BIOGRAFIJA Dr. Dušan V. Tripković je rođen 28.04.1979. godine u ...

BIOGRAFIJA Dr. Dušan V. Tripković je rođen 28.04.1979. godine u ... BIOGRAFIJA Dr. Dušan V. Tripković je rođen 28.04.1979. godine u ...

ihtm.bg.ac.rs
from ihtm.bg.ac.rs More from this publisher
20.02.2013 Views

BIOGRAFIJA Dr. Dušan V. Tripković je rođen 28.04.1979. godine u Beogradu. Završio je osovnu školu “Vladislav Ribnikar” i “Prvu beogradsku gimnaziju” u Beogradu, sa odličnim uspehom. Tehnološko metalurški fakultet upisao je 1998. godine (smer hemijsko inženjerstvo), a diplomirao je 26.11.2003. godine sa prosečnom ocenom 8,53 i ocenom diplomskog rada 10. Diplomski rad je uradio u Razvojno-istraživačkom centru “Haldor Topsoe” u Kopenhagenu, Danska. Magistraske studije upisao je na Tehnološko metalurškom fakultetu 2003. godine. Nakon položenih ispita sa prosečnom ocenom 10, eksperimentalni deo teze je uradio na Institute of Catalysis and Surface Chemistry, Polish Academy of Science, Krakov, Poljska. Zvanje Magistar tehničkih nauka stekao je 2005. godine odbranivši magistarski rad na temu “Karakterizacija katalizatora dobijenih modifikovanjem staklastog ugljenika platinom” na Tehnološko metalurškom fakultetu. Doktorsku disertaciju “Uticaj morfologije površine platinskih materijala na elektrokatalitičku aktivnost u gorivim spregovima” uradio je u Nacionalnoj laboratoriji Argonne, Čikago, SAD, a odbranio je na Tehnološko metalurškom fakultetu u Beogradu 2008. godine i time stekao zvanje Doktor tehničkih nauka Od 1 januara 2003. godine zaposlen je na Institutu za hemiju, tehnologiju i metalurgiju, Centar za elektrohemiju, u Beogradu. Trenutno zvanje Kandidata je viši naučni saradnik. Od 2003 god angažovan je na projektima osnovnih istraživanja iz hemije finansiranih od strane Ministarstva U okviru međunarodne saradnje učestvovao je i učestvuje na dva bilateralna projekta sa Poljskom i četiri američka projekta. U toku 2004/2005 godine bio je na studijskom boravku u Institute of Catalysis and Surface Chemistry, Polish Academy of Science, Krakov, Poljska. Na doktorskim studijama je bio 2005/2007 godine u Argonne National Laboratory, Čikago, SAD. U istoj insituciji boravio je kao post doktor u periodu 2009/2011 god. Dobitnik je nagrade ”Panta Tutundžić” za izuzetan uspeh na poslednjoj godini studija na TMF-u i nagrade Nacionalne laboratorije „Argonne“ za prijavljen patent (oktobar 2012.) Govori tečno engleski jezik i ima školsko znanje francuskog jezika.

<strong>BIOGRAFIJA</strong><br />

<strong>Dr</strong>. <strong>Dušan</strong> V. <strong>Tripković</strong> <strong>je</strong> <strong>rođen</strong> <strong>28.04.1979.</strong> <strong>godine</strong> u Beogradu. Završio <strong>je</strong> osovnu<br />

školu “Vladislav Ribnikar” i “Prvu beogradsku gimnaziju” u Beogradu, sa odličnim uspehom.<br />

Tehnološko metalurški fakultet upisao <strong>je</strong> 1998. <strong>godine</strong> (smer hemijsko inžen<strong>je</strong>rstvo), a<br />

diplomirao <strong>je</strong> 26.11.2003. <strong>godine</strong> sa prosečnom ocenom 8,53 i ocenom diplomskog rada 10.<br />

Diplomski rad <strong>je</strong> uradio u Razvojno-istraživačkom centru “Haldor Topsoe” u Kopenhagenu,<br />

Danska. Magistraske studi<strong>je</strong> upisao <strong>je</strong> na Tehnološko metalurškom fakultetu 2003. <strong>godine</strong>.<br />

Nakon položenih ispita sa prosečnom ocenom 10, eksperimentalni deo teze <strong>je</strong> uradio na<br />

Institute of Catalysis and Surface Chemistry, Polish Academy of Science, Krakov, Poljska.<br />

Zvan<strong>je</strong> Magistar tehničkih nauka stekao <strong>je</strong> 2005. <strong>godine</strong> odbranivši magistarski rad na temu<br />

“Karakterizacija katalizatora dobi<strong>je</strong>nih modifikovan<strong>je</strong>m staklastog ugl<strong>je</strong>nika platinom” na<br />

Tehnološko metalurškom fakultetu. Doktorsku disertaciju “Uticaj morfologi<strong>je</strong> površine<br />

platinskih materijala na elektrokatalitičku aktivnost u gorivim spregovima” uradio <strong>je</strong> u<br />

Nacionalnoj laboratoriji Argonne, Čikago, SAD, a odbranio <strong>je</strong> na Tehnološko metalurškom<br />

fakultetu u Beogradu 2008. <strong>godine</strong> i time stekao zvan<strong>je</strong> Doktor tehničkih nauka<br />

Od 1 januara 2003. <strong>godine</strong> zaposlen <strong>je</strong> na Institutu za hemiju, tehnologiju i<br />

metalurgiju, Centar za elektrohemiju, u Beogradu. Trenutno zvan<strong>je</strong> Kandidata <strong>je</strong> viši naučni<br />

saradnik.<br />

Od 2003 god angažovan <strong>je</strong> na pro<strong>je</strong>ktima osnovnih istraživanja iz hemi<strong>je</strong> finansiranih<br />

od strane Ministarstva<br />

U okviru međunarodne saradn<strong>je</strong> učestvovao <strong>je</strong> i učestvu<strong>je</strong> na dva bilateralna pro<strong>je</strong>kta<br />

sa Poljskom i četiri američka pro<strong>je</strong>kta.<br />

U toku 2004/2005 <strong>godine</strong> bio <strong>je</strong> na studijskom boravku u Institute of Catalysis and<br />

Surface Chemistry, Polish Academy of Science, Krakov, Poljska. Na doktorskim studijama <strong>je</strong><br />

bio 2005/2007 <strong>godine</strong> u Argonne National Laboratory, Čikago, SAD. U istoj insituciji<br />

boravio <strong>je</strong> kao post doktor u periodu 2009/2011 god.<br />

Dobitnik <strong>je</strong> nagrade ”Panta Tutundžić” za izuzetan uspeh na poslednjoj godini studija<br />

na TMF-u i nagrade Nacionalne laboratori<strong>je</strong> „Argonne“ za prijavl<strong>je</strong>n patent (oktobar 2012.)<br />

Govori tečno engleski <strong>je</strong>zik i ima školsko znan<strong>je</strong> francuskog <strong>je</strong>zika.


BIBLIOGRAFIJA<br />

<strong>Dr</strong>. <strong>Dušan</strong> V. <strong>Tripković</strong> <strong>je</strong> izuzetno uspešan u karakterizaciji površine<br />

katalitičkih materijala za primenu u gorivim spregovima korišćen<strong>je</strong>m spektroskopskih<br />

(FTIR) i mikroskopskih tehnika (STM i AFM). Osmišl<strong>je</strong>nom pripremom površina i<br />

primenom ovih tehnika <strong>Dr</strong>. <strong>Dušan</strong> <strong>Tripković</strong> <strong>je</strong> prvi u svetu detektovao površinska<br />

ostrva kao aktivna mesta na platinskim monokristalnim katalizatorima, čime <strong>je</strong><br />

otvoren put dizajniranju superiornih katalizatora.<br />

Naučni doprinos <strong>Dr</strong>. <strong>Dušan</strong>a V. <strong>Tripković</strong>a pored pronalaska aktivnih mesta<br />

na površini monokristala <strong>je</strong> i sinteza i karakterizacija katalizatora na bazi Pt sa<br />

klasterima Ni(OH)2 koji poseduju za red veličine veću aktivnost i značajno<br />

unapređenu stabilnost od čiste Pt koja se trenutno koristi za reakciju izdvajanja<br />

vodonika u alkalnim i hlor-alkalnim elektrolizerima. Ovaj pronalazak potencijalno<br />

može dovesti do uštede pri proizvodnji vodonika u vrednosti od više stotina miliona<br />

dolara na godišnjoj bazi i to samo u SAD (prijavl<strong>je</strong>n patent).<br />

Po pitanju gorivih spregova, koji i predstavljaju glavnu oblast n<strong>je</strong>govog<br />

interesovanja, <strong>Dr</strong>. <strong>Dušan</strong> <strong>Tripković</strong> <strong>je</strong> pokušao da pronađe rešen<strong>je</strong> za dva ključna<br />

problema komercjalizaci<strong>je</strong> ovih uređaja; stablinost u agresivnim uslovima pod kojima<br />

rade i aktivnost katalitičkog materijala za reakciju redukci<strong>je</strong> kiseonika. Sa tim u vezi<br />

kandidat <strong>je</strong> predstavio i sasvim novi metod zaštite katalitizatora, na bazi ori<strong>je</strong>ntisanog<br />

sloja calih[4]arene molekula, od extremnih uslova u kojima se mogu naći tokom<br />

puštanja u pogon/gašenja gorivih spregova. Takođe <strong>je</strong> i dizajnirao Pt-bimetalni<br />

katalizator u vidu tankog filma nanesenog na nosač koji pokazu<strong>je</strong> za red veličine veću<br />

aktivnost i značajno unapređenu stabilnost od aktuelnog Pt/C kazalizatora za reakciju<br />

redukci<strong>je</strong> kiseonika.<br />

<strong>Dr</strong>. <strong>Dušan</strong> V. <strong>Tripković</strong> <strong>je</strong> objavio 30 naučnih radova od kojih: 18 radova u<br />

prvoj kategoriji vrhunskih međunarodnih časopisa (M21) sa izuzetno visokim impakt<br />

faktorima, 1 rad u istaknutim časopisima međunarodnog značaja (M22), 10 radova u<br />

međunarodnim časopisima (M23) i 1 rad u nacionalnom časopisu (M51). Ukupni<br />

IF=215.335. Kandidat takođe ima i 26 saopštenja na međunarodnim (M34) i 4 na<br />

nacionalnim naučnim skupovima (M61).<br />

Od izbora u zvan<strong>je</strong> viši naučni saradnik objavio <strong>je</strong>: 8 radova kategori<strong>je</strong> M21,<br />

1 rad kategori<strong>je</strong> M22 i 5 radova kategori<strong>je</strong> M23. Ukupni IF=172.876. Kandidat ima<br />

16 saopštenja na međunarodnim skupovima iz kategorija M33 i M34.<br />

<strong>Dr</strong>. <strong>Dušan</strong> V. <strong>Tripković</strong> <strong>je</strong> učestvovao i aktivno učestvu<strong>je</strong> na šest<br />

međunarodnih pro<strong>je</strong>kata, dva u okviru bilateralne saradn<strong>je</strong> IHTM-a i Poljske<br />

akademi<strong>je</strong> nauka, tri Američka pro<strong>je</strong>kta finansirana od strane; University of Chicago<br />

and Argonne National laboratory, US Department of Energy i Basic Energy Science i<br />

pro<strong>je</strong>kat koji predstavlja saradnju američke Nacionalne laboratori<strong>je</strong> Argonne i<br />

razvojnog centra Toyote.<br />

Februara 2009. <strong>godine</strong> po pozivu <strong>je</strong> održao predavan<strong>je</strong> na Technical<br />

University of Denmark, Kopenhagen, Danska.<br />

Citiranost <strong>Dr</strong>. <strong>Dušan</strong>a V. <strong>Tripković</strong>a prema izveštaju servisa Scopus iznosi<br />

358, odnosno 330 bez autocitata, na dan 25 januar 2013.<br />

<strong>Dr</strong>. <strong>Dušan</strong> V. <strong>Tripković</strong> ima i <strong>je</strong>dan prijavl<strong>je</strong>n patent za koji <strong>je</strong> dobio nagradu<br />

američke Nacionalne laboratori<strong>je</strong> „Argonne“.


III. SPISAK NAUČNIH RADOVA I SAOPŠTENJA OD IZBORA U ZVANJE<br />

VIŠI NAUČNI SARADNIK<br />

M 21 – Radovi objavl<strong>je</strong>ni u vrhunskim časopisima međunarodnog značaja (8x8=64)<br />

M21/1. Bostjan Genorio, Dusan Strmcnik, Ram Subbaraman, Dusan Tripkovic,<br />

Goran Karapetrov, Vojislav R. Stamenkovic, Stane Pejovnik and Nenad M.<br />

Markovic “Selective catalyst for the hydrogen oxidation and oxygen<br />

reduction reactions by patterning of platinum with calix[4]arene<br />

molecules” Nature Materials (12), 998-1003, (2010)<br />

[IF: 29.920 (2010); oblast: Chemistry, Physical, 1/127]<br />

M21/2 Bostjan Genorio, Ram Subbaraman, Dusan Strmcnik, Dusan Tripkovic,<br />

Vojislav R. Stamenkovic, and Nenad M. Markovic “ Tailoring the selectivity<br />

and stability of chemically modified platinum nanocatalysts to design<br />

highly durable anodes for PEM fuel cells” Angewante Chemie, 50(24),<br />

5468-72, (2011)<br />

[IF: 13.455 (2011); oblast: Chemistry, Multidisciplinary, 7/154<br />

M21/3 Chao Wang, Miaofang Chi, Dongguo Li, Dusan Strmcnik, Dennis van der<br />

Vliet,Guofeng Wang,Vladimir Komanicky, Kee-Chul Chang, Arvydas P.<br />

Paulikas, Dusan Tripkovic, John Pearson,Karren L. More, Nenad M.<br />

Markovic, and Vojislav R. Stamenkovic ” Design and Synthesis of<br />

Bimetallic Electrocatalyst with Multilayered Pt-Skin Surfaces” Journal of<br />

the American Chemical Society,133(36),14396-403, (2011)<br />

[IF: 9.907 (2011); oblast: Chemistry, Multidisciplinary, 11/154]<br />

M21/4 Ram Subbaraman, Dusan Tripkovic, Dusan Strmcnik, Kee-Chul Chang,<br />

Masanobu Uchimura, Arvydas P. Paulikas, Vojislav Stamenkovic, Nenad M.<br />

Markovic “Enhancing Hydrogen Evolution Activity in Water Splitting by<br />

Tailoring Li + -Ni(OH)2-Pt Interfaces”, Science, vol. 334 no. 6060 pp. 1256-<br />

1260, (2011)<br />

[IF: 31.201 (2011); oblast: Multidisciplinary Sciences, 2/56]<br />

M21/5 Dennis van der Vliet, Chao Wang, Dongguo Li, Arvydas P. Paulikas,Jeffrey<br />

Greely,Rees B. Rankin, Dusan Strmcnik, Dusan Tripkovic, Nenad M.<br />

Markovic, and Vojislav R. Stamenkovic “Unique Electrochemical<br />

Adsorption Properties of Pt-skin Surfaces”, Angewante Chemie,Vol 51,<br />

Issue 13, 3139–3142, (2012)<br />

[IF: 13.455 (2011); oblast: Chemistry, Multidisciplinary, 7/154]


M21/6 Ram Subbaraman, Dusan Tripkovic,Kee-Chul Chang,Dusan Strmcnik,<br />

Arvydas P.Paulikas,Pussana Hirunsit, Maria Chan, Jeff Greeley, Vojislav<br />

Stamenkovic and Nenad M.Markovic” Trends in activity for the water<br />

electrolyser reactions on 3dM (Ni,Co,Fe,Mn) hydr(oxy)oxide<br />

catalysts”,Nature Materials 11, (6), 550-557 (2012) DOI: 10.1038/nmat3313<br />

[IF: 32.841 (2011); oblast: Chemistry, Physical, 1/134]<br />

M21/7 Dennis van der Vliet, Chao Wang, Dusan Tripkovic, Dusan Strmcnik, Xiao<br />

Zhang, Mark Debe, Radoslav Atanososki, Nenad Markovic and Vojislav<br />

Stamenkovic “Mesostructured Thin Films as Electrocatalysts with<br />

Tunable Composition and Surface Morphology”, Nature Materials, (2012),<br />

DOI: 10.1038/nmat3457<br />

[IF: 32.841 (2011); oblast: Chemistry, Physical, 1/134]<br />

M21/8 Ram Subbaraman, N. Danilovic, P. P. Lopes, D. Tripkovic, D. Strmcnik, V.<br />

R. Stamenkovic, and N. M. Markovic “Origin of Anomalous Activities for<br />

Electrocatalysts in Alkaline Electrolytes”, J. Phys. Chem. C, 116 (42), pp<br />

22231–22237 (2012), DOI: 10.1021/jp3075783<br />

[IF: 4.805 (2011); oblast: Chemistry, Physical, 26/134]<br />

M22 – Radovi objavl<strong>je</strong>ni u istaknutim časopisima međunarodnog značaja (1x5=5)<br />

M22/1 Stevanovic I. Sanja, Tripkovic V. Dusan, Rogan R. Jelena, Popovic Dj.<br />

Ksenija, Lovic D. Jelena, Tripkovic V. Amalija and Jovanovic M Vladislava,<br />

“Microwave-assisted polyol synthesis of carbon-supported platinumbased<br />

bimetallic catalysts for ethanol oxidation”, Journal of Solid State<br />

Electrochemistry, vol. 16 br. 10, str. 3147-3157, (2012)<br />

[IF: 2.234 (2010); oblast: Electrochemistry, 13/26]<br />

M23 – Radovi objavl<strong>je</strong>ni u časopisima međunarodnog značaja (3x3=9)<br />

M23/1 S. Stevanovic, D. Tripkovic, J. Rogan, D. Minic, A. Gavrilovic, A. Tripkovic<br />

and V. M. Jovanovic: “Enhanced Activity in Ethanol Oxidation of Pt3Sn<br />

Electrocatalysts Synthesized by Microwave Irradiation”, Russ. J. Phys.<br />

Chem., 85(13),2299-2304, (2011)<br />

[IF=0.459 (2011); oblast: Chemistry, Physical, 125/134]


M23/2 S. Stevanovic, D. Tripkovic, D. Poleti, J. Rogan, A. Tripkovic and<br />

V.M.Jovanovic: “Microwave Synthesis and Characterization of PT and<br />

Pt-Rh-Sn Electrocatalyst for Ethanol Oxidation” J. Serb. Chem.Soc. 76<br />

(12), 1673-1685, (2011)<br />

[IF= 0.879 (2011); oblast: Chemistry – multidisciplinary, 103/154]<br />

M23/3 Jelena D.Lovic, Dusan V. Tripkovic, Ksenija DJ.Popovic, Vladislava M.<br />

Jovanovic and Amalija Tripkovic “Electrocatalytic Properties of Pt-Bi<br />

Electrodes Towards the Electrooxidation of Formic Acid”, J. Serb. Chem.<br />

Soc., DOI:10.2298/JSC121012138L<br />

[IF= 0.879 (2011); oblast: Chemistry – multidisciplinary, 103/154]<br />

M23/4 Dongguo Li, Chao Wang, Dusan Tripkovic, Shouheng Sun, Nenad M.<br />

Markovic, and Vojislav R. Stamenkovic “Surfactant Removal for Colloidal<br />

Nanoparticles from Solution Synthesis: The Effect on Catalytic<br />

Performance”, ACS Catalysis, 2 (7), pp 1358–1362 (2012),<br />

DOI:10.1021/cs300219j<br />

[IF dobija 2013 god.; oblast: Chemistry-Physical, 130/134]<br />

M23/5 J.D.Lovic, M.D.Obradovic, D.V.Tripkovic, K,Dj.Popovic,V.M.Jovanovic,<br />

S.LJ. Gojkovic, A.V.Tripkovic “High Activity and Stability of Pt2Bi<br />

Catalyst in Formic Acid Oxidation”, Springer Electrocatalysis, (2012),<br />

DOI: 10.1007/sl12678-012-0099-9<br />

[IF dobija 2013 god.: oblast: Electrochemistry]<br />

M33- Radovi saopšteni na skupu međunarodnog značaja štampani u celini (1x1 = 1 )<br />

M33/1 S. Stevanović, D. <strong>Tripković</strong>, J. Rogan, D. Minić, A. Gavrilović, A.<strong>Tripković</strong>,<br />

V.M. Jovanović: ″Microvawe assisted sinthesys of Pt and Pt3Sn<br />

electrocatalysts for ethanol oxidation″ , 10th International Conference on<br />

Fundamental and Applied Aspects of Physical Chemistry, Proceedings E-O-2,<br />

Belgrade, Serbia, 21-24 September 2010.


M34- Radovi saopšteni na skupu međunarodnog značaja štampani u izvodu<br />

(15x0.5=7.5 )<br />

M34/1 Dusan Strmcnik, Dusan Tripkovic, Dennis Van der Vliet, Jeffrey P. Greeley,<br />

Alexander Brownrigg, Christopher Lucas, Goran Karapetrov, Vojislav<br />

Stamenkovic, and Nenad Markovic „Active Sites for PEM Fuel Cell<br />

Reactions“, 216th ECS Meeting, Abstract #868, Vienna, Austria, Oct 6 2009<br />

M34/2 C. Wang, D. Li, D. van der Vliet, D. Strmcnik, D. Tripkovic, N. Markovic,<br />

and V. Stamenkovic: ”Advanced Electrocatalysts: - From Extended to<br />

Nanoscale Surfaces”, 220th ECS Abstract #947, Boston, MA, Oct 11 2011<br />

M34/3 Dusan Strmcnik, Kensaku Kodama, Dusan Tripkovic,Dennis Vander<br />

Vliet,Chao Wang,Vojislav Stamenkovic,and Nenad M. Markovic: “Design<br />

Catalytic Properties of Electrochemical Interfaces”, 217 th ECS Meeting<br />

Abstract #1755, Vancouver, Canada, Apr 27 2010<br />

M34/4 Dusan Tripkovic,Dusan Strmcnik,Dennis F. Van der Vliet,Chao Wang,Nenad<br />

M. Markovic,and Vojislav R. Stamenkovic: “In-situ Infrared Spectroscopy<br />

at Solid-Liquid Interfaces as a Tool for Evaluation of Nanoscale Surface<br />

Morphology”, 221 th ECS Meeting Abstract #1570, Seattle, WA, May 9 2012<br />

M34/5 Chao Wang,Dusan Strmcnik,Dusan Tripkovic,Dennis Van der Vliet,Nenad<br />

Markovic,and Vojislav R. Stamenkovic ” Electrocatalysis on Well-Defined<br />

Solid-Liquid Interfaces”, 219 th ECS Meeting Abstract 1924, Montreal, QC,<br />

Canada, May 2 2011<br />

M34/6 Dusan Strmcnik,Ramachandran Subbaraman,Dusan Tripkovic,Kee-Chul<br />

Chang,Nemanja Danilovic,Dennis F. Van der Vliet,Pietro Lopes,Arvydas Paul<br />

Paulikas,Vojislav R. Stamenkovic,and Nenad M. Markovic: ” Controlling<br />

Reactivity of Electrochemical Interfaces by Tuning Non-covalent<br />

Interactions”, 221 th ECS Meeting Abstract #1560, Seattle, WA, May 7 2012<br />

M34/7 Dusan Strmcnik,Dusan Tripkovic, Ramachandran Subbaraman, Nemanja<br />

Danilovic, Dennis F. Van der Vliet, Arvydas Paul Paulikas, Vojislav R.<br />

Stamenkovic and Nenad M. Markovic” Fundamental investigations of<br />

precious metal stability in energy conversion systems”, 221 th ECS Meeting<br />

Abstract #1532, Seattle, WA, May 10 2012<br />

M34/8 Ramachandran Subbaraman, Dusan Tripkovic, Dusan Strmcnik, Gustav K.<br />

Wiberg, Jakub S. Jirkovsky, Chao Wang, Vojislav R. Stamenkovic and Nenad<br />

M. Markovic: ” Electrochemical Interfaces for Energy Conversion and<br />

Storage”, 221 th ECS Meeting Abstract #549, Seattle, WA, May 7 2012


M34/9 S. Stevanović, D. <strong>Tripković</strong>, V. <strong>Tripković</strong>, D. Minić, A.Gavrilović, K.<br />

Popović, A. <strong>Tripković</strong> and V.M. Jovanović: ″Insight of Sn influence on<br />

formic acid oxidation at Pt based catalysts″, 63rd Meeting of International<br />

Society of Electrochemistry, Prague (Czech Republic), Book of Abstracts S05-<br />

027, August 2012<br />

M34/10 J.D. Lović, M.D. Obradović, D.V. <strong>Tripković</strong>, K.Dj. Popović, V.M.<br />

Jovanović, S.Lj. Gojković and A.V. <strong>Tripković</strong>: ″High Activity and Stability<br />

of Pt2Bi Catalyst in Formic Acid Oxidation″, 63rd Meeting of International<br />

Society of Electrochemistry, Prague (Czech Republic), Book of Abstracts S05-<br />

063, August 2012<br />

M34/11 S. Stevanović, D. <strong>Tripković</strong>, J. Rogan, J. Lović, K. Popović, A. <strong>Tripković</strong><br />

and V.M. Jovanović: ″Ehanol oxidation on carbon supported platinum<br />

based bimetallic catalysts synthesized by microwave assisted polyol<br />

procedure″, 63rd Meeting of International Society of Electrochemistry,<br />

Prague ( Czech Republic ), Book of Abstracts S05-052, August 2012<br />

M34/12 Chao Wang, Dennis Van der Vliet, Dusan Tripkovic, Dusan Strmcnik,<br />

Dongguo Li, Nenad M. Markovic, and Vojislav R. Stamenkovic: ”Advanced<br />

Electrocatalysts for PEM Fuel Cells”, Abstract 1650, PRiME 2012.<br />

Honolulu, Hawaii October 7-12, 2012<br />

M34/13 Nemanja Danilovic, Ramachandran Subbaraman, Dusan Strmcnik, Dusan<br />

Tripkovic, Kee-Chul Chang, Arvydas Paul Paulikas, Vojislav R.<br />

Stamenkovic, Debbie J. Myers, and Nenad M. Markovic „Electrocatalysts for<br />

the Oxygen Evolution Reaction“, 221 th ECS Meeting, Abstract #1510,<br />

Seattle, WA, May 9 2012<br />

M33/14 R. Subbaraman, D. Tripkovic, G.K. Wiberg, J.S. Jirkovsky and N. M.<br />

Markovic: ” Li-Air- Just a <strong>Dr</strong>eam or Future Reality”,DOE-Chine Meeting-<br />

Beyond Li-ion systems, Boston, MS, September 2012, (Invited)<br />

M33/15 D. Tripkovic D. Strmcnik, R. Subbaraman, D. V. Stamenkovic and N. M.<br />

Markovic: “Tailored Nanomaterials for Clean Energy Conversion and<br />

Storage International Symposium”, Tsukuba, Japan, March 2012 (Invited)


III. ANALIZA RADOVA<br />

Rad M21/1 U radu <strong>je</strong> prikazan novi pristup dizajniranju katalizatora za gorive ćelija<br />

sa polimernom membranom koji mora biti vođen sa dva <strong>je</strong>dnako važna principa:<br />

optimizacija katalitičke aktivnosti i dugoročna stabilnost metalnih katalizatora i<br />

njihovih nosača u radnoj sredini (elektrolitu). Postoji širok spektar metoda ko<strong>je</strong> se<br />

koriste za unapređivan<strong>je</strong> katalitičke aktivnosti, međutim, metode za poboljšan<strong>je</strong><br />

stabilnosti ugl<strong>je</strong>ničnog nosača i katalizatora su ograničene, a naročito tokom gašenja i<br />

pal<strong>je</strong>nja gorivog sprega. Pod ovim okolsnostima na stabilnost katodnog materijala<br />

(oksidaci<strong>je</strong> ugl<strong>je</strong>nika) snažno utiče redukcija kiseonika (Orr) na anodnoj strani. U<br />

ovom radu pokazano <strong>je</strong> da hemijski modifikovana platina sa samo ori<strong>je</strong>ntisanim<br />

monoslo<strong>je</strong>m calix[4]arene molekula može biti primen<strong>je</strong>na kao selektivni anodi<br />

materijal koji efektivno suzbija Orr a da pri tome ne inhibira reakciju oksidaci<strong>je</strong><br />

vodonika. Ovako dizajnirani katalitički materijal ispoljava izuzetno veliku stabilnost<br />

čak i pri na<strong>je</strong>kstremnijim uslovima u kojima se može naći tokom rada gorivog sprega.<br />

Rad M21/2 U ovom radu nastavl<strong>je</strong>no <strong>je</strong> ispitivan<strong>je</strong> efekta hemijskog modifikovanja<br />

platine monoslo<strong>je</strong>m calix[4]arene molekula ko<strong>je</strong> da<strong>je</strong> vrlo stabilnu anodu u gorivom<br />

spregu usled mogućnosti da efektivno suzbija reakciju oksidaci<strong>je</strong> kiseonika, a da pri<br />

tome ne utiče na maksimalnu aktivnost za reakciju oksidaci<strong>je</strong> vodonika. Najveći<br />

doprinos ovog rada <strong>je</strong> to sto <strong>je</strong> pokazano da <strong>je</strong> pomenuti efekat univerzalan i da se<br />

proteže od visoko uređenih monokristalnih površina pa sve do komercijalnih<br />

nanokatalizatora.<br />

Rad M21/3 Napredak u dizajniranju heterogenih katalizatora oslanja se na sposobnost<br />

modifikovanja strukture materijala na nanoskali. U ovom radu dizajniran <strong>je</strong> i<br />

napravl<strong>je</strong>n katalizator na bazi platine sa višeslojnom bimetalnom strukturom koji<br />

pokazu<strong>je</strong> izuzetnu elektrokatalitičku aktivnost u reakciji redukci<strong>je</strong> kiseonika (Orr).<br />

Ovakva struktura prvo <strong>je</strong> testirana na tankoslojnim površinama sa tačno određenim<br />

sastavom, a zatim implementirana u izradu nanokatalizatora korišćen<strong>je</strong>m organske<br />

sinteze. Elektrokemijska istraživanja za Orr su pokazala da nakon dužeg izlaganja<br />

reakcijskim uslovima, Pt-bimetalni katalizator sa višeslojnom strukturom pokazu<strong>je</strong> za<br />

red veličine veću aktivnost, dok <strong>je</strong> pri tome i stabilnost katalizatora značajno<br />

unapred<strong>je</strong>na u odnosu na aktivnosti i stabilnost konvencionalnog Pt katalizatora.<br />

Znatno poboljšane katalitičke aktivnosti i stabilnosti pokazu<strong>je</strong> veliki potencijal pri<br />

izradi superiornih komercijalnih nanokatalizatora za Orr.<br />

Rad M21/4 Poboljšan<strong>je</strong> spore kinetike elektrokemijske redukci<strong>je</strong> vode do<br />

molekularnog vodonika u alkalnim sredinama <strong>je</strong> ključni faktor za sman<strong>je</strong>n<strong>je</strong> visokih<br />

nadpotencijala i sa njima povezanih energetskih gubitaka u alkalnim i hlor-alkalnim<br />

elektrolizerima. U ovom radu pokazano <strong>je</strong> da kontrolisanim raspoređivan<strong>je</strong>m nanoklastera<br />

Ni(OH)2 po površini platine manifestu<strong>je</strong> povećanja katalitičke aktivnosti za<br />

faktor 8 kod vodonične reakci<strong>je</strong> u poređenju sa najmodernijim metalnim i metaloksidnim<br />

katalizatorima. Kako bi se objasnila dobi<strong>je</strong>na aktivnost postavl<strong>je</strong>na <strong>je</strong> nova<br />

teorija da bifunkcionalnim mehanizmom, ivice Ni(OH)2 klastera podstiču disocijaciju<br />

vode i nastanak intermedi<strong>je</strong>ra koji se adsorbuju na obližn<strong>je</strong> atome platine i nakon<br />

rekombinaci<strong>je</strong> daju konačan proizvod - molekularni vodonik. Pokazano <strong>je</strong> i da<br />

nastanak intermedi<strong>je</strong>ra može dodatno biti favorizovano putem Li + indukovane<br />

destabilizaci<strong>je</strong> HO-H veze, što na kraju dovodi do ukupnog povećanja aktivnosti za<br />

faktor 10.


Rad M21/5 Legiran<strong>je</strong>m platine sa drugim ne plemenitim metalom i formiran<strong>je</strong>m Ptskin<br />

strukture dolazi do značajne promene adsorpcionih svojstava u odnosu na čistu<br />

platinu. Naime, potencijal adsorpci<strong>je</strong> vodonika i oksidnih vrsta <strong>je</strong> pomeren i sman<strong>je</strong>n<br />

u odnosu na čistu Pt u istoj potencijalnoj regiji. Čin<strong>je</strong>nica da <strong>je</strong> količina naelektrisanja<br />

dobi<strong>je</strong>na oksidacijom punog monosloja CO približno ista na aniliranim Pt3M (Mmetal)<br />

površinama kao kod čiste Pt potvrđu<strong>je</strong> da se površina sastoji od čiste Pt.<br />

Sman<strong>je</strong>na adsorpcija vodonika (Hupd) na površini se može koristiti kao metoda za<br />

potvrdu Pt-skin strukture. Sa druge strane proračun na osnovu količine adsorbovanog<br />

vodonika predstavlja problem pri određivanju elektrohemijski aktivne površine na<br />

nanočesticama sa Pt-skin strukturom. U ovom radu, kako bi se izbegla greška<br />

prilikom proračuna korišćena <strong>je</strong> i metoda oksidaci<strong>je</strong> monosloja CO uporedo sa<br />

metodom proračuna aktivne površine Pt na osnovu količine adsorbovanog vodonika.<br />

Rad M21/6 Dizajn i sinteza materijala za efikasanu elektrohemijsku transformaciju<br />

vode do molekularnog vodonika i hidroksilnih jona do kiseonika u alkalnim<br />

sredinama <strong>je</strong> od neprocenjive važnosti za sman<strong>je</strong>n<strong>je</strong> gubitaka energi<strong>je</strong> u alkalnim<br />

elektrolizerima. U ovom radu, korišćen<strong>je</strong>m 3D-M hydr(oksi)oksida, strogo<br />

definisanih u pogledu stehiometri<strong>je</strong> i morfologi<strong>je</strong> u reakcijama izdvajanja vodonika<br />

(HER) i redukci<strong>je</strong> kiseonika (OER), određene su ukupne katalitičke aktivnosti za ove<br />

reakci<strong>je</strong> kao funkcija fundamentalne zavisnosti, deskriptora, OH-M2 + δ jačina veze<br />

(0 ≤ δ ≤ 1,5). Ovaj odnos pokazu<strong>je</strong> trendove u reaktivnosti (Mn


Rad M22/1 Katalizatori na bazi Pt, PtRh i PtSn nanetih na ugl<strong>je</strong>nik razvi<strong>je</strong>ne<br />

površine sintetisani su pomoć u mikrotalasne-poliol metode i testirani za elektrooksidaciju<br />

etanola. Katalizatori su okarakterisani korišćen<strong>je</strong>m: termogravimetrijske<br />

analize (TGA), X-ray difrakci<strong>je</strong> (XRD), skenirajuće tunelirajuće mikroskopi<strong>je</strong> (STM),<br />

TEM, i EDX tehnika. Rezultati su pokazali da <strong>je</strong> PtSn/C približno tri puta aktivniji<br />

dok <strong>je</strong> PtRh/C pokazao samo neznatno poboljšan<strong>je</strong> u odnosu na Pt/C katalizator.<br />

Potencijal početka oksidaci<strong>je</strong> etanola na PtSn katalizatoru predtsavlja najniži<br />

zabeleženi potencijal u literaturi. Hronoamerometrijska merenja su pokazala da <strong>je</strong><br />

PtSn/C znatno stabilniji od Pt/C katalizatora.<br />

Rad M23/1. U ovom radu sintetisani su katalizatori na bazi Pt i Pt3Sn korišćen<strong>je</strong>m<br />

mikrotalasnog zračenja. Nakon sinteze katalizatori su naneti na nosač od ugl<strong>je</strong>nika<br />

razvi<strong>je</strong>ne površine i ispitani u reakciji elektro-oksidaci<strong>je</strong> etanola. Katalizatori su<br />

dobi<strong>je</strong>ni primenom izmen<strong>je</strong>ne poliol metode u rastvoru etilen glikola i okarakterisani<br />

u pogledu strukture, morfologi<strong>je</strong> i sastava korišćen<strong>je</strong>m XED, STM i EDX tehnika.<br />

Pokazano <strong>je</strong> da <strong>je</strong> Pt3Sn / C katalizator vrlo aktivan za oksidaciju etanola sa početnim<br />

potencijalom pomerenim za ~ 150 mV negativni<strong>je</strong> i strujom i ~ 2 puta većom u<br />

odnosu na Pt/C katalizator.<br />

Rad M23/2 Pt i Pt-Rh-Sn katalizatori sintetisani su korišćen<strong>je</strong>m mikrotalasne-poliol<br />

metode u rastvoru etilen glikola i testirani u reakciji elektro-oksidaci<strong>je</strong> etanola.<br />

Katalizatori su okarakterisani u pogledu strukture, morfologi<strong>je</strong> i sastava korišćen<strong>je</strong>m<br />

XRD, STM i EDX tehnika. Male veličine i homogena distribucija veličina čestica oba<br />

katalizatora <strong>je</strong> rezultat korišćenja mikrotalasnog zračenja. Pokazano <strong>je</strong> da <strong>je</strong> Pt-Rh-<br />

Sn/C katalizatora <strong>je</strong> vrlo aktivan za elektro-oksidaciju etanola sa početnim<br />

potencijallom oksidaci<strong>je</strong> pomerenim za više od 150 mV negativni<strong>je</strong> i strujom<br />

oksidaci<strong>je</strong> gotovo pet puta većom u odnosu na Pt/C katalizator. Unapređena aktivnost<br />

<strong>je</strong> posledica bifuncionalnog i elektronskog efekta. Ispitana <strong>je</strong> i stabilnosti katalizatora<br />

korišćen<strong>je</strong>m hronoamerimetrijskih merenja koja <strong>je</strong> pokazala da <strong>je</strong> Pt-Rh-Sn stabilniji<br />

od Pt/C katalizatora.<br />

Rad M23/2 U ovom radu oksidacija mravl<strong>je</strong> kiselina ispitivana <strong>je</strong> na dva Pt-Bi<br />

katalizatora; Pt2Bi i polikristalnj Pt modifikovanoj ireverzibilno adsorbovanim Bi (Pt /<br />

Biirr) kako bi se uspostavila razlika između dejstva Biirr i Bi u obliku legure. Rezultati<br />

su predstavl<strong>je</strong>ni u odnosu na čistu Pt. Utvrđeno <strong>je</strong> da su oba bimetalna katalizatora<br />

aktivnija od čiste Pt, sa početnim potencijalima pomerenim ka negativnijim<br />

vrednostima i strujom većom do dva reda veličina (na 0,0 V vs SCE pod stacionarnim<br />

uslovima). Poreklo visoke aktivnosti i stabilnosti Pt2Bi katalizatora <strong>je</strong> povećana<br />

selektivnost prema direktnom putu, dehidrogenaciji mravl<strong>je</strong> kiseline uzrokovane<br />

ensemble efektom i elektroniskom modifikacijom koja dovodi do suzbijanja<br />

izluživanja Bi sa površine tokom reakci<strong>je</strong>. Međutim, iako Pt / Biirr takođe ispoljava<br />

izuzetnu početnu aktivnost u odnosu na čistu Pt, izluživan<strong>je</strong> Bi ni<strong>je</strong> inhibirano pa <strong>je</strong><br />

trovan<strong>je</strong> površine katalizatora izazvano dehidratacijom prisutno. Poređen<strong>je</strong>m<br />

inicijalnog kvazi-stabilnog stanja i potenciodinamičkih rezultata za ova dva Pt-Bi<br />

katalizatora može se zaključiti da elektronski efekat, prisutan kod legure, doprinosi<br />

rani<strong>je</strong>m početku reakci<strong>je</strong>, dok <strong>je</strong> maksimalna gustina stru<strong>je</strong> određena ensemble<br />

efektom.


Rad M23/4 Koloidne nanočestice pripreml<strong>je</strong>ne korišćen<strong>je</strong>m metode sinteze iz<br />

rastvora sa preciznom kontrolom nad parametrima kao što su veličina čestica, oblik,<br />

sastav i struktura su pokazali veliki potencijal za katalitičke aplikaci<strong>je</strong>. Međutim,<br />

ovakve koloidne nanočestice su obično prekrivene sa organskim ligandima i ne mogu<br />

se direktno koristiti kao katalizatori. U ovom radu ispitan <strong>je</strong> efekat površinskog<br />

uklanjanja hemijski vezanih organskih materija na elektrokatalitička svojstva Pt<br />

nanočestica dobi<strong>je</strong>nih organskom sintezom iz rastvora. Korišćene su raznovrsne<br />

metode za uklanjan<strong>je</strong> oleylamine surfaktanta, kao što su: žaren<strong>je</strong>, pran<strong>je</strong> sirćetnom<br />

kiselinom i UV- zračen<strong>je</strong>. Ovako tretirane nanočestice su testirane kao katalizatori za<br />

reakciju izdvajanja kiseonika (Orr). Utvrđeno <strong>je</strong> da su katalitičke performanse ko<strong>je</strong><br />

podrazumevaju elektrohemijski aktivnu površinu i katalitičku aktivnost direktno<br />

zavisne od pred tretmana. Od metoda ko<strong>je</strong> su u ovom radu korišćene žaren<strong>je</strong> do<br />

temperature od 185 ° C na vazduhu se pokazalo kao na<strong>je</strong>fikasni<strong>je</strong> za čišćen<strong>je</strong> površina<br />

bez izazivanja promena u pogledu veličine i morfologi<strong>je</strong> čestica.<br />

Rad M23/5 U ovom radu oksidacija mravl<strong>je</strong> kiseline ispitivana <strong>je</strong> na sveže<br />

pripreml<strong>je</strong>nom Pt2Bi katalizatoru koji <strong>je</strong> okarakterisan korišćen<strong>je</strong>m difrakci<strong>je</strong> X-zraka<br />

(fazni sastav), skenirajuće tunelirajuće mikroskopi<strong>je</strong> (STM) (morfologija poršine) i<br />

oksidacijom adsorbovanog sloja CO (sastav površine). Rezultati karakterizaci<strong>je</strong> su<br />

pokazali da se Pt2Bi katalizatora sastoji od dve faze-55% PtBi legure i 45% Pt. Pt2Bi<br />

katalizator posedu<strong>je</strong> visoku aktivnost i stabilnost u reakciji oksidaci<strong>je</strong> mravl<strong>je</strong> kiseline<br />

favorizovan<strong>je</strong>m direktnog puta,dehidrogenaci<strong>je</strong>, kao posledice ensemble efekta.<br />

Visoka stabilnost Pt2Bi površine <strong>je</strong> posledica suzbijanja izluživanja Bi što se vidi na<br />

osnovu minimalne promene površinske morfologi<strong>je</strong> prikazanimna STM slikama pre i<br />

posle elektrohemijskih tretmana u mravljoj kiselini. Pt2Bi se pokazao kao moćan<br />

katalizator sa gustinama stru<strong>je</strong> većim za dva reda veličina na 0,0 V i potencijalom<br />

pomerenim za ~ 0,2 V negativni<strong>je</strong> u odnosu na čistu Pt.


IV. PRILOG<br />

Prilog 1. Kopi<strong>je</strong> separata radova i saopštenja<br />

Prilog 2. kopija potvrde o citiranosti naučnih radova<br />

Prilog 3.<br />

- potvrda o prijavl<strong>je</strong>nom patentu<br />

- poziv za predavan<strong>je</strong><br />

- dokaz o rukovođenju podpro<strong>je</strong>ktom<br />

- dokaz o učestvovanju na izradi magistarskih teza i doktorskih disertacija<br />

Prilog 4. Predlog rezimea izveštaja o kandidatu za stican<strong>je</strong> naučnog zvanja<br />

naučni savetnik


Др. Душан Трипковић<br />

СПИСАК НАУЧНИХ РАДОВА И САОПШТЕЊА<br />

(2005. – 2012.)<br />

(радови и саопштења од предходног избора у звање су означени *)<br />

РАДОВИ ОБЈАВЉЕНИ У НАУЧНИМ ЧАСОПИСИМА МЕЂУНАРОДНОГ<br />

ЗНАЧАЈА (М20):<br />

Радови у врхунским међународним часописима (М21):<br />

1. J.D. Lović, A.V. <strong>Tripković</strong>, S.Lj. Gojković, K.Đ. Popović, D.V. <strong>Tripković</strong>, P.<br />

Olszewski, A. Kowal, "Kinetic study of formic acid oxidation on carbon-supported<br />

platinum electrocatayst", J. Electroanal. Chem., 581 (2005) 294.<br />

[IF=2.223 (2005); oblast: Chemistry,analytical, 17/70]<br />

2. Vladislava M. Jovanovic, Dusan Tripkovic, Amalija Tripkovic, Andrzej Kowal, Jerzy<br />

Stoch: “Oxidation of formic acid at platinum electrodeposited on polished and<br />

oxidized glassy carbon”, Electrochem. Comm., 7, 1039-1044, (2005)<br />

[IF: 3.388 (2005); oblast: Electrochemistry, 2/21]<br />

3. A. <strong>Tripković</strong>, S. Gojković, K. Popović, J.D. Lović, A. Kowal, "Study of the kinetics<br />

and influence of Biir on formic acid oxidation on Pt2Ru3 / C", Electrochim. Acta,53<br />

(2007) 887.<br />

[IF: 2.848 (2007); oblast: Electrochemistry, 7/23]<br />

4. Strmcnik D. S., Rebec P., Gaberscek M., Tripkovic D., Stamenkovic, V., Lucas C. and<br />

Markovic N. M.: “Relationship between the Surface Coverage of Spectator Species<br />

and the Rate of Electrocatalytic Reactions“, J. Phys Chem.C. 111, 18672-18678,<br />

(2008)<br />

[IF: 3.396 (2008); oblast: Chemistry, Physical, 28/113]


5. D. <strong>Tripković</strong>, S. Stevanović, A. <strong>Tripković</strong>, A. Kowal and V.M. Jovanović: “Strustural<br />

effect in electrocatalysis: Formic acid oxidation on Pt electrodeposited on glassy<br />

carbon support“, J. Electrochem. Soc., 155, B281-289, (2008)<br />

[IF: 2.387 (2006); oblast: Electrochemistry, 6/22]<br />

6. D. V. Tripkovic, D. Strmcnik, D. van der Vliet, V. Stamenkovic and N. M. Markovic:<br />

“The role of anions in surface electrochemistry“, Faraday Discuss., 140 (2008) 25-40<br />

[IF: 4.604 (2008); oblast: Chemistry, Physical, 19/113]<br />

7. D. Strmcnik, D. Tripkovic, D. van der Vliet, V. Stamenkovic and N.M. Markovic:<br />

“Adsorption of hydrogen on Pt(1 1 1) and Pt(1 0 0) surfaces and its role in the HOR“,<br />

Electrochem. Communications, 10, 1602-1605, (2008)<br />

[IF: 4.194 (2008); oblast: Electrochemistry, 2/22]<br />

8. D. Strmcnik, D. Tripkovic, D. van der Vliet, K. C. Chang, V. Komanicky, H. You, G.<br />

Karapetrov, J. Greeley, V. Stamenkovic and N. M. Markovic: “Unique activity of<br />

platinum ad-islands in the CO electrooxidation reaction“, J. American Chem. Soc.,<br />

130 (46), 15332–15339, (2008)<br />

[IF: 8.091 (2008); oblast: Chemistry, Multidisciplinary, 7/127]<br />

9. Sanja Stevanović, Vladimir Panić, <strong>Dušan</strong> <strong>Tripković</strong>, and Vladislava M. Jovanović:<br />

''Promoting effect of carbon functional groups in methanol oxidation on supported Pt<br />

catalyst'', Electrochem. Comm., 11, 18-21, (2009)<br />

[IF: 4.243 (2009); oblast: Electrochemistry, 2/24]<br />

10. A.V. <strong>Tripković</strong>, K.Dj. Popović, J.D. Lović, V.M. Jovanović, S.I. Stevanović, D.V.<br />

<strong>Tripković</strong> and A. Kowal " Promotional effect of Snad on the ethanol oxidation at<br />

Pt3Sn/C catalyst" Electrochem. Comm., 11 (2009) 1030.<br />

[IF: 4.243 (2009); oblast: Electrochemistry, 2/24]


11. *Bostjan Genorio, Dusan Strmcnik, Ram Subbaraman, Dusan Tripkovic, Goran<br />

Karapetrov, Vojislav R. Stamenkovic, Stane Pejovnik and Nenad M. Markovic<br />

“Selective catalyst for the hydrogen oxidation and oxygen reduction reactions by<br />

patterning of platinum with calix[4]arene molecules” Nature Materials (12), 998-<br />

1003, (2010)<br />

[IF: 29.920 (2010); oblast: Chemistry, Physical, 1/127]<br />

12. *Bostjan Genorio, Ram Subbaraman, Dusan Strmcnik, Dusan Tripkovic, Vojislav R.<br />

Stamenkovic, and Nenad M. Markovic “ Tailoring the selectivity and stability of<br />

chemically modified platinum nanocatalysts to design highly durable anodes for PEM<br />

fuel cells” Angewandte Chemie, 50(24), 5468-72, (2011)<br />

[IF: 13.455 (2011); oblast: Chemistry, Multidisciplinary, 7/154<br />

13. *Chao Wang, Miaofang Chi, Dongguo Li, Dusan Strmcnik, Dennis van der<br />

Vliet,Guofeng Wang,Vladimir Komanicky, Kee-Chul Chang, Arvydas P. Paulikas,<br />

Dusan Tripkovic, John Pearson,Karren L. More, Nenad M. Markovic, and Vojislav R.<br />

Stamenkovic ” Design and Synthesis of Bimetallic Electrocatalyst with Multilayered<br />

Pt-Skin Surfaces” Journal of the American Chemical Society,133(36),14396-403,<br />

(2011)<br />

[IF: 9.907 (2011); oblast: Chemistry, Multidisciplinary, 11/154]<br />

14. *Ram Subbaraman, Dusan Tripkovic, Dusan Strmcnik, Kee-Chul Chang, Masanobu<br />

Uchimura, Arvydas P. Paulikas, Vojislav Stamenkovic, Nenad M. Markovic<br />

“Enhancing Hydrogen Evolution Activity in Water Splitting by Tailoring Li + -<br />

Ni(OH)2-Pt Interfaces”, Science, vol. 334 no. 6060 pp. 1256-1260, (2011)<br />

[IF: 31.201 (2011); oblast: Multidisciplinary Sciences, 2/56]<br />

15. *Dennis van der Vliet, Chao Wang, Dongguo Li, Arvydas P. Paulikas,Jeffrey<br />

Greely,Rees B. Rankin, Dusan Strmcnik, Dusan Tripkovic, Nenad M. Markovic, and<br />

Vojislav R. Stamenkovic “Unique Electrochemical Adsorption Properties of Pt-skin<br />

Surfaces”, Angewandte Chemie,Vol 51, Issue 13, 3139–3142, (2012)<br />

[IF: 13.455 (2011); oblast: Chemistry, Multidisciplinary, 7/154<br />

16. *Ram Subbaraman, Dusan Tripkovic,Kee-Chul Chang,Dusan Strmcnik, Arvydas<br />

P.Paulikas,Pussana Hirunsit, Maria Chan, Jeff Greeley, Vojislav Stamenkovic and<br />

Nenad M.Markovic” Trends in activity for the water electrolyser reactions on 3dM<br />

(Ni,Co,Fe,Mn) hydr(oxy)oxide catalysts”,Nature Materials 11, (6), 550-557 (2012)<br />

DOI: 10.1038/nmat3313<br />

[IF: 32.841 (2011); oblast: Chemistry, Physical, 1/134]


17. *Dennis van der Vliet, Chao Wang, Dusan Tripkovic, Dusan Strmcnik, Xiao Zhang,<br />

Mark Debe, Radoslav Atanososki, Nenad Markovic and Vojislav Stamenkovic<br />

“Mesostructured Thin Films as Electrocatalysts with Tunable Composition and<br />

Surface Morphology”, Nature Materials, (2012), DOI: 10.1038/nmat3457<br />

[IF: 32.841 (2011); oblast: Chemistry, Physical, 1/134]<br />

18. *Ram Subbaraman, N. Danilovic, P. P. Lopes, D. Tripkovic, D. Strmcnik, V. R.<br />

Stamenkovic, and N. M. Markovic “Origin of Anomalous Activities for<br />

Electrocatalysts in Alkaline Electrolytes”, J. Phys. Chem. C, 116 (42), pp 22231–<br />

22237 (2012), DOI: 10.1021/jp3075783<br />

Радови у истакнутим међународним часописима (М22):<br />

[IF: 4.805 (2011); oblast: Chemistry, Physical, 26/134]<br />

1. *Stevanovic I. Sanja, Tripkovic V. Dusan, Rogan R. Jelena, Popovic Dj. Ksenija,<br />

Lovic D. Jelena, Tripkovic V. Amalija and Jovanovic M Vladislava, “Microwaveassisted<br />

polyol synthesis of carbon-supported platinum-based bimetallic catalysts for<br />

ethanol oxidation”, Journal of Solid State Electrochemistry, vol. 16 br. 10, str. 3147-<br />

3157, (2012)<br />

Радови у међународним часописима (М23):<br />

[IF: 2.234 (2010); oblast: Electrochemistry, 13/26]<br />

1. D. <strong>Tripković</strong>, V. Rado<strong>je</strong>vić, R. Aleksić: “Factors affecting the microstructure of porous<br />

ceramics“, J. Serb. Chem. Soc, 71 (3), 277-284, (2006)<br />

[IF=0.522 (2004); oblast: Chemistry – multidisciplinary, 85/124]<br />

2. A. Kowal, P. Olszewski, D. <strong>Tripković</strong>, R. Stevanović: “Nanoscale topography of<br />

GC/Pt-C and GC/Pt-Ru-C electrodes studied by means of STM, AFM and XRD<br />

methods“, Material Science Forum (RECENT DEVELOPMENTS IN ADVANCED<br />

MATERIALS AND PROCESSES), 518, 271-275, (2006)<br />

[IF=0.498 (2004); oblast: Materials Science – multidisciplinary, 137/178]


3. J.D.Lović, S. Lj. Gojković, K.Đ. Popović, D.V. <strong>Tripković</strong>, A.V. <strong>Tripković</strong>, "Structural<br />

effects in electrocatalysis: formic acid oxidation at model and real Pt catalysts"<br />

Mat.Sci.Forum, 518, (2006) 259.<br />

[IF=0.498 (2004); oblast: Materials Science – multidisciplinary, 137/178]<br />

4. Sanja Terzić, <strong>Dušan</strong> <strong>Tripković</strong>, Vladislava M. Jovanović, Amalija <strong>Tripković</strong> and<br />

Andrzej Kowal: “Effect of glassy carbon properties on electrochemical deposition of<br />

platinum nano-catalyst and its activity for methanol oxidation”, J. Serb. Chem. Soc., 72<br />

(2), 165-181, (2007)<br />

[IF=0.536 (2007); oblast: Chemistry – multidisciplinary, 95/127]<br />

5. S. Stevanović, D. <strong>Tripković</strong>, A. Kowal, D. Minić, V. M. Jovanović and A. <strong>Tripković</strong>:<br />

“Influence of surface morphology on methanol oxidation at glassy carbon supported Pt<br />

catalyst”, J. Serb. Chem. Soc., 73, 845-859, (2008)<br />

[IF=0.611 (2009); oblast: Chemistry – multidisciplinary, 91/127]<br />

6. *S. Stevanovic, D. Tripkovic, J. Rogan, D. Minic, A. Gavrilovic, A. Tripkovic and V.<br />

M. Jovanovic: “Enhanced Activity in Ethanol Oxidation of Pt3Sn Electrocatalysts<br />

Synthesized by Microwave Irradiation”, Russ. J. Phys. Chem., 85 (13), 2299-2304,<br />

(2011)<br />

[IF=0.459 (2011); oblast: Chemistry, Physical, 125/134]<br />

7. *S. Stevanovic, D. Tripkovic, D. Poleti, J. Rogan, A. Tripkovic and V.M.Jovanovic:<br />

“Microwave Synthesis and Characterization of PT and Pt-Rh-Sn Electrocatalyst for<br />

Ethanol Oxidation”, J. Serb. Chem. Soc., 76 (12), 1673-1685, (2011)<br />

[IF= 0.879 (2011); oblast: Chemistry – multidisciplinary, 103/154]<br />

8. *J.D.Lović, D.V. <strong>Tripković</strong>, K.Dj. Popović, V.M. Jovanović, A.V. <strong>Tripković</strong>,<br />

“Electrocatalytic properties of Pt-Bi electrodes towards the electrooxidation of formic<br />

acid”, J.Serb.Chem.Soc., DOI: 10.2298/JSC121012138L<br />

[IF=0.879 (2011); oblast: Chemistry – multidisciplinary, 102/152]


9. *Dongguo Li, Chao Wang, Dusan Tripkovic, Shouheng Sun, Nenad M. Markovic, and<br />

Vojislav R. Stamenkovic “Surfactant Removal for Colloidal Nanoparticles from<br />

Solution Synthesis: The Effect on Catalytic Performance”, ACS Catalysis, 2 (7), pp<br />

1358–1362 DOI:10.1021/cs300219j, (2012)<br />

[IF dobija 2013; oblast: Chemistry - Physical, 130/134]<br />

10. *J.D. Lović, M.D. Obradović, D.V. <strong>Tripković</strong>, K.Dj. Popović, V.M. Jovanović,<br />

S.Lj. Gojković, A.V. <strong>Tripković</strong> “High activity and stability of Pt2Bi catalyst in formic<br />

acid oxidation”, Springer Electrocatalysis, DOI: 10.1007/s12678-012-0099-9<br />

[IF dobija 2013; oblast: Electrochemistry]<br />

УКУПНИ М=205.5 УКУПНИ ИФ=215.335<br />

ЗБОРНИЦИ МЕЂУНАРОДНИХ НАУЧНИХ СКУПОВА (М30):<br />

Саопштење са међународног скупа штампано у целини (М33):<br />

1. * S. Stevanović, D. <strong>Tripković</strong>, J. Rogan, D. Minić, A. Gavrilović, A.<strong>Tripković</strong> and V.M.<br />

Jovanović: „Microwave assisted synthesis of Pt and Pt3Sn electrocatalysts for<br />

ethanol oxidation“ , 10th International Conference on Fundamental and Applied<br />

Aspects of Physical Chemistry, Proceedings E-O-2, Belgrade, Serbia, 21-24 September<br />

2010.


Саопштење са међународног скупа штампано у изводу (М34):<br />

1. D.<strong>Tripković</strong>, S.Terzić, V. M. Jovanović and A. Kowal: ″ Characterisation of<br />

Platinum nanoparticles electrochemically deposited on glassy carbon″, The seventh<br />

Yugoslav Matherials Research Conference, The Book of Abreacts, P.S.A.8, page 94,<br />

September 2005, Herceg-Novi, Yugoslavia<br />

2. D. <strong>Tripković</strong>, V. Rado<strong>je</strong>vić, R. Aleksić : “Factors affecting the microstructure of<br />

porous ceramics“, The seventh Yugoslav Materials Research Conference, The Book of<br />

Absreacts, P.S.A.8, page 71, September 2005, Herceg-Novi, Yugoslavia<br />

3. S. Terzić, D. <strong>Tripković</strong>, A. <strong>Tripković</strong>, A. Kowal and V. M. Jovanović: “Methanol<br />

oxidation at glassy carbon supported Pt catalyst in acid and alkaline solution –<br />

effect of support activation”, 56th Meeting of International Society of<br />

Electrochemistry, Bussan (Korea), September 2005<br />

4. S. Terzić, D. <strong>Tripković</strong>, A. <strong>Tripković</strong>, A. Kowal and V. M. Jovanović: “The effect of<br />

electrochemically treated glassy carbon on the activity of supported Pt catalyst”,<br />

56th Meeting of International Society of Electrochemistry, Bussan (Korea), September<br />

2005<br />

5. V. M. Jovanović, D. <strong>Tripković</strong> and A. <strong>Tripković</strong>: “Formic acid oxidation on glassy<br />

carbon supported platinum nanoparticles – structural effect“, 57th Meeting of<br />

International Society of Electrochemistry, Edinburgh (United Kingdom), August 2006,<br />

Book of Abstracts S10-P-24<br />

6. D. <strong>Tripković</strong>, S. Terzić, V. M. Jovanović and A. Kowal: “Electrocatalytic activity of<br />

platinum nanoparticles on oxidized support“, 57th Meeting of International Society<br />

of Electrochemistry, Edinburgh (United Kingdom), August 2006, Book of Abstracts S4-<br />

P-48<br />

7. V. Stamenkovic, C. Lucas, D. Tripkovic, D. Strmcnik and N. M. Markovic“In-Situ<br />

SXS/STM Characterization of Temperature Controlled ElectrifiedSolid-Liquid<br />

Interfaces” 210th Int. Meeting of the Electrochemical Society, November 2006,<br />

Cancun, Mexico


8. V. Stamenkovic, C. A. Lucas, D. Tripkovic, D. Strmcnik, K. C. Chang, H. You, N. M.<br />

Markovic “In-Situ characterization Temperature Controlled Electrified Solid-<br />

Liquid Interfaces”233rd American Chemical Society National Meeting, March 2007,<br />

Chicago, IL, USA<br />

9. D. Strmcnik, D. Tripkovic, D.van der Vliet, J. Greeley, V. Stamenkovic, N. M.<br />

Markovic “Active Sites for Fuel Cell reactions: Experiments and Theory” 11th<br />

International Conference on Electrified Interfaces, June 2007, Sapporo, Japan.<br />

10. V. Stamenkovic, D. Strmcnik, D. Tripkovic, D. van der Vliet, H.You, N.<br />

Markovic“Nanocatalysts Engineering on Extended and Nanoscale Surfaces”9th<br />

International Conf. of the Yugoslav Materials Research Society, September 2007,<br />

Herceg Novi, Montenegro.<br />

11. D. Tripkovic, D. Strmcnik, D. van der Vliet, V. Stamenkovic and N. M. Markovic<br />

“Active sites in the CO oxidation on Pt(111)” 59th Annual Meeting of the<br />

International Society of Electrochemistry, September 2008, Seville, Spain<br />

12. D. <strong>Tripković</strong>, S. Stevanović, Amalija <strong>Tripković</strong>, A. Kowal and V. M. Jovanović<br />

“Structural effect in electrocatalysis:Formic acid oxidation on Pt electrodeposited<br />

on glassy carbon support“, First Regional Symposium on Electrochemistry of South-<br />

East Europe, Crveni Otok, Rovinj, Istria, Croatia, May 4-8, 2008.<br />

13. D.<strong>Tripković</strong> “The influence of surface morphology of platinum materials on<br />

electrocatalytic activity toward CO oxidation” Invited lecture at Center for Atomic<br />

Scale Materials Designee, Technical University of Denmark, Department of Physics,<br />

February 2009, Copenhagen, Denmark<br />

14. *Dusan Strmcnik, Dusan Tripkovic, Dennis Van der Vliet, Jeffrey P. Greeley,<br />

Alexander Brownrigg, Christopher Lucas, Goran Karapetrov, Vojislav Stamenkovic and<br />

Nenad Markovic „Active Sites for PEM Fuel Cell Reactions“, 216th ECS Meeting,<br />

Abstract #868, Vienna, Austria, Oct 6 2009


15. *Dusan Strmcnik, Kensaku Kodama, Dusan Tripkovic, Dennis Vander Vliet, Chao<br />

Wang, Vojislav Stamenkovic and Nenad M. Markovic: “Design Catalytic Properties<br />

of Electrochemical Interfaces”, 217 th ECS Meeting, Abstract #1755, Vancouver,<br />

Canada, Apr 27 2010<br />

16. *Chao Wang, Dusan Strmcnik, Dusan Tripkovic, Dennis Van der Vliet, Nenad<br />

Markovic and Vojislav R. Stamenkovic ” Electrocatalysis on Well-Defined Solid-<br />

Liquid Interfaces”, 219 th ECS Meeting, Abstract #1924, Montreal, QC, Canada, May 2<br />

2011<br />

17. *C. Wang, D. Li, D. van der Vliet, D. Strmcnik, D. Tripkovic, N. Markovic and V.<br />

Stamenkovic: ”Advanced Electrocatalysts: - From Extended to Nanoscale<br />

Surfaces”, 220th ECS Meeting, Abstract #947, Boston, MA, Oct 11 2011<br />

18. *S. Stevanović, D. <strong>Tripković</strong>, V. <strong>Tripković</strong>, D. Minić, A.Gavrilović, K. Popović, A.<br />

<strong>Tripković</strong> and V.M. Jovanović: „Insight of Sn influence on formic acid oxidation at<br />

Pt based catalysts“, 63rd Meeting of International Society of Electrochemistry, Prague<br />

(Czech Republic), August 2012, Book of Abstracts S05-027<br />

19. *J. D. Lović, M. D. Obradović, D.V. <strong>Tripković</strong>, K. Dj. Popović, V. M. Jovanović, S<br />

.Lj. Gojković and A.V. <strong>Tripković</strong>: „High Activity and Stability of Pt2Bi Catalyst in<br />

Formic Acid Oxidation“, 63rd Meeting of International Society of Electrochemistry,<br />

Prague (Czech Republic), August 2012, Book of Abstracts S05-063<br />

20. *S. Stevanović, D. <strong>Tripković</strong>, J. Rogan, J. Lović, K. Popović, A. <strong>Tripković</strong> and V.M.<br />

Jovanović: „Ethanol oxidation on carbon supported platinum based bimetallic<br />

catalysts synthesized by microwave assisted polyol procedure“, 63rd Meeting of<br />

International Society of Electrochemistry, Prague (Czech Republic), August 2012, Book<br />

of Abstracts S05-052<br />

21. *Dusan Strmcnik, Ramachandran Subbaraman, Dusan Tripkovic, Kee-Chul Chang,<br />

Nemanja Danilovic, Dennis F. Van der Vliet, Pietro Lopes, Arvydas Paul Paulikas,<br />

Vojislav R. Stamenkovic and Nenad M. Markovic: ” Controlling Reactivity of<br />

Electrochemical Interfaces by Tuning Non-covalent Interactions”, 221 th ECS<br />

Meeting, Abstract #1560, Seattle, WA, May 7 2012


22. *Dusan Strmcnik, Dusan Tripkovic, Ramachandran Subbaraman, Nemanja Danilovic,<br />

Dennis F. Van der Vliet, Arvydas Paul Paulikas, Vojislav R. Stamenkovic and Nenad<br />

M. Markovic” Fundamental investigations of precious metal stability in energy<br />

conversion systems”, 221 th ECS Meeting, Abstract #1532, Seattle, WA, May 10 2012<br />

23. *Ramachandran Subbaraman, Dusan Tripkovic, Dusan Strmcnik, Gustav K. Wiberg,<br />

Jakub S. Jirkovsky, Chao Wang, Vojislav R. Stamenkovic and Nenad M. Markovic: ”<br />

Electrochemical Interfaces for Energy Conversion and Storage”, 221 th ECS<br />

Meeting, Abstract #549, Seattle, WA, May 7 2012<br />

24. *Nemanja Danilovic, Ramachandran Subbaraman, Dusan Strmcnik, Dusan Tripkovic,<br />

Kee-Chul Chang, Arvydas Paul Paulikas, Vojislav R. Stamenkovic, Debbie J. Myers<br />

and Nenad M. Markovic „Electrocatalysts for the Oxygen Evolution Reaction“, 221 th<br />

ECS Meeting, Abstract #1510, Seattle, WA, May 9 2012<br />

25. *Dusan Tripkovic, Dusan Strmcnik, Dennis F. Van der Vliet, Chao Wang, Nenad M.<br />

Markovic and Vojislav R. Stamenkovic: “In-situ Infrared Spectroscopy at Solid-<br />

Liquid Interfaces as a Tool for Evaluation of Nanoscale Surface Morphology”,<br />

221 th ECS Meeting, Abstract #1570, Seattle, WA, May 9 2012<br />

26. *Chao Wang, Dennis Van der Vliet, Dusan Tripkovic, Dusan Strmcnik, Dongguo Li,<br />

Nenad M. Markovic and Vojislav R. Stamenkovic: ”Advanced Electrocatalysts for<br />

PEM Fuel Cells”, PRiME 2012., Abstract 1650 Honolulu, Hawaii October 7-12, 2012<br />

27. *R. Subbaraman, D. Tripkovic, G. K. Wiberg, J. S. Jirkovsky and N. M. Markovic:”<br />

Li-Air- Just a <strong>Dr</strong>eam or Future Reality”, DOE-Chine Meeting-Beyond Li-ion systems,<br />

Boston, MS, September 2012, (Invited)<br />

28. *D. Tripkovic, D. Strmcnik, R. Subbaraman, D. V. Stamenkovic and N. M. Markovic:<br />

“Tailored Nanomaterials for Clean Energy Conversion and Storage”, International<br />

Symposium, Tsukuba, Japan, March 2012 (Invited)


РАДОВИ ОБЈАВЉЕНИ У ЧАСОПИСИМА НАЦИОНАЛНОГ ЗНАЧАЈА (М50):<br />

Радови у водећим часописима националног значаја (М51)<br />

1. S. Terzić, V. M. Jovanović, D.<strong>Tripković</strong>, A. Kowal, J. Stoch: “Elektrohemijska,<br />

mikroskopska i spektroskopska karakterizacija platine nataložene na staklasti ugl<strong>je</strong>nik“,<br />

Hemijska industrija, 61 (3), 135-141, (2007)<br />

[IF=0.177 (2009); oblast: Engineering, Chemical,118/127]<br />

ЗБОРНИЦИ СКУПОВА НАЦИОНАЛНОГ ЗНАЧАЈА (М60):<br />

Саопштење на скупу националног значаја штампано у целини (М63):<br />

Остварени<br />

М<br />

коефицјент<br />

М коефицјент од избора у звање<br />

виши научни сарадник<br />

М20 M21: 8 x 8 = 64<br />

M22: 1 x 5 = 5<br />

M23: 3 x 5 = 15<br />

М30 M33: 1 x 1 = 1<br />

M34: 15 x 0,5 = 7,5<br />

Укупни М коефицјент<br />

M21: 18 x 8 = 144<br />

M22: 1 x 5 = 5<br />

M23: 10 x 3 = 30<br />

M33: 1 x 1 = 1<br />

M34: 28 x 0,5 = 14<br />

М50 M51: 1 x 2 = 2<br />

М60 M63: 1 x 0,5 = 0.5<br />

М70 М71: 1 x 6 = 6<br />

М72: 1 x 3 = 3<br />

Укупно: 92,5 Укупно: 205,5


PIB: 100160355<br />

Matični broj: 07805497<br />

Šifra delatnosti: 073101<br />

UNIVERZITET U BEOGRADU<br />

IHTM INSTITUT ZA HEMIJU TEHNOLOGIJU I METALURGIJU<br />

CEH CENTAR ZA ELEKTROHEMIJU<br />

N<strong>je</strong>goševa 12, 11001 Beograd, Srbija<br />

Kao komentor magistarske teze Mr. San<strong>je</strong> Stevanović (Terzić) pod nazivom<br />

"Oksidacija metanola na platinskim nano-česticama elektrohemijski nataloženim na<br />

staklasti ugl<strong>je</strong>nik" odbran<strong>je</strong>ne 2007. god. na Fakultetu za fizičku hemiju u Beogradu i kao<br />

rukovodilac podpro<strong>je</strong>kta u okviru koga se izvodi doktorska disertacija koleginice Stevanović<br />

na temu ″Sinteza i karakterizacija platinskih legura za anodne reakciju u gorivim<br />

spregovima″ prihvaćena i odobrena 2009. god. (odbrana se oćeku<strong>je</strong> do leta 2013. god.)<br />

potvrdju<strong>je</strong>m da <strong>je</strong> <strong>Dr</strong>. <strong>Dušan</strong> <strong>Tripković</strong> učestvovao u rukovod<strong>je</strong>nju i izradi delova obe teze<br />

koji se odnose na karakterizaci<strong>je</strong> površina ispitivanih katalizatora (čestica) koriščen<strong>je</strong>m AFM<br />

i STM tehnika kao i oksidaci<strong>je</strong> CO. Za<strong>je</strong>dnički radovi proizašli iz magistarske teze i<br />

doktorske disertaci<strong>je</strong> su:<br />

1. Sanja Terzić, <strong>Dušan</strong> <strong>Tripković</strong>, Vladislava M. Jovanović, Amalija <strong>Tripković</strong> and<br />

Andrzej Kowal: “Effect of glassy carbon properties on electrochemical deposition of<br />

platinum nano-catalyst and its activity for methanol oxidation”, J.Serb.Chem.Soc.,<br />

72 (2) (2007) 165-181<br />

2. S. Stevanović, D. <strong>Tripković</strong>, A. Kowal, D. Minić, V.M. Jovanović and A.<strong>Tripković</strong>:<br />

“Influence of surface morphology on methanol oxidation at glassy carbon<br />

supported Pt catalyst”, J.Serb.Chem.Soc., 73 (8-9) (2008) 845-859<br />

3. Sanja Stevanović, Vladimir Panić, <strong>Dušan</strong> <strong>Tripković</strong>, and Vladislava M. Jovanović:<br />

“Promoting effect of carbon functional groups in methanol oxidation on supported<br />

Pt catalyst“, Electrochem .Comm., 11 (2009) 18-21<br />

4. A.V. <strong>Tripković</strong>, K.Dj. Popović, J.D. Lović, V.M. Jovanović, S.I. Stevanović, D.V.<br />

<strong>Tripković</strong> and A. Kowal: “Promotional effect of Snad on the ethanol oxidation at<br />

Pt3Sn/C catalyst“, Electrochem. Comm. 11 (5) (2009) 1030-1033.<br />

5. S. Stevanović, D. <strong>Tripković</strong>, D. Poleti, J. Rogan, A. <strong>Tripković</strong> and V.M. Jovanović:<br />

“Microwave synthesis and characterization of Pt and Pt–Rh–Sn electrocatalysts<br />

for ethanol oxidation“, J. Serb. Chem. Soc. 76 (12) (2011)1673–1685<br />

6. S. Stevanović, D. <strong>Tripković</strong>, J. Rogan, D. Minić, A. Gavrilović, A. <strong>Tripković</strong> and V. M.<br />

Jovanović: “Enhanced Activity in Ethanol Oxidation of Pt3Sn Electrocatalysts<br />

Synthesized by Microwave Irradiation“, Russian Journal of Physical Chemistry A, 85<br />

(13) (2011) 2299–2304<br />

7. S. Stevanović, D. <strong>Tripković</strong>, J. Rogan, K. Popović and J. Lović, A. Tripkovic and<br />

V.M.Jovanovic: "Microwave-assisted polyol synthesis of carbon-supported<br />

platinum-based bimetallic catalysts for ethanol oxidation", J. Solid State<br />

Electrochemistry. 16 (2012) 3147-3157<br />

Beograd, 14.12.2012. god.<br />

Tel/Fax: (+38111 ) 3370-389, 3370-390<br />

<strong>Dr</strong>. Vladislava M. Jovanović<br />

Naucni savetnik IHTM-CEH


~<br />

Univerzitet uBeogradu<br />

Fakultet zafizicku hemiju<br />

Sanja Terzic<br />

OKSIDACIJA MET ANOLA NA PLATINSKIM<br />

NANO-CESTICAMA ELEKTROHEMIJSKI<br />

NATALOZENIM NA STAKLASTI UGLJENIK<br />

magistarska teza<br />

Beograd, jun 2007.


