />

components, usually as the OH - ion and as the H2O<br />

molecule [1]. These hydrous components can influence<br />

the physical properties <strong>of</strong> the host phase, in some cases<br />

producing 5- or 6-orders <strong>of</strong> magnitude change in<br />

properties when incorporated in concentrations at the level<br />

<strong>of</strong> a few-hundred ppm. For decades, the detection <strong>of</strong> the<br />

hydrous components has been easily accomplished<br />

through the use <strong>of</strong> infrared spectroscopy. The<br />

determination <strong>of</strong> their quantitative concentration has<br />

proven more difficult, especially when their<br />

concentrations is low. Nevertheless, it has been shown<br />

that these components are particularly important in<br />

minerals that come from the earth’s mantle and constitute<br />

a major resevoir <strong>of</strong> “water” within planet earth. Much<br />

effort has been developed towards determining the<br />

analytical concentrations <strong>of</strong> hydrous components using<br />

methods such as quantitative infrared spectroscopy,<br />

nucular magnetic resonance (NMR) spectroscopy, thermal<br />

analysis, nuclear pr<strong>of</strong>ile analysis (NPA), and, more<br />

recently, secondary ion mass spectrometry (SIMS).<br />

Several <strong>of</strong> these analytical methods reveal a high<br />

concentration <strong>of</strong> hydrogen on the surface <strong>of</strong> crystals<br />

compared to the interior (Figure 1).<br />

Fig. 1 The hydrogen content,<br />

expressed as H2O, <strong>of</strong> a nominally<br />

anhydrous crystal <strong>of</strong> olivine,<br />

(Mg,Fe)2SiO4 determined by nuclear<br />

pr<strong>of</strong>ile analysis [1].<br />

1. Rossman GR, Reviews in Mineralogy and Geochemistry 62 (2006), 1.<br />

2. D.R. Bell, G.R. Rossman, J. Maldener, D. Endisch, F. Rauch, Journal <strong>of</strong> Geochemical Research, 108(B2), 2105,<br />

doi:10.1029/2001JB000679, 2003<br />

- 17 -


The dynamical behaviour <strong>of</strong> water and protons in perfluorinated<br />

membranes and surfactants.<br />

S. Lyonnard 1 , C. Cailleteau 1 , H. Mendil-Jakani 1 , S. Mossa 1 , Gerard Gebel 1<br />

Armel Guillermo 1 , B. Frick 2 and J. Ollivier 2<br />

1 Structures et Propriétés d’Architectures Moléculaires, UMR 5819 (CEA-CNRS-UJF)<br />

Laboratoire des Polymères Conducteurs Ioniques. INAC/SPrAM – CEA-Grenoble<br />

38054 Grenoble Cedex 9, France<br />

2<br />

Institut Laue Langevin<br />

BP<strong>15</strong>6, 6 rue Jules Horowit z,<br />

38042 Grenoble, France.<br />

- 18 -<br />

O 17 (K)<br />

Perfluorinated ionomers, the reference membranes for PEM fuel cells, are characterized by a<br />

hydrophobic/hydrophilic nanophase separation. Water is the medium for proton transfer: it plays a<br />

central role, although not yet fully elucidated, in the conduction mechanisms that need to be<br />

optimized for the ongoing development <strong>of</strong> new efficient fuel cells.<br />

The dynamical properties <strong>of</strong> the water are severely affected by the confinement in the charged<br />

matrix at the nanometric scale. In order to clarify the structure-to-transport relationship in ionomers,<br />

we have studied by Quasi-elastic Neutron Scattering and Molecular Dynamics Simulations the<br />

behaviour <strong>of</strong> water in different systems: perfluorinated membranes (Nafion® [1], aged Nafion®,<br />

and short side chain Aquivion TM ) and surfactants [2]. The latter mimic the physico-chemical<br />

properties <strong>of</strong> the real systems and <strong>of</strong>fer the advantage to self-assemble in well defined organized<br />

phases.<br />

QENS studies were performed on both time-<strong>of</strong>-flight and backscattering spectrometers at the ILL.<br />

The quasielastic spectra are discussed on the basis <strong>of</strong> a sophisticated model for confined motion [3].<br />

This unique diffusion model based on Gaussian statistics takes into account both localized<br />

translational motions and long-range diffusion. Comparison <strong>of</strong> the diffusion mechanisms and<br />

parameters (diffusion coefficients, confinement sizes and characteristic times) obtained in the<br />

various<br />

systems as a function <strong>of</strong> their hydration bring new insight on the complex molecular<br />

scenario<br />

for proton motion under confinement.<br />

[1] J-C. Perrin, S. Lyonnard and F. Volino; Quasielastic neutron scattering study <strong>of</strong> water dynamics in hydrated nafion<br />

membranes, Journal <strong>of</strong> Physical Chemistry C, 111 (2007), 3393-3404.<br />

[2] S. Lyonnard, B-A. Bruening, G. Gebel, A. Guillermo, J. Ollivier and B. Frick; Perfluorinated surfactants as model<br />

charged systems for understanding the effect <strong>of</strong> confinement on proton transport and water mobility in fuel cells<br />

membranes. A study by QENS. Proceedings <strong>of</strong> Confit 2010, EPJ, to be published.<br />

[3] F. Volino, J-C. Perrin and S. Lyonnard; Gaussian model for localized translational motion : application to<br />

incoherent neutron scattering, Journal <strong>of</strong> Physical Chemistry B, 110 (2006), 11217-11223.


O 18 (K)<br />

Tailored Tungsten Oxide Nanoparticles Produced with Hot Wire Chemical<br />

Vapor Deposition for Protonic Electrochromic and Fuel Cell Applications<br />

Anne C. Dillon, See-Hee Lee*, Rohit Deshpande, Katherine E. Hurst, Shyam Kocha, Virginia R.<br />

Anderson* and Steven M. George*<br />

National Renewable Energy Laboratory , 1617 Cole Blvd., Golden , CO 80401, USA<br />

*University <strong>of</strong> Colorado, Boulder 80309, USA<br />

The majority <strong>of</strong> the world energy consumption is derived from fossil fuels. The conversion<br />

reactions required for retrieving energy from carbon resources result in the production <strong>of</strong> green<br />

house gases and subsequent global warming effects. It is therefore necessary to develop “green”<br />

technologies that will reduce fossil fuel consumption. We<br />

have employed hot wire chemical vapor deposition<br />

(HWCVD) for the generation <strong>of</strong> crystalline WO3<br />

nanostructures at high-density. Furthermore, the<br />

morphology <strong>of</strong> the nanoparticles is easily tailored by<br />

altering the HWCVD synthesis conditions[1]. A<br />

transmission electron microscope (TEM) image <strong>of</strong> a WO3<br />

nanorod is provided in Figure 1. Electrochromic (EC)<br />

materials change optical properties (darken / lighten) by<br />

application <strong>of</strong> small electric potential difference, and are<br />

suitable for energy efficient windows, anti-glare<br />

automobile rearview mirrors and sunro<strong>of</strong>s that enable<br />

significant energy savings. There are two critical criteria<br />

for selecting an EC material. One is the time constant <strong>of</strong><br />

the ion intercalation reaction, which is limited both by the<br />

diffusion coefficient and also by the length <strong>of</strong> the<br />

diffusion path. While the former depends on the<br />

Figure 1: TEM image <strong>of</strong> HWCVDgenerated<br />

crystalline WO3 nanorods.<br />

chemical and crystal structure <strong>of</strong> the metal oxide, the latter is determined by the microstructure. In<br />

the case <strong>of</strong> a nanoparticle the smallest dimension represents the diffusion path length. Thus<br />

designing a nanostructure with a small radius while also maintaining the proper crystal structure is<br />

key to a material with fast insertion kinetics, enhanced durability and superior performance. State<strong>of</strong>-the–art<br />

amorphous WO3 films that are typically employed for commercial EC applications rely<br />

on Li + insertion rather than H + insertion as cycling in an acidic electrolyte generally results in<br />

significantly faster device degradation. However, the ability to employ protons to change the<br />

optical properties will enable significantly faster switching speeds to be demonstrated. We have<br />

shown that porous films <strong>of</strong> crystalline WO3 nanorods exhibit vastly superior electrochemical<br />

cycling stability in an acidic electrolyte and higher charge insertion with comparable coloration<br />

efficiency relative to amorphous films[2]. The WO3 nanoparticles are also promising as catalyst<br />

supports in PEM fuel cells. Tungsten oxides have long been known to enhance the oxygen<br />

reduction reaction and to decompose peroxide that attacks the carbon support and ionomer in the<br />

catalyst layer. We are currently employing atomic layer deposition[3] to anchor small Pt clusters to<br />

the WO3 nanorods for the development <strong>of</strong> an improved hydrogen fuel cell catalyst. The synthesis <strong>of</strong><br />

these metal oxide nanoparticles and their application in both EC windows and as fuel cell catalysts<br />

will be presented in detail.<br />

1. A.H. Mahan, P.A. Parilla, K.M. Jones and A.C. Dillon, Chem. Phys. Lett. 413 (2005), p. 88.<br />

2. S.-H. Lee, R. Deshpande, P.A. Parilla, K.M. Jones, B. To, A.H. Mahan and A.C. Dillon,<br />

Adv. Mat. 18 (2006), p. 763.<br />

3. A.C. Dillon, A.W. Ott, J.D. Way and S.M. George, Surf. Sci. 322 (1995) (1-3), p. 230.<br />

- 19 -


O 19 (I)<br />

Uphill gas permeation due to coupled transport in a mixed proton, oxygen<br />

vacancy and hole conducting perovskite membrane<br />

Michael Sanders 1 , Jianhua Tong 1 , Huayang Zhu 2 , Robert Kee 2 , and Ryan O’Hayre 1<br />

1 Metallurgical and Materials Engineering, Colorado School <strong>of</strong> Mines, <strong>15</strong>00 Illinois St., Golden, CO. 80401, USA<br />

2 Mechanical Engineering, Colorado School <strong>of</strong> Mines, <strong>15</strong>00 Illinois St., Golden, CO. 80401, USA<br />

Coupled transport membranes utilize the<br />

flux <strong>of</strong> one species to help drive the flux<br />

<strong>of</strong> another species, in some cases even<br />

against its own chemical potential<br />

gradient. These membrane processes are<br />

exploited in both natural systems, (where,<br />

for example, they underpin cellular<br />

membrane signal transduction [1]) and in<br />

man-made systems, (where, for example,<br />

they have been applied to a number <strong>of</strong><br />

multi-component solvent or metal<br />

separations processes [2-5]). To date,<br />

however, coupled transport membrane<br />

studies and applications have been almost<br />

exclusively confined to aqueous or liquid<br />

membrane media. While these systems<br />

have provided remarkable insights and<br />

have enabled novel separations<br />

approaches, both scientific understanding<br />

and commercial application have been<br />

Fig. 1 Permeation experiments showing uphill transport. The tilt<br />

<strong>of</strong><br />

the arrows signifies whether the transport was “uphill” or<br />

“downhill.”<br />

For case a, steam transported uphill while oxygen<br />

transported downhill; the opposite occurred in case b.<br />

limited because <strong>of</strong> the challenges associated with liquid membranes. These challenges include<br />

instability <strong>of</strong> the membrane and/or the carrier chemistry.<br />

Using the coupled defect transport inherent in a proton conducting perovskite ceramic, we have<br />

developed a membrane capable <strong>of</strong> simultaneously transporting water and oxygen, including the<br />

uphill transport <strong>of</strong> either chemical species. This is the first known example <strong>of</strong> a solid-state coupled<br />

transport membrane and is the first <strong>of</strong> any type to operate at high temperatures. The unique features<br />

<strong>of</strong> this system alleviate many <strong>of</strong> the drawbacks associated with liquid membranes and opens up the<br />

possibility <strong>of</strong> new membrane separation applications, including air separation or natural gas<br />

reforming using waste steam and heat streams, as well as the ability to controllably “pump” or<br />

“gate” the permeation <strong>of</strong> select gas species using conjugate chemical potential gradients.<br />

1. Schultz, S. & Curran, P. Coupled transport <strong>of</strong> sodium and organic solutes. Physiological Reviews 50, 637 (1970).<br />

2. Kondepudi, D. K. & Prigogine, I. Modern thermodynamics from heat engines to dissipative structures. Wiley,<br />

(2002).<br />

3. Babcock, W. C., Baker, R. W., Lachapelle, E. D. & Smith, K. L. Coupled transport membranes II: : The<br />

mechanism <strong>of</strong> uranium transport with a tertiary amine. Journal <strong>of</strong> Membrane Science 7, 71-87, (1980).<br />

4. Baker, R. W. Membrane technology and applications. (Wiley, 2004).<br />

5.<br />

Baker, R. W., Tuttle, M. E., Kelly, D. J. & Lonsdale, H. K. Coupled transport membranes : I. Copper separations.<br />

Journal<br />

<strong>of</strong> Membrane Science 2, 213-233, (1977).<br />

- 20 -


O 20<br />

Proton migration at Σ5 (310)/[001] tilt grain boundary in Y-doped BaZrO3<br />

Cheonan Kim<br />

Korea University <strong>of</strong> <strong>Technology</strong> and Education<br />

The energy barriers for proton migration in polycrystalline Y-doped BaZrO3 (BZYO) have been<br />

experimentally studied: the energy barrier for bulk was about 0.45 eV [1] and that for grain<br />

boundary was about 0.6 eV [2]. The energy barrier for proton migration in bulk BZYO has been<br />

calculated and compared with the experimentally obtained values [3]. Since the practical proton<br />

conductors are polycrystalline, and therefore protons should migrate across the grain boundaries,<br />

there has been a big demand for the energy barrier calculation about the grain boundaries <strong>of</strong> BZYO.<br />

We studied the proton migration at Σ5 (310)/[001] tilt grain boundary in BZYO using density<br />

functional theory (DFT), since the tilt grain boundary was commonly observed experimentally. An<br />

optimum Σ5 (310)/[001] tilt grain boundary was first constructed and a preferential site for an Y+3<br />

ion among the Zr sites was calculated. There were one Zr site on the grain boundary and three other<br />

Zr sites near it. The Y+3 ion preferred the first nearest Zr site <strong>of</strong>f the grain boundary among the<br />

three other Zr sites. There were five paths for proton to migrate across the grain boundary with the<br />

energy barriers in the range <strong>of</strong> 0.18 ~ 1.06 eV. However, the energy barriers for proton to arrive at<br />

the grain boundary were higher than those that needed to migrate across it. Therefore, the overall<br />

energy barrier for the easiest proton migration across the grain boundary was about 0.55 eV that<br />

was well matched with the experimentally obtained one [2].<br />

Reference<br />

[1] A. Braun, et al., J. Appl. Electrochem., 39, 471 (2009).<br />

[2] R. B. Cervera, et al., Solid State Ionics, 179, 236 (2008).<br />

[3] B. Merinov, et al., J. Chem. Phys., 130, 194707 (2009).<br />

- 21 -


O 21 (K)<br />

On the path dependence <strong>of</strong> the open-cell voltage <strong>of</strong> a galvanic cell involving<br />

a ternary or higher compound with multiple mobile ionic species under<br />

multiple chemical potential gradients<br />

Han-Ill Yoo a) and Manfred Martin b)<br />

a) WCU Hybrid Materials Program,<br />

Department <strong>of</strong> Materials Science and Engineering,<br />

Seoul National University, Seoul <strong>15</strong>1-744, Korea<br />

b) <strong>Institute</strong> <strong>of</strong> Physical Chemistry,<br />

RWTH Aachen University, D-52056 Aachen, Germany<br />

It is well known that the open-circuit<br />

voltage (U) <strong>of</strong> a galvanic cell involving a<br />

binary compound, or a higher compound<br />

with a single kind <strong>of</strong> mobile ionic species,<br />

is a state property under a gradient <strong>of</strong><br />

chemical potential <strong>of</strong> the mobile<br />

component. It is not so transparent,<br />

however, when involving a ternary or<br />

higher compound with two or more kinds <strong>of</strong><br />

mobile ions under multitude chemical<br />

potential gradients <strong>of</strong> those mobile<br />

components. We clarify this issue with a<br />

proton-oxide ion-hole mixed-conductor<br />

oxide under chemical potential gradients <strong>of</strong><br />

both water and oxygen: U is path- and<br />

history-dependent, and manifests itself<br />

along the diffusion paths <strong>of</strong> the two mobile<br />

components H and O under given boundary<br />

conditions (see Fig. 1).<br />

log( a H2 O )<br />

- 22 -<br />

-2<br />

-3<br />

-4<br />

-5<br />

-6<br />

U II = 5.7 mV<br />

IIa<br />

electrode'<br />

IIb<br />

U(t���) = 11.1 mV<br />

Ia<br />

U I = 16.6 mV<br />

electrode"<br />

-3 -2 -1 0<br />

log( a ) O2<br />

Fig. 1. The open-cell voltage U <strong>of</strong> an electrochemical<br />

cell, a � , a� SrCe Yb O a �� , a��<br />

is not a<br />

O2 H2O 0.95 0.05 3�� O2 H2O state property, but dependent on paths (I and II)<br />

between the electrode ( ‘) and (“). At the steady state,<br />

the path is chosen to be the diffusion path, leading to a<br />

unique value U(t→∞).<br />

Ib


Development <strong>of</strong> proton conducting SOFCs based on LaNbO4 electrolytes<br />

– Status in Norway<br />

Anna Magrasó 1 , Marie Laure. Fontaine 2 , Guttorm E. Syvertsen 3 , Yngve Larring 2 , Hilde L. Lein 3 ,<br />

Tor Grande 3 , Reidar Haugsrud 1 and Truls Norby 1<br />

1 Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, FERMiO, Gaustadalleen 21, NO-0349 Oslo, Norway<br />

2 SINTEF Materials and Chemistry, N-0314 Oslo, Norway<br />

3 NTNU N-7491 Trondheim, Norway<br />

O 22<br />

High temperature Proton-Conducting Solid Oxide Fuel Cells (PC-SOFCs, Fig. 1) have potential<br />

advantages compared to other fuel cell alternatives and are, as such, <strong>of</strong> interest to future green<br />

energy technologies. Ideally, electrolytes for these types <strong>of</strong> cells should exhibit pure proton<br />

conductivity as well as long-term stability in acidic atmospheres. Acceptor substituted LaNbO4<br />

exhibit proton conductivity in wet atmospheres in the order <strong>of</strong> ~10 -3 S/cm, for 1% Ca-doped<br />

LaNbO4. To reach acceptable performances in a PC-SOFC based on this material, the rather modest<br />

conductivity requires an electrolyte thickness in the micron range. However, with a long-term<br />

chemical stability and a proton transference number close to unity, LaNbO4 may still be interesting<br />

for high temperature fuel cell applications - in contrast to the more investigated acceptor-doped<br />

BaCeO3 that shows instability towards acidic atmospheres.<br />

Development <strong>of</strong> a complete fuel cell system based on LaNbO4 has been underway, with<br />

collaborative efforts between Norwegian partners: SINTEF, Norwegian University <strong>of</strong> Science and<br />

<strong>Technology</strong> and the University <strong>of</strong> Oslo. This comprises identifying materials for the electrodes, the<br />

interconnect and the sealing, optimization <strong>of</strong> the microstructures <strong>of</strong> all cell components,<br />

development <strong>of</strong> shaping processes and, design <strong>of</strong> the fuel cell stack. We have addressed the crucial<br />

technological issues <strong>of</strong> building and testing a PC-SOFC stack, as well as to reach a comprehensive<br />

fundamental understanding <strong>of</strong> the processes involved - from fabrication and behavior <strong>of</strong> the<br />

individual components to optimization <strong>of</strong> PC-SOFC fuel cell stacks.<br />

Figure 1. Schematics <strong>of</strong> the PC-SOFC assembly.<br />

The present contribution will review current developments <strong>of</strong> these systems generated within a<br />

portfolio <strong>of</strong> several project financed through internal University funding and by the Research<br />

Council <strong>of</strong> Norway. The support is gratefully acknowledged.<br />

- 23 -


Charge Transfer Protonation <strong>of</strong> a Manganese-Doped Strontium Zirconate<br />

Hiroshige Matsumoto, Takaaki Sakai, Yuya Kawasaki, Yasuhiko Sato and Tatsumi Ishihara<br />

O 23<br />

INAMORI Frontier Research Center, Kyushu University, 744 Motooka, Nishiku, Fukuoka 819-0395, JAPAN<br />

Dept. Appl. Chemistry, Faculty <strong>of</strong> Engineering, Kyushu University, 744 Motooka, Nishiku, Fukuoka 819-0395, JAPAN<br />

The proton conduction in acceptor-doped oxides is resulted from the hydration <strong>of</strong> the oxide ion<br />

vacancies. The hydration takes place by the following two steps: (i) an ambient water molecule<br />

sprits into two protons and one oxide ion on the surface <strong>of</strong> the oxide and (ii) the protons and oxide<br />

ion diffuse into the bulk <strong>of</strong> the oxide in ambipolar manner [1]. Thus, the hydration needs enough<br />

diffusivity <strong>of</strong> both protons and oxide ions as well as the negative free energy change <strong>of</strong> the<br />

hydration reaction. The hydration will not occur if the oxide ion has poor diffusivity in the oxide,<br />

and strontium zirconate will be the case [2].<br />

We report in this paper that transition metal dopant, manganese in the present case, can be used<br />

for protonation <strong>of</strong> an oxide which have poor oxygen diffusivity. We can assume manganese<br />

introduced partially to the zirconium site <strong>of</strong> SrZrO3, i.e., SrZr0.9Mn0.1O3-δ for example, being<br />

tetravalent after baked in air, corresponding to the value <strong>of</strong> δ to be zero. When the material is<br />

exposed in reducing atmosphere, e.g. wet hydrogen at a high temperature, manganese thus<br />

introduced will be reduced to the trivalent state. Due to the electrical neutrality, the reduction <strong>of</strong> the<br />

manganese must be accompanied by either desorption <strong>of</strong> oxide ion (Eq. 1) or incorporation <strong>of</strong><br />

cationic species, i.e., proton (Eq. 2).<br />

� �<br />

��<br />

1<br />

2Mn Zr � OO<br />

� 2Mn�Zr<br />

�V<br />

O � O2<br />

(1)<br />

2<br />

�<br />

�<br />

�<br />

2Mn H � 2O � 2Mn�<br />

� 2OH<br />

(2)<br />

Zr � 2 O<br />

Zr<br />

O<br />

In the latter case, reduction <strong>of</strong> the manganese results in the incorporation <strong>of</strong> protons into the oxide.<br />

This will occur preferentially when the diffusivity <strong>of</strong> oxide ion is small. The protons thus<br />

introduced are thermodynamically stable when the hydration reaction is spontaneous (ΔG


Electronic conduction <strong>of</strong> Sr-doped Ce-La monazite ceramics<br />

Hannah L. Ray, Lutgard C. De Jonghe<br />

Department <strong>of</strong> Materials Science and Engineering, University <strong>of</strong> <strong>California</strong>, Berkeley, Berkeley, CA 94720, USA<br />

Materials Science Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA<br />

Cerium orthophosphate and lanthanum orthophosphate are<br />

isostructural, but the conduction properties <strong>of</strong> these two materials<br />

are extremely different. When doped with aliovalent cations<br />

such as Sr, both materials can incorporate and conduct protons<br />

under wet, reducing conditions with conductivities <strong>of</strong> around 10 -4<br />

S/cm at 600 C , with activation energies <strong>of</strong> about 1 eV (Figure<br />

1). However, at high oxygen partial pressures, the total<br />

conductivity <strong>of</strong> CePO4 increases by over an order <strong>of</strong> magnitude,<br />

whereas that <strong>of</strong> LaPO4 decreases [1]. This increase in total<br />

conductivity under oxidizing conditions may be due to a change<br />

in the identity <strong>of</strong> the dominant charge carrier: since the cerium<br />

cation can exist in either the 3+ or the 4+ valence state, the<br />

cerium orthophosphate is thought to conduct electron holes under<br />

oxidizing conditions.<br />

The interaction between protons and electron holes in a mixed<br />

protonic and electronic conductor is interesting because <strong>of</strong> the<br />

exchange between protons and holes <strong>of</strong> the type<br />

X<br />

( SrH )<br />

� X<br />

�h�( Srh) �<br />

�H,<br />

where ( ) X<br />

SrH � represents a<br />

3�<br />

3� 3<br />

Ce Ce<br />

Ce<br />

strontium dopant on a cerium site with a charge-compensating<br />

proton bonded nearby, and ( ) represents a strontium dopant<br />

X<br />

Srh<br />

3<br />

Ce �<br />

Log (Conductivity) (S/cm<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

-6<br />

Conductivity <strong>of</strong> LaPO4 and CePO4<br />

under oxidizing and reducing conditions<br />

O 24<br />

-7<br />

0.9 1 1.1 1.2 1.3 1.4<br />

1000/T (1/K)<br />

2Sr CePO4, oxidizing<br />

2Sr CePO4, reducing<br />

2Sr LaPO4, oxidizing<br />

2Sr LaPO4, reducing<br />

Figure 1.<br />

Compiled total conductivity<br />

measurements <strong>of</strong> CePO4 and LaPO4 in<br />

oxidizing and wet reducing conditions<br />

from [1].<br />

on a cerium site with a charge-compensating hole. This type <strong>of</strong> exchange could potentially release<br />

protons from traps created by coulombic attraction to the dopant, and increase the mobility <strong>of</strong><br />

protons in a material. However, neither the defect chemistry for the incorporation <strong>of</strong> these electron<br />

hole carriers, nor the mechanism by which these carriers move through the material, is well<br />

understood.<br />

Since the cerium and lanthanum orthophosphates are isostructural, an orthophosphate incorporating<br />

both cations forms a solid solution. The change in conduction behavior on increasing the Ce<br />

content <strong>of</strong> a 2% Sr-doped LaPO4 material can clarify the electronic conduction mechanism. One<br />

possibility is a model whereby electron holes hop over a percolative network <strong>of</strong> cerium sites,<br />

essentially creating itinerant cerium valance states. This mechanism would be disrupted if La 3+ ,<br />

which cannot change valence, interrupted the cerium ion network. The total conductivity <strong>of</strong> a series<br />

<strong>of</strong> La-Ce orthophosphate solid solutions is reported under both oxidizing and reducing conditions,<br />

as well as on a step change in atmospheric conditions, and the applicability <strong>of</strong> a percolation model<br />

is discussed.<br />

1. N. Kitamura, K. Amezawa, Y. Tomii, T. Hanada, N. Yamamoto, T. Omata, and S. Otsuka-Yao-Matsuo, Journal <strong>of</strong><br />

the Electrochemical Society, <strong>15</strong>2 (4) A658-A663 (2005)<br />

- 25 -


O 25 (K)<br />

Interfacial Proton Dynamics in PEM: Theory and Molecular Modeling<br />

A. Roudgar, S. Vartak, and M. Eikerling<br />

Department <strong>of</strong> Chemistry, Simon Fraser University<br />

8888 University Drive, Burnaby, BC, Canada<br />

The presentation will focus on theoretical studies <strong>of</strong> proton transport mechanisms in polymer<br />

electrolyte membranes (PEM) for polymer electrolyte fuel cells. The main class <strong>of</strong> protonconducting<br />

PEM channel protons through random networks <strong>of</strong> water-filled nanopores. Hydrated<br />

side chains <strong>of</strong> the ionomer form charged and flexible interfacial layers relative to which protons<br />

move in nanopores; sulfonic acid head groups <strong>of</strong> the side chains release protons into water-filled<br />

channels. We consider a dense 2D array <strong>of</strong> flexible protogenic surface groups (SG) as a model for<br />

studying interfacial correlations and molecular mechanisms <strong>of</strong> proton mobility in polymeric<br />

membranes at minimal hydration. Our quantum mechanical calculations have elucidated hitherto<br />

unknown effects <strong>of</strong> molecular structure and packing density <strong>of</strong> SG on spontaneous ordering, acid<br />

dissociation, water binding, flexibility <strong>of</strong> SG, and proton dynamics. Major discoveries include a<br />

structural transition to a condensed state <strong>of</strong> the minimally hydrated array at a critical packing<br />

density <strong>of</strong> SG (1 SG per 40 Å 2 ), which is accompanied by a hydrophilic-to-hydrophobic transition.<br />

Furthermore, we have explored the mechanism <strong>of</strong> interfacial proton transport at the critical packing<br />

density. Frequency spectra, calculated with Car-Parinello molecular dynamics, suggest that local<br />

concerted tilting-rotation modes <strong>of</strong> SG trigger the lateral proton move. Adopting the metadynamics<br />

method <strong>of</strong> Laio and Parrinello, i we explored an elementary proton move at high SG density and<br />

minimal hydration. Concerted or consecutive sequences <strong>of</strong> elementary proton moves constitute an<br />

ultra-efficient mechanism <strong>of</strong> interfacial proton transport. This can be demonstrated by using proton<br />

transfer theory and a simple model <strong>of</strong> PEM with lamellar morphology. Results <strong>of</strong> calculations will<br />

be discussed in view <strong>of</strong> the design <strong>of</strong> membranes that exhibit high proton conductivity at minimal<br />

hydration.<br />

1 B. Ensing et al., J. Phys. Chem. B 109 (2005) 6687; A. Laio et al., PNAS 99 (2002) 12562.<br />

- 26 -


Proton Transport Mechanism and Pathways in Solid Acids from<br />

Experiment and Theory<br />

Boris Merinov<br />

Materials and Process Simulation Center, <strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, MC 139-74, Pasadena, <strong>California</strong>, USA<br />

O 26<br />

The future development <strong>of</strong> the fuel cell technology relies on production <strong>of</strong> proton-conducting<br />

materials that will allow fuel cells operate at medium temperature. Such materials should be solid<br />

electrolytes which show high protonic conductivity in the temperature range <strong>of</strong> ~<strong>15</strong>0–400ºC. Solid<br />

acids, alkali metal hydrogen sulfates, selenates, phosphates and arsenates, can be considered as<br />

promising candidates [1, 2] since they exhibit superprotonic conductivity (>10-2 S/cm) in the<br />

temperature range <strong>of</strong> ~<strong>15</strong>0– 250ºC [3-5]. Their transport properties are due to the presence <strong>of</strong><br />

particularly mobile acid protons and specific structural characteristics that result in formation <strong>of</strong> a<br />

dynamically disordered hydrogen bond network in the superprotonic phase. The proton diffusion in<br />

solid acids is governed by the two-step Grotthuss-type mechanism involving a proton transfer along<br />

inter-tetrahedral hydrogen bonds (intra-bond motion) and HAO4-group (A=S, Se, P, As)<br />

reorientation [6] at which the existing hydrogen bond is broken and a new one is formed. Four main<br />

processes occurring at different time scales can be distinguished in this mechanism:<br />

1) thermal vibrations <strong>of</strong> the atoms involved in hydrogen bonding,<br />

2) proton transfer from one potential minimum <strong>of</strong> the hydrogen bond to the other,<br />

3) thermal librations <strong>of</strong> the HAO4 groups, and<br />

4) reorientations <strong>of</strong> the HAO4 groups with breakage <strong>of</strong> existing hydrogen bonds and formation<br />

<strong>of</strong> new ones.<br />

Using quantum mechanical (QM) data on the potential energy surfaces, energetics <strong>of</strong> different<br />

proton motions have been estimated and compared with experimental results obtained. Based on<br />

this analysis, the proton transport mechanism in the superprotonic phase <strong>of</strong> solid acids is discussed<br />

and detailed proton diffusion pathways are proposed. A concerted motion <strong>of</strong> the protons is a<br />

characteristic feature <strong>of</strong> this mechanism.<br />

References<br />

[1] S.M. Haile, D.A. Boysen, C.R.I. Chisholm, R.B. Merle, Nature 410, 910 (2001).<br />

[2] D.A. Boysen, T. Uda T, C.R.I. Chisholm, S.M. Haile, Science 303, 68 (2004).<br />

[3] A.I. Baranov, L.A. Shuvalov, N.M. Schagina, JETP Letters 36, 459 (1982).<br />

[4] A.I. Baranov, I.P. Makarova, L.A. Muradyan, A.V. Tregubchenko, L.A. Shuvalov, V.I. Simonov, Sov. Phys.<br />

Crystallogr. 32, 400 (1987).<br />

[5] A.I. Baranov, B.V. Merinov, A.V. Tregubchenko, V.P. Khiznichenko, L.A. Shuvalov, N.M. Schagina, Solid State<br />

36, 279(1989).<br />

[6] R. Blinc, J. Dolinsek, G. Lahajnar. 1. Zupancic, L.A. Shuvalov, A.I. Baranov. Phys. Status Solidi (b) 123, K83 (<br />

1984).<br />

- 27 -


Adsorption processes related to the surface properties <strong>of</strong> LaNbO4<br />

K. Hadidi 1 , O.M. Løvvik 1 , T. Norby 2<br />

1 Department <strong>of</strong> Physics, University <strong>of</strong> Oslo, P.O. Box 1048, Blindern, NO-0316 Oslo, Norway<br />

2 Department <strong>of</strong> Chemistry and Centre for Materials Science and Nanotechnology, University <strong>of</strong> Oslo, P.O. Box 1126,<br />

Blindern, NO-0318 Oslo, Norway<br />

Acceptor-doped lanthanum ortho-niobate (LaNbO4)<br />

based materials are high temperature proton conductors<br />

[1] with a proton conductivity around ~0.001 S cm -1 .<br />

These materials are attractive candidate as an electrolyte<br />

in solid oxide fuel cells (SOFCs) and hydrogen sensors<br />

due to their low interaction with water and carbon<br />

dioxide which exist in the real conditions <strong>of</strong> operation.<br />

As a proton conductor, the surface properties <strong>of</strong><br />

lanthanum niobate related to the adsorption processes are<br />

<strong>of</strong> great interest.<br />

In this study, density-functional band structure calculations<br />

using the Vienna ab-initio simulation package (VASP) [2]<br />

were utilized to investigate the surface properties <strong>of</strong><br />

lanthanum niobate in the monoclinic (low temperature)<br />

phase [3]. From surface energy calculations comparing the<br />

(101), (010), (100) and (001) terminations, the (010) surface<br />

was found to be the most stable surface. The analogy<br />

between the electronic structure <strong>of</strong> bulk and different<br />

surfaces, presented that the electronic structure <strong>of</strong> surface is<br />

influenced by the surface instability. Thus, the adsorption<br />

energy calculations demonstrated that the most stable (010)<br />

surface is repulsive towards hydrogen adsorption (Figure 1)<br />

and<br />

oxygen excess whereas these two processes are more<br />

stable<br />

at the (101) surface.<br />

Adsorption energy (eV)<br />

Adsorption energy (ev)<br />

3.2<br />

2.8<br />

2.4<br />

2<br />

1.6<br />

1.2<br />

0.8<br />

0.4<br />

0<br />

‐0.4<br />

3.8<br />

3.4<br />

3<br />

2.6<br />

2.2<br />

1.8<br />

1.4<br />

1<br />

The most stable site<br />

after relaxation.<br />

(101) Surface<br />

0 1 2 3<br />

The most stable site<br />

after relaxation.<br />

[1] R. Haugsrud and T. Norby, Solid State Ionics 177 (2006), 1129.<br />

[2] G. Kresse and J. Furthmüller, Physical Review B 54 (1996), 11169.<br />

[3] L. Jian and C.M. Wayman, Journal <strong>of</strong> the American Ceramic Society 80 (1997), 803.<br />

- 28 -<br />

0.2 0.7 1.2 1.7 2.2 2.7<br />

Height above surface (Å)<br />

(010) Surface<br />

O 27<br />

O1Nb1<br />

O4Nb1<br />

O3‐top<br />

O3O4<br />

Nb1O3<br />

O8La2<br />

O4‐top<br />

O1‐top<br />

La1Nb1<br />

O2‐top<br />

Nb‐top<br />

re‐O3top<br />

La1‐Nb1‐<br />

La2‐Nb4<br />

Nb1‐O6<br />

Nb1‐Nb2‐<br />

La3<br />

La‐top<br />

O6‐top<br />

La3‐O6<br />

O7‐Nb1<br />

O7‐Nb2<br />

Nb‐top<br />

re‐O6top<br />

Fig. 1 The optimized adsorption energy for<br />

hydrogen at the most stable site <strong>of</strong> the (101)<br />

surface presents the stability <strong>of</strong> this surface<br />

related to the hydrogen adsorption. The most<br />

stable (010) surface is repulsive towards<br />

hydrogen adsorption.