~<br />

Mentor:<br />

<strong>Dr</strong>. <strong>Dr</strong>agica Minie, red. Prof.,<br />

Fakultet za fizicku hemiju, Beograd<br />

Komentor: <strong>Dr</strong>. Vladislava Jovanovic,<br />

naucni savetnik IHTM-a<br />

Komisija: <strong>Dr</strong>. Nikola Vukeli6, van. Prof,<br />

Fakultet za fizicku hemiju, Beograd<br />

Datum odbrane: jun. 2007.


~<br />

{66//._._..<br />

. /!J J::<br />

/17- '. ,'t. (IJ. . Jv\<br />

Na osnovu clana 227. stay 2. Statuta Univerzitet u Beogradu - Fakulteta za fizicku<br />

hemiju, Nastavno-naucno vece Fakulteta, na I redovnoj sednici, odrzanoj 19.10.2006. <strong>godine</strong>,<br />

donosi sledecu<br />

1.- Prihvata se pozitivni izvestaj 0 predlogu teme za izradu magistarske teze i odobrava se<br />

tema za izradu magistarske teze studenta San<strong>je</strong> Terzic, diplomiranog fizikohemicara,<br />

istrafivaca-pripravnika, IHTM - Centar za elektrohemiju, pod naslovom: «Oksidacija<br />

metanola na platinskim nano-cesticama elektrohemijski natalozenim na staklasti ugl<strong>je</strong>nik»,<br />

Komisi<strong>je</strong> u sastavu:<br />

]) dr <strong>Dr</strong>agica Minic,redovni profesor, Fakultet za fizicku hemiju,<br />

(!J dr Vladislava Jovanovic, naucni savetnik, IHTM - Centar za elektrohemiju,<br />

3) dr Nikola Vukeli6, docent, Fakultet za fizicku hemiju.<br />

2.- Na zahtev studenta, za mentora za izradu eve magistarske teze odreduju se<br />

1) dr <strong>Dr</strong>agica Minie, redovni profesor, i 2) dr Vladislava Jovanovic, naucni savetnik.<br />

3.- Student maze braniti magistarsku tezu u roku od tri <strong>godine</strong> od danaodobrenja teme<br />

magistarske teze, a ovaj rok, na zahtev studenta, maze biti produzen, najduzeza pet godina od<br />

dana stupanja na snagu Zakonao visokom obrazovanju, odnosno do 11.9.2010. <strong>godine</strong>. Odluku 0<br />

produzenju, na predlog mentora, donosi Nastavno-naucno vece.<br />

Po uradenoj magistarskoj tezi, student podnosi Nastavno-naucnom vecu zahtev za<br />

odbranu teze idostavlja primerak teze.<br />

Odluku dostaviti:<br />

- studentu,<br />

- mentoru,<br />

- Arhivi Fakulteta.<br />

.,',<br />

. )<br />

ODLUKU<br />

'"<br />

,)


<strong>Dr</strong>. Nenad M Markovic<br />

Argonne National Laboratory<br />

Materials Science Division<br />

Phone: (630) 252-5181<br />

Email: nmmarkovic@anl.gov<br />

Dear Madam/Sir,<br />

December 14 , 2012<br />

This letter is to confirm that <strong>Dr</strong>. Dusan Tripkovic, employed at ICTM-Department of<br />

Electrochemistry, was visiting Argonne National Laboratory first as a student (2005-2007) and<br />

then as a postdoctoral fellow from 2009 to 2011. During his stay in my group <strong>Dr</strong>. Tripkovic was<br />

working on the development and design of surface sensitive probes (scanning tunneling<br />

microscopy, STM) and vibrational spectroscopy (infraread spectroscopy, IR) capable to provide<br />

information at electrochemical interfaces at atomic and molecular levels. In addition to doing his<br />

own experiments <strong>Dr</strong>. Tripkovic was actively involved in designing STM and IR experiments for<br />

both Dusan Tripovic and Dennis Van der Vliet. As outlined below, both Strmcinik and Van der<br />

Vliet used these results in their thesis to explain structure function relations in surface chemistry<br />

of CO oxidation reaction.<br />

1. In Strmcnik’s thesis, entitled “Active Sites for PEM Fuel Cell Reactions Model and Real<br />

Systems”, STM and IR results were used to find correlations between the number of active sites<br />

on Pt single crystal surfaces and surface structure of CO. <strong>Dr</strong>. Tripkovic’s role was to help<br />

Strmcing in designing, collecting and interpreting the STM and IR results.<br />

2. In Van der Vliet’s thesis, entitled “Oxygen Reduction on Pt-based Nanoparticle Catalyst – the<br />

experimental STM and IR results van der Vliet used to find relationships between the structure<br />

of Pt single crystal and the size of naoparticles in the case of the CO oxidation reaction. As in the<br />

case with Strimcnik, <strong>Dr</strong>. Tripkovic was carrying out the measurements related to both the<br />

determination of the surface morphology of the catalysts, using STM technique as well as the IR<br />

measurements which explained anomalous dependence of the rate of CO oxidation and the size<br />

of Pt nanoparticles. <strong>Dr</strong>. Tripkovic was also actively involved in the discussions with both<br />

Strmcnik and van der Vliet, giving important suggestions in finalizing the final draft of papers


which are produced based on their experimental results. From these collaboration 4 papers were<br />

published in high impact journals including:<br />

1. Strmcnik D. S., Rebec P., Gaberscek M., Tripkovic D., Stamenkovic, V., Lucas C. and<br />

Markovic N. M.: “Relationship between the Surface Coverage of Spectator Species and<br />

the Rate of Electrocatalytic Reactions“, J. Phys Chem.C. 111, 18672-18678, (2008)<br />

2. D. V. Tripkovic, D. Strmcnik, D. van der Vliet, V. Stamenkovic and N. M. Markovic: “The<br />

role of anions in surface electrochemistry“, Faraday Discuss., 140 (2008) 25-40<br />

3. D. Strmcnik, D. Tripkovic, D. van der Vliet, V. Stamenkovic and N.M. Markovic:<br />

“Adsorption of hydrogen on Pt(1 1 1) and Pt(1 0 0) surfaces and its role in the HOR“,<br />

Electrochem. Communications, 10, 1602-1605, (2008)<br />

4. D. Strmcnik, D. Tripkovic, D. van der Vliet, K. C. Chang, V. Komanicky, H. You, G.<br />

Karapetrov, J. Greeley, V. Stamenkovic and N. M. Markovic: “Unique activity of platinum<br />

ad-islands in the CO electrooxidation reaction“, J. American Chem. Soc., 130 (46), 15332–<br />

15339, (2008)<br />

Sincerely,<br />

<strong>Dr</strong>. Nenad M. Markovic<br />

Senior Staff Scientist


Univerza v Ljubljani<br />

Fakulteta za kemijo kemijsko tehnologijo<br />

I<br />

b .~r.;.t<br />

I. :~:I\-1";j:::,~<br />

ACTIVE,SITES FOR PEM FUEL CELL<br />

REACTIONS IN MODEL AND REAL<br />

SYSTEMS<br />

AKTIVNA MESTA ZA REAKCIJE V PEM GORIVNIH<br />

CELICAH V MODELNIH IN REALNIH SISTEMIH<br />

Ph.D. DISSERTATION<br />

DOKTORSKA DISER T ACIJA<br />

Dusan Strmcnik


~<br />

Ljubljana, October 2007<br />

My work was performed under mentorship of:<br />

<strong>Dr</strong>. Nenad Markovic,<br />

Argonne National Laboratory, Chicago, Illipois, USA<br />

and<br />

Doc.<strong>Dr</strong>. Miran Gaberscek<br />

National Institute of Chemistry, Ljubljana, Slovenia


~<br />

Fuel Cell Electrocatalysis<br />

Oxygen Reductional" .Pt-based Nanoparticle<br />

Catalysts<br />

Proefscllrift<br />

ter verkrijging van<br />

de graad van Doctor aan de Universiteit Leiden,<br />

op gezag van Rector Magnificus Prof. Mr. P.F. van der Heijden,<br />

volgens besluit van bet College veer Promoties<br />

te verdedigen op dinsdag 21. september 2010<br />

klokke 16.15 Our<br />

door<br />

Dennis Franciscus van de}" Vliet<br />

geboren te Tilbmg in ] 98]


~<br />

Promotiecommissie:<br />

Promotor:<br />

CoproJnotor:<br />

Overige Lcden;<br />

Prof. <strong>Dr</strong>. M. T. M. Koper<br />

<strong>Dr</strong>. N. M. Markovic (Argonne National Lab, USA)<br />

Prof. <strong>Dr</strong>. J. Brouwer<br />

Prof. <strong>Dr</strong>. B. E. Nieuwel11mys<br />

Prof. <strong>Dr</strong>, ], W, N, Frenken<br />

Prof. <strong>Dr</strong>. G. A. Attard (University of Cardiff, UK)<br />

Prof. <strong>Dr</strong>. J. A. R. van Veen (Shell)<br />

<strong>Dr</strong>. N. P. Lebedeva (ECN)


~<br />

....<br />

0 ~ ~ ~ m ~ AWN ....<br />

------------------------------<br />

~ ~ ~ ~ ~ ~ ~ ~ ~ ~<br />

000 0 0 0 0 0 0 0<br />

~ ~ ~ ~ ~ ~ ~ ~ ~ ~<br />

~ ~ ~ ~ ~ ~ ~ ~ ~ ~<br />

""""""""""""""""""""""""""""""""""""""""""""""""""<br />

0 m ~ c ~ ffi ~ 0 I ~<br />

CD ~ ~ ~ ~ !:: ~ ~ 6 CD<br />

~. ::::r 15. ~ 4. co ::::r C/)<br />

co IlJ 0 0.. IlJ 0 -. 0<br />

~ ~ ~ ~ C/) ~ ~ ~ ~ ~<br />

.." - . IlJ CD - . IlJ IlJ Q.. ~ ~<br />

o.. ~ co CD CD n, - CD IlJ :5. ()<br />

~< ~~< -......<br />

I ~ - -<br />

::::r C/) ~ () ~ IlJ 0 ~ !::<br />

~ ~ ~ 0 ~ 3 ~ 3 IlJ ~<br />

~ 0.. - () ~ 0 ~ IlJ ~ 0..<br />

-0 ::::r::::r~ < --0..-<br />

~ co ffi CD :- e?. CD g ffi -;!'.<br />

C/) CD w. 3 : -- 0.. C/) ;+ ~<br />

w.~ o. :"0.:.<br />

: ~ :: .:: :<br />

ci -<br />

et<br />

» c<br />

r::: Q)<br />

.....<br />

::r CD<br />

Q 0<br />

" -<br />

~ 0<br />

-0 . . Ci1<br />

A Q)<br />

0 =-.<br />

:5. 0<br />

." :]<br />

~ ..<br />

0 I\)<br />

C .j::>.<br />

." C-<br />

O> 0><br />

:] :::J<br />

:< ~<br />

-


~<br />

Q(J)W<br />

-.c::-<br />

1.0'0-<br />

-,0-<br />

::s Q)<br />

"""<br />

0 Q)<br />

-3<br />

tU Q)<br />

::s :J<br />

0 -<br />

3 JJ<br />

tU ~<br />

co<br />

s:: Q)<br />

III ~.<br />

tU -<br />

0 ~<br />

- -.<br />

-, ()<br />

< -<br />

a:.z<br />

(1) ~<br />

III r<br />

-0<br />

0"'0 -. CD<br />

(1) U><br />

(Do<br />

:u<br />

~:-o 0 -<br />

0 -I<br />

tU - """ -.<br />

tU"'O - 7\<br />

'< 0<br />

III - < -.<br />

illS'><br />

5' ~<br />

tU -<br />

-(J)<br />

"'tU<br />

"""<br />

=3<br />

::s ()<br />

(1) :J<br />

(1) 7\<br />

- -<br />

(Do<br />

o.<br />

- -<br />

-.<br />

0 (J)<br />

'< Q)<br />

CD'3<br />

III CD<br />

:J<br />

7\<br />

0<br />

<<br />

p'<br />

:<<br />

::D<br />

s:<br />

!:!><br />

.,<br />

;1\<br />

0<br />

<<br />

(';z<br />

s:<br />

I\)<br />

::I: r I\) ..........<br />

I\) -. 0 -<br />

ace < a<br />

~::r~~ ~<br />

'--'» '--'<br />

[Oat; CD < . ~ .....<br />

() -. ~ c:<br />

a~ 0 Q3<br />

cr<br />

~~ii1 ~<br />

- tb"c.a. 0 ~ CD<br />

"0--; I\)<br />

"0 '"C :::!. ::-'<br />

. -"0<br />

c.u I\) " "0<br />

.j:>.m~ ::'<br />

'7' -. ~ ......<br />

c.u (") - a<br />

(}ll»o 01<br />

!'VS"< "7'"<br />

-< ~ 0<br />

~ ""(J 01<br />

-. 0 ex:><br />

::::1"0 .<br />

.,,~<br />

0 ~<br />

... -<br />

~. A<br />

0 0<br />

»~<br />

2. c...<br />

c.~<br />

00><br />

>< ::J<br />

-.0<br />

--<br />

C. ~.<br />

I» .--:><br />

o' <<br />

::::I'<br />

~<br />

G)<br />

.2.<br />

"<br />

0 <<br />

~<br />

en<br />

r<br />

:s:~-=<br />

m ::J<br />

0 0<br />

en CD<br />

- ,<br />

2 <<br />

(') = -CD<br />

C -"'"<br />

~<br />

(I) 0<br />

0..<br />

_.:-n<br />

::J'" -<br />

:;" ~<br />

_1lJ<br />

~ !!.? ~. :::..<br />

-. I» .--:> tU<br />

-Cf) fj - Ci)<br />

c.u:s:: - -. ::J<br />

-.' ...... -(0 3 -<br />

- 0 ......<br />

c.u '< . en 0<br />

::-'0- --::;: Q) ~<br />

en -f<br />

~<br />

"6'<br />

7\<br />

0 <<br />

~<br />

;x><br />

:<<br />

(I) ::J.<br />

--0<br />

(I) '"<br />

(') 0<br />

!:t<<br />

0 _.<br />

(') _0<br />

ao<br />

~~<br />

en ..... - ,<br />

:E -. °<br />

::J<br />

'«J)<br />

en 3<br />

- -.<br />

-0<br />

::J"_'"<br />

c.<br />

::::s -<br />

Q) N<br />

C"::r<br />

-1lJ (I) ::J<br />

(')(0 0 -<br />

3 «<br />

"Oil<br />

0 ~<br />

!!.!.O<br />

!:!:CD<br />

0 cr<br />

::::s _CD<br />

Q) s::<br />

::::So<br />

o.;A;<br />

en ~<br />

c ::t><br />

::1.....<br />

Q) IlJ<br />

(') ::J<br />

(I) ~<br />

3 0<br />

0 ~<br />

~ --.<br />

"g.JJ<br />

O' -~<br />

0 -<br />

CQ(J)<br />

'< fir<br />

3<br />

CD<br />

::J<br />

'"<br />

0<br />

s.<br />

°<br />

:<<br />

JJ<br />

c<br />

0<br />

(")<br />

t:<br />

3<br />

CD<br />

::3<br />

-<br />

t/)<br />

-.<br />

::3<br />

0 <<br />

CD<br />

~<br />

:$.<br />

CD<br />

:e<br />

I\) I\)<br />

(,.) I\)<br />

1\)1\)<br />

0 0<br />

0 0<br />

(j) (j)<br />

I\) I\) ... ......<br />

... 0 CD 00 ......<br />

1\)1\)1\)1\)1\)<br />

0 0 0 0 0<br />

0 0 0 0 0<br />

ex> ex> ex><br />

......- '-' '-' '-' '-' '-'<br />

Z (j)<br />

-<br />

P> ~<br />

:J r::<br />

0 (") en<br />

(") r::<br />

m JJ (J) 5"»<br />

::::; CD q- ~<br />

~ ~ § c<br />

0 0 C :J"O<br />

0..<br />

~<br />

-.<br />

P> .....<br />

- P><br />

CD -<br />

- CD<br />

0 ~<br />

'D CD<br />

0 (")<br />

CO en<br />

.....<br />

P> ;-"<br />

'D :<br />

::J'"<br />

':<<br />

- -. - CD<br />

_:J-.o=:<br />

CO en !l> CD 0<br />

~ - :J<br />

!l> -. CD 0<br />

en "0 ::::;-0 en -<br />

en o-CD c ~<br />

'< ~ 54 ~'<<br />

0 ::2: - . !l> 0..<br />

!l> CD :J 0 -.<br />

=-' CD : CD 0<br />

: ? . : , to) 1\)""<br />

....<br />

.--...<br />

I\) I\) I\) I\) I\) I\)<br />

a a 0 000<br />

a a 0 0 ......<br />

~~~~9~<br />

-I C '1J '1J en ~<br />

::J'" ::J CD -<br />

CD -. 0 0 - =<br />

.o33CDO<br />

a t:: 0 0 ~ ::J.<br />

CD CD =. =.


~<br />

SQ () -::J<br />

""' ::!" Q) -.<br />

3 (D =1'<br />

0 :3<br />

::J -. 0' 0<br />

~~<br />

~- ?:'<br />

~ ~<br />

0<br />

~<br />

CJ)<br />

a ~::J<br />

-I<br />

~.<br />

0<br />

Q.<br />

(D (')<br />

0<br />

CD<br />

~ ~<br />

0 <<br />

o'<br />

0<br />

(J)<br />

.-+<br />

~<br />

3<br />

CD<br />

::J<br />

'"<br />

0 <<br />

o'<br />

c~...... CD ", 0<br />

G)...I. I\)<br />

\<br />

- (f)-a :T <<br />

-" CD 0' :iE -<br />

Q)Q)!:J<br />

'-' - -- CD<br />

"0"""0<br />

"OCD:::J<br />

. - s:u<br />

(J1CD::J<br />

(}l~(Q<br />

0""" -<br />

'OA<br />

(}l-\<br />

-0<br />

-a<br />

~ :5,<br />

0 - z:s::<br />

!\3(J)rO1<br />

0 s:::: ~_.-<br />

~O<br />

f\) I:\) ,<br />

O~<br />

:bS':Z:<br />

~::JO> VJ - :::I<br />


~<br />

i\)3'(J)'"<br />

0--00<br />

O_CDo:>c:<<br />

CD~<br />

::::J :J<br />

6'-°0<br />

,... CD S,<br />

... "oV<br />

:3 0 -<br />

tl)-(J) - tn "<br />

Q c: -<br />

::::~~<br />

~ I» -,<br />

CD 0""0 CD "<br />

(J)<br />

CD 3<br />

0<br />

s,<br />

:-.. c:r<br />

0<br />

""I<br />

"oV<br />

-<br />

il)" 'tJ 0<br />

::J ';j' ~<br />

t"".°A<br />

~oo<br />

CD (,Q :<br />

:<br />

E<br />

:3'<br />

(J)3~<br />

g~~ -_:J<br />

CD'ji) ~<br />

~::::J -<br />

~2.~<br />

UJo~<br />

oo~. 0<br />

I C. <<br />

CDI»~<br />

~ _:J<br />

-.0<br />

""00<<br />

::,::::J~ ex> I» -<br />

.j::>. -"<br />

()11»S:;<br />

cb(,Q~<br />

()1-<br />

~m;! tn""O<br />

'\"<br />

-o~<br />

~ (i) _O'<br />

n3g,0<br />

"'" g.Q)<br />

3<br />

CD<br />

:J<br />

;>\"<br />

0<br />

<<br />

0'<br />

:<<br />

s::<br />

P><br />

.....<br />

'"<br />

0<br />

<<br />

o'<br />

z<br />

s::<br />

Bcen""""<br />

O::s~O1<br />

0 -. 3-<br />

~.g 0<br />

c... (1) ~ "<br />

a""'" c:: (; -<br />

~ :::. en<br />

3=0<br />

- .<br />

a'<<br />

i:i<br />

I»o'<br />

::s<br />

° ::s<br />

tn<br />

r:::<br />

"C<br />

"C o<br />

~<br />

(1)<br />

C.<br />

"'C<br />

-<br />

(")<br />

I»<br />

-<br />

I»<br />

-< tn<br />

-<br />

i\)"tI-I""'" 0 0" ~ . c.><br />

0 -0-<br />


~<br />

- 0 '- I\) '--"'"T1 -j I\) Z A I\) (J) r I\) 0 m =I I\) '--"'::x:J (f) I\) (J) -I ....<br />

I\) >< 0 01 I\) II) :::; ,.a::. J\) II) 0 W I\) - 0 I\) I\) >< -<br />

. CD I\) (1) 0 I\) - :::; . < J\) ~ C) (1) :::;. II) -I (D 3 A<br />

-- N- 0-- 3 - o~-o-<br />

< ~ A CD' 0 J (') Q) ~ 3 ~ CD' (1) -0 to:::;. ~ -. 0<br />

CD ~ ~ ::J co -. _. =" ~ (1) -0 - n<br />

!=2- So ~ ~ g ~~~ ~n~ ~ (j)~ So;1\ """'11) ~<br />

(j)<br />

()<br />

<br />

::; -<br />

~6'<br />

-':E<br />

gQ) - -<br />

- - 0<br />

< -0 I/) 0'<br />

::' -g - n !=J<br />

~S'-<br />

Jg!:!?<br />

~~Q)<br />

I/) 3<br />

~ n<br />

~ -. -<br />

C. »<br />

~ 0 ~ -~. '-<br />

:gg.~<br />

.-Q)<br />

ro~~<br />

I\) ::J <<br />

OJ<br />

CD<br />

(Q<br />

-(f) , - ~<br />

CD<br />

Q.<br />

t1J'-"';'.I<br />

. 0<br />

=.<br />

I/)<br />

::r'<br />

(1)<br />

c.<br />

II)<br />

I\)<br />

m<br />

~<br />

C.<br />

(1)<br />

I/)<br />

I/) - I:<br />

9:<br />

(1)!l'::J<br />

::J~<br />

II) 0<br />

- :5.<br />

3 . - "<br />

0»<br />

Q, <<br />

Cf<br />

m<br />

c.;I><br />

(1)<br />

1J<br />

0<br />

I/)<br />

;::;:<br />

o'<br />

~-g~<br />

J n ="<br />

m -. 0<br />

. (1) <<br />

I/) 0'<br />

11)-<br />

::J <<br />

c.~<br />

~o~~<br />

'::J<br />

OJ:<<br />

I\) "'D :::::0<br />

ex> - ?'"<br />

~(1)<br />

CD<br />

n<br />

§'::J<br />

Q. C.<br />

C.<br />

C"::J<br />

II) 0<br />

-<br />

-r<br />

::rc<br />

~<br />

0<br />

CD 0 '< C. 1J (1)2 c.<br />

3 >< 3 ~ - ~(/) (1)<br />

OJ -. (1) II) a - 1J<br />

+ c. (1) II) ::!: (1)<br />

~ -. II) - ::J 0. 0I/)<br />

~ ~ ::J "'D I: g,- ;::;:<br />

m c. I/) - 3 (1)S: (1)<br />

- (Q 0 n -Q) Q,<br />

~ - - a ::J (1)- 0<br />

:5. ~ (J) II) ~ ~ 6 ::J<br />

~ I/) -I -< 0 a :5. (Q<br />

~ '< ~ ~ 0 gP ~<br />

an I/) II) -Z I/)<br />

II) ~ - 11). I/)<br />

-g ~ "T1 ~ -s: '<<br />

::4- C" S '< '$,. n<br />

_. 0 I/) -. II)<br />

ill::J<br />

(j)<br />

,<br />

- ?><br />

~ CD<br />

Q.<br />

C<br />

~<br />

Q.<br />

CD<br />

II)<br />

::J<br />

C.<br />

><<br />

::x:J<br />

c<br />

3<br />

(1) - ::;<br />

0<br />

c.<br />

- II)<br />

::J<br />

c. ;::;:::!:<br />

I/)<br />

~<br />

::!:<br />

< -. - '<<br />

n<br />

@<br />

II)<br />

n<br />

0<br />

::J<br />

I/)<br />

~<br />

C"<br />

0<br />

I/) ::J<br />

I:<br />

1J<br />

1J<br />

Q-<br />

~<br />

CD<br />

~<br />

~<br />

I/) 0<br />

~<br />

3<br />

(1)<br />

(j)<br />

()<br />

0<br />

l:J<br />

C<br />

(f)<br />

-<br />

::r<br />

II)<br />

::J<br />

0-<br />

@<br />

00'<br />

;:J"'<br />

~<br />

:::::::<br />

:E<br />

:E<br />

;E<br />

[/><br />

(")<br />

0<br />

'0<br />

.:<br />

[/><br />

(-,<br />

0<br />

~(")<br />

8 tV<br />

---<br />

'0 "1<br />

~'<br />

§.<br />

,'"<br />

[/><br />

S"<br />

CD<br />

~<br />

'-< II<br />

(")<br />

,<br />

0<br />

1"rI<br />

w<br />

00<br />

w<br />

0<br />

J<br />

J<br />

J<br />

tV<br />

J<br />

Ro<br />

0 "1<br />

,;;;'


~<br />

Print citation overview<br />

< Back I Print<br />

Date of creation: 25 January 2013<br />

Author: Tripkovi?, Du?an V.<br />

Self citations of selected author are excluded.<br />

hindex = 10<br />

Of the 26 documents considered for the h-Index, 10 have been cited at least 10 times.<br />

Note: The h Index considers Scopus documents published after 1995.<br />


Enhancing Hydrogen Evolution Activity in Water Splitting by<br />

Tailoring Li + -Ni(OH) 2-Pt<br />

Interfaces<br />

Ram Subbaraman,<br />

et al.<br />

Science 334,<br />

1256 (2011);<br />

DOI: 10.1126/science.1211934<br />

This copy is for your personal, non-commercial use only.<br />

If you wish to distribute this article to others,<br />

you can order high-quality copies for your<br />

colleagues, clients, or customers by clicking here.<br />

Permission to republish or repurpose articles or portions of articles can be obtained by<br />

following the guidelines here.<br />

The following resources related to this article are available online at<br />

www.sciencemag.org (this infomation is current as of December 8, 2011 ):<br />

Updated information and services, including high-resolution figures, can be found in the online<br />

version of this article at:<br />

http://www.sciencemag.org/content/334/6060/1256.full.html<br />

Supporting Online Material can be found at:<br />

http://www.sciencemag.org/content/suppl/2011/11/30/334.6060.1256.DC1.html<br />

A list of selected additional articles on the Science Web sites related to this article can be<br />

found at:<br />

http://www.sciencemag.org/content/334/6060/1256.full.html#related<br />

This article cites 38 articles,<br />

3 of which can be accessed free:<br />

http://www.sciencemag.org/content/334/6060/1256.full.html#ref-list-1<br />

This article appears in the following sub<strong>je</strong>ct collections:<br />

Chemistry<br />

http://www.sciencemag.org/cgi/collection/chemistry<br />

Science (print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by the<br />

American Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005. Copyright<br />

2011 by the American Association for the Advancement of Science; all rights reserved. The title Science is a<br />

registered trademark of AAAS.<br />

on December 8, 2011<br />

www.sciencemag.org<br />

Downloaded from


REPORTS<br />

1256<br />

(aL † aR † |vac〉) given an expected mean number of<br />

events m 0 = d 2 Np s consistent with zero concurrence,<br />

where ps =1.0×10 −3 is the probability<br />

of generating a Stokes heralding photon and N<br />

is the number of experimental runs. Our results<br />

(X =3,N =1.9×10 14 ,andm 0 =9.1T 0.9) indicate<br />

positive concurrence at a 98 T 1% confidence level.<br />

Therefore, based on this detection of entanglement<br />

between Stokes and anti-Stokes modes, we<br />

can infer entanglement between the phonon modes<br />

of two macroscopic solids at room temperature.<br />

Finally, we examine the quality of entanglement<br />

generated between the diamonds by neglecting<br />

the vacuum component in Fig. 3, which<br />

is only caused by inefficiencies in coupling, detection,<br />

and readout of the anti-Stokes mode. To<br />

do this, we performed quantum state tomography<br />

(25) on the joint polarization state of the<br />

Stokes and anti-Stokes modes, postselecting on<br />

the detection of both photons. The reconstructed<br />

state is shown in Fig. 4, and we have subtracted<br />

accidental coincidences calculated from the Stokes<br />

and anti-Stokes singles rates. These results provide<br />

a more complete estimate of the coherence between<br />

the two modes than the interference fringes<br />

in Fig. 2. The concurrence of this subspace, 0.85,<br />

provides an estimate of the achievable entanglement<br />

between the two diamond phonon modes<br />

as the readout efficiency, coupling, and detector<br />

efficiencies approach unity (i.e., p00 → 0). Further,<br />

the fidelity to the nearest Bell state ½jHV 〉 þ<br />

jVH〉Š= ffiffi p<br />

2 is 0.91. However, in the presence of<br />

real coupling and detection losses, the existence<br />

of entanglement can only be inferred from the<br />

density matrix in Fig. 3 (22).<br />

In our experiment, short-lived quantum correlations<br />

were revealed by combining an ultrafast<br />

interferometric pump-probe scheme with<br />

photon-counting techniques. The large optical<br />

bandwidth enabled the resolution of extremely<br />

fast dynamics in the solids, and also operation<br />

at high data rates, providing sufficient statistics<br />

to establish entanglement even in the presence<br />

of losses. This approach lays the foundation for<br />

future studies of quantum phenomena in manybody,<br />

strongly interacting systems coupled to strongly<br />

decohering environments and points toward a<br />

novel platform for ultrafast quantum information<br />

processing at room temperature.<br />

References and Notes<br />

1. R. Penrose, in Mathematical Physics, A. Fokas,<br />

T. W. B. Kibble, A. Grigoriou, B. Zegarlinski, Eds.<br />

(Imperial College Press, London, 2000), pp. 266–282.<br />

2. D. P. DiVincenzo, Fortschr. Phys. 48, 771 (2000).<br />

3. S. Gigan et al., Nature 444, 67 (2006).<br />

4. J. D. Thompson et al., Nature 452, 900 (2008).<br />

5. B. Abbott et al., New J. Phys. 11, 073032 (2009).<br />

6. S. Gerlich et al., Nat Commun 2, 263 (2011).<br />

7. A. D. O’Connell et al., Nature 464, 697 (2010).<br />

8. J. D. Teufel et al., Nature 475, 359 (2011).<br />

9. J. Chan et al., Nature 478, 89 (2011).<br />

10. G. Panitchayangkoon et al., Proc. Natl. Acad. Sci. U.S.A.<br />

107, 12766 (2010).<br />

Enhancing Hydrogen Evolution Activity<br />

in Water Splitting by Tailoring<br />

Li + -Ni(OH) 2-Pt Interfaces<br />

Ram Subbaraman, 1,2 Dusan Tripkovic, 1 Dusan Strmcnik, 1 Kee-Chul Chang, 1<br />

Masanobu Uchimura, 1,3 Arvydas P. Paulikas, 1 Vojislav Stamenkovic, 1 Nenad M. Markovic 1 *<br />

Improving the sluggish kinetics for the electrochemical reduction of water to molecular hydrogen in<br />

alkaline environments is one key to reducing the high overpotentials and associated energy losses in<br />

water-alkali and chlor-alkali electrolyzers. We found that a controlled arrangement of nanometer-scale Ni<br />

(OH)2 clusters on platinum electrode surfaces manifests a factor of 8 activity increase in catalyzing the<br />

hydrogen evolution reaction relative to state-of-the-art metal and metal-oxide catalysts. In a bifunctional<br />

effect, the edges of the Ni(OH)2 clusters promoted the dissociation of water and the production of<br />

hydrogen intermediates that then adsorbed on the nearby Pt surfaces and recombined into molecular<br />

hydrogen. The generation of these hydrogen intermediates could be further enhanced via Li + -induced<br />

destabilization of the HO–H bond, resulting in a factor of 10 total increase in activity.<br />

Electrocatalysis of the hydrogen evolution<br />

reaction (HER) is critical to the operation<br />

of water-alkali electrolyzers (1–6), in which<br />

hydrogen is the main product, and chlor-alkali<br />

electrolyzers (5, 6), in which it is a side product.<br />

1<br />

Materials Science Division, Argonne National Laboratory,<br />

Lemont, IL 60559, USA. 2 Nuclear Engineering Division,<br />

Argonne National Laboratory, Lemont, IL 60559, USA. 3 Advanced<br />

Materials Laboratory, Nissan Research Center, Kanagawa<br />

237-8523, Japan.<br />

*To whom correspondence should be addressed. E-mail:<br />

nmmarkovic@anl.gov<br />

These two technologies are highly energy-intensive<br />

and are known to account for ~25 to 30% (87,600<br />

to 92,000 GWh/year) of the total electrical energy<br />

consumption by industrial processes in the United<br />

States (3, 7). The HER is also an electrochemical<br />

reaction of fundamental scientific importance; the<br />

basic laws of electrode kinetics, as well as many<br />

modern concepts in electrocatalysis, were developed<br />

and verified by examining the reaction mechanisms<br />

related to the charge transfer–induced<br />

conversion of protons (in acid solutions) and water<br />

(in alkaline solutions) to molecular hydrogen.<br />

2 DECEMBER 2011 VOL 334 SCIENCE www.sciencemag.org<br />

11. E. Collini et al., Nature 463, 644 (2010).<br />

12. L.-M. Duan, M. D. Lukin, J. I. Cirac, P. Zoller, Nature 414,<br />

413 (2001).<br />

13. T. Chanelière et al., Nature 438, 833 (2005).<br />

14. K. S. Choi, H. Deng, J. Laurat, H. J. Kimble, Nature 452,<br />

67 (2008).<br />

15. D. N. Matsukevich, A. Kuzmich, Science 306, 663 (2004).<br />

16. C. W. Chou et al., Nature 438, 828 (2005).<br />

17. G. Milburn, J. Opt. Soc. Am. B 24, 167 (2007).<br />

18. See supplementary information on Science Online.<br />

19. K. C. Lee et al., Diam. Relat. Mater. 19, 1289 (2010).<br />

20. A. Greentree, B. Fairchild, F. Hossain, S. Prawer,<br />

Mater. Today 11, 22 (2008).<br />

21. L. Prechtel et al., Nano Lett. 11, 269 (2011).<br />

22. S. J. van Enk, Phys. Rev. A 75, 052318 (2007).<br />

23. W. Wootters, Phys. Rev. Lett. 80, 2245 (1998).<br />

24. G. J. Feldman, R. D. Cousins, Phys. Rev. D Part. Fields 57,<br />

3873 (1998).<br />

25. D. F. V. James, P. G. Kwiat, W. J. Munro, A. G. White,<br />

Phys. Rev. A 64, 052312 (2001).<br />

Acknowledgments: We thank V. Vedral, A. Datta, and<br />

L. Zhang for valuable insights. This work was supported by the<br />

Royal Society, Engineering and Physical Sciences Research<br />

Council (grant GR/S82176/01), EU IP Q-ESSENCE (grant<br />

248095), EU ITN FASTQUAST, U.S. European Office of<br />

Aerospace Research and Development (grant 093020),<br />

Clarendon Fund, St. Edmund Hall, and Natural Sciences and<br />

Engineering Research Council of Canada.<br />

Supporting Online Material<br />

www.sciencemag.org/cgi/content/full/334/6060/1253/DC1<br />

Materials and Methods<br />

SOM Text<br />

References (26–36)<br />

29 July 2011; accepted 27 October 2011<br />

10.1126/science.1211914<br />

Although previous studies have helped to rationalize<br />

which surface properties govern the<br />

variations in reactivity among catalysts (8–12),<br />

many key questions concerning the HER remain<br />

unanswered. For example, it is not clear why the<br />

rate of the HER is ~2 to 3 orders of magnitude<br />

lower at pH = 13 than at pH = 1, nor why the<br />

reaction is sensitive to the catalyst surface structure<br />

in alkaline media but largely insensitive in acids<br />

(13–17). A practical implication of the slow kinetics<br />

in alkaline solution is the lower energy efficiency<br />

for both water-alkali and chlor-alkali electrolyzers.<br />

For water-alkali electrolyzers, the high overpotentials<br />

for the oxygen evolution reaction (OER) at<br />

the anode also contribute significantly overall<br />

energy losses (18). This has led to various approaches<br />

to identify catalysts for both the OER<br />

and HER. However, such design strategies<br />

have rarely been based on molecular-level<br />

understanding of the reaction pathways. In<br />

addition, the influence of noncovalent (van der<br />

Waals–type) interactions on the overall kinetics<br />

of the HER has been underexplored, particularly<br />

in light of recent studies highlighting the<br />

impact of noncovalent interactions on the rates<br />

of many electrochemical reactions such as the<br />

oxygen reduction reaction, together with CO<br />

and methanol oxidation reactions (19–22).<br />

Currently, various combinations of metals (Pt,<br />

Pd,Ir,Ru,Ag,Ni),metalalloys(Ni-Co,Ni-Mn,<br />

Ni-Mo), metal oxides (RuO 2), and Ni sulfides<br />

and phosphides are used to catalyze the conver-<br />

on December 8, 2011<br />

www.sciencemag.org<br />

Downloaded from


sion of H2O toH2 (2, 10–12, 23–27). Although<br />

Pt and Pt-based systems offer the highest activity<br />

and stability of all materials used for the HER,<br />

the benefits have not, to date, warranted the high<br />

cost associated with these materials. As a result,<br />

conventional electrolyzers generally use high–<br />

surface area Raney Ni and Ni alloys as the HER<br />

catalysts (2, 23, 24, 28). Several engineering approaches<br />

have been used to improve these nonnoble<br />

catalyst materials, such as enhancing the<br />

surface area, changing the alloy composition, and<br />

using higher catalyst loadings (~25 to 40 times<br />

the equivalent for Pt) (2). Although these approaches<br />

have offered small performance gains,<br />

key problems with the use of such non-noble materials<br />

remain, including the decrease in activity<br />

arising from both the formation of hydrides and<br />

the oxidative dissolution of the catalyst during<br />

intermittent operation (2). These materials limitations<br />

suggest that superior performance might<br />

be achieved if lower-cost Pt-based cathode materials<br />

could be developed. Indeed, by substantially<br />

increasing the activity of Pt and by also<br />

decreasing the Pt loading through the use of<br />

Pt-shell nanomaterials with non-noble cores<br />

(29, 30), it would be possible to envision the use<br />

of highly active, durable, and low-cost Pt-based<br />

HER electrodes for alkaline electrolyzers.<br />

Limitations in the catalytic activities of Pt and<br />

Pt-group metals arise from the fact that although<br />

most of these materials are good catalysts for the<br />

adsorption and recombination of the reactive hydrogen<br />

intermediates (H ad), they are generally<br />

inefficient in the prior step of water dissociation.<br />

Conversely, metal oxides (and in some cases<br />

other compounds such as sulfides), although effective<br />

for cleaving the HO–H bond, are poor at<br />

converting the resulting Had intermediates to H2<br />

(31–33). Hence, optimal HER catalyst design<br />

will depend on combining the catalytic proficiencies<br />

of metals and metal oxides by creating<br />

new bifunctional metal oxide–metal systems<br />

(metal oxides deposited on metal substrates)<br />

(34–37).<br />

Here, we report the design and performance<br />

of composite materials to facilitate different parts<br />

of the overall multistep HER process in alkaline<br />

environments: an oxide to provide the active sites<br />

for dissociation of water, and a metal to facilitate<br />

adsorption of the atomic hydrogen produced and<br />

its subsequent association to form H2 from these<br />

intermediates. This involved growth of conductive<br />

ultrathin Ni(OH)2 clusters (height 0.7 nm,<br />

width 8 to 10 nm) on both pristine Pt singlecrystal<br />

surfaces and Pt surfaces modified by twodimensional<br />

(2D) Pt ad-islands [Pt-islands/Pt(111)].<br />

We found that, relative to the corresponding Pt<br />

single-crystal surfaces, the most active Ni(OH) 2/Ptislands/Pt(111)<br />

electrodes in KOH solutions are<br />

more active for the HER by a factor of ~8 at an<br />

overpotential of –0.1 V. Further enhancement<br />

of water dissociation is achieved by the intro-<br />

Fig. 1. (A to C) STM images (60 nm by 60 nm) and CV traces for (A) Pt(111), (B) Pt(111) with 2D Pt<br />

islands, and (C) Pt(111) modified with 3D Ni(OH)2 clusters in 0.1 M KOH electrolyte. (D) HER activities for<br />

Pt(111), Pt-islands/Pt(111), and Ni(OH)2/Pt(111) surfaces in alkaline solution (a, b, and c, respectively).<br />

Corresponding HER activities for Pt(111) and Pt-islands/Pt(111) electrodes in acid solutions (a´ and b´) are<br />

shown to emphasize large initial differences between kinetics of the HER in alkaline versus acid solutions.<br />

The inset shows XANES spectra for Ni(OH)2 on Pt(111) shown for three different potentials: HER (–0.1 V),<br />

Hupd (0.1 V), and near OER (1.2 V). Also shown is the reference spectrum for Ni(OH)2. No shift in the edge<br />

energy in XANES spectra between HER and H upd regionsisobserved.<br />

REPORTS<br />

duction of solvated Li + ions into the compact<br />

portion of the double layer, resulting in a factor<br />

of 10 total increase in activity. Finally, we demonstrate<br />

that the knowledge attained by studying<br />

single-crystal surfaces can be used for the design<br />

of prospective commercial nanocatalysts for alkaline<br />

electrolyzers.<br />

As a starting point, to develop more complete<br />

structure-function relationships for the HER, we<br />

used scanning tunneling microscopy (STM) to<br />

compare the atomic structures of Pt(111) and<br />

Pt(111) modified by electrochemically deposited<br />

Pt islands, referred as Pt-islands/Pt(111), (38). In<br />

agreement with prior reports (39, 40), the image<br />

of Pt(111) in Fig. 1A displays the presence of<br />

a few randomly distributed mono-atomic steps<br />

and 2D Pt islands with diameters of 1 to 2 nm and<br />

monoatomic height. Considering that the shape<br />

of the current-potential curve in both the underpotentially<br />

deposited hydrogen (Hupd) region<br />

[defined as the state of hydrogen adsorbed at a<br />

potential that is positive of the Nernst potentialforthehydrogenreaction(33)],<br />

between<br />

0.05 to 0.35 V, and the region of reversible adsorption<br />

of hydroxyl (OH ad) species, above 0.6 V,<br />

is consistent with earlier reports for a perfect<br />

Pt(111) surface, we conclude that these defects<br />

are invisible in cyclic voltammetry (CV) traces.<br />

In the STM image of the islands shown in Fig.<br />

1B, however, the Pt adatoms can be clearly<br />

resolved as 2D features with diameters of ~1 to<br />

3 nm and a height of 1 atomic layer. The CV<br />

trace of such a surface encompasses two sharp<br />

Hupd peaks centered at 0.23 V and 0.4 V (Fig.<br />

1B). On the basis of prior studies (39, 40), we<br />

associate these peaks with hydrogen adsorption<br />

at the (111)-(111) and (111)-(100) terrace-step<br />

sites, respectively. Consistent with the higher<br />

oxophilicity of low-coordinated Pt sites, the<br />

onset of OH adsorption starts at more negative<br />

potentials on the Pt island–covered electrode<br />

than on pristine Pt(111), whereas the OHad<br />

peaks are less reversible on the former surface.<br />

After 50 potential cycles between –0.3 V and<br />

+0.3 V, the STM images and the CV traces<br />

remain the same, indicating that within this<br />

potential range, the morphologies of the Pt(111)<br />

and Pt-islands/Pt(111) surfaces are stable. Therefore,<br />

such well-defined surfaces offer a unique<br />

opportunity to correlate the kinetic rates of the<br />

HER with a truly atomically resolved surface<br />

structure.<br />

The mechanism of the HER in alkaline media<br />

is typically treated as a combination of three elementary<br />

steps: the Volmer step—water dissociation<br />

and formation of a reactive intermediate Had<br />

(2H2O+M+2e – ⇆ 2M-Had +2OH – )—followed<br />

by either the Heyrovsky step (H 2O+H ad-M + e – ⇆<br />

M+H2 +OH – ) or the Tafel recombination step<br />

(2M-H ad ⇆ 2M + H 2). Adsorbed hydrogen species<br />

Had formed at potentials negative of the Nernst<br />

reversible potential for the HER are also referred<br />

to as overpotentially deposited hydrogen (Hopd).<br />

To distinguish the different states of adsorbed<br />

hydrogen, we use a thermodynamic notation,<br />

www.sciencemag.org SCIENCE VOL 334 2 DECEMBER 2011 1257<br />

Downloaded from<br />

www.sciencemag.org on December 8, 2011


REPORTS<br />

1258<br />

referring to Hupd as the strongly adsorbed state<br />

and H ad (i.e., H opd) as a weakly adsorbed state.<br />

Any rigorous kinetic analysis of the HER lies<br />

beyond the scope of the present discussion. Rather,<br />

we focus mainly on the design of interfaces<br />

for efficient electrochemical conversion of H 2O<br />

to H2. For example, Fig. 1D shows that in alkaline<br />

solution, relative to the corresponding pristine<br />

Pt(111) surface, the Pt-islands/Pt(111)<br />

surface is ~5 to 6 times as active for the HER.<br />

Figure 1D also shows that in acid solution, the<br />

HER on the Pt-islands/Pt(111) electrode is improved<br />

by a factor of only ~1.5. In turn, this<br />

strong pH effect indicates that the low-coordinated<br />

Pt atoms may have a major effect on the ratedetermining<br />

step of the HER in alkaline solutions.<br />

Because the major difference between the<br />

reaction pathways in alkaline and acid solutions<br />

is that the hydrogen in alkaline solutions is discharged<br />

from water instead of from hydronium<br />

ions (H3O + )(13–15), we propose that the large<br />

promoting effect of low-coordinated Pt atoms in<br />

alkaline solution is due to more facile dissociative<br />

adsorption of water. In turn, this would be<br />

consistent with the Volmer reaction being the<br />

rate-determining step for the HER in alkaline<br />

electrolytes. The role of edge-step sites in accelerating<br />

dissociative adsorption of water on metal<br />

surfaces is well documented in ultrahigh-vacuum<br />

(UHV) environments (33). We therefore conclude<br />

that for materials with near-optimal M-H ad energetics<br />

(such as Pt), surface reactivity for the<br />

HER can be further improved by tailoring the active<br />

sites for more efficient dissociative adsorption<br />

of water molecules.<br />

To fulfill this requirement, we modified Pt(111)<br />

and Pt-island/Pt(111) surfaces by depositing<br />

Ni-(hydr)oxide clusters (38), as the 3d-transition<br />

metal oxides might be even more active for water<br />

dissociation than Pt defect sites (31). The facile<br />

water dissociation properties of Ni(hydroxy)<br />

oxides relative to other transition metal oxides<br />

have been well established (26–28), motivating<br />

the use of Ni(OH)2 for this work. The local symmetry,<br />

the oxidation state of Ni atoms, and the<br />

number and identities of nearest-neighbor atoms<br />

and the distances between them were determined<br />

by in situ x-ray absorption spectroscopy (XAS)<br />

measurements (41, 42). For example, from the<br />

analysis of the x-ray absorption near-edge structure<br />

(XANES) and extended x-ray absorption<br />

fine structure (EXAFS) of the XAS spectra (inset<br />

of Fig. 1D; see also fig. S3), we found Ni-O and<br />

Ni-Ni bond distances of 2.05 T 0.01 Å and 3.08 T<br />

0.01 Å, and we also determined that Ni remains<br />

mostly in the +2 valence state, even after multiple<br />

hours of holding the electrode potential at –0.1 V.<br />

Furthermore, from the edge shift (defined as the<br />

half-height energy of the normalized XANES<br />

edge step), we concluded that between –0.1 Vand<br />

+0.8 V, the change in the oxidation state of Ni is<br />

less than 0.5. These results suggest that stable<br />

Ni(OH)2 clusters are the predominant hydr(oxide)<br />

form on the Pt(111) and Pt-islands/Pt(111) surfaces,<br />

especially in the HER potential region.<br />

Because the octahedral symmetry of the a and/or<br />

b phases of Ni(OH) 2 prevents p-d hybridization,<br />

the prominent pre-edge from 1s → 3d transitions<br />

implies that the Ni(OH) 2 species are rich in defects<br />

that, according to prior reports (31–33), are<br />

particularly active for dissociative adsorption<br />

of water molecules.<br />

The surface morphologies of Ni(OH) 2/Pt(111)<br />

and Ni(OH)2/Pt-islands/Pt(111) were probed by<br />

STM and in situ surface x-ray crystal truncation<br />

rod (CTR) measurements (43, 44)(seefig.S4for<br />

the CTR data). Although atomic resolution could<br />

not be obtained, the STM image in Fig. 1C clearly<br />

shows that the Ni(OH) 2 clusters are randomly<br />

distributed across the (111) terraces. All Ni(OH)2<br />

clusters exhibited hemisphere-like shapes, with<br />

characteristic diameters of ~8 to 10 nm and heights<br />

of ~0.7 nm, the latter corresponding to two layers<br />

of Ni(OH)2. This result indicates that the oxide<br />

exhibits Volmer-Weber (VW)–type growth<br />

whereby 3D clusters of Ni(OH)2 grow even at<br />

the lowest coverages (45). VW growth, in turn, is<br />

possible if the heat of adsorption of Ni(OH)2 on<br />

Pt is lower than the cohesive energy of Ni(OH) 2.<br />

Fig. 2. (A) STM image (60 nm by 60 nm) and CV trace of the Ni(OH)2/Pt-islands/Pt(111) surface. Clusters<br />

of Ni(OH) 2 in the STM image appear ellipsoidal with particle sizes between 4 and 12 nm. (B)Comparison<br />

of HER activities with Pt(111) as the substrate. Incremental improvements in activities for the HER in 0.1 M<br />

KOH from the unmodified Pt(111) surface are shown for the hierarchical materials [ad-islands, Ni(OH)2,<br />

and their combination] as well as the double layer (addition of Li + cations). The activity for the unmodified<br />

Pt(111) surface in 0.1 M HClO 4 is shown for reference. Dashed arrow shows the activity trend.<br />

Fig. 3. Schematic representation of water dissociation, formation of M-H ad intermediates, and subsequent<br />

recombination of two Had atoms to form H2 (magenta arrow) as well as OH – desorption from<br />

the Ni(OH)2 domains (red arrows) followed by adsorption of another water molecule on the same site<br />

(blue arrows). Water adsorption requires concerted interaction of O atoms with Ni(OH) 2 (broken orange<br />

spikes) and H atoms with Pt (broken magenta spikes) at the boundary between Ni(OH)2 and Pt domains.<br />

The Ni(OH)2-induced stabilization of hydrated cations (AC + ) (broken dark blue spikes) likely occurs through<br />

noncovalent (van der Waals–type) interactions. Hydrated AC + can further interact with water molecules<br />

(broken yellow spikes), altering the orientation of water as well as the nature and strength of interaction<br />

of the oxide with water.<br />

2 DECEMBER 2011 VOL 334 SCIENCE www.sciencemag.org<br />

on December 8, 2011<br />

www.sciencemag.org<br />

Downloaded from


The surface coverage of Ni(OH)2 on Pt(111) is<br />

estimated from the STM image by measuring<br />

the area covered by the particles on the Pt(111)<br />

substrate. According to such analysis, the cluster<br />

density reached a maximum at a surface coverage<br />

of ~35%. Upon comparing STM images of<br />

Pt(111) and Ni(OH)2/Pt(111) surfaces (Fig. 1, A<br />

and C), however, it was not clear whether isolated<br />

defects or flat Pt(111) terraces were the<br />

preferred nucleation sites for Ni(OH) 2. This information<br />

was more accessible by comparing<br />

STM images of the Ni(OH) 2-free (Fig. 1B) and<br />

Ni(OH)2-covered Pt-island/Pt(111) (Fig. 2A)<br />

electrodes. The STM image in Fig. 2A was acquired<br />

after deposition of Ni(OH)2 on a Pt(111)<br />

surface modified by ~0.2 monolayer (ML) of<br />

Pt islands. We observed formation of both 3D<br />

Ni(OH)2 clusters (having a predominantly ellipsoidal<br />

shape) and oxide-free terraces. The<br />

clusters had approximately constant heights of<br />

~0.8 nm but diameters ranging from 4 to 12 nm.<br />

The same STM image, however, revealed no<br />

visible presence of 2D Pt islands, which suggests<br />

that Ni(OH)2 preferentially nucleates on the Pt<br />

surface defects and that most of the Pt islands<br />

are covered by Ni(OH)2.<br />

Having established the nature and morphology<br />

of the Ni(OH)2 clusters, we briefly summarize<br />

the role of the Ni(OH) 2 clusters in the formation<br />

of Hupd and OHad adlayers on Pt(111) and Ptislands/Pt(111)<br />

electrodes. Addition of Ni(OH) 2<br />

on the surface of Pt(111) as well as on the Pt(111)<br />

surface covered with Pt islands led to a decrease<br />

(~35%) in the coverage of Hupd. This finding<br />

suggests that the Ni(OH) 2 clusters selectively<br />

block the Pt sites corresponding to Hupd. Furthermore,<br />

the two sharp H upd peaks characteristic<br />

of hydrogen adsorption/desorption on the Pt(111)<br />

electrode modified by the 2D Pt islands are completely<br />

suppressed on the surface covered by the<br />

3D Ni(OH)2 clusters (Fig. 2A). This is additional<br />

evidence, consistent with the STM results,<br />

that defects serve as the nucleation centers for<br />

electrodeposition of Ni(OH)2. We therefore conclude<br />

that Pt islands are predominantly covered<br />

by Ni(OH)2. In contrast to the Hupd potential re-<br />

Fig. 4. (A) STM image (50 nm by 50 nm) and CV trace for Ni(OH)2/Pt-islands/Pt(110) single-crystal<br />

surface. STM image shows Ni(OH) 2 randomly distributed on inherently rough Pt(110). . (B)Comparisonof<br />

HER activities with Pt(110) as the substrate. Incremental improvements in activities for the HER in 0.1 M<br />

KOH from the unmodified Pt(110) surface are shown for the hierarchical materials [Ni(OH)2 and the<br />

combination of ad-islands with Ni(OH)2] as well as the double layer (addition of Li + cations). The activity<br />

for the unmodified Pt(110) surface in 0.1 M HClO 4 is shown for reference. Dashed arrow shows the activity<br />

trend. (C) Transmission electron micrograph image (50 nm by 50 nm) and corresponding CV trace of<br />

Ni(OH)2-free Pt-nano catalysts (TKK) with an average particle size of 5 nm. (D) Comparison of HER activities<br />

with commercial nanocatalyst Pt/C (TKK) as the substrate. Incremental improvements in activities for the<br />

HER in 0.1 M KOH from unmodified Pt/C are shown for the hierarchical materials [surface covered with<br />

Ni(OH)2] as well as the double layer (addition of Li + cations). The activity for the unmodified Pt/C surface<br />

in 0.1 M HClO4 is shown for reference. Dashed arrow shows the activity trend. All the current densities for<br />

the TKK catalyst system are normalized by the geometric surface area of the glassy carbon substrate.<br />

REPORTS<br />

gion, an enhanced adsorption of OHad, whichis<br />

accompanied by irreversible reduction of OH ad<br />

on the negative-going sweep, is observed on both<br />

electrodes, arising from the higher oxophilicity<br />

of the surface elements.<br />

Although in the presence of Ni(OH) 2 clusters<br />

there are 35% fewer Pt sites available for<br />

the HER than on the bare Pt(111) substrate, the<br />

Ni(OH)2/Pt(111) electrode is ~7 times as active<br />

for the HER relative to Pt(111) (Fig. 1D). Moreover,<br />

Fig. 2B shows that the activity is further enhanced<br />

[by a factor of ~8 relative to bare Pt(111)]<br />

on the Ni(OH)2/Pt-island/Pt(111) surface; at<br />

6 mA/cm 2 , the difference in overpotential between<br />

the HER in alkaline and acid solutions is<br />

reduced to only 100 mV. Clearly, then, on both<br />

surfaces, Ni(OH)2 must play a promoting role in<br />

the dissociation of water and thereby enhances<br />

therateofformationofHad intermediates on the<br />

metal surface. As schematically depicted in Fig.<br />

3, we propose that water adsorption requires<br />

concerted interaction of O atoms with Ni(OH)2<br />

and H atoms with Pt at the boundary between<br />

Ni(OH)2 and Pt domains. Water adsorption is<br />

then followed by water dissociation and hydrogen<br />

adsorption (Had) on the nearby vacant Pt<br />

sites. Finally, two H ad atoms on the Pt surface<br />

recombine to form H2 (H2 desorption step) and<br />

OH – desorbs from the Ni(OH) 2 domains, followed<br />

by adsorption of another water molecule<br />

on the same site.<br />

From a surface reactivity standpoint, fruitful<br />

kinetic synergy (bifunctionality) between Ni(OH) 2<br />

and Pt appears to be the key to maximizing the rate<br />

of the HER. As shown in Fig. 2B, this bifunctionality<br />

in turn brings the activity of the HER in<br />

alkaline solutions very close to the activity of Pt<br />

in acid solutions. To verify this conclusion, we<br />

have also compared the HER on Au(111) and<br />

Ni(OH)2/Au(111) in alkaline solution. The relatively<br />

weak interaction between Au and H ad offsets<br />

the benefit of the enhanced water dissociation<br />

at the Au/Ni(OH) 2 interface (38). As a result, the<br />

rate of the HER on the Ni(OH)2/Au(111) surface<br />

is much lower than on the Pt(111)/Ni(OH) 2 surface.<br />

This further emphasizes the importance of<br />

choosing the correct metal oxide–metal pairs in<br />

optimizing the kinetics of the HER. In what<br />

follows, we demonstrate that the nature of interactions<br />

in double layer is equally important, and<br />

the rate-determining Volmer step can be further<br />

enhanced by tuning the double-layer properties.<br />

Several recent studies have unambiguously<br />

shown that the rate of electrochemical reactions<br />

on Pt in alkaline solutions is controlled by the<br />

presence of alkali-metal cations (AC + )(20, 21).<br />

However, these effects have been entirely restricted<br />

to the potential region of a critical OH ad<br />

coverage (for electrode potential E > 0.6 V),<br />

the latter species serving to stabilize hydrated<br />

cations in the compact part of the double layer<br />

through noncovalent (van der Waals–type) interactions.<br />

This stabilization leads to the formation<br />

of OH ad-AC + (H 2O) x complexes that can<br />

either decrease the reactivity of Pt by blocking<br />

www.sciencemag.org SCIENCE VOL 334 2 DECEMBER 2011 1259<br />

on December 8, 2011<br />

www.sciencemag.org<br />

Downloaded from


REPORTS<br />

1260<br />

the active sites for adsorption of reactants such<br />

as O 2,H 2, and CH 3OH (20, 21) or, as in the<br />

case of the CO oxidation reaction, improve<br />

the reactivity of Pt via enhanced adsorption of<br />

OHad (22). For the present reaction, the key questions<br />

are whether hydrated alkali cations on the<br />

Ni(OH)2/Pt-islands/Pt(111) surface can be stabilized<br />

through interactions with OH species in the<br />

Ni(OH)2 cluster and, if so, whether these hydrated<br />

cations could affect the alkaline HER kinetics.<br />

For these purposes, the effect of hydrated<br />

Li cations was probed mainly because, in alkaline<br />

environments, Li + is known to interact<br />

with H 2O and OH ad more strongly than K + .In<br />

line with the previous electrochemical report<br />

(20), we found that Li + cations have no effect on<br />

the HER on Pt(111) surfaces. However, the<br />

results in Fig. 2B revealed that the HER on<br />

Ni(OH)2/Pt-islands/Pt(111) is enhanced by almost<br />

a factor of 2 in the presence of Li + cations.<br />

This increase in activity has substantially narrowed<br />

the gap between the rates of HER on Pt<br />

in acid and alkaline solution; more specifically,<br />

inspection of Fig. 2B shows that, at 5 mA/cm 2 ,<br />

the difference in overpotential between acid and<br />

alkaline environments is narrowed to only 35 mV.<br />

The fact that the activity of the HER is affected<br />

by the nature of alkali metal cations strongly<br />

suggests that Ni(OH) 2-Li + -OH-H complexes are<br />

present in the compact portion of the double<br />

layer. The presence of this complex does not completely<br />

explain the factor of 2 increase in HER activity.<br />

It is likely that the probability of water<br />

dissociation is enhanced via possible Li + -induced<br />

steric and/or electronic effects on the interfacial<br />

water structure and reactivity, as shown schematically<br />

in Fig. 3. Thus, Ni(OH) 2 plays a dual role:<br />

In addition to assisting with water dissociation, it<br />

also provides an anchor to hold the beneficial Li +<br />

ions in the compact portion of the double layer.<br />

Having established the methodology of tuning<br />

interfacial activity for the HER on Pt(111),<br />

we demonstrate that the very same guiding principles<br />

for accelerating the Volmer reaction step<br />

in alkaline solutions are equally applicable to<br />

Pt(110). For simplicity, only the CV and STM<br />

data for the most active system, Ni(OH) 2/Ptislands/Pt(110),<br />

are shown in Fig. 4A. The general<br />

characteristics (both structural and electrochemical)<br />

are similar to what was observed for the<br />

corresponding Pt(111) systems. The current denities<br />

for the HER on Pt(110) and Pt-nanocatalyst<br />

systems are presented in the logarithmic Tafel<br />

form (Fig. 4, B and D) to offer the most condensed<br />

representation of surface characterization<br />

and polarization curves. As expected, the<br />

systematic modification of Pt(110), first with Pt<br />

islands and then with Ni(OH) 2, exhibits an HER<br />

activity trend (Fig. 4B) with the same order as the<br />

driving force for dissociative adsorption of water<br />

molecules, as discussed above for the Pt(111) systems:<br />

Pt(110) < Ni(OH) 2/Pt(110)


Surface Chemistry<br />

DOI: 10.1002/anie.201107668<br />

Unique Electrochemical Adsorption Properties of Pt-Skin Surfaces**<br />

Dennis F. van der Vliet, Chao Wang, Dongguo Li, Arvydas P. Paulikas, Jeffrey Greeley,<br />

Rees B. Rankin, Dusan Strmcnik, Dusan Tripkovic, Nenad M. Markovic, and<br />

Vojislav R. Stamenkovic*<br />

PtM alloys (M = Co, Ni, Fe, etc.) have been extensively<br />

studied for their use in fuel cells, both in well-defined<br />

extended surfaces, [1] as well as in nanoparticles. [2] After the<br />

report about exceptional activity of Pt 3Ni(111)-skin surface [1a]<br />

for the oxygen reduction reaction (ORR) a lot of efforts have<br />

been made to mimic this catalytic behavior at the nanoscale.<br />

It has been shown that a Pt 3Ni(111) crystal annealed in ultrahigh<br />

vacuum (UHV) shows an oscillating segregation profile,<br />

with the outermost layer consisting of pure platinum while the<br />

second layer is enriched in nickel compared to the bulk<br />

composition. [1a,3] Such a surface we termed Pt skin, and owing<br />

to the presence of the non-noble metal in the subsurface layer<br />

it has altered electronic properties compared to the monometallic<br />

Pt single crystal with the same orientation. Accordingly,<br />

altered electronic properties induce a change in<br />

adsorption behavior, specifically a shift of surface-oxide<br />

formation to higher potentials. [1a,b] This adsorption behavior<br />

is believed to be the origin of the high activity for the ORR.<br />

On the opposite side of the potential scale, the adsorption of<br />

hydrogenated species, denoted as underpotentially adsorbed<br />

hydrogen (H upd), is also largely affected on Pt-skin surfaces. [4]<br />

Despite numerous efforts dedicated to synthesize nanocatalysts<br />

with Pt-skin-type surfaces, [2c,5] it still remains a challenge<br />

to claim their existence at the nanoscale. To systematically<br />

resolve this issue, we attempt to provide fundamental insight<br />

into the adsorption properties of well-defined Pt-skin surfaces<br />

under relevant electrochemical conditions and to transfer that<br />

knowledge to corresponding nanocatalysts.<br />

For that reason, we first examine the formation and<br />

composition of Pt-skin surfaces by low-energy ion scattering<br />

(LEIS) and scanning tunneling microscopy (STM) in UHV,<br />

[*] <strong>Dr</strong>. D. F. van der Vliet, <strong>Dr</strong>. C. Wang, D. Li, A. P. Paulikas,<br />

<strong>Dr</strong>. D. Strmcnik, <strong>Dr</strong>. D. Tripkovic, <strong>Dr</strong>. N. M. Markovic,<br />

<strong>Dr</strong>. V. R. Stamenkovic<br />

Materials Science Division, Argonne National Lab<br />

9700 S. Cass Ave, Argonne IL (USA)<br />

E-mail: vrstamenkovic@anl.gov<br />

vrstamenkovic@anl.gov<br />

D. Li<br />

Brown University<br />

1 Prospect St., Providence, RI, 12912, USA<br />

<strong>Dr</strong>. J. Greeley, <strong>Dr</strong>. R. B. Rankin<br />

Center for Nanoscale Materials, Argonne National Lab<br />

9700 S. Cass Ave, Argonne IL (USA)<br />

[**] This work was supported by the U.S. Department of Energy, Office of<br />

Science, Office of Basic Energy Sciences, under contract No. DE-<br />

AC02-06CH11357.<br />

Supporting information for this article is available on the WWW<br />

under http://dx.doi.org/10.1002/anie.201107668.<br />

and second we study the composition of the surfaces in an<br />

electrochemical environment to establish their adsorption<br />

properties. We demonstrate by cyclic voltammetry that the<br />

surface coverage of H upd on Pt skin is about half of that found<br />

on Pt(111), whereas the surface coverage of a saturated<br />

monolayer of carbon monoxide is similar for both surfaces.<br />

This is an important finding, which provides a link towards<br />

accurate determination of the electrochemically active surface<br />

area of nanoscale catalysts. The developed methodology<br />

provides additional evidence for the existence of Pt-skin<br />

surfaces on Pt-bimetallic nanocatalysts and can substantially<br />

diminish errors in the evaluation of the real surface area and<br />

catalytic activity.<br />

A thorough examination of the Pt-skin surfaces was<br />

performed in view of their importance in electrocatalysis as<br />

well as in response to recent questions and doubts in the<br />

Figure 1. Surface characterization of Pt 3M(111) surfaces in UHV: STM<br />

images of A) sputtered and B) annealed Pt 3Ni(111) surfaces. The<br />

brightness in colors is a measure of the depth profile, with each color<br />

change marking a single atomic step. Low-energy ion scattering<br />

spectra of C) sputtered (blue traces) versus annealed (red traces)<br />

surfaces of Pt 3Ni(111) and Pt 3Co(111) crystals. D) Successive LEIS<br />

spectra of Pt 3Ni(111) surface during Ne-sputtering. The graphs in (C)<br />

and (D) are shifted, and the scale bars indicate the intensities in<br />

arbitrary units. E/E 0 is the ratio of the energy of the scattered ion<br />

beam (E) and the energy of the incident ion beam (E 0).<br />

Angew. Chem. Int. Ed. 2012, 51, 3139 –3142 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

Angewandte<br />

Chemie<br />

3139


Angewandte .<br />

Communications<br />

literature [6] on if the state of the Pt-skin surface formed over<br />

the Pt 3M systems in UHV and in an electrochemical environment<br />

consists of pure Pt. Figure 1C shows the LEIS results<br />

obtained by a 1 keV Ne + ion beam for the Pt 3Ni(111) and<br />

Pt 3Co(111) crystals before and after thermal annealing in<br />

UHV. Based on these LEIS spectra it is obvious that after<br />

consecutive annealing/sputtering cycles, both Pt and Ni (or<br />

Co) are present on the surface, while the STM image in<br />

Figure 1A illustrates the roughness of the sputtered surface.<br />

However, the LEIS signal for Ni and/or Co is diminished if<br />

annealing is the final step, which is indicative for the<br />

formation of a Pt skin. The morphology of this surface is<br />

revealed by STM (Figure 1B), which displays the smooth and<br />

ordered formation that is typical for single-crystalline systems.<br />

To investigate the stability of these surfaces in UHV,<br />

more energy was applied through the monochromatic Ne + ion<br />

beam, causing a sputtering effect to take place on the topmost<br />

Pt surface layer. Consequently, the subsurface nickel and/or<br />

cobalt atoms become exposed in consecutive LEIS spectra<br />

(shown for Pt 3Ni(111) in Figure 1D; the result for Pt 3Co(111)<br />

was identical), indicating a change in surface composition<br />

from Pt-skin (100% Pt) into Pt-rich. These combined LEIS<br />

and STM results unambiguously confirm that both crystals<br />

exhibit substantial transition of surface composition/morphology<br />

upon annealing in UHV owing to complete segregation of<br />

Pt to the surface, thereby forming a full Pt skin.<br />

Cyclic voltammetry is used to examine electrochemical<br />

adsorption properties of Pt-skin surfaces. [9] As shown in<br />

Figure 2A, the onset of H upd adsorption on Pt 3Ni(111) skin is<br />

shifted towards lower potentials when compared to Pt(111).<br />

The onset is approximately 150 mV lower and shows a reversible<br />

pre-wave ahead of the main H upd adsorption owing to<br />

surface steps, which are inevitably present on Pt 3Ni(111).<br />

Moreover, the formation of surface oxides OH ad at potentials<br />

greater than 0.6 V is delayed by 100 mV versus Pt(111). [1a]<br />

Furthermore, from the integrated H upd regions it was revealed<br />

that charges are considerably different (see Table 1), i.e., in<br />

the same potential range, Pt-skin surfaces have about half of<br />

the value obtained for Pt(111).<br />

It is important to emphasize that Pt(111) and Pt 3M(111)skin<br />

surfaces have essentially identical geometric surface<br />

areas, surface compositions, and surface structures, albeit the<br />

interatomic distances may be altered by the composition of<br />

the second layer. However, the electrochemical adsorption<br />

properties between them are quite different. This is a consequence<br />

of the electronically modified structure of Pt for Ptskin<br />

surfaces by Ni or Co subsurface atoms, which leads to<br />

Table 1: Integrated charges for Pt(111), Pt 3Ni(111), Pt 3Co (111), and<br />

polycrystalline-Pt (Pt(poly)) extended surfaces obtained from cyclic<br />

voltammetry curves for H upd (Q H), CO stripping (Q CO), and the ratio<br />

between measured charges. Analysis of the charges can be found in the<br />

Supporting Information.<br />

Catalyst Q H<br />

[mCcm 2 ]<br />

Q CO<br />

[mCcm 2 ]<br />

Q CO/2Q H<br />

Pt(111) 152 315 1.04<br />

Pt 3Ni(111) 98 304 1.55<br />

Pt 3Co(111) 91 283 1.55<br />

Pt(poly) 190 386 1.02<br />

Figure 2. Electrochemical characterization of Pt(111) (black traces)<br />

and Pt 3Ni(111)-skin surfaces (red traces) by using a rotating disc<br />

electrode (RDE) in 0.1m HClO 4: A) cyclic voltammograms; B) electrooxidation<br />

of an adsorbed monolayer of CO on Pt(111) and on<br />

Pt 3Ni(111) skin. RHE = reversible hydrogen electrode.<br />

weakened interactions between Pt and adsorbates such as<br />

H upd and OH ad. The H upd integrated charge is used as<br />

conventional approach in the estimation of the electrochemical<br />

surface area; however, the discrepancy that is coming<br />

from altered electronic properties of Pt can substantially<br />

affect the accurate estimation of the real surface area of Ptalloy<br />

catalysts.<br />

Another surface-sensitive reaction that is commonly used<br />

to probe surface properties is the electro-oxidation of<br />

adsorbed carbon monoxide, well-known as CO stripping. In<br />

this reaction, the surface is first saturated with a CO adlayer<br />

at the negative potential limit. Owing to a strong Pt–CO ad<br />

interaction, CO molecules stay adsorbed on Pt surface atoms,<br />

while the electrolyte is purged with argon gas. The consecutive<br />

step is electro-oxidation of CO by sweeping the<br />

potential towards the positive limit. The obtained CO<br />

stripping voltammetric profiles of surfaces prepared in<br />

UHV are shown in Figure 2B. The onset of CO stripping on<br />

Pt 3Ni(111) skin is 150 mV lower than that on Pt(111), while<br />

the shape of the stripping peak is much broader, and the main<br />

oxidation peak is also shifted by about 100 mV. The oxidation<br />

of adsorbed CO ad proceeds at lower potentials on Pt-skin<br />

surfaces owing to a weaker interaction of the Pt surface atoms<br />

with CO, caused by the modified electronic properties, but the<br />

3140 www.angewandte.org 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2012, 51, 3139 –3142


similar charge of CO oxidation points to an equal coverage of<br />

CO. [7] This hypothesis is supported by our density functional<br />

theory calculations, which indicate that CO binding to<br />

a Pt 3Ni(111)-skin surface, with a 1:1 Pt/Ni ratio in the<br />

subsurface layer, is approximately 0.27 eV weaker than the<br />

corresponding binding on Pt(111). [10] For that reason, the<br />

reduced binding of surface oxides (OH ad) onPt 3Ni (111) skin<br />

does not cause an increase in onset potential for CO<br />

electrooxidation. In fact, since CO ad and OH ad are competing<br />

for the same Pt adsorption sites, the reduced binding energy<br />

between Pt 3Ni (111) skin and CO ad leads to a lower onset<br />

potential as the blocking effect of CO ad is diminished (see the<br />

Supporting Information for additional computational details).<br />

It is important to mention that it was demonstrated that CO ad<br />

can modify the surface composition and/or morphology; [8]<br />

such an effect would imply a stronger Pt–CO interaction,<br />

whereas we clearly note a weakened adsorption, convincingly<br />

disproving such an interaction in the Pt-skin case. This was<br />

additionally confirmed by unchanged cyclic voltammetry<br />

after CO electrooxidation.<br />

Based on the results depicted in Figure 2B and Table 1 the<br />

onset, shape, and main oxidation peak of the CO stripping<br />

curves differ between Pt 3M(111)-skin surfaces and Pt(111),<br />

while the integrated charges are nearly equal. The similar CO<br />

stripping charges add to the LEIS evidence that the outer<br />

layer of the Pt 3M(111) skin consists of only Pt atoms. Moreover,<br />

the ratio between the charges of CO stripping versus<br />

H upd was calculated to visualize the difference in adsorption<br />

properties for all surfaces (see Table 1). Both H upd and CO<br />

stripping are surface-specific probes for Pt atoms, so in case of<br />

pure Pt systems, such as Pt(111) and polycrystalline Pt, this<br />

charge ratio is identical and close to one. However, the ratio<br />

obtained for Pt 3Ni(111)- and Pt 3Co(111)-skin surfaces is 1.55,<br />

which confirms that the surface coverage of CO ad is not<br />

affected by altered electronic properties of Pt-skin surfaces.<br />

As a consequence of these results, it is obvious that the<br />

integrated H upd region cannot be solely used in estimation of<br />

the electrochemical surface area in case of Pt-skin surfaces.<br />

For that reason, special care must be taken when analyzing<br />

annealed PtM alloys in the form of high-surface-area nanoscale<br />

catalysts.<br />

As a case in point, in Table 2 the charges for H upd and CO<br />

oxidation are compared for three different catalysts of equally<br />

sized nanoparticles (NPs): monometallic Pt/C, PtNi/C, and<br />

annealed PtNi/C NPs with a Pt skin. Detailed information on<br />

the particles size distribution and composition is presented in<br />

the Supporting Information. Transmission electron microscopy<br />

(TEM) results show that the particles size of about 5 nm<br />

and shape are not affected by thermal annealing, while<br />

energy-dispersive X-ray (EDX) line scan analyses revealed<br />

the elemental distribution across the particles. In Figure S1 in<br />

the Supporting Information, a uniform distribution of Pt and<br />

Ni is confirmed for as-prepared particles. On the other hand,<br />

after the acid leaching and annealing treatments, a Pt-overlayer<br />

has emerged on PtNi nanoparticles. Representative<br />

voltammetric curves for skin-type PtNi NPs compared to Pt/C<br />

can be seen in Figure 3. The CO stripping curves in Figure 3B<br />

closely match the results obtained on single crystals (Figure<br />

2B). For the non-annealed NPs, the ratio between the<br />

Table 2: Integrated charges and calculated surface areas (ECSAs) for<br />

H upd (Q H) and CO stripping (Q CO) obtained from cyclic voltammetries of<br />

Pt/C, acid-treated PtNi/C, and annealed PtNi/C catalysts. The ratio<br />

between the integrated charges for H upd and CO stripping demonstrates<br />

the discrepancy in ECSAs and the underestimation of the real surface<br />

area if H upd is used in case of Pt-skin surfaces.<br />

Catalyst Q H<br />

[mC]<br />

ECSA H<br />

[cm 2 ]<br />

Q CO<br />

[mC]<br />

ECSA CO<br />

[cm 2 ]<br />

Q CO/2Q H<br />

Pt/C 279 1.47 545 1.41 0.98<br />

PtNi/C 292 1.54 615 1.60 1.05<br />

PtNi skin/C 210 1.10 595 1.54 1.42<br />

charges for H upd and CO stripping is similar to Pt/C (and<br />

Pt(poly)), and thus the surface-area estimation based on the<br />

H upd charge is reasonable. However, the adsorption properties<br />

of annealed PtNi particles with a Pt-skin-type surface<br />

resemble those previously established on well-defined skintype<br />

bulk crystals of Pt 3Ni(111) and Pt 3Co(111), as judged by<br />

the suppression of H upd adsorption (measured as charge).<br />

This suppression of H upd adsorption confirms that the<br />

current focus in nanoparticle research on skin-type or core–<br />

shell structures [2c,5] can potentially suffer from substantial<br />

underestimation of the electrochemically active surface area<br />

if only the integrated H upd region is used. To be certain that<br />

the electrochemical surface area is not overlooked for skintype<br />

alloy catalysts, the estimated H upd charge always has to be<br />

Figure 3. Electrochemical surface characterization of catalysts consisting<br />

of Pt/C (black dashed traces) and PtNi/C with Pt skin (red traces)<br />

by using a RDE in 0.1m HClO 4: A) cyclic voltammograms; B) CO<br />

stripping curves.<br />

Angewandte<br />

Chemie<br />

Angew. Chem. Int. Ed. 2012, 51, 3139 –3142 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org<br />

3141


Angewandte .<br />

Communications<br />

compared to the CO stripping charge. Furthermore, the<br />

observed discrepancy between surface-area estimations based<br />

on H upd and CO stripping can give an indication of the<br />

formation of a skin-type nanocatalyst alloy, because acid<br />

leached skeleton-type surfaces and nanoparticles do not show<br />

such behavior.<br />

In summary, we have demonstrated that by alloying<br />

platinum with a non-noble second metal and inducing a Ptskin-type<br />

structure, the adsorption properties of the resulting<br />

PtM alloys are significantly altered. Specifically, the adsorption<br />

of hydrogen and oxide species is shifted in potential and<br />

reduced in magnitude compared to monometallic Pt in the<br />

same potential region. Added to LEIS evidence, the fact that<br />

the charge of the oxidation of a complete monolayer of CO is<br />

similar on annealed Pt 3M surfaces as on monometallic Pt<br />

confirms the surface consists of only platinum. The suppression<br />

of H upd adsorptionon these surfaces is both a method to<br />

confirm Pt-skin formation, but also prompts a difficulty in<br />

determining the electrochemically active surface area on Ptskin-type<br />

nanoparticles. To avoid underestimation of the<br />

surface area owing to suppression of H upd adsorption, and<br />

hence overrating of specific activity, we demonstrate that CO<br />

stripping has to be used alongside the H upd charge for the<br />

determination of the electrochemically active surface area of<br />

Pt-alloy catalysts.<br />

Experimental Section<br />

Methods: After annealing cycles in UHV, the Pt 3Ni(111) crystal was<br />

transferred to the electrochemical cell in hanging meniscus mode in<br />

which only the (111) surface is exposed to the electrolyte. The crystal<br />

was protected from the airborne impurities by a drop of hydrogensaturated<br />

Milli-Q water during transfer to the electrolyte, which was<br />

deoxygenated 0.1m HClO 4. All gases are of research grade (5N5 or<br />

higher).<br />

After voltammetry confirmed a clean and stable surface, CO was<br />

inserted into the cell for one minute, while the crystal was rotated at<br />

1600 rpm. Consecutively, rotation was stopped and argon was<br />

bubbled through the cell for 45 min to remove any trace of dissolved<br />

CO. After CO was purged from the solution, two cycles were<br />

recorded: the first one being the CO stripping voltammetry, the<br />

second to verify the absence of residual CO in solution. All<br />

electrochemical measurements were performed with a scan rate of<br />

50 mVs 1 in 0.1m HClO 4 at room temperature.<br />

The electrochemical glass cell was a standard 3-electrode cell, as<br />

used in previous experiments, [1a,b] with a Pt counter and a calomel<br />

reference electrode. All potentials in this article are given versus the<br />

reversible hydrogen electrode (RHE).<br />

Nanoparticle RDE electrodes were prepared by the solvothermal<br />

method described previously. [2b] Ni(acetate) 2·4H 2O and [Pt(acac) 2]<br />

(acac = acetylacetonate) were the precursors for the PtNi nanoparticles<br />

and added in a proper ratio, to ensure a 1:1 PtNi alloy is<br />

formed. PtNi acid-treated and PtNi acid-treated/annealed nanoparticles<br />

with the average particle size of 5 nm were supported on<br />

high-surface-area carbon and their electrochemical properties were<br />

compared to Tanaka Pt/C with similar particle size (total metal<br />

content for all catalysts was 40 %). The catalysts were deposited on<br />

6 mm glassy carbon electrode and the loading in all cases was adjusted<br />

to be 12 mg Pt/cm disc.<br />

Received: October 31, 2011<br />

Revised: December 21, 2011<br />

Published online: February 20, 2012<br />

. Keywords: adsorption · alloys · electrochemistry ·<br />

nanoparticles · platinum<br />

[1] a) V. R. Stamenkovic, B. Fowler, B. S. Mun, G. F. Wang, P. N.<br />

Ross, C. A. Lucas, N. M. Markovic, Science 2007, 315, 493 – 497;<br />

b) V. R. Stamenkovic, B. S. Mun, M. Arenz, K. J. J. Mayrhofer,<br />

C. A. Lucas, G. F. Wang, P. N. Ross, N. M. Markovic, Nat. Mater.<br />

2007, 6, 241 – 247; c) V. R. Stamenkovic, B. S. Mun, K. J. J.<br />

Mayrhofer, P. N. Ross, N. M. Markovic, J. Am. Chem. Soc.<br />

2006, 128, 8813 – 8819; d) I. E. L. Stephens, A. S. Bondarenko,<br />

F. J. Perez-Alonso, F. Calle-Vallejo, L. Bech, T. P. Johansson,<br />

A. K. Jepsen, R. Frydendal, B. P. Knudsen, J. Rossmeisl, I.<br />

Chorkendorff, J. Am. Chem. Soc. 2011, 133, 5485 – 5491.<br />

[2] a) C. Wang, D. van der Vilet, K. C. Chang, H. D. You, D.<br />

Strmcnik, J. A. Schlueter, N. M. Markovic, V. R. Stamenkovic,<br />

J. Phys. Chem. C 2009, 113, 19365 – 19368; b) C. Wang, G. F.<br />

Wang, D. van der Vliet, K. C. Chang, N. M. Markovic, V. R.<br />

Stamenkovic, Phys. Chem. Chem. Phys. 2010, 12, 6933 – 6939;<br />

c) S. Chen, W. C. Sheng, N. Yabuuchi, P. J. Ferreira, L. F. Allard,<br />

Y. Shao-Horn, J. Phys. Chem. C 2009, 113, 1109 – 1125.<br />

[3] a) B. Fowler, C. A. Lucas, A. Omer, G. Wang, V. R. Stamenkovic,<br />

N. M. Markovic, Electrochim. Acta 2008, 53, 6076 – 6080;<br />

b) S. Modak, S. Gangopadhyay, Solid State Commun. 1991, 78,<br />

429 – 432.<br />

[4] H. Schulenburg, J. Durst, E. Muller, A. Wokaun, G. G. Scherer,<br />

J. Electroanal. Chem. 2010, 642, 52 – 60.<br />

[5] a) C. Wang, D. van der Vliet, K. L. More, N. J. Zaluzec, S. Peng,<br />

S. Sun, H. Daimon, G. Wang, J. Greeley, J. Pearson, A. P.<br />

Paulikas, G. Karapetrov, D. Strmcnik, N. M. Markovic, V. R.<br />

Stamenkovic, Nano Lett. 2011, 11, 919 – 926; b) P. Strasser, S.<br />

Koh, T. Anniyev, J. Greeley, K. More, C. F. Yu, Z. C. Liu, S.<br />

Kaya, D. Nordlund, H. Ogasawara, M. F. Toney, A. Nilsson, Nat.<br />

Chem. 2010, 2, 454 – 460; c) A. R. Malheiro, J. Perez, E. I.<br />

Santiago, H. M. Villullas, J. Phys. Chem. C 2010, 114, 20267 –<br />

20271; d) C. Wang, M. Chi, D. Li, D. Strmcnik, D. van der Vliet,<br />

G. Wang, V. Komanicky, K.-C. Chang, A. P. Paulikas, D.<br />

Tripkovic, J. Pearson, K. L. More, N. M. Markovic, V. R.<br />

Stamenkovic, J. Am. Chem. Soc. 2011, 133, 14396 – 14403.<br />

[6] S. Axnanda, K. D. Cummins, T. He, D. W. Goodman, M. P.<br />

Soriaga, ChemPhysChem 2010, 11, 1468 – 1475.<br />

[7] a) A. Atli, M. Abon, J. C. Bertolini, Surf. Sci. 1993, 287 – 288,<br />

110 – 113; b) J. C. Bertolini, B. Tardy, M. Abon, J. Billy, P.<br />

Delich›re, J. Massardier, Surf. Sci. 1983, 135, 117 – 127; c) G.<br />

Chiarello, A. R. Marino, V. Formoso, A. Politano, J. Chem. Phys.<br />

2011, 134, 224705.<br />

[8] K. J. Andersson, F. Calle-Vallejo, J. Rossmeisl, L. Chorkendorff,<br />

J. Am. Chem. Soc. 2009, 131, 2404 – 2407.<br />

[9] a) D. Strmcnik, D. Tripkovic, D. van der Vliet, V. Stamenkovic,<br />

N. M. Markovic, Electrochem. Commun. 2008, 10, 1602 – 1605;<br />

b) J. M. Orts, R. Gómez, J. M. Feliu, A. Aldaz, J. Clavilier,<br />

Electrochim. Acta 1994, 39, 1519 – 1524.<br />

[10] a) B. Hammer, L. B. Hansen, J. K. Nørskov, Phys. Rev. B 1999,<br />

59, 7413 – 7421;b) L. Bengtsson, Phys. Rev. B 1999, 59, 12301 –<br />

12304;c) D. Vanderbilt, Phys. Rev. B 1990, 41, 7892 – 7895;d) G.<br />

Kresse, J. Furthmuller, Comput. Mater. Sci. 1996, 6, 15 – 50;e) F.<br />

Abild-Pedersen, J. Greeley, F. Studt, P. G. Moses, J. Rossmeisl, T.<br />

Munter, T. Bligaard, J. K. Nørskov, Phys. Rev. Lett. 2007, 99,<br />

016105;f) D. Strmcnik, D. Tripkovic, D. van der Vliet, K. C.<br />

Chang, V. Komanicky, H. You, J. Greeley, V. Stamenkovic, N. M.<br />

Markovic, J. Am. Chem. Soc. 2008, 130, 15332 – 15339.<br />

3142 www.angewandte.org 2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2012, 51, 3139 –3142


© WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim<br />

REPRINT


Reprint<br />

Heterogeneous Catalysis<br />

Tailoring the Selectivity and Stability of<br />

Chemically Modified Platinum<br />

Nanocatalysts To Design Highly Durable<br />

Anodes for PEM Fuel Cells<br />

Big and small: Chemically modifying<br />

platinum with calix[4]arene yields a highly<br />

stable anode catalyst that effectively suppresses<br />

the oxidation reduction reaction<br />

without altering the maximum activity for<br />

the hydrogen oxidation reaction (see<br />

picture, Pt blue, C gray, O red, S yellow).<br />

This behavior extends from long-rangeordered<br />

stepped single-crystal surfaces to<br />

nanocatalysts.<br />

B. Genorio, R. Subbaraman, D. Strmcnik,<br />

D. Tripkovic, V. R. Stamenkovic,<br />

N. M. Markovic* 5468 – 5472<br />

Keywords: calixarenes · electrochemistry ·<br />

heterogeneous catalysis ·<br />

oxygen reduction reaction · platinum<br />

2011 – 50/24<br />

WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


Communications<br />

Heterogeneous Catalysis<br />

Angewandte<br />

Chemie<br />

DOI: 10.1002/anie.201100744<br />

Tailoring the Selectivity and Stability of Chemically<br />

Modified Platinum Nanocatalysts To Design Highly<br />

Durable Anodes for PEM Fuel Cells**<br />

Bostjan Genorio, Ram Subbaraman, Dusan Strmcnik, Dusan Tripkovic,<br />

Vojislav R. Stamenkovic, and Nenad M. Markovic*<br />

5468 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2011, 50, 5468 –5472


An ever-changing energy landscape and the global drive<br />

toward greener energy technologies have made fuel cells a<br />

focal point of numerous research initiatives. In their current<br />

form, proton-exchange-membrane fuel cells (PEMFCs) have<br />

been shown to perform well under operating conditions for<br />

both automotive and stationary applications. In order for<br />

PEMFCs to reach commercial implementation, issues such as<br />

durability under both normal and startup and shutdown<br />

conditions have to be tackled effectively. Reactivity, selectivity,<br />

and stability are the quintessential properties that need to<br />

be tailored to develop catalysts that can tackle the durability<br />

issues arising during startup and shutdown. [1] One approach to<br />

accomplish this task is to design an anode catalyst that can<br />

efficiently suppress the undesired oxygen reduction reaction<br />

(ORR; imparting selectivity) and preserve the platinum-like<br />

hydrogen oxidation reaction (HOR) activity (imparting<br />

reactivity), while remaining stable under operating conditions<br />

(imparting stability). Such an approach not only reduces the<br />

overpotential on the cathode side owing to negligible ORR<br />

currents on the anode but also prevents formation of<br />

detrimental products such as hydrogen peroxide, which can<br />

be formed under “normal” anode startup and shutdown<br />

conditions. We have shown that chemically modified electrodes<br />

(CME) consisting of self-assembled monolayers (SAMs)<br />

of calix[4]arene molecules on extended platinum singlecrystal<br />

surfaces can selectively block the ORR without<br />

affecting the HOR activities and kinetics. [2] Usually, the<br />

lessons learned from such extended surfaces have helped in<br />

the understanding of nanocatalysts that mimic the reactivity<br />

and catalytic behavior of the extended surfaces. [3] Seldom,<br />

however, can the behavior of extended surfaces be completely<br />

translated down to the nanocatalysts.<br />

Herein, we show that the platinum modified with<br />

calix[4]arene (calix) is, in fact, one of these rare examples in<br />

which the modified nanocatalyst system behaves in line with<br />

the corresponding extended-surface system. First, we demonstrate<br />

high selectivity of the HOR on calix-modified<br />

Pt(1099){10(111) ” (100)}and Pt(1 10){2(111) ” (100)} step<br />

[*] <strong>Dr</strong>. B. Genorio, <strong>Dr</strong>. D. Strmcnik, <strong>Dr</strong>. D. Tripkovic,<br />

<strong>Dr</strong>. V. R. Stamenkovic, <strong>Dr</strong>. N. M. Markovic<br />

Materials Science Division, Argonne National Laboratory<br />

Argonne, IL 60439 (USA)<br />

E-mail: nmmarkovic@anl.gov<br />

<strong>Dr</strong>. B. Genorio<br />

Faculty of Chemistry and Chemical Technology<br />

University of Ljubljana (Slovenia)<br />

<strong>Dr</strong>. R. Subbaraman<br />

Nuclear Engineering Division<br />

Argonne National Laboratory (USA)<br />

[**] This work was supported by the Director, Office of Science, Office of<br />

Basic Energy Sciences, Division of Materials Sciences, US Department<br />

of Energy under Contract No. DE-AC03-76SF00098 and the<br />

Center of Excellence Low Carbon Technologies Slovenia (CO NOT),<br />

Center of Excellence Advanced Materials and Technologies for the<br />

Future Slovenia (CO NAMASTE). R.S. is grateful for financial<br />

support from an Argonne postdoctoral fellowship. PEM = Protonexchange<br />

membrane.<br />

Supporting information for this article is available on the WWW<br />

under http://dx.doi.org/10.1002/anie.201100744.<br />

surfaces. Then, we developed a methodology to form highly<br />

selective and stable SAMs of calix molecules on commercial<br />

nanocatalysts (3M nanostructured thin film (NSTF) [4] and<br />

Tanaka 5 nm Pt/C (TKK) catalysts). We find that if the<br />

synthesis is precisely controlled, the selectivity of nanoparticles<br />

for the ORR in the presence of hydrogen under<br />

conditions relevant to PEMFC operations is almost 100 %.<br />

We start with the electrochemical characteristics of<br />

calix[4]arene-decorated Pt(1 10) and Pt(1099). As summarized<br />

in Figure 1, both stepped surfaces show characteristic<br />

cyclic voltammograms; the under-potentially deposited (H upd)<br />

hydrogen (0–0.4 V) is followed first by a double-layer region<br />

and then at E > 0.6 V by reversible (OH ad) and irreversible<br />

oxide formation. [5] On the calix-covered surfaces, however,<br />

both the H upd and OH ad regions are significantly suppressed.<br />

In line with Ref. [2], on the highly covered surfaces the<br />

number of “free” Pt sites (determined from the H upd charge) is<br />

extremely low (ca. 2–3 %). However, on the same surface the<br />

HOR is similar to calix-free Pt, thus confirming that the<br />

turnover frequency (TOF) of the hydrogen reaction is<br />

extremely high [6] and that the Pt–H 2 energetics are not<br />

affected by the adsorbed calix molecules.<br />

More importantly, we find that at E > 0.6 V, while the<br />

ORR is almost completely inhibited (these modifications are<br />

not accompanied by undesired peroxide production), [2] the<br />

HOR is under pure diffusion control. This unique selectivity is<br />

attributed to very strong ensemble effects in which the<br />

number of bare Pt sites available for adsorption of O 2 is much<br />

smaller than that for the adsorption of H 2 and the subsequent<br />

HOR. [2] We conclude therefore that the selectivity achieved<br />

using calix-modified electrodes is not affected by the presence<br />

of steps. This result is very important because it provides<br />