O 28 (K)<br />

High temperature protonic conduction in rare earth phosphates and borates<br />

K. Amezawa 1 , H. Takahashi 1 , A. Unemoto 2 , H. Kuwabara 3 , N. Kitamura 4 and T. Kawada 1<br />

1 Grad. Sch. <strong>of</strong> Environmental Studies, Tohoku Univ., 6-6-01 Aramaki-Aoba, Aoba-ku, Sendai 980-8579, Japan<br />

2 Insti. <strong>of</strong> Multidisciplinary Res. for Adv. Mat., Tohoku Univ., 2-2-1 Katahira, Aoba-ku, Sendai 980-8577, Japan/<br />

3 Japan Fine Ceramics Center, 2-4-1 Mutsuno, Atsuta-ku, Nagoya 456-8587, Japan<br />

4 Dep. <strong>of</strong> Pure and Appl. Chem., Faculty <strong>of</strong> Sci. and Tech., Tokyo Univ. <strong>of</strong> Sci., 2641 Yamazaki, Noda 278-8510, Japan<br />

Rare earth phosphates such as orthophosphate LnPO4 and metaphosphate LnP3O9 are found<br />

to exhibit protonic conduction at elevated temperatures [1, 2] and thus expected as an electrolyte <strong>of</strong><br />

fuel cells, hydrogen activity sensors, and hydrogen separation systems. Similar protonic conduction<br />

has been also demonstrated in rare earth borates [3]. These oxoacid salts-based materials become<br />

protonic conductors under moisturized atmospheres by doping lower valence (divalent) cations into<br />

rare earth sites. High temperature protonic conductors (HTPCs) based on rare earth oxoacid salts, in<br />

general, show relatively dominant protonic conduction and excellent chemical stability against acid<br />

gases, although their conductivities are lower than those <strong>of</strong> proton-conducting perovskite-type<br />

oxides. In this presentation, our recent researches on high temperature protonic conduction in rare<br />

earth phosphates and borates are reviewed.<br />

Mechanism <strong>of</strong> proton dissolution processes into the rare earth phosphates and borates has<br />

been examined by means <strong>of</strong> spectroscopic measurements, e.g. MAS-NMR, FT-Raman, and FT-IR<br />

spectroscopies in addition to conventional electrochemical measurements, e.g. measurements <strong>of</strong><br />

electrical conductivity and transport number [2]. Their defect structures were also investigated by<br />

the first principle calculation [4]. Consequently, in the case <strong>of</strong> rare earth orthophosphates, it was<br />

suggested that substitution <strong>of</strong> divalent cations for rare earth ions leads to the condensation <strong>of</strong><br />

phosphate ions, i.e. formation <strong>of</strong> P2O7 4- from two PO4 3- . Such condensation <strong>of</strong> phosphate ions is<br />

regarded as formation <strong>of</strong> oxygen deficits like formation <strong>of</strong> oxygen vacancies in conventional<br />

proton-conducting oxides. Protons are considered to be dissolved into phosphates forming hydrogen<br />

phosphate ions through the equilibrium between the condensed phosphate ions and ambient water<br />

vapor.<br />

Electrical conduction in LaP3O9 glass was also investigated [5]. LaP3O9 glass had<br />

considerably lower conductivity than LaP3O9 ceramics, and its charge carrier species were not<br />

protons. However LaP3O9 glass became to exhibit relatively higher and protonic conductivity by<br />

partial<br />

crystallization <strong>of</strong> the glass. These results indicated that crystalline state is essentially needed<br />

for<br />

high temperature protonic conduction in phosphates.<br />

References<br />

1. T. Norby and N. Christiansen., Solid State Ionics, 77, 240 (1995) .<br />

2. K. Amezawa, et al., Solid State Ionics, 145, 233 (2001); Electrochem. Solid-State Lett., 7, A511 (2004).<br />

3. K. Amezawa, et al. Solid State Ionics, 175, 575 (2004); H. Takahashi, et al., Solid State Ionics, in press.<br />

4. N. Kitamura, et al., 16th Int. Conf. on Solid State Ionics, P594 (2007).<br />

5. K. Amezawa, J. Am. Ceram. Soc., 88, 321 (2005).<br />

Acknowledgements: This work was supported by the Grant-in-Aid for Scientific Research from<br />

the Ministry <strong>of</strong> Education, Culture, Sports, Science, and <strong>Technology</strong> (MEXT) <strong>of</strong> Japan.<br />

- 29 -


Thermodynamics and proton conductivity trends in fluorite-pyrochlore<br />

structures: A study on lanthanum cerate<br />

O 29<br />

Vasileios Besikiotis 1 , Tor S. Bjørheim 1 , Sandrine Ricote 2 , Christian Kjølseth 1 , Reidar Haugsrud 1<br />

and Truls Norby 1<br />

1 Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, FERMiO, Gaustadallèen 21, NO-0349 Oslo, Norway<br />

2 Fuel Cells and Solid State Chemistry Division, Risø National Laboratory, Technical University <strong>of</strong> Denmark, P.O. 49,<br />

4000 Roskilde, Denmark<br />

What appears to be mixed conducting Ca-doped La2Ce2O7 has been applied as component in<br />

electrodes <strong>of</strong> proton ceramic fuel cells [1], and various ternary pyrochlore structured oxides with the<br />

general formula Ln2B2O7 (Ln=La-Lu and B=Ce, Zr, Sn, Ti) are currently being investigated for<br />

their proton conducting properties. We have studied the effect<br />

<strong>of</strong> A and B-site substitution on the protonic conductivity in the<br />

series La2-yCayCe2-xZrxO7 (y=0, 0.02, 0.1 and x=0, 0.5, 0.75).<br />

DFT calculations have been performed to elucidate trends in<br />

the hydration and mobility enthalpy <strong>of</strong> protons for the<br />

abovementioned series Ln2B2O7.<br />

AC conductivity <strong>of</strong> La2-yCayCe2-xZrxO7 (y=0, 0.02, 0.1 and<br />

x=0, 0.5, 0.75) vs. T, pO2, and pH2O shows that the materials<br />

are essentially ionic conductors in the whole temperature range<br />

T


Synthesis <strong>of</strong> Highly Sr-doped Lanthanum Orthophosphate and<br />

Polyphosphate in Phosphoric Acid Solutions<br />

Naoyuki Hatada, Akiko Kuramitsu, Yoshitaro Nose, and Tetsuya Uda<br />

Department <strong>of</strong> Materials Science and Engineering, Kyoto University,<br />

Yoshida Honmachi, Sakyo-ku, Kyoto 606-8501, Japan<br />

O 30<br />

In recent years, lanthanum orthophosphate (LaPO4) and lanthanum polyphosphate (LaP3O9)<br />

have received attention because <strong>of</strong> their potential application as solid electrolytes in fuel cells. They<br />

exhibit relatively high proton conductivities when doped with alkaline earth metals, such as Sr [1].<br />

For practical use, however, further enhancement <strong>of</strong> the conductivities is necessary. They are<br />

expected to increase with increasing the Sr doping level. But the solubilities <strong>of</strong> Sr in the phosphates<br />

have been reported to be comparatively low: approx. 2 mol% in LaPO4 [2]; below 5 mol% in<br />

LaP3O9 [3].<br />

To extend the solubility <strong>of</strong> Sr, we employed a low-temperature synthesis method: Synthesis<br />

from phosphoric acid solutions. In our previous experiments, 13 mol% Sr-doped LaPO4 was<br />

obtained. It was also found that the Sr doping level depended on synthesis conditions, with drastic<br />

variation. Therefore, in this research, a systematic approach was taken to reveal the optimized<br />

condition for obtaining highly Sr-doped lanthanum phosphates.<br />

Lanthanum oxide, strontium carbonate, and phosphoric acid were mixed in a crucible and<br />

held at 190 °C to obtain the initial solution. A typical molar composition <strong>of</strong> the initial solutions was<br />

La+Sr : P = 1 : <strong>15</strong>. Then the temperature and water vapor pressure were changed to precipitate<br />

lanthanum phosphates. After held for a desired period, the precipitates were washed with hot water<br />

and dried in air. Finally, the precipitates were analyzed by X-ray diffraction and ICP-AES. The<br />

compositions <strong>of</strong> the solutions were also analyzed by ICP-AES.<br />

Ratio <strong>of</strong> La precipitated as<br />

lanthanum phosphates, namely,<br />

precipitation ratio <strong>of</strong> La, was found to have<br />

a significant influence on the Sr doping<br />

level in the precipitates. As shown in Fig. 1,<br />

the Sr doping level in LaPO4 increased with<br />

decreasing the precipitation ratio <strong>of</strong> La.<br />

Meanwhile, the effect <strong>of</strong> the precipitation<br />

temperature on the Sr doping level was not<br />

significant within the temperature range <strong>of</strong><br />

120 – 250 °C.<br />

5<br />

4 %<br />

Initial Sr concentrations<br />

in solutions<br />

0<br />

0 20 40 60 80 100<br />

Precipitation ratio <strong>of</strong> La (mol%)<br />

[1] T. Norby and N. Christiansen, Solid State Ionics 77 (1995) 240.<br />

[2] K. Amezawa, Y. Tomii, and N. Yamamoto, Solid State Ionics 176 (2005) 143.<br />

[3] K. Amezawa, Y. Uchimoto, and Y. Tomii, Solid State Ionics 177 (2006) 2407.<br />

Sr/(La+Sr) in precipitates (mol%)<br />

<strong>15</strong><br />

10<br />

pH 2 O � 1 atm<br />

8 %<br />

Precipitation<br />

temperature (°C)<br />

Initial<br />

Sr<br />

concentration,<br />

Sr<br />

La+Sr<br />

120 ��<br />

250 ��<br />

(mol%)<br />

4 8<br />

Fig. 1. Sr-doping level in LaPO4 precipitates as a function<br />

<strong>of</strong> precipitation ratio <strong>of</strong> La.<br />

- 31 -<br />

��<br />

��


O 31 (K)<br />

Characterization <strong>of</strong> Structural Change in PLD-Fabricated SrZrO3 Thin Films<br />

Noriko SATA 1 , Sho TAMURA 1 , Yuki NAGAO 1 ,<br />

Hiroyuki KAGEYAMA 2 , Katsumi HANDA 3 , Katsuhiro NOMURA 2 ,<br />

Fumitada IGUCHI 1 , Hiroo YUGAMI 1<br />

1 Graduate School <strong>of</strong> Engineering, Tohoku University, Sendai 980-8579, Japan<br />

2 AIST-Kansai, Midorigaoka 1-8-31, Ikeda, Osaka 563-8577, Japan<br />

3 <strong>Institute</strong> for Chemical Research, Kyoto University, Gokasho, Uji, Kyoto 611‐0011, Japan<br />

Alkaline earth zirconates based proton-conducting oxides, such as SrZrO3 and BaZrO3, are<br />

promising because <strong>of</strong> their chemical and mechanical stability and very low electronic conduction<br />

even at low PO 2. One can expect to reduce the resistance <strong>of</strong> these materials in thin films, however, it<br />

is commonly reported that the conductivity decreases by one or two orders <strong>of</strong> magnitude and the<br />

activation energy increases in SrZrO3 and BaZrO3 thin films. The origin <strong>of</strong> this degradation in<br />

proton conductivity is not yet clearly explained.<br />

We have been working on the pulsed laser deposition (PLD) process, which is widely used to<br />

obtain epitaxial oxide thin films. Epitaxial growth is realized on an appropriate single crystalline<br />

substrate at a sufficiently high temperature, so that rearrangement <strong>of</strong> the deposited atoms is<br />

accelerated. It is found that in a suitable PLD condition, room temperature deposition on a single<br />

crystal substrate can provide epitaxial thin films by post-annealing (LT-PLD). We have found that<br />

the epitaxial thin films obtained by LT-PLD process have different structural and electrical<br />

properties from those <strong>of</strong> bulk crystals and thin films fabricated by the conventional (hightemperature)<br />

PLD process (HT-PLD). In-situ X-ray diffraction (XRD) measurements <strong>of</strong> the postannealing<br />

process <strong>of</strong> SrZrO3(SZO) in the LT-PLD showed that the epitaxial crystallization starts at<br />

560C, which is much lower than that <strong>of</strong> HT-PLD process. It is also found that 5-mol% Yttrium<br />

doping reduces the crystallization temperature by 20 degree C, which might be due to crystalline<br />

nucleation. Local environmental information in the thin films will be valuable to understand the<br />

crystallization process in LT-PLD and the origin <strong>of</strong> the difference in electrical conduction. In<br />

addition to the XRD analysis, X-ray absorption fine structure (XAFS) is performed to investigate<br />

the structural difference in the SZO thin films. Fluorescence EXAFS spectra <strong>of</strong> the Sr and Zr Kedges<br />

were obtained by a Lytle type ionization chamber at BL-9C, Photon Factory, KEK, Japan.<br />

XANES spectra <strong>of</strong> the Sr K edge indicate that the configuration (CN, bond length) <strong>of</strong> the oxygen<br />

atoms around Sr is different not only in as-deposited thin film but also in post-annealed SZO thin<br />

films <strong>of</strong> LT-PLD process compared to the crystalline SZO fabricated by HT-PLD. EXAFS analysis<br />

<strong>of</strong> those data has confirmed that the as-deposited SZO thin film has no long-range order, which is<br />

correspondent to the results in TEM observation and XRD. Zr-O inter-atomic distances obtained by<br />

the EXAFS analysis are not very different among all specimens. On the contrary, it is found that Sr-<br />

O inter-atomic distances are different from each other. Though one can obtain precise lattice<br />

parameters by XRD, the geometry <strong>of</strong> the oxygen atoms is not reflected very much. It is not easy to<br />

discuss quantitatively about the structural change in the SZO thin films by the XAFS measurements,<br />

however, the analysis has indicated a possible explanation <strong>of</strong> those data by the tilt angle change <strong>of</strong><br />

the ZrO6 octahedron in LT-PLD thin films on MgO substrates from the other crystalline SZO<br />

specimens.<br />

- 32 -


Water uptake and conduction property <strong>of</strong><br />

nano-grained yttria-doped zirconia<br />

Yasuaki Akao a , Shogo Miyoshi a , Naoaki Kuwata b , Junichi Kawamura b ,<br />

Yukiko Oyama a , Shu Yamaguchi a<br />

a: Dep. <strong>of</strong> Materials Engineering, The University <strong>of</strong> Tokyo, 7-3-1 Hongo, Bunkyo-ku, 113-8656 Tokyo, JAPAN<br />

b: IMRAM, Tohoku University, 2-1-1 Katahira, Aoba-ku, 980-8577 Sendai, JAPAN<br />

O 32<br />

One <strong>of</strong> the approaches to development <strong>of</strong> new ionic conductors would be making use <strong>of</strong> interfaces<br />

such as grain boundary and surface. The transport properties along the interfaces come to appear as<br />

macroscopic properties, when the materials are fabricated in a nano-grained structure. In our<br />

previous study 1 , protonic conduction at low temperatures has been observed on nano-grained<br />

(Zr,Y)O2 (YSZ) compacts, which have been fabricated via a combination <strong>of</strong> low-temperature<br />

synthesis <strong>of</strong> nano-powder and room-temperature high-pressure (4GPa) compaction. Protonic<br />

conduction in nano-YSZ has also been reported by Kim et al. 2 , while the specimens were prepared<br />

by spark plasma sintering. While those observations indicate a good feasibility <strong>of</strong> developing a<br />

new kind <strong>of</strong> nano-grained protonic conductors, <strong>of</strong> which the bulk properties involve negligible<br />

proton solubility or conductivity, the detailed mechanism <strong>of</strong> such low temperature proton<br />

conductivity has not been well understood. In the present study, considerations are made on<br />

thermal stability <strong>of</strong> protonic species at and in the vicinity <strong>of</strong> interfaces in nano-grained YSZ, and<br />

their roles in protonic conduction are discussed.<br />

The microstructure observed with FE-SEM indicates the existence <strong>of</strong> nano-pores, which are<br />

regarded to originate from condensation <strong>of</strong> adsorbed H2O molecules on the powder surface during<br />

the room-temperature compaction. Such a microstructure may hold free H2O molecules in its<br />

structure, as well as OH group terminating the surface and H2O molecules hydrogen-bonded to<br />

them. 1H MAS-NMR measurements distinguish three kinds <strong>of</strong> proton in the materials. The<br />

difference in thermal stability <strong>of</strong> those protonic species is observed with a thermal desorption<br />

spectroscopy (TDS) and in-situ FT-IR analysis. One <strong>of</strong> the important observations is that the<br />

species with the high thermal stability, which is regarded as OH group terminating the surface,<br />

undergoes a H/D exchange even at low temperatures. The results <strong>of</strong> conductivity measurements<br />

including the H/D isotope effect suggest that protonic conduction prevails over oxide ion<br />

conduction bellow ca 873 K. The isotope effect remains prominent even at low temperatures down<br />

to 60C, indicating that the dominant proton migration process is essentially based on the Grotthuss<br />

mechanism. This conduction behavior is consistent with the observations with TDS and in-situ FT-<br />

IR, i.e., significant mobility <strong>of</strong> protons dissociated from OH group, while the dissociation process<br />

usually seems rather unfavorable at low temperatures. To understand the observed behavior, we<br />

propose that the dissociation <strong>of</strong> protons from the surface-terminating OH groups is assisted by an<br />

interaction with H2O molecules which adsorb on the OH groups with hydrogen bond.<br />

1.<br />

Y. Akao et al., The 14th International Conference on Solid State Protonic Conductors, (2008).<br />

2.<br />

S. Kim et al., Advanced Materials 20 (2008) 556.<br />

- 33 -


Proton conductivity in acceptor doped LaVO4<br />

Morten Huse, Tor S. Bjørheim, Reidar Haugsrud<br />

Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, FERMiO, Gaustadallèen 21, NO-0349 Oslo, Norway<br />

O 33<br />

State-<strong>of</strong>-the-art proton conductors comprise alkaline-earth based perovskites, e.g. BaCeO3, SrCeO3<br />

and BaZrO3. These materials are reactive towards CO2 and other acidic species and/or exhibit high<br />

grain boundary impedances [1, 2], and there is thus an interest in new stable proton conductors. In<br />

this contribution we report on the electrical properties <strong>of</strong> LaVO4, a new proton conductor.<br />

Electrical characterization <strong>of</strong> nominally undoped LaVO4, La0.99Ca0.01VO4 and La0.95Ca0.05VO4 was<br />

performed in various partial pressures <strong>of</strong> oxygen, water vapor and hydrogen isotopes, from 300 to<br />

1100 °C by impedance spectroscopy, AC conductivity measurements (10 kHz) and EMFmeasurements.<br />

Relative permittivity was measured in a custom-made aluminium sample holder.<br />

XRD, SEM and EDS were used for structural, micro structural and compositional analysis.<br />

Generally, acceptor doped LaVO4 was a pure ionic conductor in oxidizing atmospheres in the entire<br />

measured temperature range, dominated by proton conductivity at low temperatures (T


O 34 (I)<br />

Quasielastic neutron scattering (QENS) applied to proton dynamics in<br />

bulk and confinement – the case <strong>of</strong> phosphoric acid<br />

B. Frick 1 , A. Thomas 2 , S. Lyonnard 3 , L.Vilciauskas 4 , K.D. Kreuer 4<br />

1 Institut Laue-Langevin, F-38042 Grenoble, France; 2 MPI for Colloid Research, Golm, Germany, 3 CEA,<br />

INAC/SPrAM/PCI, Grenoble,France; 4 MPI for Solid State Research, Stuttgart, Germany<br />

Quasielastic neutron scattering (QENS) has been proven to be a powerful tool for studying the local<br />

fast dynamics <strong>of</strong> materials with high proton conductivity [1, 2]. We review first more generally the<br />

information content <strong>of</strong> QENS for proton dynamics in bulk and confinement and report on progress<br />

in neutron scattering instrumentation.<br />

In the following we will then discuss recent QENS data on bulk phosphoric acid (PA) and on<br />

geometrically confined PA to get thereby a deeper insight into the proton conduction mechanisms<br />

that are still not elucidated in detail, even though PA is today an important constituent <strong>of</strong> new<br />

proton conducting membranes [3].<br />

For anhydrous bulk phosphoric acid (PA), the ns- and ps- dynamics is investigated in the<br />

temperature range from 2K to 430K (bulk PA : Tm ~ 3<strong>15</strong>K). We interpret the data by fitting to<br />

different models, e.g. jump diffusion models and by comparing to ab-initio MD-simulations carried<br />

out by one <strong>of</strong> the authors [4, 5]. Proton conduction in PA is very sensitive towards perturbations,<br />

but it is not clear to which extend these are a consequence <strong>of</strong> chemical interaction or <strong>of</strong> geometrical<br />

confinement. For this purpose PA was also confined in nano-porous SiO2 matrices with cylindrical<br />

pores <strong>of</strong> average diameters <strong>of</strong> 2.3 nm, 4.9 nm and 9.8 nm and by introducing inert and chemically<br />

interacting molecules into the hydrogen bond network <strong>of</strong> PA. All samples were investigated on the<br />

backscattering spectrometer IN16 with a sub-µeV energy resolution and also on the time-<strong>of</strong>-flight<br />

instrument IN5 at 12Å and 5Å with a resolution <strong>of</strong> 20µeV and 80µeV, thus covering 4 orders <strong>of</strong><br />

magnitude in energy transfer from relaxations in the µeV range to vibrations in the several 10meV<br />

energy range. We discuss the differences in the dynamics <strong>of</strong> confined and bulk PA as well as its<br />

dependence<br />

on pore size. The data evaluation under progress indicates an astonishingly weak<br />

influence<br />

<strong>of</strong> confinement in pores <strong>of</strong> 2-10 nm diameter on the dynamics <strong>of</strong> PA.<br />

1. P. Perrin, Lyonnard and S.M. Volino, Journal <strong>of</strong> Physical Chemistry C 111, 3393-3404 (2007).<br />

2. Th. Dippel, K. D. Kreuer, J. C. Lassègues, and D. Rodriguez, Solid State Ionics 61, 41–46 (1993).<br />

3. K. D. Kreuer, S.J. Paddison, E. Spohr, M. Schuster, Chemical Reviews 104 (2004) 4637.<br />

4. L. Vilciauskas, S. J. Paddison, and K.-D. Kreuer, J. Phys. Chem. A 113(32), 9193–9201 (2009).<br />

5. L. Vilciauskas et al., work in progress<br />

- 35 -


O 35 (I)<br />

Ambient and high-pressure synchrotron x-ray diffraction studies <strong>of</strong><br />

heating-induced structural modifications in phosphate-based solid acids<br />

Cristian E. Botez<br />

Department <strong>of</strong> Physics, University <strong>of</strong> Texas at El Paso, 500 W. University Ave., El Paso, TX 79968, USA<br />

We present results from<br />

synchrotron x-ray powder<br />

diffraction demonstrating that<br />

heating CsH2PO4 (CDP) to about<br />

260�C under a pressure <strong>of</strong> 1 GPa<br />

leads to a polymorphic transition<br />

from its room-tempearture<br />

monoclinic (P21/m) phase to this<br />

compound’s high-tempearture<br />

cubic (Pm3m) modification, while<br />

completely inhibiting CDP’s<br />

dehydration [1]. As CDP becomes<br />

superprotonic at 260�C upon<br />

heating under the same pressure<br />

[2], the above-mentioned finding<br />

adds more unambiguous evidence<br />

to confirm that CDP’s<br />

superprotonic jump is triggered by<br />

a heating induced polymorphic<br />

transition, and not by chemical<br />

modifications. RbH2PO4 (RDP) is<br />

another phosphate-based solid acid<br />

Intensity (arb. units)<br />

80000<br />

60000<br />

40000<br />

20000<br />

0<br />

RbH 2PO 4<br />

T=<strong>15</strong>0 o C<br />

30 40 50<br />

two theta (deg.)<br />

Fig. 1 Rietveld fit (solid line) to x-ray diffraction data (circles), and<br />

refined positions <strong>of</strong> the non-hydrogen atoms in the crystal structure <strong>of</strong><br />

intermediate-temperature RDP. The lower trace is the difference<br />

between the observed and the calculated patterns and the vertical bars<br />

are the Bragg reflection markers. Spacegroup is P 21/m and lattice<br />

parameters are a=7.694Å, b=6.199Å, c=4.774Å and �=109.02 deg.<br />

that exhibits a jump <strong>of</strong> its proton conductivity on heating [3], but, for RDP, much less is known (or<br />

agreed upon) in terms <strong>of</strong> the structural changes underlying its superprotonic behavior. We discuss<br />

our temperature-resolved experiments [4] aimed at uncovering the structural details <strong>of</strong> the transition<br />

from room-temperature tetragonal (I-42d) RDP to an intermediate-tempearture (120�C-200�C)<br />

monoclinic modification, including the determination <strong>of</strong> the crystal structure <strong>of</strong> the monoclinic<br />

RDP polymorph, and its isomorphism with room-temperature CDP. We also present results from<br />

high-pressure x-ray diffraction measurements that reveal the structural changes that accompany the<br />

jump in the proton conductivity <strong>of</strong> RDP at 327°C under 1GPa <strong>of</strong> pressure. Finally, we report the<br />

crystal structure <strong>of</strong> an intermediate-temperature (190�C-235�C) monoclinic KH2PO4 (KDP) phase,<br />

and discuss this result in the context <strong>of</strong> the reported lack <strong>of</strong> superprotonic behavior <strong>of</strong> this solid<br />

acid.<br />

1. C. E. Botez, J. D. Hermosillo, J. Zhang, J. Qian, Y. Zhao, J Majzlan, R. R. Chianelli, and C. Pantea, J. Chem. Phys.<br />

127, 194701(2007).<br />

2. D.A. Boysen, S.M. Haile, H. Liu, and R.A. Secco, Chem. Mater. <strong>15</strong>, 727(2003).<br />

3. D.A. Boysen, S.M. Haile, H. Liu, and R.A. Secco, Chem. Mater. 16, 693(2004).<br />

4. C. E. Botez, H. Martinez, R. J. Tackett, R. R. Chianelli, J. Zhang, and Y. Zhao, J. Phys: Condens. Matter 21,<br />

325401(2009).<br />

- 36 -


Proton dynamics <strong>of</strong> CsH2PO4 and related salts containing organic ions<br />

Hideki Maekawa 1,2,3 , Ayumu Ishikawa and Akira Kudo<br />

1 Graduate School <strong>of</strong> Engineering, Tohoku University, Sendai 980-8579, Japan<br />

2 Center for Interdisciplinary Research, Tohoku University, Sendai 980-8578, Japan<br />

3 CREST, Japan Science and <strong>Technology</strong> Corporation, Kawaguchi 332-0012, Japan<br />

To illuminate ionic conduction behavior <strong>of</strong> super<br />

protonic phase <strong>of</strong> CsH2PO4 (CDP), quasi-elastic<br />

neutron scattering (QENS), pulse field gradient<br />

nuclear magnetic resonance (PFG-NMR) and 1 H<br />

NMR spin-lattice relaxation time (T1)<br />

measurements were carried out. 1 Diffusion<br />

coefficient <strong>of</strong> protons obtained from PFG-NMR<br />

reproduced well the electrical conductivity by using<br />

Nernst-Einstein equation. Comparison <strong>of</strong> PFG-<br />

NMR and T1 provided a hopping distance <strong>of</strong><br />

protons as 2.7 Å, which agreed with the next nearest<br />

proton sites in CDP. On the other hand, QENS data<br />

was fitted well with a jump-diffusion model.<br />

Comparison <strong>of</strong> the jump diffusion coefficient and<br />

the correlation time obtained by QENS provided a<br />

librational correlation time <strong>of</strong> protons in the order<br />

<strong>of</strong> 10 -12 s. The present investigation suggests that<br />

different techniques, QENS and NMR, can observe<br />

O 36 (I)<br />

Fig. 1 Comparison <strong>of</strong> the diffusion<br />

coefficient <strong>of</strong> protons in CDP obtained<br />

from various techniques. 1<br />

protonic motions having different time constants including librational and translational motions<br />

within the hydrogen bonding network <strong>of</strong> CDP.<br />

On the other hand, a series <strong>of</strong> oxo-acidic salt composites containing organic cations has<br />

been prepared and surveyed by protonic conductivity measurement, XRD, TG-DTA in order to<br />

clarify their applicability for electrolytes at intermediate temperatures. All samples were<br />

composed as AX (A:Cs, NH4, CH3NH3, (CH3)2NH2, (CH3)4N, (C2H5)4N X2: HSO4, H2PO4;<br />

X:HSO4, H2PO4). Decrease in the crystallographic transition temperature <strong>of</strong> CsHSO4(CHS) and<br />

CDP have been observed in several composites. Also, protonic conductivities increased by about<br />

one order before the transition temperature. TG-DTA and powder XRD suggest some composites<br />

have stoichiometric compounds.<br />

1. A. Ishikawa, H. Maekawa, T. Yamamura, Y. Kawakita, K. Shibata, M. Kawai, Solid State Ionics, 179, 2345-2349<br />

(2008)<br />

- 37 -


Oxygen Reduction Kinetics at Pt | CsHSO4<br />

by Conducting Atomic Force Microscopy<br />

Mary W. Louie † , Adrian Hightower § , and Sossina M. Haile †‡<br />

† Department <strong>of</strong> Chemical Engineering, ‡ Department <strong>of</strong> Materials Science, <strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, 1200 E.<br />

<strong>California</strong> Blvd., Pasadena, CA 91125, USA<br />

§ Department <strong>of</strong> Engineering, Harvey Mudd College, Claremont, CA 91711<br />

O 37<br />

Oxygen reduction kinetics at the nanoscale Pt | CsHSO4 interface at ~<strong>15</strong>0 ºC in humidified air were<br />

characterized using conducting atomic force microscopy (AFM) combined with A.C. impedance<br />

spectroscopy and cyclic voltammetry. Impedance measurements revealed that oxygen reduction at<br />

Pt | CsHSO4 was comprised <strong>of</strong> two processes, one exponentially dependent and the other weakly<br />

dependent on overpotential. Both processes displayed near-ideal capacitive behavior which<br />

indicates a minimal distribution in their relaxation times. This ideal behavior is taken to be<br />

representative <strong>of</strong> a nanoscale interface for which spatial averaging effects are minimal and,<br />

moreover, enables the rigorous separation <strong>of</strong> multiple processes that could otherwise be convoluted<br />

in measurements using macroscale electrode geometries. Measurements <strong>of</strong> the complete currentvoltage<br />

characteristics <strong>of</strong> the Pt | CsHSO4 interface at various points across the electrolyte surface<br />

reveal a variation <strong>of</strong> the oxygen reduction<br />

kinetics with position. The overpotentialactivated<br />

process was interpreted as a<br />

charge transfer reaction, and within the<br />

Butler-Volmer framework, analysis yielded<br />

exchange coefficients (α) for charge<br />

transfer ranging from 0.1 to 0.6 and<br />

exchange currents (i0) spanning five orders<br />

<strong>of</strong> magnitude. The observed countercorrelation<br />

between i0 and α indicates that<br />

the extent to which the activation barrier<br />

decreases under bias (as indicated by the<br />

value <strong>of</strong> α) depends on the initial magnitude<br />

<strong>of</strong> that barrier under open circuit conditions<br />

(as indicated by the value <strong>of</strong> i0).<br />

Furthermore, the observation <strong>of</strong> such a<br />

correlation across multiple independent sets<br />

<strong>of</strong> data demonstrates the suitability <strong>of</strong><br />

conducting AFM for comprehensive studies<br />

<strong>of</strong> electrochemical reactions at electrolytemetal-gas<br />

boundaries.<br />

-i 0 [A]<br />

10 -13<br />

10 -14<br />

10 -<strong>15</strong><br />

10 -16<br />

10 -17<br />

10 -18<br />

10 -19<br />

10 -20<br />

No. k [N m -1 Probe Nominal<br />

_________________ ]<br />

(1) 5<br />

(2) 5<br />

(3) 5<br />

(4) 1.8<br />

(5) 3.5<br />

(6) 5<br />

0.1 0.2 0.3 0.4 0.5 0.6<br />

Figure 1. Exchange current for charge transfer plotted as a<br />

function <strong>of</strong> exchange coefficient, showing a strong countercorrelation.<br />

Parameters were extracted from a Tafel fit to the<br />

current-voltage characteristics <strong>of</strong> Pt | CsHSO4 at ~<strong>15</strong>0 °C in air.<br />

- 38 -<br />


Impact <strong>of</strong> cation nonstoichiometry on phase behavior, mictrostructure,<br />

water incorporation and proton conductivities in yttrium-doped BaZrO3<br />

Yoshihiro Yamazaki, Chih-Kai Yang, Raul Hernandez-Sanchez, and Sossina M. Haile<br />

Materials Science, <strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, 1200 E <strong>California</strong> Blvd., Pasadena, CA 91125, USA<br />

Barium zirconate, specifically that doped with yttrium, is a<br />

prominent proton conducting electrolyte because <strong>of</strong> its high<br />

bulk proton conductivity, chemical stability, and mechanical<br />

robustness. The material presents challenges, however,<br />

because <strong>of</strong> difficulties in attaining reproducible proton<br />

transport properties, and a wide scatter in bulk conductivity<br />

has been reported in the literature (Figure 1). The question<br />

addressed here is the origin <strong>of</strong> such scatter. In an earlier<br />

work, it was proposed that barium loss during high<br />

temperature sintering is responsible [1]. An alternative<br />

explanation based on the separation <strong>of</strong> perovskite into<br />

distinct phases has also been proposed [2]. Here, we report<br />

the phase behavior, microstructure, water uptake, and proton<br />

conductivities in 10 and 20 at% yttrium-doped barium<br />

zirconates, in which cation nonstoichiometry (Ba deficiency,<br />

x,<br />

in Ba1-xZr0.9Y0.1O3-δ and Ba1-xZr0.8Y0.2O3-δ) is precisely<br />

controlled<br />

using chemical solution synthesis [3].<br />

X-ray diffraction analysis <strong>of</strong> the Ba1-xZr0.8Y0.2O3-δ series<br />

reveals the presence <strong>of</strong> a single perovskite phase up to x =<br />

0.06, and the decrease in cell constant with increasing x<br />

O 38<br />

Fig. 1 Increasing Ba deficiency, x, in<br />

Ba1-xZr0.8Y0.2O3-δ lowers bulk proton<br />

conductivity. Data points indicate<br />

current measurements and solid lines<br />

without data points represent selected<br />

literature results.<br />

indicates that the barium deficiency is accommodated largely by partial incorporation <strong>of</strong> the yttrium<br />

on the A-site. At greater barium deficiency, yttria-rich phases (Zr-doped Y2O3 and Y-doped ZrO2)<br />

are observed in conjunction with essentially undoped barium zirconate. In combination with the<br />

known binary phase behavior, these results enable construction <strong>of</strong> the ternary phase diagram BaO-<br />