evidence that such behavior can be successfully translated to<br />

the most commonly used forms of nanocatalysts, which are<br />

known to contain a vast majority of such sites (steps and<br />

short-range terraces).<br />

Having established the behavior of well-defined surfaces,<br />

we move on to the most relevant electrocatalyst systems:<br />

nanocatalysts. To encompass a wide range of electrocatalyst<br />

designs and properties, we provide an analysis for the two<br />

most commonly used commercial electrocatalysts. The TKK<br />

catalyst and the 3M NSTF catalyst were both studied<br />

(Figure 2). The TKK catalyst represents supported nanocatalysts,<br />

where platinum nanoparticles 2–10 nm in diameter<br />

are supported on amorphous carbon black. NSTF catalysts,<br />

comprised of a unique catalyst structure which is free of<br />

carbon support, are usually applied directly to the membrane<br />

to provide a compact membrane electrode assembly structure<br />

(Figure 2). Aqueous electrochemical experiments conducted<br />

using the RDE/RRDE (RDE = rotating disk electrode,<br />

RRDE = rotating ring disk electrode) methods for these<br />

nanocatalysts are well-established [7] and have been shown to<br />

correlate very well with operating fuel-cell systems.<br />

We present herein results obtained from the RDE study<br />

that should be relevant for operating fuel-cell systems.<br />

Various modifications of the calix molecules were studied,<br />

including the thiolated derivatives of calix[6]arenes and<br />

calix[8]arenes (see the Supporting Information for the synthesis),<br />

but only the derivatives of the calix[4]arene family<br />

Angew. Chem. Int. Ed. 2011, 50, 5468 –5472 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org<br />

5469


Communications<br />

Figure 1. Electrochemical characteristics of calix-modified stepped surfaces: a) Pt(1099), b) Pt(110). Potentials of interest during startup and<br />

shutdown are shown in the shaded region. [1,2]<br />

Figure 2. a) SEM image of an unmodified Pt nanowhisker. b) TEM<br />

morphology of typical TKK nanocatalysts. Calix molecules are not<br />

visible by electron microscopy. Based on our previous STM study, [2] we<br />

present a schematic representation of calix-modified nanocatalysts:<br />

c) Model morphology of calix[4]arene (yellow-gray-red)-modified Pt<br />

NSTF nanowhisker (blue). d) Model for TKK nanocatalyst chemically<br />

modified with calix[4]arene.<br />

proved to be effective in achieving the selectivity, so only the<br />

results pertaining to the latter are presented. As can be seen<br />

in Figure 3, the calix[4]arene molecules are found to suppress<br />

the H upd region (0.05–0.4 V) for both NSTF and TKK<br />

catalysts. The relative coverages for similar methods of<br />

preparation are slightly different, but the net results appear<br />

to be the same: an exceptional selectivity for the HOR versus<br />

ORR. As for stepped surfaces discussed above, the diffusion-<br />

limiting currents for the HOR are observed at potentials<br />

above 0.1 V and the activities below 0.1 V are, within the<br />

experimental limits, almost identical.<br />

Furthermore, the ORR polarization curves show limited<br />

or insignificant currents in the potential region of interest for<br />

the anode-side catalyst. As was shown in the earlier study with<br />

Pt(111), [2] the peroxide yield on all extended and nanoparticle<br />

Pt–calix systems is negligible above 0.6 V, and the overall<br />

ORR behavior of these surfaces mimic ORR on uncovered or<br />

partially covered patches. [6] All of these observations suggest<br />

that SAMs of calix molecule can be used to tailor the<br />

selectivity of the nanocatalyst toward ORR while preserving<br />

the HOR activity, the goal for an ideal anode catalyst. It is<br />

also important that the established selectivity was possible<br />

only because the required number of active sites for maximal<br />

rates of the HOR is, in fact, extremely small but is sufficient to<br />

provide enough sites for the diffusion-limiting currents. [2]<br />

In addition to selectivity of CME, both thermal and<br />

electrochemical stability of these electrodes are important<br />

properties that need to be addressed to evaluate the anode<br />

catalyst s applicability to PEMFC. In order to study the<br />

stability of calix-modified electrodes, we tested a Pt–calix<br />

system in an oxygen-rich environment at 0.8 V for approximately<br />

14 h in solution at 608C. These conditions are<br />

expected to be harsher than those experienced by the<br />

electrode in a real fuel-cell system. The exposure of the<br />

anode catalyst to high potentials (E < 0.8 V for anode) in an<br />

air (oxygen)-rich atmosphere during startup and shutdown is<br />

expected to last between tens of seconds and a few minutes a<br />

day. The temperatures are expected to be similar to those<br />

used in our test conditions.<br />

Figure 4 shows the current–time relationship for the CME<br />

held at 0.8 V in an oxygen-rich atmosphere. The ORR current<br />

5470 www.angewandte.org 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2011, 50, 5468 –5472


Figure 3. Electrochemical characteristics of calix-modified Pt nanocatalysts: a) NSTF, b) TKK (50% Pt loading). Catalyst loadings were<br />

approximately 14–16 mgcm 2 ; lower loadings were also used to mimic anode catalyst performance, yielding similar qualitative results. All current<br />

densities are given with respect to the disk geometric area.<br />

actually shows a small decay, thus suggesting that there is no<br />

loss of the calix molecules from the surface owing to<br />

oxidation. (Removal of the molecules by oxidation or<br />

desorption would increase the reduction current.) A similar<br />

experiment was also performed for the nanocatalysts (TKK)<br />

modified with calix[4]arene molecules, which show qualitatively<br />

similar results. This finding suggests that the calix[4]arene-modified<br />

electrodes are stable under these operating<br />

conditions. Moreover, during the long-term experiments, the<br />

HOR (results not shown) is not affected at all.<br />

In conclusion, our CMEs prepared by modifying Pt with<br />

calix[4]arene molecules are highly stable and can effectively<br />

Figure 4. Stability of the Pt–calix system at 60 8C inO 2-saturated 0.1m<br />

HClO 4 at 0.8 V. Inset: ORR curves for unmodified surface and Pt–calix<br />

surface both before and after the stability test. Note: the HOR remains<br />

unchanged for the duration of the experiment.<br />

tune the selectivity of anode catalysts for ORR without<br />

altering the maximum activity of the HOR. This behavior is<br />

highly transformational, extending from long-range-ordered<br />

stepped single-crystal surfaces to nanocatalysts. The CME<br />

approach is not restricted to a Pt–calix system, and we<br />

envision it to provide many applications in analytical,<br />

synthetic, and materials chemistry as well as in chemical<br />

energy conversion and storage.<br />

Experimental Section<br />

Synthesis of the thiolated derivative of calix[4]arene: Adsorption of<br />

organic groups on noble-metal surfaces has been well-established for<br />

various groups (-S, -CN, -A, where A denotes the anchoring group). [8]<br />

The driving force for ordering of such large molecules is presumable<br />

governed by a synergy between the strong chemical bond between the<br />

anchoring groups and Pt surface atoms and the local steric interaction<br />

between adsorbed molecules. The calix[4]arene molecules anchoring<br />

groups are usually of the form S(R) where (R) is used to cap the thiol<br />

group on the quadrupole anchoring groups. A detailed description of<br />

the synthesis procedure as well as molecular designs considered are<br />

presented in the Supporting Information.<br />

Preparation of Pt(1099), Pt(110), and Pt(polycrystalline) surfaces<br />

and self-assembly: Pt electrodes were prepared by inductive<br />

heating for 10 min at approximately 1100 K in an argon–hydrogen<br />

flow (3 % hydrogen). The annealed specimen was cooled slowly to<br />

room temperature in this flow stream and immediately covered by a<br />

droplet of water. The electrode was then immersed in a THF solution<br />

of calix[4]arene for 24 h, allowing the formation of a calix[4]arene<br />

SAM. The concentration of calix[4]arene in THF was 600 mm to<br />

obtain samples with very high coverages of calix on Pt surface.<br />

Coverages were estimated from the H upd measurements. The effect of<br />

coverage on ORR and HOR was previously presented. [3] The<br />

coverages can be modified by either varying the concentrations of<br />

the calix/THF solution or the exposure time to the high-concentration<br />

solution. After SAM preparation, the crystals were washed thor-<br />

Angew. Chem. Int. Ed. 2011, 50, 5468 –5472 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.angewandte.org<br />

5471


Communications<br />

oughly with deionized water before assembly and immersion in the<br />

electrochemical cell.<br />

Preparation of NSTF and TKK catalyst electrodes and their selfassembly:<br />

The catalysts were mixed with water at a concentration of<br />

1mgmL 1 . This dispersion was then ultrasonically mixed for one<br />

hour, after which a stable suspension was obtained. A glassy carbon<br />

disk (6 mm diameter) was then mechanically polished. Known<br />

volumes of the suspensions were then added using a micropipette<br />

onto the glassy carbon disk electrode. The electrode was dried at 60 8C<br />

in an inert atmosphere. The suspension was applied so that it coated<br />

the surface of the electrode very uniformly. Once dry, these electrodes<br />

were washed with water to verify the good adhesion of particles to the<br />

glassy carbon substrate. Subsequently, the electrodes were immersed<br />

in 1000 mm solution of calix in THF. We chose to use a high<br />

concentration of calix owing to the larger surface area of Pt compared<br />

to the disk electrodes. The systems were equilibrated for 24 h.<br />

Another method involved assembly of the disk electrode in a hanging<br />

meniscus arrangement with subsequent immersion of the electrode in<br />

the calix solution with rotation (600 rpm) for 4 h. Both of these<br />

methods yielded similar coverages. After equilibration, the samples<br />

were washed thoroughly with water before being immersed in the<br />

electrochemical cell. For a discussion on the relative coverages<br />

obtained for the same conditions for TKK and NSTF, please refer to<br />

the Supporting Information.<br />

RDE method, electrolytes, and electrochemical setup: After<br />

extensive rinsing, the electrode was embedded into the rotating-disk<br />

electrode (RDE) and transferred into a standard three-compartment<br />

electrochemical cell containing 0.1m HClO 4 (Sigma–Aldrich). In each<br />

experiment, the electrode was immersed at 0.07 V in solution<br />

saturated with Ar. After obtaining a stable voltammogram between<br />

0.07 and 0.7 V the polarization curve for the ORR was recorded on<br />

the disk on the disk electrode. Subsequently, oxygen was purged out<br />

of the solution and replaced with hydrogen, and HOR polarization<br />

curves were measured. Finally, the voltammetric response was again<br />

recorded in argon-purged solution to confirm that calix coverages had<br />

not changed significantly.<br />

All gases were 5N5 quality purchased from Airgas Inc. The sweep<br />

rate for all measurements was 50 mVs 1 ; for the ORR measurements,<br />

the electrode was rotated at 1600 rpm. Electrode potentials are given<br />

versus the reversible hydrogen electrode (RHE).<br />

Received: January 29, 2011<br />

Published online: May 12, 2011<br />

. Keywords:<br />

calixarenes · electrochemistry ·<br />

heterogeneous catalysis · oxygen reduction reaction · platinum<br />

[1] R. Borup, J. Meyers, B. Pivovar, Y. S. Kim, R. Mukundan, N.<br />

Garland, D. Myers, M. Wilson, F. Garzon, D. Wood, P. Zelenay, K.<br />

More, K. Stroh, T. Zawodzinski, J. Boncella, J. E. McGrath, M.<br />

Inaba, K. Miyatake, M. Hori, K. Ota, Z. Ogumi, S. Miyata, A.<br />

Nishikata, Z. Siroma, Y. Uchimoto, K. Yasuda, K. I. Kimijima, N.<br />

Iwashita, Chem. Rev. 2007, 107, 3904 – 3951; C. A. Reiser, L.<br />

Bregoli, T. W. Patterson, J. S. Yi, J. D. L. Yang, M. L. Perry, T. D.<br />

Jarvi, Electrochem. Solid-State Lett. 2005, 8, A273 – A276; J. P.<br />

Meyers, R. M. Darling, J. Electrochem. Soc. 2006, 153, A1432 –<br />

A1442; R. A. Sidik, J. Solid State Electrochem. 2009, 13, 1123 –<br />

1126.<br />

[2] B. Genorio, D. Strmcnik, R. Subbaraman, D. Tripkovic, G.<br />

Karapetrov, V. R. Stamenkovic, S. Pejovnik, N. M. Markovic,<br />

Nat. Mater. 2010, 9, 998 – 1003.<br />

[3] H. A. Gasteiger, N. M. Markovic, Science 2009, 324,48–49;V.R.<br />

Stamenkovic, B. S. Mun, M. Arenz, K. J. J. Mayrhofer, C. A.<br />

Lucas, G. F. Wang, P. N. Ross, N. M. Markovic, Nat. Mater. 2007, 6,<br />

241 – 247.<br />

[4] M. K. Debe, A. K. Schmoeckel, G. D. Vernstrorn, R. Atanasoski,<br />

J. Power Sources 2006, 161, 1002 – 1011.<br />

[5] N. M. Markovic, P. N. Ross, Jr., Surf. Sci. Rep. 2002, 45, 117 – 229.<br />

[6] D. S. Strmcnik, P. Rebec, M. Gaberscek, D. Tripkovic, V.<br />

Stamenkovic, C. Lucas, N. M. Markovic, J. Phys. Chem. C 2007,<br />

111, 18672 – 18678.<br />

[7] K. J. J. Mayrhofer, D. Strmcnik, B. B. Blizanac, V. Stamenkovic,<br />

M. Arenz, N. M. Markovic, Electrochim. Acta 2008, 53, 3181 –<br />

3188.<br />

[8] J. C. Love, L. A. Estroff, J. K. Kriebel, R. G. Nuzzo, G. M.<br />

Whitesides, Chem. Rev. 2005, 105, 1103 – 1169.<br />

5472 www.angewandte.org 2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Angew. Chem. Int. Ed. 2011, 50, 5468 –5472


Surfactant Removal for Colloidal Nanoparticles from Solution<br />

Synthesis: The Effect on Catalytic Performance<br />

Dongguo Li, †,‡ Chao Wang, † Dusan Tripkovic, † Shouheng Sun,* ,‡ Nenad M. Markovic, †<br />

and Vojislav R. Stamenkovic* ,†<br />

† Materials Science Division, Argonne National Laboratory, Argonne, Illinois 60439, United States<br />

‡ Department of Chemistry, Brown University, Providence, Rhode Island 02912, United States<br />

*S Supporting Information<br />

ABSTRACT: Colloidal nanoparticles prepared by solution<br />

synthesis with robust control over particle size, shape,<br />

composition, and structure have shown great potential for<br />

catalytic applications. However, such colloidal nanoparticles are<br />

usually capped with organic ligands (as surfactants) and cannot<br />

be directly used as catalyst. We have studied the effect of<br />

surfactant removal on the electrocatalytic performance of Pt<br />

nanoparticles made by organic solution synthesis. Various<br />

methods were applied to remove the oleylamine surfactant,<br />

which included thermal annealing, acetic acid washing, and UV-Ozone irradiation, and the treated nanoparticles were applied as<br />

electrocatalysts for the oxygen reduction reaction. It was found that the electrocatalytic performance, including electrochemically<br />

active surface area and catalytic activity, was strongly dependent on the pretreatment. Among the methods studied here, lowtemperature<br />

thermal annealing (∼185 °C) in air was found to be the most effective for surface cleaning without inducing particle<br />

size and morphology changes.<br />

KEYWORDS: nanoparticles, organic solution synthesis, surfactant removal, catalysis, oxygen reduction reaction<br />

Recently colloidal nanoparticles from aqueous 1−4 or<br />

organic, 5−7 solution synthesis have attracted increasing<br />

interest for the development of advanced catalysts. Through<br />

controlled nucleation and growth in a homogeneous solution,<br />

this method has been demonstrated to be powerful in<br />

controlling nanoparticle (NP) size, 6,8−10 shape, 1,11,12 composition<br />

13−15 and structure, 16−19 which has enabled systematic<br />

studies of these parameters in catalysis. 5,10,18−21 However, the<br />

NPs obtained by solution synthesis are usually capped by<br />

organic surfactants and cannot be directly applied as catalysts.<br />

For example, in the organic solution synthesis of Pt and Ptbased<br />

alloy NPs, oleylamine and/or oleic acid are usually used<br />

to stabilize the NPs and control NP sizes and shapes. 11,18,22,23<br />

While these ligands are indispensible in the NP synthesis, they<br />

are detrimental for catalysis as they block the access of reactant<br />

molecules to the surface atoms. Therefore, it is critical to<br />

establish a proper procedure of surfactant removal and surface<br />

cleaning, without inducing particle size and morphology<br />

changes, for application of colloidal NPs in catalysis.<br />

Here we report a systematical investigation of the effect of<br />

surfactant removal on the electrocatalytic performance of Pt<br />

NPs prepared by organic solution synthesis. Various methods<br />

were applied to remove the oleylamine surfactant, including<br />

thermal annealing, acetic acid washing, and UV-Ozone<br />

irradiation, and the efficiency of surface cleaning was examined<br />

by electrocatalytic studies of the treated catalysts for the oxygen<br />

reduction reaction (ORR). The removal of organic surfactants<br />

from the NP surface was further confirmed by thermogravi-<br />

metric analysis (TGA) and infrared adsorption spectroscopy<br />

(IRAS) studies.<br />

■ METHODS<br />

Synthesis. For the synthesis of Pt NPs, 50 mg of platinum<br />

acetylacetonate (Pt(acac) 2) was dissolved in 10 mL of<br />

oleylamine at 100 °C under Ar flow. After 20 min, a mixture<br />

of 0.2 g of borane tributylamine (Aldrich) and 5 mL of<br />

oleylamine was added into the solution. The solution<br />

temperature was raised to 120 °C and kept at this temperature<br />

for an hour. The solution was then cooled to room<br />

temperature, and 30 mL of ethanol was added. The product<br />

was collected by centrifugation (8,000 rpm, 6 min). The<br />

obtained precipitation was dispersed in hexane for further<br />

experiments.<br />

Catalyst Preparation and Surfactant Removal. The assynthesized<br />

Pt NPs were washed by ethanol and then dried in<br />

vacuum. After that, they were mixed with carbon black<br />

(Tanaka, ∼900 m 2 /g) with a mass ratio of 1:1 in chloroform<br />

(CHCl 3) by sonication. The solvent was then evaporated at<br />

room temperature in air. The obtained powder was denoted as<br />

“untreated” catalyst.<br />

The following treatments were applied to the untreated<br />

catalyst for surfactant removal: (1) Thermal annealing. The<br />

Received: January 24, 2012<br />

Published: May 21, 2012<br />

Letter<br />

pubs.acs.org/acscatalysis<br />

© 2012 American Chemical Society 1358 dx.doi.org/10.1021/cs300219j | ACS Catal. 2012, 2, 1358−1362


ACS Catalysis Letter<br />

Figure 1. TEM images of (a) as-synthesized Pt NPs, (b) Pt NPs loaded on carbon black, and (c) Pt NPs on carbon black after 185 °C annealing.<br />

Figure 2. (a) Cyclic voltammograms of various Pt/C catalysts. (b) ORR polarization curves of the annealed, the UV-ozone treated and the untreated<br />

samples. (c) Tafel plots of Pt/C catalysts at 20 mV/s, 1600 rpm, 20 °C. (d) Specific activity at 0.90 V and specific surface area.<br />

catalyst was heated in a tube furnace for 5 h at 185 °C in air;<br />

(2) Chemical washing. The catalyst was dispersed in pure acetic<br />

acid (HAc) with vigorous stirring at 75 °C for 10 h, and then<br />

collected by adding ethanol and centrifugation; (3) UV-Ozone<br />

treatment. An aqueous suspension of the untreated catalyst was<br />

deposited onto a glassy carbon electrode. After drying in air, the<br />

electrode was placed in a UV-Ozone chamber and sub<strong>je</strong>ct to<br />

UV irradiation for 30 min.<br />

Characterizations. Transmission electron microscopy<br />

(TEM) images were collected on a Philips EM 30 (200 kV).<br />

X-ray diffraction (XRD) patterns were collected on a<br />

PANalytical X’pert PRO diffractometer using Cu Kα radiation<br />

at room temperature. Thermogravimetric analysis for the Pt<br />

NPs was done with a Pyris 1 TGA/HT Lab System<br />

(PerkinElmer). Fourier transformed infrared (FTIR) spectra<br />

were collected on a Nicolet Nexus 8700 spectrometer with a<br />

MCT detector cooled by liquid nitrogen and in situ electrode<br />

potential control, following the previous reported procedures. 24<br />

1359<br />

Electrochemical Studies. All electrochemical measurements<br />

were performed in a three-compartment electrochemical<br />

cell in 0.1 M perchloric acid at room temperature. A Ag/AgCl<br />

electrode was used as the reference electrode, and a platinum<br />

wire as the counter electrode. A mirror-polished glassy carbon<br />

disk (6 mm in diameter) was used as the working electrode.<br />

The catalyst was dispersed in deionized water by sonication and<br />

30−40 μL of the suspension was deposited onto the glassy<br />

carbon electrode by pipet. The loading of Pt was controlled to<br />

be ∼20 μg/cm 2 disk. A solution of Nafion (0.1 wt %, 15 μL) was<br />

added on top of the catalyst and then dried in air. Cyclic<br />

voltammograms (CVs) were recorded in Ar saturated electrolytes<br />

with a scan rate of 50 mV/s. Polarization curves for the<br />

ORR were recorded in an oxygen saturated electrolytes at 20<br />

mV/s with an electrode rotation speed of 1600 rpm. All the<br />

potentials given were calibrated versus reversible hydrogen<br />

electrode (RHE).<br />

dx.doi.org/10.1021/cs300219j | ACS Catal. 2012, 2, 1358−1362


■ RESULTS AND DISCUSSION<br />

Monodisperse Pt NPs were synthesized by reduction of<br />

Pt(acac) 2 with borane tributylamine in an oleylamine solution<br />

(see the Methods). Here oleylamine served as both solvent and<br />

surfactant. Figure 1a shows a representative TEM image of the<br />

as-synthesized Pt NPs with a particle size of 2.8 ± 0.4 nm<br />

(Supporting Information, Figure S1a). The synthesized NPs<br />

were supported on high-surface-area carbon (Tanaka, ∼900<br />

m2 /g) and various treatments were applied to remove the<br />

organic surfactants and clean the surface. These include mildtemperature<br />

heating (160−200 °C) in air, 25 chemical washing<br />

by acetic acid, 26 and UV-Ozone irradiation. 27 No significant<br />

change in particle size or morphology was found after these<br />

treatments, as shown by the TEM images presented in Figure 1,<br />

for example, for the 185 °C thermal treatment in air (Figure<br />

1b,c and Supporting Information, Figure S1b). X-ray powder<br />

diffraction patterns of the as-synthesized Pt NPs, 185 °C<br />

annealed and HAc treated catalysts were presented in<br />

Supporting Information, Figure S2, which shows no significant<br />

particle size change or coarsening after these treatments.<br />

Catalytic performance of the various treated catalysts was<br />

examined by the RDE method. Figure 2 summarizes the results<br />

of electrochemical studies. CVs show typical Pt like features<br />

with under potential deposited hydrogen (Hupd) peaks at E <<br />

0.4 V (Figure 2a). Among the various treated catalysts, the 185<br />

°C annealed one shows more pronounced Hupd peaks than the<br />

others. As the catalyst loading was consistently controlled to be<br />

∼20 μg/cm2 disk, this difference directly corresponds to the<br />

variation in specific surface area, namely, the electrochemically<br />

active surface area (ECSA, estimated from the Hupd charges by<br />

ECSA = QHupd/210 μC/cm2 ) normalized by the mass of Pt in<br />

the catalyst. As shown in Figure 2d, the 185 °C annealed<br />

catalyst gave a specific surface area of ∼330 cm2 /gPt, versus<br />

∼270 cm2 /gPt for the HAc washed, 240 cm2 /gPt for the UV-<br />

Ozone treated, and 190 cm2 /gPt for the untreated catalyst.<br />

Moreover, the electrocatalytic activity was found to be also<br />

strongly dependent on the methods of surfactant removal. Both<br />

the polarization curves (Figure 2b) and Tafel plots (Figure 2c)<br />

show that the 185 °C annealed catalyst has the highest ORR<br />

activity, reaching 0.64 mA/cm2 at 0.9 V, compared to 0.49 mA/<br />

cm2 for the HAc washed, 0.29 mA/cm2 for the UV-Ozone<br />

treated, and 0.22 mA/cm2 total loss in weight, ∼45%, corresponds to the weight ratio of<br />

surfactants in the NPs. On the basis of the TGA studies, we can<br />

conclude that the thermal annealing at 185 °C in air is efficient<br />

for removing the organic surfactants from the Pt NP surface. It<br />

is also important to point out that the employed carbon black is<br />

stable in the surfactant removal process, as no weight change<br />

was observed in TGA for the carbon support at 185 °C.<br />

After surfactant removal, surface atoms on the Pt NPs were<br />

exposed and became electrochemically active. This was further<br />

confirmed by in situ electrochemical IRAS studies, in which<br />

surface conditions of the NPs as well as the CO-Pt interaction<br />

for the untreated catalyst. Both the<br />

results of specific surface areas and specific activities indicate<br />

that the efficiency of surfactant removal follows the trend 185<br />

°C annealing > HAc washing > UV-Ozone (Figure 2d). It has<br />

to be pointed out that potential cycling can bring additional<br />

cleaning for the catalysts. For example, after one hundred<br />

potential cycles, the HAc washed Pt/C shows comparable<br />

electrocatalytic performance as the thermal annealed one<br />

(Supporting Information, Figure S3). However, such potential<br />

cycling had very limited effect on the UV-Ozone treated and<br />

untreated catalysts. In the following discussion, we will focus on<br />

the study of surfactant removal by thermal annealing to confirm<br />

the effectiveness of surfactant removal.<br />

Figure 3 shows the weight loss curves for the process of<br />

surfactant removal by 185 °C annealing in air for the assynthesized<br />

Pt NPs and carbon support. It was found that a<br />

weight loss of up to 44% took place during the isothermal<br />

process at 185 °C for the NPs, because of the removal of<br />

organic surfactants. Not much organic substances were left after<br />

the 185 °C stage, as further increase of the temperature up to<br />

900 °C did not lead to substantial loss in addition (∼1%). The<br />

24<br />

was examined. Figure 4a shows the IR spectra of the annealed<br />

and untreated NPs collected at 0.1 V in 0.1 M HClO4 electrolyte. The insets of closeup view highlight the difference<br />

between the two spectra in the region from 2800 cm−1 to 3000<br />

cm−1 . The bands at 2850 cm−1 and 2920 cm−1 observed in the<br />

untreated NPs can be associated to the C−H stretching<br />

vibrations in oleylamine molecules. 28 After the thermal<br />

annealing, the C−H bands disappeared, suggesting successful<br />

removal of surfactants from the particle surface. Figure 4b<br />

shows the differentiation spectra obtained by subtracting the IR<br />

spectra recorded at 0.1 V with the spectra at 0.9 V (see<br />

Supporting Information, Figure S4 for the primary spectra).<br />

Considering that the adsorbed CO (COad) is oxidized at high<br />

potentials (E = 0.9 V) while at low potentials CO is strongly<br />

adsorbed on Pt, the differentiation spectra are thus able to<br />

exclusively tell the situation of CO adsorption on the particle<br />

surface. The untreated NPs do not exhibit any observable<br />

feature in the wavenumber range from 1800 cm−1 to 2800 cm−1 because of the blocked CO adsorption on the particle surface.<br />

The annealed NPs show a sharp peak at 2060 cm−1 which can<br />

be assigned to atop-bonded CO on Pt, 29 indicating that the<br />

surface atoms were activated by the thermal annealing and<br />

became accessible for CO adsorption. The positive peak around<br />

2340 cm−1 can be associated with CO2 which was the product<br />

of CO oxidation. Consistent with the observations from<br />

electrochemical and TGA studies (Figure 2 and 3), the IRAS<br />

results demonstrate that organic surfactants present on the NP<br />

surface, which block Hupd and CO adsorption and ORR, can be<br />

sufficiently removed by the thermal annealing at a mild<br />

temperature in air.<br />

■ CONCLUSION<br />

ACS Catalysis Letter<br />

Figure 3. TGA of the as-synthesized Pt NPs and carbon support.<br />

We have studied the effect of surfactant removal on the<br />

electrocatalytic performance of Pt nanoparticles made by<br />

1360<br />

dx.doi.org/10.1021/cs300219j | ACS Catal. 2012, 2, 1358−1362


ACS Catalysis Letter<br />

Figure 4. (a) FTIR spectra of adsorbed CO on untreated and annealed Pt NPs. The spectra were recorded after CO adsorption at 0.1 V. The insets<br />

show the region of the spectra that is typical for surfactant molecules. (b) Subtracted IR spectra at 0.9 and 0.1 V for annealed and untreated samples.<br />

organic solution synthesis. Various methods existing in the<br />

literature were applied to remove the oleylamine surfactant, and<br />

the dependence of electrocatalytic performance for the ORR on<br />

the treatment was found to follow the trend: 185 °C annealing<br />

in air > HAc washing > UV-Ozone. The effectiveness of<br />

surfactant removal and surface cleaning by thermal annealing in<br />

air was further confirmed by TGA and IRAS studies. Our work<br />

revealed the importance of pretreatment in catalyst preparation<br />

and would have great implication for the development of<br />

advanced catalysts with colloidal nanoparticles from solution<br />

■synthesis.<br />

ASSOCIATED CONTENT<br />

*S Supporting Information<br />

Further details are given in Figures S1−S4. This material is<br />

available<br />

■<br />

free of charge via the Internet at http://pubs.acs.org.<br />

AUTHOR INFORMATION<br />

Corresponding Author<br />

*E-mail: ssun@brown.edu (S.S.), vrstamenkovic@anl.gov<br />

(V.R.S.).<br />

Funding<br />

This work was conducted at Argonne National Laboratory, a<br />

U.S. Department of Energy, Office of Science Laboratory,<br />

operated by UChicago Argonne, LLC, under contract no. DE-<br />

AC02−06CH11357. This research was sponsored by the U.S.<br />

Department of Energy, Office of Energy Efficiency and<br />

Renewable Energy, Fuel Cell Technologies Program. Microscopy<br />

research was conducted at the Electron Microscopy<br />

Center for Materials Research at Argonne.<br />

1361<br />

Notes<br />

The authors declare no competing financial interest.<br />

■ REFERENCES<br />

(1) Ahmadi, T. S.; Wang, Z. L.; Green, T. C.; Henglein, A.; El-Sayed,<br />

M. A. Science 1996, 272, 1924−1925.<br />

(2) Lee, H.; Habas, S. E.; Kweskin, S.; Butcher, D.; Somorjai, G. A.;<br />

Yang, P. Angew. Chem., Int. Ed. 2006, 45, 7824−7828.<br />

(3) Habas, S. E.; Lee, H.; Radmilovic, V.; Somorjai, G. A.; Yang, P.<br />

Nat. Mater. 2007, 6, 692−697.<br />

(4) Lim, B.; Jiang, M.; Camargo, P. H. C.; Cho, E. C.; Tao, J.; Lu, X.;<br />

Zhu, Y.; Xia, Y. Science 2009, 324, 1302−1305.<br />

(5) Wang, C.; Daimon, H.; Lee, Y.; Kim, J.; Sun, S. J. Am. Chem. Soc.<br />

2007, 129, 6974−6975.<br />

(6) Tsung, C.-K.; Kuhn, J. N.; Huang, W.; Aliaga, C.; Hung, L.-I.;<br />

Somorjai, G. A.; Yang, P. J. Am. Chem. Soc. 2009, 131, 5816−5822.<br />

(7) Zhang, J.; Yang, H.; Fang, J.; Zou, S. Nano Lett. 2010, 10, 638−<br />

644.<br />

(8) Puntes, V. F.; Krishnan, K. M.; Alivisatos, A. P. Science 2001, 291,<br />

2115−2117.<br />

(9) Sun, S.; Zeng, H. J. Am. Chem. Soc. 2002, 124, 8204−8205.<br />

(10) Wang, C.; van der Vliet, D.; Chang, K.-C.; You, H.; Strmcnik,<br />

D.; Schlueter, J. A.; Markovic, N. M.; Stamenkovic, V. R. J. Phys. Chem.<br />

C 2009, 113, 19365−19368.<br />

(11) Wang, C.; Hou, Y.; Kim, J.; Sun, S. Angew. Chem., Int. Ed. 2007,<br />

46, 6333−6335.<br />

(12) Sun, Y.; Xia, Y. Science 2002, 298, 2176−2179.<br />

(13) Swafford, L. A.; Weigand, L. A.; Bowers, M. J.; McBride, J. R.;<br />

Rapaport, J. L.; Watt, T. L.; Dixit, S. K.; Feldman, L. C.; Rosenthal, S.<br />

J. J. Am. Chem. Soc. 2006, 128, 12299−12306.<br />

(14) Wang, C.; Chi, M.; Wang, G.; van der Vliet, D.; Li, D.; More, K.;<br />

Wang, H.; Schlueter, J. A.; Markovic, N. M.; Stamenkovic, V. R. Adv.<br />

Funct. Mater. 2011, 21, 147−152.<br />

dx.doi.org/10.1021/cs300219j | ACS Catal. 2012, 2, 1358−1362


ACS Catalysis Letter<br />

(15) Chen, W.; Kim, J.; Sun, S.; Chen, S. J. Phys. Chem. C 2008, 112,<br />

3891−3898.<br />

(16) Cheng, K.; Peng, S.; Xu, C.; Sun, S. J. Am. Chem. Soc. 2009, 131,<br />

10637−10644.<br />

(17) Yavuz, M. S.; Cheng, Y.; Chen, J.; Cobley, C. M.; Zhang, Q.;<br />

Rycenga, M.; Xie, J.; Kim, C.; Song, K. H.; Schwartz, A. G.; Wang, L.<br />

V.; Xia, Y. Nat. Mater. 2009, 8, 935−939.<br />

(18) Wang, C.; van der Vliet, D.; More, K. L.; Zaluzec, N. J.; Peng, S.;<br />

Sun, S.; Daimon, H.; Wang, G.; Greeley, J.; Pearson, J.; Paulikas, A. P.;<br />

Karapetrov, G.; Strmcnik, D.; Markovic, N. M.; Stamenkovic, V. R.<br />

Nano Lett. 2011, 11, 919−926.<br />

(19) Wang, C.; Daimon, H.; Sun, S. Nano Lett. 2009, 9, 1493−1496.<br />

(20) Lim, B.; Jiang, M.; Camargo, P. H. C.; Cho, E. C.; Tao, J.; Lu,<br />

X.; Zhu, Y.; Xia, Y. Science 2009, 324, 1302−1305.<br />

(21) Zhang, J.; Yang, H.; Fang, J.; Zou, S. Nano Lett. 2010, 10, 638−<br />

644.<br />

(22) Sun, S. Adv. Mater. 2006, 18, 393−403.<br />

(23) Mazumder, V.; Chi, M.; More, K. L.; Sun, S. J. Am. Chem. Soc.<br />

2010, 132, 7848−7849.<br />

(24) Stamenković, V.; Arenz, M.; Ross, P. N.; Marković, N.M.J.<br />

Phys. Chem. B 2004, 108, 17915−17920.<br />

(25) Liu, Z.; Shamsuzzoha, M.; Ada, E. T.; Reichert, W. M.; Nikles,<br />

D. E. J. Power Sources 2007, 164, 472−480.<br />

(26) Lee, Y. H.; Lee, G.; Shim, J. H.; Hwang, S.; Kwak, J.; Lee, K.;<br />

Song, H.; Park, J. T. Chem. Mater. 2006, 18, 4209−4211.<br />

(27) Chen, W.; Kim, J.; Sun, S.; Chen, S. Phys. Chem. Chem. Phys.<br />

2006, 8, 2779−2786.<br />

(28) Xu, Z.; Shen, C.; Hou, Y.; Gao, H.; Sun, S. Chem. Mater. 2009,<br />

21, 1778−1780.<br />

(29) Chang, S.-C.; Weaver, M. J. Surf. Sci. 1990, 238, 142−162.<br />

1362<br />

dx.doi.org/10.1021/cs300219j | ACS Catal. 2012, 2, 1358−1362


Electrocatalysis<br />

DOI 10.1007/s12678-012-0099-9<br />

High Activity and Stability of Pt2Bi Catalyst in Formic<br />

Acid Oxidation<br />

J. D. Lović & M. D. Obradović & D. V. <strong>Tripković</strong> &<br />

K. Dj. Popović & V. M. Jovanović & S. Lj. Gojković &<br />

A. V. <strong>Tripković</strong><br />

# Springer Science+Business Media, LLC 2012<br />

Abstract Formic acid oxidation was studied on a new<br />

prepared Pt2Bi characterized by X-ray diffraction spectroscopy<br />

(phase composition), scanning tunneling microscopy<br />

(STM) (surface morphology), and CO ads stripping voltammetry<br />

(surface composition). Bulk composition of Pt2Bi<br />

revealed two phases—55% PtBi alloy and 45% Pt. Estimated<br />

contribution of pure Pt on the Pt 2Bi surface (43.5%)<br />

determined by COads stripping voltammetry corresponds<br />

closely to bulk composition. Pt2Bi reveals high activity<br />

and stability in formic acid oxidation. High activity originates<br />

from the fact that formic acid oxidation proceeds<br />

completely through dehydrogenation path based on an ensemble<br />

effect. The high stability of Pt 2Bi surface is induced<br />

by the suppression of Bi leaching as it was evidenced by<br />

insignificant changes of surface morphology and surface<br />

roughness shown by STM images before and after electrochemical<br />

treatment in formic acid containing solution. Pt2Bi<br />

is found to be powerful catalyst exhibiting up to two orders<br />

of magnitude larger current densities at 0.0 V and onset<br />

potential shifted for ∼0.2 V to more negative value relative<br />

to Pt under steady-state condition.<br />

Keywords Formic acid . Electrochemical oxidation . Pt2Bi<br />

catalyst . Fuel cell<br />

J. D. Lović : M. D. Obradović : D. V. <strong>Tripković</strong> : K. D. Popović :<br />

V. M. Jovanović : A. V. <strong>Tripković</strong> (*)<br />

ICTM-Institute of Electrochemistry, University of Belgrade,<br />

N<strong>je</strong>goševa 12, P.O.Box 473, 11000 Belgrade, Serbia<br />

e-mail: amalija@tmf.bg.ac.rs<br />

S. L. Gojković<br />

Faculty of Technology and Metallurgy, University of Belgrade,<br />

Karnegi<strong>je</strong>va 4, P.O.Box 3503, 11000 Belgrade, Serbia<br />

Introduction<br />

Polymer electrolyte membrane fuel cell using formic acid as<br />

a fuel, i.e., direct formic acid fuel cell has been attracting a<br />

lot of attention in the last few years due to its advantages<br />

over direct methanol fuel cell such as less positive onset<br />

potential of the reaction and lower crossover trough the<br />

polymer membrane [1, 2].<br />

It has been widely accepted that formic acid on Pt is<br />

oxidized to CO 2 via a dual path mechanism [3] involving<br />

a reactive intermediate in the main path–dehydrogenation<br />

and adsorbed COads in the parallel path–dehydration. Since<br />

potential of CO ads formation is lower than potential of formic<br />

acid dehydrogenation, COads remains on the Pt surface<br />

until it can be oxidized at high positive potentials (∼0.65 V<br />

vs. saturated calomel electrode (SCE)). This susceptibility<br />

of Pt to poisoning species significantly reduces its catalytic<br />

performance at low potentials [4]. To solve this problem, Ptbased<br />

bimetallic catalysts were used. Although bifunctional<br />

and electronic effects influence oxidative removal of COads<br />

from the Pt sites, the approach recently proposed based on<br />

the ensemble effect which increases selectivity of Pt toward<br />

dehydrogenation path seems to be a better solution [5].<br />

Intermetallic PtBi [6–10] as well as PtBi alloy [11, 12]<br />

were proposed as powerful catalysts for formic acid oxidation.<br />

The origin of their catalytic activity was related to the<br />

electronic effect [13, 14], to enhanced tolerance to CO<br />

poisoning compared with Pt [15], and to bifunctional effect<br />

[11]. However, they are not stable in acidic media. It has<br />

been reported that instability of Pt-Bi bimetallic catalysts is<br />

attributed to the leaching/dissolution of Bi from the alloy<br />

matrix [16]. This leaching occurs through the oxidation of<br />

the less-noble metal generating a Pt rich surface. The onset<br />

potential of Bi leaching from PtBi corresponds closely to<br />

that predicted from the Pourbaix diagram for elemental Bi


and proceeds all along anodic potentials. Moreover, the adsorption<br />

of previously dissolved Bi could be playing a role as<br />

well. High activity of PtBi alloy relative to polycrystalline Pt<br />

electrode in formic acid oxidation was also discussed in terms<br />

of electrochemically detected underpotential phenomenon of<br />

readsorbed Bi on Pt [11]. However, leaching of Bi causes<br />

instability of PtBi catalysts which is the main problem for<br />

practical application of any Pt-Bi catalysts.<br />

In this work, activity and stability of new prepared specific<br />

Pt2Bi catalyst consisting of two phases (45% pure Pt<br />

and 55% PtBi alloy) in formic acid oxidation was studied.<br />

Experimental<br />

Pt2Bi Alloy Preparation<br />

Pt2Bi was obtained by the induction melting of the constituents<br />

in quartz container under purified Argon atmosphere.<br />

According to Bi–Pt phase diagram [17], Pt starts to dissolve<br />

in liquid Bi already at 271.4° C. At about 500 ° C, fast<br />

dissolution of Pt in liquid Bi was observed. Temperature<br />

was increased slowly up to 1,000 ° C at which clear alloy<br />

meniscus was formed. The meniscus size and shape varied<br />

upon changes of electromagnetic field, attesting intensive<br />

mixing of liquid alloy at 1,000–1,100 ° C. Finally, the alloy<br />

was cooled and stored in Ar atmosphere.<br />

Electrode Characterization<br />

The Pt2Bi alloy was characterized by X-ray diffraction<br />

(XRD). XRD measurements were carried out with a<br />

Bruker/AXS diffractometer using CuKα source operating<br />

at 30 mA and 45 kV and a graphite monochromator. The<br />

polished sample was placed in a Plexiglas holder. Data were<br />

collected over the range 10–135° 2θ using CuKα radiation<br />

with a 0.04° step size and counting time of 2 s.<br />

The quantitative analysis of phase content and crystallite<br />

size calculations were carried out by multiphase Rietveld<br />

refinement using Topas 2.1 software [18].<br />

Surface morphology of Pt2Bi electrode before and after<br />

formic acid oxidation was examined by scanning tunneling<br />

microscopy (STM) using a NanoScope III D (Veeco, USA)<br />

microscope. The STM images were obtained in the height<br />

mode using a Pt–Ir tip (current set-point, it—1 nA; bias<br />

voltage, Vb—100 mV). Fifteen different parts at each electrode<br />

surface were analyzed.<br />

Electrochemical Measurements<br />

Pt2Bi alloy (A00.42 cm 2 ) and polycrystalline Pt discs<br />

(A00.196 cm 2 ) were used in a rotating disc assembly. The<br />

experiments were carried out at the rotation rate of 2,000 rpm.<br />

Prior to each experiment, the electrodes were mirror-polished<br />

(1–0.05 μm Buehler alumina). The surfaces were rinsed with<br />

Millipore water, sonicated for 2–3 min, and rinsed again with<br />

high purity water.<br />

All experiments were carried out on as-prepared catalysts.<br />

Three-compartment electrochemical glass cells with Pt wire as<br />

the counter electrode and SCE as the reference electrode was<br />

used. All the potentials are expressed on the scale of SCE. The<br />

electrolyte contained 0.1 M H2SO4 as a supporting electrolyte<br />

and 0.125 M HCOOH prepared with high purity water (Millipore,<br />

18 MΩ cm resistivity) and p.a. grade chemicals (Merck).<br />

The electrolyte was deaerated by bubbling of nitrogen. Upon<br />

addition of HCOOH at –0.2 V, potentiodynamic (ν050 mV s –1 ),<br />

quasi steady-state measurements (ν01 mVs –1 ), or chronoamperometric<br />

measurements were carried out. The experiments<br />

were conducted at 295±0.5 K. A VoltaLab PGZ 402 (Radiometer<br />

Analytical, Lyon, France) was employed.<br />

For the CO stripping measurements, pure CO was bubbled<br />

through the electrolyte for 20 min while keeping the electrode<br />

potential at –0.2 V vs. SCE. After purging the electrolyte by<br />

N2 for 30 min to eliminate the dissolved CO, the adsorbed CO<br />

wasoxidizedinananodicscanat50mVs −1 . Two subsequent<br />

voltammograms were recorded to verify the completeness of<br />

the CO oxidation. Real surface area of the Pt electrode was<br />

calculated from the charge of CO stripping corrected for<br />

background currents. Assuming charge of 420 μC cm −2 for<br />

the CO monolayer adsorption, roughness factor of 1.4±0.1<br />

was estimated. The real surface area of Pt 2Bi electrode was<br />

estimated assuming the same roughness factor as for Pt electrode,<br />

which is to be expected since both electrodes were<br />

polished in the same way. The specific activity of Pt and Pt 2Bi<br />

electrodes for formic acid oxidation are normalized using<br />

these values of the surface area.<br />

The CO stripping charge was also determined for Pt 2Bi<br />

electrode and corrected for the background currents to eliminate<br />

the contribution of the double layer charge, as well as<br />

Bi oxidation charge. Since CO does not adsorb at PtBi (1:1)<br />

alloy [9, 11, 14, 19], the charge under the CO stripping peak<br />

on Pt2Bi surface corresponds to the surface area of Pt phase<br />

and may serve for determination of the surface composition.<br />

Results and Discussion<br />

Electrode Characterization<br />

Phase Composition<br />

Electrocatalysis<br />

XRD characterization of the Pt 2Bi alloy was performed to<br />

determine its phase composition (Fig. 1).<br />

Diffraction pattern for Pt2Bi sample reveals two crystal<br />

phases: platinum (fcc, cubic system, space group Fm-3 m,<br />

no. 225) and platinum bismuth PtBi (hcp, hexagonal system,


Electrocatalysis<br />

Fig. 1 XRD patterns of Pt 2Bi, PtBi, and polycrystalline Pt catalysts.<br />

Markers represent position of peaks for pure Pt and PtBi alloy<br />

space group P63/mmc, no. 194, ICSD 58845). No other<br />

crystalline phases were found.<br />

The phase composition of the sample was calculated using<br />

the Rietveld refinement as 45% and 55% for Pt and PtBi<br />

phases, respectively. The slight difference between intended<br />

(nominal) and calculated composition can be ascribed to the<br />

sample texture which influences the Rietveld analysis and is<br />

within usual errors for this method. The lattice constants of Pt<br />

fcc phase were refined as a03.9279(2) and for PtBi hcp phase<br />

as a04.3375(7) and c05.5107(3).<br />

As revealed by the XRD analysis, phase composition of the<br />

Pt2Bi sample is consistent with the Bi–Pt phase diagram<br />

(Pt2Bi → Pt+PtBi) [17]. Accordingly, the diffraction patterns<br />

for polycrystalline Pt and PtBi alloy are presented in Fig. 1 to<br />

support qualitatively two-phase composition of Pt2Bi catalyst.<br />

CO ads Stripping Voltammetry<br />

COads stripping voltammetry recorded at Pt2Bi and Pt are<br />

displayed in Fig. 2. The onset potential of CO ads oxidation<br />

as well as the peak position at Pt2Bi is slightly shifted to more<br />

Fig. 2 CO stripping voltammograms on Pt 2Bi and Pt catalysts (first<br />

positive-going sweeps) in 0.1 M H2SO4 solution corrected for background<br />

current. Inset:basevoltammogramofPt 2Bi with CO ads stripping<br />

voltammogram and the voltammogram recorded upon holding the potential<br />

at −0.2 V for 20 min in CO-free solution. Scan rate, 50 mV s –1<br />

negative potentials relative to Pt, indicating the presence of<br />

some electronic modification of Pt surface atoms, capable for<br />

CO adsorption, by Bi. This is in accordance with a small shift<br />

of oxide reduction peak to the same direction at the base<br />

voltammogram for Pt2Bi regarding Pt (not shown). Since<br />

CO does not adsorb at Bi [20] and PtBi [14, 19] alloys,<br />

oxidation of CO occurs only on Pt domains. The sharpness<br />

and the symmetry of the COads stripping peak generally<br />

reflects the uniformity of Pt surface [21]. The shoulder on<br />

CO stripping voltammogram recorded at Pt2Bi could indicate<br />

the existence of two states of Pt atoms in the Pt phase on this<br />

electrode surface. Sites on the edge of Pt domains being in<br />

contact with PtBi alloy phase binds COads stronger than other<br />

Pt sites. Stronger adsorption of CO and OH species on Pt<br />

atoms in contact with PtBi sites arises from the some electronic<br />

modification of Pt surface atoms by Bi [22, 23].<br />

Since CO adsorbs only on Pt domains of Pt2Bi electrode,<br />

the charge under the COads stripping peak can be used for<br />

determining the surface area of pure Pt. Comparing this surface<br />

area to the estimated surface area of Pt2Bi, contribution of<br />

43.5% of Pt domains on the surface is obtained, which is close<br />

to 45% Pt calculated from XRD measurement.<br />

The fact that the methods for determination of phase and<br />

surface composition (XRD and COads stripping voltammetry,<br />

respectively), resulting approximately the same contribution<br />

of Pt, demonstrates clearly that adsorbed CO prevents leaching<br />

of Bi which should lead to Pt enriched surface. This<br />

statement can be recognized by contrasting the base voltammogram<br />

of Pt 2Bi and the voltammogram recorded upon holding<br />

the potential at –0.2 V for 20 min in CO-free solution with<br />

COads stripping voltammogram (inset in Fig. 2). The increase<br />

of the currents at potentials E>0.1 Vobserved in forward scan<br />

of the voltammogram recorded after the potential was kept in


hydrogen region without CO in solution compared with base<br />

voltammogram is a consequence of additional oxidation of<br />

leached Bi. On the contrary, COads stripping voltammogram<br />

recorded in the potential region before COads oxidation commences<br />

coincides with the base voltammogram demonstrating<br />

that adsorbed CO prevents leaching of Bi. Effect of surface<br />

stabilization with COads enables the accurate estimation of the<br />

real surface area by calculation of CO ads stripping charge<br />

which is the only way to do it since the hydrogen region at<br />

Pt2Bi base voltammogram is featureless.<br />

Oxidation of Formic Acid<br />

Potentiodynamic Measurements<br />

Cyclic voltammograms for formic acid oxidation on Pt2Bi and<br />

Pt are presented in Fig. 3. Potentiodynamic profile of Pt shows<br />

well-established feature of HCOOH oxidation on Pt [24]. In the<br />

forward scan, small currents reach a plateau at ∼0.3 V followed<br />

by ascending currents starting at ∼0.45 V and the maximum<br />

at ∼0.6 V. Oxidation of formic acid on Pt proceeds through two<br />

parallel pathways, i.e., dehydrogenation and dehydration, both<br />

generating CO2 as the final reaction product. The dehydrogenation,<br />

assigned as the direct path, is based on the oxidation of<br />

formate [25], the active intermediate containing two oxygen<br />

atoms, which does not need adsorbed water or other oxygencontaining<br />

species [5, 26] to be oxidized to CO2. On the other<br />

hand, dehydration assumes forming COads, well-known as<br />

poisoning species at low potentials, which can be oxidized to<br />

CO2 only by oxygen-containing species formed at higher<br />

potentials. The currents below 0.45 V at the voltammogram<br />

for formic acid oxidation on Pt are related to dehydrogenation<br />

path on the surface partially covered by COads produced by<br />

dehydration of formic acid molecules. At higher potentials,<br />

Fig. 3 Cyclic voltammograms for the oxidation of 0.125 M HCOOH<br />

in 0.1 M H 2SO 4 solution on Pt 2Bi and Pt catalysts. Scan rate,<br />

50 mV s –1<br />

formic acid is oxidized on Pt sites being released upon COads<br />

oxidation with oxygen-containing species. The current peak<br />

at ∼0.6 V is a result of inactive Pt-oxide formation which finally<br />

stops the reaction at ∼0.8 V. In the backward scan, currents<br />

begin to rise at ∼0.7 V, simultaneously with the onset of Pt<br />

oxide reduction, and then decreases at potential more negative<br />

than ∼0.45 V since COads cannot be further oxidized due to a<br />

lack of oxygen-containing species.<br />

On Pt2Bi, catalyst formic acid oxidation commences<br />

at ∼–0.2 V, i.e., more than 0.2 V earlier than on Pt. The<br />

currents increase reaching a peak at ∼0.35 V which is ∼30<br />

times higher than a plateau on Pt and then diminishes up<br />

to positive potential limit. The onset of oxide reduction<br />

at ∼0.6 V in the backward scan (voltammogram is given in<br />

the inset in Fig. 2) causes an increase of formic acid oxidation<br />

currents attaining the maximum at ∼0.35 V.<br />

As Bi does not adsorb formic acid [11, 27], oxidation of<br />

formic acid occurs on pure Pt domains and on Pt atoms on<br />

PtBi domains. Bell-shaped voltammogram for formic acid<br />

oxidation suggests that the reaction on Pt2Bi proceeds through<br />

dehydrogenation path with the dehydration path completely<br />

suppressed. This statement is supported by a remarkable increase<br />

of current bellow 0.35 V, where the activity of Pt<br />

electrode is poor because of high surface coverage with COads,<br />

as well as by the absence of the peak at ∼0.6 V appearing on<br />

Pt. Decrease of the activity at potential higher than 0.35 V is<br />

induced by blocking the Pt sites, needed for decomposition of<br />

adsorbed formate, with oxygen-containing species [5, 26].<br />

Generally, high activity of Pt2Bi catalyst as bimetallic<br />

catalyst could be explained by bifunctional effect, ensemble<br />

effect, and electronic effect. However, bearing in mind all<br />

the data mentioned above, it is reasonable to assume that<br />

increased selectivity toward dehydrogenation path on Pt2Bi<br />

compared with Pt is caused by ensemble effect originating<br />

from the interruption of continuous Pt sites by Bi atoms. In<br />

favor of this statement is the fact that the voltammetric<br />

profile of the formic acid oxidation on Pt 2Bi (Fig. 3) does<br />

not indicate any adsorption or oxidation of CO, although<br />

Pt2Bi is capable of adsorbing CO (Fig. 2).<br />

The electronic modification of Pt by Bi, based on charge<br />

redistribution in a PtBi phase of Pt2Bi catalyst, enhances the<br />

affinity towards formic acid adsorption thus increasing interaction<br />

of formic acid molecules with the catalyst surface<br />

[14], which results in lowering the onset potential of the<br />

reaction compared with Pt.<br />

As formic acid oxidation on Pt 2Bi proceeds through<br />

dehydrogenation path, bifunctional mechanism is not relevant<br />

for interpretation of high catalyst activity.<br />

Chronoamperometric Measurements<br />

Electrocatalysis<br />

Current density at the constant potential of 0.2 V was recorded<br />

over 30 min on Pt2Bi and Pt catalysts and presented in Fig. 4.


Electrocatalysis<br />

Fig. 4 Chronoamperometric curves for the oxidation of 0.125 M<br />

HCOOH at 0.2 V in 0.1 M H2SO4 solution on Pt2Bi and Pt catalysts<br />

The higher initial current density on Pt2Bi than on Pt is in<br />

accordance with potentiodynamic measurements (Fig. 3). The<br />

initial current on Pt2Bi decreases slightly, and at the end of<br />

experiment, its value is about 20 times higher than the current<br />

on Pt. This confirms high activity and stability of Pt2Bi<br />

catalyst also under the steady-state conditions.<br />

Quasi-steady-State Measurements<br />

Kinetics of HCOOH oxidation on Pt2Bi and Pt under the quasisteady-state<br />

conditions is presented in Fig. 5 in a form of the<br />

Tafel plots. The order of activity for the catalysts is the same as<br />

under the potentiodynamic condition. Pt2Bi exhibits up to two<br />

orders of magnitude larger current densities at 0.0 V than Pt.<br />

Fig. 5 Tafel plots for oxidation of 0.125 M HCOOH in 0.1 M H 2SO 4<br />

solution on Pt 2Bi and Pt catalysts. Scan rate, 1 mV s –1<br />

Fig. 6 Cyclic voltammograms for the HCOOH oxidation (first and<br />

20th sweep) on Pt 2Bi catalyst in 0.1 M H 2SO 4 solution. Inset:<br />

corresponding basic voltammograms. Scan rate, 50 mV s –1<br />

Fig. 7 Cyclic voltammograms of Pt recorded before (dashed line) and<br />

after (solid line) replacement of Pt 2Bi in 0.1 M H 2SO 4 solution (a) and<br />

in 0.125 M HCOOH solution (b). Scan rate, 50 mV s –1


Fig. 8 STM images of Pt 2Bi catalyst (300×300×50 nm) before and after formic acid oxidation<br />

The Tafel slope of 120 mV dec –1 indicates that HCOOH<br />

oxidation proceeds on CO ads free surfaces through dehydrogenation<br />

path. Larger Tafel slope of ∼150 mV dec −1 obtained on<br />

Pt suggests that HCOOH oxidation takes place through dehydrogenation<br />

path as on Pt 2Bi but on the surface partially<br />

coveredbyCOads produced in dehydration path occurring in<br />

parallel. The same value of Tafel slope was obtained for<br />

HCOOH oxidation on Pt/C [28]. On both catalysts, first electron<br />

transfer is the rate-determining step in the reaction.<br />

Stability of Pt2Bi<br />

Potentiodynamic Measurements Cyclic voltammograms<br />

(first and 20th sweeps) recorded on Pt2Bi catalyst in a formicacid-containing<br />

solution are shown in Fig. 6. Activity of Pt2Bi<br />

electrode continuously decreases with cycling up to 0.8 V<br />

during initial five to seven sweeps but remains unchanged with<br />

further cycling. On the contrary, cycling of Pt2Bi in base electrolyte<br />

leads to enhancement of currents related to oxidation of<br />

Bi species, indicating some surface decomposition caused by Bi<br />

leaching/dissolution process (inset in Fig. 6). It appears, consequently,<br />

that stability of Pt 2Bi during oxidation of formic acid<br />

could be induced by the presence of formic acid in the electrolyte.<br />

To test this assumption, Pt2Bi electrode was sub<strong>je</strong>cted to<br />

potential cycling in base electrolyte, and after 20 cycles, the<br />

electrode was replaced by Pt electrode (experiment 1).<br />

The same procedure was repeated in the electrolytecontaining<br />

formic acid (experiment 2). The voltammetric profiles<br />

of Pt electrode after the experiments 1 and 2, along with<br />

Electrocatalysis<br />

the voltammograms of Pt electrode in fresh base and formicacid-containing<br />

electrolytes are presented in Fig. 7. Results of<br />

the experiment 1 (Fig. 7a) show slightly reduced charge of<br />

hydrogen adsorption/desorption, indicating underpotential deposition<br />

of previously leached/dissolved Bi. The voltammogram<br />

of formic acid oxidation recorded on Pt in experiment 2<br />

almost retraces the characteristic profile of pure Pt, suggesting<br />

that leaching of Bi is suppressed in the presence of formic acid<br />

(Fig. 7b). It should be noted that the experiment performed<br />

after replacing of PtBi electrode with Pt [11] revealedsignificant<br />

Bi leaching under the same experimental conditions,<br />

meaning that Pt2Bi is more resistant to Bi leaching than PtBi.<br />

STM Measurements In order to test whether the surface morphology<br />

of Pt2Bi changes during formic acid oxidation, STM<br />

measurements were carried out before and after the reaction.<br />

Typical STM images of the electrode before and after the<br />

experiment are presented in Fig. 8. The image of polished<br />

surface pointes out inhomogeneity of the material structure<br />

illustrated by larger and smaller domains separated by both<br />

wide and narrow cavities, as a consequence of catalyst<br />

preparation. A similar image is recorded for the electrode<br />

after the oxidation of formic acid. Moreover, roughness<br />

analysis of 15 different parts at the electrode surface, both<br />

before and after the experiment, reveals comparable surface<br />

roughness values (Rms). The mean Rms value for the polished<br />

electrode before formic acid oxidation is 2.72 with<br />

standard deviation of ±0.28, while the mean Rms for the<br />

electrode after 20 cycles in HCOOH-containing sulfuric


Electrocatalysis<br />

acid is 3.13 with standard deviation of ±0.46. Thus, based<br />

on this small difference in Rms, it could be reasonable to<br />

assume that no significant leaching of Bi occurred during<br />

the formic acid oxidation.<br />

These STM analyses, which refer to insignificant change<br />

of surface morphology and roughness, are in accordance<br />

with previously determined suppression of Bi leaching during<br />

formic acid oxidation. It appears, consequently, that<br />

Pt2Bi surface becomes kinetically stabilized due to the competition<br />

between the oxidation of formic acid at the electrode/solution<br />

interface and Bi leaching, i.e., corrosion/<br />

oxidation process of the electrode surface itself [29]. Accordingly,<br />

the main reasons for high stability of Pt2Bi catalyst<br />

is the suppression of Bi leaching, as well as inhibition of<br />

dehydration path in the reaction of formic acid oxidation.<br />

Conclusion<br />

Bulk composition of Pt 2Bi characterized by XRD revealed two<br />

phases—55% PtBi alloy and 45% Pt. Estimated contribution<br />

of pure Pt on the Pt2Bi surface (43.5%) determined by COads<br />

stripping voltammetry corresponding closely to bulk composition<br />

indicates that adsorbed CO prevents leaching of Bi.<br />

Pt2Bi exhibits high activity and stability in formic acid<br />

oxidation. High activity is caused by the fact that formic<br />

acid oxidation proceeds completely through dehydrogenation<br />

path. Increased selectivity toward dehydrogenation is<br />

caused by an ensemble effect.<br />

Stability of Pt2Bi surface in formic acid oxidation is<br />

induced by preventing Bi leaching/dissolution during formic<br />

acid oxidation, as well as by the absence of surface poisoning<br />

by COads species due to inhibition of dehydration path.<br />

Pt2Bi is powerful catalyst for formic acid oxidation exhibiting<br />

up to two orders of magnitude larger current densities<br />

at 0.0 V and onset potential negatively shifted for ∼0.2 V<br />

compared with Pt.<br />

Acknowledgments This work was financially supported by the Ministry<br />

of Education and Science, Republic of Serbia, contract no. H-172060.<br />

References<br />

1. X. Wang, J.-M. Hu, I.-M. Hsing, J Electroanal Chem 562, 73<br />

(2004)<br />

2. J. Willsau, J. Heitbaum, Electrochim Acta 31, 943 (1986)<br />

3. A. Capon, R. Parsons, J. Electroanal. Chem. 44, 1 (1973)<br />

4. P.K. Babu, H.S. Kim, J.H. Chung, E. Oldfield, A. Wieckowski, J<br />

Phys Chem B 108, 20228 (2004)<br />

5. A. Cuesta, M. Escudero, B. Lanova, H. Baltruschat, Langmuir 25,<br />

6500 (2009)<br />

6. D. Volpe, E. Casado-Rivera, L. Alden, C. Lind, K. Hagerdon, C.<br />

Downie, C. Korzniewski, F.J. DiSalvo, H.D. Abruna, J Electrochem<br />

Soc 151, A971 (2004)<br />

7. E. Casado-Rivera, D.J. Volpe, L. Alden, C. Lind, C. Downie, T.<br />

Vazquez-Alvarez, A.C.D. Angelo, F.J. DiSalvo, H.D. Abruna, J<br />

Am Chem Soc 126, 4043 (2004)<br />

8. C. Roychowdhury, F. Matsumoto, V.B. Zeldovich, S.C. Warren, P.F.<br />

Mutolo, M.J. Ballesteros, U. Wiesner, H.D. Abruna, F.J. DiSalvo,<br />

Chem Mater 18, 3365 (2006)<br />

9. H. Wang, L. Alden, F.J. DiSalvo, H.D. Abruna, Phys Chem Chem<br />

Phys 10, 3739 (2008)<br />

10. J. Sanabria-Chinchilla, H. Abe, F.J. DiSalvo, H.D. Abruna, Surf<br />

Sci 602, 1830 (2008)<br />

11. A.V. <strong>Tripković</strong>, K.D. Popović, R.M. Stevanović, R. Socha, A.<br />

Kowal, Electrochem Comm 8, 1492 (2006)<br />

12. E. Herrero, A. Fernandez-Vega, J.M. Feliu, A. Aldaz, J Electroanal<br />

Chem 350, 73 (1993)<br />

13. L.R. Alden, D.K. Han, F. Matsumoto, H.D. Abruna, F.J. DiSalvo,<br />

Chem Mater 18, 5591 (2006)<br />

14. E. Casado-Rivera, Z. Gal, A.C.D. Angelo, C. Lind, F.J. DiSalvo,<br />

H.D. Abruna, Chem Phys Chem 4, 193 (2003)<br />

15. N. de-los-Santos-Alvarez, L.R. Alden, E. Rus, H. Wang, F.J. DiSalvo,<br />

H.D. Abruna, J Electroanal Chem 626, 14(2009)<br />

16. D.R. Blasini, D. Rochefort, E. Fachini, L.R. Alden, F.J. DiSalvo,<br />

C.R. Cabrera, H.D. Abruna, Surf Sci 600, 2670 (2006)<br />

17. H. Okamoto (ed) (1990) “Bi–Pt phase diagram” APD program. In:<br />

Binary alloy phase diagrams. ASM International, Materials Park,<br />

OH. p. 778<br />

18. A.X.S. Bruker, TOPAS V2.1: general profile and structure analysis<br />

software for powder diffraction data, user manual. Bruker AXS,<br />

Karlsruhe, Germany (2003)<br />

19. M. Oana, R. Hoffmann, H.D. Abruna, F.J. DiSalvo, Surf Sci 574,1<br />

(2005)<br />

20. T.J. Schmidt, B.N. Grgur, R.J. Behm, N.M. Marković, P.N. Ross<br />

Jr., Phys Chem Chem Phys 2, 4379 (2000)<br />

21. N.P. Lebedeva, M.T.M. Koper, E. Herrero, J.M. Feliu, R.A. van<br />

Santen, J Electroanal Chem 487, 37 (2000)<br />

22. J.R. Kitchin, J.K. Nørskov, M.A. Barteau, J.G. Chen, Phys Rev<br />

Lett 93, 156801 (2004)<br />

23. J.R. Kitchin, J.K. Nørskov, M.A. Barteau, J.G. Chen, J Chem Phys<br />

120, 10240 (2004)<br />

24. D. Jarvi, E.M. Stuve (1998) In: J. Lipkowski and P.N. Ross (eds)<br />

Electrocatalysis. New York, Wiley-VCH<br />

25. A. Miki, S. Ye, M. Osawa, Chem Commun 1500 (2002)<br />

26. M. Arenz, V. Stamenković, T.J. Schmidt, K. Wandelt, P.N. Ross,<br />

N.M. Marković, Phys Chem Chem Phys 5, 4242 (2003)<br />

27. N. Kapur, B. Shan, J. Hyun, L. Wang, S. Yang, J.B. Nicholas, K.<br />

Cho, Mol Simul 37, 648 (2011)<br />

28. J.D. Lović, A.V. <strong>Tripković</strong>, S.L. Gojković, K.D. Popović, D.V.<br />

<strong>Tripković</strong>, P. Olszewski, A. Kowal, J Electroanal Chem 581, 294<br />

(2005)<br />

29. Y. Liu, M.A. Lowe, F.J. DiSalvo, H.D. Abruna, J Phys Chem C<br />

114, 14929 (2010)


Received: May 24, 2011<br />

Published: July 19, 2011<br />

ARTICLE<br />

pubs.acs.org/JACS<br />

Design and Synthesis of Bimetallic Electrocatalyst with Multilayered<br />

Pt-Skin Surfaces<br />

Chao Wang, † Miaofang Chi, ‡ Dongguo Li, †,§ Dusan Strmcnik, † Dennis van der Vliet, † Guofeng Wang, ||<br />

Vladimir Komanicky, †,z Kee-Chul Chang, † Arvydas P. Paulikas, † Dusan Tripkovic, † John Pearson, †<br />

KarrenL.More, ‡ Nenad M. Markovic, † and Vojislav R. Stamenkovic* ,†<br />

†<br />

Materials Science Division, Argonne National Laboratory, Argonne, Illinois 60439, United States<br />

‡<br />

Division of Material Science and Technology, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United States<br />

§<br />

Department of Chemistry, Brown University, Providence, RI 02912, United States<br />

Department of Mechanical Engineering and Materials Science, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States<br />

z<br />

Safarik University, Faculty of Science, Kosice 04154, Slovakia<br />

)<br />

bS Supporting Information<br />

ABSTRACT: Advancement in heterogeneous catalysis relies<br />

on the capability of altering material structures at the nanoscale,<br />

and that is particularly important for the development of highly<br />

active electrocatalysts with uncompromised durability. Here, we<br />

report the design and synthesis of a Pt-bimetallic catalyst with<br />

multilayered Pt-skin surface, which shows superior electrocatalytic<br />

performance for the oxygen reduction reaction (ORR).<br />

This novel structure was first established on thin film extended<br />

surfaces with tailored composition profiles and then implemented<br />

in nanocatalysts by organic solution synthesis. Electrochemical<br />

studies for the ORR demonstrated that after prolonged<br />

exposure to reaction conditions, the Pt-bimetallic catalyst with<br />

multilayered Pt-skin surface exhibited an improvement factor of more than 1 order of magnitude in activity versus conventional Pt<br />

catalysts. The substantially enhanced catalytic activity and durability indicate great potential for improving the material properties by<br />

fine-tuning of the nanoscale architecture.<br />

’ INTRODUCTION<br />

The foreground of sustainable energy is built up on a renewable<br />

and environmentally compatible scheme of chemical-electrical<br />

energy conversion. One of the key processes for such energy<br />

conversion is the electrocatalytic reduction of oxygen, the cathode<br />

reaction in fuel cells 1 and metal air batteries, 2 where an electrocatalyst<br />

is used to accelerate the course of ORR. Current electrocatalysts<br />

used for this reaction are typically in the form of dispersed Pt<br />

nanoparticles (NPs) on amorphous high-surface-area carbon. Considering<br />

the high cost and limited resource of Pt, large-scale applications<br />

of these renewable energy technologies demand substantial<br />

improvement of the catalyst performance so that the amount of Pt<br />

needed can be significantly reduced. For example, a 5-fold improvement<br />

of catalytic activity for the ORR is required for the commercial<br />

implementation of fuel cell technology in transportation. 3<br />

Our recent work on well-defined extended surfaces has shown<br />

that high catalytic activity for the ORR can be achieved on<br />

Pt bimetallic alloys (Pt 3M, M = Fe, Co, Ni, etc.), due to the altered<br />

electronic structure of the Pt topmost layer and hence reduced<br />

adsorption of oxygenated spectator species (e.g., OH )onthe<br />

surface. 4 It was also found that in an acidic electrochemical environment<br />

the non-noble 3d transition metals are dissolved from the near-<br />

surface layers, which leads to the formation of Pt-skeleton surfaces.<br />

Moreover, the thermal treatment of Pt3M alloys in ultra high vacuum<br />

(UHV) has been shown to induce segregation of Pt and formation of<br />

a distinctive topmost layer that was termed Pt-skin surface. However,<br />

the same treatment did not cause Pt to segregate over PtM alloys<br />

with high content (g50%) of non-Pt elements. 4b,5 Recently, we<br />

further demonstrated the surfacing of an ordered Pt(111)-skin over<br />

Pt3Ni(111) single crystal with 50% of Ni in the subsurface layer. This<br />

unique nanosegregated composition profile was found to be responsibleforthedramatically<br />

enhanced ORR activity. 6<br />

On the basis of these findings, it could be envisioned that the<br />

most advantageous nanoscale architecture for a bimetallic electrocatalyst<br />

would correspond to the segregated Pt-skin composition<br />

profile established on extended surfaces. A lot of effort has<br />

thus been dedicated, 7 but it still remains elusive, to finely tune the<br />

Pt-bimetallic nanostructure to achieve this desirable surface<br />

structure and composition profile. Major obstacles reside not<br />

only in the difficulty to manipulate elemental distribution at the<br />

nanoscale, but also in the fundamental differences in atomic<br />

r 2011 American Chemical Society 14396 dx.doi.org/10.1021/ja2047655 | J. Am. Chem. Soc. 2011, 133, 14396–14403


Journal of the American Chemical Society ARTICLE<br />

structures, electronic properties, and catalytic performance between<br />

extended surfaces and confined nanomaterials. For example, in<br />

an attempt to induce surface segregation, high-temperature (>600 °C)<br />

annealing is typically applied for Pt-based alloy nanocatalysts.<br />

While improvement in specific activity is obtained, such treatment<br />

usually causes particle sintering and loss of electrochemical<br />

surface area (ECSA). 7b,8 Besides that, the surface coordination<br />

of nanomaterials is quite different from that of bulk materials;<br />

that is, the surface of NPs is rich in corner and edge sites, which<br />

have a smaller coordination number than the atoms on longrange<br />

ordered terraces of extended surfaces. 9 These low-coordination<br />

surface atoms are considered as preferential sites for the<br />

adsorption of oxygenated spectator species (e.g., OH ) 10 and<br />

thus become blocked for adsorption of molecular oxygen and<br />

inactive for the ORR. 9b,11 Additionally, due to strong Pt O<br />

interaction, these sites are more vulnerable for migration and<br />

dissolution, resulting in poor durability and fast decay of the<br />

catalyst. 12 The latter effect is even more pronounced in Pt bimetallic<br />

systems, considering that more undercoordinated atoms<br />

are present on the skeleton surfaces formed after the depletion of<br />

nonprecious metals from near-surface regions. 4a,13 Therefore, a<br />

systematic approach with all of these factors integrally considered<br />

becomes necessary to pursue the design and synthesis of<br />

advanced bimetallic catalysts.<br />

Our focus in this study has been placed on the fine-tuning of<br />

Pt bimetallic nanostructure aiming to achieve the advantageous<br />

Pt-skin surface structure and composition profile established on<br />

extended surfaces. We started with Pt thin films of controlled<br />

thickness deposited over PtNi substrate to explore the correlation<br />

between the surface composition profile and catalytic performance.<br />

These findings were then applied for guiding the synthesis of<br />

nanocatalysts with the optimized structure. The outcome of such<br />

effort is an advanced Pt bimetallic catalyst with altered nanoscale<br />

architecture that is highly active and durable for the ORR.<br />

’ EXPERIMENTAL SECTION<br />

Thin Film Preparation. Pt films were deposited at room temperature<br />

on PtNi substrates (6 mm in diameter), which were set 125 mm<br />

away from DC sputter magnetrons in 4 mTorr argon gas (base vacuum<br />

1 10 7 Torr). The Pt source rate (0.32 Å/s) was determined by quartz<br />

crystal microbalance, and an exposure duration of 7.0 s was calibrated for<br />

the nominal thickness of 2.2 2.3 Å for a monolayer of Pt. The film<br />

thickness was derived from the exposure time of computer-controlled<br />

shutters during sputtering.<br />

NP and Catalyst Synthesis. In a typical synthesis of PtNi NPs,<br />

0.67 mmol of Ni(ac)2 3 4H2O was dissolved in 20 mL of diphenyl either<br />

in the presence of 0.4 mL of oleylamine and 0.4 mL of oleic acid.<br />

0.33 mmol of 1,2-tetradodecanediol was added, and the formed solution<br />

washeatedto80°C under Ar flow. After a transparent solution formed, the<br />

temperature was further raised to 200 °C, where 0.33 mmol of Pt(acac)2<br />

dissolved in 1.5 mL of dichlorobenzene was in<strong>je</strong>cted. The solution was<br />

kept at this temperature for 1 h and then cooled to room temperature.<br />

An amount of 60 mL of ethanol was added to precipitate the NPs, and<br />

the product was collected by centrifuge (6000 rpm, 6 min). The<br />

obtained NPs were further washed by ethanol two times and then<br />

dispersed in hexane. The as-synthesized PtNi NPs were deposited on<br />

high surface area carbon (∼900 m 2 /g) by mixing the NPs with carbon<br />

black (Tanaka, KK) in hexane or chloroform with a 1:1 ratio in mass.<br />

This mixture was sonicated for 1 h and then dried under nitrogen flow.<br />

The organic surfactants were removed by thermal treatment at<br />

150 200 °C in an oxygenated atmosphere. The obtained catalyst is<br />

denoted as “as-prepared PtNi/C”. For the acid treatment, ∼10 mg of the<br />

as-prepared PtNi/C catalyst was mixed with 20 mL of 0.1 M HClO4 that<br />

has been used as electrolyte in electrochemical measurements. After<br />

overnight exposure to the acidic environment, the product was collected<br />

by centrifuge and washed three times by deionized water. Such NPs are<br />

named as “acid treated PtNi/C”. The acid treated PtNi/C was further<br />

annealed at 400 °C to reduce low-coordinated surface sites, and the<br />

obtained catalyst is termed as “acid treated/annealed PtNi/C”.<br />

Microscopic Characterization. TEM images were collected on a<br />

Philips EM 30 (200 kV) equipped with EDX functionality. XRD patterns<br />

were collected on a Rigaku RTP 300 RC machine. STEM and elemental<br />

analysis were carried out on a JEOL 2200FS TEM/STEM with a CEOS<br />

aberration (probe) corrector. The microscope was operated at 200 kV in<br />

HAADF-STEM mode equipped with a Bruker-AXS X-Flash 5030 silicon<br />

drift detector. The probe size was ∼0.7 Å and the probe current was<br />

∼30 pA during HAADF-STEM imaging. When accumulating EDX data,<br />

the probe current was increased to ∼280 pA and the probe size was<br />

∼2 Å. The presented EDX data were confirmed to be from “e-beam<br />

damage-free” particles by comparing STEM images before and after<br />

EDX acquisition.<br />

X-ray Absorption Spectroscopy. X-ray fluorescence spectra of<br />

at the Ni K and Pt L 3 edges were acquired at bending magnet beamline<br />

12-BM-B at the Advanced Photon Source (APS), Argonne National<br />

Laboratory. The incident radiation was filtered by a Si(111) doublecrystal<br />

monochromator (energy resolution ΔE/E = 14.1 10 5 ) with a<br />

double mirror system for focusing and harmonic re<strong>je</strong>ction. 14 All of the<br />

data were taken in fluorescence mode using a 13-element Germanium<br />

array detector (Canberra), which was aligned with the polarization of the<br />

X-ray beam to minimize the elastic scattering intensity. Co and Ge filters<br />

(of 6 absorption length in thickness) were applied in front of the<br />

detector to further reduce the elastic scattering intensity for the Ni K and<br />

Pt L 3 edges, respectively. The Ni K and Pt L 3 edge spectra were<br />

calibrated by defining the zero crossing point of the second derivative<br />

of the XANES spectra for Ni and Pt reference foils as 8333 and 11564 eV,<br />

respectively. The background was subtracted using the AUTOBK<br />

algorithm, 15 and data reduction was performed using Athena from the<br />

IFEFFIT software suite. 16 A scheme of the homemade in situ electrochemical<br />

cell and setup at beamline was shown in the Supporting<br />

Information, Figure S9.<br />

Electrochemical Characterization. The electrochemical measurements<br />

were conducted in a three-compartment electrochemical cell<br />

with a rotational disk electrode (RDE, 6 mm in diameter) setup (Pine)<br />

and a Autolab 302 potentiostat. A saturated Ag/AgCl electrode and a Pt<br />

wire were used as reference and counter electrodes, respectively. 0.1 M<br />

HClO 4 was used as electrolyte. The catalysts were deposited on glassy<br />

carbon electrode substrate and dried in Ar atmosphere without using any<br />

ionomer. The loading was controlled to be 12 μgPt/cm 2 disk for PtNi/C<br />

nanocatalysts. All of the potentials given in the discussion were against<br />

reversible hydrogen electrode (RHE), and the readout currents were<br />

recorded with ohmic iR drop correction during the measurements. 17<br />

’ RESULTS AND DISCUSSION<br />

Pt films of various thicknesses, that is, 1 7 atomic monolayers<br />

(ML), were deposited in a vacuum by sputtering on PtNi<br />

(Pt:Ni = 1:1) substrate and then transferred to an electrochemical<br />

cell for further characterizations (see the Experimental<br />

Section). The as-sputtered Pt films consist of randomly distributed<br />

Pt nanoclusters (


Journal of the American Chemical Society ARTICLE<br />

Figure 1. Electrochemical studies on the Pt thin films deposited over PtNi substrate by RDE: (a) cyclic voltammograms, (b) polarization curves,<br />

and (c) summary of specific activities and corresponding improvement factors (vs polycrystalline Pt surface) for the Pt films of various<br />

thicknesses. Cyclic voltammograms were recorded in Ar saturated 0.1 M HClO4 electrolyte with a sweeping rate of 50 mV/s. Polarization curves<br />

were recorded in the same electrolyte under O2 saturation with a sweep rate of 20 mV/s. Specific activities were presented as kinetic currents<br />