ZrO2-YO1.5. Proton conductivity is found to decrease monotonically with increasing barium<br />

deficiency. In the single-phase region this attributed to the dopant partitioning, as evidenced by a<br />

decrease in water uptake measured by thermogravimetry in both the Ba1-xZr0.9Y0.1O3-δ and<br />

Ba1-xZr0.8Y0.2O3-δ series. The sharp decrease in conductivity upon traversing the two phase region<br />

reflects the low H2O uptake in fully non-doped barium zirconate. Thus, precise control <strong>of</strong> the cation<br />

stoichiometry indicates the requirements for obtaining reproducible bulk proton conductivity in<br />

yttrium-doped barium zirconates, while revealing mechanisms by which Ba deficiency lowers the<br />

bulk proton conductivity. The potential for utilizing anomalous X-ray scattering (AXS) to directly<br />

determine the extent <strong>of</strong> dopant partitioning is discussed.<br />

1. P. Babilo, T. Uda and S.M. Haile, Journal <strong>of</strong> Materials Research 22 (2007), 1322.<br />

2. A.K. Azad, C. Savaniu, S.W. Tao, S. Duval, P. Holtappels, R.M. Ibberson and J.T.S. Irvine, Journal <strong>of</strong> Materials<br />

Chemistry 18 (2008), 3414.<br />

3.<br />

Y. Yamazaki, R. Hernandez-Sanchez and S.M. Haile, Chemistry <strong>of</strong> Materials 21 (2009), 2755.<br />

- 39 -


Cost-effective solid-state reactive sintering method for proton-conducting<br />

ceramics<br />

Jianhua Tong, Daniel Clark, Ryan O’Hayre<br />

Metallurgical & Materials Engineering, Colorado School <strong>of</strong> Mines, <strong>15</strong>00 Illinois Street, Golden, CO 80401, USA<br />

O 39<br />

High-performance proton-conducting yttrium-doped barium zirconates and cerates have been<br />

successfully fabricated from inexpensive raw carbonate and oxide precursor powders using a<br />

recently developed solid-state reactive sintering (SSRS) process. This SSRS process can be briefly<br />

described as follows. With the addition <strong>of</strong> small amount (less than 2wt%) <strong>of</strong> a sintering additive to<br />

raw material powders, high-quality phase-pure, dense perovskite ceramics can be produced in a<br />

straightforward process involving a single ball-milling, dry pressing, and high-temperature sintering<br />

procedure. Compared to conventional polymer sol-gel or conventional solid reaction methods, this<br />

modified technique reduces materials costs by a factor <strong>of</strong> 10 and simplifies fabrication since the<br />

processes <strong>of</strong> perovskite phase formation, pellet densification, and grain growth all occur in a single<br />

sintering step, greatly decreasing fabrication time and cost. In addition, this technique greatly<br />

enhances the ability to obtain high-quality, dense, large-grain sized ceramics, bringing the<br />

commercialization <strong>of</strong> proton conducting ceramics closer to reality.<br />

Exemplifying the capabilities <strong>of</strong> this approach, fully dense, large-grained (~5μm) pellets <strong>of</strong><br />

BaZr0.8Y0.2O3-δ (BZY20) have been produced at the remarkably low sintering temperature <strong>of</strong><br />

1350 o C. An excellent proton conductivity <strong>of</strong> 33mS/cm was achieved for 2wt% NiO modified<br />

BZY20 pellets under room temperature water humidified UHP Ar atmosphere at 600 o C. Moreover,<br />

the mechanisms <strong>of</strong> the formation, ceramic densification, and grain growth were fully explored<br />

through a suite <strong>of</strong> comprehensive characterization. The second phase <strong>of</strong> BaY2NiO5 formed from<br />

reaction <strong>of</strong> the sintering additive and raw materials during the SSRS process is believed to play a<br />

critical role in the densification process. The SSRS technique has also been applied to NiO modified<br />

BaCe0.8Y0.2O3-δ (BCY20), where again BaY2NiO5 second-phase formation was identified and<br />

implicated in the perovskite phase conversion and densification process. In the BCY20 system, the<br />

second-phase BaY2NiO5 was found to preferentially locate in grain boundary regions and nanosized<br />

nickel metal particles were also found to enrich in the grain boundary region after reduction <strong>of</strong> the<br />

ceramic in H2, leading to a significant commensurate increase in conductivity. The proton<br />

conductivity for a BCY20 pellet modified with 1wt% NiO was determined to be 2.7mmS/cm at<br />

500 o C under room temperature water humidified UHP Ar atmosphere, which was 7 times higher<br />

than the proton conductivity for a control sample obtained from high-quality polymeric sol-gel<br />

powder under identical conditions.<br />

- 40 -


Solid State NMR studies <strong>of</strong> doped BaSnO3 and BaZrO3 protonic<br />

conductors: Defect trapping and ionic mobility<br />

Lucienne Buannic, 1 Frédéric Blanc, 1 Clare P. Grey 1,2<br />

1 Department <strong>of</strong> Chemistry, State University <strong>of</strong> New York, Stony Brook, NY 11794-3400, USA<br />

2 Department <strong>of</strong> Chemistry, University <strong>of</strong> Cambridge, Cambridge, UK<br />

One <strong>of</strong> the key problems that prevent the extensive use<br />

<strong>of</strong> Solid Oxide Fuel Cells (SOFC) is the high operating<br />

temperatures - 800 to 1000 °C - required to activate the<br />

anionic conduction through the solid electrolyte. New<br />

materials presenting good protonic conductivities could<br />

replace the current electrolytes and lower the operating<br />

temperature from 400 to 200 °C. These protonic<br />

conductors usually adopt a perovskite structure ABO3<br />

where doping <strong>of</strong> the B 4+ cation by a lower valence<br />

cation like Y 3+ or Sc 3+ creates oxygen vacancies in the<br />

structure. Hydration <strong>of</strong> these materials occurs through<br />

the reaction <strong>of</strong> a molecule <strong>of</strong> water and an oxygen<br />

vacancy in order to form hydroxyl groups. These<br />

protonic defects are responsible for the protonic<br />

conductivity in these perovskites. The cation ordering<br />

following substitution on the B site can account for the<br />

vacancy trapping in the dry material which will in turn<br />

influence the hydration property <strong>of</strong> the material.<br />

In order to understand the protonic conductivity<br />

exhibited by these materials, the first step is to explain<br />

cation ordering and vacancy trapping in the dry<br />

structure while the second stage includes looking at the<br />

protonic defects and their location in the hydrated<br />

materials. We examined a series <strong>of</strong> BaZr1-xYxO3-δ and<br />

BaSn1-xYxO3-δ (0.10 ≤ x ≤ 0.50), and BaZr1-xScxO3-δ<br />

(0.05 ≤ x ≤ 0.30) by solid-state NMR, an excellent tool<br />

to observe the different local environments. Study <strong>of</strong> the<br />

dry structures by 17 O, 91 Zr, 119 Sn, 45 Sc and 89 Y showed a<br />

random distribution <strong>of</strong> the substituting cations in BaZr1xScxO3-δ<br />

and BaZr1-xYxO3-δ but a clear ordering in<br />

BaSn1-xYxO3-δ. In addition, preliminary results allowed<br />

us to identify some <strong>of</strong> the protonic conduction pathways<br />

in hydrated BaZr1-xScxO3-δ by 1 H NMR.<br />

O 40<br />

Fig. 1: Variable Temperature 1 H MAS NMR Hahn echo spectra <strong>of</strong> BaZr1-xScxO3-x/2(OH)y. (a) x = 0.05, y = 0.01;<br />

(b) x = 0.<strong>15</strong>, y = 0.10; (c) x = 0.30, y = 0.20. Top: spectra obtained at 300 K., bottom: spectra obtained at 110 K.<br />

- 41 -


The thermodynamics and kinetics <strong>of</strong> the dehydration <strong>of</strong> CsH2PO4 studied<br />

in the presence <strong>of</strong> SiO2<br />

Ayako Ikeda and Sossina M. Haile<br />

Materials Science, <strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, 1200 E <strong>California</strong> Blvd., Pasadena, CA 91125, USA<br />

The thermodynamics and kinetics <strong>of</strong> dehydration and hydration<br />

<strong>of</strong> CsH2PO4 are investigated by thermogravimetric, differential<br />

scanning calorimetry and X-ray diffraction analysis in the<br />

temperature range 200 ~ 400 ºC and water partial pressure range<br />

0.06 ~ 0.89 atm. Addition <strong>of</strong> SiO2 powder to CsH2PO4 prevents<br />

coarsening <strong>of</strong> the solid acid during heat treatment (upper figure)<br />

and accelerates the dehydration and hydration reactions. Despite<br />

the morphological impact, no chemical reaction is observed<br />

between CsH2PO4 and SiO2. Accordingly, use <strong>of</strong> SiO2 permits<br />

access to thermodynamic behavior on moderate timescales.<br />

The<br />

pH<br />

2O<br />

–T phase diagram determined from this work (lower<br />

figure) confirms the occurrence <strong>of</strong> a partially dehydrated liquid,<br />

CsH2-2xPO4-x(l), under high-temperature, high-humidity<br />

conditions, in addition to CsH2PO4(s) and CsPO3(s). The phase<br />

boundaries are each determined with good accuracy as a result<br />

<strong>of</strong> the elimination <strong>of</strong> kinetic effects. The triple point connecting<br />

the three phases is located at p = 0.35 ± 0.2 atm and T = 267.5<br />

H 2O<br />

± 1.0 °C. The thermodynamic stability <strong>of</strong> CsH2PO4 and<br />

CsH2-2xPO4-x(l), the latter <strong>of</strong> which exists over the composition<br />

range x = 0.38 ~ 0.48, were ascertained from hydration studies<br />

carried out under the appropriate temperature and water partial<br />

pressure conditions. The effect <strong>of</strong> grain size, water partial<br />

pressure and temperature on the kinetics <strong>of</strong> hydration from<br />

CsPO3 to CsH2PO4 are reported.<br />

O 41<br />

Figure 1. (upper) Morphology <strong>of</strong><br />

post-dehydration sample <strong>of</strong> SiO2 and<br />

(formerly) CsH2PO4; (lower) p – H 2O<br />

T phase diagram representing the<br />

dehydration behavior <strong>of</strong> CsH2PO4,<br />

compared to earlier literature studies.<br />

1. Y. Taninouchi, T. Uda, Y. Awakura, A. Ikeda, and S. M. Haile, Journal <strong>of</strong> Materials Chemistry 17 (2007), 3182<br />

2. Y. Taninouchi, T. Uda, and Y. Awakura, Solid State Ionics 178 (2008), 1648.<br />

3. Y. Taninouchi, T. Uda, and Y. Awakura, Journal <strong>of</strong> The Electrochemical Society <strong>15</strong>6 (2009), B572<br />

- 42 -


Enhancement <strong>of</strong> proton conductivity<br />

in highly oriented poly(aspartic acid) thin film<br />

Yuki Nagao, *,† Jun Matsui, ‡ Takashi Abe, §<br />

Hitoshi Yamamoto, || Tokuji Miyashita, ‡ Noriko Sata, † Hiroo Yugami †<br />

† Department <strong>of</strong> Mechanical Systems and Design, Graduate School <strong>of</strong> Engineering, Tohoku University, Sendai 980-<br />

8579, Japan, ‡ <strong>Institute</strong> <strong>of</strong> Multidisciplinary Research for Advanced Materials (IMRAM), Tohoku University, Sendai<br />

980-8577, Japan, § Department <strong>of</strong> Mechanical and Production Engineering, Faculty <strong>of</strong> Engineering, Niigata<br />

University, Niigata 950-2181, Japan, || Department <strong>of</strong> Macromolecular Science, Graduate School <strong>of</strong> Science, Osaka<br />

University, 1-1 Yamadaoka Suita, Osaka 565-0871, Japan<br />

The design <strong>of</strong> highly proton-conductive solid electrolytes is<br />

essential to many applications in the field <strong>of</strong> solid-state ionics. One<br />

fundamental approach to creating highly proton-conductive material<br />

is chemical modification. Usually, sulfonic acid groups are used as a<br />

proton conductive group because <strong>of</strong> their excellent proton<br />

conductivity. However, the high acidity <strong>of</strong> sulfonic groups restricts<br />

the polymer backbone to fluoro or aromatic groups, which result in a<br />

high cost for production. So a new method to produce highly protonconductive<br />

material has been desired since a long time. Ionic<br />

conductivity enhancement in 2D systems has been recognized for<br />

more than 20 years. Recently, we have been reported a several<br />

enhancement phenomena <strong>of</strong> the proton conductivity in thin<br />

films.[1,2] An application <strong>of</strong> this phenomenon may show promise <strong>of</strong><br />

a new method to produce highly proton-conductive material. In this<br />

study, we investigated the proton transport properties <strong>of</strong> a partially<br />

protonated poly(aspartic acid)/sodium polyaspartate (P-Asp) thin<br />

Log (� / S cm -1)<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

-6<br />

60-nm-thick film<br />

Pelletized P-Asp<br />

P 01<br />

-7<br />

298 K<br />

-8<br />

40 50 60 70<br />

RH / %<br />

Fig. 1 The difference <strong>of</strong> the<br />

proton conductivities <strong>of</strong> the<br />

pelletized P-Asp and 60-nmthick<br />

film on MgO(100)<br />

substrate at each relative<br />

humidity.<br />

film prepared by spin coating. We also report that the orientation <strong>of</strong> the amide groups, which act as<br />

a proton conducting channel, induced the enhancement effect.<br />

The difference <strong>of</strong> the proton conductivities <strong>of</strong> the pelletized P-Asp and 60-nm-thick film on<br />

MgO(100) substrate at each relative humidity (RH) was shown in Fig. 1. The proton conductivity <strong>of</strong><br />

60-nm-thick thin film exhibits 3.4 × 10 -3 at 298 K and 70 % RH. This conductivity is relatively high<br />

for a proton conductor in which the protons <strong>of</strong> carboxylic acid groups serve as proton sources. The<br />

proton conductivity <strong>of</strong> 60-nm-thick film is ~10 times higher than that <strong>of</strong> the pelletized sample at the<br />

same temperature and RH. An enhancement phenomenon <strong>of</strong> the proton conductivity was also<br />

observed when the 60-nm-thick thin film was prepared on amorphous SiO2 substrate.[2] The<br />

activation energy for proton conduction <strong>of</strong> 60-nm-thick film is 0.53 eV, which becomes 0.12 eV<br />

smaller than that <strong>of</strong> the pelletized P-Asp. [2]<br />

The origin for this enhancement <strong>of</strong> the proton conductivity by the thin film was not due to the<br />

difference in adsorbed water amount in this case, which was determined by quartz crystal<br />

microbalance (QCM) method. Incident-angle dependence <strong>of</strong> polarized IR absorption spectra was<br />

measured to evaluate the orientation <strong>of</strong> molecular chains in the 60-nm-thick film <strong>of</strong> P-Asp on a<br />

MgO substrate. The result suggests that the C=O bonds <strong>of</strong> the amide groups are oriented mostly<br />

perpendicular to substrate surface in the 60-nm-thick film.<br />

1. Y. Nagao, N. Naito, F. Iguchi, N. Sata and H. Yugami, Solid State Ionics 180 (2009), 589.<br />

2. Y. Nagao, F. Iguchi, N. Sata and H. Yugami, Solid State Ionics 181 (2010), 206.<br />

‐ 43 ‐


P 02<br />

Crystallinity and morphology <strong>of</strong> PVDF-HFP-based proton-exchange<br />

membranes embedding polytyrenesulfonic acid-grafted silica particles<br />

M. Maréchal, 1,4*<br />

F. Niepceron 2,3<br />

, J. Bigarré 2<br />

, H. Mendil-Jakani 1 , G. Gebel 1<br />

1) Commissariat à l’Energie Atomique, Matter Science Department, Grenoble, France<br />

2) Commissariat à l’Energie Atomique, Department <strong>of</strong> Materials, Le Ripault, Monts, France<br />

3) Laboratoire des Matériaux Macromoléculaires, INSA Lyon, Villeurbanne, France<br />

4) Laboratoire d’Electrochimie et de Physicochimie des Matériaux et Interfaces, Grenoble, France<br />

Novel proton-exchange membrane (PEM) materials based on a hybrid organic-inorganic<br />

formulation have been investigated [1]. The water retention properties around 100°C is one <strong>of</strong> their<br />

advantages. In the present contribution, we intend to report on the incorporation <strong>of</strong> original acidfunctionalized<br />

inorganic nanoparticles, polystyrenesulfonic acid-grafted silica particles [1, 2], in an<br />

inert PVDF-HFP matrix. The proton conductive characteristics relied exclusively on the particle<br />

phase gives promising results. Membranes with different amounts <strong>of</strong> particles have been prepared<br />

by<br />

evaporation and recasting techniques. These membranes were characterised and their<br />

morphologies<br />

have been investigated.<br />

Intensité / u.a.<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01 0.1 1<br />

q / A -1<br />

Figure: A log-log representation <strong>of</strong> a X-ray scattering curve from a membrane with a 50 wt.%<br />

loading<br />

Finally, the performance <strong>of</strong> membrane-electrode assemblies (MEAs), using selected hybrid<br />

membranes, was evaluated by single cell fuel cell tests. Remarkably, such hybrid membrane systems<br />

exhibited up to 1.2 W/cm 2<br />

, at 80 °C using non-hydrated gas feeds.<br />

[1] F. Niepceron, B.Lafitte, H. Galiano, J. Bigarré, E. Nicol, J-F Tassin, J. Membr. Sci. 338 (2009) 100.<br />

[2] X. Chen, D. P. Randall, C. Perruchot, J. F. Watts, T. E. Patten, T. V. Werne, S. P. Armes, J. Colloid Interface Sci.<br />

257 (2003) 56.<br />

‐ 44 ‐


Proton mobility in tetragonal and monoclinic LaNbO4<br />

through a second-order phase transition<br />

Kazuaki Toyoura*, Harald Fjeld, Reidar Haugsrud, and Truls Norby<br />

Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, FERMiO, Gaustadalléen 21, NO-0349 Oslo, Norway<br />

LaNbO4 is an electrolyte candidate for proton conducting<br />

fuel cells. It exists as two polymorphs, the high<br />

temperature tetragonal (t-LaNbO4) and the low temperature<br />

monoclinic (m-LaNbO4) where the transition between the<br />

two is a second-order displacive phase transition. The<br />

crystal structure <strong>of</strong> m-LaNbO4 changes gradually as a<br />

function <strong>of</strong> temperature. In this work, the relation between<br />

the structural change and the migration enthalpy <strong>of</strong> protons<br />

has been investigated by means <strong>of</strong> first principles<br />

calculations and conductivity measurements.<br />

Protons generally reside around oxide ions in oxides, to<br />

form O-H bonding (~ 1 Å). Therefore, proton migration in<br />

oxides can be considered as a sequence <strong>of</strong> proton jumps<br />

from one oxide ion to another. Our theoretical study<br />

clarified that protons migrate in similar paths for t- and m-<br />

LaNbO4. Figure 1 shows the trajectory <strong>of</strong> the ratedetermining<br />

proton jump in the tetragonal phase. In the<br />

first four (0–3) and the last four (5–8) positions, the<br />

distance between the migrating proton and the first nearest<br />

neighbor oxide ion is kept constant, ~ 1 Å. In position 4, in<br />

contrast, the migrating proton resides not around an oxide<br />

ion, but between two oxide ions. This kind <strong>of</strong> migration,<br />

i.e. rotation and hopping, has been reported in other proton<br />

conducting oxides, such as BaCeO3 and BaZrO3. The<br />

potential barriers are 0.41 eV (39 kJ/mol) in t-LaNbO4 and<br />

0.62 eV (60 kJ/mol) in m-LaNbO4, respectively.<br />

Figure 2 shows the bulk conductivity <strong>of</strong> Sr-doped LaNbO4<br />

as a function <strong>of</strong> inverse temperature. The migration<br />

enthalpy <strong>of</strong> protons is 35 kJ/mol in the tetragonal phase,<br />

which is in good agreement with the computational value,<br />

39 kJ/mol. In the monoclinic region, the experimental<br />

conductivity is gradually approaching the theoretical line<br />

corresponding to the monoclinic phase at 0 K. This is<br />

reasonable because <strong>of</strong> the displacive phase transition. The<br />

migration enthalpy increases gradually with decreasing<br />

temperature,<br />

from 35 to 57 kJ/mol in the range <strong>of</strong> 520-205<br />

° C.<br />

* (Current affiliation) Department <strong>of</strong> Materials Science and Engineering,<br />

Kyoto University, Yoshida, Sakyo, Kyoto 606-8501, Japan<br />

‐ 45 ‐<br />

O<br />

O<br />

O<br />

O<br />

Nb<br />

O<br />

O<br />

7 8<br />

6<br />

5<br />

0<br />

1 2<br />

4<br />

3<br />

Nb<br />

O<br />

O<br />

Fig. 1. The rate-determining path<br />

<strong>of</strong> proton migration in t-LaNbO4.<br />

Fig. 2. The bulk conductivity <strong>of</strong><br />

Sr-doped LaNbO4.<br />

P 03


NMR measurements on proton mobility in nano-crystalline YSZ<br />

Judith Hinterberg a , Alina Adams b , Bernhard Blümich b , Martin Wilkening c , Paul Heitjans c ,<br />

Sangtae Kim d , Zuhair A. Munir d , Roger A. De Souza a , Manfred Martin a<br />

a <strong>Institute</strong> <strong>of</strong> Physical Chemistry, RWTH Aachen University (Germany)<br />

b <strong>Institute</strong> <strong>of</strong> Macromolecular Chemistry, RWTH Aachen University (Germany)<br />

c <strong>Institute</strong> <strong>of</strong> Physical Chemistry and Electrochemistry, Leibniz University <strong>of</strong> Hannover (Germany)<br />

d Department <strong>of</strong> Chemical Engineering and Materials Science, University <strong>of</strong> <strong>California</strong> (USA)<br />

Nano-crystalline cubic 9.5 mol% YSZ shows proton conductivity at room temperature [1,2]. In this<br />

work we report preliminary Nuclear Magnetic Resonance (NMR) results on water saturated, dense,<br />

nanostructured YSZ samples. Static NMR as well as magic angle spinning NMR spectra show two<br />

distinct signals. Their temperature dependent behaviour and their linewidths suggest that one can be<br />

attributed to (free) water adsorbed on the surface <strong>of</strong> the sample and the other one to mobile protons<br />

within the sample. This interpretation is supported by comparison with measurements on a single<br />

crystalline sample. Motional narrowing is observed for the signal originating from protons inside the<br />

sample. The analysis <strong>of</strong> temperature and field dependent T1 relaxation measurements points towards<br />

diffusion<br />

in a confined geometry. This supports our previous experimental results which indicated<br />

conduction<br />

<strong>of</strong> protons along the grain boundaries [3].<br />

1. S.Kim, U. Anselmi-Tamburini, H. Jung Park, M. Martin, Z.A. Munir, Adv.Mat. 20, 556-559 (2008)<br />

2. H.J. Avila-Paredes, J. Zhao, S. Wang, M. Pietrowski, R.A. De Souza, A. Reinholdt, Z.A.Munir, M. Martin, S. Kim,<br />

J. Mater. Chem. 20, 990-994 (2010)<br />

3. S. Kim, H.J. Avila-Paredes, S. Wang, C-T. Chen, R.A. De Souza, M. Martin, Z.A. Munir, PCCP 11, 3035-3038<br />

(2009)<br />

‐ 46 ‐<br />

P 04


P 05<br />

Determination <strong>of</strong> the amount <strong>of</strong> proton in proton conducting alumina by DC<br />

polarization method<br />

Yuji Okuyama 1 , Noriaki Kurita 2 , Daisuke Sato 2 , Hisashi Douhara 2 and Norihiko Fukatsu 2<br />

1 Intellectual property Division, Industry-Academia- Government Collaboration Center,<br />

Nagoya <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, Gokiso-cho, Showa-ku, Nagoya 466-8555, Japan<br />

2 Department <strong>of</strong> Materials Science and Engineering, Graduate School <strong>of</strong> Engineering, Omohi College,<br />

Nagoya <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, Gokiso-cho, Showa-ku, Nagoya 466-8555, Japan<br />

Acceptor doped α-alumina exhibits protonic conduction at high temperature. Although the conductivity <strong>of</strong><br />

alumina is lower than that <strong>of</strong> the other proton conducting oxide, the transport number <strong>of</strong> proton is maintained<br />

higher even at elevated temperature (1073-1673 K). Therefore, this material can be used favorably as the<br />

electrolyte <strong>of</strong> the galvanic-cell type hydrogen sensor for metal melting processes.<br />

The chemical potential pr<strong>of</strong>ile in ionic conductor under polarized state is determined by the continuity <strong>of</strong> ionic<br />

and electronic currents. Internal chemical potential distribution is established uniquely when the chemical<br />

potentials at both electrodes and the applied voltage are given. In this study, the maximum solubility <strong>of</strong> the<br />

dissolved hydrogen C was determined for commercial polycrystalline alumina at temperature range <strong>of</strong> 1273-<br />

� �i<br />

H<br />

<strong>15</strong>73 K by measuring the amount <strong>of</strong> the desorbed hydrogen when the potential pr<strong>of</strong>ile <strong>of</strong> hydrogen in alumina was<br />

changed by changing the applied voltage. The amount <strong>of</strong> desorbed hydrogen was measured by a hydrogen pump<br />

reported elsewhere [1].<br />

The equilibrium reaction for hydrogen dissolution is represented as follow,<br />

� �<br />

1/2H 2 � h � Hi<br />

(1)<br />

For alumina equilibrated with H2O-H2 atmosphere, the relative charge <strong>of</strong> acceptor dopant, M� is almost perfectly<br />

�1 � �1/2<br />

compensated with proton, therefore, the concentration <strong>of</strong> positive hole is given as K C � p . Applying to the<br />

Al<br />

H H<br />

i 2<br />

calculation method by Coudhury and Patterson[2], the distribution <strong>of</strong> hydrogen potential can be represented as<br />

follow,<br />

�1/<br />

2<br />

�1/<br />

2<br />

x � r � p � � � �<br />

H x�<br />

x / A r p x�L<br />

/ A<br />

� �<br />

2 ( )<br />

H<br />

� �<br />

2 ( )<br />

ln ln<br />

�<br />

(2)<br />

� �1/<br />

2 � � �1/<br />

2<br />

L<br />

�<br />

� r � pH<br />

( x�0)<br />

/ A � � r � pH<br />

( x�0)<br />

/ A<br />

2<br />

2 �<br />

where, r is the ratio <strong>of</strong> the current <strong>of</strong> positive hole to that <strong>of</strong> proton. A can be estimated from the mobility <strong>of</strong> proton<br />

and positive hole and the equilibrium constant, K, <strong>of</strong> the reaction represented by Eq. (1).<br />

1 �1<br />

A � m m K<br />

(3)<br />

�<br />

� �<br />

h Hi<br />

The parameter r may be estimated from the applied voltage by using the following relation.<br />

1/<br />

2<br />

RT � p<br />

�<br />

H x�<br />

L � A/<br />

r<br />

� � � 2 ( )<br />

E<br />

�<br />

apply ln (4)<br />

F � 1/<br />

2 �<br />

� pH<br />

( x�<br />

0)<br />

� A/<br />

r<br />

2 �<br />

The value, A, was determined by the emf measurement.<br />

The increment <strong>of</strong> the concentration <strong>of</strong> positive hole upon changing the polarization voltage can then be<br />

calculated based on the change <strong>of</strong> the chemical potential pr<strong>of</strong>ile <strong>of</strong> hydrogen, and which must be equal to the small<br />

amount <strong>of</strong> desorbed hydrogen, �n<br />

� , therefore,<br />

Hi ( E1<br />

�E<br />

2 )<br />

L<br />

L<br />

�1<br />

�<br />

�1<br />

/ 2<br />

�1<br />

/ 2<br />

�n � � K C �S<br />

( ( pH<br />

( ) ( x)<br />

) dx ( pH<br />

( ) ( x)<br />

) dx)<br />

Hi ( E1�E2<br />

)<br />

Hi<br />

�0<br />

2 E �<br />

2 �0<br />

2 E<br />

(5)<br />

1<br />

where S and L are the sectional area and the thickness <strong>of</strong> the sample, respectively. According to Eq.(5), the<br />

0<br />

concentration <strong>of</strong> proton, C , was determined employing the equilibrium constant reported in previous study[3].<br />

The value mol cm -3 �7<br />

2. 0 �1.<br />

6�10<br />

was obtained.<br />

H i �<br />

[1]N.Kurita, K.Otsuka, N.Fukatsu and T. Ohashi, Solid State Ionics, 79(1995) 358-365.<br />

[2]N. S. Coudhury and J. W. Patterson, J. Electrochem. Soc.(1970) 1384.<br />

[3]Y. Okuyama et al. Solid State Ionics 180 (2009)175-182.<br />

‐ 47 ‐


1<br />

Dynamic absorption behavior <strong>of</strong> hydrogen within perfluorosulfonic acid<br />

polymer electrolyte membranes during exposure to water vapor<br />

Bun Tsuchiya 1<br />

, Shinji Nagata 2 , Kesami Saito 2 , Tatsuo Shikama 2<br />

Department <strong>of</strong> General Education, Faculty <strong>of</strong><br />

Science and <strong>Technology</strong>, Meijo University, 1-501, Shiogamaguchi,<br />

Tempaku-ku, Nagoya 468-8502, Japan<br />

stitute for Materials Research, Tohoku University, 2-1-1, Katahira, Aoba-ku, Sendai 980-8577, Japan<br />

2 In<br />

r <strong>of</strong> hydrogen<br />

He 2+ Perfluorosulfonic acid (PFSA) polymer electrolyte membranes (PEMs) have potential for use as<br />

hydrogen separators from water (H2O) for hydrogen gas production systems because <strong>of</strong> having<br />

extremely high protonic proton conductivity on the condition <strong>of</strong> operating temperature around 373<br />

K and high humidity. The property <strong>of</strong> water absorption into the polymer electrolyte membranes is<br />

one <strong>of</strong> the most significant parameters for the development <strong>of</strong> the hydrogen gas production systems<br />

and greatly contributes to increments <strong>of</strong> the protonic conductivity and re-combination rate for<br />

hydrogen molecule formation. Therefore, it is essential to understand dynamic behavio<br />

absorbed within the polymer electrolyte membranes during exposure to water vapor.<br />

The PFSA membranes used in the present study were Aciplex-SF-1004® which were ones <strong>of</strong><br />

typical proton conducting polymers, fabricated by Asahi Kasei Corporation. The Pt electrodes with<br />

a 0.1 �m thick were deposited on both sides <strong>of</strong> the PFSA membranes using an electron-beam<br />

sputtering device. The protonic conduction, water absorption characteristics and hydrogen behavior<br />

at the bulk, radical defects, and interface between the membranes and Pt electrode in the membranes,<br />

exposed immediately at several temperatures <strong>of</strong> 293-393 K from vacuum to air with a relative<br />

humidity <strong>of</strong> approximately 50 %, were investigated in-situ by direct current (DC) resistance and<br />

alternative current (AC) impedance measurements. In addition, the concentration <strong>of</strong> hydrogen (H)<br />

retained near the surface (about 400 nm in depth) <strong>of</strong> the membranes were measured using an elastic<br />

recoil detection (ERD) technique with 2.8 MeV ion probe beams after isochronal annealing for<br />

10 min and exposure to water vapor at 293 K.<br />

The proton conductivities <strong>of</strong> the water vapor-induced PFSA membranes at 293 K gradually<br />

increased with increasing the air exposure time, and became approximately four orders <strong>of</strong><br />

magnitude higher when reached up to saturation at approximately 400 min. The hydrogen content in<br />

the near-surface region <strong>of</strong> the membranes at the exposure time <strong>of</strong> 60 min was also observed to<br />

increase to approximately five times that <strong>of</strong> the initial polymer by means <strong>of</strong> the ERD technique. In<br />

particular, the presence <strong>of</strong> some intrinsic defects such as peroxy free radicals and fluorocarbon<br />

radicals<br />

which have been produced in the membranes significantly play an important role for the<br />

increase<br />

<strong>of</strong> the proton conductivity, and can allow hydrogen to be dissociated<br />

from water, and to<br />

migrate<br />

more smoothly at the interface between the membranes and electrode.<br />

‐ 48 ‐<br />

P 06


P 07<br />

Preparation <strong>of</strong> Multilayered Thin Film Fuel Cell Using Titanium Oxide as<br />

Anodic Catalyst via Layer-by-Layer Assembly<br />

Hisatoshi Sakamoto 1 , Yusuke Daiko 2 , Hiroyuki Muto 1,3 ,<br />

Mototsugu Sakai 3 , and Atsunori Matsuda 1,3<br />

(1) Department <strong>of</strong> Environmental and Life Sciences, Toyohashi University <strong>of</strong> <strong>Technology</strong>,<br />

Hibarigaoka, Tempaku, Toyohashi, Aichi 441-8580, JAPAN<br />

(2) Department <strong>of</strong> Materials Science and Chemistry, University <strong>of</strong> Hyogo,<br />

Shosha, Himeji, Hyogo 670-2101, JAPAN<br />

(3) Department <strong>of</strong> Electrical and Electronic Information Engineering, Toyohashi University <strong>of</strong> <strong>Technology</strong>,<br />

Hibarigaoka, Tempaku, Toyohashi, Aichi 441-8580, JAPAN<br />

Direct methanol fuel cells (DMFCs) have attractive potentials <strong>of</strong> power source for portable<br />

devices such as cell-phones and laptop computers. However, the poisoning <strong>of</strong> catalysts by carbon<br />

monoxide and the fuel cross-over are serious problems for the fuel cell property <strong>of</strong> DMFCs. The Pt-<br />

Ru-supported carbon (C/Pt-Ru) is generally used to anodic catalyst. On the other hand, the<br />

photocatalysts such as anatase titanium oxide (TiO2) and/or noble metal-doped TiO2 were applied to<br />

methanol oxidation and water-splitting for the production <strong>of</strong> hydrogen gas (H2) [1]. In addition, the<br />

TiO2 photocatalyst was employed in an anodic catalyst with Pt-Ru catalyst to promote the methanol<br />

oxidation by UV excitation [2]. Layer-by-layer (LbL) assembly is a convenient technique to form<br />

the polymeric film with a thickness <strong>of</strong> nano-meter order, and that is useful to deposit nano-particles.<br />

In this study, membrane-electrode assembly (MEA) using the TiO2 nano-particles as a catalyst was<br />

prepared on porous glass substrates by LbL method.<br />

The illustration <strong>of</strong> multilayered thin film fuel cell was shown in Fig.1. First, the gold collecting<br />

electrode was sputtered on porous glass substrate, and then anatase TiO2 nano-particles (P-25) were<br />

deposited by the intermediary <strong>of</strong> primer layer. Continuously, a multilayerd polyelectrolyte film<br />

consisting <strong>of</strong> anionic poly(2-acrylamido-2-methyl-1-propanesulfonic acid) (PAMPS) and cationic<br />

poly(allylamine hydrochloride) (PAH) was also deposited on the TiO2 layer. A cathodic catalyst <strong>of</strong><br />

Pt nano-particles was prepared by photoreduction <strong>of</strong> hexachloroplatinic acid (H2PtCl6), and then<br />

deposited on the multilayered electrolyte by using the PDDA as a binder. Finally, an electron<br />

conductive polymer <strong>of</strong> poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) (PEDOT:PSS) was<br />

spin-coated as a cathodic collecting electrode. A<br />

CH3OH aqueous solution was supplied through the<br />

O2 (air) UV light<br />

porous glass substrate, and UV light was irradiated<br />

from the cathodic side to excite the layered TiO2.<br />

Total thickness <strong>of</strong> MEA was about 650 nm. UV<br />

light at 365 nm <strong>of</strong> center wavelength was passed<br />

though about 70 % until the TiO2 layer. Electrical<br />

generation was confirmed and fuel cell performance<br />

was enhanced in proportion to the CH3OH<br />

concentration and the irradiation intensity <strong>of</strong> UV light.<br />

[1] G. Wu et al., Int. J. Hydrogen Energy, 33, (2008) 1243.<br />

[2] K. Drew et al., J. Phys. Chem. B, 109, (2005) 11851.<br />

‐ 49 ‐<br />

Electrode<br />

Cathodic catalyst ; Pt<br />

Multilayered electrolyte<br />

Anodic catalyst ; TiO2 Electrode<br />

Porous glass<br />

Fuel ( CH3OH aq. )<br />

Fig. 1. Diagrammatic illustration <strong>of</strong> multilayered<br />

thin film fuel cell using the anatase TiO2<br />

nanoparticles as a anodic catalyst.