normalized by ECSAs obtained from integrated Hupd, except that for the annealed 3 ML Pt/PtNi surface which was based on COad stripping<br />

polarization curve.<br />

which had confirmed the superior catalytic properties of<br />

systems with 50% of Ni in subsurface layers. Figure 1 summarizes<br />

the results of electrochemical studies for these thin films by<br />

rotating disk electrode (RDE). Cyclic voltammograms (CVs,<br />

Figure 1a) of the as-sputtered films correspond to polycrystalline<br />

Pt (poly-Pt) with similar, but slightly enlarged, underpotentially<br />

deposited hydrogen (Hupd) regions (E < 0.4 V) due to the<br />

rougher surfaces. Consistent with our previous findings, 6 the<br />

onset of Pt OH ad formation has anodic shifts for most of the Pt<br />

films (e5 ML) as compared to poly-Pt (more visible in the CVs<br />

shown in the Supporting Information, Figure S1, with currents<br />

normalized by the electrochemical surface area (ECSA) obtained<br />

from integrated Hupd region). Correspondingly, similar positive<br />

shifts are also present in the polarization curves for the ORR<br />

(Figure 1b). The largest shift of ∼30 mV was obtained for the Pt<br />

films with thicknesses of three atomic layers. Measured specific<br />

activities at 0.95 V, expressed as kinetic current normalized by the<br />

ECSA, show that the thinner films (e3 ML) have improvement<br />

factors of ∼2.5 versus poly-Pt surface, which is in line with the<br />

previous results on polycrystalline Pt3M bulk alloys with the<br />

skeleton type of surfaces. 4b Reduced enhancement was observed<br />

for thicker Pt films, for example, improvement factor of 1.7 for<br />

5MLofPt,whilethespecific activity measured for the 7 ML<br />

film was close to that of poly-Pt. It should be noted here that<br />

for the as-sputtered films,1MLofPtmaynotbeableto<br />

entirely cover Ni atoms in the alloy substrate and protect<br />

them from dissolution, whereas addition of a second and/or<br />

third layer can effectively diminish this process. Along the<br />

same lines, this may also be the reason that the surfaces with 2 or<br />

3 ML of Pt were found to be more active than that with 1 ML.<br />

These findings revealed that bimetallic systems with Pt-skeleton<br />

near-surface formation of up to three atomic layers in thickness<br />

are also capable of efficiently harvesting the beneficial properties<br />

of bimetallic alloys, while protecting the subsurface Ni from<br />

leaching out.<br />

Because the as-sputtered skeleton type of surfaces have abundant<br />

low-coordination sites 4a that are detrimental to the ORR, we have<br />

applied thermal treatment to investigate potential surface restructuring<br />

and further catalytic improvement. A moderate temperature<br />

of ∼400 °C was chosen as it was determined to be optimal for<br />

Pt bimetallic nanocatalysts. 8a In Figure 1a, the CV of annealed<br />

3 ML Pt/PtNi surface is also shown. The suppressed Hupd region<br />

and an even larger positive shift of the Pt OHad peak (Figure S1)<br />

indicates the formation of Pt-skin type of surface, which is smoother<br />

and less oxophilic with significantly reduced number of lowcoordination<br />

surface atoms. 5,6 Additional proof of the transition<br />

toward Pt-skin is provided by the measured boost in specific<br />

activity for the ORR (Figure 1c), reaching an improvement factor<br />

of more than 5 with respect to poly-Pt (Figure 1d). Moreover,<br />

this high catalytic activity was based on the ECSA estimated from<br />

CO stripping polarization curves, not Hupd. The ECSA estimated<br />

from integrated Hupd charge was found to be substantially smaller<br />

than that obtained from the electrochemical oxidation of adsorbed<br />

CO monolayer (Figure S2), which was not observed on unannealed<br />

Pt-skeleton surfaces. Such a difference can only be interpreted<br />

in terms of the altered electronic properties of the Pt-skin surface<br />

14398 dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. 2011, 133, 14396–14403


Journal of the American Chemical Society ARTICLE<br />

Figure 2. Representative transmission electron microscopy (TEM)<br />

images for (a,b) the as-synthesized PtNi NPs, (c) the as-prepared, and<br />

(d) the acid treated/annealed PtNi/C catalysts. (d) X-ray diffraction<br />

(XRD) patterns for the PtNi/C catalysts in comparison with commercial<br />

Pt/C (∼6 nm in particle size).<br />

that have affected the adsorption of hydrogenated species, but<br />

not the Pt COad interaction. 6<br />

The above studies of Pt thin films over PtNi substrate, as well<br />

as the knowledge previously acquired from poly-/single-crystalline<br />

surfaces 4 6 and nanocatalysts, 8a,11b,13 have led us to a novel<br />

approach toward the design and synthesis of Pt bimetallic<br />

catalysts with Pt terminated surfaces. The initial step in this<br />

approach should involve the synthesis of monodisperse and<br />

homogeneous PtNi NPs followed by intentional depletion of<br />

Ni from the surface, producing a skeleton type of surface<br />

structure. The final step is supposed to be the thermal treatment<br />

aimed to induce the transition from Pt-skeleton into Pt-skin type<br />

of structure by surface relaxation and restructuring. For that<br />

purpose, PtNi NPs were synthesized by simultaneous reduction of<br />

platinum acetylacetonate, Pt(acac)2, and nickel acetate, Ni(ac)2,<br />

in an organic solution (see the Experimental Section). 13,18<br />

Figure 2a and b shows representative transmission electron<br />

microscopy (TEM) images of the as-synthesized PtNi NPs<br />

prepared with a molar ratio of 1:2 between the Pt and Ni<br />

precursors. The NP size was controlled to be ∼5 nm 11b with a<br />

very narrow size distribution, as evidenced by the formation of<br />

various types of super lattices after drying the nanoparticle<br />

suspension (in hexane) under ambient conditions. 19 The final<br />

composition was characterized by energy-dispersive X-ray spectroscopy<br />

(EDX), which confirmed an atomic ratio of Pt/Ni ≈ 1/1<br />

(Figure S3). The as-synthesized NPs were incorporated into<br />

carbon black (∼900 m 2 /g) via a colloidal-deposition method,<br />

and the organic surfactants were efficiently removed by thermal<br />

treatment. 11b Such as-prepared PtNi/C catalyst was first treated by<br />

acid to dissolve the surface Ni atoms 13 (Figure S3) and then<br />

annealed at 400 °C. These consecutive treatments were expected<br />

to bring on the Pt-skin type of surface over the substrate with 50%<br />

of Ni, which otherwise would not be possible because complete<br />

segregation of Pt only takes place in Pt3M systems. 5 TEM images<br />

of the acid treated/annealed catalyst do not show notable<br />

changes in morphology (Figure 2c and d), except a slight decrease<br />

(∼0.3 nm) in average particle size (Figures S4). Additionally,<br />

X-ray diffraction (XRD) analysis was used to characterize the<br />

crystal structure of the NPs. As compared to the commercial Pt/C<br />

catalyst (Tanaka, ∼6 nm), both the as-prepared and the acid<br />

treated/annealed PtNi/C systems show a face-centered cubic (fcc)<br />

pattern with noticeable shifts (e.g., ∼1° for (111) peak) toward<br />

high angle (Figure 2e), corresponding to a decrease of lattice<br />

constant due to alloying between Pt and Ni. 20 The XRD pattern of<br />

the acid treated/annealed NPs has sharper peaks as compared to<br />

the as-prepared one, which indicates the increased crystallinity<br />

after annealing. These observations, in addition to the absence of<br />

peaks for separate Pt or Ni phases, show that the bimetallic catalyst<br />

preserved the alloy properties after the applied treatments.<br />

The nanostructures and composition profiles of the PtNi/C<br />

catalysts were characterized by atomically resolved aberrationcorrected<br />

high angle annular dark field-scanning transmission<br />

electron microscopy (HAADF-STEM) in combination with energy<br />

dispersive X-ray spectroscopy (EDX). Figure 3a shows representative<br />

HAADF-STEM images taken along the Æ110æ zone axis of the<br />

as-prepared (left), acid treated (middle), and acid treated/annealed<br />

(right) PtNi/C catalysts, with the intensity profiles along Æ001æ<br />

directions across the single particles shown in Figure 3b. As<br />

compared to the benchmark intensity profiles calculated for ideal<br />

octahedral alloy NPs of the same size and orientation (see the<br />

Supporting Information), NP exposed to acid shows 3 4peakson<br />

each side stretching above the standards, indicating the formation of<br />

a Pt-rich overlayer. This feature was preserved after annealing, but<br />

with 2 3 Pt-rich peaks on each side, corresponding to a reduced Pt<br />

overlayer thickness due to restructuring and smoothing (Figure 3b).<br />

These findings were further confirmed by EDX analysis. By<br />

scanning the e-beam across the particle while simultaneously<br />

analyzing the generated X-rays, composition line profiles were<br />

obtained for the NPs (Figure 3c). It can be seen that the distribution<br />

of Pt and Ni in the as-prepared catalyst was highly intermixed and<br />

the sketched trend lines were almost identical, indicating a homogeneous<br />

alloy nature of the catalyst particles. The treated catalysts<br />

have substantially broader distribution of Pt than Ni, with a<br />

difference of ∼1 nm (at the half-maximum of the trend lines) for<br />

the acid treated and ∼0.6 nm for the acid treated/annealed catalyst.<br />

Hence, both the intensity and the composition line profiles show<br />

that multilayered Pt-rich surface structure was formed by acid<br />

treatment and preserved after annealing.<br />

The microscopic characterizations strongly point toward surface<br />

restructuring in the bimetallic catalyst upon annealing. This was<br />

additionally depicted by atomistic simulation of the nanostructure<br />

evolution sub<strong>je</strong>ct to the acid and annealing treatments (Figure 3d<br />

for overviews and Figure 3e for cross-section views; see the<br />

Supporting Information for more details). It shows that removing<br />

Ni atoms from the surface led to the formation of a Pt-skeleton<br />

overlayer with a thickness of up to 3 atomic layers. Further relaxation<br />

of low-coordination surface atoms resulted in a multilayered Pt-skin<br />

surface, whereas the PtNi core was barely affected. It is important to<br />

mention that the relaxation process is expected to induce preferential<br />

formation of highly active (111) surface 6 (labeled by arrows in<br />

14399 dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. 2011, 133, 14396–14403


Journal of the American Chemical Society ARTICLE<br />

Figure 3. Microscopic characterization and theoretical simulation of nanostructure evolution in the PtNi/C catalysts: (a) Representative high-angle<br />

annular dark-field scanning transmission electron microscopy (HAADF-STEM) images taken along the zone axis Æ110æ, as confirmed by the fast Fourier<br />

transfer (FFT) patterns of the STEM images (shown as insets); (b) background subtracted, normalized intensity line profiles extracted for the regions<br />

marked in (a); (c) composition line profiles (normalized for Pt L peaks) obtained by energy-dispersive X-ray spectroscopy (EDX) with an electron<br />

beam (∼2 Å in spot size) scanning across individual catalyst particles; (d) overview; and (e) cross-section views of the nanostructures depicted by<br />

atomistic particle simulation. The figure is also organized in columns for the as-prepared (left), acid treated (middle), and acid treated/annealed (right)<br />

PtNi/C catalysts, respectively.<br />

Figure 3e), due to the higher atomic coordination, that is, lower<br />

surface energy, of this facet as compared to others.<br />

To gain more insights into the nanostructure evolution,<br />

especially the correlation of surface structures to their electrochemical<br />

properties, we have carried out in situ X-ray absorption<br />

near edge structure (XANES) studies of the nanocatalysts (see the<br />

Experimental Section and Supporting Information for details).<br />

Figure 4a and b shows the normalized XANES spectra collected<br />

at Ni K and Pt L3 edges, under the ORR-relevant conditions<br />

(∼1.0 V). As compared to the spectra of reference foils, Pt and Ni<br />

edge positions were found to correspond to the bulk oxidation<br />

state of zero for both elements in the treated catalysts. 21 It is<br />

intriguing to see that the acid treated catalyst shows higher white<br />

line intensity than does the acid treated/annealed catalyst at the<br />

Ni edge, which is caused by the presence of a small amount of<br />

NiO underlying the highly corrugated Pt-skeleton surface<br />

morphology in the acid treated catalyst, 22 whereas subsurface<br />

Ni in the acid treated/annealed catalyst was well protected.<br />

The distinction in surface structure between the two treated<br />

catalysts is even more visible at the Pt edge, where a slightly lower<br />

white line intensity for the acid treated/annealed catalyst<br />

corresponds to a reduced amount of platinum oxides under the<br />

same conditions, and, more fundamentally, less oxophilic surface<br />

with larger average surface coordination number. 5,22a,23 The<br />

findings from XANES provide additional evidence for the<br />

formation of surface relaxed multilayered Pt-skin in the acid<br />

treated/annealed catalyst and its superiority in protecting the<br />

inner Ni from leaching out.<br />

On the basis of these results, we have managed to achieve the<br />

desirable nanoscale architecture established on PtNi supported<br />

Pt films, that is, multilayered Pt-skin over a particle core with 50%<br />

of Ni. Considering what was revealed from the studies on<br />

extended surfaces, the obtained nanocatalyst should show superior<br />

catalytic performance for the ORR, which was examined by RDE<br />

14400 dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. 2011, 133, 14396–14403


Journal of the American Chemical Society ARTICLE<br />

Figure 4. In situ X-ray absorption and electrochemical studies of the PtNi/C catalysts. (a,b) XANES spectra for the PtNi/C catalysts recorded at Ni K<br />

and Pt L3 edges with an electrode potential of 1.0 V, in comparison with standard spectra of Ni, NiO, and Pt. (c) Cyclic voltammograms, (d) polarization<br />

curves, and (e) Tafel plots with the specific activity (jk, kinetic current density) as a function of electrode potential, in comparison with the commercial<br />

Pt/C catalyst. Estimation of ECSA was based on integrated H upd for the Pt/C and acid treated PtNi/C catalysts, and CO ad stripping polarization curve<br />

for the acid treated/annealed PtNi/C catalyst.<br />

measurements (see the Experimental Section). Figure 4c e<br />

summarizes the electrochemical studies for the three types of<br />

nanocatalysts. It can be seen from the voltammograms (Figure 4c)<br />

that the Hupd region (E < 0.4 V) of the acid treated/annealed<br />

catalyst is largely suppressed versus the acid treated sample. On the<br />

positive side of the potential scale, the onset of oxide formation<br />

for the acid treated/annealed catalyst is shifted positively by<br />

about 20 mV versus that for the acid treated catalyst, and more<br />

than 50 mV with respect to Pt/C. Similar shifts are also seen for<br />

the reduction peaks in the cathodic scans. Such peak shifts are<br />

representative of a less oxophilic catalyst surface due to the<br />

formation of multilayered Pt-skin structure, and further corresponding<br />

to remarkable enhancement in the ORR activity as<br />

evidenced by the polarization curves shown in Figure 4d and the<br />

Tafel plots, Figure 4e. These findings are reminiscent of those on<br />

extended surfaces (Figure 1) and from in situ XANES studies<br />

(Figure 4b). At 0.95 V, the specific activity of the acid treated/<br />

annealed PtNi/C reaches 0.85 mA/cm 2 , as compared to<br />

0.35 mA/cm 2 for the acid treated specimen and 0.13 mA/cm 2<br />

for Pt/C. This translates into improvement factors versus Pt/C<br />

of 3 and over 6 for the acid treated and acid treated/annealed<br />

PtNi/C catalysts, respectively, which is also in line with the<br />

results obtained on extended surfaces (Figure 1c). Therefore,<br />

the electrochemical studies of nanocatalysts validated that the<br />

scheme of the near surface architecture established on extended<br />

surfaces had been successfully applied to nanocatalysts by forming<br />

a multilayered Pt-skin surface. Remarkably, the ECSA of this<br />

catalyst obtained from integrated H upd region was over 30%<br />

lower than that from CO stripping (Figure S2), which also confirms<br />

the formation of Pt-skin type of surface in the nanocatalyst.<br />

Moreover, the developed Pt bimetallic catalyst with the<br />

unique nanoscale architecture does not only show enhanced<br />

catalytic activity, but also improved catalyst durability for the<br />

ORR. Figure 5 summarizes the electrochemical results for the<br />

PtNi/C catalysts before and after 4000 potential cycles between<br />

0.6 and 1.1 V at 60 °C. Both the acid treated and the acid treated/<br />

annealed PtNi/C catalysts had minor losses (∼10%) in ECSA<br />

after cycling, in comparison to a substantial drop (∼40%) for Pt/<br />

C (Figure 5a). An additional observation was that the acid<br />

treated/annealed PtNi/C had only 15% loss in specific activity,<br />

in contrast to 57% for the acid treated catalyst and 38% for Pt/C<br />

(Figure 5b). We have also applied in situ XANES to monitor the<br />

catalyst structures in the durability studies (Figure 5d and e, and<br />

more details in the Supporting Information). Not surprisingly,<br />

the acid treated/annealed PtNi/C does not show visible changes,<br />

whereas reduction of absorption at the Ni edge was observed for<br />

the acid treated PtNi/C during and after potential cycling. These<br />

findings are in line with the elemental analysis of the PtNi/C<br />

catalysts after the durability experiments, which indicate no loss<br />

for the Ni content in the acid treated/annealed catalyst in<br />

contrast to the significant loss of Ni in the acid treated catalyst<br />

(Figures S5). It is thus assured that the multilayered Pt-skin<br />

formation has indeed provided complete protection of the Ni<br />

inside the catalyst and enabled the sustained high catalytic<br />

activity under fuel cell operating conditions. On the basis of<br />

that, in addition to diminished number of vulnerable undercoordinated<br />

Pt surface atoms after annealing, multilayered<br />

Pt-skin formation is also thick enough to protect subsurface<br />

Ni from dissolution that otherwise occurs through the placeexchange<br />

mechanism 12a (Figures 5 and S5). At the same time,<br />

14401 dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. 2011, 133, 14396–14403


Journal of the American Chemical Society ARTICLE<br />

Figure 5. Summary of electrochemical durability studies obtained by RDE before and after 4000 potential cycles between 0.6 and 1.1 V for the Pt/C<br />

and PtNi/C catalysts in 0.1 M HClO4 at 0.95 V and 60 °C: (a) specific surface area, (b) specific activity, and (c) mass activity. Activity improvement<br />

factors versus Pt/C before and after cycling were also shown for specific and mass activities in (b) and (c). Parts (d) and (e) show the XANES spectra<br />

recorded for the acid treated and acid treated/annealed PtNi/C catalysts at Ni K edge, at 1.0 V before and after potential cycling. Estimation of ECSA<br />

was based on integrated Hupd for the Pt/C and acid treated PtNi/C catalysts, and COad stripping polarization curve for the acid treated/annealed<br />

PtNi/C catalyst.<br />

the multilayered Pt-skin is thin enough to maintain typical skin-like<br />

properties (discussed above), which originates from altered<br />

electronic structures due to the presence of a desirable amount of<br />

subsurface Ni. As a result, the PtNi/C catalyst with multilayered<br />

Pt-skin surfaces exhibited improvement factors in mass activity of<br />

more than 1 order of magnitude after elongated potential cycling<br />

versus the Pt/C catalyst (Figure 5c).<br />

’ SUMMARY<br />

We have demonstrated the design and synthesis of an<br />

advanced Pt bimetallic catalyst, which simultaneously achieves<br />

high catalytic activity and superior durability for the ORR. The<br />

developed catalyst contains a unique nanoscale architecture with<br />

a PtNi core of 50 at% Ni and a multilayered Pt-skin surfaces. This<br />

structure was built up through synergistic studies of extended<br />

surfaces and nanocatalysts, with critical parameters such as<br />

particle size, thermal treatment, particle sintering, alloy composition,<br />

and elemental composition profile integrally designed and<br />

optimized. Delicate structure function correlation in the bimetallic<br />

electrocatalysts with composite nanostructures has been<br />

comprehensively resolved by employing state-of-the-art electron<br />

microscopy and in situ X-ray spectroscopy characterization. Our<br />

findings have immense implications for the development of<br />

heterogeneous catalysts and nanostructure engineering toward<br />

advanced functional materials.<br />

’ ASSOCIATED CONTENT<br />

bS Supporting Information. Additional material characterization<br />

and theoretical analysis. This material is available free of<br />

charge via the Internet at http://pubs.acs.org.<br />

’ AUTHOR INFORMATION<br />

Corresponding Author<br />

vrstamenkovic@anl.gov<br />

’ ACKNOWLEDGMENT<br />

This work was conducted at Argonne National Laboratory, a<br />

U.S. Department of Energy, Office of Science Laboratory,<br />

operated by UChicago Argonne, LLC, under contract no. DE-<br />

AC02-06CH11357. It was sponsored by the U.S. Department of<br />

Energy, Office of Energy Efficiency and Renewable Energy, Fuel<br />

Cell Technologies Program. Microscopy research was conducted<br />

14402 dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. 2011, 133, 14396–14403


Journal of the American Chemical Society ARTICLE<br />

at the Electron Microscopy Center for Materials Research at<br />

Argonne, and ORNL’s SHaRE User Facility sponsored by the<br />

Scientific User Facilities Division, Office of Basic Energy<br />

Sciences, the U.S. Department of Energy. XANES were accomplished<br />

at the Advanced Photon Source at Argonne. We thank<br />

<strong>Dr</strong>. Cindy Chaffee for the help on setup at APS, and <strong>Dr</strong>. Sonke<br />

Seifert and Byeongdu Lee from APS for valuable discussion on<br />

X-ray absorption experiments.<br />

’ REFERENCES<br />

(1) Vielstich, W.; Lamm, A.; Gasteiger, H. A. Handbook of Fuel Cells:<br />

Fundamentals, Technology, and Applications; Wiley: Chichester, England;<br />

Hoboken, NJ, 2003; p 4.<br />

(2) (a) Abraham, K. M.; Jiang, Z. J. Electrochem. Soc. 1996, 143,1–5.<br />

(b) Armand, M.; Tarascon, J. M. Nature 2008, 451, 652–657.<br />

(3) Gasteiger, H. A.; Kocha, S. S.; Sompalli, B.; Wagner, F. T. Appl.<br />

Catal., B 2005, 56, 9–35.<br />

(4) (a) Stamenkovic, V.; Mun, B. S.; Mayrhofer, K. J. J.; Ross, P. N.;<br />

Markovic, N. M.; Rossmeisl, J.; Greeley, J.; Norskov, J. K. Angew. Chem.,<br />

Int. Ed. 2006, 45, 2897–2901. (b) Stamenkovic, V. R.; Mun, B. S.; Arenz,<br />

M.; Mayrhofer, K. J. J.; Lucas, C. A.; Wang, G. F.; Ross, P. N.; Markovic,<br />

N. M. Nat. Mater. 2007, 6, 241–247.<br />

(5) Stamenkovic, V. R.; Mun, B. S.; Mayrhofer, K. J. J.; Ross, P. N.;<br />

Markovic, N. M. J. Am. Chem. Soc. 2006, 128, 8813–8819.<br />

(6) Stamenkovic, V. R.; Fowler, B.; Mun, B. S.; Wang, G. F.; Ross,<br />

P. N.; Lucas, C. A.; Markovic, N. M. Science 2007, 315, 493–497.<br />

(7) (a) Strasser, P.; Mani, P.; Srivastava, R. J. Phys. Chem. C 2008,<br />

112, 2770–2778. (b) Chen, S.; Sheng, W. C.; Yabuuchi, N.; Ferreira,<br />

P. J.; Allard, L. F.; Shao-Horn, Y. J. Phys. Chem. C 2009, 113, 1109–1125.<br />

(c) Chen, J. Y.; Lim, B.; Lee, E. P.; Xia, Y. N. Nano Today 2009, 4,81–95.<br />

(d) Zhang, J.; Yang, H. Z.; Fang, J. Y.; Zou, S. Z. Nano Lett. 2010,<br />

10, 638–644. (e) Strasser, P.; Koh, S.; Anniyev, T.; Greeley, J.; More, K.;<br />

Yu, C. F.; Liu, Z. C.; Kaya, S.; Nordlund, D.; Ogasawara, H.; Toney,<br />

M. F.; Nilsson, A. Nat. Chem. 2010, 2, 454–460.<br />

(8) (a) Wang, C.; Wang, G. F.; van der Vliet, D.; Chang, K. C.;<br />

Markovic, N. M.; Stamenkovic, V. R. Phys. Chem. Chem. Phys. 2010,<br />

12, 6933–6939. (b) Schulenburg, H.; Muller, E.; Khelashvili, G.; Roser,<br />

T.; Bonnemann, H.; Wokaun, A.; Scherer, G. G. J. Phys. Chem. C 2009,<br />

113, 4069–4077.<br />

(9) (a) van Hardeveld, R.; Hartog, F. Surf. Sci. 1969, 15, 189–230.<br />

(b) Kinoshita, K. J. Electrochem. Soc. 1990, 137, 845–848.<br />

(10) (a) Han, B. C.; Miranda, C. R.; Ceder, G. Phys. Rev. B 2008,<br />

77, 075410. (b) Strmcnik, D. S.; Tripkovic, D. V.; van der Vliet, D.;<br />

Chang, K. C.; Komanicky, V.; You, H.; Karapetrov, G.; Greeley, J.;<br />

Stamenkovic, V. R.; Markovic, N. M. J. Am. Chem. Soc. 2008, 130,<br />

15332–15339.<br />

(11) (a) Mayrhofer, K. J. J.; Blizanac, B. B.; Arenz, M.; Stamenkovic,<br />

V. R.; Ross, P. N.; Markovic, N. M. J. Phys. Chem. B 2005,<br />

109, 14433–14440. (b) Wang, C.; van der Vilet, D.; Chang, K. C.;<br />

You, H. D.; Strmcnik, D.; Schlueter, J. A.; Markovic, N. M.; Stamenkovic,<br />

V. R. J. Phys. Chem. C 2009, 113, 19365–19368.<br />

(12) (a) Komanicky, V.; Chang, K. C.; Menzel, A.; Markovic, N. M.;<br />

You, H.; Wang, X.; Myers, D. J. Electrochem. Soc. 2006, 153, B446–B451.<br />

(b) Borup, R.; et al. Chem. Rev. 2007, 107, 3904–3951.<br />

(13) Wang, C.; Chi, M.; Wang, G.; van der Vliet, D.; Li, D.; More, K.;<br />

Wang, H.-H.; Schlueter, J. A.; Markovic, N. M.; Stamenkovic, V. R. Adv.<br />

Funct. Mater. 2011, 21, 147–152.<br />

(14) Beno, M. A.; Engbretson, M.; Jennings, G.; Knapp, G. S.;<br />

Linton, J.; Kurtz, C.; Rutt, U.; Montano, P. A. Nucl. Instrum. Methods<br />

Phys. Res., Sect A 2001, 467, 699–702.<br />

(15) Newville, M.; Livins, P.; Yacoby, Y.; Rehr, J. J.; Stern, E. A. Phys.<br />

Rev. B 1993, 47, 14126–14131.<br />

(16) (a) Newville, M. J. Synchrotron Radiat. 2001, 8, 322–324.<br />

(b) Ravel, B.; Newville, M. J. Synchrotron Radiat. 2005, 12, 537–541.<br />

(17) (a) Newman, J. J. Electrochem. Soc. 1966, 113, 2. (b) van der Vliet,<br />

D.; Strmcnik, D. S.; Wang, C.; Stamenkovic, V. R.; Markovic, N. M.;<br />

Koper, M. T. M. J. Electroanal. Chem. 2010, 647, 29–34.<br />

(18) Ahrenstorf, K.; Heller, H.; Kornowski, A.; Broekaert, J. A. C.;<br />

Weller, H. Adv. Funct. Mater. 2008, 18, 3850–3856.<br />

(19) Kiely, C. J.; Fink, J.; Brust, M.; Bethell, D.; Schiffrin, D. J. Nature<br />

1998, 396, 444–446.<br />

(20) Li, Y.; Zhang, X. L.; Qiu, R.; Qiao, R.; Kang, Y. S. J. Phys. Chem.<br />

C 2007, 111, 10747–10750.<br />

(21) Bunker, G. Introduction to XAFS: A Practical Guide to X-ray<br />

Absorption Fine Structure Spectroscopy; Cambridge University Press:<br />

Cambridge, UK; New York, 2010; viii, 260 pp.<br />

(22) (a) Muker<strong>je</strong>e, S.; Srinivasan, S.; Soriaga, M. P.; Mcbreen, J.<br />

J. Electrochem. Soc. 1995, 142, 1409–1422. (b) Rodriguez, J. A.; Hanson,<br />

J. C.; Frenkel, A. I.; Kim, J. Y.; Perez, M. J. Am. Chem. Soc. 2002,<br />

124, 346–354.<br />

(23) (a) Imai, H.; Izumi, K.; Matsumoto, M.; Kubo, Y.; Kato, K.;<br />

Imai, Y. J. Am. Chem. Soc. 2009, 131, 6293–6300. (b) Friebel, D.; Miller,<br />

D. J.; O’Grady, C. P.; Anniyev, T.; Bargar, J.; Bergmann, U.; Ogasawara,<br />

H.; Wikfeldt, K. T.; Pettersson, L. G. M.; Nilsson, A. Phys. Chem. Chem.<br />

Phys. 2011, 13, 262–266. (c) Russell, A. E.; Rose, A. Chem. Rev. 2004,<br />

104, 4613–4635.<br />

14403 dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. 2011, 133, 14396–14403


1 Design and Synthesis of Bimetallic Electrocatalyst with Multilayered<br />

2 Pt-Skin Surfaces<br />

3 Chao Wang, † Miaofang Chi, ‡ Dongguo Li, †,§ Dusan Strmcnik, † Dennis van der Vliet, † Guofeng Wang, ||<br />

4 Vladimir Komanicky, † Kee-Chul Chang, † Arvydas P. Paulikas, † Dusan Tripkovic, † John Pearson, †<br />

5 KarrenL.More, ‡ Nenad M. Markovic, † and Vojislav R. Stamenkovic* ,†<br />

6<br />

7<br />

8<br />

9<br />

†<br />

Materials Science Division, Argonne National Laboratory, Argonne, Illinois 60439, United States<br />

‡<br />

Division of Material Science and Technology, Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, United States<br />

§<br />

Department of Chemistry, Brown University, Providence, RI 02912, United States<br />

Department of Mechanical Engineering and Materials Science, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States<br />

)<br />

10 bS Supporting Information<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

ABSTRACT: Advancement in heterogeneous catalysis relies<br />

on the capability of altering material structures at the nanoscale,<br />

and that is particularly important for the development of highly<br />

active electrocatalysts with uncompromised durability. Here, we<br />

report the design and synthesis of a Pt-bimetallic catalyst with<br />

multilayered Pt-skin surface, which shows superior electrocatalytic<br />

performance for the oxygen reduction reaction (ORR).<br />

This novel structure was first established on thin film extended<br />

surfaces with tailored composition profiles and then implemented<br />

in nanocatalysts by organic solution synthesis. Electrochemical<br />

studies for the ORR demonstrated that after prolonged<br />

exposure to reaction conditions, the Pt-bimetallic catalyst with<br />

multilayered Pt-skin surface exhibited an improvement factor of more than 1 order of magnitude in activity versus conventional Pt<br />

catalysts. The substantially enhanced catalytic activity and durability indicate great potential for improving the material properties by<br />

fine-tuning of the nanoscale architecture.<br />

27 ’ INTRODUCTION<br />

Journal of the American Chemical Society | 3b2 | ver.9 | 17/8/011 | 7:51 | Msc: ja-2011-047655 | TEID: dmadmin | BATID: 00000 | Pages: 7.66<br />

28 The foreground of sustainable energy is built up on a renew-<br />

29 able and environmentally compatible scheme of chemical-electrical<br />

30 energy conversion. One of the key processes for such energy<br />

31 conversion is the electrocatalytic reduction of oxygen, the<br />

32 cathode reaction in fuel cells 1 and metal air batteries, 2 where<br />

33 an electrocatalyst is used to accelerate the course of ORR.<br />

34 Current electrocatalysts used for this reaction are typically in<br />

35 the form of dispersed Pt nanoparticles (NPs) on amorphous<br />

36 high-surface-area carbon. Considering the high cost and limited<br />

37 resource of Pt, large-scale applications of these renewable energy<br />

38 technologies demand substantial improvement of the catalyst<br />

39 performance so that the amount of Pt needed can be significantly<br />

40 reduced. For example, a 5-fold improvement of catalytic activity<br />

41 for the ORR is required for the commercial implementation of<br />

42 fuel cell technology in transportation. 3<br />

43 Our recent work on well-defined extended surfaces has shown<br />

44 that high catalytic activity for the ORR can be achieved on<br />

45 Pt bimetallic alloys (Pt 3M, M = Fe, Co, Ni, etc.), due to the<br />

46 altered electronic structure of the Pt topmost layer and hence<br />

47 reduced adsorption of oxygenated spectator species (e.g., OH )<br />

48 on the surface. 4 It was also found that in an acidic electrochemical<br />

49 environment the non-noble 3d transition metals are dissolved<br />

from the near-surface layers, which leads to the formation of<br />

Pt-skeleton surfaces. Moreover, the thermal treatment of Pt3M alloys in ultra high vacuum (UHV) has been shown to induce<br />

segregation of Pt and formation of a distinctive topmost layer<br />

that was termed Pt-skin surface. However, the same treatment<br />

did not cause Pt to segregate over PtM alloys with high content<br />

(g50%) of non-Pt elements. 4b,5 Recently, we further demonstrated<br />

the surfacing of an ordered Pt(111)-skin over Pt3Ni(111) single crystal with 50% of Ni in the subsurface layer. This unique<br />

nanosegregated composition profile was found to be responsible<br />

for the dramatically enhanced ORR activity. 6<br />

On the basis of these findings, it could be envisioned that the<br />

most advantageous nanoscale architecture for a bimetallic electrocatalyst<br />

would correspond to the segregated Pt-skin composition<br />

profile established on extended surfaces. A lot of effort has<br />

thus been dedicated, 7 but it still remains elusive, to finely tune the<br />

Pt-bimetallic nanostructure to achieve this desirable surface<br />

structure and composition profile. Major obstacles reside not<br />

only in the difficulty to manipulate elemental distribution at the<br />

nanoscale, but also in the fundamental differences in atomic<br />

Received: May 24, 2011<br />

ARTICLE<br />

pubs.acs.org/JACS<br />

r XXXX American Chemical Society A dx.doi.org/10.1021/ja2047655 | J. Am. Chem. Soc. XXXX, XXX, 000–000<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

62<br />

63<br />

64<br />

65<br />

66<br />

67<br />

68<br />

69


Journal of the American Chemical Society ARTICLE<br />

70 structures, electronic properties, and catalytic performance between<br />

71 extended surfaces and confined nanomaterials. For example, in<br />

72 an attempt to induce surface segregation, high-temperature (>600 °C)<br />

73 annealing is typically applied for Pt-based alloy nanocatalysts.<br />

74 While improvement in specific activity is obtained, such treatment<br />

75 usually causes particle sintering and loss of electrochemical<br />

76 surface area (ECSA). 7b,8 Besides that, the surface coordination<br />

77 of nanomaterials is quite different from that of bulk materials;<br />

78 that is, the surface of NPs is rich in corner and edge sites, which<br />

79 have a smaller coordination number than the atoms on long-<br />

80 range ordered terraces of extended surfaces. 9 These low-coordi-<br />

81 nation surface atoms are considered as preferential sites for the<br />

82 adsorption of oxygenated spectator species (e.g., OH ) 10 and<br />

83 thus become blocked for adsorption of molecular oxygen and<br />

84 inactive for the ORR. 9b,11 Additionally, due to strong Pt O<br />

85 interaction, these sites are more vulnerable for migration and<br />

86 dissolution, resulting in poor durability and fast decay of the<br />

87 catalyst. 12 The latter effect is even more pronounced in Pt bi-<br />

88 metallic systems, considering that more undercoordinated atoms<br />

89 are present on the skeleton surfaces formed after the depletion of<br />

90 nonprecious metals from near-surface regions. 4a,13 Therefore, a<br />

91 systematic approach with all of these factors integrally considered<br />

92 becomes necessary to pursue the design and synthesis of<br />

93 advanced bimetallic catalysts.<br />

94 Our focus in this study has been placed on the fine-tuning of<br />

95 Pt bimetallic nanostructure aiming to achieve the advantageous<br />

96 Pt-skin surface structure and composition profile established on<br />

97 extended surfaces. We started with Pt thin films of controlled<br />

98 thickness deposited over PtNi substrate to explore the correlation<br />

99 between the surface composition profile and catalytic performance.<br />

100 These findings were then applied for guiding the synthesis of<br />

101 nanocatalysts with the optimized structure. The outcome of such<br />

102 effort is an advanced Pt bimetallic catalyst with altered nano-<br />

103 scale architecture that is highly active and durable for the ORR.<br />

104 ’ EXPERIMENTAL SECTION<br />

105 Thin Film Preparation. Pt films were deposited at room tempera-<br />

106 ture on PtNi substrates (6 mm in diameter), which were set 125 mm<br />

107 away from DC sputter magnetrons in 4 mTorr argon gas (base vacuum<br />

108 1 10 7 Torr). The Pt source rate (0.32 Å/s) was determined by quartz<br />

109 crystal microbalance, and an exposure duration of 7.0 s was calibrated for<br />

110 the nominal thickness of 2.2 2.3 Å for a monolayer of Pt. The film<br />

111 thickness was derived from the exposure time of computer-controlled<br />

112 shutters during sputtering.<br />

113 NP and Catalyst Synthesis. In a typical synthesis of PtNi NPs,<br />

114 0.67 mmol of Ni(ac)2 3 4H2O was dissolved in 20 mL of diphenyl either<br />

115 in the presence of 0.4 mL of oleylamine and 0.4 mL of oleic acid.<br />

116 0.33 mmol of 1,2-tetradodecanediol was added, and the formed solution<br />

117 washeatedto80°C under Ar flow. After a transparent solution formed, the<br />

118 temperature was further raised to 200 °C, where 0.33 mmol of Pt(acac)2<br />

119 dissolved in 1.5 mL of dichlorobenzene was in<strong>je</strong>cted. The solution was<br />

120 kept at this temperature for 1 h and then cooled to room temperature.<br />

121 An amount of 60 mL of ethanol was added to precipitate the NPs, and<br />

122 the product was collected by centrifuge (6000 rpm, 6 min). The<br />

123 obtained NPs were further washed by ethanol two times and then<br />

124 dispersed in hexane. The as-synthesized PtNi NPs were deposited on<br />

125 high surface area carbon (∼900 m 2 /g) by mixing the NPs with carbon<br />

126 black (Tanaka, KK) in hexane or chloroform with a 1:1 ratio in mass.<br />

127 This mixture was sonicated for 1 h and then dried under nitrogen flow.<br />

128 The organic surfactants were removed by thermal treatment at<br />

129 150 200 °C in an oxygenated atmosphere. The obtained catalyst is<br />

130 denoted as “as-prepared PtNi/C”. For the acid treatment, ∼10 mg of the<br />

as-prepared PtNi/C catalyst was mixed with 20 mL of 0.1 M HClO4 that<br />

has been used as electrolyte in electrochemical measurements. After<br />

overnight exposure to the acidic environment, the product was collected<br />

by centrifuge and washed three times by deionized water. Such NPs are<br />

named as “acid treated PtNi/C”. The acid treated PtNi/C was further<br />

annealed at 400 °C to reduce low-coordinated surface sites, and the<br />

obtained catalyst is termed as “acid treated/annealed PtNi/C”.<br />

Microscopic Characterization. TEM images were collected on a<br />

Philips EM 30 (200 kV) equipped with EDX functionality. XRD patterns<br />

were collected on a Rigaku RTP 300 RC machine. STEM and elemental<br />

analysis were carried out on a JEOL 2200FS TEM/STEM with a CEOS<br />

aberration (probe) corrector. The microscope was operated at 200 kV in<br />

HAADF-STEM mode equipped with a Bruker-AXS X-Flash 5030 silicon<br />

drift detector. The probe size was ∼0.7 Å and the probe current was<br />

∼30 pA during HAADF-STEM imaging. When accumulating EDX data,<br />

the probe current was increased to ∼280 pA and the probe size was<br />

∼2 Å. The presented EDX data were confirmed to be from “e-beam<br />

damage-free” particles by comparing STEM images before and after<br />

EDX acquisition.<br />

X-ray Absorption Spectroscopy. X-ray fluorescence spectra of<br />

at the Ni K and Pt L3 edges were acquired at bending magnet beamline<br />

12-BM-B at the Advanced Photon Source (APS), Argonne National<br />

Laboratory. The incident radiation was filtered by a Si(111) doublecrystal<br />

monochromator (energy resolution ΔE/E = 14.1 10 5 ) with a<br />

double mirror system for focusing and harmonic re<strong>je</strong>ction. 14 All of the<br />

data were taken in fluorescence mode using a 13-element Germanium<br />

array detector (Canberra), which was aligned with the polarization of the<br />

X-ray beam to minimize the elastic scattering intensity. Co and Ge filters<br />

(of 6 absorption length in thickness) were applied in front of the<br />

detector to further reduce the elastic scattering intensity for the Ni K and<br />

Pt L3 edges, respectively. The Ni K and Pt L3 edge spectra were<br />

calibrated by defining the zero crossing point of the second derivative<br />

of the XANES spectra for Ni and Pt reference foils as 8333 and 11564 eV,<br />

respectively. The background was subtracted using the AUTOBK<br />

algorithm, 15 and data reduction was performed using Athena from the<br />

IFEFFIT software suite. 16 A scheme of the homemade in situ electrochemical<br />

cell and setup at beamline was shown in the Supporting<br />

Information, Figure S9.<br />

Electrochemical Characterization. The electrochemical measurements<br />

were conducted in a three-compartment electrochemical cell<br />

with a rotational disk electrode (RDE, 6 mm in diameter) setup (Pine)<br />

and a Autolab 302 potentiostat. A saturated Ag/AgCl electrode and a Pt<br />

wire were used as reference and counter electrodes, respectively. 0.1 M<br />

HClO4 was used as electrolyte. The catalysts were deposited on glassy<br />

carbon electrode substrate and dried in Ar atmosphere without using any<br />

ionomer. The loading was controlled to be 12 μgPt/cm 2 disk for PtNi/C<br />

nanocatalysts. All of the potentials given in the discussion were against<br />

reversible hydrogen electrode (RHE), and the readout currents were<br />

recorded with ohmic iR drop correction during the measurements. 17<br />

’ RESULTS AND DISCUSSION<br />

Pt films of various thicknesses, that is, 1 7 atomic monolayers<br />

(ML), were deposited in a vacuum by sputtering on PtNi<br />

(Pt:Ni = 1:1) substrate and then transferred to an electrochemical<br />

cell for further characterizations (see the Experimental<br />

Section). The as-sputtered Pt films consist of randomly distributed<br />

Pt nanoclusters (


Journal of the American Chemical Society ARTICLE<br />

Figure 1. Electrochemical studies on the Pt thin films deposited over PtNi substrate by RDE: (a) cyclic voltammograms, (b) polarization curves,<br />

and (c) summary of specific activities and corresponding improvement factors (vs polycrystalline Pt surface) for the Pt films of various<br />

thicknesses. Cyclic voltammograms were recorded in Ar saturated 0.1 M HClO4 electrolyte with a sweeping rate of 50 mV/s. Polarization curves<br />

were recorded in the same electrolyte under O2 saturation with a sweep rate of 20 mV/s. Specific activities were presented as kinetic currents<br />

normalized by ECSAs obtained from integrated Hupd, except that for the annealed 3 ML Pt/PtNi surface which was based on COad stripping<br />

polarization curve.<br />

193 which had confirmed the superior catalytic properties of<br />

F1 194 systems with 50% of Ni in subsurface layers. Figure 1 summarizes<br />

195 the results of electrochemical studies for these thin films by<br />

196 rotating disk electrode (RDE). Cyclic voltammograms (CVs,<br />

197 Figure 1a) of the as-sputtered films correspond to polycrystalline<br />

198 Pt (poly-Pt) with similar, but slightly enlarged, underpotentially<br />

199<br />

200<br />

deposited hydrogen (Hupd) regions (E < 0.4 V) due to the<br />

rougher surfaces. Consistent with our previous findings, 6 the<br />

201<br />

202<br />

onset of Pt OHad formation has anodic shifts for most of the Pt<br />

films (e5 ML) as compared to poly-Pt (more visible in the CVs<br />

203 shown in the Supporting Information, Figure S1, with currents<br />

204 normalized by the electrochemical surface area (ECSA) obtained<br />

205<br />

206<br />

from integrated Hupd region). Correspondingly, similar positive<br />

shifts are also present in the polarization curves for the ORR<br />

207 (Figure 1b). The largest shift of ∼30 mV was obtained for the Pt<br />

208 films with thicknesses of three atomic layers. Measured specific<br />

209 activities at 0.95 V, expressed as kinetic current normalized by the<br />

210 ECSA, show that the thinner films (e3 ML) have improvement<br />

211 factors of ∼2.5 versus poly-Pt surface, which is in line with the<br />

212<br />

213<br />

previous results on polycrystalline Pt3M bulk alloys with the<br />

skeleton type of surfaces. 4b Reduced enhancement was observed<br />

214 for thicker Pt films, for example, improvement factor of 1.7 for<br />

215 5MLofPt,whilethespecificactivity measured for the 7 ML<br />

216 film was close to that of poly-Pt. It should be noted here that<br />

217 for the as-sputtered films,1MLofPtmaynotbeableto<br />

218 entirely cover Ni atoms in the alloy substrate and protect<br />

219 them from dissolution, whereas addition of a second and/or<br />

220 third layer can effectively diminish this process. Along the<br />

same lines, this may also be the reason that the surfaces with 2 or<br />

3 ML of Pt were found to be more active than that with 1 ML.<br />

These findings revealed that bimetallic systems with Pt-skeleton<br />

near-surface formation of up to three atomic layers in thickness<br />

are also capable of efficiently harvesting the beneficial properties<br />

of bimetallic alloys, while protecting the subsurface Ni from<br />

leaching out.<br />

Because the as-sputtered skeleton type of surfaces have abundant<br />

low-coordination sites 4a that are detrimental to the ORR, we have<br />

applied thermal treatment to investigate potential surface restructuring<br />

and further catalytic improvement. A moderate temperature<br />

of ∼400 °C was chosen as it was determined to be optimal for<br />

Pt bimetallic nanocatalysts. 8a In Figure 1a, the CV of annealed<br />

3 ML Pt/PtNi surface is also shown. The suppressed Hupd region<br />

and an even larger positive shift of the Pt OHad peak (Figure S1)<br />

indicates the formation of Pt-skin type of surface, which is smoother<br />

and less oxophilic with significantly reduced number of lowcoordination<br />

surface atoms. 5,6 Additional proof of the transition<br />

toward Pt-skin is provided by the measured boost in specific<br />

activity for the ORR (Figure 1c), reaching an improvement factor<br />

of more than 5 with respect to poly-Pt (Figure 1d). Moreover,<br />

this high catalytic activity was based on the ECSA estimated from<br />

CO stripping polarization curves, not Hupd. The ECSA estimated<br />

from integrated Hupd charge was found to be substantially smaller<br />

than that obtained from the electrochemical oxidation of adsorbed<br />

CO monolayer (Figure S2), which was not observed on unannealed<br />

Pt-skeleton surfaces. Such a difference can only be interpreted<br />

in terms of the altered electronic properties of the Pt-skin surface<br />

C dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. XXXX, XXX, 000–000<br />

221<br />

222<br />

223<br />

224<br />

225<br />

226<br />

227<br />

228<br />

229<br />

230<br />

231<br />

232<br />

233<br />

234<br />

235<br />

236<br />

237<br />

238<br />

239<br />

240<br />

241<br />

242<br />

243<br />

244<br />

245<br />

246<br />

247<br />

248


Journal of the American Chemical Society ARTICLE<br />

Figure 2. Representative transmission electron microscopy (TEM)<br />

images for (a,b) the as-synthesized PtNi NPs, (c) the as-prepared, and<br />

(d) the acid treated/annealed PtNi/C catalysts. (d) X-ray diffraction<br />

(XRD) patterns for the PtNi/C catalysts in comparison with commercial<br />

Pt/C (∼6 nm in particle size).<br />

249 that have affected the adsorption of hydrogenated species, but<br />

250 not the Pt CO ad interaction. 6<br />

251 The above studies of Pt thin films over PtNi substrate, as well<br />

252<br />

253<br />

as the knowledge previously acquired from poly-/single-crystalline<br />

surfaces 4 6 and nanocatalysts, 8a,11b,13 have led us to a novel<br />

254 approach toward the design and synthesis of Pt bimetallic<br />

255 catalysts with Pt terminated surfaces. The initial step in this<br />

256 approach should involve the synthesis of monodisperse and<br />

257 homogeneous PtNi NPs followed by intentional depletion of<br />

258 Ni from the surface, producing a skeleton type of surface<br />

259 structure. The final step is supposed to be the thermal treatment<br />

260 aimed to induce the transition from Pt-skeleton into Pt-skin type<br />

261 of structure by surface relaxation and restructuring. For that<br />

262 purpose, PtNi NPs were synthesized by simultaneous reduction of<br />

263<br />

264<br />

platinum acetylacetonate, Pt(acac) 2, and nickel acetate, Ni(ac) 2,<br />

in an organic solution (see the Experimental Section). 13,18<br />

F2 265 Figure 2a<br />

and b shows representative transmission electron<br />

266 microscopy (TEM) images of the as-synthesized PtNi NPs<br />

267<br />

268<br />

prepared with a molar ratio of 1:2 between the Pt and Ni<br />

precursors. The NP size was controlled to be ∼5 nm 11b with a<br />

269 very narrow size distribution, as evidenced by the formation of<br />

270<br />

271<br />

various types of super lattices after drying the nanoparticle<br />

suspension (in hexane) under ambient conditions. 19 The final<br />

272 composition was characterized by energy-dispersive X-ray spec-<br />

273 troscopy (EDX), which confirmed an atomic ratio of Pt/Ni ≈ 1/1<br />

274<br />

275<br />

(Figure S3). The as-synthesized NPs were incorporated into<br />

carbon black (∼900 m 2 /g) via a colloidal-deposition method,<br />

276<br />

277<br />

and the organic surfactants were efficiently removed by thermal<br />

treatment. 11b 278<br />

Such as-prepared PtNi/C catalyst was first treated<br />

by acid to dissolve the surface Ni atoms 13 (Figure S3) and then<br />

279 annealed at 400 °C. These consecutive treatments were expected<br />

to bring on the Pt-skin type of surface over the substrate with 280<br />

50% of Ni, which otherwise would not be possible because<br />

complete segregation of Pt only takes place in Pt3M systems.<br />

281<br />

282<br />

5<br />

TEM images of the acid treated/annealed catalyst do not show 283<br />

notable changes in morphology (Figure 2c and d), except a slight 284<br />

decrease (∼0.3 nm) in average particle size (Figures S4). 285<br />

Additionally, X-ray diffraction (XRD) analysis was used to 286<br />

characterize the crystal structure of the NPs. As compared to 287<br />

the commercial Pt/C catalyst (Tanaka, ∼6 nm), both the as- 288<br />

prepared and the acid treated/annealed PtNi/C systems show a 289<br />

face-centered cubic (fcc) pattern with noticeable shifts (e.g., ∼1° 290<br />

for (111) peak) toward high angle (Figure 2e), corresponding to<br />

a decrease of lattice constant due to alloying between Pt and Ni.<br />

291<br />

292<br />

20<br />

The XRD pattern of the acid treated/annealed NPs has sharper 293<br />

peaks as compared to the as-prepared one, which indicates the 294<br />

increased crystallinity after annealing. These observations, in 295<br />

addition to the absence of peaks for separate Pt or Ni phases, 296<br />

show that the bimetallic catalyst preserved the alloy properties 297<br />

after the applied treatments.<br />

298<br />

The nanostructures and composition profiles of the PtNi/C 299<br />

catalysts were characterized by atomically resolved aberration- 300<br />

corrected high angle annular dark field-scanning transmission 301<br />

electron microscopy (HAADF-STEM) in combination with 302<br />

energy dispersive X-ray spectroscopy (EDX). Figure 3a shows 303 F3<br />

representative HAADF-STEM images taken along the Æ110æ 304<br />

zone axis of the as-prepared (left), acid treated (middle), and 305<br />

acid treated/annealed (right) PtNi/C catalysts, with the intensity 306<br />

profiles along Æ001æ directions across the single particles shown 307<br />

in Figure 3b. As compared to the benchmark intensity profiles 308<br />

calculated for ideal octahedral alloy NPs of the same size and 309<br />

orientation (see the Supporting Information), NP exposed to 310<br />

acid shows 3 4 peaks on each side stretching above the 311<br />

standards, indicating the formation of a Pt-rich overlayer. This 312<br />

feature was preserved after annealing, but with 2 3 Pt-rich peaks 313<br />

on each side, corresponding to a reduced Pt overlayer thickness 314<br />

due to restructuring and smoothing (Figure 3b). These findings 315<br />

were further confirmed by EDX analysis. By scanning the e-beam 316<br />

across the particle while simultaneously analyzing the generated 317<br />

X-rays, composition line profiles were obtained for the NPs 318<br />

(Figure 3c). It can be seen that the distribution of Pt and Ni in the 319<br />

as-prepared catalyst was highly intermixed and the sketched 320<br />

trend lines were almost identical, indicating a homogeneous 321<br />

alloy nature of the catalyst particles. The treated catalysts have 322<br />

substantially broader distribution of Pt than Ni, with a difference 323<br />

of ∼1 nm (at the half-maximum of the trend lines) for the acid 324<br />

treated and ∼0.6 nm for the acid treated/annealed catalyst. 325<br />

Hence, both the intensity and the composition line profiles show 326<br />

that multilayered Pt-rich surface structure was formed by acid 327<br />

treatment and preserved after annealing.<br />

328<br />

The microscopic characterizations strongly point toward sur- 329<br />

face restructuring in the bimetallic catalyst upon annealing. This 330<br />

was additionally depicted by atomistic simulation of the nano- 331<br />

structure evolution sub<strong>je</strong>ct to the acid and annealing treatments 332<br />

(Figure 3d for overviews and Figure 3e for cross-section views; 333<br />

see the Supporting Information for more details). It shows 334<br />

that removing Ni atoms from the surface led to the formation 335<br />

of a Pt-skeleton overlayer with a thickness of up to 3 atomic 336<br />

layers. Further relaxation of low-coordination surface atoms 337<br />

resulted in a multilayered Pt-skin surface, whereas the PtNi core 338<br />

was barely affected. It is important to mention that the relaxation 339<br />

process is expected to induce preferential formation of highly<br />

active (111) surface<br />

340<br />

341<br />

6 (labeled by arrows in Figure 3e), due to the<br />

D dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. XXXX, XXX, 000–000


Journal of the American Chemical Society ARTICLE<br />

Figure 3. Microscopic characterization and theoretical simulation of nanostructure evolution in the PtNi/C catalysts: (a) Representative high-angle<br />

annular dark-field scanning transmission electron microscopy (HAADF-STEM) images taken along the zone axis Æ110æ, as confirmed by the fast Fourier<br />

transfer (FFT) patterns of the STEM images (shown as insets); (b) background subtracted, normalized intensity line profiles extracted for the regions<br />

marked in (a); (c) composition line profiles (normalized for Pt L peaks) obtained by energy-dispersive X-ray spectroscopy (EDX) with an electron<br />

beam (∼2 Å in spot size) scanning across individual catalyst particles; (d) overview; and (e) cross-section views of the nanostructures depicted by<br />

atomistic particle simulation. The figure is also organized in columns for the as-prepared (left), acid treated (middle), and acid treated/annealed (right)<br />

PtNi/C catalysts, respectively.<br />

342 higher atomic coordination, that is, lower surface energy, of this<br />

343 facet as compared to others.<br />

344 To gain more insights into the nanostructure evolution,<br />

345 especially the correlation of surface structures to their electro-<br />

346 chemical properties, we have carried out in situ X-ray absorption<br />

347 near edge structure (XANES) studies of the nanocatalysts (see the<br />

348 Experimental Section and Supporting Information for details).<br />

F4 349 Figure 4a<br />

and b shows the normalized XANES spectra collected<br />

350<br />

351<br />

at Ni K and Pt L3 edges, under the ORR-relevant conditions<br />

(∼1.0 V). As compared to the spectra of reference foils, Pt and Ni<br />

352<br />

353<br />

edge positions were found to correspond to the bulk oxidation<br />

state of zero for both elements in the treated catalysts. 21 It is<br />

354 intriguing to see that the acid treated catalyst shows higher white<br />

355 line intensity than does the acid treated/annealed catalyst at the<br />

356 Ni edge, which is caused by the presence of a small amount of<br />

357<br />

358<br />

NiO underlying the highly corrugated Pt-skeleton surface<br />

morphology in the acid treated catalyst, 22 whereas subsurface<br />

Ni in the acid treated/annealed catalyst was well protected.<br />

The distinction in surface structure between the two treated<br />

catalysts is even more visible at the Pt edge, where a slightly lower<br />

white line intensity for the acid treated/annealed catalyst<br />

corresponds to a reduced amount of platinum oxides under the<br />

same conditions, and, more fundamentally, less oxophilic surface<br />

with larger average surface coordination number. 5,22a,23 The<br />

findings from XANES provide additional evidence for the<br />

formation of surface relaxed multilayered Pt-skin in the acid<br />

treated/annealed catalyst and its superiority in protecting the<br />

inner Ni from leaching out.<br />

On the basis of these results, we have managed to achieve the<br />

desirable nanoscale architecture established on PtNi supported<br />

Pt films, that is, multilayered Pt-skin over a particle core with 50%<br />

of Ni. Considering what was revealed from the studies on<br />

extended surfaces, the obtained nanocatalyst should show superior<br />

catalytic performance for the ORR, which was examined by RDE<br />

E dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. XXXX, XXX, 000–000<br />

359<br />

360<br />

361<br />

362<br />

363<br />

364<br />

365<br />

366<br />

367<br />

368<br />

369<br />

370<br />

371<br />

372<br />

373<br />

374<br />

375


Journal of the American Chemical Society ARTICLE<br />

Figure 4. In situ X-ray absorption and electrochemical studies of the PtNi/C catalysts. (a,b) XANES spectra for the PtNi/C catalysts recorded at Ni K<br />

and Pt L3 edges with an electrode potential of 1.0 V, in comparison with standard spectra of Ni, NiO, and Pt. (c) Cyclic voltammograms, (d) polarization<br />

curves, and (e) Tafel plots with the specific activity (jk, kinetic current density) as a function of electrode potential, in comparison with the commercial<br />

Pt/C catalyst. Estimation of ECSA was based on integrated H upd for the Pt/C and acid treated PtNi/C catalysts, and CO ad stripping polarization curve<br />

for the acid treated/annealed PtNi/C catalyst.<br />

376 measurements (see the Experimental Section). Figure 4c e<br />

377 summarizes the electrochemical studies for the three types of<br />

378 nanocatalysts. It can be seen from the voltammograms (Figure 4c)<br />

379 that the Hupd region (E < 0.4 V) of the acid treated/annealed<br />

380 catalyst is largely suppressed versus the acid treated sample. On the<br />

381 positive side of the potential scale, the onset of oxide formation<br />

382 for the acid treated/annealed catalyst is shifted positively by<br />

383 about 20 mV versus that for the acid treated catalyst, and more<br />

384 than 50 mV with respect to Pt/C. Similar shifts are also seen for<br />

385 the reduction peaks in the cathodic scans. Such peak shifts are<br />

386 representative of a less oxophilic catalyst surface due to the<br />

387 formation of multilayered Pt-skin structure, and further corre-<br />

388 sponding to remarkable enhancement in the ORR activity as<br />

389 evidenced by the polarization curves shown in Figure 4d and the<br />

390 Tafel plots, Figure 4e. These findings are reminiscent of those on<br />

391 extended surfaces (Figure 1) and from in situ XANES studies<br />

392 (Figure 4b). At 0.95 V, the specific activity of the acid treated/<br />

393 annealed PtNi/C reaches 0.85 mA/cm 2 , as compared to<br />

394 0.35 mA/cm 2 for the acid treated specimen and 0.13 mA/cm 2<br />

395 for Pt/C. This translates into improvement factors versus Pt/C<br />

396 of 3 and over 6 for the acid treated and acid treated/annealed<br />

397 PtNi/C catalysts, respectively, which is also in line with the<br />

398 results obtained on extended surfaces (Figure 1c). Therefore,<br />

399 the electrochemical studies of nanocatalysts validated that the<br />

400 scheme of the near surface architecture established on extended<br />

401 surfaces had been successfully applied to nanocatalysts by form-<br />

402 ing a multilayered Pt-skin surface. Remarkably, the ECSA of this<br />

403 catalyst obtained from integrated H upd region was over 30%<br />

404 lower than that from CO stripping (Figure S2), which also con-<br />

405 firms the formation of Pt-skin type of surface in the nanocatalyst.<br />

Moreover, the developed Pt bimetallic catalyst with the 406<br />

unique nanoscale architecture does not only show enhanced 407<br />

catalytic activity, but also improved catalyst durability for the 408<br />

ORR. Figure 5 summarizes the electrochemical results for the 409 F5<br />

PtNi/C catalysts before and after 4000 potential cycles between 410<br />

0.6 and 1.1 V at 60 °C. Both the acid treated and the acid treated/ 411<br />

annealed PtNi/C catalysts had minor losses (∼10%) in ECSA 412<br />

after cycling, in comparison to a substantial drop (∼40%) for Pt/ 413<br />

C (Figure 5a). An additional observation was that the acid 414<br />

treated/annealed PtNi/C had only 15% loss in specific activity, 415<br />

in contrast to 57% for the acid treated catalyst and 38% for Pt/C 416<br />

(Figure 5b). We have also applied in situ XANES to monitor the 417<br />

catalyst structures in the durability studies (Figure 5d and e, and 418<br />

more details in the Supporting Information). Not surprisingly, 419<br />

the acid treated/annealed PtNi/C does not show visible changes, 420<br />

whereas reduction of absorption at the Ni edge was observed for 421<br />

the acid treated PtNi/C during and after potential cycling. These 422<br />

findings are in line with the elemental analysis of the PtNi/C 423<br />

catalysts after the durability experiments, which indicate no loss 424<br />

for the Ni content in the acid treated/annealed catalyst in 425<br />

contrast to the significant loss of Ni in the acid treated catalyst 426<br />

(Figures S5). It is thus assured that the multilayered Pt-skin 427<br />

formation has indeed provided complete protection of the Ni 428<br />

inside the catalyst and enabled the sustained high catalytic 429<br />

activity under fuel cell operating conditions. On the basis of 430<br />

that, in addition to diminished number of vulnerable under- 431<br />

coordinated Pt surface atoms after annealing, multilayered 432<br />

Pt-skin formation is also thick enough to protect subsurface 433<br />

Ni from dissolution that otherwise occurs through the placeexchange<br />

mechanism<br />

434<br />

435<br />

12a (Figures 5 and S5). At the same time,<br />

F dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. XXXX, XXX, 000–000


Journal of the American Chemical Society ARTICLE<br />

Figure 5. Summary of electrochemical durability studies obtained by RDE before and after 4000 potential cycles between 0.6 and 1.1 V for the Pt/C<br />

and PtNi/C catalysts in 0.1 M HClO4 at 0.95 V and 60 °C: (a) specific surface area, (b) specific activity, and (c) mass activity. Activity improvement<br />

factors versus Pt/C before and after cycling were also shown for specific and mass activities in (b) and (c). Parts (d) and (e) show the XANES spectra<br />

recorded for the acid treated and acid treated/annealed PtNi/C catalysts at Ni K edge, at 1.0 V before and after potential cycling. Estimation of ECSA<br />

was based on integrated Hupd for the Pt/C and acid treated PtNi/C catalysts, and COad stripping polarization curve for the acid treated/annealed<br />

PtNi/C catalyst.<br />

436 the multilayered Pt-skin is thin enough to maintain typical skin-like<br />

437 properties (discussed above), which originates from altered<br />

438 electronic structures due to the presence of a desirable amount of<br />

439 subsurface Ni. As a result, the PtNi/C catalyst with multilayered<br />

440 Pt-skin surfaces exhibited improvement factors in mass activity of<br />

441 more than 1 order of magnitude after elongated potential cycling<br />

442 versus the Pt/C catalyst (Figure 5c).<br />

443 ’ SUMMARY<br />

444 We have demonstrated the design and synthesis of an<br />

445 advanced Pt bimetallic catalyst, which simultaneously achieves<br />

446 high catalytic activity and superior durability for the ORR. The<br />

447 developed catalyst contains a unique nanoscale architecture with<br />

448 a PtNi core of 50 at% Ni and a multilayered Pt-skin surfaces. This<br />

449 structure was built up through synergistic studies of extended<br />

450 surfaces and nanocatalysts, with critical parameters such as<br />

451 particle size, thermal treatment, particle sintering, alloy composi-<br />

452 tion, and elemental composition profile integrally designed and<br />

453 optimized. Delicate structure function correlation in the bime-<br />

454 tallic electrocatalysts with composite nanostructures has been<br />

455 comprehensively resolved by employing state-of-the-art electron<br />

456 microscopy and in situ X-ray spectroscopy characterization. Our<br />

457 findings have immense implications for the development of<br />

heterogeneous catalysts and nanostructure engineering toward<br />

advanced functional materials.<br />

’ ASSOCIATED CONTENT<br />

bS Supporting Information. Additional material characterization<br />

and theoretical analysis. This material is available free of<br />

charge via the Internet at http://pubs.acs.org.<br />

’ AUTHOR INFORMATION<br />

Corresponding Author<br />

vrstamenkovic@anl.gov<br />

’ ACKNOWLEDGMENT<br />

This work was conducted at Argonne National Laboratory, a<br />

U.S. Department of Energy, Office of Science Laboratory,<br />

operated by UChicago Argonne, LLC, under contract no. DE-<br />

AC02-06CH11357. It was sponsored by the U.S. Department of<br />

Energy, Office of Energy Efficiency and Renewable Energy, Fuel<br />

Cell Technologies Program. Microscopy research was conducted<br />

G dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. XXXX, XXX, 000–000<br />

458<br />

459<br />

460<br />

461<br />

462<br />

463<br />

464<br />

465<br />

466<br />

467<br />

468<br />

469<br />

470<br />

471<br />

472<br />

473


Journal of the American Chemical Society ARTICLE<br />

474 at the Electron Microscopy Center for Materials Research at<br />

475 Argonne, and ORNL’s SHaRE User Facility sponsored by the<br />

476 Scientific User Facilities Division, Office of Basic Energy<br />

477 Sciences, the U.S. Department of Energy. XANES were accom-<br />

478 plished at the Advanced Photon Source at Argonne. We thank<br />

479 <strong>Dr</strong>. Cindy Chaffee for the help on setup at APS, and <strong>Dr</strong>. Sonke<br />

480 Seifert and Byeongdu Lee from APS for valuable discussion on<br />

481 X-ray absorption experiments.<br />

482 ’ REFERENCES<br />

483 (1) Vielstich, W.; Lamm, A.; Gasteiger, H. A. Handbook of Fuel Cells:<br />

484 Fundamentals, Technology, and Applications; Wiley: Chichester, England;<br />

485 Hoboken, NJ, 2003; p 4.<br />

486 (2) (a) Abraham, K. M.; Jiang, Z. J. Electrochem. Soc. 1996, 143,1–5.<br />

487 (b) Armand, M.; Tarascon, J. M. Nature 2008, 451, 652–657.<br />

488 (3) Gasteiger, H. A.; Kocha, S. S.; Sompalli, B.; Wagner, F. T. Appl.<br />

489 Catal., B 2005, 56, 9–35.<br />

490 (4) (a) Stamenkovic, V.; Mun, B. S.; Mayrhofer, K. J. J.; Ross, P. N.;<br />

491 Markovic, N. M.; Rossmeisl, J.; Greeley, J.; Norskov, J. K. Angew. Chem.,<br />

492 Int. Ed. 2006, 45, 2897–2901. (b) Stamenkovic, V. R.; Mun, B. S.; Arenz,<br />

493 M.; Mayrhofer, K. J. J.; Lucas, C. A.; Wang, G. F.; Ross, P. N.; Markovic,<br />

494 N. M. Nat. Mater. 2007, 6, 241–247.<br />

495 (5) Stamenkovic, V. R.; Mun, B. S.; Mayrhofer, K. J. J.; Ross, P. N.;<br />

496 Markovic, N. M. J. Am. Chem. Soc. 2006, 128, 8813–8819.<br />

497 (6) Stamenkovic, V. R.; Fowler, B.; Mun, B. S.; Wang, G. F.; Ross,<br />

498 P. N.; Lucas, C. A.; Markovic, N. M. Science 2007, 315, 493–497.<br />

499 (7) (a) Strasser, P.; Mani, P.; Srivastava, R. J. Phys. Chem. C 2008,<br />

500 112, 2770–2778. (b) Chen, S.; Sheng, W. C.; Yabuuchi, N.; Ferreira,<br />

501 P. J.; Allard, L. F.; Shao-Horn, Y. J. Phys. Chem. C 2009, 113, 1109–1125.<br />

502 (c) Chen, J. Y.; Lim, B.; Lee, E. P.; Xia, Y. N. Nano Today 2009, 4,81–95.<br />

503 (d) Zhang, J.; Yang, H. Z.; Fang, J. Y.; Zou, S. Z. Nano Lett. 2010,<br />

504 10, 638–644. (e) Strasser, P.; Koh, S.; Anniyev, T.; Greeley, J.; More, K.;<br />

505 Yu, C. F.; Liu, Z. C.; Kaya, S.; Nordlund, D.; Ogasawara, H.; Toney,<br />

506 M. F.; Nilsson, A. Nat. Chem. 2010, 2, 454–460.<br />

507 (8) (a) Wang, C.; Wang, G. F.; van der Vliet, D.; Chang, K. C.;<br />

508 Markovic, N. M.; Stamenkovic, V. R. Phys. Chem. Chem. Phys. 2010,<br />

509 12, 6933–6939. (b) Schulenburg, H.; Muller, E.; Khelashvili, G.; Roser,<br />

510 T.; Bonnemann, H.; Wokaun, A.; Scherer, G. G. J. Phys. Chem. C 2009,<br />

511 113, 4069–4077.<br />

512 (9) (a) van Hardeveld, R.; Hartog, F. Surf. Sci. 1969, 15, 189–230.<br />

513 (b) Kinoshita, K. J. Electrochem. Soc. 1990, 137, 845–848.<br />

514 (10) (a) Han, B. C.; Miranda, C. R.; Ceder, G. Phys. Rev. B 2008,<br />

515 77, 075410. (b) Strmcnik, D. S.; Tripkovic, D. V.; van der Vliet, D.;<br />

516 Chang, K. C.; Komanicky, V.; You, H.; Karapetrov, G.; Greeley, J.;<br />

517 Stamenkovic, V. R.; Markovic, N. M. J. Am. Chem. Soc. 2008, 130,<br />

518 15332–15339.<br />

519 (11) (a) Mayrhofer, K. J. J.; Blizanac, B. B.; Arenz, M.; Stamenkovic,<br />

520 V. R.; Ross, P. N.; Markovic, N. M. J. Phys. Chem. B 2005,<br />

521 109, 14433–14440. (b) Wang, C.; van der Vilet, D.; Chang, K. C.;<br />

522 You, H. D.; Strmcnik, D.; Schlueter, J. A.; Markovic, N. M.; Stamenkovic,<br />

523 V. R. J. Phys. Chem. C 2009, 113, 19365–19368.<br />

524 (12) (a) Komanicky, V.; Chang, K. C.; Menzel, A.; Markovic, N. M.;<br />

525 You, H.; Wang, X.; Myers, D. J. Electrochem. Soc. 2006, 153, B446–B451.<br />

526 (b) Borup, R.; et al. Chem. Rev. 2007, 107, 3904–3951.<br />

527 (13) Wang, C.; Chi, M.; Wang, G.; van der Vliet, D.; Li, D.; More, K.;<br />

528 Wang, H.-H.; Schlueter, J. A.; Markovic, N. M.; Stamenkovic, V. R. Adv.<br />

529 Funct. Mater. 2011, 21, 147–152.<br />

530 (14) Beno, M. A.; Engbretson, M.; Jennings, G.; Knapp, G. S.;<br />

531 Linton, J.; Kurtz, C.; Rutt, U.; Montano, P. A. Nucl. Instrum. Methods<br />

532 Phys. Res., Sect A 2001, 467, 699–702.<br />

533 (15) Newville, M.; Livins, P.; Yacoby, Y.; Rehr, J. J.; Stern, E. A. Phys.<br />

534 Rev. B 1993, 47, 14126–14131.<br />

535 (16) (a) Newville, M. J. Synchrotron Radiat. 2001, 8, 322–324.<br />

536 (b) Ravel, B.; Newville, M. J. Synchrotron Radiat. 2005, 12, 537–541.<br />

(17) (a) Newman, J. J. Electrochem. Soc. 1966, 113, 2. (b) van der Vliet,<br />

D.; Strmcnik, D. S.; Wang, C.; Stamenkovic, V. R.; Markovic, N. M.;<br />

Koper, M. T. M. J. Electroanal. Chem. 2010, 647, 29–34.<br />

(18) Ahrenstorf, K.; Heller, H.; Kornowski, A.; Broekaert, J. A. C.;<br />

Weller, H. Adv. Funct. Mater. 2008, 18, 3850–3856.<br />

(19) Kiely, C. J.; Fink, J.; Brust, M.; Bethell, D.; Schiffrin, D. J. Nature<br />

1998, 396, 444–446.<br />

(20) Li, Y.; Zhang, X. L.; Qiu, R.; Qiao, R.; Kang, Y. S. J. Phys. Chem.<br />

C 2007, 111, 10747–10750.<br />

(21) Bunker, G. Introduction to XAFS: A Practical Guide to X-ray<br />

Absorption Fine Structure Spectroscopy; Cambridge University Press:<br />

Cambridge, UK; New York, 2010; viii, 260 pp.<br />

(22) (a) Muker<strong>je</strong>e, S.; Srinivasan, S.; Soriaga, M. P.; Mcbreen, J.<br />

J. Electrochem. Soc. 1995, 142, 1409–1422. (b) Rodriguez, J. A.; Hanson,<br />

J. C.; Frenkel, A. I.; Kim, J. Y.; Perez, M. J. Am. Chem. Soc. 2002,<br />

124, 346–354.<br />

(23) (a) Imai, H.; Izumi, K.; Matsumoto, M.; Kubo, Y.; Kato, K.;<br />

Imai, Y. J. Am. Chem. Soc. 2009, 131, 6293–6300. (b) Friebel, D.; Miller,<br />

D. J.; O’Grady, C. P.; Anniyev, T.; Bargar, J.; Bergmann, U.; Ogasawara,<br />

H.; Wikfeldt, K. T.; Pettersson, L. G. M.; Nilsson, A. Phys. Chem. Chem.<br />

Phys. 2011, 13, 262–266. (c) Russell, A. E.; Rose, A. Chem. Rev. 2004,<br />

104, 4613–4635.<br />

H dx.doi.org/10.1021/ja2047655 |J. Am. Chem. Soc. XXXX, XXX, 000–000<br />

537<br />

538<br />

539<br />

540<br />

541<br />

542<br />

543<br />

544<br />

545<br />

546<br />

547<br />

548<br />

549<br />

550<br />

551<br />

552<br />

553<br />

554<br />

555<br />

556<br />

557<br />

558


Origin of Anomalous Activities for Electrocatalysts in Alkaline<br />

Electrolytes<br />

Ram Subbaraman, † N. Danilovic, † P. P. Lopes, †,‡ D. Tripkovic, † D. Strmcnik, † V. R. Stamenkovic, †<br />

and N. M. Markovic* ,†<br />

† Materials Science Division, Argonne National Laboratory, Lemont, Illinois 60439 United States<br />

‡ Instituto de Química de Saõ Carlos/USP, C.P.780, CEP 13560-970, Saõ Carlos, SP, Brazil<br />

*S Supporting Information<br />

ABSTRACT: Pt extended surfaces and nanoparticle electrodes<br />

are used to understand the origin of anomalous activities<br />

for electrocatalytic reactions in alkaline electrolytes as a<br />

function of cycling/time. Scanning tunneling microscopy<br />

(STM) of the surfaces before and after cycling in alkaline<br />

electrolytes was used to understand the morphology of the<br />

impurities and their impact on the catalytic sites. The nature of<br />

the contaminant species is identified as 3d-transition metal<br />

cations, and the formation of hydr(oxy)oxides of these<br />

elements is established as the main reason for the observed<br />

behavior. We find that, while for the oxygen reduction reaction<br />

(ORR) and the hydrogen oxidation reaction (HOR) the blocking of the sites by the undesired 3d-transition metal<br />

hydr(oxy)oxide species leads to deactivation of the reaction activities, the CO oxidation reaction and the hydrogen evolution<br />

reaction (HER) can have beneficial effects from the same impurities, the latter being dependent on the exact nature of the<br />

adsorbing species. These results show the significance of impurities present in real electrolytes and their impact on<br />

electrocatalysis.<br />

■ INTRODUCTION<br />

Alkaline energy conversion devices (alkaline fuel cells, AFCs)<br />

and fuel generation devices (alkaline electrolyzers, AEs) have<br />

garnered significant attention lately due to the applicability of<br />

low-cost materials as catalysts in comparison to their acid<br />

counterparts. For example, while in PEM fuel cells Pt-based<br />

catalysts are required for efficient transformation of chemical<br />

energy of hydrogen and oxygen into electricity, 1,2 in alkaline<br />

electrolytes efficient energy conversion in AFCs can be<br />

achieved with much cheaper nonprecious materials, including<br />

Ag and abundantly available transition metal elements. 2−4 A<br />

major limiting factor thus has been the development of robust<br />

alkaline anion-exchange membranes, for which some significant<br />

progress is being made. 5,6 The same is true for the fuel<br />

production reactions, where the hydrogen evolution reaction<br />

(HER) and the oxygen evolution reaction (OER) can<br />

efficiently be carried out on metal surfaces modified with<br />

inexpensive 3d transition metal (3d-TM) oxides, especially for<br />

the HER or on 3d-TM compounds, such as oxides and<br />

sulfides. 7−9 In addition to such “material factors” (e.g., strong<br />

covalent metal−adsorbate interactions), the reactivity of<br />

interfaces in alkaline environments is also governed via rather<br />

weak noncovalent interactions between hydrated cations<br />

(located in the double layer at ca. 3.5 Å 10 (ORR),<br />

) and covalently<br />

bound oxygenated species. This issue has been recently<br />

discussed especially for understanding the role of the<br />

noncovalent interactions in the oxygen reduction reaction<br />

11 the oxidation of small organic molecules, 11,12 the<br />

hydrogen oxidation reaction (HOR), 13,11 and the HER. 13,14<br />

The influence of noncovalent reactions on kinetic rates can be<br />

either catalytic (as in a case of the HER and CO oxidation<br />

reaction) or inhibiting (e.g., ORR, HOR, and the oxidation of<br />

methanol), a fact that significantly contributes to the richness of<br />

methods that can be used to control the activity of<br />

electrochemical interfaces in alkaline environments.<br />

A central complication involved with experimentally<br />

exploring the role of covalent and noncovalent interactions in<br />

alkaline solutions, however, is that the measured rates are<br />

influenced substantially (or even predominantly) by the<br />

presence of varying levels of impurities. The puzzling role of<br />

impurities in alkaline electrochemical measurements has been a<br />

topic of interest for over 25 years. Conway 15−17 addressed<br />

some of the issues employing recrystallization protocols to<br />

purify electrolytes to remove some transition elements, while<br />

Spendelow et al. 18 demonstrated the use of sacrificial adsorbers ,<br />

in particular for Fe. However, the impact of the impurity on the<br />

electrocatalytic activities, particularly for O2 and H2 reactions,<br />

was seldom studied. Recently, effort has been focused on<br />

identifying the role of impurities from the reaction of the glass<br />

components of the electrochemical cell and how they affect the<br />

Received: July 31, 2012<br />

Revised: September 5, 2012<br />

Published: September 19, 2012<br />

Article<br />

pubs.acs.org/JPCC<br />

© 2012 American Chemical Society 22231 dx.doi.org/10.1021/jp3075783 | J. Phys. Chem. C 2012, 116, 22231−22237


The Journal of Physical Chemistry C Article<br />