Hydration kinetics <strong>of</strong> proton-conducting zirconates upon a change <strong>of</strong><br />

temperature in wet atmosphere<br />

Jung In Yeon and Han-Ill Yoo<br />

WCU Hybrid Materials Program,<br />

Department <strong>of</strong> Materials Science and Engineering<br />

Seoul National University, Seoul <strong>15</strong>1-744, Korea<br />

It is reported [Kudo et al., Solid State Ionics 179 (2008) 851] that when the temperature <strong>of</strong> proton<br />

conducting SrZr0.9Y0.1O3-δ is lowered from 700 o C to 600 o C in a wet atmosphere, its electrical<br />

conductivity first decreases with decreasing temperature, followed by “extremely” slow,<br />

monotonic increase to a saturated value. This square-root (√ )-shaped variation with time has been<br />

interpreted as being due to the decrease <strong>of</strong> mobility <strong>of</strong> protons with decreasing temperature,<br />

followed by “extremely” slow, single-fold chemical diffusion <strong>of</strong> H2O. We will demonstrate that it<br />

is not the case: the hydration proceeds by decoupled tw<strong>of</strong>old chemical diffusion <strong>of</strong> H and O, but<br />

the effect <strong>of</strong> faster H-diffusion on the conductivity may be overridden by the effect <strong>of</strong> mobility<br />

change when temperature change is slower. Otherwise, the tw<strong>of</strong>old relaxation would manifest<br />

itself. The tw<strong>of</strong>old isothermal and isoactivity hydration/dehydration relaxations are quantitatively<br />

analyzed to evaluate the two chemical diffusivities for the system <strong>of</strong> BaZr0.8Y0.2O3-δ in the range<br />

<strong>of</strong> water activity -4.0≤ log(PH 2O/atm) ≤-2.3 at 700 o C and = -0.67, and in the range <strong>of</strong> temperature<br />

<strong>of</strong> 600 o C to 800 o C in the 0 o C-water-saturated air atmosphere.<br />

‐ 50 ‐<br />

P 08


“In-situ” high temperature neutron diffraction study <strong>of</strong> lanthanum<br />

tungstate: a proton conductor with a fluorite-type structure<br />

Anna Magrasó 1 , Istaq Ahmed 2 , Reidar Haugsrud 1<br />

1 Dept. Chemistry, University <strong>of</strong> Oslo, SMN/FERMiO, Gaustadalleen 21, NO-0349 Oslo, Norway<br />

2 Dept. Chemical and Biological Engineering, Chalmers University <strong>of</strong> <strong>Technology</strong>, SE-412 96, Göteborg, Sweden<br />

Lanthanum tungstate, “La6WO12”, has been reported to<br />

exhibit appreciable proton conductivity at intermediate<br />

temperatures and mixed ionic-electronic conductivities at<br />

higher temperatures [1]. The material cannot be made phasepure<br />

with a nominal atomic ratio between La and W equal to 6,<br />

but a solid-solution <strong>of</strong> single-phase compositions with La/W<br />

ratios between 5.3 and 5.7 has been reported [2]. The lanthanum<br />

tungstate crystallizes in a fluorite-type structure, and it has<br />

recently been shown that the formula is more correctly<br />

represented as La7-xW1+yO16-δ (x=0.37, y=0.17, δ =2.57 for a<br />

La/W nominal ratio <strong>of</strong> 5.6: LWO56) [2]. According to this<br />

structure refinement, the oxygen sublattice is inherently<br />

deficient and these positions are suggested to interact with water<br />

vapor dissolving charge compensating protonic and oxygen<br />

defects, yielding the materials conductivity characteristics. A<br />

possible defect model will be presented. In addition to the<br />

influence <strong>of</strong> protons on its properties, a peculiar effect was<br />

observed when La was sought substituted by Ca [1]; the<br />

presence <strong>of</strong> Ca seemingly decreased the ionic contribution to the<br />

conductivity relative to the electronic one.<br />

In order to further clarify the structural and functional<br />

properties <strong>of</strong> this material, in-situ neutron diffraction<br />

experiments at elevated temperatures under both nominally<br />

dried and humidified Ar atmospheres have been performed at<br />

the ISIS facility, Rutherford Appleton Laboratory (UK) on<br />

single-phase undoped and 2% Ca-doped lanthanum<br />

Fig. 1 Structure <strong>of</strong> LWO56<br />

according to ref [2].<br />

lattice parameter (Å)<br />

11.28<br />

11.26<br />

11.24<br />

11.22<br />

11.2<br />

11.18<br />

11.16<br />

100 300 500 700 900<br />

T ( 0 C)<br />

DRY AR<br />

WET AR<br />

P 09<br />

Fig. 2. Variation <strong>of</strong> lattice parameter for<br />

undoped LWO56 between <strong>15</strong>0 and 800<br />

°C in flowing Ar, or wetted (D2O)- Ar.<br />

tungstate. The main focus <strong>of</strong> the investigation is to locate crystallographic positions <strong>of</strong> both cations and<br />

anions and correlate their implications on the proton conductivity, for instance, filling <strong>of</strong> oxygen<br />

positions, cation displacements, or changes in the coordination environment. The role <strong>of</strong> acceptor<br />

substitution to the behavior <strong>of</strong> this material will be discussed.<br />

1. R. Haugsrud, Solid State Ionics 178 (2007) 555-560.<br />

2. A. Magrasó, Carlos Frontera, David Marrero-López, Pedro Núñez, Dalton Transactions (2009) 10273 – 10283.<br />

This work was supported by the “N-INNER” project (187160/S30) <strong>of</strong> the Research Council <strong>of</strong> Norway.<br />

‐ 51 ‐


Periodic long range proton conduction pathways in<br />

pseudo-cubic and orthorhombic perovskites<br />

Maria A. Gomez a , Dylan Shepardson b , Luong T. Nguyen a,b , Tolu Kehinde b<br />

a Department <strong>of</strong> Chemistry, Mount Holyoke College, 50 College Street, South Hadley, MA<br />

01075, USA<br />

b Department <strong>of</strong> Mathematics and Statistics, Mount Holyoke College, 50 College Street,<br />

South Hadley, MA 01075, USA<br />

Earlier, we investigated the effect <strong>of</strong> doping on N step<br />

periodic proton conductions pathways in pseudo-cubic<br />

BaZr0.875Y0.125O3 [1] and orthorhombic perovskites<br />

SrZr0.875Y0.125O3 and SrZr0.875Al0.125O3 [2]. In the 900-1300K<br />

temperature range, overall average limiting barriers over the<br />

N step periodic pathways considered are 0.3, 0.4, and 0.6 eV<br />

for Y/BaZrO3, Y/SrZrO3, and Al/SrZrO3, respectively, in<br />

good relative agreement with experiment. There are many<br />

contributing pathways for each system. Further, the<br />

simulation box size limits the N step periodic pathways<br />

found. In this paper, we expand the simulation cell<br />

periodically and use vertex and color coding to find probable<br />

pathways through the expanded system. Comparison <strong>of</strong> these<br />

paths with the smaller simulation box N step periodic paths<br />

provides us with insight on how to interpret pathways found<br />

with different size simulation boxes.<br />

1. M. A. Gomez, M. Chunduru, L. Chigweshe, L. Foster, S. J. Fensin,<br />

K. M. Fletcher, L. E. Fernandez, The effect <strong>of</strong> yttrium dopant on the<br />

proton conduction pathways <strong>of</strong> BaZrO3, a cubic perovskite, Journal<br />

<strong>of</strong> Chemical Physics 132 (2010) 214709.<br />

2. M. A. Gomez, M. Chunduru, L. Chigweshe, K. M. Fletcher, The<br />

effect <strong>of</strong> dopant at the Zr site on the proton conduction pathways <strong>of</strong><br />

SrZrO3, an orthorhombic perovskite, Journal <strong>of</strong> Chemical Physics, Submitted.<br />

‐ 52 ‐<br />

Zr<br />

Y<br />

O<br />

P 10<br />

Figure 2: (a) shows a 17 step pathway in<br />

Y/BaZrO3and (b) shows a 16 step<br />

pathway in the Al/SrZrO3. The former<br />

stays on doped faces while the latter<br />

stays in undoped regions.<br />

H


Proton Conductivity and Stability <strong>of</strong> Ba2In2O5 in Hydrogen Containing<br />

Atmospheres<br />

Jasna Jankovic a, b , David P. Wilkinson a, b , Rob S. Hui b<br />

a Department <strong>of</strong> Chemical and Biological Engineering, University <strong>of</strong> British Columbia – 2360 East Mall, Vancouver,<br />

BC, Canada, V6T 1Z3<br />

b National Research Council – <strong>Institute</strong> for Fuel Cell Innovation – 4250 Wesbrook, Vancouver, BC, Canada, V6T 1W5<br />

P 11<br />

Ba2In2O5 is one <strong>of</strong> the oxygen deficient brownmillerite structured ceramic materials widely<br />

investigated for both oxygen ion conduction and proton conduction in oxidizing atmospheres [1-4].<br />

However, its electrochemical properties have not been studied in hydrogen containing atmospheres,<br />

especially in the intermediate temperature range between 100 o C and 500 o C, a most desirable range<br />

for many practical applications such as fuel cells. In this work the total electrical conductivity <strong>of</strong><br />

Ba2In2O5 in hydrogen containing atmospheres (with and without humidification) and contribution <strong>of</strong><br />

proton conductivity to the total conductivity were<br />

investigated by AC impedance spectroscopy and<br />

EMF method in the temperature range between<br />

100 o C and 500 o C. A surprisingly high and stable<br />

electrical conductivity <strong>of</strong> over 0.3 S/cm was<br />

achieved in the temperature range <strong>of</strong> 300 o C to<br />

480 o C in a 50%vol H2 - 50%vol N2 atmosphere,<br />

as shown in Fig.1. EMF measurements revealed a<br />

proton transport number <strong>of</strong> one at 100 o C and<br />

200 o C, while 0.84 and 0.74 at 300 o C and 350 o C,<br />

respectively. Ba2In2O5 shows chemical stability in<br />

hydrogen containing atmospheres at temperatures<br />

below 500 o C, while decomposing at above 500 o C.<br />

This previously unreported conductivity in a H2<br />

atmosphere creates an opportunity for use <strong>of</strong><br />

Ba2In2O5 as a proton conductive material for a<br />

range<br />

<strong>of</strong> intermediate temperature electrochemical<br />

devices.<br />

log� (S/cm)<br />

1. G. B. Zhang, D. M. Smyth, Solid State Ionics 82 (1995) <strong>15</strong>3-160<br />

2. T. Yajima, H. Iwahara, H. Uchida, Solid State Ionics 47 (1991) 117<br />

3. G.B. Zhang, D.M. Smyth, Solid State Ionics 82 (1995) 161 -172<br />

4. N. Bonanos, Solid State Ionics 53-56 (1992) 967-974<br />

‐ 53 ‐<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

-6<br />

-7<br />

-8<br />

500 400 300 200 100 o C<br />

Nafion<br />

-9<br />

1.20 1.60 2.00<br />

1000/T(K<br />

2.40 2.80<br />

-1 )<br />

Fig. 1 Comparison <strong>of</strong> Ba2In2O5 conductivity in different<br />

atmospheres: ● in dry air; △ in dry N2; ◊ in<br />

50%H2/50%N2; ♦ in 48%H2/49%N2/3% steam; □ sample<br />

decomposed.


Proton Solvation and Transport in Hydrated Nafion<br />

Shulu Feng and Gregory A. Voth<br />

Department <strong>of</strong> Chemistry, James Franck <strong>Institute</strong>, and Computation <strong>Institute</strong>, University <strong>of</strong> Chicago, Chicago,<br />

IL, USA<br />

Proton solvation properties and transport mechanism were studied in hydrated nafion using the selfconsistent<br />

multistate empirical valence bond (SCI-MS-EVB) method. It was found that sulfonate<br />

groups affect proton solvation as well as proton distribution by stabilizing the Zundel-like structure<br />

in their first solvation shells. Diffusion rates, Arrhenius activation energies, and transport pathways<br />

were calculated and analyzed to characterize the nature <strong>of</strong> proton transport. Similar to experimental<br />

observations, activation energy drops quickly with increasing water content when water loading<br />

level is smaller than ~ 10 H2O/SO3 - . The transport pathway is strongly correlated with sulfonate<br />

groups, which serve as a kind <strong>of</strong> transit station for the protons.<br />

‐ 54 ‐<br />

P 12


Conductivity study <strong>of</strong> dense BaZr0.9Y0.1O3-δ obtained by<br />

spark plasma sintering<br />

Sandrine Ricote * , Nikolaos Bonanos * , Hsiang-Jen Wang * and Bernard A. Boukamp **<br />

* Fuel Cells and Solid State Chemistry Division, Risø National Laboratory for Sustainable Energy, Technical University<br />

<strong>of</strong> Denmark, P.O. 49, 4000 Roskilde. Denmark.<br />

** University <strong>of</strong> Twente, Faculty <strong>of</strong> Science and <strong>Technology</strong>, MESA + <strong>Institute</strong> for Nanotechnology, P.O. Box 217, 7500<br />

AE Enschede, The Netherlands.<br />

Many studies on proton conductors at intermediate<br />

temperature are conducted on yttrium doped barium zirconate<br />

because <strong>of</strong> its good stability and satisfactory proton conductivity<br />

at temperatures around 600�C. Two issues, however,<br />

are still under investigation: the fabrication <strong>of</strong> very dense<br />

samples and the understanding <strong>of</strong> the high grain boundary<br />

resistivity.<br />

To obtain 95% dense samples requires sintering at<br />

temperature about 1700�C for several hours, resulting in<br />

evaporation <strong>of</strong> barium. The use <strong>of</strong> sintering aid has also been<br />

tried, but leads to contamination <strong>of</strong> the grain boundaries. This<br />

study presents the spark plasma sintering as a powerful<br />

method to get 99.8% dense and transparent BaZr0.9Y0.1O3-δ<br />

(BZY10) samples [1].<br />

The purity <strong>of</strong> the sample has been investigated by XRD and<br />

microscopy. The cubic lattice parameter <strong>of</strong> the single phase<br />

compound has been evaluated to a slightly higher value than<br />

the one commonly determined (4.2108Å), and the grain size<br />

has been found to be between 100 and 800 nm. The HRTEM<br />

pictures reveal clean grain boundaries (Figure 1).<br />

Conductivity measurements have been performed in wet<br />

reducing atmosphere to determine the bulk and grain<br />

boundary activation energy <strong>of</strong> proton diffusion (Figure 2).<br />

The higher value for the grain boundaries seems to be only<br />

explained by a space charge layer, as neither evaporation <strong>of</strong><br />

barium nor impurity or second phase have been detected in<br />

the grain boundaries. The activation energy found in this<br />

work falls nicely on a plot <strong>of</strong> the bulk activation energy<br />

versus the lattice parameter published by Azad et al. [2] on<br />

BZY10 sintered by different processes, showing that a larger<br />

lattice<br />

parameter gives a lower activation energy for proton<br />

transport.<br />

‐ 55 ‐<br />

� (S.cm -2 )<br />

Fig. 1: HRTEM picture <strong>of</strong> BZY10.<br />

1e-2<br />

1e-3<br />

1e-4<br />

1e-5<br />

1e-6<br />

1e-7<br />

Total<br />

Bulk<br />

GB<br />

Specific GB<br />

o<br />

T ( C)<br />

P 13<br />

1nm<br />

800 600 500 400 300 200<br />

1e-8<br />

0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2<br />

1000/T (K -1 )<br />

Fig. 2: Temperature dependence <strong>of</strong> the<br />

conductivity, N2/H2 (9%), 0.0<strong>15</strong> atm H2O.<br />

1. S. Ricote, G. Caboche, C. Estournes, N. Bonanos, Journal <strong>of</strong> Nanomaterials (2008), Article ID 354258.<br />

2. A.K. Azad, C. Savaniu, S.W. Tao, S. Duval, P. Holtappels, R.M. Ibberson and J.T.S. Irvine, Journal <strong>of</strong> Materials<br />

Chemistry 18 (2008), 3414.


Conductivity study <strong>of</strong> A- and B-site (co-)doped LaNbO4<br />

Mariya Ivanova 1 , Sandrine Ricote 2 , Wilhelm A. Meulenberg 1 , R. Haugsrud 3 , T. Norby 3<br />

P 14<br />

1 Forschungszentrum Jülich GmbH, <strong>Institute</strong> <strong>of</strong> Energy Research 1, D-52425 Jülich, Germany<br />

2 Fuel Cells and Solid State Chemistry Division, Risø National Laboratory for Sustainable Energy, Technical University<br />

<strong>of</strong> Denmark, P.O. 49, 4000 Roskilde, Denmark<br />

3 University <strong>of</strong> Oslo, FERMiO, Gaustadalleen 21, NO-0349 Oslo, Norway<br />

Acceptor-doped lanthanum niobate (LaNbO4) represents a<br />

class <strong>of</strong> ceramic materials that exhibits proton conduction at<br />

elevated temperatures accompanied by a high chemical and<br />

thermal stability under a variety <strong>of</strong> conditions. Due to their<br />

good mechanical properties, high red-ox stability and<br />

compatibility with NiO and Ni, these niobates are <strong>of</strong> interest<br />

for application as electrolyte in proton conducting solid<br />

oxide fuel cells (PC-SOFC).<br />

In order to improve the conductivity <strong>of</strong> these materials,<br />

various doping and co-doping schemes can be implemented.<br />

As shown in the literature, A-site doping for example with<br />

1% Ca improves the proton conductivity <strong>of</strong> LaNbO4 up to a<br />

value <strong>of</strong> approximately 1·10 -3 S·cm -1 at 1000 °C [1].<br />

Different doping schemes <strong>of</strong> LaNbO4 were attempted in the<br />

present study in order to: i) test the influence <strong>of</strong> doping<br />

elements on both grain interior and grain boundary<br />

conductivities; and ii) correlate the conductivity <strong>of</strong> doped<br />

materials with both the chemical and structural features <strong>of</strong><br />

the host-guest matrix.<br />

Materials developed in this study have a chemical<br />

composition corresponding to the general formula<br />

ln � tot [S.cm -1 ]<br />

-8<br />

-10<br />

-12<br />

-14<br />

-16<br />

-18<br />

T [�C]<br />

1000 800 600 400<br />

LN-Ga<br />

LN-Ge<br />

LN-In<br />

LCN-Ga<br />

LCN-Ge<br />

LCN-In<br />

LBN-Ga<br />

LBN-Ge<br />

LBN-In<br />

0.8 1.0 1.2 1.4 1.6 1.8<br />

1000/T [K -1 ]<br />

La1-xAxNb1-yByO4-δ where A stands for Ca or Ba and B for Ga, Ge or In (x=0 or 0.01; y=0.01). The<br />

investigated compounds were synthesized via the solid state route. Samples with relative densities<br />

higher than 95% were obtained after sintering at <strong>15</strong>00°C/10hrs. The impedance spectroscopic study<br />

was carried out in a wet (PH2O=0.0<strong>15</strong> atm) mixture <strong>of</strong> 9%H2 and N2 in the temperature range from<br />

300°C to 900°C. Grain boundary and grain interior conductivities were deconvoluted from the<br />

impedance spectra yielding several tendencies in analyzing the conductivity-doping elementcorrelation.<br />

Amongst the developed doped materials, La0.99Ca0.01Nb0.99Ga0.01O4-δ showed the highest<br />

total conductivity <strong>of</strong> 4·10 -4 S·cm -1 at 900 °C (Figure 1) which is however lower than the reference<br />

value <strong>of</strong> approximately ~9·10 -4 S·cm -1 measured for La0.99Ca0.01NbO4-δ at 900 °C [1].<br />

1. R. Haugsrud and T. Norby, Solid State Ionics 177 (2006), 1129.<br />

‐ 56 ‐<br />

1E-4<br />

1E-5<br />

1E-6<br />

1E-7<br />

1E-8<br />

Fig. 1: Total conductivity <strong>of</strong> (co-)doped<br />

LaNbO4 as a function <strong>of</strong> temperature.<br />

� tot [S.cm -1 ]


Preparation <strong>of</strong> Nafion 117�-SnO2 Composite Membranes using an<br />

Ion-Exchange Method<br />

C. F. Nørgaard a , U. G. Nielsen b , E. M. Skou a<br />

a<br />

<strong>Institute</strong> <strong>of</strong> Chemical Engineering, Biotechnology and Environmental <strong>Technology</strong>, Faculty <strong>of</strong> Engineering, University<br />

<strong>of</strong> Southern Denmark, 5230 Odense M, Denmark<br />

b<br />

Department <strong>of</strong> Physics and Chemistry, Faculty <strong>of</strong> Science, University <strong>of</strong> Southern Denmark, 5230 Odense M, Denmark<br />

Nafion 117�-SnO2 composite membranes were successfully prepared using an ion-exchange<br />

method. SnO2 was incorporated into Nafion 117� membranes by ion-exchange in solutions <strong>of</strong><br />

SnCl2 . 2 H2O in methanol, followed by oxidation to SnO2 in air. The content <strong>of</strong> SnO2 proved<br />

controllable by adjusting the concentration <strong>of</strong> the ion-exchange solution.<br />

The prepared nanocomposite membranes were characterized by XRD and 119 Sn MAS NMR while<br />

the<br />

in-plane proton conductivity was found to decrease with SnO2 content when evaluated with EIS.<br />

However,<br />

the conductivity was comparable to Nafion� at SnO2 contents below 8 wt%.<br />

‐ 57 ‐<br />

P <strong>15</strong>


A study <strong>of</strong> Pt electrodes on proton conducting Ca-doped LaNbO4<br />

Ragnar Strandbakke, Truls Norby, Reidar Haugsrud<br />

Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, FERMiO/SMN, NO-0349 Oslo, Norway<br />

The polarization resistance <strong>of</strong> Pt point electrodes on a La0.995Ca0.005NbO4 (LCNO) proton conductor<br />

has been investigated by electrical impedance spectroscopy in the temperature range 600 – 700 ºC in<br />

wet O2, as a function <strong>of</strong> pH2O in 2.5 % O2 + 97.5 % Ar, and as a function <strong>of</strong> pO2 in wet O2 + Ar<br />

mixtures. surface processes are being p<br />

hydroxide ions from ambient water and<br />

adsorbed surface oxygen is suggested as<br />

governing the overall electrode reaction<br />

under oxidizing conditions. H2O / D2O<br />

isotope exchange measurements are in<br />

agreement with the proposed wateroxygen<br />

reaction, suggesting proton<br />

transfer in both surface and interface<br />

reactions.<br />

(1/RS) shows a dependency <strong>of</strong> roughly<br />

pH2O 1/3 roposed as rate limiting, and the formation <strong>of</strong> surface<br />

and pO2<br />

Fig. 1: 1/ RS (three‐phase boundary length specific)<br />

as a function <strong>of</strong> pH2O and pO2 at 700 °C.<br />

1/5 at 700 ºC, see Fig. 1.<br />

The activation energy was 1.09 (± 0.04)<br />

eV in wet O2, and total electrode<br />

resistance at 700 ºC was approximately<br />

3·10<br />

f<br />

4 Ωcm referring to unit triple phase<br />

boundary length.<br />

The surface related part <strong>of</strong> the<br />

polarization resistance RS could be<br />

fitted to a Langmuir-type adsorption<br />

model, representing surface exchange o oxygen and water, followed by hydroxide formation<br />

reaction.<br />

‐ 58 ‐<br />

P 16


Proton Conductivity in BCY20-Pd Ionic Hybrid Material<br />

Archana Subramaniyan, Jianhua Tong, Ryan O’Hayre and Nigel Sammes<br />

Department <strong>of</strong> Metallurgical and Materials Engineering, Colorado School <strong>of</strong> Mines, <strong>15</strong>00 Illinois St., Golden, CO<br />

80401, USA<br />

Acceptor doped alkaline earth cerate perovskites<br />

generate hydroxyl-type defects at elevated<br />

temperatures in atmospheres containing water vapor<br />

or hydrogen. The hydroxyl-type defects lead to<br />

protonic conduction in these perovskites. Hence,<br />

they are potential candidates for applications such<br />

as fuel cells, hydrogen sensors and electrochemical<br />

reactors. This work aims to fabricate and<br />

characterize a “super ionic hybrid composite<br />

material” combining proton conducting oxides and<br />

a hydrogen permeable metal. Incorporation <strong>of</strong> the<br />

hydrogen permeable metal nano-particles in the<br />

proton conducting ceramic matrix at volume<br />

fractions below the electronic percolation threshold<br />

manipulates the space charge layer at the<br />

metal/ceramic interfaces, and is thereby<br />

hypothesized to increase the proton conduction<br />

significantly. The super ionic hybrid composite<br />

pellets are fabricated by mixing 20 mol% yttria<br />

doped barium cerate (BCY20) with palladium<br />

powder. As a first step in this two-step process,<br />

BCY20 pellets were fabricated via powders<br />

synthesized from a polymeric sol-gel route and<br />

characterized by XRD, SEM and Raman<br />

spectroscopy. A new phase, identified to be yttria<br />

P 17<br />

Fig. 1 Conductivities <strong>of</strong> BCY20 pellets sintered at<br />

1450˚C for 24 h in air with (open symbol) and<br />

without BCY20 powder bath (closed symbol).<br />

doped ceria (Y0.2Ce0.8O1.9) was observed on the surface <strong>of</strong> the BCY20 pellets when sintered above<br />

1250˚C for 24 hours. On sintering the pellets in a BCY20 powder bath, this phase was eliminated.<br />

From the electrical characterization <strong>of</strong> BCY20 pellets (figure 1) sintered with and without powder,<br />

the effect <strong>of</strong> the second phase on ionic conduction was observed. Also, the need for the<br />

elimination <strong>of</strong> the same is proposed. The synthesis and sintering <strong>of</strong> BCY20 was thereby optimized<br />

to obtain a phase pure material with the desired density. As a final step, super ionic hybrid pellets<br />

were fabricated by mixing synthesized phase pure BCY20 powders with palladium nano particles.<br />

The structural, microstructural, physical and electrical characterization <strong>of</strong> these BCY20/Pd<br />

composite ceramics will be discussed in detail.<br />

‐ 59 ‐


Defect chemistry and transport properties <strong>of</strong> oxide protonic perovskite<br />

materials with transition metals on B-site: BaZr1-xPrxO3-δ<br />

K. E. J. Eurenius, T. Kikuchi, M. Tamaru, S. Miyoshi, and S. Yamaguchi<br />

Department <strong>of</strong> Materials Engineering, The University <strong>of</strong> Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, JAPAN<br />

The development <strong>of</strong> highly reactive cathode materials for proton conducting fuel cells<br />

(PCFCs) is a necessary component that has so far been missed to date i . Mixed electron and hole<br />

conductors could greatly reduce cathode polarization because <strong>of</strong> extended 2-dimensional reaction<br />

surface in comparison with a limited reaction area <strong>of</strong> triple phase boundary for composite cathodes.<br />

One direction for the mixed conductivity is to employ the solid solution phase between AB 3+ xB 4+ 1xO3-(h<br />

● )x and AB 3+ xB 4+ 1-xO3H + x ii,iii . However, this idea has not been successful, since the materials<br />

examined earlier have shown poor or no proton conductivity. An additional route is to investigate<br />

site percolation systems doped with a multivalent transition metal on the B-site, such as AM1-xyB<br />

3+ xB 4+ yO3H. This composition has not yet<br />

been successful due to poor hole and proton<br />

conductivity iv .<br />

The aim <strong>of</strong> this study is to<br />

illustrate a complete picture based on<br />

defect chemistry by Brouwer diagram<br />

for such a system and hence to analyze<br />

the characteristics <strong>of</strong> a site-percolation<br />

properties. Further, the possibility <strong>of</strong><br />

strong correlation between localized<br />

electrons and both holes and protonic<br />

defects in combination with possible<br />

two-fold percolation <strong>of</strong> electrons and<br />

protons via the transition metal and<br />

oxygen network is discussed.<br />

Log<br />

[Defect]<br />

P 18<br />

Log (pO2/atm) Log (pH2O/atm)<br />

Fig.1. Illustration <strong>of</strong> defects under dry (pO2) and wet (pH2O) conditions.<br />

As an example for the examination <strong>of</strong> the present idea, the experimental results <strong>of</strong> BaZr1-xPrxO3-δ (x<br />

= 0.1 - 0.9) by electrochemical measurements for transference number, conductivity and Seebeck<br />

coefficient as well as the electron spectroscopy measurements and DFT calculations have been<br />

analyzed based on three-dimensional Brouwer diagram: A complex conduction behavior, where the<br />

material with hole conductivity in O2 atmosphere due to the nature <strong>of</strong> PrO6 octahedra, turns into a<br />

mixed conductor <strong>of</strong> protons, oxide ions and holes under reducing atmospheres, is analyzed using the<br />

3D-Brouwer diagram estimated. The moderate range <strong>of</strong> mixed proton and hole conductivity, which<br />

shows a cubic dependence with dopant concentration, is due to the strong correlation between<br />

localized electrons on Pr ions and both positively charged electron holes and protons on oxide ions<br />

sublattice B.<br />

1<br />

S. M. Haile Actra Materialia 51 (2003) 5981<br />

1<br />

H. Matsumoto, T. Shimura, T. Higuchi, T. Otake, Y. Sasaki, K. Yashiro, A. Kaimai, T. Kawada, J. Mizusaki,<br />

Electrochemistry, 72 (2004) 861<br />

1<br />

H. Matsumoto, T. Shimura, T. Higuchi, H. Tanaka, K. Katahira, T. Otake, T. Kudo, K. Yashiro, A. Kaimai, T. Kawada,<br />

J. Mizusaki, Journal <strong>of</strong> The Electrochemical Society, <strong>15</strong>2 (2005) A488<br />

1<br />

S. Mimuro, S. Shibako, Y. Oyama, K. Kobayashi, T. Higuchi, S. Shin, S. Yamaguchi, Solid State Ionics, 178 (2007)<br />

647<br />

‐ 60 ‐


Microstructural analysis <strong>of</strong> Y doped BaZrO3<br />

Yukiko Oyama a , Takao Tsurui b , Miyoshi Shogo c , and Shu Yamaguchi c<br />

a: Department <strong>of</strong> Applied Chemistry and Chemical Engineering, Toyama National College <strong>of</strong> <strong>Technology</strong><br />

13 Hongo, Toyama-shi, Toyama, 939-8630, Japan<br />

b: Tohoku University <strong>Institute</strong> <strong>of</strong> Materials Research, Tohoku University<br />

2-1-1Katahira, Aoba-ku, Sendai, 980-8577, Japan<br />

c: Department <strong>of</strong> Materials Engineering, Graduate School <strong>of</strong> Engineering, The University <strong>of</strong> Tokyo, 7-3-1 Hongo,<br />

Bunkyo-ku, Tokyo, 113-8656, Japan<br />

P 19<br />

Yttrium doped BaZrO3 (BZ) is one <strong>of</strong> the candidate materials for the proton conducting electrolyte<br />

with a superior chemical stability and the highest conductivity among perovskite oxide systems.<br />

Our previous study for the samples synthesized by conventional solid reaction method sintered at<br />

1600 ºC shows the phase separation <strong>of</strong> two BaZrO3 phased; a simple cubic BaZrO3 and perovskite<br />

phase (BZ(II)) with an ordered super cell structure with structural units <strong>of</strong> BaZrO3 and BaYO2.5.<br />

The samples with the composition <strong>of</strong> BaO:ZrO2:YO1.5(=1/2Y2O3)=3:2:1 and 9:4:8 have been<br />

prepared by three different synthetic processes <strong>of</strong> conventional solid state reaction (SSR) method,<br />

Pechini method, and Sol-Gel method using metal alkoxides as source materials, annealed for<br />

various period <strong>of</strong> time. Microstructure observation <strong>of</strong> Ba(Zr1-xYx)O3-d by a transmission electron,<br />

electron beam diffraction (JEM-4000EX, JEOL), and energy dispersive X-ray spectroscopy (JEM-<br />

3000F, JEOL) are carried out to examine the phase formation in the polycrystalline samples<br />

prepared by those preparatory methods.<br />

TEM micrographs <strong>of</strong> SSR samples annealing at 1600 ˚C for 10h are shown in Fig 1 (a) and (c). The<br />

primitive lattice <strong>of</strong> simple cubic BaZrO3 (Fig. 1 (a)), and nanocrystal grains (c) are observed. A<br />

nanocrystal grains (Fig. 1 (c)) and the superlattice phase (Fig. 1 (d)) are both observed in addition to<br />

the primitive lattice phase from the most <strong>of</strong> the samples except for sample. Amorphous phase (Fig.<br />

1 (b)) is observed with supper lattice phase from SSR 100h and 200h samples. The primitive lattice<br />

is only observed from the composition around the solubility limit <strong>of</strong> about 8 mol% Y in BaZrO3<br />

phase. The compositions <strong>of</strong> superlattice phase are located between BaO:ZrO2:YO1.5=3:2:1 and<br />

9:4:8 as shown in Fig. 2. The nano-grain crystals with almost the same compositions to the initial<br />

one are observed from samples by Pechini and Sol-Gel method. The existence <strong>of</strong> stable nano-grain<br />

crystals even after 200h annealing suggests the contribution <strong>of</strong> liquid phase for the phase formation.<br />

The phase formation process <strong>of</strong> Y doped BaZrO3 phase based on the phase diagram for BaO-ZrO2-<br />

YO1.5 system is also discussed by taking into account <strong>of</strong> the presence <strong>of</strong> liquid phase at the sintering<br />

temperature.<br />

(a)<br />

(c)<br />

2nm<br />

2nm<br />

(b)<br />

(d)<br />

2nm<br />

2nm<br />

Fig.1 TEM micrographs with electron beam diffraction patterns <strong>of</strong><br />

Ba 3 Zr 2 YO 8.5 synthesized by solid state reaction method annealed<br />

at 1600°C for (a) and (c) 10h, and (b) and (d) 100h. The electron<br />

diffraction image indicates (a) basic cubic pattern, (b) amorphous,<br />

(c) nano crystal, and (d) super lattice, respectively.<br />

- 61 -<br />

ZrO2<br />

BaO<br />

3:2:1<br />

■ 3:2:1 by SSR<br />

■ 3:2:1 by Pechini<br />

9:4:8 by Sol-Gel<br />

9:4:8<br />

YO1.5<br />

Fig.2 Composition mapping <strong>of</strong> super lattice phase<br />

about 3;2;1 and 9:4:8 samples annealed at 1600°C<br />

on the BaO-ZrO2-YO1.5 ternary phase diagram.


Site Selectivity <strong>of</strong> Dopants in BaZr1-yMyO3-δ (M = Dy, Eu, Sm)<br />

and Measurement <strong>of</strong> Their Water Contents and Conductivities<br />

Donglin HAN a , Yoshitaro NOSE a , Kozo SHINODA b , Tetsuya UDA a<br />

a Department <strong>of</strong> Materials Science and Engineering, Kyoto University,<br />

Yoshida Honmachi, Sakyo-ku, Kyoto 606-8501, Japan<br />

b <strong>Institute</strong> <strong>of</strong> Multidisciplinary Research for Advanced Materials, Tohoku University,<br />

Katahira 2-1-1, Aoba-ku, Sendai 980-8577, Japan<br />

BaZrO3 doped with trivalent rare-earth cations has<br />

relatively high protonic conductivity in humidified<br />

atmosphere. However, it is reported that the rareearth<br />

cations, which are named as amphoteric<br />

dopants, can occupy both A and B sites in crystal<br />

lattice <strong>of</strong> perovskite oxides, consuming and<br />

generating the oxide ion vacancies, respectively.<br />

Since the protonic conductivity in doped BaZrO3<br />

depends on the concentration <strong>of</strong> protons formed<br />

by dissolving hydroxide ions into the oxide ion<br />

vacancies, the concentration <strong>of</strong> oxide ion<br />

P 20<br />

Fig. 1 Lattice volume difference, defined as lattice<br />

volume <strong>of</strong> the Ba-rich sample minus that <strong>of</strong> the Bapoor<br />

sample, with reported data <strong>of</strong> doped BaTiO3 [1].<br />

vacancies is one <strong>of</strong> the vital factors in the protonic conducting. Therefore, the site selectivity <strong>of</strong><br />

trivalent dopants in doped BaZrO3 is important on its protonic conductivity.<br />

First, the lattice volumes <strong>of</strong> Ba1±0.01Zr0.99M0.01O3-δ (M = Sc, Y, Dy, Eu, Sm) were investigated. As<br />

shown in Fig. 1, when doped with Eu or Sm, the difference <strong>of</strong> the lattice volume between Ba-rich<br />

and Ba-poor samples was larger than that <strong>of</strong> nondoped<br />

ones. However, when doped with Sc, Y or<br />

Dy, the lattice volume difference was close to or<br />

smaller than that <strong>of</strong> the non-doped ones. Therefore,<br />

it is considered that comparing with Sc, Y and Dy, it<br />

is relatively easy for Eu and Sm to occupy the A-site<br />

when doping into BaZrO3. Then, the water contents,<br />

namely concentrations <strong>of</strong> hydroxide ions, <strong>of</strong> the<br />

BaZr0.8M0.2O2.9 samples hydrated in the humidified<br />

O2 or Ar atmosphere were measured, as shown in<br />

Fig. 2. The results revealed that when hydrated in Fig. 2 Proton concentration <strong>of</strong> BaZr0.8M0.2O2.9 (M =<br />

Sc, Y, Dy, Eu, Sm) hydrated in humidified O2 or Ar.<br />

humidified O2 atmosphere, the concentrations <strong>of</strong><br />

hydroxide ions <strong>of</strong> the samples doped with Dy, Eu or Sm were lower than those <strong>of</strong> the samples doped<br />

with Sc or Y. And when hydrated in humidified Ar atmosphere, comparing with the ones hydrated<br />

in humidified O2 atmosphere, no significant difference in concentration <strong>of</strong> hydroxide ions was<br />

observed, expect the sample <strong>of</strong> BaZr0.8Dy0.2O2.9, which exhibited a great elevation in concentration<br />

<strong>of</strong> hydroxide ions when hydrated in the humidified Ar atmosphere. At last, the measurement <strong>of</strong><br />

conductivities was performed. It was observed that for the BaZr0.8M0.2O2.9 (M = Sc, Y) samples,<br />

there was little difference in conductivities measured in the atmosphere <strong>of</strong> humidified O2 or Ar.<br />

However, by altering the atmosphere from humidified O2 to Ar, a dramatic elevation in the<br />

conductivity <strong>of</strong> BaZr0.8Dy0.2O2.9 was observed. It is considered to be attributed reasonably to the<br />

higher concentration <strong>of</strong> hydroxide ions in humidified Ar atmosphere than that in humidified O2<br />

atmosphere, which was confirmed as shown in Fig. 2.<br />

1. Y. Tsur, A. Hitomi, I. Scrymgeour and C.A. Randall, Japan Journal <strong>of</strong> Applied Physics 40 (2001), 255.<br />

‐ 62 ‐


Proton conductors based on the double perovskite Ba2YNbO6<br />

Xiaoxiang Xu, Elena Konysheva, Shanwen Tao, John T.S. Irvine<br />

School <strong>of</strong> Chemistry, University <strong>of</strong> St Andrews, St Andrews, Fife KY16 9ST, UK<br />

The potential application <strong>of</strong> a proton conducting material is mainly dependent upon its magnitude <strong>of</strong><br />

conductivity and associated working temperatures. The most desirable working temperatures for<br />

proton conductors, as has been identified by Norby, lie from 200 o C to 700 o C, where electrode<br />

kinetics in fuel cells are fast and insensitive to poisoning [1]. But no satisfactory proton conductors<br />

have been found in this range yet. Y doped BaCeO3 is a potential material as it shows promising<br />

conductivity at a temperature as low as 600 o C. Its poor chemical stability toward CO2 and H2O<br />

however, renders a large restriction on its application as the electrolyte in fuel cells as CO2 and H2O<br />

are normal products during energy conversions. It has been reported that certain types <strong>of</strong> mixed<br />

perovskites can be protonic conductors. They demonstrate comparable proton conductivity as simple<br />

perovskite with less electronic contributions and much higher stability toward CO2 and H2O. A<br />

typical example is Ba3Ca1.18Nb1.82O9-δ whose activation energy ~0.54 eV is comparable with the<br />

simple perovskite BaCe0.9Y0.1O3-δ [2]. Their chemical formula can be written in general form<br />

A3B’B”2O9 and A2B’B”O6, where A and B” are divalent ions and pentavalent ions, while B’ is<br />

divalent for the formal but trivalent for the latter. It was found that B site disordering as well as Ba 2+<br />

at A sites are favorable for higher proton conductivity [3]. Here we develop a double perovskite<br />

structure material Ba2YNbO6 whose structure and electrical properties were investigated after<br />

doping or introduction <strong>of</strong> cation non-stoichiometry. Large polarizable cations such as Ba 2+ and Y 3+<br />

tend to form a high symmetry structure. The highest proton conductivity achieved is 1.38×10 -3 S/cm<br />

at 540°C in wet 5% H2/Ar (humidified in water at room temperature) for sample Ba2Y1.08Nb0.88O6-δ,<br />

which is comparable with primitive perovskite proton conductors。<br />

1. T. Norby, Solid State Ionics 125 (1999), 1.<br />

2. K.C. Liang, Y. Du, and A.S. Nowick, Solid State Ionics 69 (1994), 117.<br />

3.<br />

A.S. Nowick, Y. Du, and K.C. Liang, Solid State Ionics 125 (1999), 303.<br />

‐ 63 ‐<br />

P 21


Atmosphere dependent temperature evolution phase transition <strong>of</strong><br />

BaCe0.9Y0.1O2.95 - A neutron powder diffraction study<br />

Abul K. Azad 1 , Angela Kruth 2 , John T.S. Irvine 1<br />

1<br />

School <strong>of</strong> Chemistry, University <strong>of</strong> St Andrews, Fife KY16 9ST, UK<br />

2<br />

Department <strong>of</strong> Chemistry, University <strong>of</strong> Aberdeen, Meston Walk, Aberdeen AB24 3UE, UK<br />

P 22<br />

Proton conducting solid electrolytes with high protonic conductivity with high transport number<br />

have a wide range <strong>of</strong> technological applications in fuel cells, batteries, sensors,<br />

hydration/dehydration <strong>of</strong> hydrocarbons, electrolysers. Yttrium doped BaCeO3 has been widely<br />

investigated as a proton conducting material [1]. Earlier neutron powder diffraction measurements<br />

probing the influence <strong>of</strong> water content on structure have started with wet or dry samples and heated<br />

these up in vacuum with gradual loss <strong>of</strong><br />

water in conditions that do not quite match<br />

those <strong>of</strong> the greatest technical relevance<br />

[2,3]. In this study high-resolution neutron<br />

powder diffraction data were collected at<br />

different temperatures under hydrogen or<br />

oxygen under high and low humidities. The<br />

samples were loaded in quartz tubes in dry<br />

or wet conditions, reducing pressure at<br />

ambient to allow measurements in situ at<br />

approximately 1 bar. Rietveld refinement <strong>of</strong><br />

neutron diffraction data showed that the<br />

structure changed from orthorhombic (space<br />

group Pbnm) to rhombohedral (R-3c) on<br />

heating from room temperature to 973 K.<br />

Figure 1 shows the Rietveld refinement<br />

Fig. 1. Rietveld refinement pr<strong>of</strong>ile <strong>of</strong> the neutron<br />

powder diffraction data <strong>of</strong> BaCe0.9Y0.1O3-� collected<br />

at room temperature in wet hydrogen atmosphere.<br />

pr<strong>of</strong>ile <strong>of</strong> BaCe0.9Y0.1O3-d collected at room temperature in wet hydrogen atmosphere. The unit-cell<br />

volume under oxygen was found to be higher than under hydrogen at the same temperatures.<br />

Oxygen site occupancies were directly determined as functions <strong>of</strong> temperature for dry and wet<br />

conditions in both oxidising and reducing conditions in order to probe actual oxygen/hydroxyl<br />

contents under the conditions relevant to most fuel cell or transport studies.<br />

[1] K.D. Kreuer, Solid State Ionics, 97 (1997), 1.<br />

[2] K. Knight, Solid State Ionics, 127 (2000), 43.<br />

[3] I. Ahmed, C.S. Knee, M. Karlsson, S.-G. Eriksson, P.F. Henry, A. Matic, D. Engberg and L. Börjesson, Journal <strong>of</strong><br />

Alloys and Compounds 450 (2008), 103.<br />

‐ 64 ‐


The effect <strong>of</strong> cation non-stoichiometry <strong>of</strong> proton conducting LaNbO4<br />

Guttorm E. Syvertsen 1 , Anna Magrasó 2 , Mari-Ann Einarsrud 1 , and Tor Grande 1<br />

1 Department <strong>of</strong> Material Science and Engineering, Norwegian University <strong>of</strong> Science and<br />

<strong>Technology</strong>, NO-7491 Trondheim, Norway<br />

2 Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, FERMiO, Gaustadalléen 21, NO-0349 Oslo, Norway<br />

The proton conductivity <strong>of</strong> LaNbO4 is enhanced by<br />

minor substitution <strong>of</strong> La with alkaline earth ions [1-2].<br />

The enhanced proton conductivity is rationalised by<br />

hydration <strong>of</strong> oxygen vacancies, but interfacial space<br />

charges may contribute as well. The solubility <strong>of</strong><br />

alkaline earth oxides in LaNbO4 is still under debate,<br />

and possible supersaturation and minor deviations from<br />

the nominal stoichiometric composition, may influence<br />

the electrical conductivity <strong>of</strong> the bulk materials. Here<br />

the influence <strong>of</strong> variations in cation non-stoichiometry<br />

<strong>of</strong> LaNbO4 is reported. LaNbO4 produced by spray<br />

pyrolysis [3] were systematically doped with small<br />

amounts <strong>of</strong> La 3+ , Nb 5+ and Ca 2+ oxide precursors. The<br />

sintering properties <strong>of</strong> the doped LaNbO4-powders were<br />

investigated by dilatometry, and 4-point dc-conductivity<br />

measurements were performed at elevated temperatures<br />

in controlled atmosphere. Minor variations in the cation<br />

stoichiometry were shown to have a pronounced effect<br />

on both the sintering properties as well as the electrical<br />

conductivity. The effects <strong>of</strong> the La-, Nb-, and Caadditions<br />

are discussed with regards to mass transport<br />

during sintering and on the electrical conductivity.<br />

Finally the solubility <strong>of</strong> La2O3, Nb2O5 and alkaline earth<br />

oxides in LaNbO4 is discussed with respect to possible<br />

formation <strong>of</strong> secondary phases due to the cation nonstoichiometry.<br />

1. R. Haugsrud and T. Norby, Nature Materials 5 (2006) (3),<br />

193.<br />

2. R. Haugsrud and T. Norby, Solid State Ionics 177 (2006) (13-<br />

14), 1129.<br />

3. T. Mokkelbost, I. Kaus, R. Haugsrud, T. Norby, T. Grande<br />

and M.-A. Einarsrud, J. Am. Ceram. Soc. 91 (2008) (3), 879.<br />

4. A.M. Frolov and A.A. Evdokimov, Russ. J. Inorg. Chem.<br />

(Engl. Transl.), 32 (1987), 1771.<br />

‐ 65 ‐<br />

P 23<br />

Fig. 1 10 kHz total conductivities <strong>of</strong> selected samples<br />

<strong>of</strong> LaNbO4 with excess CaO.<br />

Fig. 2 Part <strong>of</strong> the CaO-La2O3-Nb2O5 phase diagram<br />

showing the investigated compositions. Phase diagram<br />

was redrawn from Frolov and Evdokimov [4].


Solid State NMR Studies <strong>of</strong> CsH2PO4, a Protonic Conductor for<br />

Intermediate-Temperature Solid Oxide Fuel Cells (IT-SOFCs)<br />

Gunwoo Kim, 1 Frédéric Blanc, 1 and Clare P. Grey 1,2<br />

1 Department <strong>of</strong> Chemistry, State University <strong>of</strong> New York, Stony Brook, NY 11794-3400, USA<br />

2 Department <strong>of</strong> Chemistry, University <strong>of</strong> Cambridge, Lensfield Road, Cambridge, CB2 1EW, UK<br />

There are strong demands for clean energy sources<br />

to solve environmental issues. Fuel cells are a<br />

promising alternative to conventional, internal<br />

combustion engines as they are more efficient and<br />

cleaner. Among the various electrolytes used for<br />

IT-SOFCs, solid inorganic acids (e.g. CsH2PO4,<br />

CsHSO4)[1, 2] are drawing significant interest as<br />

they present unusually high protonic conductivity<br />

over the phase transition state, referred to as<br />

superprotonic phase. Although the conduction<br />

mechanism has not yet been completely addressed,<br />

it is generally accepted that the rotation <strong>of</strong> the<br />

tetrahedral anion, H2PO4 - facilitates proton<br />

conduction.[3] In order to fully understand the<br />

detailed mechanism, solid state NMR spectroscopy<br />

was used to examine the dynamics <strong>of</strong> the hydrogen<br />

onded protons and the rotational behavior <strong>of</strong> the<br />

trahedral anion in CsH PO .[4] The motions <strong>of</strong><br />

per<br />

at<br />

H line<br />

lysi<br />

O4 -<br />

b<br />

te 2 4<br />

protons were observed using<br />

spin e<br />

measur and H- P double resonance<br />

chniques to provide further insight into the<br />

mechanism <strong>of</strong> conduction.<br />

1 H and 2 H variabletem<br />

ature (VT) NMR experiments including<br />

relax ion time measurements. Different timescale<br />

for the motion was obtained by simulations <strong>of</strong> the<br />

2<br />

shape, 2-site exchange <strong>of</strong> 1 H spectra and<br />

ana s <strong>of</strong> relaxation times. The motion <strong>of</strong> the<br />

H2P anion was also investigated by using 31 P<br />

cho VT experiments, relaxation time<br />

1 31<br />

ements<br />

te<br />

P 24<br />

1. Boysen, D.A., et al., High-Performance Solid Acid Fuel<br />

Cells Through Humidity Stabilization. Science, 2004.<br />

303(5654): p. 68-70.<br />

2. Haile, S.M., et al., Solid acids as fuel cell electrolytes. Nature, 2001. 410(6831): p. 910-913.<br />

3. Haile, S.M., et al., Solid acid proton conductors:<br />

from laboratory<br />

curiosities to fuel cell electrolytes. Faraday<br />

Discussions, 2007. 134: p. 17-39.<br />

4. Yamada, K., et al., Superprotonic conductor CsH2PO4<br />

studied by 1H, 31P NMR and X-ray diffraction. Solid<br />

State Ionics, 2004. 175(1-4): p. 557-562.<br />

‐ 66 ‐<br />

Figure 1. Variable-temperature 1 H MAS spectra <strong>of</strong><br />

CsH2PO4 at a spinning speed <strong>of</strong> 14 kHz at 11.4 T,<br />

showing the effect <strong>of</strong> temperature on the two distinct<br />

environments. The isotropic peaks at room temperature<br />

are labeled.


First-Principles Studies <strong>of</strong> Proton - Ba tions in LaPO4<br />

Nicole Adelstein<br />

1,4 1,5 2,5<br />

, Jian Feng , Jeffrey A. Reimer , Jeffrey<br />

B. Neaton,<br />

1,3<br />

3<br />

and Lutgard C. De Jonghe<br />

‐ 67 ‐<br />

2+ Interac<br />

1 Materials Sciences Division<br />

2 Energy and Environmental <strong>Technology</strong> Division<br />

3 Molecular Foundry<br />

Lawrence Berkeley National Laboratory, 1 Cyclotron Road, Berkeley CA 94720, USA<br />

4 Department <strong>of</strong> Material Science and Engineering<br />

5 Department <strong>of</strong> Chemical Engineering<br />

U. <strong>California</strong> at Berkeley, Berkeley CA 94720, USA<br />

The<br />

proton‐dopant interaction in Ba‐doped LaPO4 is studied with first‐principles density<br />

functional<br />

theory to explore the role played by dopants in proton conduction. Using large<br />

supercells,<br />

we find that 3% Ba‐doping results in a local effective<br />

negative charge near the<br />

dopant that attracts protons. Protons near Ba dopants bind<br />

to oxygen much more strongly<br />

relative to sites far from the dopant, suggesting the prot<br />

room<br />

temperature. ped La<br />

e .3 Å,<br />

1H{ 137 on should sit near Ba at<br />

Ba NMR double resonance experiments on<br />

a related system, a Ba‐do<br />

metaphosphate glass, support this claim, finding that the av rage proton‐Ba distance is 3<br />

in good agreement with our calculations for Ba‐doped LaPO4.<br />

Fig. 1 The proton sits bounded<br />

to an oxygen. Sites are chosen<br />

based on the distance to the<br />

barium dopant.<br />

P 25


High Temperature Protonic Conduction Properties <strong>of</strong><br />

(La,Sr)MO3 – SrZrO3 (M = Cr, Mn and Fe) Solid Solutions<br />

Atsushi Unemoto A , Koji Amezawa B , and Tatsuya Kawada B<br />

A <strong>Institute</strong> <strong>of</strong> Multidisciplinary Research for Advanced Materials, Tohoku University<br />

B Graduate School <strong>of</strong> Environmental Studies, Tohoku University<br />

P 26<br />

To date, although protonic conductivity <strong>of</strong> the rare-earth doped SrZrO3 is widely<br />

investigated, the number <strong>of</strong> reports regarding the transition metal-doped one is limited to only a few<br />

reports. In our recent work on the electrical conductivity measurements <strong>of</strong> SrZr0.99Fe0.01O3-δ, it was<br />

found that the transport number <strong>of</strong> proton is relatively high even at higher temperatures, for<br />

instance, 0.75 at 1173 K in H2 – 1.9%H2O [1]. Since the solubility <strong>of</strong> Fe in SrZrO3 is limited to be at<br />

most 2.5 % [2], we additionally doped La in A-site to increase the doping level <strong>of</strong> Fe in B-site,<br />

aiming to increase the protonic conductivity. However, it resulted in only drastic decrease <strong>of</strong> the<br />

electrical conductivity and transport number <strong>of</strong> protons [3]. In this study, we examined the electrical<br />

conductivities <strong>of</strong> (La,Sr)MO3 – SrZrO3 (M = Cr and Mn) by two-probe ac technique to investigate<br />

the effect <strong>of</strong> co-doping <strong>of</strong> La and transition metal M (M = Cr, Mn and Fe) into SrZrO3.<br />

Figure 1 shows the electrical conductivities <strong>of</strong> (La0.9Sr0.1MO3-δ)0.05(SrZrO3)0.95 for M = Cr<br />

and Mn in H2 – 1.9%H2O and in D2 – 1.9%D2O at 1073 – 1273 K. In the case <strong>of</strong> M = Mn, the<br />

conductivity degraded gradually with time in lower oxygen partial pressure, suggesting<br />

decomposition <strong>of</strong> the oxide. Thus, the measurements could be performed only limited oxygen<br />

partial pressure range. Remarkable isotope effect <strong>of</strong> H and D was observed for both oxides similarly<br />

to the case <strong>of</strong> M = Fe [3], suggesting remarkable protonic conduction. It was also found that the<br />

total conductivity increased by decreasing oxygen partial pressure, corresponding to decrease <strong>of</strong><br />

mean valence <strong>of</strong> Cr in the oxide.<br />

When major positive defect is oxygen vacancy, the proton concentration is expected to be<br />

12<br />

proportional to pH2O. Thus, by assuming the linear relation between the proton conductivity and the<br />

concentration, the contribution <strong>of</strong> protonic conductivity could be separated under constant oxygen<br />

o 12<br />

partial pressure by the equation, � � ����<br />

��<br />

�� � pH2O,<br />

at temperature from 1273 – 1073 K. �� H ��is<br />

the<br />

o<br />

conductivities except for protonic one and � � � H<br />

is a constant. Transport number <strong>of</strong> proton was<br />

found to depend on oxygen partial pressure<br />

similarly to the case <strong>of</strong> M = Fe [3]. At 1173 K<br />

in log( pH2 O / atm) = 0.019, it varied from 0.63<br />

to 0.73 in the oxygen partial pressure range <strong>of</strong><br />

log( p / atm) = -<strong>15</strong>.58 to -19.57.<br />

O2<br />

1. A. Unemoto, A. Kaimai, K. Sato, N. Kitamura, K.<br />

Yashiro, H. Matsumoto, J. Mizusaki, K. Amezawa,<br />

T. Kawada, Solid State Ionics 181 (2010) 868.<br />

2. A. Unemoto, A. Kaimai, K. Sato, K. Yashiro, H.<br />

Matsumoto, J. Mizusaki, K. Amezawa, T. Kawada,<br />

Solid State Ionics 178 (2008) 1663.<br />

3. A. Unemoto, A. Kaimai, K. Sato, N. Kitamura, K.<br />

Yashiro, H. Matsumoto, J. Mizusaki, K. Amezawa,<br />

T. Kawada (In Preparation).<br />

‐ 68 ‐<br />

Figure 1. Electrical conductivity <strong>of</strong> (La0.9Sr0.1MO3δ)0.05(SrZrO3)0.95<br />

(M = Cr, Mn and Fe) in humidified hydrogen.<br />

Closed and open symbols are the values in H2 – 1.9%H2O and<br />

D2 – 1.9%D2O, respectively.


Electrical Conductivity and Defect Structure <strong>of</strong> Sr-Doped Nd3PO7<br />

Atsushi Unemoto A , Koji Amezawa B , and Tatsuya Kawada B<br />

A <strong>Institute</strong> <strong>of</strong> Multidisciplinary Research for Advanced Materials, Tohoku University<br />

B Graduate School <strong>of</strong> Environmental Studies, Tohoku University<br />

P 27<br />

A series <strong>of</strong> phosphate-based proton conductors is a promising candidate as a solid electrolyte<br />

[1-3] because they are expected to have high tolerance against ambient gases such carbon dioxide<br />

and water vapor contrary to proton conducting cerates [4]. In this study, we chose the rare-earth<br />

oxyphosphate, (Nd,Sr)3PO7, as a phosphate-based proton conductor. Although crystal structure<br />

analysis [5] and luminescent properties [6] <strong>of</strong> the rare-earth oxyphosphates are revealed so far, the<br />

electrical conduction properties are still unclear.<br />

Powders <strong>of</strong> the rare earth oxyphosphates, (Nd1-xSrx)3PO7-δ (x = 0 and 0.03), were prepared by<br />

a conventional solid state reaction route. The sintered bodies were obtained by a spark plasma<br />

sintering (SPS) technique. Platinum-paste was painted on both surfaces as electrodes. The electrical<br />

conductivities were evaluated by a two-probe ac technique in oxygen environments.<br />

The electrical conductivity <strong>of</strong> Sr-doped Nd3PO7 showed higher values than that <strong>of</strong> non-dope<br />

one by approximately one – half order <strong>of</strong> magnitude in 1.0%O2 – 1.9%H2O and unhumidified<br />

1.0%O2. This suggests that doping <strong>of</strong> Sr into Nd3PO7 enhanced the electrical conductivity.<br />

Figure 1 shows the electrical conductivity <strong>of</strong> (Nd0.97Sr0.03)3PO7-δ as a function <strong>of</strong> oxygen<br />

partial pressure in unhumidified and humidified environments. It was found from the figure that the<br />

electrical conductivity in humidified oxygen is smaller than that in unhumidified one, obviously at<br />

temperatures below 1073 K and as oxygen partial pressure decreased. When major positive defect is<br />

12 14<br />

oxygen deficit, proton and electron hole concentrations are expected to depend on pH2O and pO2 ,<br />

respectively. Thus, the electrical conductivity positively depends on water vapor pressure in proton<br />

conducting phosphates [3]. On the other hand, when major positive defect is proton, the oxygen<br />

�1 �1 2 14<br />

deficit and electron hole concentrations are expected to depend on pH2O and pH2O pO2 , respectively.<br />

As seen in the figure, the electrical conductivity decreased by increasing the water vapor pressure<br />

under constant oxygen partial pressure at temperatures below 1073 K while it increased as oxygen<br />

partial pressure increased under constant water vapor pressure. Considering the gas partial pressure<br />

dependences <strong>of</strong> the electrical conductivity, it might be concluded that protonic defect is induced in<br />

oxyphosphates, however, the ceramics does not allow their significant conduction. This means,<br />

major ionic carrier is oxide ion rather than<br />

proton in (Nd0.97Sr0.03)PO7-δ.<br />

1. T. Norby, N. Christiansen, Solid State Ionics 77 (1995)<br />

240.<br />

2. K. Amezawa, S. Kjelstrup, T. Norby, Y. Ito, J.<br />

Electrochem. Soc. 145 (1998) 3313.<br />

3. K. Amezawa, H. Maekawa, Y. Tomii, N. Yamamoto,<br />

Solid State Ionics 145 (2001) 233.<br />

4. H. Matsumoto, S. Okada, S. Hashimoto, K. Sasaki, R.<br />

Yamamoto, M. Enoki, T. Ishihara, Ionics 13 (2007)<br />

93.<br />

5. E. G. Tselebrovskaya, B. F. Dzhurinskii, O. I.<br />

Lyamina, Inorg. Mater. 33 (1997) 52.<br />

6. S. Ku, J. Zhang, J. Lumin. 122-123 (2007) 500.<br />

‐ 69 ‐<br />

Figure 1. Electrical conductivity <strong>of</strong> (Nd0.97Sr0.03)3PO7-δ as<br />

a function <strong>of</strong> oxygen partial pressure at 973 – 1273 K.


Mechanochemical Synthesis <strong>of</strong> Proton Conductive CsHSO4-Azole Composites<br />

for Medium Temperature Dry Fuel Cells<br />

Song-yul Oh 1 , Toshihiro Yoshida 1 , Go Kawamura 2 , Hiroyuki Muto 1,2 , and Atsunori Matsuda 1,2<br />

(1) Department <strong>of</strong> Environmental and Life Sciences, Toyohashi University <strong>of</strong> <strong>Technology</strong>,<br />

Tempaku-cho, Toyohashi, Aichi 441-8580, JAPAN<br />

(2) Department <strong>of</strong> Electrical and Electronic Information Engineering, Toyohashi University <strong>of</strong> <strong>Technology</strong>, Tempakucho,<br />

Toyohashi, Aichi 441-8580, JAPAN<br />

P 28<br />

Inorganic solid acids, such as sulfates and phosphates, have been widely studied because <strong>of</strong><br />

their promising proton conductivity for the application to electrochemical devices and unique phasetransition<br />

to super-protonic state. 1,2) However, despite their high proton conductivity in the superprotonic<br />

regions, they show much lower proton conductivity at temperatures lower than the<br />

transition temperature.<br />

On the other hand, we have shown that a solid state reaction involving mechanochemical<br />

treatment using a high-energy ball mill is a promising way to improve the proton conductivity <strong>of</strong><br />

inorganic solid acids and to synthesize a new class <strong>of</strong> solid acid composites. 3,4) In this study,<br />

inorganic-organic composites were synthesized by the solid state mechanochemical method, and<br />

their proton conductivity and structure studies were performed. CsHSO4 (CHS) was chosen as an<br />

inorganic acidic compound to investigate its super-protonic behavior, and three kinds <strong>of</strong> azoles,<br />

imidazole (Iz, Tm = 89 o C), 1,2,4-triazole (Tz, Tm = 123 o C) and benzimidazole (Bz, Tm = 171 o C),<br />

were examined as organic basic compounds to use their self-dissociation behaviors.<br />

80CHS·20Azole composites (in mol. %) were prepared to improve the proton conductivity,<br />

especially in a temperature range lower than Tc <strong>of</strong> CHS in a dry atmosphere. According to this<br />

approach, it is expected that newly developed hydrogen bonding and acid-base ionic cluster between<br />

CHS and azoles involve promising proton conductivity in a wide range <strong>of</strong> temperature under<br />

anhydrous atmosphere.<br />

The azoles such as Iz, Tz and Bz are capable basic components that can act as a proton solvent,<br />

like water, due to its amphoteric (proton donor–acceptor sites) nature, intermolecular hydrogen bond<br />

formation and self-dissociation characteristics. 5) In Fig. 1, the CHS-Azole composites show higher<br />

proton conductivities, especially in<br />

temperature regions lower than the phase<br />

transition temperatures <strong>of</strong> their raw substances.<br />

Furthermore, the CHS-Iz and CHS-Tz show a<br />

value range close to 4.8×10 -4 to 1.9×10 -3 Scm -<br />

1 with in a wide temperature range <strong>of</strong> from 50<br />

to 160 o C, in a dry N2 atmosphere. These<br />

results, including their structural studies,<br />

indicate that chemical interactions between<br />

CHS and azoles were newly formed, which<br />

improve proton transfer by protonic hopping.<br />

Detailed results on the structural studies and<br />

proton conductivities <strong>of</strong> CHS-Azole<br />

composites with various molar ratios will be<br />

presented and discussed.<br />

[1] T. Norby, Solid State Ionics 125 (1999) 1.<br />

[2] S.M. Haile, et al., Nature 410 (2001) 910.<br />

[3] A. Matsuda, et al., Solid State Ionics 176 (2005) 2899.<br />

[4] A. Matsuda, et al., Solid State Ionics 177 (2006) 2421.<br />

[5] K.D. Kreuer, Chem. Mater. 8 (1996) 610<br />

Fig. 1. The temperature dependence <strong>of</strong> conductivity <strong>of</strong><br />

various CHS-Azole composites. CHS, Iz, Tz and Bz<br />

are for CsHSO4, imidazole, 1,2,4-triazole and<br />

benzimidazole, respectively.<br />

‐ 70 ‐


Micro solid oxide fuel cells with perovskite type proton conductive thin<br />

electrolytes<br />

Fumitada Iguchi, Kensuke Kubota, Syuji Tanaka, Noriko Sata,<br />

Masayoshi Esashi and Hiroo Yugami<br />

Graduate School <strong>of</strong> Engineering, Tohoku University<br />

Aoba 6-6-01, Aramaki, Aoba-ku, Sendai 980-8579, Japan<br />

P 29<br />

We fabricated micro - sized solid oxide fuel cells (m-SOFC), which based on proton conductive<br />

electrolytes, i.e. Y-doped BaZrO3 and Sr-doped LaScO3 on silicon wafers and perform power<br />

generation test <strong>of</strong> those cells. In this study, modified cell design with flat thin electrolytes and<br />

trapezoid shape flow channels were adopted to resolve the problems <strong>of</strong> mechanical stability. The<br />

schematic illustration <strong>of</strong> the cell design is shown in fig.1. As an electrolyte, <strong>15</strong>mol% Y-doped<br />

BaZrO3 or 20mol% Sr-doped LaScO3 were used, and at the both sides <strong>of</strong> the electrolyte, Pt-Pd<br />

porous electrodes were fabricated as electrodes.<br />

As shown in fig.2, free standing m-SOFC cells (300mm) with <strong>15</strong>mol% Y-doped BaZrO3 (BZY<strong>15</strong>)<br />

proton conductive thin electrolyte, were successfully fabricated on the (100) Si wafer. Because<br />

residual stress in the electrolyte was kept at compressive during operation to prevent mechanical<br />

failures, the electrolyte was significantly distorted. But, it is consistence during thermal cycles up to<br />

several times.<br />

The power generation test was performed in the temperature range <strong>of</strong> 286 o C to 430 o C with<br />

humidified H2 as anode gas and air as cathode gas. Observed open circuit voltage around 400 o C was<br />

over 1000mV. So, it is confirmed that anode and cathode were successfully separated by the<br />

electrolyte as same as conventional SOFC. The maximum power density was obtained at 386 o C and<br />

the value was 0.03 mW /cm 2 . The value is quite lower than leading power density (861mW /cm 2 at<br />

450 o C[1], <strong>15</strong>0mW /cm 2 [2] at 550 o C) using YSZ a electrolyte. In this study, although m-SOFC can<br />

work well, it is clear that the performance <strong>of</strong> the m-SOFC should be improved.<br />

Figure 1 schematic illustration <strong>of</strong> cell design Fig.2 �‐SOFC (□300mm)<br />

‐ 71 ‐<br />

100�m<br />

References<br />

[ 1] P.-C. Su, C.-C. Chao, J.H. Shim, R. Fasching and F.B. Prinz, Nano Letters 8 (2008) (8), p. 2289.<br />

[ 2] Anja Bieberle-Hutter et al, Journal <strong>of</strong> Power Sources 177 (2008), p.123.