electrochemical behavior of polycrystalline Pt electrodes. 19<br />

While the use of non-glass-based cells have addressed some of<br />

the inconsistencies, it has been found that depending on the<br />

prehistory of electrolyte preparation, temperature of electrolyte,<br />

duration of experiment, scan rates, and applied potential<br />

windows, adsorption properties and the reaction activities may<br />

vary substantially even if experiments were performed in<br />

fluoroethylenepolymer-based electrochemical cells. Consequently,<br />

unraveling both the true nature of these impurities<br />

as well as how they may affect the key electrochemical<br />

processes in alkaline solutions should be of paramount<br />

importance for the development of technologies that can<br />

efficiently transform and store energy at electrochemical<br />

interfaces at high pH values.<br />

Here, in order to probe the impurities effect on the surface<br />

electrochemistry in alkaline electrolytes, we employ the use of<br />

well-defined metal single-crystal surfaces in a glass-free cell<br />

(made with fluoropolymer). The use of single crystal surfaces<br />

for “detecting” impurities in electrochemical environments is a<br />

well-established tactic, as has been demonstrated where they<br />

are used for monitoring the effects of trace levels of halides,<br />

nitrates, and other anions on adsorption and catalytic properties<br />

of metal−electrolyte interfaces in acidic media. 1 Combining<br />

scanning tunneling microscopy (STM) with traditional electrochemical<br />

methods, applied to the Pt(111) surface, we identify<br />

the signature for the presence of the impurities and establish<br />

the role of these impurities present in the alkaline electrolyte on<br />

both the adsorption and reaction properties of the electrode for<br />

the various reactions. Furthermore, intentionally adding<br />

compounds of the 3d elements (Ni, Co, Cr, Fe), we further<br />

establish the nature of the common impurities present in the<br />

alkaline electrolyte as that of the 3d-TMs. Lastly, it is also<br />

shown that the formation of 3d-TM hydr(oxy)oxides is mainly<br />

responsible for the anomalous behavior observed in alkaline<br />

electrocatalytic measurements, and depending on the nature of<br />

these elements, we can observe beneficial or poisoning effects<br />

for<br />

■<br />

the various reactions.<br />

EXPERIMENTAL SECTION<br />

A standard three-electrode cell made from fluoroethylenepolymer<br />

(FEP) was used for the experiments. The chemical<br />

solutions were prepared from KOH obtained from different<br />

sources (GFS, Sigma Aldrich, Fluka, and Alfa Aesar) and Milli-<br />

Q deionized (DI) water. The perchlorate salts of Ni, Co, Fe,<br />

Mn, and Cr were all purchased from Sigma Aldrich at the<br />

highest purity levels available and were made into 0.01 M<br />

solution with the DI water. Small volumes of this solution were<br />

added to the electrochemical cell to study the effect of the<br />

concentrations. Six millimeter disk electrodes were used for all<br />

experiments. The electrodes were prepared by radio frequency<br />

(RF) annealing at ∼1100 °C in a 3% H2−Ar gas mixture for 7<br />

min. The samples were transferred into the electrochemical cell<br />

with the surface protected with a drop of DI water and<br />

immersed under potential control at 0.05 V vs reversible<br />

hydrogen electrode (RHE). The Pt/C catalyst obtained from<br />

Tanaka (TKK) was mixed with water in the concentration of 1<br />

mg/mL. This dispersion was then ultrasonically mixed for 1 h,<br />

following which a stable suspension was obtained. A glassy<br />

carbon disk (6 mm diameter) was then mechanically polished.<br />

Known volumes of the suspensions were then added using a<br />

micro pipet onto the glassy carbon disk electrode. The<br />

electrode was dried at 60 °C in an inert atmosphere. The<br />

suspension was applied so that it coats the surface of the<br />

22232<br />

electrode very uniformly. Once dry, these electrodes were<br />

washed with water to ensure/verify the good adhesion of<br />

particles to the glassy carbon substrate, after which they were<br />

introduced into the alkaline cell. During the alkaline ORR,<br />

HOR measurements, in order to ensure that particles had not<br />

been dislodged during the experiments, the underpotentially<br />

deposited hydrogen based charge was measured both before<br />

and after in 0.1 M HClO4. Given that the 3d-TM hydr(oxy)oxides<br />

are very soluble in acid solutions, we were able to obtain<br />

accurate measures of the surface area.<br />

A standard rotating disk electrode (RDE) setup with Ag/<br />

AgCl reference (−0.96 V vs RHE), was used for electrochemical<br />

measurements. All the results reported in the<br />

manuscript are versus the RHE. A Pt counter electrode was<br />

used for all the experiments. The sweep rates used in the cyclic<br />

voltammetry (CV) experiments were 50 mV s−1 , while the<br />

rotation rate was either 1600 or 2500 rpm. Typical experiments<br />

were conducted at this sweep rate. We also tried slower sweep<br />

rates, which showed similar behavior at much smaller number<br />

of cycles and are not reported here. Most experiments were<br />

performed at room temperature (RT) except for those<br />

mentioned in the Results and Discussion section conducted<br />

at 60 °C. For HOR and HER experiments, the potential was<br />

swept in the cathodic direction from the hold potential; the<br />

data presented is taken from first sweep curves. First scans were<br />

used for deriving baselines for clean conditions. The first scans<br />

are always obtained within 1−2 min of introduction of the<br />

electrode into the electrolyte under potential control. This was<br />

helpful to protect the electrode from electrolyte impurities.<br />

Ohmic resistaances 20 were corrected for all the data reported<br />

here using the Autolab PGSTAT 302N potentiostat. Electrolyte<br />

resistance was also measured with AC impedance spectroscopy.<br />

The gases used were research grade (5N) Ar and H2.CO cylinders were purchased from Airgas at research plus grade<br />

(aluminum container, to avoid Fe contamination due to<br />

stainless<br />

■<br />

steel materials).<br />

RESULTS AND DISCUSSION<br />

We begin by presenting the CVs obtained for Pt(111) under<br />

three different experimental conditions (Figure 1a): the first<br />

and 25th sweep in “clean” 0.1 M KOH at RT, the 15th sweep in<br />

1 M KOH, and the 10th sweep for 1 M KOH at 60 °C. The CV<br />

recorded on Pt(111) (see corresponding STM image) during<br />

the first sweep shows all the expected characteristics of a wellordered<br />

surface in “clean” KOH environments: reversible<br />

adsorption of hydrogen (Hupd: 0.05 < E < 0.4), formation of the<br />

double layer between 0.4 and 0.6 V, and adsorption of hydroxyl<br />

species (OHad: 0.6 < E < 0.95 V). However, after 25 cycles<br />

within the same potential region, three distinctive CV features<br />

are observed: (i) the Hupd potential region is suppressed; (ii)<br />

OHad formation starts at more negative potential; and (iii)<br />

initial adsorption of the OHad becomes highly irreversible.<br />

Similar distortions are observed for the other two cases<br />

considered here as well. In the past, such variation in the OH<br />

region has often been associated with an irreversible oxide<br />

formed on the Pt surface with unknown stoichiometries. 21,22<br />

Importantly, Figure 1b reveals that the degree of Hupd and<br />

OHad distortion is strongly dependent on both the concentration<br />

as well as temperature of the electrolyte. For example, in<br />

1 M KOH, only 15 cycles were sufficient to produce a similar<br />

degree of changes as previously observed after 25 potential<br />

cycles in dilute electrolyte solutions. The temperature of the<br />

electrolyte plays an even bigger role, i.e., at 60 °C, merely 10<br />

dx.doi.org/10.1021/jp3075783 | J. Phys. Chem. C 2012, 116, 22231−22237


The Journal of Physical Chemistry C Article<br />

Figure 1. Comparison of STM and cyclic voltammograms for Pt(111)<br />

electrode. (a) STM image of as-prepared Pt(111) surface after the first<br />

sweep at RT. The CV for this surface is also shown. Also shown are<br />

the CVs after 25 sweeps in 0.1 M KOH at RT, 10 sweeps in 0.1 M<br />

KOH at 60 °C, and 15 sweeps in 1 M KOH at RT. CVs exhibit<br />

distortion of adsorption properties of both H upd and OH ad. Note the<br />

scan window was shortened for the 1 M KOH in the H upd region due<br />

to reference electrode correction. (b) STM image of Pt(111) electrode<br />

cycled in 0.1 M KOH with 25 ppm of Co 2+ cations. Also shown are the<br />

CVs after 25 sweeps in 0.1 M KOH in solutions containing different<br />

cations such as Ni 2+ ,Co 2+ , and Cr 2+ . Fluka KOH electrolyte, ultra high<br />

purity, was used for these spiking measurements (c) STM image of the<br />

cycled electrode (Figure 1b) after cycling in CO saturated solution.<br />

Also shown are the polarization curves for the CO oxidation reaction<br />

for the fresh as-prepared electrode as well as the consecutive cycles of<br />

the “contaminated” electrode. CV at the end of “cleaning” protocol is<br />

shown in Figure S3. We note in passing that, holding the contaminated<br />

electrode in CO-saturated solution at 0.05 V does not show any<br />

22233<br />

Figure 1. continued<br />

change in the surface coverage, indicative of the absence of a “simple<br />

CO displacement” mechanism. Enhanced CO oxidation activities are<br />

observed in the presence of 3d-TM hydr(oxy)oxides on the surface.<br />

potential cycles were sufficient to completely transform the<br />

perfect to rather anomalous pseudocapacitive features of H upd<br />

and OH ad. Interestingly, with the exception of the CV, which is<br />

recorded during the very first potential sweep, the anomalous<br />

shapes display a close resemblance to current−potential curves<br />

recorded on Pt(111) modified with 3d-TM hydr(oxy)oxide<br />

clusters. 23 The chemical nature of such species as a function of<br />

potential was also established previously for different 3dmetals.<br />

23 In turn, this was the first indication that the observed<br />

changes might be due to the adsorption of 3d-TM elements,<br />

which are known to be present in alkaline electrolytes at ppm−<br />

ppb levels even in the most clean alkaline salts.<br />

Here, the existence of trace levels (0.5−20 μg/L depending<br />

on the supplier and batch of the electrolyte salt; averages for all<br />

the cations and their levels are shown in Table S1, Supporting<br />

Information) of 3d-TM elements in dilute alkaline electrolytes<br />

was confirmed by using inductively coupled plasma mass<br />

spectrometry (ICPMS). To further confirm that the 3d-TM<br />

impurities are the main contributors to the observed anomalous<br />

behavior of Pt(111) in Figure 1a, we intentionally “spiked” the<br />

electrolyte with small concentrations (25 ppm) of 3d-TM<br />

cations. As in ref 23, CV/STM results are used to establish the<br />

correlation between the coverage of 3d-TM hydr(oxy)oxide<br />

clusters and their effects on H upd and OH ad formation.<br />

Inspection of Figure 1b indicates several features of interest<br />

in this regard. First, the STM results for the Pt(111)/Co 2+<br />

system (used as a representative of the 3d-TMs consider in this<br />

work) reveals that after potential cycling in solution containing<br />

ppm-levels of Co 2+ , nanoclusters (ca. 2 atomic layers thick) are<br />

uniformly distributed across the (111) terrace. Second, the<br />

adsorption of OH ad (position of irreversible peaks) follows the<br />

trends in oxophilicity of the 3d elements, consistent with what<br />

was reported previously. 23 To further probe the nature of<br />

clusters, we use the CO oxidation reaction, which is known to<br />

be strongly dependent on a nature and surface coverage by<br />

hydroxyl species. A key observation in Figure 1c is that the<br />

most active surface is covered with relatively high coverage of<br />

Co-hydr(oxy)oxide clusters (see STM image in Figure 1b and<br />

the polarization curve in Figure 1c). Moreover, there is STM<br />

evidence that deactivation in the subsequent sweeps is closely<br />

related to the disappearance of the very same clusters (Figure<br />

1c); as evident from the 15th sweep recorded in CO saturated<br />

solution. It is therefore plausible that, during the CO oxidation,<br />

OH species associated with the 3d-TM clusters are consumed<br />

and, as a result, the remaining Co species become<br />

thermodynamically unstable and dissolve into the solution.<br />

We notice in passing, after ∼30 potential cycles, that Co(OH) 2<br />

clusters are barely present on the surface (see Figure S3). The<br />

underlying CO oxidation mechanism can also be extended to<br />

Pt−Co bimetallic systems; e.g., rather than affecting segregation<br />

of Pt to the surface in Pt−Co bimetallic alloys, 24 a key role of<br />

CO is simply to “clean” the Pt−Co surface of the Cohydr(oxy)oxides<br />

species, which are formed when non-noble Co<br />

metal atoms are exposed to alkaline electrolytes and positive<br />

potentials. We conclude, therefore, that the anomalous behavior<br />

of CVs in the “butterfly” potential region (0.6 < E < 0.9 V)<br />

appears to be due to the presence of 3d-TM hydr(oxy)oxides<br />

dx.doi.org/10.1021/jp3075783 | J. Phys. Chem. C 2012, 116, 22231−22237


The Journal of Physical Chemistry C Article<br />

Figure 2. Polarization curves corresponding to the ORR and HOR on Pt surfaces. (a) Effect of cycling on the ORR and HOR activities of the<br />

Pt(111) surface in 0.1 M KOH as a function of cycling at room temperature. Activities are found to decrease with cycling, consistent with<br />

contamination of the surface by the adsorption of 3d-TM hydr(oxy)oxides. (b) Comparison of the ORR activity stability at 60 °C for Pt-poly<br />

surfaces. Pt(111) surface exhibits significant deactivation within few cycles at such high temperatures. Inset shows the voltammetric features’<br />

evolution with cycling for the Pt-poly electrode at RT. Pt-poly exhibits deactivation of ∼200 mV for the half-wave potential with cycling at high<br />

temperature, which is higher, compared to ∼120 mV at 30 °C.<br />

rather than the formation of Pt-oxide with yet unknown<br />

stoichiometry, as has been considered previously in the<br />

literature. 21,22 Given that these experiments were conducted<br />

in a glass-free cell, the main source of contamination is the<br />

electrolyte itself, and the role of glass components is secondary<br />

in nature, contrary to what was thought previously. 19,25 We<br />

would like to point out that the CO oxidation was studied here<br />

primarily as a probe reaction, and our results indicate that CO<br />

oxidation reaction is the least affected experiment with cycling<br />

in alkaline solutions. In fact, CO oxidation or cycling appears to<br />

be the possible protocol for cleaning the electrodes in alkaline<br />

solutions.<br />

It is important to note that the time required for the<br />

complete distortion of the CVs is also a function of<br />

hydrodynamic conditions such as stirring of electrolyte, i.e.,<br />

the number of CVs required for complete distortion in “clean”<br />

0.1 M KOH is less than ∼12 cycles. A similar effect is also<br />

observed with decreasing the sweep rates and/or increasing the<br />

potential hold times before experiments. The use of rotating<br />

disk electrode (RDE) enhances the mass transport of both<br />

reactants (H 2,O 2,H 2O) and the 3d-TM impurities present in<br />

the bulk of the electrolyte. As we demonstrate further below,<br />

interplay between these two mass-transport-dependent processes<br />

will, in turn, govern the rate HOR/HER and ORR. In<br />

general, in alkaline solutions, the rate of electrochemical<br />

reaction (current density i) at a constant electrode potential<br />

(E) can be governed by the following rate expression: 11<br />

i = nFK c [1 − Θ − Θ ] exp<br />

@E1 1 r cov noncov<br />

*<br />

− Δ<br />

(<br />

G<br />

)<br />

kT<br />

where n is the number of electrons, K 1 is a constant that<br />

incorporates all constant variables and the rate constants for a<br />

particular reaction, F is Faraday’s constant, c r is the<br />

concentration of reactant species in a solution (H 2,O 2 and<br />

H 2O), Θ cov and Θ noncov represent the fraction of the surface<br />

masked by site-blocking covalently and noncovalently bound<br />

“spectator” species. The exponential ΔG* term (ΔG = ΔG* cov<br />

+ ΔG* noncov) corresponds to the standard Gibb’s free energy<br />

change required to form products from the reactants and the<br />

active intermediates.<br />

(1)<br />

22234<br />

In the following, we use this rate expression to discuss the<br />

puzzling cycling-/temperature-induced activity variations in<br />

alkaline solutions at constant rotation rates. As shown in Figure<br />

2a, the activities for the HOR and the ORR are found to<br />

decrease with potential cycling. Notice that the difference in<br />

activity between the first and tenth potential cycles is massive,<br />

consistent with 3d-TM impurity species blocking the sites<br />

required for H 2 and O 2 adsorption; thus affecting the reaction<br />

through a (1 − Θ cov − Θ noncov) term. The deactivation for the<br />

reactions is large enough to prevent achieving the mass<br />

transport limited currents in some cases for both the HOR and<br />

the ORR.<br />

Because the activities of the ORR and the HOR on Pt(111)<br />

at high temperatures change dramatically (even ∼3−4 cycles<br />

exhibit complete deactivation), the temperature effects will be<br />

presented only for the Pt-poly electrodes, which we found to be<br />

less sensitive toward poisoning effects by the impurities in the<br />

electrolyte. The behavior exhibited by the Pt-poly surface at 60<br />

°C is shown in Figure 2b. The half wave potential (potential at<br />

half the maximum current) is found to shift toward higher<br />

overpotentials much faster at higher temperatures: after 50<br />

cycles, ∼100 mV at 30 °C compared to ∼200 mV at 60 °C.<br />

Thus, from the high-temperature ORR experiments in Figure<br />

2b, we conclude that one should be very careful in deriving<br />

activation energies from Arrhenius plots in alkaline solutions.<br />

With the exception of the CO oxidation reaction, the main<br />

contribution of the 3d-TM impurities on surface activity<br />

between 0.05 to 1.0 V (the HOR and ORR) appears to be via a<br />

(1 − Θ cov − Θ noncov) term. So far, the role of the impurities has<br />

been found to be generally independent of the chemical nature<br />

of the elements (except for the CO oxidation reaction). Also,<br />

the potential regions of interest, particularly for the ORR and<br />

the HOR, are CV visible and hence can be correlated with the<br />

observed CV behaviors of the cycled electrodes.<br />

However, the situation below 0.05 V is rather different, as<br />

revealed in Figure 3 for the HER in electrolytes with intrinsic<br />

differences in the content and nature of transition metal<br />

impurities. In particular, while in the case of one electrolyte<br />

with a higher content of Co and Ni, the HER activity is found<br />

to enhanced after extensive potential cycling, under the same<br />

experimental conditions the HER is substantially deactivated in<br />

dx.doi.org/10.1021/jp3075783 | J. Phys. Chem. C 2012, 116, 22231−22237


The Journal of Physical Chemistry C Article<br />

Figure 3. Polarization curves comparing the HER activities for the<br />

Pt(111) surface cycled in two different batches of 0.1 M KOH. The<br />

first one is rich in Fe 2+ and as shown, the activity for the HER<br />

decreases with cycling, and the second electrolyte was found to have a<br />

higher content of Co 2+ compared to other TM cations and was found<br />

to exhibit higher activities for the HER. This explains the onset of the<br />

anomalous activities observed for the HER in alkaline solutions.<br />

the other electrolyte, which had a higher concentration of Fe.<br />

Such behavior is not uncommon, and has been discussed in<br />

literature, for example, Petrii and Tsirilna, 26 as well as others<br />

who have reported the deactivation of the HER activities as was<br />

compiled by Trasatti. 9 Nevertheless, in our experiments, we<br />

found that solution used in the first experiment is rich in the<br />

concentrations of Co and Ni impurities, while in the second<br />

experiment with higher concentrations of Fe. As was shown<br />

recently, 23 the nature of the transition element and hence its<br />

hydr(oxy)oxide on the surface plays a significant role in<br />

determining the activity of the surface for the HER. It was<br />

shown that the presence of less oxophilic hydroxides such as<br />

Co(OH) 2 and Ni(OH) 2 on the surface of the Pt can optimize<br />

the turnover frequency (TOF) of the water dissociation step<br />

and indeed to catalyze the HER. 14,23 On the other hand, the<br />

TOF is attenuated on highly oxophilic Fe or Mn hydr(oxy)oxides.<br />

It must be noted that while these two cases presented<br />

here exhibit qualitatively distinct features, it is often difficult to<br />

predict the trends in the HER with cycling. This is often related<br />

to the variations in the relative concentrations of the cations.<br />

Therefore, significant precautions are necessary while measuring<br />

the HER activities in alkaline solutions. For fundamental<br />

studies, in order to avoid the contributions of the impurities,<br />

activities for the HER must be either derived from the very first<br />

potential scan, or the electrolyte purity must be significantly<br />

enhanced. For practical applications, however, stable activities<br />

are seldom established (in fact, it is customary to see<br />

deactivation of the HER), signaling that alkaline electrolytes<br />

are predominantly contaminated with undesired Fe-/Mn-type<br />

impurities.<br />

So far, we have demonstrated using the Pt extended surfaces,<br />

that it is indeed possible to understand the origin of anomalous<br />

activities in alkaline electrolytes. The relatively high sensitivity<br />

of these surfaces aids in the detection and characterization of<br />

the electrochemical signatures for the adsorbed species<br />

responsible. In order to apply these principles and to<br />

demonstrate that the behavior exhibited by such surfaces is<br />

completely transformational across various length scales, we<br />

compare the results for cycling of Pt nanoparticles (see<br />

Experimental Section for preparation) in alkaline electrolytes.<br />

Figure 4a,b shows the effect of cycling on both formation of<br />

adsorbates as well as the role of CO oxidation reaction based<br />

“cleaning” on recovering the original voltammogram. Unlike<br />

the case of extended surfaces, the surface contamination<br />

process is slower, due to the lower sensitivity and lack of welldefined<br />

adsorption sites; however, within the time frame of a<br />

reasonable long-term experiment (e.g., 100 cycles), the features<br />

similar to those in Figure 1 appear: (i) suppression of the H upd<br />

and (ii) irreversible peaks in the oxide (hydroxide) region (E ><br />

0.6 V). On sub<strong>je</strong>cting this surface to the CO-cycling protocol<br />

discussed briefly in Figure 1c, we once again find that the<br />

surface adsorbed species are removed, restoring the original<br />

voltammogram. In order to confirm the difference between<br />

oxidation of the CO molecules through the hydroxyl groups on<br />

these 3d-hydr(oxy)oxides versus CO displacement of such<br />

species, we have shown the CO stripping data for the<br />

Figure 4. (a) Comparison of pristine Pt/C electrode and the one that was cycled for 100 cycles between 0.05 and 1.0 V at 50 mV/s. Slower scan<br />

rates (20 mV/s) were also tried, and similar behavior was obtained in fewer than 100 cycles. Also shown is the CV for the “cleaned” electrode, which<br />

was obtained after cycling in CO for 30 cycles at 50 mV/s. The surface essentially resembles that of the pristine electrode with small signatures for<br />

the contamination. (b) CO stripping curves for both pristine electrode and the cycled electrode recorded at 20 mV/s (corresponding CV in 4a). The<br />

cycled electrode clearly exhibits a prepeak at 0.45 V. The reduction curve exhibits the signature for the 3d-hydr(oxy)oxides present on the surface,<br />

indicating that CO does not displace these species from the surface, and the removal involves reaction of the OH groups present on these oxides<br />

with the adsorbed CO molecules. CV for the pristine electrode is also shown.<br />

22235<br />

dx.doi.org/10.1021/jp3075783 | J. Phys. Chem. C 2012, 116, 22231−22237


The Journal of Physical Chemistry C Article<br />

nanoparticles in Figure 4b. Also shown is the comparison of<br />

CO stripping for “clean” (uncycled) Pt nanoparticle electrode.<br />

If the effect of addition of CO was to purely displace the oxide<br />

species, one would not expect any difference between the<br />

“clean” and the cycled electrodes for CO stripping. The<br />

existence of a prepeak in this region is a clear indication that<br />

CO oxidation proceeds through reaction with hydroxyl species<br />

adsorbed on the adsorbates on the surface. Furthermore, the<br />

preservation of the 3d-hydr(oxy)oxide peak after CO-stripping<br />

voltammetry (at 0.5 V in Figure 4b) is another clear indication<br />

that CO cleaning does not operate via a simple displacement of<br />

adsorbed species on the surface. This procedure also provides<br />

insight into the so-called CO-annealing in alkaline environments<br />

24 where the surface segregation of Pt in the Pt-M alloy<br />

nanoparticles was assumed to lead to the formation of Pt-rich<br />

surfaces. Given the higher oxophilicity of 3d elements as well as<br />

the similarity in the voltammetric features exhibited by these<br />

alloying elements to the ones in Figure 4a, for example, it is<br />

plausible that these alloying elements are removed from the Ptalloy<br />

nanoparticles through destabilization by CO molecules.<br />

The qualitative behavior of these nanocatalysts for the ORR,<br />

HOR, and the HER mimic that of the extended surfaces, where<br />

the former reactions are affected by a third-body effect of<br />

poisoning, and the latter depends on the nature of contaminant<br />

species present in the electrolyte.<br />

In conclusion, the anomalous adsorption and catalytic<br />

behavior in alkaline environments arise from the presence of<br />

3d-TM impurities in the electrolyte. While using the FEP-based<br />

cells helps alleviate the problems that arise from reaction of the<br />

electrolyte with glass components of electrochemical cells, the<br />

intrinsic transition metal impurities present in the electrolyte<br />

play a bigger role in determining the electrochemical<br />

characteristics of metal electrodes. The use of Pt(111) as a<br />

probe was found to be very advantageous in determining the<br />

nature of 3d-TM impurities even at very low levels. We found<br />

that the ORR and HOR in KOH electrolytes are affected by a<br />

third-body effect where, irrespective of the nature of the 3d-TM<br />

elements, impurities simply act as spectators and block the<br />

surface sites (the 1 − Θ term). On the other hand, there are<br />

also desirable type of 3d-TM impurities, as in the case of<br />

Co(OH) 2/Ni(OH) 2 acting catalytically (the ΔG*) on the CO<br />

oxidation reaction and the HER. Lastly, this behavior is not<br />

restricted to Pt extended surfaces, and is found to affect the Pt<br />

nanoparticles system as well, albeit to a different degree. Taken<br />

together, our results demonstrated that extensive care must be<br />

taken while determining the adsorption and catalytic properties<br />

of metal surfaces in alkaline electrolytes, further necessitating<br />

the development of strategies to purify the alkaline electrolytes<br />

for a “real system”, and/or careful design of experimental<br />

protocols.<br />

■ ASSOCIATED CONTENT<br />

*S Supporting Information<br />

ICPMS results for various alkaline electrolytes, cleanliness of<br />

ultra-high purity electrolytes, the use of first scans for “control”<br />

for the alkaline measurements, and CO oxidation-induced<br />

cleaning of Pt(111) surfaces. This material is available free of<br />

charge via the Internet at http://pubs.acs.org.<br />

■ AUTHOR INFORMATION<br />

Corresponding Author<br />

*E-mail: nmmarkovic@anl.gov.<br />

22236<br />

Author Contributions<br />

R.S. and N.M. designed the experiments and wrote the<br />

manuscript. R.S., N.D., P.P.L., and D.S. did the electrochemical<br />

experiments. R.S. and D.T. prepared the samples and<br />

performed the STM measurements. R.S., N.M., and V.S.<br />

discussed the results. R.S. and N.M. prepared the manuscript.<br />

Notes<br />

The authors declare no competing financial interest.<br />

■ ACKNOWLEDGMENTS<br />

This work was supported by the Office of Science, Office of<br />

Basic Energy Sciences, Division of Materials Science, U.S.<br />

Department of Energy, under contract DE-AC02-06CH11357.<br />

N.D. would like to thank the Chemical Sciences and<br />

Engineering Division at Argonne National Laboratory for<br />

funding. P.P.L. would like to thanks CAPES and FAPESP for<br />

financial support.<br />

■ REFERENCES<br />

(1) Marković, N. M.; Ross, P. N. Surf. Sci. Rep. 2002, 45, 117.<br />

(2) Vielstich, W.; Lamm, A.; Gasteiger, H. A. Handbook of Fuel Cells:<br />

Fundamentals, Technology, and Applications; John Wiley & Sons: New<br />

York, 2003.<br />

(3) McLean, G. F.; Niet, T.; Prince-Richard, S.; Djilali, N. Int. J.<br />

Hydrogen Energy 2002, 27, 507.<br />

(4) Spendelow, J. S.; Wieckowski, A. Phys. Chem. Chem. Phys. 2007,<br />

9, 2654.<br />

(5) Pivovar, B.; Fedkiw, P.; Mantz; Alkaline Membrane Fuel Cell<br />

Workshop Final Report, Phoenix, AZ, 2006;113; (http://www1.eere.<br />

energy.gov/hydrogenandfuelcells/pdfs/amfc_dec2006_workshop_<br />

report.pdf).<br />

(6) Merle, G.; Wessling, M.; Nijmei<strong>je</strong>r, K. J. Membr. Sci. 2011, 377, 1.<br />

(7) Lasia, A. In Handbook of Fuel Cells: Fundamentals, Technology and<br />

Applications; Vielstich, W., Lamm, A., Gasteiger, H. A., Eds.; Wiley:<br />

New York, 2003; Vol. 2, p 416.<br />

(8) Pletcher, D.; Walsh, F. Industrial Electrochemistry; Springer: New<br />

York, 1990.<br />

(9) Trasatti, S.; Transition Metal Oxides: Versatile Materials for<br />

Electrocatalysis, In Electrochemistry of novel materials. Lipkowski, J.,<br />

Ross, P. N., Eds.; VCH: New York, 1994.<br />

(10) Strmcnik, D.; van der Vliet, D. F.; Chang, K. C.; Komanicky, V.;<br />

Kodama, K.; You, H.; Stamenkovic, V. R.; Marković, N.M.J. Phys.<br />

Chem. Lett. 2012, 2, 2733.<br />

(11) Strmcnik, D.; Kodama, K.; Van der Vliet, D.; Greeley, J.;<br />

Stamenkovic, V. R.; Marković, N.M.Nat. Chem. 2009, 1, 466.<br />

(12) Sitta, E.; Batista, B. C.; Varela, H. Chem. Commun. 2011, 47,<br />

3775.<br />

(13) Danilovic, N.; Subbaraman, R.; Strmcnik, D.; Paulikas, A.;<br />

Myers, D.; Stamenkovic, V.; Markovic, N. Electrocatalysis 2012, 1−9.<br />

(14) Subbaraman, R.; Tripkovic, D.; Strmcnik, D.; Chang, K. C.;<br />

Uchimura, M.; Paulikas, A. P.; Stamenkovic, V.; Markovic, N. M.<br />

Science 2011, 334, 1256.<br />

(15) Angerstein-Kozlowska, H.; Conway, B. E.; Barnett, B.; Mozota,<br />

J. J. Electroanal. Chem. Interfacial Electrochem. 1979, 100, 417.<br />

(16) Conway, B. E.; Tilak, B. V. Electrochim. Acta 2002, 47, 3571.<br />

(17) Conway, B. E.; Barber, J.; Morin, S. Electrochim. Acta 1998, 44,<br />

1109.<br />

(18) Spendelow, J. S.; Goodpaster, J. D.; Kenis, P. J. A.; Wieckowski,<br />

A. J. Phys. Chem. B 2006, 110, 9545.<br />

(19) Mayrhofer, K. J. J.; Crampton, A. S.; Wiberg, G. K. H.; Arenz,<br />

M. J. Electrochem. Soc. 2008, 155, P78.<br />

(20) van der Vliet, D.; Strmcnik, D. S.; Wang, C.; Stamenkovic, V. R.;<br />

Markovic, N. M.; Koper, M. J. Electroanal. Chem. 2010, 647, 29.<br />

(21) Markovic, N. M.; Schmidt, T. J.; Grgur, B. N.; Gasteiger, H. A.;<br />

Behm, R. J.; Ross, P. N. J. Phys. Chem. B 1999, 103, 8568.<br />

(22) Tripković, A. V.; Popović, K. D.; Lović, J.D.J. Serb. Chem. Soc.<br />

2001, 66, 825.<br />

dx.doi.org/10.1021/jp3075783 | J. Phys. Chem. C 2012, 116, 22231−22237


The Journal of Physical Chemistry C Article<br />

(23) Subbaraman, R.; Tripkovic, D.; Chang, K.-C.; Strmcnik, D.;<br />

Paulikas, A. P.; Hirunsit, P.; Chan, M.; Greeley, J.; Stamenkovic, V.;<br />

Markovic, N. M. Nat. Mater. 2012, 11, 550.<br />

(24) Mayrhofer, K. J. J.; Juhart, V.; Hartl, K.; Hanzlik, M.; Arenz, M.<br />

Angew. Chem., Int. Ed. 2009, 48, 3529.<br />

(25) Mayrhofer, K. J. J.; Wiberg, G. K. H.; Arenz, M. J. Electrochem.<br />

Soc. 2008, 155, P1.<br />

(26) Petrii, O. A.; Tsirlina, G. A. Electrochim. Acta 1994, 39, 1739.<br />

22237<br />

dx.doi.org/10.1021/jp3075783 | J. Phys. Chem. C 2012, 116, 22231−22237


ACCEPTED MANUSCRIPT<br />

This is an early electronic version of an as-received manuscript that has been<br />

accepted for publication in the Journal of the Serbian Chemical Society but has<br />

not yet been sub<strong>je</strong>cted to the editing process and publishing procedure applied by<br />

the JSCS Editorial Office.<br />

Please cite this article as: J. D. Lović, D. V. <strong>Tripković</strong>, K. Đ. Popović, V. M.<br />

Jovanović, A. V. <strong>Tripković</strong>, J. Serb. Chem. Soc. (2012), doi:<br />

10.2298/JSC121012138L<br />

This “raw” version of the manuscript is being provided to the authors and readers<br />

for their technical service. It must be stressed that the manuscript still has to be<br />

sub<strong>je</strong>cted to copyediting, typesetting, English grammar and syntax corrections,<br />

professional editing and authors’ review of the galley proof before it is published<br />

in its final form. Please note that during these publishing processes, many errors<br />

may emerge which could affect the final content of the manuscript and all legal<br />

disclaimers applied according to the policies of the Journal.


J. Serb. Chem. Soc. 77 (0) 1–18 (2012) UDC<br />

JSCS–5508 Original scientific paper<br />

Electrocatalytic properties of Pt–Bi electrodes towards the<br />

electrooxidation of formic acid<br />

JELENA D. LOVIĆ # , DUŠAN V. TRIPKOVIĆ # , KSENIJA Đ. POPOVIĆ # , VLADISLAVA<br />

M. JOVANOVIĆ # and AMALIJA V. TRIPKOVIĆ* #<br />

ICTM – Institute of Electrochemistry, University of Belgrade, N<strong>je</strong>goševa 12, P.O.Box 473,<br />

11000 Belgrade, Serbia<br />

(Received 12 October, revised 16 November 2012)<br />

Abstract: Formic acid oxidation was studied on two Pt-Bi catalysts, Pt 2Bi and<br />

polycrystalline Pt modified by irreversible adsorbed Bi (Pt/Bi irr) in order to<br />

establish the difference between the effects of Bi irr and Bi in alloyed state. The<br />

results were compared to pure Pt. It was found that both bimetallic catalysts<br />

were more active than Pt with the onset potentials shifted to more negative<br />

values and the currents at 0.0 V vs. SCE (under steady state conditions)<br />

improved up to two order of magnitude. The origin of Pt 2Bi high activity and<br />

stability is increased selectivity toward formic acid dehydrogenation caused by<br />

the ensemble and electronic effect and suppression of Bi leaching from the<br />

surface during formic acid oxidation. However, although Pt/Bi irr also shows<br />

remarkable initial activity compared to pure Pt, dissolution of Bi is not<br />

suppressed and the poisoning of the electrode surface induced by dehydration<br />

path is observed. Comparison of the initial quasi-steady state and<br />

potentiodynamic results obtained for these two Pt-Bi catalysts revealed that the<br />

electronic effect, existing only in the alloy, contributes earlier start of the<br />

reaction, while the maximum current density is determined by the ensemble<br />

effect.<br />

Keywords: formic acid; electrochemical oxidation; Pt 2Bi catalyst; Pt/Bi irr<br />

catalyst; fuel cell.<br />

INTRODUCTION<br />

Accepted Manuscript<br />

The electrocatalytic oxidation of small organic molecules, such as methanol,<br />

ethanol and formic acid has been extensively studied because such molecules can<br />

potentially be used as fuel in fuel cell applications. 1 Pt is an excellent catalyst for<br />

dehydrogenation of small organic molecules but, on the other hand, has several<br />

significant disadvantages: high cost and extreme susceptibility to poisoning by<br />

CO. To improve its catalyst performance and minimize its quantity in the<br />

* Corresponding author. E-mail: amalija@tmf.bg.ac.rs<br />

# Serbian Chemical Society member.<br />

doi: 10.2298/JSC121012138L<br />

1


2 FORMIC ACID OXIDATION ON Pt–Bi ELECTRODES<br />

catalyst, Pt was modified by the addition of other metals such as Ru, Rh, Pd, Sn,<br />

Pb, etc. 2–5<br />

Electrochemical oxidation of formic acid has been comprehensively<br />

investigated as the anodic reaction in direct formic acid fuel cell (DFAFC) as<br />

well as a model reaction important for understanding electrooxidation of other<br />

small organic molecules. 6,7 Mechanistic studies of the formic acid<br />

electrooxidation put forward that the reaction on Pt proceeds through two parallel<br />

paths: 8 dehydrogenation with direct oxidation to CO2 and the path consisting of a<br />

dehydration step, to yield water and adsorbed CO (COad) as a poisoning<br />

intermediate, and the subsequent oxidation of COad to CO2. Since the potential of<br />

COad formation is lower than the dehydrogenation potential, the poison remains<br />

on electrode surfaces until it is oxidized by oxygen containing species at positive<br />

potentials. 9 Thus, Pt atoms covered with poison are not available for<br />

dehydrogenation and the catalytic efficiency of the Pt surface decreases.<br />

Therefore, practical application of Pt for formic acid oxidation requires some<br />

modification of the surface. This is commonly done by alloying or by alteration<br />

of the Pt surface with adsorbed foreign metals in the amount less than a full<br />

monolayer. As the COad poison needs an ensemble of Pt surface atoms to adsorb<br />

on, modification of the local distribution of surface domains is achieved by<br />

adsorption of different adatoms, such as Bi. Irreversibly adsorbed Bi. 10–13 inhibits<br />

poison formation simultaneously enhancing dehydrogenation, 13 i.e. this<br />

modification is an efficient way to hinder the dehydration path (CO-intermediate<br />

pathway) in favor of direct path. 14 This increased selectivity for dehydrogenation<br />

has been proposed as an “ensemble effect” 15,16 in which the adsorbed Bi divides<br />

the Pt surface into small domains where only dehydrogenation can occur. A<br />

correlation between ensemble size and formic acid oxidation activity were also<br />

established. 17 According to literature data the activity of Pt catalyst modified with<br />

Bi depends on the shape of Pt nanocrystals, 18 vary with the size of particles 19 and<br />

Pt catalyst loading. 14<br />

Ordered intermetallic PtBi or PtBi2 alloys 20–23 and PtBi alloy nanoparticles 24–<br />

29 were proposed as good catalysts for formic acid oxidation. The onset potential<br />

of the reaction is significantly shifted to more negative values (by over 300 mV)<br />

and the current density is remarkably enhanced in whole potential range<br />

Accepted Manuscript<br />

compared to Pt. 30 Moreover, the PtBi surface appears to have a considerably<br />

lower sensitivity to poisoning by CO according to DEMS, 22,30 FTIR 22,31 and DFT<br />

calculations. 22,23,32 The origin of its catalytic activity was related to electronic<br />

effects. The formation of the PtBi ordered intermetallic phase results in a charge<br />

redistribution enhancing the affinity of PtBi for formic acid adsorption and<br />

producing surface oxides at low potentials, 25,33 as well as to geometric effects<br />

reducing the affinity for CO poisoning. In addition, the excellent catalytic


QUALITY OF MOLECULAR STRUCTURE DESCRIPTORS 3<br />

properties of PtBi nanoparticles can be attributed to the occurrence of the<br />

ensemble effect at the nanoscale level. 29<br />

In our previous studies, the oxidation of formic acid was investigated on<br />

single phase PtBi alloy, 34 as well as on Pt2Bi catalyst, a two-phase material<br />

consisting of PtBi alloy and pure Pt. 35 A huge increase in catalytic activity on<br />

PtBi electrode relative to polycrystalline Pt was observed and discussed in terms<br />

of electrochemically detected UPD phenomena of Bi readsorbed on Pt.<br />

Additionally, on the basis of the analysis of X-ray photoelectron spectra, it was<br />

proposed that bifunctional action of hydroxylated Bi species may contribute to<br />

the enhanced activity of PtBi alloys. 34 Pt2Bi, is found to be powerful catalyst for<br />

formic acid oxidation exhibiting high activity and stability. High activity<br />

originates from the fact that formic acid oxidation proceeds completely through<br />

dehydrogenation path, while the high stability of Pt2Bi surface is induced by the<br />

suppression of Bi leaching in the presence of formic acid. 35 It is difficult to<br />

separate the contributions of the ensemble and the electronic effect since both of<br />

them can be present in the Pt–Bi surfaces.<br />

In this work formic acid oxidation was studied on two types Pt-Bi surfaces:<br />

polycrystalline Pt modified by irreversible adsorbed Bi (designated as Pt/Biirr)<br />

and on Pt2Bi catalyst. In order to investigate the promotional role of Bi in formic<br />

acid oxidation, as well as the difference between the effect of irreversibly<br />

adsorbed Bi and Bi in alloyed state, Pt was modified with ~ 30% Biirr which<br />

correspond to the nominal content of Bi in Pt2Bi catalyst. Comparative<br />

investigation based on the effects influencing the catalytic properties of these<br />

electrodes enables better understanding of different activities between Pt2Bi<br />

catalyst and Biirr modified Pt.<br />

EXPERIMENTAL<br />

Polycrystalline Pt and Pt2Bi electrodes in the form of disc were used in this study. Pt2Bi catalyst was prepared and characterized at Institute of Catalysis and Surface Chemistry, Polish<br />

Academy of Sciences, Krakow, Poland. 35 Briefly, the catalyst was fabricated by melting of the<br />

pure elements in inert atmosphere in the proportion of Bi to Pt = 1:2 and characterized by Xray<br />

diffraction (XRD). Diffraction pattern for Pt2Bi sample reveals two crystal phases:<br />

platinum (fcc) and platinum bismuth PtBi (hcp). The phase composition of the sample was<br />

calculated using the Rietveld refinement as 45% and 55 % for Pt and PtBi phases,<br />

respectively.<br />

Accepted Manuscript<br />

Prior to each experiment, the electrodes were mirror polished (1–0.05 µm Buehler<br />

alumina). The surfaces were rinsed with high purity water (Millipore, 18 M� cm resistivity),<br />

sonicated for 2-3 min and rinsed again with ultrapure water.<br />

All experiments were carried out on as-prepared catalysts. Three-compartment<br />

electrochemical glass cells with Pt wire as the counter electrode and saturated calomel<br />

electrode (SCE) as the reference electrode was used. All the potentials are expressed on the<br />

scale of SCE. The electrolyte containing 0.1 M H 2SO 4 as a supporting electrolyte and 0.125 M<br />

HCOOH was prepared with high purity water and p.a. grade chemicals (Merck). The<br />

electrolyte was deaerated by bubbling of nitrogen. Upon addition of HCOOH at -0.2 V,


4 FORMIC ACID OXIDATION ON Pt–Bi ELECTRODES<br />

potentiodynamic (� = 50 mV s -1 ), quasi steady-state measurements (� = 1 mV s -1 ) or<br />

chronoamperometric measurements were carried out. The electrode was rotating at 2000 rpm<br />

in all the experiments.<br />

Modification of the Pt electrode was performed from 1 x 10 -5 M Bi 2O 3 in 0.1 M H 2SO 4 at<br />

open circuit potential during 45 s. After modification, the electrode was rinsed with water and<br />

transferred into a cell containing supporting electrolyte. Fraction of sites covered by Bi<br />

(denoted Pt/Bi irr) was estimated by decrease in the charge for desorption of hydrogen,<br />

assuming a charge of 210 �C cm −2 for the hydrogen monolayer adsorption.<br />

The real surface area of all as-prepared catalysts was calculated from CO stripping<br />

voltammetry. For the CO stripping measurements, pure CO was bubbled through the<br />

electrolyte for 20 min while keeping the electrode potential at -0.2 V vs. SCE. After purging<br />

the electrolyte by N 2 for 30 min to eliminate the dissolved CO, the adsorbed CO was oxidized<br />

in an anodic scan at 50 mV s −1 . Two subsequent voltammograms were recorded to verify the<br />

completeness of the CO oxidation. Real surface area of all electrodes used was estimated by<br />

the calculation of the charge from CO ad stripping voltammograms corrected for background<br />

currents. Assuming charge of 420 μC cm −2 for the CO monolayer adsorption on the Pt<br />

electrode, roughness factor of 1.4 � 0.1 was estimated. The real surface area of Pt 2Bi electrode<br />

was estimated assuming the same roughness factor as for Pt electrode, which is to be expected<br />

since both electrodes were polished in the same way. The specific activity of Pt and Pt 2Bi<br />

electrodes for formic acid oxidation are normalized using these values of the surface area.<br />

The experiments were conducted at 295 ± 0.5 K. A VoltaLab PGZ 402 (Radiometer<br />

Analytical, Lyon, France) was employed.<br />

Electrochemical characterization<br />

RESULTS AND DISCUSSION<br />

Initial voltammograms of as-prepared Pt and Pt/Biirr catalysts are presented<br />

in Fig. 1. Cyclic voltammogram for polycrystalline Pt electrode is characterized<br />

by defined region of hydrogen adsorption/desorption (E < 0.05 V), separated by<br />

double layer from the region of surface oxide formation (E > 0.45 V). The<br />

absence of well developed peaks at Pt polycrystalline in hydrogen<br />

adsorption/desorption region is caused by preparation procedure used.<br />

Figure 1 compares typical voltammograms recorded in supporting<br />

electrolyte before and after Pt modification with adsorbed bismuth. Distinctive<br />

characteristics for the presence of bismuth ad-atoms on the platinum surface are<br />

the diminution of the hydrogen adsorption/desorption charge due to the fact that<br />

hydrogen does not adsorb on Bi 36 and the appearance of peaks a and a’, as a<br />

Accepted Manuscript<br />

result of the oxidation/reduction of Bi irreversibly adsorbed onto the Pt surface,<br />

which superimpose to those corresponding to Pt oxide formation/reduction. Since<br />

only the Pt sites, unblocked by Bi are available for hydrogen adsorption, the<br />

fractional coverage by Bi ad-atoms, evaluated from the decrease in charge<br />

involved in the hydrogen desorption before and after adsorption of Bi on the Pt<br />

electrode surface, was set to be about 30%, corresponding the nominal content of<br />

Bi in Pt2Bi catalyst.


Oxidation of pre-adsorbed CO<br />

QUALITY OF MOLECULAR STRUCTURE DESCRIPTORS 5<br />

Oxidation of pre-adsorbed CO is examined on Pt, Pt2Bi and Pt/Biirr<br />

electrodes and the stripping voltammograms after the subtraction of the<br />

background current are given in Fig. 2. As one can see, the oxidation of COad<br />

starts earlier on Pt2Bi catalyst than on pure Pt. The difference in onset and peak<br />

potential of COad oxidation could be ascribed to the some electronic modification<br />

of Pt surface atoms by Bi 10,17,37,38 resulting in weaker bonding of COad. This<br />

statement is consistent with literature data based on theoretical calculations. 39,40<br />

However, the difference in onset and peak potentials between Pt/Biirr and Pt<br />

catalysts is insignificant.<br />

The CO stripping charge was also determined for Pt2Bi electrode and<br />

corrected for the background currents to eliminate the contribution of the double<br />

layer charge, as well as Bi oxidation charge. Since Bi 37 Error! Bookmark not<br />

defined. and PtBi 23,30 are inactive for CO adsorption, the oxidation of CO occurs<br />

only on Pt domains. Therefore, the charge under the COad peak at Pt2Bi reflects a<br />

process at Pt parts and can be used for determining the contribution of pure Pt in<br />

the surface composition of Pt2Bi catalyst. 35<br />

Oxidation of formic acid<br />

Potentiodynamic measurements<br />

Activity of catalysts. The activities of Pt2Bi, Pt/Biirr and Pt electrodes (first<br />

sweeps) towards formic acid oxidation are compared in Fig. 3. Cyclic<br />

voltammogram for Pt electrode shows well establish feature for the formic acid<br />

oxidation. 8 In the positive scan current slowly increases reaching a plateau at ~<br />

0.25 V followed by ascending current starting at 0.5 V which attains a maximum<br />

at ~ 0.62 V. The explanation for this behavior could be described considering<br />

dual path mechanism, i.e. dehydrogenation assigned as the direct path, based on<br />

the oxidation of formate, 41,42 and dehydration, indirect path, assumes formation of<br />

COad, both generates CO2 as the final reaction product. At low potentials<br />

HCOOH oxidizes through the direct path with the simultaneous formation of<br />

COad. Increasing coverage with COad reduces the Pt sites available for the direct<br />

path and current slowly increases reaching a plateau. Subsequent formation of<br />

oxygen-containing species on Pt enables the oxidative removal of COad, more Pt<br />

Accepted Manuscript<br />

sites become available for HCOOH oxidation and current increases until Pt<br />

oxide, inactive for HCOOH oxidation, is formed,which results in the current peak<br />

at ~0.62 V. In the backward scan the sharp increase of HCOOH oxidation current<br />

coincides with reduction of Pt oxide. The currents are much higher than in the<br />

forward sweep, because Pt surface is freed of COad.<br />

The polarization curves for bimetallic surfaces indicate quite different<br />

behavior. Figure 3 shows that the onset potential for the reaction on Pt/Biirr is<br />

about 0.1 V less positive than on Pt electrode. The current raises up to 0.35V and


6 FORMIC ACID OXIDATION ON Pt–Bi ELECTRODES<br />

reaches a peak which corresponds to the oxidation of HCOOH to CO2 in the<br />

direct path, occurring on Pt sites that are not blocked by COad. On the descending<br />

part of the curve the shoulder appears almost at the same potential as the peak on<br />

bare Pt electrode. The currents recorded in the backward direction are slightly<br />

higher, so the difference between forward and backward scan is not as large as on<br />

bare Pt. Since the peak at ~ 0.62 V arises from the HCOOH oxidation on the Pt<br />

sites being released by COad oxidation, its height is an indication of the degree of<br />

the Pt poisoning at lower potentials. Accordingly, the amount of COad formed in<br />

the indirect path on Pt/Biirr is much lower than on pure Pt electrode.<br />

The oxidation of formic acid on Pt2Bi electrode starts at ~ -0.2 V, which is<br />

0.1 V less positive than on Pt/Biirr and ~0.2 V less positive than on Pt electrode.<br />

The current raises up to 0.35 V and reaches a peak about 15 times higher than the<br />

plateau on Pt. At more positive potentials reaction currents decrease due to<br />

surface oxide formation. The absence of the shoulder on the desceding part of the<br />

curve, recorded on the Pt/Biirr, is indicating no poisoning of the alloy by COads.<br />

Bell-shaped voltammogram clearly suggests that oxidation of HCOOH on<br />

Pt2Bi proceeds through dehydrogenation path. 35 Error! Bookmark not defined.<br />

Since Bi does not adsorb HCOOH 34,40 oxidation of HCOOH occurs on Pt sites in<br />

pure Pt and Bi alloyed Pt domains. As it was shown in Fig. 3 lower activity and<br />

appearance of shoulder on Pt/Biirr compared to Pt2Bi could suggest a low<br />

coverage by COad at Pt sites on Pt/Biirr, i.e. incomplete suppression of the<br />

dehydration path although there is almost the same (nominal) amount of Bi on<br />

these two electrodes. Thus, HCOOH oxidation on Pt/Biirr proceeds<br />

predominantly by dehydrogenation path with some minor degree of<br />

dehydratation that takes place as well. Increased selectivity toward<br />

dehydrogenation path on Pt2Bi as well as on Pt/Biirr compared to Pt is caused<br />

mainly by an ensemble effect originating from diminishing of continuous Pt sites<br />

(capable for dehydration) by Bi. However, the ensemble effect on Pt/Biirr catalyst<br />

is enabled by the adsorbed Bi with practically no influence on the neighboring<br />

free Pt atoms. When Pt2Bi alloy is in question, the ensembles are created by<br />

alloyed Bi atoms incorporated in Pt lattice causing the shift in d-band center of<br />

adjacent Pt atoms. Therefore, Bi in the alloy also exhibits electronic effect that<br />

can be the reason for better performance of this catalyst resulting in higher<br />

Accepted Manuscript<br />

currents and lower onset potential.<br />

Stability of catalysts. Cyclic voltammograms (first and 20 th sweeps) recorded<br />

on Pt2Bi and Pt/Biirr catalysts in formic acid containing solution, are shown in<br />

Fig. 4. Over the potential cycling up to 0.8 V the activity of Pt2Bi electrode (Fig.<br />

4a) slowly decreases during first 5-7 cycles reaching the value of 90% and 82%<br />

of the initial currents at the potential of the maximum current and at the potential<br />

of 0.0 V in the low current region, respectively. After these first few sweeps the<br />

currents remain unchanged with further cycling (Inset in Fig. 4a). Since cycling


QUALITY OF MOLECULAR STRUCTURE DESCRIPTORS 7<br />

of Pt2Bi in supporting electrolyte leads to enhancement of currents related to<br />

oxidation of Bi species indicating some surface decomposition caused by Bi<br />

leaching/dissolution process, 35 the stability of Pt2Bi during oxidation of formic<br />

acid could be induced by the presence of HCOOH in the electrolyte. Abruna and<br />

co-workers 43 explained that the stability of Pt-Bi intermetallic surface originates<br />

from the competition between oxidation of formic acid at the electrode/solution<br />

interface and Bi leaching, i.e. corrosion/oxidation processes of the electrode<br />

surface itself. Accordingly, the main reason for high stability of formic acid<br />

oxidation current on Pt2Bi catalyst is the inhibition of dehydration path, as well<br />

as, suppression of Bi leaching. This statement was confirmed by STM imaging<br />

before and after electrochemical treatment in formic acid containing solution,<br />

which did not indicate any significant change of surface morphology and<br />

roughness after that procedure. 35<br />

Contrary to Pt2Bi alloy, as can be seen in Fig. 4(b), Pt/Biirr electrode shows<br />

significant changes with continuous cycling in the solution containing formic<br />

acid (Inset in Fig. 4b). Repetitive cycling up to 0.8 V shifts the onset potential for<br />

formic acid oxidation to more positive values, decreases the reaction currents,<br />

while anodic peak diminishes and a new peak starts to emerge and grow at ~ 0.6<br />

V. This transformation of cyclic voltammograms indicates continuous Bi<br />

dissolution and modification of the surface composition. Upon prolonged<br />

cycling, electrode surface becomes enriched in platinum and exhibits a Pt-like<br />

electrochemical behavior in acid electrolyte containing formic acid. Apparently<br />

re-adsorption of Bi species from the solution is rather low, so the initial<br />

voltammogram was never restored, which is in accordance with results obtained<br />

for formic acid oxidation on bismuth-coated mesoporous Pt microelectrodes. 44<br />

Finally, it can be concluded that initial activity at two Bi-Pt catalysts<br />

originates from the ensemble effect, but the activity of the Pt/Biirr in the reaction<br />

is diminished due to leaching of Bi. An electronic effect contributes to the early<br />

start of the reaction on alloy.<br />

Chronoamperometric measurements<br />

Chronoamperometric experiments were performed to prove activity and<br />

stability of the catalysts investigated. The current density of formic acid<br />

Accepted Manuscript<br />

oxidation was recorded as a function of time at 0.2V over 30 min (Fig. 5). The<br />

highest initial current density at 0.2 V on Pt2Bi related to the other two catalysts<br />

is in accordance with the potentiodynamic measurements (Fig. 3). The initial<br />

currents at Pt2Bi electrode decreases slightly and in observed period of time<br />

attain a value which is about 3.5 times higher than on Pt/Biirr catalyst and almost<br />

20 times higher than at Pt electrode. Chronoamperometic test confirmed high<br />

activity and stability of the Pt2Bi catalyst.


8 FORMIC ACID OXIDATION ON Pt–Bi ELECTRODES<br />

Quasi-steady state measurements<br />

The quasi-steady state measurements for formic acid oxidation at all<br />

investigated electrodes are presented in Fig. 6. The data obtained under the slow<br />

sweep conditions corroborated the difference in the activities of pure Pt, Pt<br />

modified by Bi and Pt2Bi alloy that was found under the potentiodynamic<br />

measurements.<br />

Tafel slope on Pt2Bi electrode is about 120 mV dec -1 , indicating that the first<br />

electron transfer is the rate-determining step, with transfer coefficient being<br />

about 0.5. This means that C–H bond cleavage, to form COOHad, is the slow step<br />

and determines the rate of formic acid oxidation on Pt2Bi electrodes.<br />

The Tafel slope of about 150 mV dec -1 obtained during formic acid oxidation<br />

on bare Pt indicate that formed CO was adsorbed and collected on the surface<br />

slowing down the reaction rate. The same value of Tafel slope was obtained for<br />

formic acid oxidation on carbon supported high surface area platinum. 45 Tafel<br />

slope of 135 mV dec −1 was found at Pt/Biirr and imply moderate surface coverage<br />

by COad.<br />

Comparing the activities of electrodes investigated at 0.0 V it can be seen<br />

that the current densities are enhanced 15 times at Pt/Biirr and up two orders of<br />

magnitude at Pt2Bi catalyst in respect to Pt electrode.<br />

CONCLUSION<br />

Pt2Bi and Pt/Biirr electrodes were investigated in formic acid oxidation and<br />

the results were compared to Pt electrode. The results presented indicate that Bi<br />

in alloy and irreversibly adsorbed Bi exhibit different effects on the catalytic<br />

activity.<br />

Pt2Bi is highly active for the formic acid oxidation because the<br />

dehydrogenation path is predominant in the overall reaction. Bi in the alloy not<br />

only that facilitates ensemble effect but also has electronic effect that can be the<br />

reason for better performance of this catalyst resulting in higher currents and<br />

lower onset potential. The main reason for high stability of Pt2Bi catalyst is the<br />

inhibition of dehydration path in the reaction, as well as suppression of Bi<br />

leaching indicated by insignificant change of surface morphology and roughness.<br />

On the contrary, Pt/Biirr electrode is not stabilized by formic acid oxidation,<br />

Accepted Manuscript<br />

since the desorption of Bi is not suppressed in the presence of formic acid, as<br />

well as poisoning effect induced by dehydration path. Since Biirr do not provoke<br />

any significant modification of electronic environment, modified Pt catalysts is<br />

less active than corresponding alloy.<br />

Comparing the results obtained for these two Pt-Bi catalysts, the role of the<br />

ensemble effect and electronic effect in the oxidation of formic acid on Pt-Bi<br />

electrodes can be distinguished. The electronic effect, existing only on the alloy,<br />

contributes earlier start of the reaction, while the maximum current density is


QUALITY OF MOLECULAR STRUCTURE DESCRIPTORS 9<br />

determined by the ensemble effect. During potential cycling of Pt/Biirr electrode<br />

Bi is leached form the electrode surface and the ensemble effect is reduced over<br />

time, or lost. Insight into the chronoamperometric curves confirms an advantage<br />

of alloys, i.e. necessity of alloying Pt with Bi to obtain a corrosion stable catalyst.<br />

Acknowledgement. This work was financially supported by the Ministry of Education<br />

and Science, Republic of Serbia, Contract No. H-172060.<br />

И З В О Д<br />

ЕЛЕКТРОКАТАЛИТИЧКЕ ОСОБИНЕ Pt–Bi ЕЛЕКТРОДА У ОКСИДАЦИЈИ<br />

МРАВЉЕ КИСЕЛИНЕ<br />

ЈЕЛЕНА Д. ЛОВИЋ, ДУШАН В. ТРИПКОВИЋ, КСЕНИЈА Ђ. ПОПОВИЋ, ВЛАДИСЛАВА М. ЈОВАНОВИЋ<br />

и АМАЛИЈА В. ТРИПКОВИЋ<br />

ИХТМ – Центар за електрохемију, Универзитет у Београду,Његошева 12, п. пр. 473, 11000 Београд<br />

Оксидација мравље киселине испитивана је на два типа Pt–Bi катализатора: Pt2Bi<br />

електроди и на поликристалној Pt електроди модификованој иреверзибилно<br />

адсорбованим Bi (Pt/Biirr). Активности су упређене са резултатима добијеним на чистој<br />

поликристалној Pt електроди. Циљ је био да се објасни разлика у деловању<br />

иреверзибилно адсорбованог Bi (Biirr) и Bi у легираном стању. Показано је да су оба<br />

биметална катализатора активнија од поликристалне Pt, почетак реакције је померен ка<br />

негативнијим вредностима и у поређењу са чистом Pt при стационарним условима<br />

добијене су до два реда величине веће густине струје. Разлог за велику активност и<br />

стабилност Pt2Bi електроде у оксидацији мравље киселине је одигравање реакције по<br />

главном реакционом путу (дехидроганација мравље киселине), што је изазвано ефектом<br />

трећег тела и електронским ефектом, као и спречавање излуживања Bi из елецтроде. С<br />

друге стране, иако Pt/Biirr показује значајну почетну активност у односу на Pt, ова<br />

електрода није стабилна током реакције оксидације HCOOH због континуалног<br />

растварања Bi са површине елецтроде, као и тровања површине изазваног током<br />

реакције по индиректном, дехидратационом путу. Поређењем резултата добијених на<br />

ове две Pt-Bi електроде може се објаснити улога ефекта трећег тела и електронског<br />

ефекта у оксидацији HCOOH. Наиме, електронски ефекат, који постоји само код легуре,<br />

доприноси ранијем почетку реакције, док је максимална струја одређена ефектом<br />

трећег тела. Током циклизирања Pt/Biirr електроде Bi одлази са површине и ефекат<br />

трећег тела се губи током времана. Хроноамперометријска метења указују на предност<br />

легуре, односно неопходност легирања Bi са Pt да би се добио корозионо стабилан<br />

катализатор.<br />

(Примљено 12. октобра, ревидирано 16 новембра 2012)<br />

Accepted Manuscript<br />

REFERENCES<br />

R. Parsons, T. VanderNoot, J. Electroanal. Chem., 257 (1988) 9.<br />

A.V. <strong>Tripković</strong>, J.D. Lović, K.DJ. Popović, J.Serb.Chem.Soc., 75 (2010) 1559.<br />

J.D. Lović, K.DJ. Popović, A.V. <strong>Tripković</strong>, J.Serb.Chem.Soc., 76 (2011) 1523.<br />

S. Stevanović, D. <strong>Tripković</strong>, D. Poleti, J. Rogan, A. <strong>Tripković</strong>, V. M. Jovanović,<br />

J.Serb.Chem.Soc., 76 (2011) 1673.<br />

X. Xia, T. Iwasita, J. Electrochem. Soc., 140 (1993) 2559.<br />

X. Wang, J-M. Hu, I-M. Hsing, J. Electroanal. Chem., 562 (2004) 73.<br />

A.V.<strong>Tripković</strong>, K.Đ.Popović, J.D.Lović, J.Serb.Chem.Soc., 68 (2003) 849.


10 FORMIC ACID OXIDATION ON Pt–Bi ELECTRODES<br />

A. Capon, R. Parsons, J. Electroanal. Chem., 44 (1973) 1.<br />

J.M. Feliu, E. Herrero, in: W. Vielstich, A. Lamm, H.A. Gasteiger (Eds.), Handbook of Fuel<br />

Cells-Fundamentals Technology and Applications, Vol. 2, John Wiley & Sons Ltd, New<br />

York, 2003, Ch. 42.<br />

E. Herrero, A. Fernandez-Vega, J.M. Feliu, A. Aldaz, J. Electroanal. Chem., 350 (1993) 73.<br />

J. Clavilier, A. Fernandez-Vega, J.M. Feliu, A. Aldaz, J. Electroanal. Chem., 258 (1989) 89.<br />

J. Clavilier, A. Fernandez-Vega, J.M. Feliu, A. Aldaz, J. Electroanal. Chem., 261 (1989) 113.<br />

B.-J. Kim, K. Kwon, C. K. Rhee, J. Han, T.-H. Lim, Electrochim. Acta, 53 (2008) 7744.<br />

A. Sáez, E. Expósito, J. Solla-Gullón, V. Montiel, A. Aldaz, Electrochim. Acta, 63 (2012)<br />

105.<br />

A. Lopez-Cudero, F. J. Vidal-Iglesias, J. Solla-Gullon, E. Herrero, A. Aldaz, J. M. Feliu,<br />

Phys. Chem. Chem. Phys., 11 (2009) 416.<br />

J. Clavilier, J. M. Feliu, A. Aldaz, J. Electroanal. Chem., 243 (1988) 419.<br />

J. Kim, C. K. Rhee, Electrochem. Comm., 12 (2010) 1731.<br />

Q.-S. Chen, Z.-Y. Zhou, F. J. Vidal-Iglesias, J. Solla-Gullon, J. M. Feliu, S.-G. Sun, J. Am.<br />

Chem. Soc., 133 (2011) 12930.<br />

C. Jung, T. Zhang, B.-J. Kim, J. Kim, C. K. Rhee, T.-H. Lim, Bull. Korean Chem. Soc., 31<br />

(2010) 1543.<br />

D. Volpe, E. Casado-Rivera, L. Alden, C. Lind, K. Hagerdon, C. Downie, C. Korzniewski, F.<br />

J. DiSalvo, H. D. Abruna, J. Electrochem. Soc., 151 (2004) A971.<br />

E. Casado-Rivera, D. J. Volpe, L. Alden, C. Lind, C. Downie, T. Vazquez-Alvarez, A. C. D.<br />

Angelo, F. J. DiSalvo, H. D. Abruna, J. Am. Chem. Soc., 126 (2004) 4043.<br />

H. Wang, L. Alden, F. J. DiSalvo, H. D. Abruna, Phys. Chem. Chem. Phys., 10 (2008) 3739.<br />

M. Oana, R. Hoffmann, H..D. Abruna, F..J. DiSalvo, Surf. Sci., 574 (2005) 1.<br />

C. Roychowdhury, F. Matsumoto, V. B. Zeldovich, S. C. Warren, P. F. Mutolo, M. J.<br />

Ballesteros, U. Wiesner, H. D. Abruna, F. J. DiSalvo, Chem. Mater., 18 (2006) 3365.<br />

Y. Liu, M. A. Lowe, F. J. DiSalvo, H. D. Abruna, J. Phys. Chem. C, 114 (2010) 14929.<br />

L. M. Magno, W. Sigle, P. A. van Aken, D. G. Angelescu, C. Stubenrauch, Chem. Mater., 22<br />

(2010) 6263.<br />

X. Yu, P.G. Pickup, Electrochim. Acta, 56 (2011) 4037.<br />

Y. Liu, M. A. Lowe, K. D. Finkelstein, D. S. Dale, F. J. DiSalvo, H. D. Abruna, Chem. A -<br />

Eur. J., 16 (2010) 13689.<br />

X. Ji, K. T. Lee, R. Holden, L. Zhang, J. Zhang, G. A. Botton, M. Couillard, L. F. Nazar, Nat.<br />

Chem., 2 (2010) 286.<br />

E. Casado-Rivera, Z. Gal, A.C.D. Angelo, C. Lind, F.J. DiSalvo, H.D. Abruna,<br />

Chem.Phys.Chem., 4 (2003) 193.<br />

N. de-los-Santos-Alvarez, L. R. Alden, E. Rus, H. Wang, F. J. DiSalvo, H. D. Abruna, J.<br />

Electroanal. Chem., 626 (2009) 14.<br />

Accepted Manuscript<br />

I. Pašti, S. Mentus, Phys. Chem. Chem. Phys., 11 (2009) 6225.<br />

D. R. Blasini, D. Rochefort, E. Fachini, L. R. Alden, F. J. DiSalvo, C. R. Cabrera, H. D.<br />

Abruna, Surf. Sci., 600 (2006) 2670.<br />

A. V. <strong>Tripković</strong>, K. Dj. Popović, R. M. Stevanović, R. Socha, A. Kowal, Electrochem.<br />

Comm., 8 (2006) 1492.<br />

J. D. Lović, M. D. Obradović, D. V. <strong>Tripković</strong>, K. Dj. Popović, V. M. Jovanović, S. Lj.<br />

Gojković, A. V. <strong>Tripković</strong>, Electrocatal., DOI 10.1007/s12678-012-0099-9.<br />

R. Gomez, J. M. Feliu, A. Aldaz, Electrochim. Acta, 42 (1997) 1675.


QUALITY OF MOLECULAR STRUCTURE DESCRIPTORS 11<br />

T.J. Schmidt, B.N. Grgur, R.J. Behm, N.M. Markovic, P.N. Ross, Jr., Phys. Chem. Chem.<br />

Phys., 2 (2000) 4379.<br />

B. E. Hayden, A. J. Murray, R. Parsons, D. J. Pegg, J. Electroanal. Chem., 409 (1996) 51.<br />

L.-L. Wang, D. D. Johnson, J. Phys. Chem. C, 112 (2008) 8266.<br />

N. Kapur, B. Shan, J. Hyun, L. Wang, S. Yang, J.B. Nicholas, K. Cho, Molecular Simulation,<br />

37 (2011) 648.<br />

A. Miki, S. Ye, M. Osawa, Chem. Commun., (2002) 1500.<br />

G. Sam<strong>je</strong>ske, A. Miki, S. Ye, M. Osawa, J. Phys. Chem. B, 110 (2006) 16559.<br />

1 . Y. Liu, M. A. Lowe, F. J. DiSalvo, H. D. Abruna, J. Phys. Chem. C, 114 (2010) 14929.<br />

S. Daniele, C. Bragato, D. Battistel, Electroanalysis, 24 (2012) 759.<br />

J. D. Lović, A. V. <strong>Tripković</strong>, S. Lj. Gojković, K. Dj. Popović, D. V. <strong>Tripković</strong>, P. Olszewski,<br />

A. Kowal, J. Electroanal. Chem., 581 (2005) 294.<br />

Accepted Manuscript


12 FORMIC ACID OXIDATION ON Pt–Bi ELECTRODES<br />

Figure captions<br />

Fig. 1. Initial basic voltammograms for Pt/Biirr and Pt catalysts in 0.1 M<br />

H2SO4 solution. Scan rate: 50 mV s –1 .<br />

Fig. 2. CO stripping voltammograms on Pt2Bi, Pt/Biirr and Pt catalysts<br />

(first positive going sweeps) in 0.1 M H2SO4 solution corrected for<br />

background current. Scan rate: 50 mV s –1 .<br />

Fig. 3. Cyclic voltammograms for the oxidation of 0.125 M HCOOH in<br />

0.1 M H2SO4 solution on Pt, Pt2Bi and Pt/Biirr catalysts. Scan rate:<br />

50 mV s –1 .<br />

Fig. 4. First and 20 th sweep for the oxidation of 0.125 M HCOOH in 0.1 M<br />

H2SO4 solution on (a) Pt2Bi and (b) Pt/Biirr catalysts. Insets: Effect<br />

of cycling -Plots of current density vs. number of cycles. Scan rate<br />

50 mV s –1 .<br />

Fig. 5. Chronoamperometric curves for the oxidation of 0.125 M HCOOH<br />

at 0.2 V in 0.1 M H2SO4 solution on Pt2Bi, Pt/Biirr and Pt catalysts.<br />

Fig. 6. Tafel plots for oxidation of 0.125 M HCOOH in 0.1 M H2SO4<br />

solution on Pt2Bi, Pt/Biirr and Pt catalysts. Scan rate: 1 mV s –1 .<br />

Accepted Manuscript


QUALITY OF MOLECULAR STRUCTURE DESCRIPTORS 13<br />

Fig. 1<br />

Accepted Manuscript<br />

Fig. 2


14 FORMIC ACID OXIDATION ON Pt–Bi ELECTRODES<br />

Fig. 3<br />

Accepted Manuscript


QUALITY OF MOLECULAR STRUCTURE DESCRIPTORS 15<br />

Accepted Manuscript<br />

Fig. 4


16 FORMIC ACID OXIDATION ON Pt–Bi ELECTRODES<br />

Fig. 5<br />

Accepted Manuscript<br />

Fig. 6


J. Serb. Chem. Soc. 76 (12) 1673–1685 (2011) UDC 542.913–732:546.92’97’811–<br />

JSCS–4239 44:547.262+66.094.3<br />

Original scientific paper<br />

Microwave synthesis and characterization of Pt and<br />

Pt–Rh–Sn electrocatalysts for ethanol oxidation<br />

SANJA STEVANOVIĆ 1 * # , DUŠAN TRIPKOVIĆ 2# , DEJAN POLETI 3# , JELENA<br />

ROGAN 3# , AMALIJA TRIPKOVIĆ 1# and VLADISLAVA M. JOVANOVIĆ 1<br />

1 Department of Electrochemistry, ICTM, University of Belgrade, N<strong>je</strong>goševa 12, Belgrade,<br />

Serbia, 2 Materials Science Division, Argonne National Laboratory, Argonne, IL 60439,<br />

USA and 3 Faculty of Technology and Metallurgy, University of Belgrade,<br />

Karnegi<strong>je</strong>va 4, Belgrade, Serbia<br />

(Received 5, revised 29 April 2011)<br />

Abstract: Carbon-supported Pt and Pt–Rh–Sn catalysts were synthesized by the<br />

microwave-polyol method in ethylene glycol solution and were investigated in<br />

the ethanol electro-oxidation reaction. The catalysts were characterized in<br />

terms of structure, morphology and composition employing the X-ray diffraction<br />

(XRD), scanning tunneling microscopy and energy-dispersive X-ray spectroscopy<br />

techniques. The STM analysis indicated rather uniform particles and<br />

particle sizes below 2 nm for both catalysts. The XRD analysis of the Pt/C catalyst<br />

revealed two phases, one with the main characteristic peaks of the face-<br />

-centered cubic crystal structure (fcc) of platinum and the other related to the<br />

graphite-like structure of the carbon support, Vulcan XC-72R. However, in the<br />

XRD pattern of the Pt–Rh–Sn/C catalyst, diffraction peaks for Pt, Rh or Sn<br />

could not be resolved, indicating extremely low crystallinity. The small particle<br />

sizes and homogeneous size distributions of both catalysts could be attributed<br />

to the advantages of the microwave-assisted modified polyol process in ethylene<br />

glycol solution. The Pt–Rh–Sn/C catalyst was highly active for ethanol<br />

oxidation with the onset potential shifted by more than 150 mV to more negative<br />

values and with currents nearly 5 times higher in comparison to the Pt/C<br />

catalyst. The stability tests of the catalysts, as studied by chronoamperometric<br />

experiments, revealed that the Pt–Rh–Sn/C catalyst was evidently less poisoned<br />

than the Pt/C catalyst. The increased activity of Pt–Rh–Sn/C in comparison<br />

to Pt/C catalyst was most probably promoted by the bi-functional mechanism<br />

and the electronic effect of the alloyed metals.<br />

Keywords: Pt–Rh–Sn catalyst; ethanol oxidation; polyol synthesis; microwave<br />

irradiation; STM.<br />

* Corresponding author. E-mail: sanjat@tmf.bg.ac.rs<br />

# Serbian Chemical Society member.<br />

doi: 10.2298/JSC110405166S<br />

1673<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


1674 STEVANOVIĆ et al.<br />

INTRODUCTION<br />

High surface area carbon-supported Pt and Pt-alloys are widely proposed as<br />

promising anode catalysts in direct alcohol fuel cells (DAFCs). 1 Platinum is an<br />

excellent catalyst for the adsorption and dissociation of small organic molecules.<br />

However, platinum itself is known to be rapidly poisoned by reaction intermediates,<br />

such as CO, that are formed by dehydrogenation of the alcohol molecule 2<br />

and limited ability for cleavage of C−C bonds. The most common fuel besides<br />

methanol is ethanol, which is less toxic and can be easily produced by fermentation<br />

of sugar-containing biomass. The oxidation mechanism of ethanol in acid<br />

solution may be summarized in the following schema of parallel reactions: 1<br />

CH3CH2OH→[CH3CH2OH]ad → C1ad, C2ad → CO2 (total oxidation) (1)<br />

CH3CH2OH→[CH3CH2OH]ad → CH3CHO → CH3COOH (partial oxidation) (2)<br />

Differential mass spectrometry (DEMS) and in situ infrared spectroscopy<br />

(FTIR) determined acetaldehyde (CH3CHO) and acetic acid (CH3COOH) as the<br />

main products of the oxidation of ethanol in acidic solution, with carbon dioxide<br />

(CO2) appearing at very high positive potentials. 3,4 The efficiency of C−C bond<br />

cleavage is the key to enable this reaction to be useful in fuel cell applications<br />

and thus, the major challenge is to achieve total oxidation of ethanol to CO2 at<br />

low overpotentials. Efforts in this sense have been focused on the addition of cocatalysts,<br />

such as Ru, Rh, W, Pd, Sn, etc. 1,5–10 The addition of metals such as Sn<br />

or Ru, even though beneficial for the overall electrochemical activity of the<br />

catalysts for ethanol oxidation, does not enhance the yield of CO2, i.e., C−C bond<br />

breakage. 4,11 On the other hand, the addition of Rh to Pt improves the activation<br />

for C−C bond dissociation, but does not enhance the overall electrochemical reaction.<br />

9,10 Thus, a good electrocatalyst for ethanol oxidation should have, besides<br />

Pt, both kind of metals that would improve dehydrogenation, C−C bond dissociation<br />

and CO–O coupling. 9 Different ternary catalysts have been described in<br />

the literature1 (and references therein), but although PtSn appears to enhance<br />

ethanol oxidation better than other bimetallic catalysts and presence of Rh significantly<br />

increases C−C bond breakage, there are only a few data on ternary Pt<br />

catalyst with Sn and Rh. 12–14 Colmati et al. 12 investigated the activity for<br />

ethanol oxidation of Pt–Rh–Sn catalysts and found that at potentials higher than<br />

0.45 V vs. RHE, the alloy possessed the highest activity for this reaction, while at<br />

potentials more negative than 0.45 V, the activity was lower than that of the binary<br />

PtSn catalyst. Kowal et al. 13 indicated that Pt–Rh–SnO2 exhibited higher<br />

activity than PtSnO2 even at lower potentials. The extent of activity of such<br />

catalyst depends strongly on the SnO2 content. 14<br />

In general, metal catalytic activity is considerably dependent on the particle<br />

shape, size and particle size distribution. 15 A variety of methods can be used for<br />

nanocatalyst preparation, such as wet impregnation, sonochemical method, che-<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


MICROWAVE SYNTHESIS OF Pt AND Pt–Rh–Sn ELECTROCATALYSTS 1675<br />

mical reduction of metal precursors, etc. In the last decade, Pt or Pt-based nanoclusters<br />

with small particle size and narrow size distribution have often been synthesized<br />

by the polyol method. 16 This procedure, as most of the other conventional<br />

methods, requires longer treatment of metal precursors at a high temperature.<br />

To overcome the arduous processes, in recent years, microwave irradiation<br />

has been widely used for the preparation of nanomaterials. Compared with conventional<br />

preparation methods, microwave synthesis has the advantages of very<br />

short heating time and uniform heating of the substance, leading to a small particle<br />

size, narrow particle size distribution and high purity. Yu et al. 17 suggested<br />

that these advantages could be attributed to fast homogenous nucleation and<br />

growth of metal particles.<br />

The goal of this work was to examine ethanol oxidation on a carbon-supported<br />

Pt–Rh–Sn catalyst synthesized by the microwave-assisted polyol method.<br />

This procedure for the preparation of the previously mentioned catalyst, to the<br />

best of our knowledge, has not hitherto been described in the literature.<br />

EXPERIMENTAL<br />

Preparation of Pt and Pt–Rh–Sn/C electocatalysts<br />

To prepare Pt–Rh–Sn catalyst, a mixture of 0.5 ml of 0.05 M H2PtCl6, 0.5 ml of 0.1 M<br />

SnCl2 solution and 0.5 ml 0.05M of RhCl3 was mixed with 25 ml of ethylene glycol (EG) in a<br />

100 ml beaker under magnetic stirring. Then 0.8 M NaOH was added drop wise to adjust the<br />

pH to ≈12. The same procedure was used to synthesize the Pt catalyst. In each case, the beaker<br />

was placed in the center of a domestic microwave oven and heated 60 s for the Pt and 90 s for<br />

Pt–Rh–Sn catalyst at 700 W. After microwave heating, the mixture was uniformly mixed with<br />

20 ml of an aqueous suspension of Vulcan XC-72 carbon (containing 20 mg of carbon in the<br />

case of Pt catalyst and 53.5 mg of carbon in the case of Pt–Rh–Sn catalyst) and 150 ml of 2 M<br />

H2SO4 solution for 3 h under magnetic stirring. The resulting suspension was filtered and the<br />

residue was washed with high purity water. The solid product was dried at 160 °C for 3 h<br />

under a N2 atmosphere. The metal loading for both catalysts should have been ≈20 wt. %.<br />

Thermogravimetric analysis (TGA) confirmed 19 wt. % for Pt/C, while for Pt–Rh–Sn/C, a<br />

lower loading of metal (≈11 wt. %) was found.<br />

Characterization of the Pt/C and Pt–Rh–Sn/C electrocatalysts<br />

Thermogravimetry (TG) and differential thermal analyses (DTA) were performed simultaneously<br />

(30–800 °C temperature range) on a SDT Q600 TGA/DSC instrument (TA Instruments).<br />

The heating rate was 20 °C min-1 and the sample mass was less than 10 mg. The<br />

furnace atmosphere consisted of air at a flow rate of 100 cm3 min-1 .<br />

The unsupported Pt and Pt–Rh–Sn nanoparticles were characterized by scanning tunneling<br />

microscopy (STM). Samples were prepared by applying a few drops of a diluted colloidal<br />

solution of catalyst on a hot HOPG plate. The STM characterizations were realized<br />

using a NanoScope III A (Veeco, USA) microscope. The images were obtained in the height<br />

mode using a Pt–Ir tip (set-point current, It, from 1 to 2 nA, bias voltage, Vb = –300 mV). The<br />

mean particle size and size distribution were acquired from several randomly chosen areas of<br />

the STM images containing about 50 particles.<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


1676 STEVANOVIĆ et al.<br />

X-Ray diffraction (XRD) patterns of the powder catalysts were recorded with an Ital<br />

Structure APD2000 X-ray diffractometer in Bragg–Brentano geometry using CuK α radiation<br />

(λ = 0.15418 nm) in the step-scan mode (range: 15−85° 2θ, step-time: 2.50 s, step-width:<br />

0.02°). The program PowderCell [18] was used for phase analysis and calculation of the unit<br />

cell parameters.<br />

Microstructural examination was performed by scanning electron microscopy (SEM). An<br />

XL 30 environmental scanning microscope with a field emission gun (ESEM–FEG) (manufactured<br />

by FEI, The Netherlands) equipped with an energy dispersive X-ray (EDX) spectrometer<br />

was used. The samples were inspected using 5, 10 and 20 kV acceleration voltages at<br />

magnifications of 20000× and 10000×.<br />

The electrocatalytic activity of the catalysts was investigated by potentiodynamic and<br />

chronoamperometric tests using an Autolab potentiostat/galvanostat (ECO Chemie, The<br />

Netherlands) and a three-electrode compartment cell at room temperature. The working<br />

electrode was a thin layer of Nafion-impregnated Pt/C or Pt–Rh–Sn/C catalyst applied on a<br />

glassy carbon disk electrode with a loading of 10 μg cm -2 of the catalyst counted on metal<br />

content. The thin layer was obtained from a suspension of 2 mg of the Pt–Rh–Sn/C or 1 mg of<br />

Pt/C catalyst in a mixture of 1 ml water and 50 µl of 5 % aqueous Nafion solution, prepared in<br />

an ultrasonic bath, placed onto the substrate and dried at room temperature. A Pt wire and a<br />

saturated calomel electrode (SCE) were used as the counter and reference electrode, respectively.<br />

The electrocatalytic activity of the as-prepared Pt/C and Pt–Rh–Sn/C was studied in 0.1<br />

M HClO 4 + 0.5 M C 2H 5OH solution. The electrolyte was prepared with high purity water and<br />

deaerated with N 2. Ethanol was added to the supporting electrolyte solution while holding the<br />

electrode potential at –0.2 V. The potential was then cycled up to 0.3 V, i.e., the potential<br />

range of technical interest (E < 0.4 V), at sweep rate of 20 mV s -1 . Current–time transient<br />

curves were recorded after immersion of the freshly prepared electrode in the solution at –0.2 V<br />

for 2 s followed by stepping the potential to 0.2 V and holding the electrode at this potential<br />

for 30 min.<br />

RESULTS AND DISCUSSION<br />

Catalysts characterizations<br />

The particle size and surface morphology of the unsupported Pt and Pt–Rh–Sn<br />

catalysts were characterized by STM. As observed from the top view of the STM<br />

images (Fig. 1), both catalysts had rather uniform particles of small diameter.<br />

Most of particles were spherical in shape. Cross section analysis (Fig. 1) confirmed<br />

particle sizes of < 1.7 nm for both catalysts (Table I).<br />

The Pt/C and Pt–Rh–Sn/C catalysts were characterized by X-ray powder diffraction<br />

analysis. The XRD patterns of the carbon-supported catalysts are shown<br />

in Fig. 2. Two phases were identified in the Pt/C pattern, one with the main<br />

characteristic peaks of the face-centered cubic crystal structure (fcc) of platinum<br />

(111, 200, 220 and 311) and the other with a diffraction peak at around 2θ 25°<br />

related to the graphite-like structure of the Vulcan XC-72R carbon support. The<br />

XRD peaks of the Pt–Rh–Sn/C catalyst were rather broad and diffraction peaks<br />

for Pt, Rh and Sn in the catalyst could not be separately resolved, indicating a<br />

small metal content and a very low crystallinity or an amorphous form of the catalyst.<br />

Since TGA analysis revealed the presence of only 11 wt. % of the metals<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


MICROWAVE SYNTHESIS OF Pt AND Pt–Rh–Sn ELECTROCATALYSTS 1677<br />

in the supported catalyst and the STM analysis showed very small particle size,<br />

the low metal fraction and low crystallinity could be the reason for the obtained<br />

XRD pattern of the Pt–Rh–Sn/C catalyst. The mean particle size for the Pt/C catalyst<br />

calculated by the Scherrer formula 19 was larger than that obtained by STM<br />

(Table I), possibly because unsupported catalysts were used for the STM analysis.<br />

Still, the agreement can be described as good because the values calculated<br />

by the Scherrer formula also account for a very probable lattice stress.<br />

Fig. 1. STM Images and height profiles of the Pt catalyst (50×50 nm 2 ×4 nm)<br />

and Pt–Rh–Sn catalyst (30×30 nm 2 ×4 nm).<br />

The small particle sizes and homogeneous size distributions of both catalysts<br />

could be attributed to the advantages of the microwave-assisted modified polyol<br />

process in which ethylene glycol (EG) and hydroxide are used as stabilizers. The<br />

metal salts and hydroxide react to form colloidal metal hydroxide particles,<br />

which are reduced to metal nanoclusters by EG. 16 In this process, pH value of the<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


1678 STEVANOVIĆ et al.<br />

TABLE I. Characteristics of the Pt/C and Pt–Rh–Sn/C catalysts obtained by STM, XRD and<br />

EDX analysis<br />

Catalyst<br />

Particle size, nm Unit cell parameter<br />

STM XRD (XRD), a / nm<br />

Elemental composition (Pt:Rh:Sn), at. %<br />

Nominal EDX<br />

Pt/C 1.7±0.3 2.5 0.3944<br />

Pt–Rh–Sn/C 1.2±0.3 – –<br />

25:25:50 46:32:22<br />

Fig. 2. XRD Patterns of the Pt /C and Pt–Rh–Sn/C<br />

catalysts.<br />

solution is very important for obtaining stable metal particles. Hydroxide/metal<br />

molar ratio depends on the metal in question and is characteristic of electrostatic<br />

stabilization of metal colloids by the adsorbed anions. 16 EG acting as both reaction<br />

and dispersion media can efficiently adsorb and stabilize the surface of the<br />

particles 20 and favor the production of monodispersed metal particles with good<br />

dispersivity. 17 The high viscosity of this compound also helps in preventing agglomeration<br />

of the nanoparticles and for this reason, the water content in the reaction<br />

solution effects this process as well as the reaction temperature in influencing<br />

the particle size and size distribution. 21 Since EG has a large permanent<br />

dipole, it is very susceptible to microwave irradiation, which can absorb the<br />

energy from the microwave field and the polar reaction solution is heated up to a<br />

high temperature instantaneously. 22 Nevertheless, the heating rate of EG dispersion<br />

system could be affected by parameters of microwave operation, such as<br />

irradiation time, amount of dielectric and additives. 23 The fast and uniform microwave<br />

heating reduces the temperature and concentration gradients, thus accele-<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


MICROWAVE SYNTHESIS OF Pt AND Pt–Rh–Sn ELECTROCATALYSTS 1679<br />

rating the reduction of the metal ions and the formation of metallic nuclei. 24–27<br />

In processes with plurality of pathways and activation barriers, according to Rao<br />

et al. 26 , microwaves might promote the pathway with the lowest activation barrier.<br />

EDX Analysis of the Pt–Rh–Sn/C surface composition gave 48 at. % Pt and<br />

32 at. % Rh and 20 at. % Sn for the Pt–Rh–Sn alloy (Table I), which deviates<br />

from the nominal composition in the initial mixture (25:25:50).<br />

Bearing in mind the aforesaid, the irradiation time in the microwave heating<br />

together with NaOH/metal and EG/water ratios used in the synthesis of both<br />

catalyst were probably the reason for the so small particle size obtained, and also<br />

for the yield of the metal and the catalyst composition.<br />

Electrochemical performances<br />

Ethanol oxidation was studied at the as-prepared Pt/C and Pt–Rh–Sn/C catalysts.<br />

The cyclic voltammogram for Pt–Rh–Sn/C after only two cycles to characterize<br />

roughly the surface but to avoid significant dissolution of Rh and Sn is<br />

shown in Fig. 3. The cyclic voltammograms for Pt/C are given as steady state and<br />

after two cycles for comparison. The steady state CV for Pt/C was similar to<br />

those for polycrystalline platinum or other Pt catalysts supported on high surface<br />

area carbon, with a well-defined region of hydrogen adsorption/desorption, separated<br />

by a double layer from the region of surface oxide formation. These regions<br />

were not well-defined at the beginning of cycling since the surface was still not<br />

fully reconstructed. The CV for Pt–Rh–Sn/C was similar to the CV for PtRh/C10 i / mA<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

-0.1<br />

-0.2<br />

Pt-Rh-Sn/C 2 nd cycle<br />

Pt/C<br />

Pt/C 2 nd cycle<br />

-0.3<br />

-0.4 -0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

E / V vs. SCE<br />

Fig. 3. Basic voltammograms of the Pt/C and Pt–Rh–Sn/C catalysts in<br />

0.1 M HClO 4, v = 50 mV s -1 .<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


1680 STEVANOVIĆ et al.<br />

and is characterized by a large single peak in the hydrogen adsorption/desorption<br />

region, which Lima et al. 10 associated with adsorption/desorption of hydrogen on<br />

the intermetallic phase of PtRh. The shift of the reduction peak for Pt–Rh–Sn/C<br />

to more a negative potential value in comparison to Pt/C could be an indication<br />

of alloyed Pt and Rh. The larger double layer in the case of Pt–Rh–Sn/C compared<br />

to Pt/C was due to the lower metal content (11 mass %) of this catalyst in<br />

comparison to Pt/C (20 mass %).<br />

The electrocatalytic activities of the as-prepared catalysts were studied in 0.1<br />

M HClO4 + 0.5 M C2H5OH solution and the positive scan voltammetric curves<br />

are presented in Fig. 4. The Pt–Rh–Sn/C catalyst was highly active in ethanol<br />

oxidation with the onset potential at approximately –0.15 V (shifted by ≈0.15 V<br />

towards more negative potentials compared to Pt/C) and rapid kinetics. The hydrogen<br />

adsorption/desorption peaks were clearly suppressed because the ethanol<br />

adsorption displaced the adsorbed hydrogen from the interface. The current densities,<br />

calculated on Pt content, throughout the studied potential region were five<br />

times higher for the Pt–Rh–Sn/C catalyst in comparison to the Pt/C catalyst. The<br />

stability of the catalysts was studied in chronoamperometric experiments and the<br />

results are presented in Fig. 5. The higher initial current density at 0.2 V on the<br />

Pt–Rh–Sn/C catalyst in comparison to the Pt/C catalyst is in accordance with the<br />

-1<br />

j / mA mg Pt<br />

100<br />

50<br />

0<br />

Pt/C<br />

Pt-Rh-Sn/C<br />

-0.2 -0.1 0.0 0.1 0.2 0.3<br />

E / V (SCE)<br />

Available online at www.shd.org.rs/JSCS<br />

Fig. 4. Potentiodynamic curves for<br />

the oxidation of 0.5 M C 2H 5OH at<br />

the Pt/C and Pt–Rh–Sn/C catalysts in<br />

0.1 M HClO 4, v = 20 mV s -1 .<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


MICROWAVE SYNTHESIS OF Pt AND Pt–Rh–Sn ELECTROCATALYSTS 1681<br />

potentiodynamic measurements. The initial current at Pt–Rh–Sn/C stabilizes at a<br />

value that was significantly higher than that for the Pt/C catalyst. The Pt–Rh–<br />

–Sn/C catalyst is evidently more active and less poisoned than the Pt/C catalyst.<br />

-1<br />

j / mA mg Pt<br />

50<br />

40<br />

30<br />

20<br />

10<br />

Pt/C<br />

Pt-Rh-Sn/C<br />

0<br />

0 300 600 900 1200 1500 1800<br />

t / s<br />

Fig. 5. Chronoamperometric curves for the oxidation of 0.5M C2H5OH at 0.2 V<br />

at the Pt/C and Pt–Rh–Sn/C catalysts in 0.1 M HClO4. The better activity for the EOR of the ternary Pt–Rh–Sn catalyst in comparison<br />

to Pt/C catalyst must be related to the formation of the ternary alloy. Thus,<br />

the presence of Sn and Rh in the catalyst can promote ethanol oxidation by an<br />

electronic effect in the Pt-based electrode material affecting the adsorption properties<br />

of the surface, making this system less prone to poisoning by organic<br />

species than pure Pt. Sn or its oxides can supply surface oxygen-containing species<br />

at lower potentials by activation of the interfacial water molecule necessary<br />

to complete the oxidation of the adsorbed reaction intermediates leading to<br />

carbon dioxide, in the situation that the C−C bond was broken, or to the formation<br />

of acetic acid. 12,29 This oxidative removal of CO-like species strongly<br />

adsorbed on adjacent Pt active sites proceeds through the so-called bi-functional<br />

mechanism. According to density functional theory (DFT) calculations, 14 the role<br />

of Rh was to adsorb and stabilize the key intermediate CH2CH2O in this route,<br />

which leads to the cleavage of C−C bonds. A back donation from the Rh d-band<br />

electrons to Pt was proposed. Thus, the presence of Pt could modify the electronic<br />

structure of Rh by partially emptying its d-band states enabling its strong<br />

bonding to CH2CH2O, while the activity of Pt was lowered, thereby preventing<br />

the partial oxidation of ethanol on the Pt sites. 14<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


1682 STEVANOVIĆ et al.<br />

In comparison with other Pt–Rh–Sn/C catalysts described in the literature 12–<br />

14 in which the atomic fraction of Sn was equal or higher than that of Pt, while<br />

the atomic fraction of Rh was much smaller than both Pt and Sn, the presently<br />

studied Pt–Rh–Sn/C catalyst with highest fraction of Pt and lowest fraction of Sn<br />

exhibited a comparably high activity for ethanol oxidation.<br />

Colmati et al. 12 studied ethanol oxidation at Pt–Rh–Sn/C catalysts with<br />

molar ratios 1:1:1 and 1:0.3:1, prepared by formic acid oxidation. The ternary<br />

catalyst with the lower fraction of Rh was found to have a higher activity, but<br />

both the Pt–Rh–Sn/C catalysts exhibited a significantly lower activity in comparison<br />

with PtSn/C catalyst at lower potentials (< 0.45 V RHE) and the opposite<br />

at higher potential values. The higher activity of PtSn/C was ascribed to the presence<br />

of SnO2 that could supply oxygen-containing species for the oxidative<br />

removal of CO and CH3CO adsorbed on adjacent Pt active sites. The introduction<br />

of Rh changes the geometry (Pt−Pt bond distances) and electronic (Pt d-band<br />

vacancy) structure of PtSn catalysts, which could improve the adsorption of ethanol<br />

and the cleavage of C−C bonds, increasing in this way the activity for ethanol<br />

oxidation at higher potentials. 12 For ethanol oxidation, Kowal et al. 13 and Li et<br />

al. 14 used Pt–Rh–SnO2/C catalysts prepared by the polyol method with conventional<br />

heating. XRD Analysis of these catalysts showed the presence of two<br />

phases, Pt–Rh alloy and SnO2. The catalysts were highly active at lower potentials<br />

(< 0.5 V RHE). The catalysts were synthesized with different ratios of the<br />

components (Pt:Rh:Sn from 3:1:2 to 3:1:6) and the highest activity for ethanol<br />

oxidation and capability to split C−C bonds, as revealed by infrared reflectionadsorption<br />

spectroscopy (IRRAS), was achieved with Pt–Rh–Sn 3:1:4. 14<br />

Although XRD analysis of the present catalyst did not give any information<br />

except for its very low crystallinity, some reasonable assumptions can be made<br />

based on literature data. It can be assumed that, similarly to other Pt–Rh–Sn<br />

catalysts, the Pt was alloyed with Rh since, according to the Pt–Rh phase diagram,<br />

29 these two metals form a solid solution at any ratio. XPS and XRD analyses<br />

of bimetallic PtSn catalysts prepared by microwave or conventional heating<br />

of ethylene glycol solutions of H2PtCl6 and SnCl2 salts indicated significant<br />

amounts of SnO2 and rather low degree of alloying. 25,30 Thus, it can be assumed<br />

that the Sn in the present Pt–Rh–Sn catalyst existed mainly as SnO2. Both assumptions<br />

mean that the prepared catalyst should be similar to the catalysts<br />

obtained by Li et al. 14 but with a larger fraction of Rh and a lower fraction of Sn,<br />

but still exhibiting a comparable performance. It was shown that the addition of a<br />

small quantity of Sn greatly enhanced the electrooxidation of ethanol at low<br />

potentials. 28,31 On the other hand, a low amount of Rh added to Pt only slightly<br />

improved the CO2 yield in ethanol oxidation, while the optimum catalyst composition<br />

for C−C bond breakage and CO2 formation was Pt:Rh 1:1 or even better<br />

3:1. 9,10 Hence, the high activity for ethanol oxidation of the present Pt–Rh–Sn/C<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


MICROWAVE SYNTHESIS OF Pt AND Pt–Rh–Sn ELECTROCATALYSTS 1683<br />

catalyst can be explained by a balanced action and well-tuned content of all three<br />

components. In addition, the particle size effect cannot be ignored since very<br />

small particles of Pt–Rh–Sn/C catalyst contribute to an increase of the active<br />

surface area of the catalyst.<br />

CONCLUSIONS<br />

A microwave-assisted polyol method was used to prepare carbon supported<br />

Pt and Pt–Rh–Sn nanoparticles with high electrocatalytic activities for the ethanol<br />

electrooxidation reaction.<br />

The structural (XRD) and surface characterization (STM) of the catalysts<br />

revealed that catalysts with small particles and a rather uniform size distribution<br />

were synthesized by this method. This could be attributed to the advantages of<br />

the microwave-assisted modified polyol process in ethylene glycol solution.<br />

The electrochemical measurements revealed a high activity of the prepared<br />

Pt–Rh–Sn/C catalyst for ethanol oxidation. This catalyst had five times higher<br />

oxidation currents and significantly lower reaction onset potential than the Pt/C<br />

catalyst. Chronoamperometric measurements confirmed notably less poisoning of<br />

the Pt–Rh–Sn/C catalyst than of the Pt/C catalyst. Although with significantly<br />

higher fraction of Rh and lower fraction of Sn in comparison to other Pt–Rh–Sn<br />

catalysts described in the literature, the catalyst prepared in the present study<br />

exhibited a similar shift of onset potential to negative values as well as lower<br />

poisoning. The increased activity of Pt–Rh–Sn/C catalyst in comparison to Pt/C<br />

catalyst was most probably promoted by the bi-functional mechanism and the<br />

electronic effect of the alloyed metals.<br />

Acknowledgements. This work was financially supported by the Ministry of Education<br />

and Science of the Republic of Serbia, Contract No. 172060.<br />

ИЗВОД<br />

МИКРОТАЛАСНА СИНТЕЗА И КАРАКТЕРИЗАЦИЈА Pt И Pt–Rh–Sn КАТАЛИЗАТОРА<br />

ЗА ОКСИДАЦИЈУ ЕТАНОЛА<br />

САЊА СТЕВАНОВИЋ 1 , ДУШАН ТРИПКОВИЋ 2 , ДЕЈАН ПОЛЕТИ 3 , ЈЕЛЕНА РОГАН 3 , АМАЛИЈА ТРИПКОВИЋ 1<br />

И ВЛАДИСЛАВА М. ЈОВАНОВИЋ 1<br />

1 2<br />

IHTM – Centar za elektrohemiju, Wego{eva 12, Beograd, Materials Science Division, Argonne National<br />

Laboratory, Argonne, IL 60439, USA i 3Tehnolo{ko–metalur{ki fakultet,<br />

Univerzitet u Beogradu, Karnegi<strong>je</strong>va 4, Beograd<br />

Pt и Pt–Rh–Sn катализатори на угљенику развијене површине су синтетизовани полиол-<br />

-микроталасним поступком у раствору етиленгликола и испитивани за реакцију елетрохемијске<br />

оксидације етанола у киселој средини. Катализатори су окарактерисани структурно,<br />

морфолошки и по саставу коришћењем XRD, STM и EDX техника. STM анализа је потврдила<br />

да су Pt и Pt–Rh–Sn честице униформне величине и пречника мањег од 2 nm. XRD анализа<br />

Pt/C катализатора показала је присуство две фазе, једне са главним карактеристичним<br />

пиковима за пљосно-центрирану кубну кристалну структуру платине (111, 200, 220 и 311) и<br />

друге са дифракционим пиком на 2θ око 25° карактеристичним за хексагоналну структуру<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


1684 STEVANOVIĆ et al.<br />

вулкана XC-72R (угљеничног носача). XRD анализа Pt–Rh–Sn/C катализатора није показала<br />

карактеристичне пикове, што је индикација веома мале кристаличности катализатора. Активност<br />

катализатора испитивана је потенциодинамичким и хроноамперометријским мерењима.<br />

Pt–Rh–Sn/C катализатор је веома активан за оксидацију етанола са почетком реакције<br />

на потенцијалима за око 150 mV помереним ка негативнијим вредностима и струјама које су<br />

око пет пута веће у поређењу са Pt/C катализатором. Стабилност катализатора испитивана<br />

хроноамперометријски показала је да се Pt–Rh–Sn/C катализатор мање трује од Pt/C катализатора.<br />

Мала величина и хомогена дистрибуција честица могу се приписати предностима<br />

микроталасне синтезе и модификованог полиол поступка у раствору етиленгликола. Већа активност<br />

Pt–Rh–Sn/C катализатора у поређењу са Pt/C катализатором последица је би-функционалног<br />

механизма и електронског (лиганд) ефекта метала у синтетизованој легури.<br />

REFERENCES<br />

(Примљено 5., ревидирано 29. априла 2011)<br />

1. E. Antolini, J. Power Sources 170 (2007) 1<br />

2. A. Lopez-Cudero, J. Solla-Gullon, E. Herrero, A. Aldaz, J. M. Feliu, J. Electroanal.<br />

Chem. 644 (2010) 117<br />

3. H. Wang, Y. Jusys, R. J. Behm, J. Power Sources 154 (2006) 351<br />

4. Q. Wang, G. Q. Sun, L. H. Jiang, Q. Xin, S. G. Sun, Y. X. Jiang, S. P. Chen, Z. Jusys, R.<br />

J. Behm, Phys. Chem. Chem. Phys. 9 (2007) 2686<br />

5. S. Rousseau, C. Coutanceau, C. Lamy, J. M. Leger, J. Power Sources 18 (2006) 158<br />

6. W. Zhou, Z. Zhou, S. Song, W. Li, G. Sun, P. Tsiakaras, Q. Xin, Appl. Catal., B 46<br />

(2003) 273<br />

7. W. J. Zhou, S. Q. Song, W. Z. Li, Z. H. Zhou, G. Q. Sun, Q. Xin, S. Douvartzides, P.<br />

Tsiakaras, J. Power Sources 140 (2005) 50<br />

8. H. Li, G. Sun, L. Cao, L. Jiang, Q. Xin, Electrochim. Acta 52 (2007) 6622<br />

9. J. P. I. de Souza, S. L. Queriroz, K. Bergamaski, E. R. Gonzalez, F. C. Nart, J. Phys.<br />

Chem. B 106 (2002) 9825<br />

10. F. H. Lima, D. Profeti, W. H. Lizcano-Valbuena, E. A. Ticianelli, E. R. Gonzales, J.<br />

Electroanal. Chem. 617 (2008) 121<br />

11. A. V. <strong>Tripković</strong>, J. D. Lović, K. Dj. Popović, J. Serb. Chem. Soc. 75 (2010) 1559<br />

12. F. Colmati, E. Antolini, E. R. Gonzalez, J. Alloys Compd. 456 (2008) 264<br />

13. A. Kowal, S. Lj. Gojković, K. S. Lee, P. Olszewski, Y. E. Sung, Electrochem. Comm. 11<br />

(2009) 724<br />

14. M. Li, A. Kowal, K. Sasaki, N. Marinkovic, D. Su, E. Korach, P. Liu, R. R. Adzic,<br />

Electrochim. Acta 55 (2010) 4331<br />

15. J. Perez, V. A. Paganin, E. Antolini, J. Electroanal. Chem. 654 (2011) 108<br />

16. Y. Wang, J. Zhang, X. Wang, J. Ren, B. Zuo, Y. Tang, Top. Catal. 35 (2005) 35<br />

17. W. Yu, W. Tu, H. Liu, Langmuir 15 (1999) 6<br />

18. W. Kraus, G. Nolze, PowderCell for Windows, V.2.4, Federal Institute for Materials<br />

Research and Testing, Berlin, Germany, 2000<br />

19. H. P. Klug, L. E. Alexander, X-Ray diffraction procedures, 2 nd ed., Wiley, New York,<br />

1974, p. 687<br />

20. C. Feldmann, C. Metzmacher. J. Mater. Chem. 11 (2001) 2603<br />

21. S. L. Knupp, W. Li, O. Paschos, T. M. Murray, J. Snyder, P. Haldar, Carbon 46 (2008)<br />

1276<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


MICROWAVE SYNTHESIS OF Pt AND Pt–Rh–Sn ELECTROCATALYSTS 1685<br />

22. W. X. Tu, H. F. Liu, Chem. Mater. 12 (2000) 564<br />

23. S. Song, J. Liu, J. Shi, H. Liu, V. Maragou, Y. Wang, P. Tsiakaras, Appl. Catal., B 103<br />

(2011) 223<br />

24. W. Tu, H. Liu, J. Mater. Chem. 10 (2000), p. 2207.<br />

25. Z. Liu, B. Guo, L. Hong, T. H. Lim, Electrochem. Comm. 8 (2006) 83<br />

26. K. J. Rao, B. Vaidhyanatham, M. Ganguli, P. A. Ramakrishnan, Chem. Mater. 11 (1999)<br />

882<br />

27. W. X. Chen, J. Y. Lee, Z. L. Liu, Chem. Comm. (2002) 2588<br />

28. F. C. Simoes, D. M. Dos Anjos, F. Vigier, J. M. Leger, F. Hahn, C. Coutanaceau, E. R.<br />

Gonyaley, G. Tremiliosi-Filho, A. R. De Andrade, P. Olivi, K. B. Kokoh, J. Power<br />

Sources 11 (2007) 1567<br />

29. M. Hansen, K. Anderko, Constitution of binary alloys, 2 nd ed., McGraw-Hill, New York,<br />

1958<br />

30. Z. Liu, L. Hong, S. W. Tay, Mater. Chem. Phys. 105 (2007) 222<br />

31. C. Lamy, S. Rousseau, E. M. Belgsir, C. Coutanceau, J. M. Leger, Electrochim. Acta 49<br />

(2004) 3901.<br />

Available online at www.shd.org.rs/JSCS<br />

_______________________________________________________________________________________________________________________________________<br />

2011 Copyright (CC) SCS


Microwave-assisted polyol synthesis<br />

of carbon-supported platinum-based<br />

bimetallic catalysts for ethanol oxidation<br />

S. Stevanović, D. <strong>Tripković</strong>, J. Rogan,<br />

K. Popović, J. Lović, A. <strong>Tripković</strong> &<br />

V. M. Jovanović<br />

Journal of Solid State<br />

Electrochemistry<br />

Current Research and Development in<br />

Science and Technology<br />

ISSN 1432-8488<br />

Volume 16<br />

Number 10<br />

J Solid State Electrochem (2012)<br />

16:3147-3157<br />

DOI 10.1007/s10008-012-1755-y<br />

1 23


Your article is protected by copyright and<br />

all rights are held exclusively by Springer-<br />

Verlag. This e-offprint is for personal use only<br />

and shall not be self-archived in electronic<br />

repositories. If you wish to self-archive your<br />

work, please use the accepted author’s<br />

version for posting to your own website or<br />

your institution’s repository. You may further<br />

deposit the accepted author’s version on a<br />

funder’s repository at a funder’s request,<br />

provided it is not made publicly available until<br />

12 months after publication.<br />

1 23


J Solid State Electrochem (2012) 16:3147–3157<br />

DOI 10.1007/s10008-012-1755-y<br />

ORIGINAL PAPER<br />

Microwave-assisted polyol synthesis of carbon-supported<br />

platinum-based bimetallic catalysts for ethanol oxidation<br />

S. Stevanović & D. <strong>Tripković</strong> & J. Rogan & K. Popović &<br />

J. Lović & A. <strong>Tripković</strong> & V. M. Jovanović<br />

Received: 10 October 2011 /Revised: 18 April 2012 /Accepted: 18 April 2012 /Published online: 5 May 2012<br />

# Springer-Verlag 2012<br />

Abstract High surface area carbon-supported Pt, PtRh, and<br />

PtSn catalysts were synthesized by microwave-assisted polyol<br />

procedure and tested for ethanol oxidation in perchloric acid.<br />

The catalysts were characterized by thermogravimetric analysis<br />

(TGA), X-ray diffraction (XRD), scanning tunnelling microscopy<br />

(STM), TEM, and EDX techniques. STM analysis of<br />

unsupported catalysts shows that small particles (∼2 nm)with<br />

a narrow size distribution are obtained. TEM and XRD examinations<br />

of supported catalysts revealed an increase in particle<br />

size upon deposition on carbon support (diameter∼3nm).The<br />

diffraction peaks of the bimetallic catalysts in X-ray diffraction<br />

patterns are slightly shifted to lower (PtSn/C) or higher<br />

(PtRh/C) 2θ values with respect to the corresponding peaks at<br />

Pt/C catalyst as a consequence of alloy formation. Oxidation of<br />

ethanol is significantly improved at PtSn/C with the onset<br />

potential shifted for∼150 mV to more negative values and the<br />

increase of activity for approximately three times in comparison<br />

to Pt/C catalyst. This is the lowest onset potential found for<br />

ethanol oxidation at PtSn catalysts with a similar composition.<br />

Chronoamperometric measurements confirmed that PtSn/C is<br />

notably less poisoned than Pt/C catalyst. PtRh/C catalyst<br />

exhibited mild enhancement of overall electrochemical reaction<br />

in comparison to Pt/C.<br />

S. Stevanović : D. <strong>Tripković</strong> : K. Popović : J. Lović :<br />

A. <strong>Tripković</strong> : V. M. Jovanović (*)<br />

ICTM, Department of Electrochemistry, University of Belgrade,<br />

N<strong>je</strong>goševa 12,<br />

Belgrade, Serbia<br />

e-mail: vlad@tmf.bg.ac.rs<br />

J. Rogan<br />

Faculty of Technology and Metallurgy, University of Belgrade,<br />

Karnegi<strong>je</strong>va 4,<br />

Belgrade, Serbia<br />

Author's personal copy<br />

Keywords PtSn/C . PtRh/C . Ethanol oxidation . Polyol<br />

synthesis . Microwave irradiation<br />

Introduction<br />

Limited sources of fossil fuels as well as the request for<br />

reduction of environmental pollution led to the development<br />

of alternative energy sources, such as fuel cells. Although the<br />

best performance, so far, has been achieved using hydrogen as<br />

fuel, its storage, handling, and distribution appear to be important<br />

barriers for direct and widespread application. Direct use of<br />

liquid fuels, such as methanol, ethanol, and formic acid, has<br />

been studied as a convenient alternative. Ethanol is receiving<br />

increasing attention due to its low toxicity and large quantity<br />

production from the fermentation of biomass as a renewable<br />

biofuel [1]. However, to oxidize ethanol efficiently, it is necessary<br />

to develop a catalyst capable for converting it completely<br />

to CO2. Pt is an excellent catalyst for dehydrogenation of small<br />

organic molecules but, on the other hand, has several significant<br />

drawbacks: high cost, extreme susceptibility to poisoning<br />

by CO, and, in the case of ethanol, the main products of its<br />

oxidation are acetaldehyde and acetic acid, while CO2 is produced<br />

at high potentials [2]. Efforts to improve catalyst performance<br />

and minimize quantity of Pt in the catalyst have been<br />

centered on the addition of metals such as Ru, Rh, W, Pd, Sn,<br />

etc. [1–10].<br />

PtSn is one of the extensively studied and among the most<br />

active bimetallic catalysts for ethanol oxidation. The activity<br />

of PtSn is attributed to a bifunctional mechanism in which<br />

alcohol is adsorbed and dehydrogenated at Pt, while added<br />

metal supplies OH at significantly lower potentials compared<br />

to Pt, providing oxidation of adsorbed reaction intermediates<br />

(CO-like species strongly adsorbed on adjacent Pt active sites)<br />

leading to carbon dioxide, in the situation that the C–C bond


was broken and the formation of acetaldehyde and acetic acid<br />

was suppressed [11, 12]. Besides, the electronic interaction<br />

between Pt and alloyed metal results in a weaker bond of<br />

adsorbed species on Pt and contributes to enhanced activity of<br />

these catalysts [12–16]. While there is no doubt in the beneficial<br />

role of the bifunctional mechanism in increased activity<br />

of PtSn catalyst for ethanol oxidation, this view on alloying<br />

effect is rather controversial. According to some authors,<br />

higher alloying degree leads to a larger increase in activity<br />

of the PtSn catalyst [14, 15], whereas for others, unchanged<br />

lattice parameter of Pt in nonalloyed Pt-SnO x catalysts enables<br />

remarkable promotion of ethanol oxidation [17–19]. However,<br />

although affecting the overall electrochemical activity in ethanol<br />

oxidation, addition of Sn does not enhance the yield of<br />

CO2, i.e., C–C bond braking as revealed by differential mass<br />

spectrometry (DEMS) and in situ infrared spectroscopy (FTIR)<br />

[2, 7, 13].<br />

On the other hand, addition of Rh to Pt leads to increased<br />

ratio of CO2/acetaldehyde and CO2/acetic acid as evidenced by<br />

DEMS and FTIR techniques, which is ascribed to improved<br />

activation of C–C bond dissociation [9, 20]. At the same time,<br />

the overall electrochemical reaction is not enhanced by the<br />

PtRh/C catalyst possibly because of a stronger CO–Rh bond<br />

and/or slower dehydrogenation of ethanol at Rh in comparison<br />

to Pt [21].<br />

In general, electrocatalytic performance of the Pt-based<br />

catalysts considerably depends not only on the nature of the<br />

added metal but also on a variety of conditions of the<br />

synthesis procedures which determine surface composition<br />

and morphology of synthesized materials. In the last decade,<br />

polyol method has been often used for the preparation of Pt<br />

or Pt-based nanoclusters with small particle size and narrow<br />

size distribution [22]. In this procedure, ethylene glycol<br />

(EG) and hydroxide are used as stabilizers. EG acting as<br />

both reaction and dispersion media can efficiently adsorb<br />

and stabilize the surface of the particles [23] and favor<br />

formation of metal particles with good dispersivity [24,<br />

25]. Since EG has high permanent dipole, it is very susceptible<br />

to microwave irradiation, which can take up the energy<br />

from the microwave field and get the polar reaction solution<br />

heated up to high temperature instantaneously [26]. The fast<br />

and uniform microwave heating reduces the temperature and<br />

concentration gradients, thus accelerating the reduction of<br />

the metal ions and enabling homogeneous nucleation and<br />

shorter crystallization time leading to formation of small<br />

uniform metal particles [27–30].<br />

In this work, as-prepared carbon-supported nanoparticles,<br />

Pt, PtRh, and PtSn catalysts, synthesized by microwaveassisted<br />

polyol method, were characterized for ethanol electrooxidation<br />

reaction. Although microwave-assisted synthesis<br />

of Pt/C and PtSn/C catalyst has been described in<br />

literature [28, 31], to our knowledge, only recently have<br />

such catalysts been studied for ethanol oxidation [32]. This<br />

Author's personal copy<br />

3148 J Solid State Electrochem (2012) 16:3147–3157<br />

reaction has not been examined by PtRh/C catalyst prepared<br />

by microwave irradiation so far.<br />

Experimental<br />

Preparation of the catalysts<br />

Stabile Pt, PtRh, and PtSn nanoparticles were prepared by<br />

polyol method. Briefly, equal volumes of 0.05 M water solutions<br />

of required metal precursor salts (H 2PtCl 6 alone or<br />

either with SnCl2 or RhCl3) were mixed with EG, and NaOH<br />

was added dropwise. The solutions mixture was constantly<br />

stirred. In this way, EG/water ratio was 20/1, NaOH/metal<br />

ratio 8/1, and pH of the solutions adjusted to over 12. The<br />

prepared solutions were heated in a microwave oven at 700 W<br />

for 60 s (Pt) or 90 s (bimetallic). After microwave heating,<br />

colloidal solutions were uniformly mixed with water suspension<br />

of high-area carbon (Vulcan XC-72) and 2 M H2SO4<br />

solution and stirred for 3 h. The resulting suspension was<br />

filtered, and the residue was rinsed with high purity water.<br />

The solid product was dried at 160 °C for 3 h in N2 atmosphere.<br />

In all cases, the amount of metal and carbon was<br />

adjusted to the loading of 20 mass% of the catalyst.<br />

Characterization of the catalysts<br />

The thermogravimetric (TGA) and differential thermal (DTA)<br />

analyses were performed simultaneously (30–800 °C range)<br />

on a SDT Q600 TGA/DSC instrument (TA Instruments). The<br />

heating rates were 20 °C min −1 , and the sample mass was less<br />

than 10 mg. The furnace atmosphere consisted of air at a flow<br />

rate of 100 cm 3 min −1 .<br />

Unsupported Pt, PtRh, and PtSn nanoparticles were characterizedbyscanningtunnellingmicroscopy(STM).Samples<br />

were prepared by applying a few drops of a diluted colloidal<br />

solution of a catalyst on a hot HOPG plate. STM characterizations<br />

were performed using a NanoScope III A (Veeco,<br />

USA) microscope. The images were obtained in the height<br />

mode using a Pt-Ir tip (set-point current, it, from1to2nA,<br />

bias voltage, Vb0−300 mV). The mean particle size and size<br />

distribution were acquired from a few randomly chosen areas<br />

in the STM images containing about 100 particles.<br />

The X-ray diffraction (XRD) patterns of the powder catalysts<br />

were recorded with an Ital Structure APD2000 X-ray<br />

diffractometer in a Bragg–Brentano geometry using CuKα<br />

radiation (λ01.5418 Å) and step-scan mode (range: 15−85 °<br />

2θ, step-time: 2.50 s, step-width: 0.02 °). The program PowderCell<br />

[33] was used for phase analysis and calculation of<br />

unit cell parameters.<br />

Structural examination of the catalysts was performed by<br />

EDX technique coupled with scanning electron microscopy<br />

usingJEOLJSM-6610(USA)instrumentwithX-Max


(silicon drift) detector and super atmospheric thin window<br />

(SATW) applying 20 kV. The measurements were performed<br />

at ten different regions of each sample.<br />

TEM analysis of the supported catalysts was carried out<br />

using a Philips CM12 TEM microscope (The Netherlands)<br />

operated at 100 kV. The samples were prepared by ultrasonically<br />

dispersing the catalysts powders in ethanol and applying<br />

a drop of the suspension onto the carbon-coated copper<br />

grid. The size distribution of the catalysts particles were<br />

created based on 450 particles from a few different areas<br />

of the sample for each catalyst.<br />

Electrochemical measurements<br />

All of the electrochemical experiments were performed at<br />

room temperature in a three-electrode compartment electrochemical<br />

cell with a Pt wire as the counter electrode and a<br />

bridged saturated calomel electrode (SCE) as reference. The<br />

working electrode was a thin layer of Nafion-impregnated<br />

Pt/C, PtRh/C, or PtSn/C catalysts applied on a polished<br />

glassy carbon disk electrode with 20 μg/cm 2 loading of<br />

the catalyst. The thin layer was obtained from a suspension<br />

of 2 mg of the respective catalyst in a mixture of 1 ml water<br />

and 50 μl of 5 % aqueous Nafion solution, prepared in an<br />

ultrasonic bath, placed onto the substrate and dried at room<br />

temperature.<br />

The electrocatalytic activity of as-prepared Pt/C, PtRh/C,<br />

and PtSn/C was studied in 0.1 M HClO4+0.5 M C2H5OH<br />

solution. Ethanol was added to the supporting electrolyte<br />

solution while holding the electrode potential at −0.2 V. The<br />

potential was then cycled up to 0.3 V, i.e., the potential<br />

range of technical interest (E


Fig. 1 STM images, crosssection<br />

analysis, and<br />

corresponding particle size distribution<br />

diagrams of Pt, PtRh,<br />

and PtSn unsupported catalysts<br />

(50×50×4 nm)<br />

Author's personal copy<br />

3150 J Solid State Electrochem (2012) 16:3147–3157<br />

Pt<br />

PtRh<br />

PtSn<br />

N/N total (%)<br />

N/N total (%)<br />

N/N total (%)<br />

10<br />

8<br />

6<br />

4<br />

2<br />

d = 1.4 ± 0.3<br />

0<br />

0 1 2 3<br />

d (nm)<br />

10<br />

8<br />

6<br />

4<br />

2<br />

d = 2.5 ± 0.3<br />

0<br />

1 2 3 4<br />

d (nm)<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

0 1 2 3<br />

d (nm)<br />

d = 1.1 ± 0.4


Fig. 2 XRD patterns of supported Pt/C, PtRh/C, and PtSn/C catalysts<br />

shift of diffraction peaks of PtSn/C and PtRh/C catalysts with<br />

respect to Pt/C and the values of their lattice parameters should<br />

indicate the alloy formation between Pt and added metals but<br />

in low degree. The average crystallite size of each catalyst is<br />

calculated by Scherrer formula [36], and they are mutually<br />

similar, but for each catalyst larger than obtained by STM<br />

(Table 1). The reason for this difference is most probably<br />

because unsupported catalysts have been used for STM analysis,<br />

while XRD examinations have been performed with<br />

supported catalysts. Since the process of particle deposition<br />

onto the carbon support involves thermal treatment, it is<br />

possible that sintering of the particles occurred. Still, the<br />

agreement can be described as considerably good.<br />

Typical TEM images of supported Pt-based catalysts (Pt/<br />

C, PtRh/C, and PtSn/C) and corresponding histograms of<br />

particle size distribution are displayed in Fig. 3. Observation<br />

of these images points out that catalyst particles are uniformly<br />

distributed on the carbon support. From the particle<br />

size histograms, it is obvious that the supported catalysts<br />

have similar particle size of ~3 nm which is in a very good<br />

agreement with the mean particles size evaluated from the<br />

Table 1 Pt/C, PtRh/C, and PtSn/C catalyst parameters obtained by<br />

STM, XRD, TEM, and EDX analyses<br />

Catalyst Pt/C PtRh/C PtSn/C<br />

STM a Particle size (nm) 1.4±0.3 2.5±0.3 1.1±0.4<br />

XRD Lattice parameter<br />

a (nm)<br />

0.3944 0.3934 0.3963<br />

XRD Particle size<br />

(nm)<br />

2.5 2.6 2.6<br />

TEM Particle size (nm) 2.7±0.8 3.1±0.7 2.4±0.7<br />

Elemental composition<br />

Nominal (at.%)<br />

50:50 50:50<br />

Elemental composition<br />

EDX (at.%)<br />

60.16:39.84 53.18:46.82<br />

a STM analysis of unsupported catalyst particles<br />

Author's personal copy<br />

J Solid State Electrochem (2012) 16:3147–3157 3151<br />

XRD patterns. This confirms sintering of the metal particles<br />

obtained by microwave irradiation upon their deposition<br />

onto the carbon support and explains the difference with<br />

STM analysis of unsupported catalysts.<br />

Although no peaks for metallic Rh or Sn or their oxides<br />

were detected by XRD, their presence should not be excluded,<br />

since they might be present in amorphous state or have a<br />

smaller particle size. Moreover, they should be present in a<br />

significant amount because EDX and TGA analyses of the<br />

catalysts have indicated practically no loss in the catalysts<br />

components. TGA revealed 20, 18, and 16 wt.% for Pt/C,<br />

PtSn/C, and PtRh/C catalyst, respectively. EDX compositions<br />

of bimetallic PtSn/C and PtRh/C catalysts (Table 1)<br />

slightly deviate from the nominal values.<br />

Electrochemical performances<br />

Ethanol oxidation was studied at as-prepared Pt/C,<br />

PtRh/C, and PtSn/C catalysts. Therefore, the initial voltammograms,<br />

i.e., only 1 cycle for each of the catalysts<br />

have been recorded and presented in Fig. 4. Steady state<br />

cyclic voltammogram for Pt/C is given in the insert to<br />

illustrate similarity of this catalyst with polycrystalline<br />

platinum or other Pt catalysts supported on high area<br />

carbon [8, 13]. Cyclic voltammograms for Pt/C and<br />

PtSn/C are characterized by a defined region of hydrogen<br />

adsorption/desorption, separated by a double layer<br />

from the region of surface oxide formation. In the<br />

hydrogen adsorption/desorption region, the cyclic voltammogram<br />

for PtSn/C catalyst is similar to the voltammogram<br />

for Pt/C but with smaller currents due to a<br />

smaller amount of Pt sites on the surface. There is no<br />

clear boundary between hydrogen adsorption/desorption,<br />

double layer, and surface oxide regions in the cyclic<br />

voltammogram for the PtRh/C catalyst. The voltammogram<br />

reminds us of the one for such catalyst obtained<br />

by Lima et al. [9, 21], characterized by a large single<br />

peak in hydrogen adsorption/desorption region which<br />

the authors associated to adsorption/desorption of hydrogen<br />

on a PtRh intermetallic phase. Shift of reduction<br />

peak for PtRh/C to a more negative potential value in<br />

comparison to Pt/C should be an indication of more<br />

stable oxides of this catalyst [21].<br />

The real surface area of the as-prepared catalysts was<br />

evaluated from CO stripping measurements (Fig. 5), assuming<br />

linear bonding of CO ads [2, 9, 37]. This enabled calculation<br />

of particle size of Pt/C and PtRh/C catalysts<br />

supposing homogenous distribution of spherical-shaped particles<br />

from the equation [38]:<br />

d ¼ 6ooo<br />

p S<br />

where S (square meter per gram) is specific surface area and ρ


Author's personal copy<br />

3152 J Solid State Electrochem (2012) 16:3147–3157<br />

Fig. 3 TEM images and particle size distributions of supported Pt/C, PtRh/C, and PtSn/C catalysts<br />

N/N total (%)<br />

N/N total (%)<br />

N/N total (%)<br />

30<br />

20<br />

10<br />

0<br />

50<br />

40<br />

30<br />

20<br />

10<br />

d = 2.7 ± 0.8<br />

1 2 3<br />

d (nm)<br />

4 5 6<br />

Pt/C<br />

d = 3.1 ± 0.7<br />

0<br />

0 1 2 3<br />

d (nm)<br />

4 5 6<br />

50<br />

40<br />

30<br />

20<br />

10<br />

PtRh/C<br />

d = 2.4 ± 0.7<br />

0<br />

0 1 2 3<br />

d (nm)<br />

4 5 6<br />

PtSn/C


Fig. 4 Initial basic voltammograms of as-prepared Pt/C, PtRh/C, and<br />

PtSn/C catalysts in 0.1 M HClO 4, v050 mV/s. Insert shows steady<br />

state CV of Pt/C catalyst and high-area carbon (Vulcan XC-72)<br />

is the density of platinum or alloy. The density of alloy was<br />

calculated according to Chen et al. [39] from the equation:<br />

1<br />

¼<br />

ρPtM cPt þ<br />

ρPt cM ρM where χ is the mass fraction of the metal in alloy.<br />

Since CO does not adsorb at Sn [40], the real surface area<br />

for the PtSn/C catalyst refers only to Pt, and therefore,<br />

particle size for this catalyst cannot be evaluated. On the<br />

contrary, as CO adsorbs at Rh, the calculated real surface<br />

area of the PtRh/C catalyst then correlates to whole catalyst<br />

surface and enables particle size calculation. The data are<br />

presented in Table 2. The values obtained for the real surface<br />

area of the catalysts confirm that the surface composition<br />

corresponds to the bulk composition. So, for PtSn/C catalyst,<br />

the real surface area, which refers only to Pt, is 52 % of<br />

the real surface area of the Pt/C catalyst that coincides well<br />

with EDX result on the bulk composition of the PtSn/C<br />

catalyst. On the other hand, the values obtained for the<br />

particle size, being slightly higher than those obtained by<br />

Fig. 5 CO ads stripping curves for as-prepared Pt/C, PtRh/C, and PtSn/<br />

C catalysts in 0.1 M HClO 4, v050 mV/s<br />

Author's personal copy<br />

J Solid State Electrochem (2012) 16:3147–3157 3153<br />

Table 2 Particle size, real and specific surface areas of Pt/C, PtRh/C,<br />

and PtSn/C catalysts calculated from CO ads stripping and TEM<br />

analyses<br />

Catalyst Pt/C PtRh/C PtSn/C<br />

Real surface area (cm 2 ) 3.43 3.74 1.79<br />

Mean particle size (nm) 3.3 3.4 n/a<br />

Specific surface area (m 2 /g) 85.8 93.5 81.4 a<br />

Specific surface area (TEM) b (m 2 /g) 88.3 91.7 155.6<br />

a<br />

Calculated only on Pt content<br />

b<br />

S was calculated taking into account particle size and their fractions,<br />

i.e., size distribution<br />

TEM, confirm agglomeration and sintering of the initial<br />

particles during deposition onto the carbon support. Also,<br />

a good agreement between values for the specific surface<br />

areas for Pt/C and PtRh/C, calculated using particle size<br />

obtained from TEM analysis and from the values for real<br />

surface area obtained by CO stripping, points out mainly<br />

monodispersion of the catalyst particles over the substrate.<br />

The large difference between these values for PtSn/C catalyst,<br />

is because the real surface area from CO stripping,<br />

refers only to Pt.<br />

Oxidation of adsorbed CO monolayer enables not only<br />

the estimation of catalyst active surface area but also reveals<br />

the catalyst tolerance to CO. Observing the curves presented<br />

in Fig. 5, it can be noticed that at Pt/C, oxidation of CO<br />

commences somewhat before 0.4 V versus SCE and proceeds<br />

through a single sharp peak with current maximum at<br />

0.55 V. The onset potential for CO oxidation at PtSn/C is<br />

shifted to lower values for more than 0.3 V, and the oxidation<br />

curve is significantly broadened as well as split into two<br />

peaks (shoulders). The latter one is centered at the potential<br />

value very close to the value of Pt/C, while the former is<br />

shifted for about 100 mV more negative and corresponds to<br />

CO oxidation peaks at Pt-Sn catalysts [12, 15]. Negative<br />

shift of the onset potential of CO oxidation at PtSn/C catalyst<br />

is attributed to the presence of oxygen-containing species<br />

at Sn at lower potentials in comparison to Pt as well as<br />

to the electronic effect of Sn on Pt [12, 15]. Two signals in<br />

the potential peak during oxidation of CO can be explained<br />

by low alloying degree of this catalyst and thus, part of<br />

nonalloyed Pt. Contrary to the case of PtSn/C, CO oxidation<br />

at PtRh/C commences at practically the same potential as at<br />

Pt/C. Since CO adsorption energy for Pt-CO is 125 kJ/mol<br />

and for Rh-CO is 134 kJ/mol [21], the addition of Rh to Pt<br />

in the PtRh catalyst might increase CO bond energy at PtRh,<br />

and although OH adsorbs at Rh at lower potentials in comparison<br />

to Pt, the oxidation of CO at PtRh is retarded.<br />

The activity of the as-prepared Pt/C, PtRh/C, and PtSn/C<br />

catalysts for ethanol oxidation was evaluated from potentiodynamic<br />

and chronoamperometric measurements in 0.1 M


HClO4+0.5 M C2H5OH solution. Potentiodynamic measurements<br />

were carried out in the potential range of technical<br />

interest, i.e., up to 0.3 V. This is also the range where no<br />

leaching of the second metal (Sn or Rh) can occur [5]. One<br />

anodic sweep at each catalyst was recorded in the extended<br />

potential region up to 0.8 V as well, in order to follow<br />

completeness of ethanol oxidation. The first anodic sweep<br />

recorded at each catalyst in low potential region is presented<br />

in Fig. 6. While PtSn/C exhibits significant activity related<br />

to Pt/C, there is almost no difference in electrocatalytic<br />

activity between PtRh/C and Pt/C catalysts. Ethanol oxidation<br />

commences at −0.15 V at as-prepared PtSn/C and<br />

proceeds with approximately three times higher currents in<br />

comparison to Pt/C. At higher potentials in the extended<br />

potential region (Fig. 7), PtSn/C remains more active than<br />

Pt/C, whereas PtRh/C losses its activity and becomes inferior<br />

in comparison to Pt/C. It should be mentioned here that<br />

Pt/C alone with onset potential for ethanol oxidation as well<br />

as the potential of anodic peak located at values lower for<br />

more than 0.1 V exhibits better performance compared to Pt/<br />

C catalysts described in literature [11, 32, 41, 42]. Analysis<br />

of the voltammetric curves in Figs. 6 and 7 point out that the<br />

Fig. 6 Potentiodynamic curves in low potential region for the oxidation<br />

of 0.5 M C 2H 5OH at as-prepared Pt/C, PtRh/C, and PtSn/C<br />

catalysts in 0.1 M HClO 4, v020 mV/s<br />

Author's personal copy<br />

3154 J Solid State Electrochem (2012) 16:3147–3157<br />

Fig. 7 Potentiodynamic curves for the oxidation of 0.5 M C 2H 5OH at<br />

as-prepared Pt/C, PtRh/C, and PtSn/C catalysts in 0.1 M HClO 4, v0<br />

20 mV/s<br />

addition of Sn in PtSn/C catalyst leads to a significant<br />

increase in activity in comparison to Pt/C specially at potentials<br />

of technical interest (lower than 0.3 V). The catalyst<br />

retains higher activity even in the extended potential region,<br />

but the current density increases less with the increase in<br />

potential. The initial potential for ethanol oxidation at asprepared<br />

PtSn/C surface and the potential of anodic peak at<br />

0.6 V are remarkably shifted to lower values in comparison<br />

to other PtSn catalysts with similar composition [5, 6, 8, 11,<br />

32, 42, 43].<br />

As already mentioned, the presence of Sn can promote<br />

ethanol oxidation by an electronic effect and/or by facilitating<br />

a bifunctional mechanism. Activation of interfacial water<br />

molecules leading to formation of OH species required<br />

for the oxidation of adsorbates generated by dehydrogenation<br />

of ethanol at adjusted Pt sites proceeds in Sn at lower<br />

potentials in comparison to Pt. In this way, by oxidation of<br />

C1,ad fragments (mostly COad), formed by splitting of C–C<br />

bond, to CO2 and C2,ad species with the intact C–C bond,<br />

i.e., acetaldehyde to acetic acid, a beneficial effect of Sn is in<br />

refreshing active Pt surface and supplying enough Pt active<br />

sites for adsorption and dissociation of ethanol [11, 32, 42].


Fig. 8 Chronoamperometric curves for the oxidation of 0.5 M<br />

C 2H 5OH at as-prepared Pt/C, PtRh/C, and PtSn/C catalysts in 0.1 M<br />

HClO 4 at 0.2 V versus SCE<br />

Further, addition of Sn to Pt results in expansion of lattice<br />

parameter and elongation of bonding structure that changes<br />

geometric environment. The extended lattice parameter may<br />

enable C–C bond cleavage and thus, improve the catalytic<br />

activity [42]. Regarding the electronic effect, since Sn has<br />

four valence electrons available for modification of unfilled<br />

d band states of Pt atoms, this charge transfer from Sn atoms<br />

to neighboring Pt atoms may induce a weaker bond between<br />

adsorbates and Pt atoms. In the case of adsorbates with<br />

carbon atom, the poisoning of Pt should be reduced and<br />

adsorption of ethanol could also be diminished [42]. However,<br />

this charge transfer leads to easier formation of Sn<br />

oxide species, and these oxygen-containing species can<br />

facilitate electrooxidation of adsorbates on Pt sites at lower<br />

potentials.<br />

XRD analysis of our PtSn/C catalyst reveals that the<br />

addition of Sn induced only slight extension of Pt–Pt distances<br />

resulting in rather low alloying degree. Since TG and<br />

Fig. 9 di/dt versus time plots (from t00tot0120 s) for as-prepared Pt/<br />

C, PtRh/C, and PtSn/C catalysts<br />

Author's personal copy<br />

J Solid State Electrochem (2012) 16:3147–3157 3155<br />

Fig. 10 Long-term stability of PtSn/C and PtRh/C catalysts in 0.1 M<br />

HClO 4+0.5 M C 2H 5OH versus number of scans, v020 mV/s (current<br />

values are at 0.3 V)<br />

EDX analyses indicated negligible loss of catalyst components,<br />

we could reasonably assume that significant quantity<br />

of nonalloyed Sn is present in the catalyst and on its surface.<br />

XPS and XRD analyses of bimetallic PtSn catalysts prepared<br />

by microwave-assisted polyol method identify a significant<br />

amount of SnO2 and rather low alloying degree [28,<br />

31, 32]. Thus, it is very likely that Sn in our PtSn/C catalyst<br />

is present mainly as SnO2. The increased activity of our<br />

PtSn/C in comparison to Pt/C catalyst, therefore, is most<br />

probably enabled by a balanced ratio of Pt ensembles for<br />

ethanol dehydrogenation and C–C bond splitting and tin<br />

oxide for OH formation, i.e., by bifunctional mechanism<br />

although the electronic effect of alloyed Sn could play some<br />

role as well. In addition, high activity of as-prepared PtSn/C,<br />

and also Pt/C, catalyst could be supported by the presence of<br />

Pt adislands as well. These Pt adislands, i.e., highly undercoordinated<br />

Pt atoms on the catalyst surface are detected by<br />

STM on Pt single crystals and their remarkable activity in<br />

the oxidation of CO at lower potentials confirmed by FTIR<br />

[44]. Since polycrystalline Pt should have a number of such<br />

undercoordinated Pt atoms, they might as well contribute to<br />

the high activity of the catalysts.<br />

Contrary to PtSn/C, PtRh/C catalyst exhibits similar activity<br />

for ethanol oxidation as Pt/C at potentials of technical<br />

interest, i.e., up to 0.3 V (Fig. 6). At higher potentials, the<br />

faradeic currents at PtRh/C are lower in comparison to Pt/C<br />

(Fig. 7). Such behavior of the catalyst is in accordance with<br />

literature data [9, 20, 21] for PtRh catalysts of similar Pt to<br />

Rh molar ratio, with the exception for onset potential, since<br />

the oxidation of ethanol commences at our PtRh/C at lower<br />

values. According to Bergamaski et al. [20, 21] Rh is less<br />

efficient for dehydrogenation of ethanol than Pt, and the<br />

high-energy barrier for dehydrogenation can lower reaction<br />

rate at surfaces containing Rh. Another reason for a low


eaction rate at PtRh catalysts might be difficulty in electrooxidation<br />

of CO on Rh because of stronger O–Rh bonding<br />

in comparison to Pt–O that can result in higher activation<br />

energy for CO–O coupling and thus hinder CO oxidation<br />

[20]. In this sense, small particle size plays an important<br />

role, since small particles are more oxophylic. Still, DEMS<br />

and FTIR studies revealed improved CO2 and decreased<br />

acetaldehyde yields at PtRh surfaces that should mean that<br />

the role of Rh in PtRh catalysts is more likely to increase C–<br />

C bond splitting and not provide oxygen more readily for<br />

CO electrooxidation [20].<br />

To examine the poisoning tolerance of our Pt/C, PtRh/C, and<br />

PtSn/C catalysts during ethanol oxidation, chronoamperometric<br />

experiments have been performed, and the results are presented<br />

in Fig. 8. The PtSn/C catalyst exhibits higher initial current<br />

density at 0.2 V, while Pt/C and PtRh/C demonstrate similar<br />

behavior with lower currents, which is in accordance with<br />

potentiodynamic measurements (Fig. 6). In Pt/C and PtRh/C<br />

catalysts, contrary to PtSn/C, current decay rapidly and reaches<br />

low steady state values in a few minutes. Current decreases<br />

slowly at PtSn/C and stabilizes at the value which is significantly<br />

higher than for two other catalysts in the experimental<br />

time period. This proves that PtSn/C is considerably less poisoned<br />

than Pt/C or PtRh/C. Lower poisoning of PtSn/C is<br />

confirmed by di/dt as a function of t for short period of time<br />

(t from0to2min.)(Fig.9). A smaller value of the slope means<br />

lower initial poisoning of the electrode surface [45]. Thus, it<br />

appears that the addition of Sn to Pt facilitates lower poisoning<br />

compared to Pt/C or PtRh/C, since the experimental slope is<br />

lower by factors 1.29 and 1.33, respectively.<br />

In addition, long-term stability of PtSn/C and PtRh/C<br />

catalysts was tested for ethanol oxidation in 0.1 M HClO4<br />

during 150 cycles in the potential range of technical interest,<br />

i.e., between −0.2 V and 0.3 V. The results are presented in<br />

Fig. 10 as the current value at 0.3 V versus number of the<br />

cycle. For both catalysts after initial decrease during the first<br />

5 cycles, decline in current values is significantly retarded<br />

and for PtSn/C at the end of the test is about 70 % lower in<br />

comparison to the fifth cycle, while for PtRh/C, it is about<br />

50 %. This loss in catalyst activity may be due to the<br />

poisoning of the surface or to leaching of the added nonnoble<br />

metal. To test leaching of Sn and Rh, the PtSn/C and<br />

PtRh/C electrodes were examined by EDX after the experiments<br />

and the results are shown in Fig. 10 as well. These<br />

analyses revealed minor leaching of both metals from bimetallic<br />

catalysts and confirmed their stability.<br />

Conclusions<br />

Ethanol oxidation reaction was studied at carbon-supported<br />

Pt/C, PtRh/C, and PtSn/C catalysts prepared by microwaveassisted<br />

polyol procedure. According to lattice constants<br />

Author's personal copy<br />

3156 J Solid State Electrochem (2012) 16:3147–3157<br />

obtained from XRD analysis, low alloyed PtSn and PtRh<br />

catalysts were obtained. STM examinations revealed that<br />

unsupported catalysts have rather uniform small particles of<br />

approximately 2 nm. This should be attributed to the advantages<br />

of microwave-assisted polyol process in EG solution.<br />

During deposition of catalyst particles onto the carbon substrate<br />

and thermal treatment, sintering and agglomeration<br />

occur, and supported catalysts as determined by TEM have<br />

somewhat larger particles of approximately 3 nm.<br />

It is found that the activity of Pt-based bimetallic catalyst<br />

for ethanol oxidation greatly depends on secondary metal<br />

and the electrode potential. While addition of Sn to Pt leads<br />

to the significant enhancement of ethanol oxidation and<br />

lower poisoning of electrode surface, addition of Rh only<br />

modestly changes overall electrochemical reaction. The onset<br />

potential for ethanol oxidation at our PtSn/C catalyst is<br />

remarkably shifted to lower values in comparison to other<br />

PtSn catalysts with similar composition. The increased activity<br />

of PtSn/C catalyst is mainly due to a bifunctional<br />

mechanism, enabled most probably by nonalloyed Sn<br />

(SnO2), although an electronic effect of low alloyed PtSn<br />

could play some role as well.<br />

Acknowledgments This work was financially supported by the Ministry<br />

of Education and Science, Republic of Serbia, Contract No.<br />

172060.<br />

References<br />

1. Antolini E (2007) J Power Sources 170:1–12<br />

2. Wang Q, Sun GQ, Jiang LH, Xin Q, Sun SG, Jiang YX, Chen SP,<br />

Jusys Z, Behm RJ (2007) Phys Chem Chem Phys 9:2686–2696<br />

3. Zhou W, Zhou Z, Song S, Li W, Sun G, Tsiakaras P, Xin Q (2003)<br />

App Catal B 46:273–285<br />

4. Tsiakaras PE (2007) J Power Sorces 171:107–112<br />

5. Kowal A, Gojković SLj, Lee K-S, Olszewski P, Sung Y-E (2009)<br />

Electrochem Commun 11:724–727<br />

6. Colmati F, Antolini E, Gonzalez ER (2008) J Alloys Compd<br />

456:264–270<br />

7. Wang H, Jusys Z, Behm RJ (2006) J Power Sources 154:351–359<br />

8. Li H, Sun G, Cao L, Jiang L, Xin Q (2007) Electrochim Acta<br />

52:6622–6629<br />

9. Lima FHB, Profeti D, Lizcano-Valbuena WH, Ticianelli EA, Gonzalez<br />

ER (2008) J Electroanal Chem 617:121–129<br />

10. Sen Gupta S, Datta J (2006) J Electroanal Chem 594:65–72<br />

11. Simões FC, Dos Anjos DM, Vigier F, Léger J-M, Hahn F, Coutanaceau<br />

C, Gonzalez ER, Tremiliosi-Filho G, De Andrade AR,<br />

Olivi P, Kokoh KB (2007) J Power Sources 167:1–10<br />

12. Vigier F, Coutanceau C, Hahn F, Belgsir EM, Lamy C (2004) J<br />

Electroanal Chem 563:81–89<br />

13. Colmenares L, Wang H, Jusys Z, Jiang L, Yan S, Sun GQ, Behm<br />

RJ (2006) Electrochim Acta 52:221–233<br />

14. Zhu M, Sun G, Xin Q (2009) Electrochim Acta 54:1511–1518<br />

15. Godoi DRM, Perez J, Villullas HM (2010) J Power Sources<br />

195:3394–3401<br />

16. Liu P, Logadottir A, Norskov JK (2003) Electrochim Acta<br />

48:3731–3742


17. Jiang L, Sun G, Sun S, Liu J, Tang S, Li H, Zhou B, Xin Q (2005)<br />

Electrochim Acta 50:5384–5389<br />

18. Delime F, Léger J-M, Lamy C (1999) J Appl Electrochem<br />

29:1249–1254<br />

19. Jiang L, Colmenares L, Jusys Z, Sun GQ, Behm RJ (2007) Electrochim<br />

Acta 53:377–389<br />

20. De Souza JPI, Queiroz SL, Bergamaski K, Gonzalez ER, Nart FC<br />

(2002) J Phys Chem B 106:9825–9830<br />

21. Bergamaski K, Gonzalez ER, Nart FC (2008) Electrochim Acta<br />

53:4396–4406<br />

22. Wang Y, Zhang J, Wang X, Ren J, Zuo B, Tang Y (2005) Topics in<br />

Catalysis 35:35–41<br />

23. Feldmann C, Metzmacher C (2001) J Mater Chem 11:2603–2606<br />

24. Yu W, Tu W, Liu H (1999) Langmuir 15:6–9<br />

25. Knupp SL, Li W, Paschos O, Murray TM, Snyder J, Haldar P<br />

(2008) Carbon 46:1276–1284<br />

26. Tu W, Liu H (2000) Chem Mater 12:564–567<br />

27. Tu W, Liu H (2000) J Mater Chem 10:2207–2211<br />

28. Liu Z, Guo B, Hong L, Lim TH (2006) Electrochem Commun<br />

8:83–90<br />

29. Rao KJ, Vaidhyanathan B, Ganguli M, Ramakrishnan PA (1999)<br />

Chem Mater 11:882–895<br />

30. Chen WX, Lee JY, Liu Z (2002) Chem Commun 2588–2589<br />

31. Liu Z, Hong L, Tay SW (2007) Mater Chem Phys 105:222–228<br />

32. Wang Y, Song S, Andreadis G, Liu H, Tsiakaras P (2011) J Power<br />

Sources 196:4980–4986<br />

Author's personal copy<br />

J Solid State Electrochem (2012) 16:3147–3157 3157<br />

33. Kraus W, Nolze G, (2000) PowderCell for Windows, V.2.4,<br />

Federal Institute for Materials Research and Testing Berlin,<br />

Germany<br />

34. Von Weast RC (1966) Handbook of chemistry and physics, 47th<br />

edn. The chemical Rubber Co., Cleveland, OH<br />

35. Fievet F, Lagier JP, Blin B, Beaudoin B, Figlarz M (1989) Solid<br />

State Ionics 32–33:198–205<br />

36. Klug HP, Alexander LE (1974) X-ray diffraction procedures 2nd<br />

ed. Wiley, New York<br />

37. García G, Silva-Chong JA, Guillén-Villafuerte O, Rodríguez JL,<br />

González ER, Pastor E (2006) Catal Today 116:415–421<br />

38. Gloaguen F, Léger JM, Lamy C, Marmann A, Stimming U, Vogel<br />

R (1999) Electrochim Acta 44:1805–1816<br />

39. Chen YA, Bandeira IN, Rowe OM, Min G (1994) J Mater Sci<br />

Letters 13:1051–1053<br />

40. Shubina TE, Koper MTM (2002) Electrochim Acta 47:3621–3628<br />

41. Lamy C, Rousseau S, Belgsir EM, Coutanceau C, Léger J-M<br />

(2004) Electrochim Acta 49:3901–3908<br />

42. Kim JH, Choi SM, Nam SH, Seo MH, Choi SH, Kim WB (2008)<br />

Appl Catal B 82:89–102<br />

43. Jiang L, Sun G, Zhou Z, Zhou W, Xin Q (2004) Catal Today 93–<br />

95:665–670<br />

44. Strmcnik DS, Tripkovic DV, Van der Vliet D, Chang KC, Komanicky<br />

V, You H, Karapetrov G, Greeley JP, Stamenkovic VR, Markovic NM<br />

(2008) J Am Chem Soc 130:15332–15339<br />

45. Delime F, Léger JM, Lamy C (1998) J Appl Electrochem 28:27–35


ARTICLES<br />

PUBLISHED ONLINE: 6 MAY 2012 | DOI: 10.1038/NMAT3313<br />

Trends in activity for the water electrolyser<br />

reactions on 3d M(Ni,Co,Fe,Mn)<br />

hydr(oxy)oxide catalysts<br />

Ram Subbaraman 1,2 , Dusan Tripkovic 1 , Kee-Chul Chang 1 , Dusan Strmcnik 1 , Arvydas P. Paulikas 1 ,<br />

Pussana Hirunsit 3 , Maria Chan 3 , Jeff Greeley 3 , Vojislav Stamenkovic 1 and Nenad M. Markovic 1 *<br />

Design and synthesis of materials for efficient electrochemical transformation of water to molecular hydrogen and of hydroxyl<br />

ions to oxygen in alkaline environments is of paramount importance in reducing energy losses in water–alkali electrolysers.<br />

Here, using 3d-M hydr(oxy)oxides, with distinct stoichiometries and morphologies in the hydrogen evolution reaction (HER)<br />

and the oxygen evolution reaction (OER) regions, we establish the overall catalytic activities for these reaction as a function of<br />

a more fundamental property, a descriptor, OH–M 2+δ bond strength (0 ≤ δ ≤ 1.5). This relationship exhibits trends in reactivity<br />

(Mn


NATURE MATERIALS DOI: 10.1038/NMAT3313 ARTICLES<br />

a<br />

Normalized absorption (a.u.)<br />

c<br />

1.5<br />

1.0<br />

0.5<br />

0<br />

5<br />

i (mA cm –2 ) 10<br />

0<br />

0 20<br />

Energy shift (eV)<br />

Pt(111)<br />

Pt(111) + CoOOH<br />

25 nm 0.8 E (V) 1.7<br />

E = ¬0.1 V<br />

Co(OH) 2 ref<br />

E = +1.4 V<br />

CoOOH ref<br />

40<br />

OHad (μC cm–2 θ<br />

)<br />

300<br />

200<br />

100<br />

150<br />

i (μA cm –2 )<br />

0<br />

–150<br />

Pt(111)<br />

Pt(111) + Co(OH) 2<br />

0 0.5 1.0<br />

15 nm<br />

E (V)<br />

i (μA cm –2 )<br />

300<br />

200<br />

100<br />

M2+ δO δ(OH)<br />

2-δ<br />

M = Mn<br />

M = Fe<br />

M = Ni<br />

Pt(111)<br />

0<br />

Ni Co<br />

0<br />

Fe<br />

0.5 E (V) 1.0<br />

0<br />

Mn<br />

Figure 1 | Characterization of M 2+δ O δ (OH)2−δ/Pt(111) systems using XANES, CV and STM measurements. a, Sample XAS spectra for Co 2+δ O δ (OH)2−δ<br />

on Pt(111) surface for two different potentials E = −0.1 V and E = +1.4 V. Comparisons of reference samples at these potentials are also shown. The<br />

comparison reveals that δ = 0 at −0.1 V and δ = 1 at 1.4 V. b, STM for Co(OH)2/Pt(111) in the HER region. Polarization curves for the HER for this surface are<br />

also shown (50 mV s −1 ). The characteristic height of the clusters shown is ∼5.8 Å with a diameter for 7–8 nm. c, STM for CoOOH/Pt(111) is shown along<br />

with the polarization curves for the OER. The characteristic height of the clusters shown is ∼5.6 Å with a diameter for 15–22 nm. Pure Pt(111) polarization<br />

curves are shown for comparison in both Fig. 1b and c. d, Comparison of OHad charge as a function of oxophilicity of the metal oxide cation (M) for same<br />

coverages of the M 2+δ O δ (OH)2−δ on Pt(111). Inset shows the comparison of voltammograms for bare Pt(111) surfaces with Co 2+δ O δ (OH)2−δ/Pt(111)<br />

surface. Enhanced adsorption of OHad is observed as the larger area under the anodic peak at 0.6 V as well as the early onset of the OHad butterfly region.<br />

Table 1 | XAS analysis for M 2+δ O δ (OH)2−δ systems.<br />

Element HER potential region (E < 0.0 V) OER potential region (E ∼ 1.4 V)<br />

b<br />

d<br />

M n+ N M–O (Å) M n+ N M–O (Å)<br />

Mn 2.0 ± 0.1 1.6 ± 0.3 2.10 ± 0.04 3.5 ± 0.1 3.3 ± 0.4 1.89 ± 0.03<br />

Fe 2.7 ± 0.1 2.5 ± 0.5 1.93 ± 0.03 3.0 ± 0.1 3.4 ± 0.3 1.95 ± 0.02<br />

Co 2.2 ± 0.1 1.8 ± 0.5 2.02 ± 0.05 2.9 ± 0.2 3.9 ± 0.5 1.89 ± 0.02<br />

Ni 2.1 ± 0.1 2.5 ± 0.4 2.06 ± 0.03 2.3 ± 0.1 3.2 ± 0.6 2.02 ± 0.03<br />

M n+ —The valence state of the 3d element; N—first shell co-ordination number; M–O (Å)—characteristic bond distance between the 3d metal centre and the oxygen anion. We show that, depending on<br />

the nature of the 3d metal, different step changes in oxidation states are observed in different potential regions. The rate of change of oxidation states with potential are found to be dependent on the<br />

nature of the elements. In the HER, with the exception of Fe, all elements are in the +2 state. On the other hand, in the OER region, with the exception of Ni, all elements are in an oxidation state > +3. We<br />

note, Fe, with its complex redox chemistry at these potentials, is known to exist in multiple forms 27 , such as Fe(II) and Fe(III) oxides and hydroxides, and therefore exhibits a valence state between 2.5 and<br />

3.0. Nickel, exhibiting lower oxidation state at this potential, is not surprising, because for the Ni-modified surfaces we found that, at 1.4 V, the phase of Ni(OH)2 present on the Pt(111) surface resembles<br />

that of β-Ni(OH)2 and this phase of nickel hydroxide is known to remain stable up to high potentials (


ARTICLES<br />

latter dimension corresponds to approximately two layers of<br />

electrically conductive Co(OH)2. The distribution of the clusters<br />

over the entire surface indicates that the clusters grow in a threedimensional<br />

(3D) (Volmer–Weber 28 ) fashion, seldom achieving<br />

a complete monolayer coverage. Importantly, after recording 50<br />

cyclic voltammograms (CVs) between −0.3V±0.4 V, the STM<br />

images of all 3d-M hydr(oxy)oxide/Pt(111) systems remain the<br />

same, indicating that in this potential range the morphology of<br />

the surface is stable. In contrast, for an electrode held at 1.55 V<br />

the STM image in Fig. 1c indicates significant sintering of the<br />

CoOOH nanoparticles; a distribution of sizes ranging from ∼15<br />

to 25 nm, with approximately constant heights of two layers. This<br />

morphology was found to be stable in the OER region (potentialand<br />

time-independent). Given that similar morphological changes<br />

were exhibited by the other M 2+δ O δ (OH)2−δ/Pt(111) systems, we<br />

conclude that these well-defined, at atomic/molecular level, surface<br />

structures form the basis for any predictive ability in tailoring<br />

catalysts to have desirable reactivity for the HER and the OER in<br />

alkaline environments.<br />

We begin our electrochemical characterization by comparing<br />

the CVs of the Pt(111) and Co(OH)2/Pt(111) surfaces (inset in<br />

Fig. 1b). As was the case for the XAS and STM analyses, the CV<br />

behaviour of Co(OH)2 is a fair representation of the other 3d-M<br />

hydr(oxy)oxide systems (shown in inset of Fig. 1d). Consistent<br />

with earlier reports for Pt(111) (refs 29,30), cyclic voltammetry<br />

of Pt(111) (Fig. 1b) shows that the adsorption of underpotentially<br />

deposited hydrogen (defined as the state of hydrogen adsorbed<br />

at a potential that is positive of the Nernst potential for the<br />

hydrogen reaction, Hupd) between 0.05 and 0.35 V is followed<br />

first by a wide double-layer potential region and then by the<br />

formation of an OHad adlayer (usually termed as the ‘butterfly’<br />

region) between 0.6 and 0.95 V. Figure 2 (bottom inset) shows<br />

that the surface coverage of the Hupd (�Hupd) is reduced by ∼45%<br />

on a Co(OH)2/Pt(111) electrode. We also found that other 3d-M<br />

hydr(oxy)oxide covered surfaces behave in a similar manner (inset<br />

Fig. 1d), suggesting they act as a third body, selectively blocking<br />

the adsorption of Hupd without affecting the Pt–Hupd energetics.<br />

Thus, determination of the �Hupd from CVs of modified Pt(111)<br />

surfaces enables accurate determination of surface coverage by<br />

the 3d-M hydr(oxy)oxide clusters. In the following, Pt(111) is<br />

always modified with nearly identical (∼45%) amounts of these<br />

clusters (Fig. 1d), thereby enabling the use of a well-characterized<br />

M 2+δ O δ (OH)2−δ/Pt(111) electrode (0 ≤ δ ≤ 1.5).<br />

In contrast to the Pt–Hupd bonding, Fig. 1b shows that the<br />

effects of a Co 2+δ O δ (OH)2−δ/Pt(111)-modified Pt(111) surface on<br />

adsorption of OH − are significant. Three distinct features are<br />

noteworthy: (1) the increase in the charge under the OHad (�OHad)<br />

region between 0.6 V < E < 0.95 V, (2) negative shifts for both<br />

the onset of the OH − adsorption and its desorption, and (3)<br />

OHad adsorption/desorption features are strongly irreversible. The<br />

magnitude of the changes in the OHad charge is greater than what<br />

can occur on an unmodified Pt(111) surface; so the oxide clusters’<br />

interaction with the OH must provide an important contribution<br />

to the observed value of �OHad. Similar behaviour is exhibited<br />

by the other 3d elements considered here (Fig. 1d); however,<br />

the charge corresponding to the formation of OHad is strongly<br />

dependent on the nature of the 3d element and they follow the<br />

order Ni < Co < Fe < Mn. This result clearly demonstrates that<br />

the oxophilicity of the 3d elements is strongly dependent on their<br />

‘nobility’ (position in the periodic table). This, in turn, follows<br />

the oxophilicity of the 3d-M hydr(oxy)oxides, a conclusion that<br />

is supported by our density functional theory (DFT) calculations,<br />

specifically for Co and Mn oxide clusters. The DFT analyses show<br />

stronger adsorption of OH on model two Co(OH)2/Pt(111) films<br />

(treated as a film with two layers of Co(OH)2 on the Pt(111) surface)<br />

than for clean Pt(111). Furthermore, the variation in binding<br />

NATURE MATERIALS DOI: 10.1038/NMAT3313<br />

energy moving from Co-based to Mn-based systems is consistent<br />

with the observed oxophilicity derived from the �OHad values (see<br />

Supplementary Information for details).<br />

We thus have a series of well-characterized electrodes showing a<br />

linear variation in the �OHad with the nature of the 3d-M elements<br />

(Fig. 1d). The trend in the values of �OHad as a function of the<br />

3d-M cation indicates that this quantity correlates well with the<br />

OHad–M 2+δ bonding. As such, this result is an excellent basis for<br />

finding variations in the kinetics of electrochemical reactions, if<br />

the rates of these reactions are controlled by the OHad–M 2+δ bond<br />

strength. It should be pointed out, however, that the physical<br />

processes that are associated with the formation of OHad on metal<br />

and metal oxide surfaces have been the sub<strong>je</strong>ct of controversy even<br />

on well-defined Pt single crystals 30–33 . Nevertheless, there is only<br />

some consensus that in alkaline solution the pseudocapacitance<br />

observed in the CVs of Pt(111) in the range 0.6 < E < 0.95 is<br />

just reversible OH − adsorption on the (111) terrace sites. Given<br />

that the thermodynamic analyses for the OHad adsorption seen<br />

in CVs is not straightforward 34 , relatively little is known about<br />

the true nature of OHad or how the adsorption energies of OHad<br />

may vary between different adsorption sites (that is, terraces versus<br />

defects). As discussed previously 35,36 , OHad species that are present<br />

at potentials below 0.6 V, are found exclusively on the defect sites.<br />

The relatively small number of such defects on Pt(111) makes<br />

them invisible in the CV owing to the larger pseudocapacitive<br />

contributions from the Hupd. To overcome this limitation of the CV<br />

method, it is customary to ‘probe’ the OHad with electrochemical<br />

reactions for which the OHad is a reactant. For our purposes here, we<br />

use the CO oxidation reaction 35,36 to verify/validate the oxophilicity<br />

trends observed in the butterfly region at potentials below 0.6 V.<br />