Room-Temperature Protonic conduction in Nanocrystalline films <strong>of</strong> Yttria-<br />

Stabilized Zirconia<br />

Sangtae Kim, Hugo J. Avila-Paredes, Zuhair A. Munir<br />

Department <strong>of</strong> Chemical Engineering and Materials Science, University <strong>of</strong> <strong>California</strong>, Davis, CA 95616, USA<br />

We have recently reported that doped ZrO2 and CeO2<br />

ceramics conduct protons under wet atmosphere at<br />

relatively low temperatures (< 200�C) when their grain<br />

size becomes smaller than 100 nm [1-3]. The protonic<br />

conductivity increases rapidly with decreasing grain size,<br />

implying that protons indeed migrate along the grain<br />

boundaries under such conditions [2].In this contribution,<br />

we present the results <strong>of</strong> our investigation on protonic<br />

conduction in 1 �m-thick polycrystalline films <strong>of</strong> 8 mol%<br />

yttria-stabilized zirconia (YSZ) with an average grain size<br />

<strong>of</strong> 17 nm prepared on sapphire substrate using a spincoating<br />

process [4]. Protonic conductivity <strong>of</strong> the films is<br />

found to be higher by over an order <strong>of</strong> magnitude (at<br />

30�C ) than earlier reported values for nanocrystalline<br />

YSZ bulk ceramics with a similar grain size. This value is<br />

comparable to the oxygen-ionic conductivity <strong>of</strong> this<br />

material presented at about 400 �C. These results<br />

suggest that, besides grain size, the structural<br />

characteristics <strong>of</strong> the grain boundaries may also play an<br />

important role in determining protonic transport in this<br />

material, and possibly other nanocrystalline solid<br />

electrolytes<br />

(SEs), and that the protonic conductivity can<br />

further<br />

be enhanced by optimizing such characteristics <strong>of</strong><br />

the<br />

grain boundaries. Aa<br />

aa<br />

Acknowledgement<br />

The authors are grateful to Pr<strong>of</strong>. H.U. Anderson for providing the<br />

YZS film samples and also thank Pr<strong>of</strong>. Barrera-Calva for AFM<br />

measurements. The authors appreciate fruitful discussion with Pr<strong>of</strong>.<br />

M. Martin and Dr. R. A. De Souza.<br />

‐ 72 ‐<br />

� t T / (S cm -1 K)<br />

10 0<br />

10 -2<br />

10 -4<br />

10 -6<br />

10 -8<br />

10 -10<br />

600 400<br />

T / °C<br />

200<br />

b<br />

YSZ Film (17 nm)<br />

YSZ Pellet (13 nm)<br />

YSZ Pellet (100 nm)<br />

P 30<br />

P H2O = 2x10 -2 bar<br />

0.8 1.2 1.6 2.0 2.4 2.8 3.2 3.6<br />

1000T -1 /K -1<br />

Fig. 1 A SEM image <strong>of</strong> 1-�m thick YSZ<br />

film with a grain size <strong>of</strong> ~17 nm on a<br />

sapphire substrate (a); the conductivities<br />

measured form nano-structured samples<br />

indicated under wet air in the temperature<br />

range <strong>of</strong> 30 – 600 �C (b).<br />

1. S. Kim, U. Anselmi-Tamburini, H. J.Park, M. Martin and Z. A. Munir, ,Adv. Materi., 2008, 20, 556.<br />

2. H. J. Avila-Paredes, J. Zhao, S. Wang, M. Pietrowski, A. Reinholdt, Z. A. Munir, M. Martin, and S. Kim, J. Mater.<br />

Chem., 2010, 90, 990.<br />

3. S. Kim, H. J. Avila-Paredes, S. Wang, C. T. Chen, R. A. De Souza, M. Martin, Z. A. Munir, Phys. Chem. Chem.<br />

Phys., 2009, 11, 3035.<br />

4. Hugo J. Avila-Paredes, Enrique Barrera-Calva, Harlan U. Anderson, Roger A. De Souza, Manfred Martin, Zuhair<br />

A. Munir and Sangtae Kim, J. Mater. Chem., 2010, DOI: 10.1039/c0jm00051e.


P 31<br />

Fabrication <strong>of</strong> BaCe0.9Y0.1O3-� thin film on Pd substrate by UV-MOD<br />

Koichi Asano a , Yoshihiro Kozawa b , Yoshihiro Mugikura a,b , and Takao Watanabe a,b<br />

a Central Research <strong>Institute</strong> <strong>of</strong> Electric Power Industry, 2-6-1 Nagasaka, Yokosuka 240-0196, JAPAN<br />

b Yokohama National University, 79-7 Tokiwadai, Hodogaya, Yokohama 240-8501, JAPAN<br />

Introduction – Solid oxide fuel cell (SOFC) has been focused on the development <strong>of</strong> intermediate<br />

temperature (IT) range <strong>of</strong> 873-1073 K considering a long-term operating and low material cost. We<br />

have applied accepter-doped BaCeO3 to an electrolyte for IT-SOFC because it shows relatively-high<br />

proton conductivity and the cell performance <strong>of</strong> proton conductor is expected higher than that <strong>of</strong><br />

oxide ion conductor around IT region. However, in a CO2-containg atmosphere as a reformed<br />

natural gas, BaCeO3 doesn’t always posses sufficient chemical stability and loses its proton<br />

conduction. We have overcome that BaCeO3 is covered with a dense Pd 1) . Pd protected BaCeO3<br />

from the reformed gas, and function at the same time as an anode. On the other hand, the reduction<br />

in operating temperature <strong>of</strong> SOFC leads to rapid increase in the bulk resistance <strong>of</strong> the electrolyte.<br />

Therefore, thin films as an electrolyte are generally considered. However, thermal resistances <strong>of</strong> Pd<br />

and cathode materials are normally lower than the sintering temperature <strong>of</strong> BCYO. In this study, the<br />

thin film to achieve the IT-SOFC that applied BaCe0.9Y0.1O3-� (BCYO) as the proton conductor was<br />

fabricated on dense Pd substrate at low sintering temperature by metal organic deposition (MOD)<br />

method.<br />

Experimental – The metal organic acid salt solutions<br />

(Kojundo Chemical Laboratory Co. Ltd., Japan) as<br />

precursor materials were applied to Pd substrate<br />

(Pd/Ag=75/25%, φ18 mm, 0.2 mm t ) by a spin-coating<br />

technique and were pyrolyzed at 523 K. This process<br />

was repeated 8 times and it was then sintered at 1123<br />

K, 0.5h in air.<br />

Result and discussion – BCYO film layer was<br />

fabricated on Pd substrate with good adhesion. The<br />

thickness <strong>of</strong> thin film is about 0.8 �m. However, the<br />

BCYO thin film had low density as an electrolyte as<br />

shown in Fig.1(a). The UV-MOD as a new process,<br />

which was added with ultraviolet lamp irradiation<br />

(254 nm) after spin-coating at room temperature, was<br />

applied. The thickness <strong>of</strong> thin film is about 0.45 �m as<br />

shown in Fig.1(b). It was found that the BCYO thinfilm<br />

fabricated by UV-MOD method was achieved<br />

high density compared with the MOD method. The<br />

XRD pattern <strong>of</strong> BCYO is shown in Fig.2. BCYO and<br />

Pd are mainly identified and BCYO single phase was<br />

formed on Pd substrate. But it seems that the<br />

conductivity <strong>of</strong> BCYO was rather low, 10 -4 Scm -1 at<br />

873 K in wet hydrogen atmosphere because the small<br />

amount <strong>of</strong> PdO was formed by sintering process.<br />

Reference – [1] K. Asano, M. Kawakami, Y. Mugikura, and T. Watanabe, ECS Trans., 7, (1), 993 (2007).<br />

‐ 73 ‐<br />

(a)<br />

Cover layer<br />

BCYO<br />

(b)<br />

Cover layer<br />

BCYO<br />

Pd substrate Pd substrate<br />

Fig.1 Cross-sectional SEM images <strong>of</strong> BCYO thin<br />

film on Pd substrate milled by FIB.<br />

(a) MOD, (b)UV-MOD.<br />

Fig.2 XRD pattern <strong>of</strong> BCYO thin film on<br />

Pd substrate.


First principles calculations <strong>of</strong> defect equilibria in BaZrO3<br />

Akihide Kuwabara 1 , Craig A. J. Fisher 1 , Hiroki Moriwake 1 ,<br />

Fumiyasu Oba 2 , Katsuyuki Matsunaga 2 , Isao Tanaka 1,2<br />

Nanostructures Research Laboratory, Japan Fine Ceramics Center, 2-4-1 Mutsuno, Nagoya, 456-8587, Japan<br />

Department <strong>of</strong> Materials Science, Kyoto University, Yoshida, Kyoto, 594-8587, Japan<br />

P 32<br />

Acceptor-doped BaZrO3 exhibits high proton conductivity at elevated temperatures in wet<br />

atmospheres [1]. It may thus be a candidate for the electrolyte in a proton conducting solid oxide<br />

fuel cell. Acceptor doping <strong>of</strong> BaZrO3 forms oxide ion vacancies in order to maintain electrical<br />

neutrality. In wet atmospheres, the vacancies react with water vapor, whereby the vacancies are<br />

filled and protons are introduced into the lattice. Theoretical calculations based on density functional<br />

theory have been systematically performed to evaluate the defect energetics <strong>of</strong> BaZrO3 [2-3].<br />

However, defect equilibria involving cation defects are not<br />

yet well understood. The aim <strong>of</strong> the present study is to<br />

clarify defect formation behavior.<br />

Total energy calculations were carried out using<br />

VASP code [4]. Electron-ion interactions were represented<br />

by the projector augmented wave (PAW) method, and the<br />

generalized gradient approximation (GGA) was chosen for<br />

the exchange-correlation potential. Plane waves up to an<br />

energy cut-<strong>of</strong>f <strong>of</strong> 400 eV were used as the basis set for wave<br />

functions. First, a primitive cell <strong>of</strong> BaZrO3 was fully<br />

optimized, and then supercells containing 320 atoms (4 � 4<br />

� 4 unit cells) were constructed for calculations <strong>of</strong> defective<br />

systems. In this study, we focused on defects in pure<br />

BaZrO3. Only the Γ point was adopted as k-point sampling<br />

for the supercells. All atomic positions in defective systems<br />

were fully relaxed until all residual forces were less than<br />

0.02 eV/Å. Details <strong>of</strong> the calculation methodologies <strong>of</strong><br />

point defect formation energy are available elsewhere [5-6].<br />

As shown in Figure 1, native defects <strong>of</strong> BaZrO3,<br />

��<br />

oxygen vacancies ( VO ) and Ba vacancies ( V� � Ba ), are likely<br />

to form under both Zr-rich and Ba-rich conditions at 1900 K.<br />

From the point <strong>of</strong> view <strong>of</strong> electrical neutrality, the major<br />

��<br />

defects are partial Schottky pairs <strong>of</strong> V ��Ba � VO<br />

. Under the Zr-<br />

��<br />

rich condition, anti-site defects <strong>of</strong> Zr ions on Ba sites ( ZrBa<br />

)<br />

can also be formed.<br />

Figure 1. Dependence <strong>of</strong> calculated defect<br />

formation energy on Fermi energy under (a)<br />

Ba-rich and (b) Zr-rich conditions at T =<br />

1900 K, pO2 = 0.1 atm and pH2O = 0.02<br />

atm. In the horizontal axis, the valence band<br />

top is set to be 0 eV. Conduction band<br />

bottom is 5.4 eV [7].<br />

1. S. Imashuku, T. Uda, Y. Nose, G. Taniguchi, Y. Ito, and Y.<br />

Awakura, Journal<br />

<strong>of</strong> The Electrochemical Society, <strong>15</strong>6 (2009) B1.<br />

2. M. E. Björketun, P. G. Sundell and G. Wahnström, Physical Review B, 76 (2007) 054307.<br />

3. M. E. Björketun, P. G. Sundell and G. Wahnström, Faraday Discussions, 134 (2007) 247.<br />

4. G. Kresse and J. Furthmüller, Physical Revew B, 54 (1996) 11169.<br />

5. A. Kuwabara and I. Tanaka, Journal <strong>of</strong> Physical Chemistry B, 108 (2004) 9168.<br />

6. A. Kuwabara, R. Haugsrud, S Stølen and T. Norby, Physical Chemistry Chemical Physics, 11 (2009) 5550.<br />

7. G. Łupina, J. Dąbrowski, P. Dudek, G. Kozłowski, P. Zaumseil, G. Lippert, O. Fursenko, J. Bauer, C. Baristiran, I.<br />

Costina, H.-J. Müssig, L. Oberbeck, and U. Schröder, Applied Physics Letters, 94 (2009) <strong>15</strong>2903.<br />

‐ 74 ‐


On the symmetry <strong>of</strong> defects in perovskite-type protonic conductors:<br />

A Sc-45 NMR study<br />

Itaru Oikawa 1 , Mariko Ando 1 , Hajime Kiyono 2 , Masataka Tansho 3 , Tadashi Shimizu 3 ,<br />

and Hideki Maekawa 1<br />

1 Graduate School <strong>of</strong> Engineering, Tohoku University, Aramaki Aoba 6-6-02, Sendai, 980-8579, Japan<br />

2 Graduate School <strong>of</strong> Engineering, Hokkaido University, N13, W8, Kita-ku, Sapporo, 060-8628, Japan<br />

3 National <strong>Institute</strong> for Materials Science, Sakura 3-13, Tsukuba, 305-0003, Japan<br />

Introduction<br />

Perovskite-type protonic conductors AB1-xMxO3-� (M: rare-earth elements) are a candidate<br />

for electrolytes <strong>of</strong> intermediate temperature solid oxide fuel cells due to high protonic conductivity.<br />

Protonic conductivity varies with A and B site cations as well as dopant cations. Nuclear magnetic<br />

resonance (NMR) can provide local structural information around specific ions in oxide materials.<br />

In the present study, defects around dopant cations in perovskite-type protonic conductors were<br />

investigated<br />

by Sc-45 NMR.<br />

Experimental<br />

Samples <strong>of</strong> AB1-xScxO3-� (A = Ba, Ca and Sr, B = Ce and Zr) were prepared by solid state<br />

reaction. Raw materials were mixed, pelletized and calcined at 1473 K in air. Calcined pellets were<br />

ground, mixed by a ball mill and sintered at 1723-1873 K in air. Sintered samples were vacuum<br />

dried at 1173 K. Sc-45 magic-angle spinning (MAS)-NMR were carried out at JNM-ECA930<br />

spectrometer (21.8 T) with Sc-45 resonance frequency 225.78 MHz. NMR spectra were collected by<br />

single pulse experiment with pulse length <strong>of</strong> 1 �s at spinning speed 22 kHz.<br />

Results and discussion<br />

Sc-45 MAS-NMR spectra <strong>of</strong> vacuum dried II: ScO ScO6-H I: ScO<br />

6-H I: ScO 6<br />

AB0.90Sc0.10O3-� (A = Ba, Ca and Sr, B = Ce and Zr) III: ScO BaZrO<br />

5<br />

3<br />

showed obvious compositional dependence, as shown in *<br />

* *<br />

Fig. 1. Peaks I, II and III were assigned to sixcoordinated<br />

Sc (ScO6), six-coordinated Sc with protons<br />

III II I BaCeO 3<br />

*<br />

*<br />

reside vicinity (ScO6-H) and five-coordinated Sc with<br />

one oxygen vacancy (ScO5), respectively. The second-<br />

III II SrCeO 3 I<br />

order quadrupole coupling alters the line shape with its<br />

*<br />

*<br />

constants CQ. The line shape change <strong>of</strong> ScO5 (hatched<br />

III I CaZrO 3<br />

peaks in Fig. 1) was interpreted as a result <strong>of</strong> symmetry<br />

change <strong>of</strong> ScO5. CQ <strong>of</strong> peak III obtained from the peak<br />

300 250 200 <strong>15</strong>0 100 50 0<br />

simulation decreased in the order <strong>of</strong> BaZrO3, BaCeO3,<br />

chemical shift / ppm<br />

SrCeO3 and CaZrO3. We found that ScO5 polyhedra<br />

Fig. 1 Sc-45 MAS-NMR spectra <strong>of</strong> vacuum<br />

change its symmetry with A and B site cations for the dried AB0.90Sc0.10O3-� (A = Ba, Ca and Sr, B<br />

first time, and it will open a new paradigm to defect = Ce and Zr). Asterisks indicate spinning<br />

engineering based on the symmetry <strong>of</strong> oxygen polyhedra. sidebands.<br />

Acknowledgements<br />

H.M. wishes to express his sincere gratitude for the financial support provided by CREST,<br />

JST under “Novel Measuring and Analytical <strong>Technology</strong> Contributions to the Elucidation and<br />

Application <strong>of</strong> Materials.” and by KAKENHI “(Grant No. 21360314)” MEXT, Japan.<br />

‐ 75 ‐<br />

P 33


Performances <strong>of</strong> Reversible SOFCs with BaZr0.6Co0.4O3-δ as air electrode<br />

Fei He, Ranran Peng, Changrong Xia<br />

Department <strong>of</strong> Material Science and Engineering,<br />

University <strong>of</strong> Science and <strong>Technology</strong> <strong>of</strong> China, Hefei, 230026, P.R.China<br />

Introduction: Reversible solid oxide fuel cells with proton<br />

conducting electrolyte (H-RSOFC) have attracted much<br />

attention because <strong>of</strong> its special characteristics[1]. In H-<br />

RSOFCs, major part <strong>of</strong> polarization resistances comes from<br />

air electrodes due to their low catalytic activities. In this<br />

paper, BaZr0.6Co0.4O3-δ were prepared by citric method, and<br />

applied as single phase air electrode for H-RSOFCs. The<br />

electro-performance <strong>of</strong> such H-RSOFCs was investigated<br />

with BaCe0.5Zr0.3Y0.2O3-δ as electrolyte. The Reversible<br />

solid oxide fuel cells with structure <strong>of</strong> Ni-BZCY|| BZCY||<br />

BZCO were fabricated and tested in a home-developed-celltesting<br />

system with 30%H2O-air and H2 introduced into air<br />

electrode and hydrogen electrode, respectively.<br />

Results and Disscussion: The performances <strong>of</strong> RSOFC<br />

were investigated in both SOFC mode and SOEC mode at<br />

600, 650 and 700 o C, as shown in Fig. 1. The open circuit<br />

voltages (OCV) <strong>of</strong> cells are 0.99, 0.97, 0.95V at 600, 650<br />

and 700 o C, respectively, indicate the electrolyte is very<br />

dense. At SOFC mode (the right part <strong>of</strong> Fig.1), the current<br />

densities increase with temperature, is about 300mAcm -2 at<br />

0.7V and 700 o C. On the other hand, the performance <strong>of</strong> cells<br />

in the SOEC mode also increases with temperature (see the<br />

left panel <strong>of</strong> Fig. 1), and the electrolysis current densities at<br />

cell voltage <strong>of</strong> 1.3V is -530 mAcm -2 at 700 o C. In Fig. 2. the<br />

impedance spectra <strong>of</strong> the H-RSOFCs with BZCO single<br />

phase electrode and BCZY/ Sm0.5Sr0.5O3-δ(SSC) composite<br />

electrode were compared at 700 o C, as shown in Fig. 2. The<br />

bulk resistance <strong>of</strong> the cell with BZCO electrode is bigger<br />

than which with BCZY/SSC electrode, however, the<br />

interfacial polarization resistance <strong>of</strong> the cell with BZCO<br />

lectrode is 0.2Ωcm 2 e<br />

, about half <strong>of</strong> that with BCZY/SSC<br />

electrode.<br />

1. M. Ni, M. K. H. Leung, D. Y. Leung, International Journal <strong>of</strong><br />

Hydrogen Eneryg 33(2008) 4040-4047<br />

‐ 76 ‐<br />

P 34<br />

Fig. 1 I-V curves <strong>of</strong> RSOFCs measured at<br />

various temperatures.<br />

Fig. 2 The impedance spectra <strong>of</strong> the H-RSOFCs<br />

with BZCO and BCZY/SSC electrode were<br />

compared at 700 o C.


Space charge effect and dopant segregation in acceptor-doped BaZrO3<br />

proton conductors<br />

Mona Shirpour 1) , Behnaz Rahmati 2) , Wilfried Sigle 2) ,<br />

Peter A. van Aken 2) , Rotraut Merkle 1) , and Joachim Maier 1)<br />

1) Max Planck <strong>Institute</strong> for Solid State Research, Heisenbergstr. 1, D-70569 Stuttgart, Germany<br />

2) Max Planck <strong>Institute</strong> for Metals Research, Heisenbergstr. 3, D-70569 Stuttgart, Germany<br />

While the bulk proton conductivity is high, the grain<br />

boundaries <strong>of</strong> acceptor-doped BaZrO3 are highly<br />

resistive, resulting in a low total conductivity in<br />

polycrystalline samples [1-4]. One possible<br />

explanation for this blocking character is the presence<br />

<strong>of</strong> an excess positive charge in the grain boundary<br />

core and depletion <strong>of</strong> protons, holes and oxygen<br />

vacancies in the adjacent space charge zone. To gain<br />

evidence for this space charge model, Y-doped<br />

BaZrO3 (sintered by SPS “Spark Plasma Sintering”:<br />

1600°C, 5 min, 50 MPa) was reduced using a layer <strong>of</strong><br />

metallic zirconium (Fig. 1). The absence <strong>of</strong> blocking<br />

grain boundaries in the reduced (Fig. 2), nconducting<br />

ceramic is an evidence for a positively<br />

charged core (electron accumulation in the space<br />

charge zone).<br />

The grain boundary properties depend strongly on<br />

processing conditions. The grain boundary<br />

conductivity <strong>of</strong> Sc- or Y- doped BaZrO3 sintered by<br />

SPS increases by a factor <strong>of</strong> 10 to 1000 at 300°C with<br />

additional high-temperature annealing (1700°C for 20<br />

h). Quantitative energy-dispersive X-ray<br />

spectroscopy in the TEM showed a significant<br />

segregation <strong>of</strong> negatively charged dopants to the<br />

grain boundary region (core and/or to the space<br />

charge layers). The detected dopant segregation is<br />

more pronounced for the grain boundaries with<br />

higher conductivity increase. In addition, different<br />

contributions<br />

<strong>of</strong> segregation driving forces in Y-<br />

doped<br />

and Sc-doped BaZrO3 are considered.<br />

P 35<br />

Fig. 1 SPS-6at%Y-BaZrO3 (a) before reduction (b)<br />

after reduction and removing the oxidized Zr layer<br />

1.<br />

2.<br />

K. D. Kreuer, Annual Review <strong>of</strong> Materials Research 33<br />

(2003), 333.<br />

F. Iguchi, T. Yamada, N. Sata, T. Tsurui, and H. Yugami,<br />

Solid State Ionics 177 (2006), 2381.<br />

Fig. 2 (a) Blocking grain boundaries in 6at%Ydoped<br />

BaZrO3 in dry N2 and at 24 °C (b) absence <strong>of</strong><br />

blocking grain boundaries in reduced sample<br />

3. Y. Yamazaki, R. Hernandez-Sanchez, and S. M. Haile,<br />

Chemistry <strong>of</strong> Materials 21 (2009), 2755.<br />

4. C. Kjolseth, H. Fjeld, O. Prytz, P. I. Dahl, C. Estournes, R. Haugsrud, and T. Norby, Solid State Ionics 181 (2010),<br />

268.<br />

‐ 77 ‐


Synthesis and characterization <strong>of</strong> BaCe0.7Zr0.1Y0.1Yb0.1O3-δ proton<br />

conducting ceramic<br />

Siwei Wang, Fei Zhao, Lingling Zhang, and Fanglin Chen*<br />

Department <strong>of</strong> Mechanical Engineering, University <strong>of</strong> South Carolina, Columbia, SC 29208, USA<br />

BaCe0.7Zr0.1Y0.1Yb0.1O3-δ proton conducting material<br />

for intermediate temperature solid oxide fuel cells<br />

(IT SOFCs) applications has been prepared by a<br />

modified pechini method [1]. Pure perovskite<br />

structured materials were obtained with calcination<br />

temperatures T≥1000 o C. The thermal expansion<br />

coefficient (TEC) for BaCe0.7Zr0.1Y0.1Yb0.1O3-δ is<br />

9.1-9.7×10 -6 K -1 from 25 to 1100 o C, which is<br />

favorably matched with that <strong>of</strong> Pr based cathode<br />

materials [2]. The microstructures <strong>of</strong> the samples<br />

sintered at different temperatures from 1350 to<br />

<strong>15</strong>50 o C were analyzed, revealing denser<br />

microstructure with higher sintering temperatures.<br />

The correlation between the microstructure and the<br />

conductivity was investigated. The sample sintered at<br />

1400 o C showed the highest conductivity (Figure 1),<br />

indicating that the beneficial effects on conductivity<br />

from denser samples was counteracted by the<br />

segregation <strong>of</strong> secondary phases in the grain<br />

boundaries<br />

at higher sintering temperatures [3].<br />

1. F. Zhao, Q. Liu, S. Wang, K. Brinkman and F. Chen,<br />

International Journal <strong>of</strong> Hydrogen Energy 35 (9) 4258.<br />

2. F. Tietz, Ionics 5 (1999) (1) 129.<br />

3. J. Lv, L. Wang, D. Lei, H. Guo and R.V. Kumar, Journal <strong>of</strong><br />

Alloys and Compounds 467 (2009) (1-2) 376.<br />

[*] Corresponding Author: (Tel.) 803-777-4875; (E-mail):<br />

chenfa@cec.sc.edu<br />

‐ 78 ‐<br />

ln(��) (S.K/cm)<br />

ln(��) (S.K/cm)<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

P 36<br />

Temperature (<br />

800 750 700 650 600 550 500 450<br />

o C)<br />

(a) Wet Air<br />

sample A<br />

sample B<br />

sample C<br />

sample D<br />

sample E<br />

0.9 1.0 1.1 1.2 1.3 1.4<br />

1000/ T (1/K)<br />

Temperature (<br />

800 750 700 650 600 550 500 450<br />

o C)<br />

(b)<br />

Wet H 2<br />

sample A<br />

sample B<br />

sample C<br />

sample D<br />

sample E<br />

1.0<br />

0.9 1.0 1.1 1.2<br />

1000/ T (1/K)<br />

1.3 1.4<br />

Fig. 1 Arrhenius plots <strong>of</strong> conductivity for<br />

BaCe0.7Zr0.1Y0.1Yb0.1O3-δ samples<br />

sintered at different temperatures in (a)<br />

wet air and (b) wet H2, sample A-E<br />

indicating sintering temperatures from<br />

1350-<strong>15</strong>50 o C, at 50 o C intervals.


High flux ND-study <strong>of</strong> the structural phase transition in LaNbO4<br />

Christopher S. Knee 1 , Morten Huse 2 , Truls Norby 2 , Sten G. Eriksson 3 , Reidar Haugsrud 2<br />

1 Department <strong>of</strong> Chemistry, University <strong>of</strong> Gothenburg, SE-412 96 Göteborg, Sweden<br />

2 Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, FERMiO, Gaustadalléen 21, NO-0349 Oslo, Norway<br />

3 Department <strong>of</strong> Environmental Inorganic Chemistry, Chalmers Univ. <strong>of</strong> <strong>Technology</strong>, SE-412 96 Göteborg, Sweden<br />

P 37<br />

In the last decade renewed interest in LaNbO4 has emerged due to reports <strong>of</strong> appreciable proton<br />

conductivity and high proton transference number [1] as well as expected high stability in CO2 and<br />

CO2-H2O atmospheres. Consequently, LaNbO4 has been proposed as a candidate electrolyte<br />

material for proton conducting SOFCs. LaNbO4 exhibits a second order phase transition at ≈ 520 °C<br />

from a monoclinic structure (� ≈ 94 ° at RT) to a tetragonal<br />

cell (�� = 90 °). This has a strong impact on the material’s<br />

conductivity (Fig. 1). The figure suggests that the enthalpy <strong>of</strong><br />

proton mobility is significantly higher in the low temperature<br />

monoclinic phase. This is supported by a recent computational<br />

study on LaNbO4 [2] which also identifies an inter-tetrahedral<br />

pathway as the rate limiting step for proton migration. Here,<br />

we use data collected on the high flux neutron diffractometer<br />

D20 at the Institut Laue Langevin to accurately map the<br />

changes in the La-O and NbO4 environments through the<br />

second order phase transition. We find that the oxygen to<br />

oxygen distance between neighboring NbO4 tetrahedra, which<br />

is suggested as the rate limiting step for proton migration,<br />

decreases with increasing temperature up to the transition<br />

temperature, from where this separation was constant. To<br />

explain the difference in proton mobility in the two<br />

polymorphs we suggest that as the oxygen-oxygen distance<br />

decreases the proton jump distance decreases and<br />

consequently the activation barrier for inter-tetrahedral proton<br />

transfer decreases.<br />

Fig. 1 a) Conductivity vs. 1/T in various<br />

atmospheres [1]<br />

This investigation was supported by the Research Council <strong>of</strong> Norway (grant number 187160) within the N-INNER<br />

program.<br />

1.<br />

R. Haugsrud, T. Norby, Solid State Ionics 177 (2006), p. 1129.<br />

2.<br />

H. Fjeld, K. Toyoura, R. Haugsrud, T. Norby, Phys. Chem. Chem. Phys., (2010) DOI:10.1039/c002851g.<br />

‐ 79 ‐


Performance <strong>of</strong> SOFC with proton-conductor BaCe0.7In0.2Yb0.1O3-δ<br />

electrolyte<br />

Fei Zhao, Latoya Dixon and Fanglin Chen *<br />

Department <strong>of</strong> Mechanical Engineering, University <strong>of</strong> South Carolina, Columbia, SC 29208, USA<br />

Recently significant efforts have been focused on<br />

the development <strong>of</strong> electrolyte materials for protonconducting<br />

SOFCs. The strategy <strong>of</strong> co-doping has<br />

been successfully applied to BaCeO3 with M<br />

(M=Ta, Ti, In, Nb and Sn) and Y as the co-dopants<br />

and proved to maintain both high conductivity and<br />

good stability [1, 2]. In this work, the proton<br />

conductor BaCe0.7In0.2Yb0.1O3-δ (BCIYb) was<br />

fabricated by co-doping In and Yb at the Ce-site <strong>of</strong><br />

BaCeO3. Pure perovskite structure was obtained<br />

with calcination temperatures T≥1000 o C. The<br />

conductivities <strong>of</strong> BCIYb pellets sintered at 1450 o C<br />

for 5h were obtained in different atmospheres<br />

(shown in Fig. 1). The performance <strong>of</strong> the single<br />

cell with BCIYb as electrolyte was also<br />

investigated. As shown in Fig. 2, the cell showed<br />

aximum power outputs <strong>of</strong> 0.170, 0.238 and 0.303<br />

cm -2 at 600, 650 and 700 o m<br />

W<br />

C, respectively.<br />

1. K. Xie, R. Q. Yan and X. Q. Liu, Journal <strong>of</strong> Alloys and<br />

Compounds 479 (2009) L40.<br />

2.<br />

F. Zhao, Q. Liu, S.W. Wang, K. Brinkman and F. Chen,<br />

International Journal <strong>of</strong> Hydrogen Energy 35 (2010), 4258.<br />

[*] Corresponding Author: (Tel.) 803-777-4875;<br />

(E-mail): chenfa@cec.sc.edu<br />

‐ 80 ‐<br />

Fig. 1 Temperature (T) dependence <strong>of</strong> total<br />

conductivity (σ) <strong>of</strong> BaCe0.7In0.2Yb0.1O3-δ in<br />

different atmospheres.<br />

P 38<br />

Fig. 2 Performance <strong>of</strong> the single cell under<br />

humidified hydrogen atmosphere at different<br />

temperatures.


Ionic conduction in Mg–Al and Zn-Al layered double hydroxide<br />

intercalated with inorganic anions<br />

Kiyoharu Tadanaga, Yoshihiro Furukawa, Akitoshi Hayashi, and Masahiro Tatsumisago<br />

Department <strong>of</strong> Applied Chemistry, Graduate School <strong>of</strong> Engineering, Osaka Prefecture University,<br />

Sakai, Osaka 599-8531, Japan<br />

Layered double hydroxides (LDHs) are inorganic<br />

layered materials consisting <strong>of</strong> positively charged<br />

metal hydroxide layers and interlayer anions as<br />

well as water molecules. LDHs have been widely<br />

studied as ion-exchangers and trapping agents for<br />

anionic contaminants owing to their anionexchangeable<br />

property. In this study, the ionic<br />

conductivity <strong>of</strong> Mg-Al LDHs and Zn-Al LDHs<br />

intercalated with various inorganic anions such as<br />

CO3 2- , Cl - , Br - , NO3 - and SO4 2- was evaluated.<br />

The ionic conductivity <strong>of</strong> LDHs was found to be<br />

affected by interlayer anions, but the effects <strong>of</strong><br />

anions were different between Mg-Al and Zn-Al<br />

LDH systems. Figure 1 shows temperature<br />

dependence <strong>of</strong> conductivities <strong>of</strong> Mg-Al LDH<br />

intercalated with CO3 2- , Cl - , Br - , NO3 - and SO4 2-<br />

under 80% relative humidity. In Mg–Al LDH<br />

system, Mg-Al LDH intercalated with Br -<br />

exhibited the highest ionic conductivity, and the<br />

conductivity decreased in the order <strong>of</strong> Br - >Cl -<br />

=NO3 - = CO3 2- > SO4 2- [1]. The ionic conductivity<br />

<strong>of</strong> Mg-Al LDH intercalated with Br - was 1.1 ×<br />

10 - 2 S cm -1 at 80°C under 80% relative humidity.<br />

Figure 2 shows conductivities <strong>of</strong> Zn-Al LDH<br />

intercalated with CO3 2- , Cl - , Br - , NO3 - and SO4 2-<br />

under 80% relative humidity. In the Zn-Al LDHs,<br />

the ionic conductivity is rather low compared with<br />

that <strong>of</strong> Mg-Al LDHs. Zn-Al LDH intercalated<br />

with NO3 - exhibited the highest ionic conductivity,<br />

and the conductivity decreased in the order <strong>of</strong><br />

NO3 - > Cl - > Br - = SO4 2- > CO3 2- . The<br />

electromotive force for the water vapor<br />

concentration cell using Mg–Al and Zn-Al LDHs<br />

showed the same behavior with that using an anion<br />

exchange<br />

membrane, indicating that these LDHs<br />

are<br />

a hydroxide ion conductor.<br />

1. Y. Furukawa, K. Tadanaga, A. Hayashi, and M.<br />

Tatsumisago, Solid State Ionics (2010),<br />

DOI: 10.1016/j.ssi.2010.05.032 .<br />

‐ 81 ‐<br />

Conductivity / S cm -1<br />

Conductivity / S cm -1<br />

10 -1<br />

10 -2<br />

10 -3<br />

10 -4<br />

2.8 2.9 3 3.1 3.2 3.3<br />

1000/T / K -1<br />

10 -3<br />

10 -4<br />

10 -5<br />

10 -6<br />

Mg3-Al CO3 2�� Mg3-Al Cl<br />

LDH<br />

�� LDH<br />

Mg3-Al SO4 2�� Mg3-Al NO3<br />

LDH<br />

�� Mg3-Al Br<br />

LDH<br />

�� LDH<br />

2.8 2.9 3 3.1 3.2 3.3<br />

1000/T / K -1<br />

P 39<br />

Fig. 1. Temperature dependence <strong>of</strong> conductivities <strong>of</strong><br />

Mg-Al LDH intercalated with CO3 2- , Cl - , Br - ,<br />

NO3 - and SO4 2- under 80% relative humidity<br />

Zn2-Al CO3 2�� Zn2-Al Cl<br />

LDH<br />

�� Zn2-Al NO3<br />

LDH<br />

�� LDH<br />

Zn2-Al SO4 2��LDH Zn2-Al Br �� LDH<br />

Fig. 2. Temperature dependence <strong>of</strong> conductivities <strong>of</strong><br />

Zn-Al LDH intercalated with CO3 2- , Cl - , Br - , NO3 -<br />

and SO4 2- under 80% relative humidity


Determination <strong>of</strong> the enthalpy <strong>of</strong> hydration <strong>of</strong><br />

oxygen vacancies by TG-DSC<br />

Christian Kjølseth 1 , Andreas Løken 1 , Lin-Yung Wang 2 , Reidar Haugsrud 1 and Truls Norby 1<br />

1 Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, FERMiO, Gaustadallèen 21, NO-0349 Oslo, Norway<br />

2 Department <strong>of</strong> Chemistry, National Taiwan University, No.1, Lane 4, Roosevelt Rd., Taipei, Taiwan (R.O.C.)<br />

P 40<br />

Accurate determination <strong>of</strong> the hydration thermodynamics is vital for the fundamental understanding<br />

and further progression towards application <strong>of</strong> high temperature proton conductors. Hydration<br />

thermodynamics have traditionally been determined indirectly, based on modeling <strong>of</strong> experimental<br />

data from the water up-take in thermogravimetry or from partial and total conductivity<br />

measurements - relying on defect chemical models and <strong>of</strong>ten resulting in unreliable and correlated<br />

values (see e.g. [1,2]). In this contribution we report on combined thermogravimetry (TG) and<br />

Differential Scanning Calorimetry (DSC), TG-DSC, as a novel experimental approach to measure<br />

directly the standard molar hydration enthalpies <strong>of</strong> high temperature proton conductors. The<br />

technique is demonstrated for several high temperature proton conductors, e.g. two <strong>of</strong> the most<br />

studied proton conducting perovskite, BaZr0.9Y0.1O3-δ (BZY10) and BaCe0.9Y0.1O3-δ (BCY10). The<br />

measurements were conducted isothermally by parallel recording <strong>of</strong> the weight and heat exchanges<br />

associated with step changes in the water vapor pressure.<br />

Heat exchange / mW mg -1<br />

0,05<br />

0,00<br />

-0,05<br />

-0,10<br />

-0,<strong>15</strong><br />

-0,20<br />

BaZr 0.9 Y 0.1 O 3-�<br />

45 50 55 60 65 70<br />

Time/min<br />

100,2<br />

100,1<br />

100,0<br />

99,9<br />

99,8<br />

99,7<br />

Weight change / %<br />

Heat exchange / mW mg -1<br />

‐ 82 ‐<br />

0,12<br />

0,10<br />

0,08<br />

0,06<br />

0,04<br />

0,02<br />

0,00<br />

-0,02<br />

-0,04<br />

-0,06<br />

BaCe 0.9 Y 0.1 O 3-�<br />

45 50 55 60 65 70<br />

Time/min<br />

Fig. 1 TG‐DSC signals recording hydration <strong>of</strong> BaZr0.9Y0.1O3‐δ and BaCe0.9Y0.1O3‐δ<br />

at 600 °C.<br />

Fig. 1 shows the typical response in the TG-DSC signals following a shift in the gas from 1 atm N2<br />

to 1 atm H2O for BZY10 and BCY10 at 600 °C. The relative weight <strong>of</strong> the specimen increases<br />

abruptly as a consequence <strong>of</strong> the interaction with the water vapour. Simultaneously, we experience a<br />

peak in the DSC signal representing an exothermic process. The obtained molar hydration enthalpy<br />

<strong>of</strong> BZY10 was investigated at 300 - 900 °C and was found constant at -81±4 kJ mol -1 (per mole <strong>of</strong><br />

H2O) over this range. The molar hydration enthalpy <strong>of</strong> BCY10 was investigated at 600 °C and found<br />

to be -170±6 kJ mol -1 . Both values are in good agreement with the majority <strong>of</strong> literature data for the<br />

two materials obtained more indirectly by equilibrium methods.<br />

1. R. Haugsrud and T. Norby, Nature Materials 5 (2006) (3), p. 193.<br />

2. K.D. Kreuer et al., Solid State Ionics 145 (2001) (1-4), p. 295.<br />

100,1<br />

100,0<br />

99,9<br />

99,8<br />

99,7<br />

Weight change / %


Proton Transfer and Hydration in 3M Ionomers with Different Protogenic<br />

Groups<br />

Jeffrey K. Clark, and Stephen J. Paddison<br />

Department <strong>of</strong> Chemical and Biomolecular Engineering, University <strong>of</strong> Tennessee, Knoxville, TN 37996, USA<br />

P 41<br />

Proton exchange membranes (PEMs) are the electrolyte in current hydrogen fuel cells and function<br />

as not only the separator <strong>of</strong> the electrodes and reactant gases but importantly as the internal ion<br />

conductor. Efficient operation <strong>of</strong> these energy conversion devices in diverse applications (vehicular,<br />

portable, and stationary) places demands on the PEM which include: long-time thermal and<br />

chemical stability (including resistance to oxidation and degradation by reactive species) at<br />

temperatures as high as 120°C, and high proton conductivity (≈ 10 -1 Scm -1 ) under low humidity<br />

conditions (25-50% relative humidity). Although a large number <strong>of</strong> strategies have been devised in<br />

the<br />

pursuit to design membrane materials that meet these requirements current PEM fuel cells still<br />

utilize perfluorosulfonic acid (PFSA) ionomers such as Nafion®.<br />

We present a molecular modeling study on the<br />

role <strong>of</strong> local hydration, side‐chain connectivity, and<br />

protogenic group separation in effecting proton<br />

dissociation in oligomeric fragments <strong>of</strong> 3M<br />

ionomers at minimal hydration. Two types <strong>of</strong><br />

ionomers, each consisting <strong>of</strong> a polytetrafluoro‐<br />

ethylene (PTFE) backbone, were considered in this<br />

study. These are: (1) perfluorosulfonic acid (PFSA)<br />

membranes at different equivalent weights with<br />

two pendant side‐chains <strong>of</strong> distinct separation<br />

(with chemical formula CF3CF(‐O(CF2)4SO3H)(CF2‐<br />

CF2)nCF(‐O(CF2)4SO3H)CF3, with n = 1 and 2) and<br />

(2) imide fragments <strong>of</strong> structural isomers with<br />

multiple and distinct pendant acid groups (with<br />

chemical formula CF3CF2CF(‐<br />

O(CF2)4SO2(NH)SO2C6H4SO3H)CF3 with the<br />

sulfonic acid group located in either the meta or the ortho position). The initial ‘dry’<br />

optimizations <strong>of</strong> these structures are shown in Figure 1. Fully optimized structures <strong>of</strong> these<br />

fragments with and without the addition <strong>of</strong> water molecules at the B3LYP/6‐311G** level<br />

revealed that both side‐chain connectivity and protogenic group separation, along with local<br />

hydration, are key contributors to proton dissociation in these membranes. Specifically,<br />

interaction between protogenic groups via hydrogen‐bonding networks proved to be<br />

necessary for proton dissociation at low levels <strong>of</strong> hydration.<br />

Figure 1. Fully optimized ‘dry’ 3M oligomeric fragments: (a) EW 590 g/mol PFSA where n=1, (b) EW 690 g/mol<br />

PFSA where n=2, (c) meta bis acid, and (d) ortho bis acid. Atom color in the structures: grey–carbon, red–oxygen,<br />

yellow–fluorine, orange –sulfur, blue–nitrogen, and white–hydrogen.<br />

‐ 83 ‐


Electrical Conductivity and Defect Structure <strong>of</strong> Y-doped BaZrO3<br />

Ho-Il Ji, Jong-Ho Lee * , Hae-Ryoung Kim, Ji-Won Son,<br />

Hae-Weon Lee and Byung-Kook Kim<br />

High Temperature Energy Materials Center, Korea <strong>Institute</strong> <strong>of</strong> Science and <strong>Technology</strong>, Seoul 136-791, Korea<br />

Rare earth doped BaZrO3 is one <strong>of</strong> the promising proton conducting oxides due to its high proton<br />

conductivity and fair chemical stability. However, poor sinterability <strong>of</strong> doped BaZrO3 has been a<br />

critical barrier to be employed as a solid electrolyte <strong>of</strong> fuel cell. For this reason there have been<br />

many efforts to improve the sinterability <strong>of</strong> doped BaZrO3 by adopting sintering aids such as CuO,<br />

ZnO.<br />

In this study, we measured the electrical conductivity <strong>of</strong> Y doped BaZrO3 (BZY) with- and without<br />

a sintering aid, (Cu, Zn) as a function <strong>of</strong> temperature, oxygen partial pressure and water vapor<br />

pressure and investigated the influence <strong>of</strong> sintering aids on its electrical property and defect<br />

structure. For a more quantitative defect structural analysis, proton concentrations were precisely<br />

measured by a Thermo-Gravimetric Analysis (TGA). On the basis <strong>of</strong> conductivity and TGA results,<br />

equilibrium concentrations <strong>of</strong> each ionic and electronic charge carrier was calculated and presented<br />

in a three dimensional space (X axis: oxygen partial pressure, Y axis: water vapor pressure, Z axis:<br />

ionic & electronic conductivity or concentration <strong>of</strong> ionic & electronic charge carriers).<br />

‐ 84 ‐<br />

P 42


Enhanced Sinterability <strong>of</strong> Y-doped BaZrO3 Powder Synthesized with CuO,<br />

ZnO Addition as a Sintering Aid<br />

Jong-Ho Lee, Seong-Jeong Hong, Ho-Il Ji, Jong-Sung Park, Hae-Ryoung Kim,<br />

Ji-Won Son, Hae-Weon Lee, and Byung-Kook Kim *<br />

High Temperature Energy Materials Center, Korea <strong>Institute</strong> <strong>of</strong> Science and <strong>Technology</strong>, Seoul 136-791, Korea<br />

Rare earth doped BaZrO3 is known to be very difficult to sinter (for example, ρ < 0.95ρth when it is<br />

sintered at 1700 o C for 10h) even though it is one <strong>of</strong> promising proton conducting oxides with high<br />

proton conductivity and good chemical stability. Such low sinterability <strong>of</strong> BaZrO3 has given many<br />

technical challenges for fuel cell application such as difficulty in thin and dense film formation<br />

under moderate sintering temperature without any chemical interaction with other cell components.<br />

Recently, it was reported that the addition <strong>of</strong> ZnO significantly enhanced the sinterability <strong>of</strong> Ydoped<br />

BaZrO3 (BZY) which might give proper processing accessibility to use it as a solid<br />

electrolyte <strong>of</strong> fuel cells. However it is still not well understood how this sintering aid influences on<br />

the physico-chemical properties <strong>of</strong> BZY. In the present study, we report the effects <strong>of</strong> Zn and Cu<br />

addition on the sinterability and the proton conductivity <strong>of</strong> BZY. The possible sintering mechanism<br />

as well as the defect structural changes <strong>of</strong> BZY due to the addition <strong>of</strong> sintering aids will be also<br />

discussed.<br />

‐ 85 ‐<br />

P 43


Intermediate temperature solid oxide fuel cells based<br />

on a thin BaZr1-xYxO3-δ proton conductor electrolyte<br />

Daniele Pergolesi, Emiliana Fabbri, and Enrico Traversa<br />

International Research Center for Materials Nanoarchitectonics (MANA), National <strong>Institute</strong> for Materials Science<br />

(NIMS), 1-1 Namiki, Tsukuba, Ibaraki, 305-0044 Japan<br />

P 44<br />

High temperature proton conductors have been the subject <strong>of</strong> extensive investigation because <strong>of</strong><br />

their potential application as electrolytes in intermediate temperature solid oxide fuel cells (IT-<br />

SOFCs). Among this class <strong>of</strong> materials, Y-doped BaZrO3 (BZY) exhibit excellent chemical<br />

stability, but BZY total proton conductivity is too low for practical use.<br />

The low conductivity <strong>of</strong> BZY derives from its poor sinterability, which leads to the presence <strong>of</strong> a<br />

large amount <strong>of</strong> resistive grain boundaries. The possibility <strong>of</strong> overcoming low sinterability and high<br />

grain boundary resistance, may lead to the development <strong>of</strong> an excellent electrolyte for IT-SOFCs.<br />

Pulsed laser deposition (PLD) allows growing thin films with high density and an excellent degree<br />

<strong>of</strong> structural quality.<br />

Using PLD, anode-supported solid oxide fuel cells (SOFCs) based on thin BaZr0.8Y0.2O3-δ (BZY)<br />

electrolyte films were fabricated on sintered NiO-BZY composite anodes (Figure 1). After in-situ<br />

reduction <strong>of</strong> NiO to Ni, the anode substrates became porous, while retaining good adhesion with the<br />

electrolyte. A slurry-coated composite cathode<br />

made <strong>of</strong> La0.6Sr0.4Co0.2Fe0.8O3-δ (LSCF) and<br />

cathode<br />

BaCe0.9Yb0.1O3-δ (BCYb), specifically developed<br />

for proton conducting electrolytes, was used to<br />

assemble fuel cell prototypes. Depositing by PLD<br />

100 nm thick LSCF porous films onto the BZY<br />

thin films was essential to improve the<br />

BZY electrolyte<br />

cathode/electrolyte adhesion.<br />

The fuel cell performance <strong>of</strong> the developed<br />

anode<br />

device was examined in the IT range by means <strong>of</strong><br />

I-V<br />

tests and electrochemical impedance<br />

spectroscopy,<br />

showing promising results.<br />

‐ 86 ‐<br />

Fig. 1 Ba deficiency, x, lowers bulk proton<br />

conductivity in yttrium-doped barium zirconates.


Does the Increase in Y-Dopant Concentration Improve the Proton<br />

Conductivity <strong>of</strong> BaZr1-xYxO3-δ Fuel Cell Electrolytes?<br />

Fabbri Emiliana 1 , Daniele Pergolesi 1 , Silvia Licoccia 2 and Enrico Traversa 1<br />

P 45<br />

International Research Center for Materials Nanoarchitectonics (MANA), National <strong>Institute</strong> for Materials Science<br />

(NIMS)1-1 Namiki, Tsukuba, Ibaraki 305-0044 Japan<br />

Department <strong>of</strong> Chemical Science and <strong>Technology</strong>, University <strong>of</strong> Rome Tor Vergata, Via della Ricerca Scientifica, 00133<br />

Rome, Italy<br />

Among high temperature proton conductors, BaZrO3-based oxides <strong>of</strong>fer excellent chemical stability<br />

against reaction with CO2 and H2O. Yttrium acts as an ideal dopant for barium zirconate. Although<br />

its substitution in the B site <strong>of</strong> the perovskite structure causes local lattice distortions, it almost does<br />

not change the acid/base properties <strong>of</strong> the coordinating oxygen. This results in the highest<br />

conductivity among doped barium zirconate oxides [1].<br />

The strategy that we have followed to<br />

improve the electrical properties <strong>of</strong><br />

barium zirconate is the optimization <strong>of</strong><br />

doping conditions, in terms <strong>of</strong> proper<br />

selection <strong>of</strong> the dopant concentration. In<br />

this study we investigated single phase<br />

BaZr1-xYxO3-δ (BZYx) oxides, with x<br />

ranging between 0.2 and 0.5, produced<br />

by means <strong>of</strong> wet-chemical synthesis<br />

procedures. Such high Yttrium dopant<br />

concentrations (x larger than 0.4) have<br />

never been explored before to the best <strong>of</strong><br />

our knowledge. Proton uptake, sintering<br />

behavior, phase stability and electrical<br />

properties <strong>of</strong> BZYx were studied as<br />

function <strong>of</strong> Y concentration, providing a<br />

comprehensive understanding <strong>of</strong> the<br />

effect <strong>of</strong> high dopant concentrations on<br />

the<br />

properties <strong>of</strong> Y-doped barium<br />

zirconate<br />

proton conductor electrolyte.<br />

1.<br />

K.D. Kreuer, Annu. Rev. Mater. Res. 33 (2003), 333.<br />

Fig. 1 Proton concentration and electrical conductivity as a<br />

function <strong>of</strong> Y concentration for BaZr1‐xYxO3‐δ<br />

‐ 87 ‐


Computational Modeling <strong>of</strong> Transport Phenomena in Polymer Electrolyte<br />

Membranes, Nafion and Hydrocarbon Membrane<br />

Yoong-Kee Choe<br />

National <strong>Institute</strong> <strong>of</strong> Advanced Industrial Science & <strong>Technology</strong> (AIST), Umezono 1-1-1, Tsukuba, Japan<br />

P 46<br />

Results <strong>of</strong> the first-principles molecular dynamics (MD) simulations carried out to investigate<br />

various transport properties in polymer electrolyte<br />

membranes (PEMs) are presented. As is well known,<br />

polymer electrolyte membrane fuel cells have attracted a<br />

lot <strong>of</strong> interest because <strong>of</strong> their potential application for<br />

automobile and mobile devices. Since transport properties<br />

arising from proton and water diffusion affects the<br />

performance <strong>of</strong> fuel cells, full comprehension <strong>of</strong> chemistry<br />

and physics <strong>of</strong> these phenomena at an atomistic level is<br />

necessary. To probe the nature <strong>of</strong> the transport properties,<br />

we applied the techniques <strong>of</strong> the first-principles MD to<br />

Nafion, a representative perfluorosulfonic acid based<br />

membrane, and other hydrocarbon membranes. We found<br />

that difference in the value <strong>of</strong> proton diffusion coefficient<br />

with respect to water content inside PEMs is related to the<br />

different character <strong>of</strong> proton hopping occurring in water<br />

hydrogen bond network. Moreover, the results <strong>of</strong> the<br />

first‐principles simulations allow us to explain the<br />

relationship between the proton dynamics and electro‐<br />

osmosis (co‐transport <strong>of</strong> water as a result <strong>of</strong> proton<br />

conduction) at an atomistic level. Our first‐principles<br />

MD study also shows that a hydrophillic functional<br />

group found in various hydrocarbon membranes makes<br />

a negative influence on the proton transport under low<br />

hydration condition. The results <strong>of</strong> the simulation<br />

indicate that such a functional group hinders effective<br />

Fig. 1 Snapshots <strong>of</strong> the first-principles<br />

hydration<br />

around the sulfonic group in PEM and show tsimulation ha t it causes <strong>of</strong> Nafion Grötthuss with transport high water <strong>of</strong><br />

proton<br />

to be ineffective.<br />

content<br />

1. Y. -K Choe, E. Tsuchida, T. Ikeshoji, S. Yamakawa and S. Hyodo, Journal <strong>of</strong> Physical Chemistry (B) 112 (2008),<br />

1<strong>15</strong>86.<br />

2. Y. -K Choe, E. Tsuchida, T. Ikeshoji, S. Yamakawa and S. Hyodo, Physical Chemistry Chemical Physics 11<br />

(2009), 3892.<br />

3.<br />

Y. -K Choe, E. Tsuchida, T. Ikeshoji, A. Ohira and K. Kidena, Journal <strong>of</strong> Physical Chemistry (B) 114 (2010), 2411.<br />

‐ 88 ‐


Study <strong>of</strong> Proton Transport using Reactive Molecular Dynamics<br />

Myvizhi Esai Selvan, David J. Keffer, Shengting Cui, Stephen J. Paddison<br />

Department <strong>of</strong> Chemical and Biomolecular Engineering, University <strong>of</strong> Tennessee, 327 Dougherty Engineering,<br />

Knoxville, TN 37996, USA<br />

We have developed a new reactive molecular dynamics algorithm (RMD) to model the structural<br />

diffusion <strong>of</strong> proton as a chemical reaction with input from both quantum mechanical and<br />

macroscopic models. Structural diffusion is incorporated in a classical MD simulation via three<br />

steps: (i) trigger satisfaction; (ii) instantaneous reaction; and (iii) local equilibration. We have<br />

applied it to study proton transport in several systems including: bulk water, aqueous HCl, and<br />

hydrated polymer electrolyte membranes (PEMs).<br />

Proton transport in hydrated PEMs is influenced by the confinement into regions or domains that are<br />

only a few nanometers in dimension and by the acidity <strong>of</strong> the sulfonic acid groups. The independent<br />

effect <strong>of</strong> acidity and confinement on proton transport has been examined by studying the aqueous<br />

HCl systems (0.22 – 0.83 M) and water filled carbon nanotubes (radius from 5.42 – 10.85 Å)<br />

respectively. The presence <strong>of</strong> the chloride ion in aqueous HCl solutions disrupts the environment <strong>of</strong><br />

a proton both structurally and energetically leading to a reduced probability for proton transfer. We<br />

determined that structural diffusion <strong>of</strong> a proton decreases with an increase in HCl concentration.<br />

When the algorithm was implemented to study the proton diffusion through water filled carbon<br />

nanotubes, the model showed that enhanced confinement drastically reduces structural diffusion by<br />

disrupting the energetic ‘balance’ around the Zundel (H5O2 + ) cation.<br />

Structural diffusion <strong>of</strong> a proton in a hydrated PEM can be categorized into three chemical reactions<br />

based on the degree <strong>of</strong> hydration <strong>of</strong> the membrane and location <strong>of</strong> the hydrated protons. Proton<br />

transport along the center <strong>of</strong> the aqueous channel is similar to that <strong>of</strong> bulk water (in terms <strong>of</strong> the<br />

transition state) while it is different at the hydrophobic/hydrophilic interfacial regions as observed at<br />

low hydration levels. We are currently trying to understand the individual contribution <strong>of</strong> these<br />

reactions to the total charge diffusion and the correlation between structural and vehicular<br />

components.<br />

‐ 89 ‐<br />

P 47


Hydrogen transport in LaNbO4-LaNb3O9 composites<br />

Wen Xing 1 , Guttorm E. Syvertsen 2 , and Reidar Haugsrud 1<br />

1 Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, FERMiO, Gaustadalleen 21, NO-0349 Oslo, Norway<br />

2 NTNU N-7491 Trondheim, Norway<br />

P 48<br />

Mixed protonic and electronic conductors have potentials for serving as membranes in hydrogen<br />

separation technology. However, state-<strong>of</strong>-the-art materials possessing these functional properties<br />

suffer from low chemical stability under operating conditions. The electrical properties <strong>of</strong> some <strong>of</strong><br />

the phases in the La2O3-Nb2O5 system have been subject to intensive characterization. During these<br />

studies, it has been realized that small amounts <strong>of</strong> excess Nb in the proton conductor LaNbO4 leads<br />

to segregation <strong>of</strong> the n-type electronic conductor LaNb3O9 along the grain boundaries. Combination<br />

<strong>of</strong> the transport properties <strong>of</strong> these two materials may yield an interesting behavior with respect to<br />

the ambipolar proton-electron conductivities, and as such be a potential candidate system for<br />

hydrogen gas separation membranes.<br />

Disc specimens <strong>of</strong> LaNbO4-LaNb3O9 composites (10 and<br />

30 vol% LaNb3O9) were fabricated by spark plasma<br />

sintering. The composition and microstructure have been<br />

characterized by SEM-EDS and HRTEM. The electrical<br />

conductivity and the hydrogen flux were characterized as<br />

a function <strong>of</strong> the temperature, oxygen and water vapor<br />

partial pressures. H/D isotope exchange was utilized to<br />

verify transport <strong>of</strong> protons.<br />

Figure 1 shows that the hydrogen flux varies with the<br />

feed side hydrogen pressure with a (pH2) 1/2 dependency.<br />

The temperature dependency <strong>of</strong> the flux density<br />

corresponds to an activation energy <strong>of</strong> ~140 kJ/mol,<br />

similarly to the activation energy for electronic<br />

conduction in LaNbO4 phase, indicating that the hydrogen<br />

permeation might be limited by electron transport at<br />

measurement conditions. Moreover, the amount <strong>of</strong><br />

hydrogen in the permeate increases when changing<br />

permeate gas from dry to wet Ar indicating that also<br />

water splitting and oxide ion conductivity are at play.<br />

Ln(J ,�mol/cm H2 2 Ln(J ,�mol/cm s)<br />

H2 2 s)<br />

-8.5 1000°C<br />

960°C<br />

940°C<br />

-9.0 900°C<br />

-9.5<br />

-10.0<br />

-10.5<br />

-5 -4 -3 -2 -1 0<br />

Ln(p H Ln(p 2<br />

feed H2 feed<br />

, atm)<br />

1/2<br />

Figure 1. Hydrogen flux as a function <strong>of</strong> the<br />

hydrogen partial pressure at the feed side<br />

In this contribution we will discuss the results from these measurements in view <strong>of</strong> ambipolar<br />

transport theory to compare the measured flux values with estimates made from the conductivity<br />

data <strong>of</strong> the composites and the individual material.<br />

This work was initiated through the Big CO2 consortium, GES acknowledges funding through the<br />

NanoPCFC project financed by the Research Council <strong>of</strong> Norway (182090/S10)<br />

‐ 90 ‐


Poly(p-phenylene sulfone)s with High Ion Exchange Capacity: Ionomers with<br />

Unique Microstructural and Transport Features<br />

C. C. de Araujo 1 , K. D. Kreuer 1 , M. Schuster 2 , G. Portale 3 , and J. Maier 1<br />

1. Max-Planck-Institut für Festkörperforschung, Heisenbergstraβe 1, D-70590 Stuttgart, Germany<br />

2. FuMATech, Am Grubenstollen 11, D-66386 St. Ingbert, Germany<br />

2.DUBBLE, BM26 at ESRF, 6 rue Jules Horowitz, BP220, F- 38043 Grenoble, Cedex, France<br />

c.araujo@fkf.mpg.de<br />

Current attempts to increase the operation temperature <strong>of</strong> PEM fuel cells are also driving the<br />

development <strong>of</strong> membranes with high proton conductivity and stability at high temperature and low<br />

humidification [1,2]. Poly(phenylene) ionomers which contain merely sulfone units (-SO2-)<br />

connecting the phenyl rings and in which each phenyl ring is sulfonated (-SO3H) have been<br />

showing to be an important advance in this direction [3-5]. The high degree <strong>of</strong> sulfonation <strong>of</strong> this<br />

polymer leads to the development <strong>of</strong> a microstructure characterized by very narrow hydrated,<br />

hydrophilic domains which are well connected on longer scales. These features together with high<br />

absolute water uptakes at given relative humidities and the high charge carrier concentration<br />

corresponding to the high ion exchange capacity (IEC ~ 4.5 meq./g) result in very high proton<br />

conductivities but also low water transport coefficients (water diffusion and presumably also<br />

electroosmotic drag and permeation). Compared to the transport properties <strong>of</strong> Nafion, these trends<br />

increase with increasing water content and with increasing temperature. For a relative humidity <strong>of</strong><br />

RH = 30 % and a temperature <strong>of</strong> T = 135 °C, the proton conductivity is found to be seven times<br />

higher than the conductivity <strong>of</strong> Nafion under the same conditions.<br />

Understanding the properties <strong>of</strong> this highly sulfonated polymer with very high ion-exchange<br />

capacity is <strong>of</strong> great interest considering that this is a potential material for applications in block<br />

copolymer and polymer blend synthesis. A complete transport study in this new class <strong>of</strong> sulfonated<br />

polymer will be presented including: proton conductivity measured by ac impedance spectroscopy,<br />

diffusion coefficients and electroosmotic drag determined by PFG-NMR and Electrophoretic-NMR,<br />

respectively, and microstructure studies by SAXS.<br />

1. K.-D. Kreuer, S. J. Paddison, E. Spohr, and M. Schuster, Chem. Rev. 104, 4637–4678 (2004)<br />

2. K. D. Kreuer, M. Schuster, B. Obliers, O. Diat, U. Traub, A. Fuchs, U. Klock, S. J. Paddison, and J. Maier, J. Power<br />

Sources 178, 499–509 (2008).<br />

3. M. Schuster, K.-D. Kreuer, H. T. Andersen, and J. Maier, Macromolecules 40, 598–607 (2007).<br />

4. M. Schuster, C. C. de Araujo, V. Atanasov, H. T. Andersen, K.-D. Kreuer, and J. Maier, Macromolecules 42, 3129–<br />

3137 (2009).<br />

5. C. C. de Araujo, K. D. Kreuer, M. Schuster, G. Portale, H. Mendil-Jakani, G. Gebel, and J. Maier, Phys. Chem. Chem.<br />

Phys. 11, 3305–3312 (2009).<br />

‐ 91 ‐<br />

P 49


Rule <strong>of</strong> superprotonic phase transition in CsxRb1-xH2PO4<br />

Yasumitsu Matsuo 1) , Junko Hatori 1) , Yukihiko Yoshida 2) and Seiichiro Ikehata 2)<br />

1) Faculty <strong>of</strong> Science and Engineering, Setsunan University, 17-8 Ikeda-Nakamachi, Neyagawa, Osaka 572-8508, Japan<br />

2) Faculty <strong>of</strong> Science, Tokyo University <strong>of</strong> Science, 1-3 Kagurazaka, Shinjyuku-ku, Tokyo 162-8601, Japan<br />

‐ 92 ‐<br />

P 50<br />

It is well known that the hydrogen-bonded material, CsH2PO4 (CDP), undergoes a superprotonic<br />

phase transition from the low-temperature ferroelastic phase to the high-temperature paraelastic<br />

phase at around 501 K. It is also known that the mixed crystals CsxRb1-xH2PO4 (x=0~1) also exhibit<br />

the superprotonic conductivity. In these mixed crystals, the temperature <strong>of</strong> superprotonic phase<br />

transition increases by the increase <strong>of</strong> x. Furthermore, we have indicated that the precursor effect in<br />

proton conductivity in CDP is closely related to the component d <strong>of</strong> a spontaneous strain tensor,<br />

which is calculated from the following equation using the lattice constants at ferroelastic phase.<br />

1<br />

tan( / 2 2arctan(<br />

m /<br />

2<br />

a b<br />

d � � �<br />

In the present work, we have prepared with the mixed crystals and investigated the rule <strong>of</strong> the<br />

superprotonic phase transition in the CDP-type compounds.<br />

Figure 1 shows the relation between the superprotonic phase transition temperature Tc and the<br />

component d. It is evident that Tc linearly increases with the decrease <strong>of</strong> the component d 2 . This<br />

result indicates that the strain is an important<br />

factor for the appearance <strong>of</strong> superprotonic<br />

conductivity in the CDP-type compounds. In<br />

addition, by the thermodynamic analyses, it was<br />

found that the relation between the strain<br />

component d and transition temperature Tc is<br />

described with the terms <strong>of</strong> the changes in the<br />

proton kinetic energy ΔK and entropy ΔS by the<br />

superprotonic phase transition as follows;<br />

1<br />

�K<br />

�Tc<br />

�S<br />

� kd<br />

2<br />

(2)<br />

From these results, it is deduced that the<br />

transition temperature Tc in CDP-type compounds<br />

is determined by the interrelation between the<br />

increase in the kinetic energy <strong>of</strong> proton, the<br />

2<br />

d 2<br />

1.160<br />

1.<strong>15</strong>0<br />

1.140<br />

m<br />

))<br />

×10 –2<br />

(1)<br />

x = 1<br />

x = 0.9 x = 0.8<br />

x = 0.75<br />

x = 0.6<br />

x = 0.5<br />

x = 0.4<br />

220 240 260<br />

T s ( o C)<br />

Cs xRb 1–xH 2PO 4<br />

x = 0.2<br />

x = 0.25<br />

increase in entropy and the release <strong>of</strong> strain energy. Fig. 1 Relation between Tc and d 2<br />

i<br />

S. M. Haile Actra Materialia 51 (2003) 5981<br />

ii<br />

H. Matsumoto, T. Shimura, T. Higuchi, T. Otake, Y. Sasaki, K. Yashiro, A. Kaimai, T. Kawada, J. Mizusaki,<br />

Electrochemistry, 72 (2004) 861<br />

iii<br />

H. Matsumoto, T. Shimura, T. Higuchi, H. Tanaka, K. Katahira, T. Otake, T. Kudo, K. Yashiro, A. Kaimai, T.<br />

Kawada, J. Mizusaki, Journal <strong>of</strong> The <strong>15</strong>2 (2005) A488<br />

iv<br />

S. Mimuro, S. Shibako, Y. Oyama, K. Kobayashi, T. Higuchi, S. Shin, S. Yamaguchi, Solid State Ionics, 178 (2007)<br />

647


Hydration and protonic conductivity in LaAsO4 based ceramics<br />

Tor S. Bjørheim, Truls Norby and Reidar Haugsrud<br />

Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, FERMiO, Gaustadallèen 21, NO-0349 Oslo, Norway<br />

Significant proton conduction has been<br />

reported for a variety <strong>of</strong> rare-earth orthophosphates<br />

[1-3], -vanadates [4], -niobates<br />

[5] and -tantalates [6]. A natural<br />

continuation along similar lines <strong>of</strong><br />

compounds is the rare-earth ortho-arsenates,<br />

LnAsO4.<br />

Incorporation and transport <strong>of</strong> protonic<br />

defects has been studied in nominally<br />

undoped and 1 and 3 mol% Sr-doped<br />

LaAsO4 ceramics. The AC impedance <strong>of</strong> the<br />

materials was measured as a function <strong>of</strong><br />

temperature (1<strong>15</strong>0 - 400 °C), (1 - 1·10 -<br />

5<br />

atm) and (0.025 – 3·10 -5 p H 2O<br />

atm).<br />

p 2<br />

H O<br />

The bulk conductivity decreases with<br />

decreasing p H within the whole<br />

2O<br />

temperature range, and with decreasing pO2<br />

at the highest temperatures. A significant<br />

isotope effect shows that protons contribute<br />

to the total conductivity at all temperatures<br />

under wet conditions and completely<br />

dominate at temperatures below ~850 °C.<br />

The remaining contributions are attributed to<br />

oxide ions, and also electron holes at the<br />

highest temperatures (see Fig. 1).<br />

Based on the results, the defect structure <strong>of</strong><br />

Sr-doped LaAsO4 was concluded to be<br />

dominated by oxide ion vacancies in the<br />

4-<br />

form <strong>of</strong> pyro-arsenate ions, As2O7<br />

, in dry<br />

atmospheres at high temperatures, and by<br />

protonic defects in the form <strong>of</strong> hydrogen<br />

-<br />

arsenate groups, HAsO 4 , in wet<br />

atmospheres. Hydration thermodynamics<br />

extracted by modeling <strong>of</strong> the conductivity<br />

data are comparable to those determined for<br />

other LnXO4 (X=P,V,Nb) compounds.<br />

- 93 -<br />

P 51<br />

Fig. 1: Measured total and modeled partial<br />

conductivities <strong>of</strong> 3 mol% Sr-doped LaAsO4<br />

Activation energies for the mobility <strong>of</strong><br />

protons are slightly lower than those<br />

determined for LaPO4. However, the partial<br />

proton conductivities <strong>of</strong> 4·10 -6 to 5·10 -5<br />

S/cm between 500 and 925 °C for 3 mol%<br />

Sr-doped LaAsO4 are lower than those<br />

determined for acceptor doped LaPO4.<br />

The<br />

authors gratefully acknowledge the Norwegian<br />

Chemical<br />

Society for financial support<br />

1. T. Norby and N. Christiansen, Solid State<br />

Ionics 77 (1995), p. 240.<br />

2. K. Amezawa, Y. Tomii and N. Yamamoto,<br />

Solid State Ionics 176 (2005) (1-2), p. 135.<br />

3. K. Amezawa, H. Maekawa, Y. Tomii and N.<br />

Yamamoto, Solid State Ionics 145 (2001) (1-<br />

4), p. 233.<br />

4. M. Huse, T. Norby and R. Haugsrud, in<br />

preparation, 2010.<br />

5. R. Haugsrud and T. Norby, Nat. Mater. 5<br />

(2006) (3), p. 193.<br />

6.<br />

R. Haugsrud and T. Norby, J. Am. Ceram.<br />

Soc. 90 (2007) (4), p. 1116.


Investigation <strong>of</strong> hydrogen permeation in mixed conductor LaWOX<br />

(La/W = 5.6)<br />

Skjalg Erdal and Reidar Haugsrud<br />

Department <strong>of</strong> Chemistry, University <strong>of</strong> Oslo, Centre for Material Science and Nanotechnology, FERMiO,<br />

Gaustadallèen 21, NO-0349 Oslo, Norway<br />

To date, no known material combines<br />

sufficiently high ambipolar proton-electron<br />

conductivity with chemical and thermal<br />

stability to be used as H2 separation<br />

membranes under typical conditions for<br />

methane steam reforming or coal gasification.<br />

In an effort to uncover materials with sufficient<br />

hydrogen uptake and transport to possibly<br />

serve in hydrogen separation membranes, the<br />

hydrogen permeation through mixed conductor<br />

lanthanum tungstate (LWO) has been<br />

investigated.<br />

LWO was synthesized via a wet chemical<br />

gelling and combustion method. The nominal<br />

La to W ratio (5.6/1) was chosen to ensure we<br />

stay within the compositional window <strong>of</strong> single<br />

phase LWO, as identified by Magrasó et. al<br />

[1].<br />

Figure 1a presents the hydrogen permeativity<br />

in LWO as a function <strong>of</strong> temperature, from<br />

1000 to 700 °C, for various feeds. The<br />

hydrogen flux increases with increasing<br />

temperature and feed side pH2 in the entire<br />

temperature range. Above about 850 °C, the<br />

flux is higher for a feed consisting <strong>of</strong> wet<br />

H2/He than for dry H2/He, but this picture is<br />

reversed as we go below 850 °C, presumably<br />

due to the more reducing conditions needed to<br />

maintain the n-type conductivity at lower<br />

temperatures. The results further support<br />

earlier speculations that the lanthanide<br />

tungstates (in [2]: Er, Gd) have the ability to<br />

incorporate protons through a partial reduction<br />

<strong>of</strong> the oxide in contact with hydrogen gas [2].<br />

Acknowledgements: This work was supported by the<br />

FRINAT project 171<strong>15</strong>7/V30 “Hydrogen in oxides<br />

(HYDROX)” <strong>of</strong> the Research Council <strong>of</strong> Norway.<br />

Permeativity H 2 (mL/min cm)<br />

10 -3<br />

10 -4<br />

10 -5<br />

0.0008 0.0009 0.0010<br />

1/T (K -1 Feed side:<br />

Wet H2 Dry H /He 2<br />

Wet H /He 2<br />

a)<br />

)<br />

Permeativity H 2 (mL/min cm)<br />

10 -3<br />

10 -4<br />

10 -5<br />

10 -3<br />

1000 °C<br />

900 °C<br />

800 °C<br />

10 -2<br />

10 -1<br />

�pH2 (atm)<br />

P 52<br />

10 0<br />

b)<br />

Fig. 1 a) Hydrogen permeativity in lanthanum tungstate as<br />

a function <strong>of</strong> inverse absolute temperature. Three different<br />

feed side gas compositions are shown. b) Double<br />

logarithmic plot <strong>of</strong> hydrogen permeativity in lanthanum<br />

tungstate as a function <strong>of</strong> hydrogen partial pressure<br />

gradient over membrane. Conducted at 1000, 900 and 800<br />

°C. Dry Ar supplied to permeate side in both graphs.<br />

Figure 1b shows the H2 permeativity as a<br />

function <strong>of</strong> the difference in hydrogen partial<br />

pressure over the membrane, at 1000-800 °C.<br />

For a given temperature, the hydrogen<br />

permeativity increases with increasing ΔpH2,<br />

and this dependency is steeper towards the lower<br />

partial pressure gradients.<br />

Effects <strong>of</strong> water vapor in the permeate side Ar<br />

have also been determined. At T > 800 °C, the<br />

detected hydrogen concentration in the permeate<br />

increases with increasing pH2O on the permeate<br />

side. This can be understood in a context <strong>of</strong><br />

water splitting as a consequence <strong>of</strong> oxygen ion<br />

conductivity - being negligible below 800 °C,<br />

and rapidly increasing with increasing<br />

temperature.<br />

1. Magrasó, A.; Frontera, C.; Marrero-López,<br />

D.;<br />

Núñez,<br />

P. Dalton Transactions, 2009.<br />

2. R. Haugsrud Solid State Ionics, 178 (2007) 555–<br />

560.<br />

-94


Fabrication <strong>of</strong> Ca-doped Lanthanum Niobate electrolyte film by<br />

Electrophoretic Deposition for PCFC applications<br />

Francesco Bozza and Nikolaos Bonanos<br />

Fuel Cells and Solid State Chemistry Division, Risø National Laboratory for Sustainable Energy, Technical University<br />

<strong>of</strong> Denmark, P.O. Box 49, 4000 Roskilde, Denmark<br />

Electrophoretic Deposition (EPD) has been applied to the preparation<br />

<strong>of</strong> a dense calcium-doped lanthanum niobate electrolyte<br />

film. The EPD technique consists <strong>of</strong> charging the particles to be<br />

deposited by a suitable suspending medium. When an electric<br />

field is applied to the suspension, the particles migrate to one <strong>of</strong><br />

the electrodes and deposit [1, 2]. Powder <strong>of</strong> La0.995Ca0.005NbO4,<br />

a high temperature proton conductor [3], was suspended in a<br />

solution <strong>of</strong> acetylacetone and iodine. The effects <strong>of</strong> suspension<br />

composition and deposition conditions were analyzed in order to<br />

identify a suitable set <strong>of</strong> EPD process parameters. The powders<br />

were deposited on a composite substrate <strong>of</strong> La0.995Ca0.005NbO4,<br />

NiO, binder and graphite. With this technique, a dense 8 µm film<br />

<strong>of</strong> lanthanum niobate supported on a porous La0.995Ca0.005NbO4<br />

/NiO substrate was obtained after sintering at 1200°C (Figure 1).<br />

Moreover the technique was found to be effective for a mixture<br />

<strong>of</strong> NiO and La0.995Ca0.005NbO4 which, after sintering, would<br />

form a supporter <strong>of</strong> La0.995Ca0.005NbO4/NiO anode functional<br />

layer. The morphology and gas tightness <strong>of</strong> the film are<br />

investigated.<br />

This work has been funded by the EU within the FP7 project Efficient<br />

and robust fuel cell with novel ceramic proton conducting electrolyte<br />

(EFFIPRO), grant agreement 227560.<br />

1. P. Sarkar, P. S. Nicholson, J. Am. Ceram. Soc. 79 (1996), 1987.<br />

2. F. Bozza, R. Polini, E. Traversa, J. Am. Ceram. Soc. 92 (2009), 1999.<br />

3. R. Haugsrud, T. Norby, Nature Materials 5 (2006), 193.<br />

- 95 -<br />

P 53<br />

Fig. 1 Cross sectional SEM <strong>of</strong> a dense<br />

layer <strong>of</strong> La0.995Ca0.005NbO4 deposited on<br />

a porous substrate <strong>of</strong> La0.995Ca0.005NbO4/<br />

NiO and sintered in air at 1200°C.


Ab initio Molecular Dynamics study <strong>of</strong> proton dynamics and mobility in<br />

phosphoric acid<br />

Linas Vilciauskas a , Gabriel Bester a , Klaus-Dieter Kreuer a , Mark E. Tuckerman b<br />

and Stephen J. Paddison c<br />

a Max-Planck-Institut fü̈ r Festkö̈ rperforschung, Heisenbergstraße 1, D-70569 Stuttgart, Germany<br />

b Department <strong>of</strong> Chemistry and Courant <strong>Institute</strong> <strong>of</strong> Mathematical Sciences, New York University, New York, New York<br />

10003, USA<br />

c Department <strong>of</strong> Chemical and Biomolecular Engineering, University <strong>of</strong> Tennessee, Knoxville, Tennessee 37996, USA<br />

P 54<br />

Proton transport is one <strong>of</strong> the most fundamental and crucial processes for all living creatures as<br />

well as in a multitude <strong>of</strong> phenomena in chemistry and physics such as the function <strong>of</strong> many devices<br />

like proton exchange membrane fuel cells [1]. Phosphoric acid above its melting point is probably<br />

the best intrinsic proton conductor, having exceptionally high proton conductivity due to a high<br />

degree <strong>of</strong> self-dissociation (~7 %) and, despite its extremely high viscosity (~120 cP), proton<br />

mobility only comparable to that <strong>of</strong> aqueous systems. The development <strong>of</strong> new water-free PEM fuel<br />

cell materials calls for a fundamental understanding <strong>of</strong> these systems. It is clear that the proton<br />

conduction mechanism is dominated by the structure diffusion however the exact details on how<br />

many molecules [2] are involved in the elementary processes as well as what is its rate limiting step<br />

must still be elucidated. The limited availability <strong>of</strong> experimental tools capable <strong>of</strong> probing the<br />

corresponding time and length scales makes the molecular modeling techniques indispensable. The<br />

ab initio molecular dynamics (AIMD) techniques were exceptionally successful in studying the<br />

proton transport processes in diverse media. We performed a set <strong>of</strong> Car-Parrinello 3 molecular<br />

dynamics simulations, employing a popular BLYP exchange-correlation functional and norm<br />

conserving Troullier-Martins pseudopotentials to treat the core electrons. The kinetic energy cut-<strong>of</strong>f<br />

<strong>of</strong> 100 Ry was used to expand the electronic wavefunctions around the �-point and the time step <strong>of</strong><br />

0.1 a.u. was used to propagate the dynamics. The calculated radial distribution functions and<br />

diffusion constants show a very good agreement with the available experimental data [3,4]. The<br />

D H /D P ratio is ~5 showing that the intermolecular proton transfer is dominating over the vehicular<br />

transport. The rotational dynamics <strong>of</strong> the phosphate groups in H3PO4 is clearly much more<br />

constrained that in the case <strong>of</strong> similar systems with fast proton transport such as CsHSO4, CsH2PO4<br />

and probably H3PO3 as well. Most <strong>of</strong> this effect might be attributed to the high concentration <strong>of</strong><br />

protons and the ‘frustrated’ H-bond network (3 proton donors vs. just 1 proton acceptor), but<br />

interestingly, this doesn't hinder the proton dynamics clearly showing that the full rotational<br />

reorientation or ‘tumbling’ <strong>of</strong> the PO4 group is not necessary for the successful proton transfer (PT)<br />

event. In addition, the strong presence <strong>of</strong> Grotthuss-like PT events as evident from the simulation<br />

indicates a strong Coulombic coupling between the protons over the O-P-O bonds, which is also<br />

accelerating the overall dynamics. The calculated PT free energy barrier is just 1.8 kcal/mol,<br />

confirming the fact that PT is completely governed by the solvent dynamics. The protons show a<br />

fast hopping between the donor and acceptor on a time scale <strong>of</strong> 440 fs and only a little angular<br />

rotation on a much slower time scale <strong>of</strong> 3.6 ps is necessary for the successful event. Moreover, there<br />

are still a lot <strong>of</strong> very interesting open questions left to answer when dealing with this and similar<br />

systems<br />

such as what is the role <strong>of</strong> equilibrium water, which is even present in the nominally dry<br />

acid<br />

or how is everything affected by changing the ratio <strong>of</strong> proton donor/ acceptor sites (e.g. H3PO3).<br />

1. Kreuer, K.D.; Paddison, S.J.; Spohr, E.; Shuster, M.; Chem. Rev. 2004, 104, 4637.<br />

2. Vilciauskas, L.; Paddison, S. J.; Kreuer, K. D.; J. Phys. Chem. A 2009, 113, 9193.<br />

3. Car, R.; Parrinello, M.; Phys. Rev. Lett. 1985, 55, 2471.<br />

4. Dippel, Th.; Kreuer, K.D.; Lassègues, J.C.; Rodriguez, D.; Solid State Ionics 1993, 61, 41.<br />

5. Aihara, Y.; Sonai, A.; Hattori, M.; Hayamizu, J. Phys Chem. B 2006, 110, 24999.<br />

- 96 -


CsH2PO4 nanoparticle synthesis via electrohydrodynamic atomization –<br />

aerosol size measurements<br />

Áron Varga † , Andrew J. Downard ‡ , Hyung Wan Do § , Richard C. Flagan ‡ , Sossina M. Haile †‡<br />

† Department <strong>of</strong> Materials Science, ‡ Department <strong>of</strong> Chemical Engineering, § Department <strong>of</strong> Electrical Engineering,<br />

<strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, 1200 E. <strong>California</strong> Blvd., Pasadena, CA 91125, USA<br />

P 55<br />

It has been shown that solid acid fuel cells based on CsH2PO4 (CDP) are performance limited by the<br />

oxygen electroreduction reaction. 1 Increasing the length <strong>of</strong> the electrocatalytically active triple<br />

phase boundaries by creating a 3-dimensional, interconnected, porous structure <strong>of</strong> CDP and Pt<br />

nanoparticles is thought to improve the utilization <strong>of</strong> Pt catalyst and hence increase fuel cell<br />

performance. 2 Electrohydrodynamic atomization (EHA), a process where an aqueous solution <strong>of</strong><br />

CDP is sprayed under high voltage, has emerged as a suitable method <strong>of</strong> generating CDP<br />

nanoparticles and directly depositing them onto the substrate <strong>of</strong> choice. 3<br />

The EHA method has an extremely large parameter space, which includes solution properties such<br />

as conductivity and surface tension, spray parameters, such as voltage, flow rate, and geometrical<br />

aspects. To optimize the process relative<br />

to these parameters, an aerosol particle<br />

sizing system, consisting <strong>of</strong> a 210 Po<br />

neutralizer, a tubular differential mobility<br />

analyzer (TDMA), and a condensation<br />

particle counter (CPC) has been<br />

constructed. The highly charged,<br />

polydisperse aerosol generated by EHA is<br />

introduced into the neutralizer, reducing<br />

the charge per particle to a uniform value.<br />

Based on the electrophoretic mobility <strong>of</strong><br />

the particles, the TDMA selects a narrow<br />

size range to pass through to the CPC,<br />

where the concentration is counted. Fine-<br />

tuning the parameters has lead to the<br />

generation <strong>of</strong> 41 nm diameter CDP<br />

particles – a value approached that <strong>of</strong> the<br />

Figure 1 CsH2PO4 (CDP) particle size distribution as generated<br />

via electrohydrodynamic atomization and measured with a<br />

differential mobility analyzer<br />

Pt catalyst particles. The result holds promise for increases in Pt utilization and hence solid acid fuel<br />

cell performance.<br />

1 S. M. Haile, C. R. I. Chisholm, K. Sasaki, D. A. Boysen and T. Uda, Faraday Discussions (2007) 134, 17-39.<br />

2 C. R. I Chisholm, D. A Boysen, A. B. Papandrew, S. Zecevic, S. Cha, K. Sasaki, A. Varga, K. P. Giapis and S. M.<br />

Haile, Interface Magazine (2009) 18, 53-59.<br />

3 A. Varga, N. A. Brunelli, K. P. Giapis, S. M. Haile, Journal <strong>of</strong> Materials Chemistry (2010) 20, 6309 – 63<strong>15</strong>.<br />

-97-


Cobalt and yttrium doped barium zirconates as mixed conducting<br />

cathodes for SOFC<br />

Björn Björnsson, Juan C. Lucio-Vega, Yoshihiro Yamazaki, and Sossina M. Haile<br />

Materials Science, <strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, 1200 E <strong>California</strong> Blvd., Pasadena, CA 91125, USA<br />

P 56<br />

Mixed ionic and electronic conductors have been shown to be highly effective as cathode materials<br />

in solid oxide fuel cells based on oxide-ion conducting electrolytes [1, 2]. The mixed transport<br />

relaxes the restriction that electrochemical reactions occur at the triple-phase boundary between the<br />

gas phase, the electrolyte, and the cathode electrocatalyst. Instead, reactions over the entirety <strong>of</strong> the<br />

cathode electrocatalyst surface can be supported. By analogy, it is anticipated that mixed proton–<br />

electronic hole conductors will be suitable for the oxygen electroreduction reaction [3].<br />

over the entire oxide-gas interface. As the first step to test such<br />

a hypothesis, we introduced cobalt into 20 at% yttrium-doped barium zirconate, a well-known<br />

proton conducting oxide. Cobalt addition has been found to enhance electronic hole conduction<br />

under oxidizing atmospheres and at temperatures relevant to intermediate temperature solid oxide<br />

fuel cells (400-500 °C) [4]. Single phase-perovskite phase is observed here in BaZr0.8-xCoxY0.2O3-δ<br />

up to x = 0.3 (at 1300 °C), for materials prepared by chemical solution methods. The results <strong>of</strong><br />

transport measurements and thermogravimetric analysis are presented.<br />

2H � � 1/2O (g) � 2h 2 � � H O(g) 2<br />

1. S.B. Adler, Chemical Reviews 104 (2004) (10), p. 4791.<br />

2. Z.P. Shao and S.M. Haile, Nature 431 (2004) (7005), p. 170.<br />

3. R. Peng, T. Wu, W. Liu, X. Liu and G. Meng, Journal <strong>of</strong> Materials Chemistry 20 (2010), p. 6218.<br />

4. M.A. Azimova and S. McIntosh, Solid State Ionics 180 (2009) (2-3), p. 160.<br />

- 98 -


High-temperature phase behavior in the Rb3H(SO4)2-RbHSO4<br />

pseudo-binary system<br />

Chatr Panithipongwut, Sossina M. Haile<br />

Materials Science, <strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, 1200 E <strong>California</strong> Blvd., Pasadena, CA 91125 USA<br />

P 57<br />

Many solid acids in the M3H(XO4)2 family (M = Cs, Rb, NH4, X = S, Se) transform from a<br />

monoclinic phase at room temperature to a superprotonic, trigonal phase at elevated temperatures.<br />

Previous studies <strong>of</strong> the conductivity <strong>of</strong> Rb3H(SO4)2 revealed a sharp increase at 205 °C [1,2], as is<br />

typically observed for polymorphic, superprotonic transitions. In contrast to other M3H(XO4)2<br />

compounds, however, Rb3H(SO4)2 undergoes a disproportionation reaction at this temperature to<br />

Rb2SO4 and an unknown phase with composition intermediate between Rb3H(SO4)2 and RbHSO4<br />

[2]. A comprehensive study using in situ high-temperature X-ray powder diffraction, thermal<br />

analysis (simultaneous thermogravimetric and calorimetric analysis) and electrical impedance<br />

measurements <strong>of</strong> the pseudo-binary system Rb3H(SO4)2 - RbHSO4 has been carried out here to<br />

determine the composition and characteristics <strong>of</strong> the intermediate phase and the overall phase<br />

behavior.<br />

In combination, these studies indicate that the intermediate phase, formerly described as<br />

RbmHn(SO4)p(s) with p = (m+n)/2, is Rb5H3(SO4)4. Solid state mixtures <strong>of</strong> Rb3H(SO4)2 and RbHSO4<br />

with 66.7 % RbHSO4 (corresponding to the stoichiometry <strong>of</strong> Rb5H3(SO4)4) show the formation <strong>of</strong><br />

the new compound at temperatures as low as 130 °C. At 185 °C, Rb5H3(SO4)4 apparently undergoes<br />

a transformation to a higher conductivity phase. Solid state mixtures <strong>of</strong> Rb3H(SO4)2 and RbHSO4<br />

rich in Rb3H(SO4)2 reveal a conductivity anomaly at 185 °C, attributed to the polymorphic<br />

transition <strong>of</strong> Rb5H3(SO4)4 and a second anomaly at 205 °C, corresponding to the disproportionation<br />

<strong>of</strong> the remaining Rb3H(SO4)2, consistent with the previous studies. Mixtures rich in RbHSO4<br />

display an anomaly at 170 °C, tentatively attributed to a polymorphic transition in RbHSO4,<br />

followed by a second anomaly at 185 °C, corresponding to the polymorphic transition in the new<br />

compound. Rb5H3(SO4)4 is not stable at ambient temperatures and eventually disproportionates into<br />

the end-members Rb3H(SO4)2(s) and RbHSO4(s). Neither the high temperature (> 185 °C) nor low<br />

temperature (~ 125 to 185 °C) forms <strong>of</strong> Rb5H3(SO4)4 are isostructural with any known M5H3(XO4)4<br />

compound.<br />

1. A.I. Baranov, V.V. Dolbinina, E.D. Yakushkin, V.Y. Vinnichenko, V.H. Schmidt and S. Lanceros-Mendez,<br />

Ferroelectrics 217 (1998), 285.<br />

2. L.A. Cowan, R.M. Morcos, N. Hatada, A. Navrotsky and S.M. Haile, Solid State Ionics 179 (2008), 305.<br />

-99-


The superprotonic phase transition and hysteresis behavior <strong>of</strong><br />

Cs1-xKxH2PO4<br />

Daniil Kitchaev, Ayako Ikeda, and Sossina M. Haile<br />

Materials Science, <strong>California</strong> <strong>Institute</strong> <strong>of</strong> <strong>Technology</strong>, 1200 E <strong>California</strong> Blvd., Pasadena, CA 91125, USA<br />

CsH2PO4 (CDP) as a fuel cell electrolyte suffers<br />

from a large hysteresis and cell volume change at its<br />

superprotonic transition – it is predicted that by<br />

modifying its lattice parameters via doping these<br />

difficulties can be avoided. CDP was doped with<br />

KH2PO4 (KDP) and the properties <strong>of</strong> the resulting<br />

doped electrolyte were studied over a range <strong>of</strong><br />

compositions. The solubility <strong>of</strong> potassium in CDP<br />

was determined by XRD and EDS measurements<br />

and found to be ~20% near superprotonic<br />

temperatures. The superprotonic transition<br />

temperature and reverse transition <strong>of</strong> CKDP (Cs1xKxH2PO4)<br />

were evaluated in this solubility range<br />

by DSC. It was found that the superprotonic<br />

transition temperature decreases with increased<br />

potassium content. This behavior contrasts that in<br />

the CRDP (Cs1-xRbxH2PO4) system, in which the<br />

superprotonic transition temperature increases with<br />

rubidium content. (Figure 1)[1][2] Furthermore, the<br />

hysteresis <strong>of</strong> the superprotonic transition has been<br />

found to decrease with potassium incorporation<br />

(Figure 2), from 13.6 o C for pure CDP [2], to 9.2 o C<br />

for Cs0.7K0.3H2PO4, yet again deviating from<br />

recorded CRDP behavior.<br />

[1] Y. K. Taninouchi, T. Uda, Y. Awakura, A. Ikeda and S. M. Haile,<br />

“Dehydration Behavior <strong>of</strong> the Superprotonic Conductor CsH 2PO 4 at<br />

Moderate Temperatures: 230 to 260°C,” J. Mater. Chem. 17, 3182-<br />

3189 (2007). (Invited)<br />

[2] Louie, Mary W., Kislitsyn, Mikhail. Bhattacharya, Kaushik.<br />

Haile, Sossina M., “Phase Transformation and Hysteresis Behavior in<br />

Cs 1-xRb xH 2PO 4”. Solid State Ionics 181 (2010) 173-179. Elsevier.<br />

-100-<br />

P 58<br />

Figure 1: Cs1-xKxH2PO4 and Cs1-xRbxH2PO4<br />

superprotonic transition temperatures<br />

Figure 2: Superprotonic phase transition<br />

hysteresis in Cs0.7K0.3H2PO4 observed by DSC<br />

with varying humidification.


Authors Index<br />

Author Program No. …Page<br />

Underline: Presenting Author<br />

[A]<br />

Abe, T. P01...43<br />

Adams, A. P04...46<br />

Adelstein, N. P25...67<br />

Ahmed, I. O09...10 O10...11<br />

P09...51<br />

Akao, Y. O32...33<br />

Amezawa, K. O28K...29 P26...68<br />

P27...69<br />

Anderson, V. O18K...19<br />

Ando, M. P33...75<br />

Argun, A. O02I...3<br />

Asano, K. P31...73<br />

Avila-Paredes, H. P30...72<br />

Azad, A. O09...10 P22...64<br />

[B]<br />

Baranek, P. O13...14<br />

Besikiotis, V. O29...30<br />

Bester, G. P54...96<br />

Bigarré, J. P02...44<br />

Bjørheim, T. O29...30 O33...34<br />

P51...93<br />

Björnsson, B. P56...98<br />

Blanc, F. O05K...6 O40...41<br />

P24...66<br />

Blümich, B. P04...46<br />

Bonanos, N. P13...55 P53...95<br />

Botez, C. O35I...36<br />

Boucher, F. O13...14<br />

Boukamp, B. P13...55<br />

Bozza, F. P53...95<br />

Buannic, L. O05K...6 O40...41<br />

[C]<br />

Cailleteau, C. O16K...17<br />

Chen, C. O11...12<br />

Chen, F. P36...78 P38...80<br />

Chisholm, C. O06I...7<br />

Choe, Y. P46...88<br />

Clark, D. O39...40<br />

Clark, J. P41...83<br />

Cui, S. P47...89<br />

[D]<br />

Daiko, Y. P07...49<br />

Danel, C. O11...12<br />

De Araujo, C. O03...4 P49...91<br />

De Jonghe, L. O24...25 P25...67<br />

De Souza, R. P04...46<br />

Deshpande, R. O18K...19<br />

Dillon, A. O18K...19<br />

Dixon, L. P38...80<br />

Do, H. P55...97<br />

Douhara, H. P05...47<br />

Downard, A. P55...97<br />

[E]<br />

Eikerling, M. O25K...26<br />

Einarsrud, M. P23...65<br />

Emiliana, F. P45...87<br />

Erdal, S. P52...94<br />

Eriksson, S. O09...10 O10...11<br />

P37...79<br />

Esai Selvan, M. P47...89<br />

Esashi, M. P29...71<br />

Eurenius, K. P18...60<br />

[F]<br />

Fabbri, E. P44...86<br />

Feng, J. P25...67<br />

Feng, S. P12...54<br />

Fisher, C. P32...74<br />

Fjeld, H. P03...45<br />

Flagan, R. P55...97<br />

Fontaine, M. O22...23<br />

Frick, B. O16K...17 O34I...35<br />

Fukatsu, N. P05...47<br />

Furukawa, Y. P39...81<br />

[G]<br />

Gebel, G. O16K...17 P02...44<br />

George, S. O18K...19<br />

Gomez, M. P10...52<br />

Grande, T. O22...23 P23...65<br />

Grey, C. O05K...6 O40...41<br />

P24...66<br />

Guillermo, A. O16K...17


[H]<br />

Habenicht, B. O17...18<br />

Hadidi, K. O27...28<br />

Haile, S. O37...38 O38...39<br />

O41...42 P55...97<br />

P56...98 P57...99<br />

P58...100<br />

Hammond, T. O02I...3<br />

Han, D. P20...62<br />

Hatada, N. O30...31<br />

Hatakeyama, A. O07...8<br />

Hatori, J. P50...92<br />

Haugsrud, R. O33...34 O22...23<br />

O29...30 P03...45<br />

P09...51 P14...56<br />

P16...58 P37...79<br />

P40...82 P48...90<br />

P51...93 P52...94<br />

Hayashi, A. P39...81<br />

He, F. P34...76<br />

Heitjans, P. P04...46<br />

Hernandez-Sanchez, R. O38...39<br />

Hightower, A. O37...38<br />

Hinterberg, J. P04...46<br />

Hong, S. P43...85<br />

Hui, R. P11...53<br />

Hurst, K. O18K...19<br />

Huse, M. O33...34 P37...79<br />

[I]<br />

Iguchi, F. O31K...32 P29...71<br />

Ikeda, A. O41...42 P58...100<br />

Ikehata, S. P50...92<br />

Irvine, J. O09...10 P21...63<br />

P22...64<br />

Ishihara, T. O23...24<br />

Ishikawa, A. O36I...37<br />

Islam, M. O12K...13<br />

Ivanova, M. P14...56<br />

[J]<br />

Jankovic, J. P11...53<br />

Ji, H. P42...84 P43...85<br />

Joubert, O. O13...14<br />

[K]<br />

Karlsson, M. O10...11<br />

Kawada, T. O28K...29 P26...68<br />

P27...69<br />

Kawamura, J. O32...33<br />

Kawamura, G. P28...70<br />

Kawasaki, Y. O23...24<br />

Kee, R. O19I...20<br />

Keffer, D. P47...89<br />

Kehinde, T. P10...52<br />

Kikuchi, T. P18...60<br />

Kim, B. P42...84 P43...85<br />

Kim, C. O20...21<br />

Kim, G. P24...66<br />

Kim, H. P42...84 P43...85<br />

Kim, S. O11...12 P04...46<br />

P30...72<br />

Kitamura, N. O28K...29<br />

Kitchaev, D. P58...100<br />

Kiyono, H. P33...75<br />

Kjølseth, C. O29...30 P40...82<br />

Klock, U. O03...4<br />

Knee, C. O10...11 P37...79<br />

Kocha, S. O18K...19<br />

Konysheva, E. P21...63<br />

Kozawa, Y. P31...73<br />

Kreuer, K. O03...4 O34I...35<br />

P49...91 P54...96<br />

Kruth, A. P22...64<br />

Kubota, K. P29...71<br />

Kudo, A. O36I...37<br />

Kuramitsu, A. O30...31<br />

Kurita, N. P05...47<br />

Kurosu, K. O07...8<br />

Kuwabara, A. P32...74<br />

Kuwabara, H. O28K...29<br />

Kuwata, N. O32...33<br />

[L]<br />

Larring, Y. O22...23<br />

Lee, H. P42...84 P43...85<br />

Lee, J. P42...84 P43...85<br />

Lee, S. O18K...19<br />

Lein, H. O22...23<br />

Licoccia, S. P45...87<br />

Løken, A. P40...82<br />

Louie, M. O37...38<br />

Løvvik, O. O27...28<br />

Lucio-Vega, J. P56...98<br />

Lyonnard, S. O16K...17 O34I...35


[M]<br />

Maekawa, H. O07...8 O36I...37<br />

P33...75<br />

Magrasó, A. O22...23 P09...51<br />

P23...65<br />

Maier, J. O03...4 P35...77<br />

P49...91<br />

Maréchal, M. P02...44<br />

Marrony, M. O13...14<br />

Martin, M. O21K...22 P04...46<br />

Matic, A. O10...11<br />

Matsuda, A. P07...49 P28...70<br />

Matsui, J. P01...43<br />

Matsumoto, H. O23...24<br />

Matsunaga, K. P32...74<br />

Matsuo, Y. P50...92<br />

Mendil-Jakani, H. O16K...17 P02...44<br />

Merinov, B. O26...27<br />

Merkle, R. P35...77<br />

Meulenberg, W. P14...56<br />

Middlemiss, D. O05K...6<br />

Miyashita, T. P01...43<br />

Miyoshi, S. O32...33 P18...60<br />

Moriwake, H. P32...74<br />

Mossa, S. O16K...17<br />

Mugikura, Y. P31...73<br />

Munir, Z. P04...46 P30...72<br />

Muto, H. P07...49 P28...70<br />

[N]<br />

Nagao, Y. P01...43 O31K...32<br />

Nagata, S. P06...48<br />

Neaton, J. P25...67<br />

Nguyen, L. P10...52<br />

Nielsen, U. P<strong>15</strong>...57<br />

Niepceron, F. P02...44<br />

Norby, T. P51...93 O22...23<br />

O27...28 O29...30<br />

P03...45 P14...56<br />

P16...58 P37...79<br />

P40...82<br />

Nørgaard, C. P<strong>15</strong>...57<br />

Nose, Y. O30...31 P20...62<br />

[O]<br />

O’Hayre, R. O19I...20 O39...40<br />

P17...59<br />

Oba, F. P32...74<br />

Oh, S. P28...70<br />

Oikawa, I. P33...75<br />

Okuyama, Y. P05...47<br />

Ollivier, J. O16K...17<br />

Oyama, Y. O32...33 P19...61<br />

[P]<br />

Paddison, S. O17...18 P41...83<br />

P47...89 P54...96<br />

Panithipongwut, C. P57...99<br />

Park, J. P43...85<br />

Peng, R. P34...76<br />

Pergolesi, D. P44...86 P45...87<br />

Portale, G. O03...4 P49...91<br />

[Q]<br />

[R]<br />

Rahmati, B. P35...77<br />

Rasim, K. O13...14<br />

Ray, H. O24...25<br />

Reimer, J. P25...67<br />

Reiter, G. O04...5<br />

Ricote, S. O29...30 P13...55<br />

P14...56<br />

Rossman, G. O<strong>15</strong>I...16<br />

Roudgar, A. O25K...26<br />

[S]<br />

Saito, K. P06...48<br />

Sakai, M. P07...49<br />

Sakai, T. O23...24<br />

Sakamoto, H. P07...49<br />

Sammes, N. P17...59<br />

Sanders, M. O19I...20<br />

Sata, N. O31K...32 P01...43<br />

P29...71<br />

Sato, Y. O23...24<br />

Sato, D. P05...47<br />

Schuster, M. P49...91<br />

Shepardson, D. P10...52<br />

Shikama, T. P06...48<br />

Shimizu, T. P33...75<br />

Shinoda, K. P20...62<br />

Shirpour, M. P35...77<br />

Shogo, M. P19...61<br />

Sigle, W. P35...77<br />

Skou, E. P<strong>15</strong>...57<br />

Son, J. P42...84 P43...85<br />

Strandbakke, R. P16...58


Subramaniyan, A. P17...59<br />

Syvertsen, G. O22...23 P23...65<br />

P48...90<br />

[T]<br />

Tadanaga, K. P39...81<br />

Takahashi, H. O28K...29<br />

Takamura, H. O07...8<br />

Tamaru, M. P18...60<br />

Tanaka, I. P32...74<br />

Tanaka, S. P29...71<br />

Tansho, M. P33...75<br />

Tao, S. P21...63<br />

Tatsumisago, M. P39...81<br />

Thomas, A. O34I...35<br />

Tong, J. O19I...20<br />

P17...59<br />

O39...40<br />

Toyoura, K. P03...45<br />

Traversa, E. O08K...9<br />

P45...87<br />

P44...86<br />

Tsuchiya, B. P06...48<br />

Tsurui, T. P19...61<br />

Tuckerman, M. O01K...2 P54...96<br />

[U]<br />

Uda, T. O30...31 P20...62<br />

Unemoto, A. O28K...29 P26...68<br />

P27...69<br />

[V]<br />

Van Aken, P. P35...77<br />

Varga, Á. P55...97<br />

Vartak, S. O25K...26<br />

Verbraeken, M. O09...10<br />

Viana, H. O09...10<br />

Vilciauskas, L. P54...96 O34I...35<br />

Virkar, A. O14K...<strong>15</strong><br />

Voth, G. P12...54<br />

[W]<br />

Wang, H. P13...55<br />

Wang, L. P40...82<br />

Wang, S. P36...78<br />

Watanabe, T. P31...73<br />

Wilkening, M. P04...46<br />

Wilkinson, D. P11...53<br />

[X]<br />

Xia, C. P34...76<br />

Xing, W. P48...90<br />

Xu, X. P21...63<br />

[Y]<br />

Yamaguchi, S. Plenary ...1 O32...33<br />

P18...60 P19...61<br />

Yamamoto, H. P01...43<br />

Yamazaki, Y. O38...39 P56...98<br />

Yang, C. O38...39<br />

Yeon, J. P08...50<br />

Yoo, H. O21K...22 P08...50<br />

Yoshida, T. P28...70<br />

Yoshida, Y. P50...92<br />

Yugami, H. P01...43 P29...71<br />

[Z]<br />

Zhang, L. P36...78<br />

Zhao, F. P36...78 P38...80<br />

Zhu, H. O19I...20

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!