The reaction mechanism for CO oxidation reaction has<br />

been well established as following the Langmuir–Hinshelwood<br />

(L–H) pathway for Pt bimetallic systems such as PtSn and<br />

PtMo (refs 30,37). These catalysts are bi-functional in nature,<br />

where the CO adsorbs exclusively on the Pt sites and the<br />

OHad groups are present exclusively on the more oxophilic<br />

Sn/Mo sites that facilitate the oxidative removal of CO. In<br />

line with these systems, for the case of Pt(111) modified by<br />

M 2+δ O δ (OH)2−δ clusters, it is reasonable to propose that, whereas<br />

CO is adsorbed exclusively on the Pt sites (CObulk ↔ Pt–COad),<br />

the OHad species adsorb preferentially on 3d-M hydr(oxy)oxides<br />

(OH − + M 2+δ O δ (OH)2−δ ↔ OHad–M 2+δ O δ (OH)2−δ + e − ). The<br />

presence of OHad can then be tested simply by monitoring<br />

the rate of CO oxidation at the constant electrode potential<br />

through a L–H type reaction (Pt–COad +OHad −M 2+δ O δ (OH)2−δ +<br />

2OH − ↔ HCO −<br />

3 + e − + H2O). To see the L–H reaction in<br />

action, we present the polarization curves for CO oxidation on<br />

Co 2+δ O δ (OH)2−δ/Pt(111) and bare Pt(111) as typical examples; the<br />

corresponding activities for other 3d-element-modified electrodes<br />

are summarized in Fig. 2. Clearly, the CO oxidation current on<br />

Co 2+δ O δ (OH)2−δ/Pt(111) is shifted negatively by about 0.3 V with<br />

respect to the bare Pt(111) surface (see top inset of Fig. 2 and<br />

Supplementary Fig. S6), indicating that this surface behaves as a<br />

bi-functional catalyst. As depicted schematically in Fig. 2 (bottom<br />

inset), the reaction proceeds along the perimeters of COad islands<br />

and neighbouring M 2+δ sites. The fact that the onset potential for<br />

the CO oxidation reaction on Co-modified Pt(111) is observed at<br />

0.1 V (see Supplementary Fig. S6), strongly suggests that oxygenated<br />

species must be present on the Co 2+δ O δ (OH)2−δ defect sites at these<br />

potentials. Although the same conclusion commonly holds true for<br />

all other M 2+δ O δ (OH)2−δ/Pt(111) systems, the reactivity of these<br />

surfaces for the CO oxidation reaction is found to be dependent on<br />

the nature of the 3d element.<br />

In general, the kinetics of the CO oxidation reaction on<br />

these surfaces is expected to be a function of both the Pt–COad<br />

and OHad–M 2+δ energetics 30 . However, if we assume that the<br />

552 NATURE MATERIALS | VOL 11 | JUNE 2012 | www.nature.com/naturematerials<br />

© 2012 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT3313 ARTICLES<br />

(mV)<br />

@ 1 mA cm ¬2<br />

η<br />

600<br />

750<br />

900<br />

ΔG Pt–COad = constant<br />

Pt (111)<br />

Pt(111)–Ni2+ δO δ(OH)<br />

2-δ<br />

HCO 3 –<br />

δδ δ<br />

OH –<br />

COb COb OH<br />

OHad OHad –<br />

M2+ O (OH) 2-δ<br />

M<br />

CO CO<br />

ad ad<br />

2+ δ δ O (OH) 2-<br />

Pt<br />

Pt(111)–Co2+ δO δ(OH)<br />

2-δ<br />

HCO<br />

–<br />

3<br />

i (mA cm –2 )<br />

2<br />

1<br />

0<br />

Co2+ Pt(111)<br />

δO δ(OH)<br />

2- δ /Pt (111)<br />

0 0.5 1.0<br />

E (V)<br />

0 50 100 150 200<br />

OHad ¬ Hupd (μC cm–2 θ θ<br />

)<br />

δ<br />

Pt(111)–Fe2+ δO δ(OH)<br />

2-δ<br />

Pt(111)–Mn2+ δO δ(OH)<br />

2-δ<br />

Figure 2 | Trend in overpotential for CO oxidation is shown as a function of the 3d transition elements. The elements are arranged in the order of their<br />

oxophilicity from Mn to Ni. Pt is shown in the figure as a reference. Top inset: a comparison of the polarization curves for Pt(111) and Pt(111) with 40% Co<br />

hydr(oxy)oxides for the CO oxidation reaction. As can clearly be seen, the onset potentials for CO oxidation are shifted ∼300 mV negative from those of<br />

the bare Pt(111) surface. Bottom inset: a schematic showing the L–H mechanism for the CO oxidation reaction. CO from bulk is found to adsorb on the free<br />

Pt site near the oxide clusters. OHad is formed by either adsorption of OH − from the electrolyte and/or a change in oxidation state of the cluster cation<br />

M 2+δ . In the presence of COad and OHad in each others vicinity, reaction between COad and OHad species the occurs forming an intermediate which is<br />

eventually converted to (bi)-carbonates. The free energy for Pt–COad is fixed, which enables the treatment of these bi-functional metal-oxide/metal<br />

catalysts as a ‘pseudo’ mono-functional catalyst with a singular descriptor OHad–M 2+δ .<br />

Pt–COad interaction is independent of the nature of the 3d-M<br />

hydr(oxy)oxide (which is a reasonable assumption, considering that<br />

even Pt–Hupd energies are not affected by the presence of these),<br />

then the rate of the CO oxidation reaction should depend only on<br />

the OHad–M 2+δ energetics. Thus, by fixing the Pt–COad energetics,<br />

we can treat the reaction as a ‘pseudo’ mono-functional reaction<br />

that is controlled by the descriptor related to OHad–M 2+δ bond<br />

strength. Indeed, Fig. 2 reveals that the rate of the CO oxidation<br />

reaction is inversely proportional to the OHad–M 2+δ bond strength,<br />

that is, the activity increases in the order: Mn < Fe < Co < Ni. We<br />

propose that, for a strong interaction, such as that for OHad–Mn 2+δ ,<br />

the CO oxidation reaction is inhibited because of the relatively low<br />

reactivity of OHad. In fact, the OHad–Mn 2+δ bond is so strong that<br />

even pure Pt is more active than the Mn 2+δ O δ (OH)2−δ/Pt(111).<br />

In contrast, for Ni hydr(oxy)oxides that bind OHad neither too<br />

strongly nor too weakly, we observe the maximum activity for<br />

the CO bulk oxidation. The similarity in the trends observed for<br />

the activity for the CO oxidation reaction and the OHad–M 2+δ<br />

interaction strength in the butterfly region, confirms that the same<br />

guiding principle, namely the oxophilicity of the 3d-M cation, is<br />

valid in the Hupd region as well. This indicates that the oxophilicity<br />

trends derived in Figs 1d and 2 are valid between 0.05 and 0.95 V.<br />

By studying the OER (E > 1.5 V) and the HER (E < 0.0V), where<br />

OHad is involved in the reaction pathway, we will further probe if<br />

this trend in OHad–M 2+δ holds true in the potential regions where<br />

these two reactions occur.<br />

The OER is much more complex than the CO oxidation<br />

reaction, involving the formation of many reaction intermediates<br />

before OH − is completely transformed to O2/H2O. So far, many<br />

descriptors have been used to express the overall catalytic activities<br />

of the OER as functions of some physicochemical properties of<br />

the catalyst, including the oxygen adsorption energy in gas phase<br />

environments, the catalyst–OH bond strength, the number of d<br />

electrons, and the lattice spacing in metal electrodes 15,17,18,23,38,39 .<br />

Although these studies identify key surface properties that govern<br />

reactivity for various catalysts, the fundamental understanding<br />

of the elementary processes involved in the OER is still lacking.<br />

This is because the previous experimental results were acquired<br />

on high-surface-area catalysts, which are complex owing to the<br />

presence of ill-defined catalyst centres. In contrast, our well-defined<br />

3d-M hydr(oxy)oxide clusters on Pt(111) provide an excellent<br />

opportunity to establish the guiding principles for designing the<br />

active sites for the OER. A rigorous kinetic analysis of the OER lies<br />

outside the scope of this discussion and will not be presented here.<br />

Rather, we will focus on understanding the reactivity trends for the<br />

OER on M 2+δ O δ (OH)2−δ/Pt(111) surfaces and how these trends<br />

may help in the future design of effective OER catalysts.<br />

As a starting point, we compare the kinetic rates of the OER<br />

on the Pt(111)-oxide electrode (dubbed hereafter as PtO) and the<br />

CoOOH/PtO electrode (shown as inset in Fig. 1c); as before, the<br />

Co system is representative of the other M 2+δ O δ (OH)2−δ systems.<br />

Figure 1c shows that the onset of the OER on the CoOOH/PtO<br />

electrode is shifted by ∼0.25 V, to more negative potentials,<br />

compared with PtO. These differences may reflect variations in the<br />

energetics (activation energies and/or the enthalpies of adsorption)<br />

for the formation of active intermediates at these two surfaces, the<br />

NATURE MATERIALS | VOL 11 | JUNE 2012 | www.nature.com/naturematerials 553<br />

© 2012 Macmillan Publishers Limited. All rights reserved<br />

250


ARTICLES<br />

(mV)<br />

@ 5 mA cm –2<br />

η<br />

450<br />

600<br />

750<br />

Pt (111)<br />

PtO–Ni2+ δO<br />

δ(OH)<br />

2-δ<br />

OH –<br />

O 2<br />

OH ad<br />

PtO/AuO<br />

Intermediate<br />

M2+ δδ δ O (OH) 2¬ δ<br />

PtO–CoO(OH)<br />

OH ad<br />

H 2 O<br />

OH –<br />

NATURE MATERIALS DOI: 10.1038/NMAT3313<br />

I (mA cm –2 )<br />

Au(111)<br />

CoOOH + Au(111)<br />

Pt(111)<br />

CoOOH + Pt(111)<br />

PtO–FeO(OH)<br />

PtO–Mn2+ 1.0 1.5<br />

E (V)<br />

2.0<br />

δO<br />

δ (OH) 2-δ<br />

Ni Co Fe Mn<br />

Figure 3 | Trend in overpotential for the oxygen evolution reaction (OER) is shown as a function of the 3d transition elements. The elements are arranged<br />

in the order of their oxophilicity from Mn to Ni. Pt is shown in the figure as a reference. Top inset: a comparison of polarization curves for Pt(111) and<br />

Au(111) with 40% CoOOH for the OER. As can clearly be seen, the two potential curves are identical, suggesting a limited or no role played by the noble<br />

metal substrate for this reaction. As a result, this reaction is classified as a mono-functional reaction, and the main descriptor (as can be clearly seen from<br />

the trend) is still the OHad–M 3+ interaction. Bottom inset: a schematic showing the OER. OH − from the bulk is found to adsorb on the free catalyst site on<br />

the oxide clusters. The adsorbed OH groups OHad react with other such groups to form a reaction intermediate (re-combination), which is then further<br />

oxidized to O2 and H2O.<br />

exact values of which are unknown. Nevertheless, the low activity of<br />

PtO, consistent with earlier reports 10,14 , is indicative of the weaker<br />

OHad–PtO interaction, and thus it seems that the rate-determining<br />

step for such noble metal catalysts could be the formation of<br />

OHad–PtO intermediates (OH − +PtO ↔ OHad −PtO+e − ). Along<br />

the same lines, the significant activity enhancement of the OER<br />

on the CoOOH/PtO electrode could be due to the enhanced<br />

interaction of OH − with CoOOH. To verify this, we have also<br />

compared the OER on the CoOOH/AuO electrode. The fact that<br />

the rate of reaction on both these surfaces is the same (top inset<br />

of Fig. 3) is confirmation that the OER reaction rate is controlled<br />

only by interaction of reactants and reaction intermediates with<br />

CoOOH and not the metal substrate. Given that the same is<br />

also true for the other M 2+δ O δ (OH)2−δ/Pt(111) electrodes, we<br />

conclude that these catalysts are purely mono-functional, a fact<br />

which is further confirmed by the monotonic variation in OER<br />

activities as a function of ‘loading’ of the CoOOH clusters (see<br />

Supplementary Section S4, Fig. S5). As summarized in Fig. 3, the<br />

OER on M 2+δ O δ (OH)2−δ/PtO exhibits activities increasing from<br />

Mn to Ni hydr(oxy)oxides. This suggests that the overall reaction<br />

rates are driven by the strength of the interaction between the two<br />

oxidic species, the recombination step being the rate determining<br />

(2OHad–M 2+δ O δ (OH)2−δ → products), rather than by the initial<br />

adsorption step (OH − +M 2+δ O δ (OH)2−δOHad–M 2+δ O δ (OH)2−δ +<br />

e − ), which, as shown in Fig. 1d, exhibits the opposite trend<br />

compared with that in Fig. 3. Certainly, we are aware that more than<br />

one recombined species might be formed during OER chemistry<br />

on oxides 10,11 , but those mechanistic details do not affect our<br />

10<br />

5<br />

0<br />

interpretations of the OER trends. In particular, too strong an<br />

interaction between 3d-M hydr(oxy)oxides and OHad can lead to<br />

an adverse effect, wherein the reaction intermediates are stabilized,<br />

leading to a lower turnover frequency (defined as the number<br />

of complete reaction events per site per second). This leads to<br />

‘poisoning’ of the surface and a concomitant decrease in OER<br />

activities, as shown schematically in Fig. 3 (bottom inset). Thus, for<br />

the 3d elements considered here, Ni, with its optimal interaction<br />

strength with OHad, satisfies the Sabatier principle for catalyst<br />

design. Considering that the reactivity trends observed for the OER<br />

(Ni > Co > Fe > Mn) match that observed for the CO reactivity,<br />

we can conclude that the OHad–M 2+δ interaction (oxophilicity)<br />

trends can be extended up to the OER potential regions. In what<br />

follows, we demonstrate that the trends in the energetics between<br />

OHad–M 2+δ for OH − , produced as the water dissociation product,<br />

which are relevant in the hydrogen evolution potential region, are<br />

similar to that for the OHad formed from the supporting electrolyte<br />

between 0.05 and 2.0 V.<br />

Surprisingly, so far, the trends in activity for the HER in<br />

alkaline solutions have never been established independently from<br />

those found for the HER in acidic media 40,41 . Given that, in both<br />

environments, the measured kinetic rate of the HER exhibits a<br />

volcano type behaviour 13,42 when plotted as a function of the<br />

hydrogen adsorption energy, it has been customary to use the<br />

same descriptor for the development of cathode catalysts for the<br />

alkaline HER catalysts. Very recently, however, it was established<br />

that there are both qualitative and quantitative differences between<br />

alkaline and acid HER catalysis 43 . Whereas in acid solutions the<br />

554 NATURE MATERIALS | VOL 11 | JUNE 2012 | www.nature.com/naturematerials<br />

© 2012 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT3313 ARTICLES<br />

(mV)<br />

@ 5 mA cm ¬2<br />

η<br />

200<br />

400<br />

600<br />

ΔG pt-Had = constant<br />

Pt(111)<br />

Pt(111)–Ni(OH) 2<br />

H2 OH ¬ OH ¬<br />

Pt(111)–Co(OH) 2<br />

H 2 O H 2 O<br />

M<br />

Had Had 2+ δδ O δ(OH) 2¬<br />

M δ<br />

2+ δ O δ(OH)<br />

2¬ δ<br />

Pt<br />

Ni<br />

I (mA cm –2 )<br />

0<br />

–6<br />

Au(111) Co(OH)2<br />

+ Au(111)<br />

Pt(111)<br />

Co(OH) 2<br />

+ Pt(111)<br />

Pt(111)–Fe2+ –10<br />

–0.60 –0.26<br />

E (V)<br />

0<br />

δO δ (OH) 2- δ<br />

Pt(111)–Mn (OH) 2<br />

Co Fe Mn<br />

Figure 4 | Trend in overpotential for the hydrogen evolution reaction (HER) is shown as a function of the 3d transition elements. The elements are<br />

arranged in order of their oxophilicity from Mn to Ni. Pt is shown in the figure as a reference. Top inset: a comparison of polarization curves for Pt(111) and<br />

Au(111) with 40% Co(OH)2 for the HER. As can clearly be seen, the Au(111)/oxide surface is significantly less active than the Pt(111) counterpart. This<br />

essentially establishes the role played by the Pt–Hupd descriptor. On fixing this interaction, by using Pt(111) as the main substrate, we have focused on the<br />

‘pseudo’ mono-functional reaction with the OH ∗ ad –M2+δ interaction as the main descriptor. Bottom inset: a schematic showing the HER. Water from the<br />

bulk of the electrolyte dissociatively adsorbs on the oxide cluster, forming OH ∗ ad intermediate on the oxide cluster, along with forming Had intermediate<br />

formed on the Pt substrate. The Had groups re-combine to form H2. Depending on the OHad–M 2+δ strength, the OH ∗ ad is either stabilized (for Mn2+ , Fe 2+δ )<br />

or destabilized (Ni 2+ , Co 2+ ) on the oxide clusters, which is found to dictate the turnover frequencies for these catalysts.<br />

reaction is controlled mainly by the hydrogen recombination (the<br />

Tafel step), in alkaline solutions the kinetics are determined by<br />

a delicate balance between the water dissociation (the Volmer<br />

step) and concomitant interaction of water dissociation products<br />

with a surface. This was recognized by studying the HER on<br />

the Ni(OH)2/Pt systems 43 , where, in a bi-functional effect, the<br />

edges of Ni(OH)2 clusters promote the dissociation of water<br />

(H2O ↔ H + OH − + e − ). The dissociation step is then followed<br />

by H adsorption on the nearby Pt surfaces (H ↔ Pt–Had) and<br />

by adsorption of OH − on Ni(OH)2 (see bottom inset in Fig. 4).<br />

The kinetics of the HER will depend both on the rate of Had<br />

recombination, which is optimized on the Pt substrate 43 , and on the<br />

rate of desorption of OHad to accommodate the adsorption of H2O<br />

on Ni(OH)2 clusters. Although information regarding kinetics of<br />

the Tafel step on metal surfaces has been studied in great detail 40,41 ,<br />

the importance of OH − ↔ OHad–M 2+δ O δ (OH)2−δ energetic, for<br />

the OH − produced from the water dissociation step, for the<br />

alkaline HER is largely unexplored. The presence of a bi-functional<br />

mode of catalysis for the HER is also confirmed by the observed<br />

enhancement for the HER activities on Co(OH)2/Pt(111) systems<br />

(Supplementary Fig. S5). Further confirmation of the bi-functional<br />

mechanism was achieved by observing a distinct maximum in<br />

the activity versus coverage of Co(OH)2 as well as by comparing<br />

the Co(OH)2/Au(111) systems with their Pt(111) counterparts<br />

(Supplementary Section S4 and Fig. S5). Thus, the overall rate<br />

of the HER may, in principle, be controlled by optimizing the<br />

density and the nature of the sites required for dissociation of<br />

water on M 2+δ O δ (OH)2−δ, as well as the OH − –M 2+δ O δ (OH)2−δ and<br />

metal–Had energetics.<br />

Here, using M 2+δ O δ (OH)2−δ/Pt(111) surfaces, we fix the<br />

descriptor related to the adsorption energetics of Pt–Had, which<br />

in turn lets us treat the HER as a ‘pseudo’ mono-functional<br />

reaction: controlled by the descriptor related to OHad–M 2+δ bond<br />

strength. As summarized in Fig. 4, a monotonic relationship<br />

exists between the HER activity and the OHad–M 2+δ energetics,<br />

with the most active catalysts being Ni(OH)2/Pt(111) and the<br />

least active Mn(OH)2/Pt(111). On the basis of the observed<br />

catalytic trends, it is clear that a balance must be found<br />

between the transition state energies for water dissociation<br />

and the final state energies of adsorbed OHad–M 2+δ O δ (OH)2−δ<br />

(refs 44,45). According to the standard Brønsted–Evans–Polanyitype<br />

principles 46 , this results in lower activation barriers for<br />

water dissociation, while also resulting in poisoning of the sites<br />

required for re-adsorption of water molecules. The net result<br />

is that the turnover frequency substantially decreases (below<br />

bare Pt activity) for the Fe and Mn hydr(oxy)oxides on which<br />

OHad is more strongly adsorbed. In essence, this suggests that<br />

the Fe and Mn behave purely as spectators, blocking the Pt<br />

active sites for transforming H2O to H2. The best combination<br />

among the 3d elements considered here, is found for the<br />

Ni(OH)2/Pt(111), which has the most favourable balance between<br />

facilitating water dissociation and preventing ‘poisoning’ with<br />

OHad (water dissociation product), together with the optimal<br />

Pt–Had energetics. The activity trends derived for the HER using<br />

a series of 3d-M cations with different interaction strength<br />

with OHad clearly establishes the presence of OHad in the HER<br />

region (E < 0 V). Most probably, the active sites for OHad<br />

are the defects, which are known to be very active for water<br />

NATURE MATERIALS | VOL 11 | JUNE 2012 | www.nature.com/naturematerials 555<br />

© 2012 Macmillan Publishers Limited. All rights reserved


ARTICLES<br />

dissociation 45,47 . We emphasize that the exact nature of OHad<br />

(electrosorption valency and free energy of adsorption) on the<br />

M 2+δ O δ (OH)2−δ cluster is not known unambiguously. However,<br />

the fact that the reactivity trends for the HER (Ni > Co ><br />

Fe > Mn), on surfaces with constant Pt–Had interaction, are<br />

identical to the trends in oxophilicity, established from the CO<br />

oxidation reaction, OH − adsorption in the butterfly region and<br />

the OER above 1.6 V, strongly validates the use of OHad–M 2+δ<br />

interaction strength as the descriptor controlling the HER on these<br />

M 2+δ O δ (OH)2−δ/Pt(Au) systems.<br />

In summary, we have used well-characterized M 2+δ O δ (OH)2−δ<br />

/Pt(111) catalyst surfaces (M = Ni, Co, Fe, Mn) to establish<br />

clear trends in activity for the HER and the OER of a complex<br />

oxide system, where the conventional methods of comparing highsurface-area<br />

materials present many uncertainties. We determined,<br />

using the OHad–M 2+δ interaction as the primary descriptor, that<br />

the activity for the HER (bi-functional) and the OER (monofunctional)<br />

for these 3d-M hydr(oxy)oxide systems follows the<br />

order Ni > Co > Fe > Mn. The increasing OER activities for<br />

the 3d-M systems (always greater than Pt), as a function of the<br />

decreasing strength of the OHad–M n+ interaction, provides us<br />

with the necessary toolkit to tune transition metal oxide catalysts<br />

for the OER. We anticipate that by further, descriptor-guided<br />

tuning of these oxides, even greater improvements in alkaline<br />

electrocatalyst performance will be possible for the design of the<br />

elusive ‘active centres’ in catalysis.<br />

Methods<br />

Electrode preparation: extended surface electrode preparation. Pt(111) and<br />

Au(111) electrodes were prepared by inductive heating for 15 min at ∼1,100 K<br />

(1,000 K for Au) in an argon hydrogen flow (3% hydrogen) 35 . The electrode<br />

behaviours are known to be significantly influenced by the preparation procedure;<br />

preparation by inductive heating leads to formation of defects such as ad-islands<br />

on the surface. The number of such defects is low enough to be electrochemically<br />

invisible. The annealed specimens were cooled slowly to room temperature<br />

under an inert atmosphere and immediately covered with a droplet of deionized<br />

(DI) water. Electrodes were then assembled into a rotating disk electrode (RDE)<br />

ensemble. Voltammograms were recorded in argon-saturated electrolytes. The<br />

Ag/AgCl reference electrode was used, but all potentials in the paper are shown<br />

versus the reversible hydrogen electrode (RHE).<br />

Metal hydroxide deposition. All substrates considered in this work (freshly<br />

prepared extended surfaces, ad-island covered surfaces, as well as high-surface-area<br />

catalyst electrodes) were washed thoroughly and introduced into an electrolyte<br />

containing various concentrations of transition metal perchlorates/chlorides.<br />

Concentration ranges tested include 150–1,000 ppm. Following the introduction<br />

of the electrode, the oxide layers were then deposited either potentiostatically<br />

(held at potentials above 0.6 V) or by cycling between Hupd and the OHad regions.<br />

Typical coverages of 30–40% were obtained within 10 min of treatment. For higher<br />

coverages, higher concentrations as well as longer times were used. After deposition,<br />

the electrodes were rinsed and introduced into the clean electrochemical cell. The<br />

Hupd of the modified surface is compared against that of the bare surface to estimate<br />

the effective surface coverage of the oxide species.<br />

Chemicals. All alkali metal hydroxides and perchlorate salts used in our<br />

experiments were obtained in the highest purity from Sigma Aldrich. Electrolytes<br />

used for our experiments, 0.1 M KOH/LiOH, were prepared with Millipore DI<br />

water. All gases (argon, oxygen, hydrogen) were of 5N5 quality purchased from<br />

Airgas. A typical three-electrode fluoro ethylene propylene (FEP) polymer based<br />

cell was used to avoid contamination from glass components. Experiments were<br />

controlled using an Autolab PGSTAT 302N potentiostat. Gold or platinum<br />

wires were used as counter electrodes for studying hydrogen evolution reaction.<br />

Precautions were taken to prevent significant accumulation of dissolved counter<br />

electrode ions near the working electrode.<br />

Electrochemical measurements. After extensive rinsing, the electrode was<br />

embedded into the RDE and transferred into a standard three-compartment<br />

electrochemical cell containing 0.1 M KOH/LiOH (Sigma-Aldrich). In each<br />

experiment, the electrode was immersed at 0.05 V in a solution saturated with<br />

argon. After obtaining a stable cycle between 0.05 and 0.7 V the electrolyte<br />

was saturated with H2, following which polarization curves for the HER were<br />

recorded on the disc electrodes between 0.05 V and −0.4 V. The lower potential<br />

limits were chosen so as to avoid significant bubble formation, as well as to<br />

NATURE MATERIALS DOI: 10.1038/NMAT3313<br />

minimize the extended dissolution of the counter electrode. The concentration<br />

of Pt after 1 h of HER measurements was found to be less than 1 ppm in the<br />

working electrode compartment. All polarization curves were corrected for<br />

the infrared contribution within the cell. OER measurements were carried out<br />

by cycling the electrode up to 1.7 V versus RHE. Potential hold experiments<br />

were also carried out for the oxide/Pt(111) systems to study the stability of the<br />

oxide clusters at these potentials, as well as for preparation of samples for the<br />

STM measurements.<br />

STM method. For the as-prepared and the modified surfaces, the STM images<br />

were acquired with a Digital Instruments Multi-Mode Dimension STM controlled<br />

by a Nanoscope III control station. During the measurement, the microscope with<br />

the sample was enclosed in a pressurized cylinder with a CO atmosphere. For<br />

further details, see ref. 35.<br />

XAS/XANES measurements. The X-ray absorption spectroscopy (XAS) data<br />

were acquired at bending magnet beamline 12-BM-B at the Advanced Photon<br />

Source (APS), Argonne National Laboratory. The synchrotron radiation was<br />

filtered by a double crystal Si(111) monochromator with a double mirror system<br />

for focusing and harmonic re<strong>je</strong>ction. A custom-made in situ transmission<br />

electrochemical X-ray cell with a 6 mm diameter Pt(111) single crystal and Ag/AgCl<br />

reference electrode was used in a grazing incidence geometry. A 13-element Ge<br />

detector (CANBERRA) was used to measure the fluorescence yield. Z-1 filters<br />

and grazing incidence geometry was used to minimize the elastic scattering<br />

intensity. The monochrometer calibration was monitored by simultaneously<br />

measuring the same element reference foil in front of a Si diode and looking at<br />

the air-scattered beam.<br />

Received 14 November 2011; accepted 23 March 2012;<br />

published online 6 May 2012<br />

References<br />

1. <strong>Dr</strong>esselhaus, M. S. & Thomas, I. L. Alternative energy technologies. Nature<br />

414, 332–337 (2001).<br />

2. Gratzel, M. Photoelectrochemical cells. Nature 414, 338–344 (2001).<br />

3. Schlapbach, L. & Zuttel, A. Hydrogen-storage materials for mobile applications.<br />

Nature 414, 353–358 (2001).<br />

4. Steele, B. C. H. & Heinzel, A. Materials for fuel-cell technologies. Nature 414,<br />

345–352 (2001).<br />

5. Tarascon, J. M. & Armand, M. Issues and challenges facing rechargeable<br />

lithium batteries. Nature 414, 359–367 (2001).<br />

6. Lefèvre, M., Proietti, E., Jaouen, F. & Dodelet, J-P. Iron-based catalysts with<br />

improved oxygen reduction activity in polymer electrolyte fuel cells. Science<br />

324, 71–74 (2009).<br />

7. Gasteiger, H. A. & Markovi, N. M. Just a dream—or future reality? Science 324,<br />

48–49 (2009).<br />

8. Lasia, A. in Handbook of Fuel Cells: Fundamentals, Technology and Applications<br />

Vol. 2 (eds Vieistich, W., Lamm, A. & Gasteiger, H. A.) 416 (Wiley, 2003).<br />

9. Moorhouse, J. (ed.) Modern Chlor-Alkali Technology (Wiley, 2001).<br />

10. Hoare, J. P. The Electrochemistry of Oxygen (Interscience, 1968).<br />

11. Kinoshita, K. & Society, E. Electrochemical Oxygen Technology (Wiley, 1992).<br />

12. Birry, L. & Lasia, A. Studies of the hydrogen evolution reaction on<br />

Raney nickel—molybdenum electrodes. J. Appl. Electrochem. 34,<br />

735–749 (2004).<br />

13. Lasia, A. & Rami, A. Kinetics of hydrogen evolution on nickel electrodes.<br />

J. Electroanal. Chem. Interfacial Electrochem. 294, 123–141 (1990).<br />

14. Birss, V. I. & Damjanovic, A. Oxygen evolution at platinum electrodes in<br />

alkaline solutions. J. Electrochem. Soc. 134, 113–117 (1987).<br />

15. Ardizzone, S., Fregonara, G. & Trasatti, S. ‘Inner’ and ‘outer’ active surface of<br />

RuO2 electrodes. Electrochim. Acta 35, 263–267 (1990).<br />

16. Lyons, M. E. G. & Burke, L. D. Mechanism of oxygen reactions at porous oxide<br />

electrodes. Part 1.—Oxygen evolution at RuO2 and RuxSn1–xO2 electrodes in<br />

alkaline solution under vigorous electrolysis conditions. J. Chem. Soc. Faraday<br />

Trans. 1 83, 299–321 (1987).<br />

17. Trasatti, S. Electrodes of Conductive Metallic Oxides (Elsevier, 1980).<br />

18. Man, I. C. et al. Universality in oxygen evolution electrocatalysis on oxide<br />

surfaces. ChemCatChem 3, 1159–1165 (2011).<br />

19. Sergio, T. Physical electrochemistry of ceramic oxides. Electrochim. Acta 36,<br />

225–241 (1991).<br />

20. Lyons, M. E. G. & Brandon, M. P. A comparative study of the oxygen evolution<br />

reaction on oxidised nickel, cobalt and iron electrodes in base. J. Electroanal.<br />

Chem. 641, 119–130 (2010).<br />

21. Jaramillo, T. F. et al. Identification of active edge sites for electrochemical H2<br />

evolution from MoS2 nanocatalysts. Science 317, 100–102 (2007).<br />

22. Kanan, M. W. & Nocera, D. G. In situ formation of an oxygen-evolving<br />

catalyst in neutral water containing phosphate and Co 2+ . Science 321,<br />

1072–1075 (2008).<br />

23. Bockris, J. O. M. & Otagawa, T. The electrocatalysis of oxygen evolution on<br />

perovskites. J. Electrochem. Soc. 131, 290–302 (1984).<br />

556 NATURE MATERIALS | VOL 11 | JUNE 2012 | www.nature.com/naturematerials<br />

© 2012 Macmillan Publishers Limited. All rights reserved


NATURE MATERIALS DOI: 10.1038/NMAT3313 ARTICLES<br />

24. Suntivich, J., May, K. J., Gasteiger, H. A., Goodenough, J. B. & Shao-Horn, Y.<br />

A perovskite oxide optimized for oxygen evolution catalysis from molecular<br />

orbital principles. Science 334, 1383–1385 (2011).<br />

25. Russell, A. E. & Rose, A. X-ray absorption spectroscopy of low temperature<br />

fuel cell catalysts. Chem. Rev. 104, 4613–4636 (2004).<br />

26. Totir, D., Mo, Y., Kim, S., Antonio, M. R. & Scherson, D. A. In situ Co K-edge<br />

X-ray absorption fine structure of cobalt hydroxide film electrodes in alkaline<br />

solutions. J. Electrochem. Soc. 147, 4594–4597 (2000).<br />

27. Pourbaix, M. in Atlas of Electrochemical Equilibria in Aqueous Solutions<br />

(ed. Pourbaix, M.) 644 (NACE, 1974).<br />

28. Campbell, C. T. Bimetallic surface chemistry. Annu. Rev. Phys. Chem. 41,<br />

775–837 (1990).<br />

29. Clavilier, J., Faure, R., Guinet, G. & Durand, R. Preparation of monocrystalline<br />

Pt microelectrodes and electrochemical study of the plane surfaces cut in<br />

the direction of the {111} and {110} planes. J. Electroanal. Chem. Interfacial<br />

Electrochem. 107, 205–209 (1979).<br />

30. Markovi, N. M. & Ross, P. N. Jr Surface science studies of model fuel cell<br />

electrocatalysts. Surf. Sci. Rep. 45, 117–229 (2002).<br />

31. Strmcnik, D. et al. Effects of Li + , K + , and Ba 2+ cations on the ORR at model and<br />

high surface area Pt and Au surfaces in alkaline solutions. J. Phys. Chem. Lett.<br />

2, 2733–2736 (2011).<br />

32. Ahmed, M. et al. Unprecedented structural sensitivity toward average<br />

terrace width: Nafion adsorption at Pt{hkl} electrodes. J. Phys. Chem. C 115,<br />

17020–17027 (2011).<br />

33. Van der Niet, M. J. T. C., den Dunnen, A., Juurlink, L. B. F. & Koper, M. T. M.<br />

Co-adsorption of O and H2O on nanostructured platinum surfaces: Does OH<br />

form at steps? Angew. Chem. Int. Ed. 122, 6572–6575 (2010).<br />

34. Marković, N. M. et al. Effect of temperature on surface processes at the<br />

Pt(111)-liquid interface:? Hydrogen adsorption, oxide formation, and CO<br />

oxidation? J. Phys. Chem. B 103, 8568–8577 (1999).<br />

35. Strmcnik, D. S. et al. Unique activity of platinum adislands in the CO<br />

electrooxidation reaction. J. Am. Chem. Soc. 130, 15332–15339 (2008).<br />

36. Schmidt, T. J., Ross, P. N. & Markovic, N. M. Temperature-dependent surface<br />

electrochemistry on Pt single crystals in alkaline electrolyte: Part 1: CO<br />

oxidation. J. Phys. Chem. B 105, 12082–12086 (2001).<br />

37. Markovic, N. R. & Ross, P. N. New electrocatalysts for fuel cells from model<br />

surfaces to commercial catalysts. Cattech 4, 110–126 (2000).<br />

38. Rossmeisl, J., Qu, Z. W., Zhu, H., Kroes, G. J. & Nørskov, J. K. Electrolysis of<br />

water on oxide surfaces. J. Electroanal. Chem. 607, 83–89 (2007).<br />

39. Norskov, J. K., Bligaard, T., Rossmeisl, J. & Christensen, C. H. Towards the<br />

computational design of solid catalysts. Nature Chem. 1, 37–46 (2009).<br />

40. Conway, B. E. & Tilak, B. V. Interfacial processes involving electrocatalytic<br />

evolution and oxidation of H2, and the role of chemisorbed H. Electrochim. Acta<br />

47, 3571–3594 (2002).<br />

41. Markovic, N. M., Sarraf, S. T., Gasteiger, H. A. & Ross, P. N. Hydrogen<br />

electrochemistry on platinum low-index single-crystal surfaces in alkaline<br />

solution. J. Chem. Soc. Faraday Trans. 92, 3719–3725 (1996).<br />

42. Greeley, J., Jaramillo, T. F., Bonde, J., Chorkendorff, I. & Norskov, J. K.<br />

Computational high-throughput screening of electrocatalytic materials for<br />

hydrogen evolution. Nature Mater. 5, 909–913 (2006).<br />

43. Subbaraman, R. et al. Enhancing hydrogen evolution activity in<br />

water splitting by tailoring Li + /Ni(OH)2/Pt interfaces. Science 334,<br />

1256–1260 (2011).<br />

44. Henrich, V. E. & Cox, P. A. The Surface Science of Metal Oxides (Cambridge<br />

Univ. Press, 1994).<br />

45. Henderson, M. A. The interaction of water with solid surfaces: Fundamental<br />

aspects revisited. Surf. Sci. Rep. 46, 1–308 (2002).<br />

46. Bligaard, T. et al. The Brønsted–Evans–Polanyi relation and the volcano curve<br />

in heterogeneous catalysis. J. Catal. 224, 206–217 (2004).<br />

47. Thiel, P. A. & Madey, T. E. The interaction of water with solid surfaces:<br />

Fundamental aspects. Surf. Sci. Rep. 7, 211–385 (1987).<br />

48. Kim, M-S. & Kim, K-B. A study on the phase transformation of<br />

electrochemically precipitated nickel hydroxides using an electrochemical<br />

quartz crystal microbalance. J. Electrochem. Soc. 145, 507–511 (1998).<br />

Acknowledgements<br />

Supported by the Office of Science, Office of Basic Energy Sciences, Division of<br />

Materials Sciences, US Department of Energy, under contract DE-AC02-06CH11357.<br />

R.S. would like to acknowledge the Argonne National Laboratory post-doctoral<br />

fellowship for his funding.<br />

Author contributions<br />

R.S. and N.M.M. developed the idea and designed the experiments. R.S., D.T., K.C.C. and<br />

A.P.P. performed the experiments and data analyses. R.S., N.M.M., D.S., K.C.C., J.G. and<br />

V.S. discussed the results. R.S. and N.M.M. co-wrote the paper.<br />

Additional information<br />

The authors declare no competing financial interests. Supplementary information<br />

accompanies this paper on www.nature.com/naturematerials. Reprints and permissions<br />

information is available online at www.nature.com/reprints. Correspondence and<br />

requests for materials should be addressed to N.M.M.<br />

NATURE MATERIALS | VOL 11 | JUNE 2012 | www.nature.com/naturematerials 557<br />

© 2012 Macmillan Publishers Limited. All rights reserved


ARTICLES<br />

PUBLISHED ONLINE: 11 NOVEMBER 2012 | DOI: 10.1038/NMAT3457<br />

Mesostructured thin films as electrocatalysts with<br />

tunable composition and surface morphology<br />

Dennis F. van der Vliet 1† , Chao Wang 1† , Dusan Tripkovic 1 , Dusan Strmcnik 1 , Xiao Feng Zhang 2 ,<br />

Mark K. Debe 3 , Radoslav T. Atanasoski 3 , Nenad M. Markovic 1 and Vojislav R. Stamenkovic 1 *<br />

Among the most challenging issues in technologies for electrochemical energy conversion are the insufficient activity of the<br />

catalysts for the oxygen reduction reaction, catalyst degradation and carbon-support corrosion. In an effort to address these<br />

barriers, we aimed towards carbon-free multi/bimetallic materials in the form of mesostructured thin films with tailored<br />

physical properties. We present here a new class of metallic materials with tunable near-surface composition, morphology<br />

and structure that have led to greatly improved affinity for the electrochemical reduction of oxygen. The level of activity for<br />

the oxygen reduction reaction established on mesostructured thin-film catalysts exceeds the highest value reported for bulk<br />

polycrystalline Pt bimetallic alloys, and is 20-fold more active than the present state-of-the-art Pt/C nanoscale catalyst.<br />

Over the past decades, extensive research has been devoted to<br />

the development of technologies that can effectively convert<br />

energy and become economically viable for use by the<br />

general public. Great expectations are held for technologies such<br />

as fuel cells and lithium–air batteries that rely on electrochemical<br />

processes. In both cases, satisfactory energy density can be attained;<br />

however, a major challenge lies in the insufficient activity and<br />

durability of the materials that are employed at present as cathode<br />

catalysts for electrochemical reduction of oxygen. These limitations<br />

inevitably lead to a lower operating efficiency of the devices,<br />

which highlights the need for the development of more active<br />

and durable oxygen reduction reaction (ORR) catalysts 1–12 . In the<br />

case of fuel cells, most of the research centres on platinum, the<br />

best monometallic catalyst for the ORR. At the present state of<br />

development, an approximately fivefold reduction in Pt content<br />

is necessary to meet cost requirements for large-scale automotive<br />

applications 5 . Pt-based alloys have already made an impact in<br />

fuel-cell catalyst design by decreasing the amount of platinum while<br />

improving activity and durability 12–18 , which places these materials<br />

at the focus of intensive fundamental and applied research on both<br />

extended (bulk) 17–25 and nanoscale systems 7,14–16,26–34 . The main<br />

challenge in that effort is linked to the possibility of achieving<br />

the unique structural and compositional profile of the Pt3Ni(111)<br />

alloy, which was established from single-crystal studies 12,17 . Such<br />

a profile was obtained on extended surfaces by thermal annealing<br />

that facilitates thermodynamically driven segregation of Pt to form a<br />

pure ordered surface layer, denoted as Pt(111)-skin. The electronic<br />

structure of Pt(111)-skin is altered by the subsurface layer of PtNi<br />

(in 1:1 ratio) and is responsible for the extreme ORR activity, which<br />

is nearly two orders of magnitude higher than the state-of-the-art<br />

Pt/C catalyst. Consequently, the ability to mimic the compositional<br />

profile and structure of Pt-skin in high-surface-area catalysts would<br />

bring unprecedented benefits to technologies that rely on the<br />

ORR. However, despite numerous attempts, this goal has not been<br />

achieved yet for practical catalysts.<br />

Here we present a new class of materials based on mesostructured<br />

multimetallic thin films with adjustable structure and<br />

composition, which have been tailored to emulate the distinctive<br />

properties of the Pt(111)-skin, to be employed in electrochemical<br />

devices. The design of these materials relied on our previous work<br />

related to well-defined extended and nanoscale surfaces in the form<br />

of PtM alloys (M = Ni, Co, Fe, V, Ti; refs 17–20). We aimed<br />

towards catalysts that can bridge the world of extended surfaces<br />

with superior activity and nanoscale systems with high specific<br />

surface area, to harvest maximal utilization of precious metals. Such<br />

synergy is foreseen to be present at the mesoscale, which implies not<br />

only a specific length scale, but rather a principle of operating in<br />

between different physical regimes that exhibit distinct functional<br />

behaviour 35 . Considering that mesoscale materials chemistry is still<br />

in its infancy, it is expected that this field will open pathways in<br />

materials design that arise from the rational control of physicochemical<br />

properties and functionality of mesostructured systems.<br />

In particular, for electrocatalytic materials, most previous work<br />

has emphasized either achievement of high surface area through<br />

small particle size, or the attainment of a better understanding of<br />

fundamental properties through the use of extended surfaces. From<br />

such studies, it is well known that there are substantial differences<br />

in catalytic properties between nanoscale and bulk materials. The<br />

benefits of targeting mesoscale architectures between these extremes<br />

have scarcely been explored, especially in the sense of transferring<br />

superior characteristics from extended surfaces to practical<br />

materials. In view of that, instead of using discrete nanoparticles<br />

(3–5 nm) supported on high-surface-area carbon 26–28,32–34,36 , we deployed<br />

continuous Pt and Pt-alloy nanostructured thin films 29,37–41<br />

(NSTF) over an oriented array of molecular solid whiskers by<br />

physical vapour deposition. Specifically, planar magnetron sputter<br />

deposition was used to deposit thin metal films with a wide range<br />

in composition. Such NSTF catalysts provide good surface area<br />

utilization and eliminate issues related to carbon-support corrosion<br />

and contact resistance at the carbon/metal interface that would lead<br />

to poor utilization and degradation of the catalyst 14 . The capability<br />

to control the deposition rate, as well as the combination and<br />

order of constituents, signifies sputter deposition as an effective<br />

tool to form thin films with desirable thickness, composition profile<br />

1 Materials Science Division, Argonne National Laboratory, Argonne, Illinois 60439, USA, 2 Hitachi High Technologies America, Pleasanton, California<br />

94588, USA, 3 Fuel Cell Components Program, 3M, St Paul, Minnesota 55144, USA. † These authors contributed equally to this work.<br />

*e-mail: vrstamenkovic@anl.gov.<br />

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials 1<br />

© 2012 Macmillan Publishers Limited. All rights reserved.


ARTICLES<br />

c<br />

a b<br />

I (μA)<br />

I (mA)<br />

20<br />

0<br />

¬20<br />

0.02<br />

0.00<br />

¬0.02<br />

0.2<br />

Pt as deposited<br />

Pt (111)<br />

0.4 0.6 0.8<br />

E (V)<br />

Pt 3 Ni (111)<br />

As-deposited PtNi/GC<br />

Annealed PtNi/GC<br />

5 nm<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

E (V versus RHE)<br />

I (μA)<br />

Specific activity (mA cm ¬2 pt)<br />

8<br />

6<br />

4<br />

2<br />

0<br />

NATURE MATERIALS DOI: 10.1038/NMAT3457<br />

Pt thin film annealed<br />

Pt (111)<br />

0.2 0.4 0.6<br />

E (V)<br />

0.8<br />

Pt 3 Ni(111)-skin<br />

Pt poly<br />

As-deposited<br />

Annealed<br />

ΔT<br />

Polycrystalline<br />

ordering<br />

Pt thin film<br />

on GC<br />

(111)<br />

Polycrystalline<br />

(111)<br />

ordering<br />

5 nm<br />

ΔT<br />

PtNi thin film<br />

on GC<br />

Figure 1 | CV and STM images of Pt and Pt alloy 20-nm thin films deposited on a glassy carbon substrate. a, As-deposited Pt thin film (solid line) and<br />

Pt(111) (dashed line). b, Annealed Pt thin film (solid line) and Pt(111) (dashed line). c, CV profile of as-deposited (blue line), annealed PtNi thin film (red<br />

line) and Pt3Ni(111)-skin (dashed line). d, Specific activities measured by RDE in 0.1 M HClO4 with 1,600 rpm, 20 mV s −1 at 0.95 V with corresponding<br />

improvements factors versus polycrystalline Pt.<br />

and surface roughness. However, the surface structure of a thin<br />

film, an important catalytic parameter 33,42–44 , cannot be altered<br />

by this method to match those established on single-crystalline<br />

systems. For that reason, we attempted a thorough examination of<br />

thin-film properties on extended, flat, non-crystalline and chemically<br />

inert substrates such as a mirror-polished glassy carbon<br />

surface. This approach brings an extra level of control in terms of<br />

defined geometric surface area and surface roughness factor that is<br />

unattainable in the case of nanoscale substrates. Consequently, our<br />

efforts have been directed towards exploring structural transitions<br />

in polycrystalline thin films.<br />

The first step comprised the deposition of a pure Pt thin film<br />

onto an ultrahigh-vacuum-cleaned glassy carbon substrate, which<br />

was followed by thermal annealing in a reductive atmosphere<br />

(see Methods). The morphology of the Pt film was validated by<br />

scanning tunnelling microscopy (STM) as shown in Fig. 1a,b. The<br />

difference between as-deposited versus annealed Pt films indicates a<br />

substantial change in the thin-film surface morphology due to rearrangement<br />

of the Pt topmost atoms towards the (111) structure with<br />

minimum surface energy. The as-deposited Pt film has a corrugated<br />

nanostructured three-dimensional surface morphology with an average<br />

grain size of ∼5 nm, whereas the morphology of the annealed<br />

thin film has been transformed into a smooth two-dimensional surface<br />

with large 20 × 100 nm hexagonal (111) facets. In accordance<br />

with the STM results, the characteristic surface features are also<br />

confirmed by electrochemical cyclic voltammetry (CV). Figure 1a<br />

reveals that the CV profile of the as-deposited thin-film surface<br />

d<br />

20<br />

0<br />

¬20<br />

matches the one established for bulk polycrystalline Pt. On the other<br />

hand, Fig. 1b shows that the CV profile of the annealed Pt thin film<br />

underwent extensive transformation from typical polycrystalline<br />

into Pt(111)-like with characteristic fingerprint features between<br />

0.5 and 0.9 V; the so-called butterfly region that corresponds to<br />

adsorption/desorption processes of OHad on Pt(111) facets (see<br />

Supplementary Information). Therefore, it is evident from both<br />

STM and CV that the annealed extended thin film consists of<br />

predominantly (111) facets encompassing the entire surface. In fact,<br />

the degree of resemblance in electrochemical signature between the<br />

annealed thin-film surface and single-crystal Pt confirms that the<br />

(111) facets are both large and interconnected. The synergy between<br />

the surface structure, domain size and functionality defines that the<br />

thin-film surface has a distinct mesostructured morphology. These<br />

findings clearly demonstrate the feasibility of controlling surface<br />

ordering of extended Pt thin films deposited over a non-crystalline<br />

substrate, that is, without the use of templates for epitaxial growth.<br />

Instead of building the crystal lattice from a seed or underlying<br />

crystalline substrate, individual randomly oriented nanoscale grains<br />

coalesce and form large well-ordered (111) facets. All of that greatly<br />

expands the potential for utilization of thin-film materials and<br />

introduces thermal annealing in a controlled atmosphere as a<br />

compelling tool in the fine tuning of a thin film’s structure and<br />

hence electrocatalytic properties.<br />

In the following steps we proceed towards a bimetallic PtNi thin<br />

film with the same thickness to mimic the composition profile of the<br />

Pt3Ni(111) system and to replicate its unique catalytic properties.<br />

2 NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials<br />

© 2012 Macmillan Publishers Limited. All rights reserved.<br />

20<br />

16<br />

12<br />

8<br />

4<br />

Improvement factor versus Pt-poly


NATURE MATERIALS DOI: 10.1038/NMAT3457 ARTICLES<br />

a b<br />

z<br />

y<br />

x<br />

50 nm<br />

Pt M N<br />

Substrate growth Physical vapour deposition Nanostructured thin films<br />

c d<br />

e<br />

15 nm 10 nm 10 nm<br />

Figure 2 | HRSEM and TEM micrographs of the NSTF whiskers. Centre: schematic illustration of the vacuum protocol that has been used for the growth of<br />

aligned perylene red substrate, which is then coated by a metallic thin film with an adjustable thickness and composition profile (Pt; M and/or<br />

N = Ni,Co,Fe,Ti,V). a, HRSEM snapshot of a group of whiskers that indicates their length, shape and alignment after thin-film deposition. b, HRSEM<br />

close-up of an intentionally broken single whisker that demonstrates the thickness of the metallic film over perylene red substrate. c, HRTEM close-up of a<br />

single whisker side, which reveals growth of whiskerettes along the whisker. d, HRSEM insight into the whisker’s surface showing a close-packed formation<br />

of whiskerette tips of 5 nm in diameter that facilitates a highly corrugated morphology. e, TEM micrograph of a whisker side that confirms the grained<br />

texture of the sputtered thin film and the average diameter of the whiskerettes.<br />

The results from the electrochemical measurements in Fig. 1c,d<br />

confirm that as for monometallic Pt, the polycrystalline nature<br />

of the as-deposited alloy thin film is predominantly transformed<br />

into a Pt(111)-skin-like surface. This is obvious from both the<br />

CV profile of the annealed alloy thin film that resembles the<br />

one obtained on Pt3Ni(111) and its superior catalytic activity<br />

for the ORR, which was up to now obtained exclusively on the<br />

Pt3Ni(111)-skin surface 17,45 . The combination of the Pt(111)-skinlike<br />

voltammetry and the marked increase in the ORR activity<br />

proves that surface ordering from randomly oriented towards (111)<br />

is indeed feasible for bimetallic thin films, and demonstrates that<br />

the catalytic improvement obeys the same mechanism as previously<br />

reported for Pt-bimetallic single-crystal surfaces; that is, electronic<br />

modification of the topmost Pt layer leads to extreme catalytic<br />

enhancement solely for the (111) orientation 19 . Therefore, the<br />

ORR-specific activity, which equals 70% of the value established<br />

for the most active catalyst, Pt3Ni(111)-skin, serves as a descriptor<br />

that (111)-skin facets are dominating on the annealed thin-film<br />

surface. Together this demonstrates the twofold power of annealing<br />

in facilitating the formation of the mesostructured alloy thinfilm<br />

morphology, characterized by both an energetically more<br />

favourable surface state rich in (111) facets, and the desired<br />

compositional profile.<br />

It is these results that provide the driving force for a shift towards<br />

corresponding thin-film-based high-surface-area materials.<br />

A Pt-alloy NSTF catalyst is deposited by magnetron sputtering<br />

20 nm<br />

over an array of molecular solid whiskers, composed of an organic<br />

pigment N , N -di(3,5-xylyl)perylene-3,4:9,10 bis(dicarboximide);<br />

hereafter denoted as perylene red 29,37–39,46 . Figure 2 illustrates the<br />

step-by-step deposition process of the thin metal films onto the<br />

perylene red support, as well as high-resolution scanning electron<br />

microscopy (HRSEM) and transmission electron microscopy<br />

(TEM) micrographs of the NSTF whiskers. These images reveal<br />

a detailed insight into critical parameters of the NSTF such as<br />

metallic film thickness, length, shape and surface morphology. A<br />

single whisker measures on average about 800 nm in length, and<br />

the film thickness is 5–20 nm. In Fig. 2, it is clearly visible that<br />

on the sides of a single whisker, smaller metal alloy whiskerettes<br />

are formed with a diameter of ∼5 nm (Fig. 2c), and a close-up<br />

of a broken whisker (Fig. 2b) illustrates the metallic film/shell<br />

that surrounds the perylene red substrate. Surface-specific HRSEM<br />

in Fig. 2b,d unveils that the side walls along the whisker have a<br />

very rough surface morphology, consisting mainly of whiskerette<br />

tips bonded closely to each other to produce densely packed<br />

corncob-like features, providing the validity of terming this material<br />

a NSTF. It is important to emphasize that the highly grained texture<br />

of the NSTF side walls made of closely packed whiskerettes is<br />

also confirmed in the TEM micrograph in Fig. 2e. Such structural<br />

parameters greatly affect the functional properties of the NSTF,<br />

and therefore the ability to control and tune them along with the<br />

near-surface compositional profile can lead towards a substantial<br />

gain in catalytic performance.<br />

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials 3<br />

© 2012 Macmillan Publishers Limited. All rights reserved.


ARTICLES<br />

a<br />

b<br />

c<br />

Nanostructured<br />

thin films<br />

Surface<br />

modification<br />

and<br />

substrate<br />

evaporation<br />

Mesostructured<br />

thin films<br />

30 nm<br />

20 nm<br />

15 nm<br />

20 nm<br />

20 nm<br />

20 nm<br />

NATURE MATERIALS DOI: 10.1038/NMAT3457<br />

Figure 3 | In situ transformation from nanostructured into mesostructured thin film during annealing. a–c, Middle: HRTEM images of progressive<br />

annealing between room temperature and 400 ◦ C on a single whisker that capture the ordering from randomly oriented to a homogeneous structure with<br />

visible crystalline domains. Right: HRTEM close-up of the transformation for the same near-surface region during annealing, from highly corrugated into a<br />

flat and smooth surface morphology. Left: schematic illustration of the mesoscale ordering during annealing and formation of the mesostructured thin film<br />

catalyst, associated with the corresponding HRSEM insets.<br />

In what follows, we apply an experimental approach combined<br />

with the knowledge related to highly active well-defined singlecrystalline<br />

and extended thin-film surfaces to develop mesostructured<br />

thin-film electrocatalysts with advanced properties. In situ<br />

HRSEM and TEM are simultaneously employed during NSTF<br />

annealing in a controlled atmosphere. This allows us to visualize<br />

real-time structural changes at the atomic level and to follow<br />

rearrangements of the surface and sub-surface morphology of<br />

thin-film materials. This insight is invaluable in the fine-tuning of<br />

the materials’ properties. Figure 3 depicts in situ results obtained<br />

during thermal annealing of a single PtNi-NSTF whisker. The NSTF<br />

catalyst is mounted onto the HRTEM heating stage and is introduced<br />

to a reductive atmosphere of argon and hydrogen gases. As<br />

the specimen is gradually heated, no change in surface morphology<br />

is observed as depicted in Fig. 3a, which reveals the initial stage and<br />

a close-up of the grained highly corrugated whisker side wall and<br />

its surface. Once the temperature reaches 300 ◦ C, we start to follow<br />

real-time restructuring of the thin film’s morphology. Figure 3b<br />

captures the onset of the surface transformation, which appears as a<br />

smoothening of the near-surface regions. The steady-state structure<br />

is achieved after 30 min and is shown in Fig. 3c. These images<br />

illustrate that the densely packed organization with the initial<br />

three-dimensional surface morphology is being transformed into<br />

a more homogeneous, flat and ordered two-dimensional thin-film<br />

material with clearly observable crystalline features in its walls.<br />

This thermodynamically driven transition releases stress and strain<br />

of the as-deposited thin film and leads towards the state with<br />

minimum surface energy without compromising the overall shape<br />

and dimension of the whisker. As for Pt thin films on glassy carbon,<br />

the initial nanostructured surface morphology that originated from<br />

× 10<br />

× 10<br />

× 10<br />

2 nm<br />

2 nm<br />

2 nm<br />

the closely bonded whiskerettes’ tips is transformed into a smooth<br />

continuous film with large crystalline domains (20–40 nm). Specifically,<br />

randomly oriented nanoscale grains coalesce and give rise to a<br />

mesostructured thin film with unique physicochemical properties;<br />

therefore, the materials after this treatment will be referred to as<br />

mesostructured thin films (Meso-TF). Close inspection of HRTEM<br />

micrographs after applied thermal treatment (see Supplementary<br />

Information) confirms that emerged facets with (111) structure<br />

prevail on the surface whereas undercoordinated sites are diminished,<br />

which also has important implications towards improved<br />

stability. As a side effect, the perylene red substrate is removed<br />

during this procedure (details can be found in the Supplementary<br />

Information). In addition to HRTEM/SEM studies, X-ray diffraction<br />

measurements, which show enhanced alloying and an increase<br />

in the number of (111)-oriented domains on the Meso-TF, are<br />

presented in the Supplementary Information.<br />

The final step in the characterization is to obtain the electrochemical<br />

signature and compare adsorption and catalytic properties<br />

between different classes of thin-film materials and the state-ofthe-art<br />

Pt/C catalyst by rotating-disc electrode (RDE), see Methods<br />

section. As expected, from the CV profile depicted in Fig. 4 we find<br />

that the smooth morphology of the Meso-TF slightly lowers the<br />

electrochemically active surface area (ECSA), from ∼11 m 2 g −1 for<br />

the NSTF to ∼9 m 2 g −1<br />

Pt for the Meso-TF. This implies that most of<br />

the inner portion of the whiskers, which has been vacated by the<br />

perylene red, is not electrochemically active, presumably owing to<br />

lack of penetration of the electrolyte into the hollow of the whisker<br />

(see closed whisker end in Fig. 3). As shown in Fig. 4a, the CV profile<br />

of PtNi NSTF whiskers exhibits similar behaviour to monometallic<br />

Pt NSTF with clearly visible polycrystalline Pt features due to<br />

4 NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials<br />

© 2012 Macmillan Publishers Limited. All rights reserved.


NATURE MATERIALS DOI: 10.1038/NMAT3457 ARTICLES<br />

a c<br />

I (mA cm ¬2 )<br />

I (mA)<br />

0.4<br />

0.0<br />

¬0.4<br />

¬0.8<br />

b<br />

PtNi Meso-TF<br />

d<br />

0.0<br />

Pt NSTF<br />

¬0.5<br />

¬1.0<br />

¬1.5<br />

PtNi Meso-TF<br />

PtNi NSTF<br />

Pt NSTF<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

E (V versus RHE)<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

E (V versus RHE)<br />

E (V versus RHE)<br />

i kin (mA cm pt ¬2 )<br />

1.00<br />

0.95<br />

PtNi Meso-TF<br />

PtNi NSTF<br />

Pt NSTF<br />

Pt/C<br />

0.90<br />

0.01 0.1<br />

ikin (mA cm ¬2 )<br />

2.4<br />

1.8<br />

1.2<br />

0.6<br />

Pt/C Ptpoly<br />

Pt-<br />

NSTF<br />

ordering<br />

Nano Meso<br />

PtNi-<br />

NSTF<br />

1 2 3<br />

ΔT<br />

PtNi-<br />

Meso-TF<br />

Figure 4 | CV on the Pt-based thin-film catalysts. a, Cyclic voltammograms of Pt-NSTF, PtNi-NSTF and PtNi-Meso-TF. b, The ORR polarization curves.<br />

c, Corresponding Tafel plots (Tafel slopes are determined at potentials higher than the half-wave potential (E1/2; potential at which I = 1/2 Idiff), to avoid<br />

diffusion- and solution-resistance-induced errors). d, Specific activities measured at 0.95 V and improvement factor versus Pt-poly (and Pt-NSTF).<br />

the adsorption–desorption processes of underpotentially deposited<br />

hydrogen (Hupd). However, the Hupd region of PtNi Meso-TF is<br />

significantly different with a characteristic flat plateau (Fig. 4a),<br />

which confirms that the surface has a relatively large contribution<br />

of (111) facets compared with the highly corrugated sputtered<br />

thin film that is rich in low-coordinated Pt sites. This is also<br />

in good agreement with HRTEM and X-ray diffraction results.<br />

Moreover, the onset of surface oxide formation is shifted positively<br />

in the following order: Pt-NSTF < PtNi-NSTF < PtNi Meso-TF.<br />

Accordingly, the ORR polarization curves, shown in Fig. 4b, follow<br />

the same trend in activity. Figure 4c and Fig. 4d summarize the<br />

kinetic current densities (specific activities per ECSA of Pt) as<br />

Tafel plots and a bar graph, respectively. As specific activity is a<br />

fundamental property of a material that reflects its intrinsic catalytic<br />

performance, as opposed to mass activity, which emphasizes the<br />

optimized dispersion of a material, our focus has been placed<br />

on boosting specific activity. This approach leads to a higher<br />

turnover frequency (the measure of activity per active site), which<br />

may result in better utilization of Pt, culminating in higher mass<br />

activity. Considering the large increase in specific activity, we report<br />

values measured at 0.95 V to avoid diffusion-induced errors in<br />

kinetic current densities. The order of specific activity becomes<br />

apparent, with Pt/C being the least active, followed by Pt-NSTF<br />

and polycrystalline Pt. One can observe a significant increase in<br />

activity for PtNi Meso-TF, accompanied by a decrease in Tafel<br />

slope from ∼70 mV dec −1 for monometallic Pt to ∼40 mV dec −1 .<br />

This value is considerably lower than those commonly reported<br />

for Pt-based catalysts in the literature 19,47 , but it is in line with the<br />

value obtained on Pt3Ni(111)-skin 47 . The activity of PtNi Meso-TF<br />

exhibits an improvement factor of over 8 versus Pt-poly and<br />

Pt-NSTF. Furthermore, when compared with the state-of-the-art<br />

conventional Pt/C catalyst, the specific activity of the PtNi Meso-TF<br />

achieves an unprecedented 20-fold enhancement. Even though<br />

optimal film thickness, alloy composition and total Pt loading will<br />

be extensively studied in the future, the measured improvement<br />

expressed in A/mgPt corresponds to a mass activity that is already<br />

three times higher than the US Department of Energy technical<br />

target 12 . Together, the flat voltammetric curves, the trend in specific<br />

activity, the low Tafel slope and the structural characterizations<br />

strongly suggest that the annealed PtNi Meso-TF has a Pt-skin-type<br />

near-surface structure.<br />

To review the findings on thin-film-based mesostructured<br />

catalysts and to merge them into the same chart with nanoscale<br />

systems and bulk materials, we present in Fig. 5 the ORR activity<br />

map for different classes of Pt alloys, that is, from nanoparticles<br />

dispersed on high-surface-area carbon, to polycrystalline bulk<br />

materials and to single-crystalline alloys of Pt3Ni(hkl) surfaces 19 .<br />

This map shows a huge span in intrinsic specific activities among<br />

materials of the same bulk elemental composition that differ in<br />

form and surface structure. It also demonstrates the importance<br />

of controlling fundamental properties that determine catalytic<br />

performance; specifically that the ability to alter physical parameters<br />

such as particle size, near-surface composition profile, morphology<br />

and surface structure can lead to great improvements in functional<br />

properties of real catalysts. Notably, we have explored a number<br />

of NSTF catalysts with different compositions as summarized in<br />

Fig. 5; however, for the sake of brevity in this report we have<br />

presented only the results for the PtNi in detail. The activity values<br />

are given for Pt alloys with different transition metals associated<br />

with the atomic number (Z). The main features in Fig. 5 are<br />

designated activity regions for different classes of materials. Metallic<br />

nanoparticles of Pt and Pt alloys dispersed on a high-surface-area<br />

carbon support exhibit profoundly lower activities compared with<br />

their polycrystalline bulk counterparts. The assigned region that<br />

reflects the activity range of metallic nanoparticles is based on<br />

the literature data reported for Pt-alloys obtained by conventional<br />

impregnation methods. The next level in activity is reserved for<br />

extended bulk polycrystalline systems, where the specific activity<br />

of Pt3M-alloys can be improved by a factor of three versus<br />

Pt-poly. As mentioned above, the capability to control the surface<br />

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials 5<br />

© 2012 Macmillan Publishers Limited. All rights reserved.<br />

8<br />

6<br />

4<br />

2<br />

Improvement factor versus Pt-poly


ARTICLES<br />

Activity improvement factor versus Pt-poly<br />

NATURE MATERIALS DOI: 10.1038/NMAT3457<br />

Pt3Ni(111) 20.0 70<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

1.0<br />

Single-crystal alloys<br />

Polycrystalline alloys<br />

Pt 3 Ti<br />

Pt-skin<br />

Pt-skeleton<br />

NSTF<br />

Pt 3 V<br />

Metallic nanoparticles dispersed in carbon<br />

Pt 3 Fe<br />

Pt 3 Co<br />

NSTF<br />

Pt 3 Co<br />

Pt 3 CoNi<br />

NSTF<br />

Mesoscale<br />

ordering<br />

22 24 26 28<br />

Atomic number (z)<br />

PtNi<br />

Meso-TF<br />

Figure 5 | Activity map for the ORR obtained for different classes of Pt-based materials. Improvement factors are given on the basis of activities<br />

compared with the values for polycrystalline Pt and the state-of-the-art Pt/C catalyst established by RDE measurements in 0.1 M HClO4 at 0.95 V.<br />

structure leads to an extra boost in activity, and hence the highest<br />

ORR activity ever measured was obtained for the Pt3Ni(111)skin<br />

surface. On the basis of the values depicted in Fig. 5, the<br />

NSTF catalysts can successfully mimic the catalytic behaviour of<br />

polycrystalline bulk materials, while Pt-alloy mesostructured thin<br />

films exceed the range designated for polycrystalline systems. This<br />

is the first practical catalyst to approach the levels of activity<br />

previously reserved only for bulk single-crystalline surfaces, owing<br />

to the formation of a surface and near-surface structure similar<br />

to that of the ideal Pt3Ni(111)-skin. These bimetallic Meso-TF<br />

materials preserve sufficiently high specific surface area, which<br />

enables better utilization of precious metals. Moreover, Pt-based<br />

catalysts with mesoscale features also avoid the activity losses that<br />

are caused by the higher fraction of low-coordinated surface atoms<br />

that are present in nanoscale catalysts 14 . Consequently, thin-film<br />

electrocatalysts are hampered neither by the stability issues that<br />

accompany the use of high-surface-area carbon support, nor by<br />

the loss of active surface area due to particle agglomeration. The<br />

mesostructured thin films, therefore, unite the beneficial properties<br />

of both the nanoscale and the extended bulk systems, and lead<br />

to new design rules for producing highly active and durable<br />

electrocatalysts. These findings provide a proof of concept that the<br />

ability to tailor the composition, morphology and structure of the<br />

thin-film-based materials at the mesoscale allows the harvesting<br />

of maximal performance from the employed constituents. This is<br />

a seeding study on mesoscale ordering in electrocatalysts where<br />

further advancement towards an increased presence of (111) facets<br />

PtNi<br />

NSTF<br />

Pt 3 Ni<br />

Pt 3 Ni<br />

NSTF<br />

30<br />

Pt/C<br />

and optimization of the thin-film thickness may lead to additional<br />

improvements in specific and mass activities.<br />

We report on a new class of mesostructured catalysts based<br />

on thin films with an adjustable composition profile and surface<br />

morphology. These materials are in the form of metallic thin films<br />

with properties that have been tailored to improve the activity<br />

for the ORR. The obtained ORR activity is the highest ever<br />

measured on non-bulk catalysts owing to the beneficial near-surface<br />

compositional profile and its highly crystalline surface morphology.<br />

The exceptional properties of this Meso-TF are comparable to<br />

extended single-crystalline surfaces and improvement factors in<br />

kinetic activity of 8 versus polycrystalline Pt and 20 versus Pt/C<br />

are observed. The substantial advances in catalytic performance<br />

are obtained through structural mesoscale ordering of the thin<br />

film induced by thermal annealing in a reductive atmosphere. The<br />

approach as developed can be applied to generate a wide range<br />

of (electro)catalysts with tailored structure/composition, ultralow<br />

precious metal content and superior functional properties such as<br />

activity and durability.<br />

Methods<br />

Thin-film deposition. Thin metal films were deposited by planar magnetron<br />

sputter deposition on the ultrahigh-vacuum-cleaned surface of a mirror-polished<br />

glassy carbon substrate of 6 mm in diameter (base vacuum 1 × 10 −10 torr). The<br />

deposition rate was set to 0.3 Å s −1 by a quartz-crystal microbalance and an<br />

exposure of 7 s was calibrated for the nominal thickness of 2.2 ∼ 2.3 Å for<br />

a monolayer of Pt. The film thickness was derived from the exposure time of<br />

computer-controlled shutters during deposition. The thickness of all thin films in<br />

6 NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials<br />

© 2012 Macmillan Publishers Limited. All rights reserved.<br />

Pt<br />

78<br />

20<br />

15<br />

10<br />

5<br />

Activity improvement factor versus Pt/C


NATURE MATERIALS DOI: 10.1038/NMAT3457 ARTICLES<br />

this study was 20 nm. In the case of NSTF catalysts, consecutive layers of platinum<br />

and the metal of choice were deposited onto the NSTF layer of oriented organic<br />

pigment (perylene red) whiskers also by planar magnetron sputter deposition<br />

in vacuum 48 . The deposition process covered each of the perylene red whiskers<br />

with a thin metallic film. Both the monometallic Pt and the Pt-alloy catalyst were<br />

obtained by this method. The Meso-TF were obtained by thermal annealing of<br />

NSTF at 400 ◦ C in a hydrogen-rich atmosphere. The temperature was increased in<br />

increments of 20 ◦ C per 5 min and the whole process lasted 2 h.<br />

Electrochemical measurements. An Autolab PGSTAT 30 with FI20, ECD,<br />

ADC and SCAN GEN modules was used for the electrochemical measurements.<br />

Perchloric acid diluted with MilliQ water to 0.1 M was the electrolyte in all cases.<br />

The gases used were research grade (5N5+) argon and oxygen. In all experiments,<br />

a silver–silver chloride was the reference electrode. However, all potentials<br />

referred to in this paper are converted to the pH-independent reversible hydrogen<br />

electrode scale. We repeated all experiments 8 times to confirm reproducibility,<br />

and to improve the accuracy in the determination of kinetic activities. Kinetic<br />

current densities were obtained from the measured ORR polarization curves<br />

in accordance with the Koutecky–Levich equation: I −1<br />

ORR = I−1<br />

kinetic + I−1<br />

diffusion . The<br />

ECSA of the nanocatalysts was determined by integrating both the Hupd part of<br />

the CV profile, and the polarization curve obtained by oxidation of a monolayer<br />

of adsorbed carbon monoxide to avoid underestimation of the surface area due<br />

to altered hydrogen adsorption properties. All catalysts were deposited on a<br />

RDE made of glassy carbon and the loading of the nanoscale thin-film catalysts<br />

was adjusted to be 60 µgPt cm −2<br />

disc, whereas the loading for Pt/C obtained from<br />

TKK was 12 µgPt cm −2<br />

disc. Kinetic current densities as reported are normalized<br />

by ECSA in all cases.<br />

Microscopy. A Hitachi H-9500 environmental transmission electron microscope<br />

operated at 300 kV was used to perform the microstructural characterization and<br />

in situ heating TEM study. Powder samples were attached to the heating zone of a<br />

Hitachi gas-in<strong>je</strong>ction-heating holder. Images of nanoparticles were first recorded<br />

at room temperature, followed by heating of the specimen inside the microscope<br />

chamber with a vacuum level of about 10 −4 Pa. A CCD (charged-coupled device)<br />

camera was used to monitor the microstructural evolution and record images<br />

and videos. Each heating temperature was held for at least 10 min for detailed<br />

structural characterization, including morphology and atomic structure. A Hitachi<br />

SU70 high-resolution field-emission SEM was used for routine nanoparticle<br />

sample inspection. For the detailed surface morphology study at the nanometre<br />

scale, a Hitachi S-5500 ultrahigh-resolution cold field-emission SEM delivered<br />

a much higher resolution power (0.4 nm secondary electron image resolution<br />

at 30 kV) than normal SEM because of the specially designed ob<strong>je</strong>ctive lens.<br />

On both SU70 and S-5500, secondary electron images were taken at 15 kV or<br />

30 kV to reveal the surface morphology of both the as-deposited, as well as the<br />

annealed nanoparticles.<br />

Received 8 February 2012; accepted 11 September 2012;<br />

published online 11 November 2012<br />

References<br />

1. Borup, R. et al. Scientific aspects of polymer electrolyte fuel cell durability and<br />

degradation. Chem. Rev. 107, 3904–3951 (2007).<br />

2. Wagner, F. T., Lakshmanan, B. & Mathias, M. F. Electrochemistry and the<br />

future of the automobile. J. Phys. Chem. Lett. 1, 2204–2219 (2010).<br />

3. Adzic, R. R. et al. Platinum monolayer fuel cell electrocatalysts. Top. Catal. 46,<br />

249–262 (2007).<br />

4. Bruce, P. G., Freunberger, S. A., Hardwick, L. J. & Tarascon, J. M. Li-O2 and<br />

Li-S batteries with high energy storage. Nature Mater. 11, 19–29 (2012).<br />

5. Gasteiger, H. A., Kocha, S. S., Sompalli, B. & Wagner, F. T. Activity benchmarks<br />

and requirements for Pt, Pt-alloy, and non-Pt oxygen reduction catalysts for<br />

PEMFCs. Appl. Catal. B-Environ. 56, 9–35 (2005).<br />

6. Arico, A. S., Bruce, P., Scrosati, B., Tarascon, J. M. & Van Schalkwijk, W.<br />

Nanostructured materials for advanced energy conversion and storage devices.<br />

Nature Mater. 4, 366–377 (2005).<br />

7. Strasser, P. et al. Lattice-strain control of the activity in dealloyed core–shell<br />

fuel cell catalysts. Nature Chem. 2, 454–460 (2010).<br />

8. Greeley, J. et al. Alloys of platinum and early transition metals as oxygen<br />

reduction electrocatalysts. Nature Chem. 1, 552–556 (2009).<br />

9. Zhang, J., Sasaki, K., Sutter, E. & Adzic, R. R. Stabilization of platinum<br />

oxygen-reduction electrocatalysts using gold clusters. Science 315,<br />

220–222 (2007).<br />

10. Nilekar, A. et al. Bimetallic and ternary alloys for improved oxygen reduction<br />

catalysis. Top. Catal. 46, 276–284 (2007).<br />

11. Gorlin, Y. & Jaramillo, T. F. A bifunctional nonprecious metal catalyst<br />

for oxygen reduction and water oxidation. J. Am. Chem. Soc. 132,<br />

13612–13614 (2010).<br />

12. Debe, M. K. Electrocatalyst approaches and challenges for automotive fuel cells.<br />

Nature 486, 43–51 (2012).<br />

13. Snyder, J., Fujita, T., Chen, M. W. & Erlebacher, J. Oxygen reduction in<br />

nanoporous metal-ionic liquid composite electrocatalysts. Nature Mater. 9,<br />

904–907 (2010).<br />

14. Wang, C. et al. Design and synthesis of bimetallic electrocatalyst with<br />

multilayered Pt-skin surfaces. J. Am. Chem. Soc. 133, 14396–14403 (2011).<br />

15. Wang, C. et al. Multimetallic Au/FePt3 nanoparticles as highly durable<br />

electrocatalyst. Nano Lett. 11, 919–926 (2010).<br />

16. Ferreira, P. J. et al. Instability of Pt/C electrocatalysts in proton exchange<br />

membrane fuel cells—A mechanistic investigation. J. Electrochem. Soc. 152,<br />

A2256–A2271 (2005).<br />

17. Stamenkovic, V. R. et al. Improved oxygen reduction activity on Pt3Ni(111)<br />

via increased surface site availability. Science 315, 493–497 (2007).<br />

18. Stamenkovic, V. R. et al. Trends in electrocatalysis on extended and nanoscale<br />

Pt-bimetallic alloy surfaces. Nature Mater. 6, 241–247 (2007).<br />

19. Stamenkovic, V., Schmidt, T. J., Ross, P. N. & Markovic, N. M. Surface<br />

composition effects in electrocatalysis: Kinetics of oxygen reduction<br />

on well-defined Pt3Ni and Pt3Co alloy surfaces. J. Phys. Chem. B 106,<br />

11970–11979 (2002).<br />

20. Stamenkovic, V. R., Mun, B. S., Mayrhofer, K. J. J., Ross, P. N. & Markovic,<br />

N. M. Effect of surface composition on electronic structure, stability, and<br />

electrocatalytic properties of Pt-transition metal alloys: Pt-skin versus<br />

Pt-skeleton surfaces. J. Am. Chem. Soc. 128, 8813–8819 (2006).<br />

21. Koh, S. & Strasser, P. Electrocatalysis on bimetallic surfaces: Modifying<br />

catalytic reactivity for oxygen reduction by voltammetric surface dealloying.<br />

J. Am. Chem. Soc. 129, 12624–12625 (2007).<br />

22. Zhang, J. L., Vukmirovic, M. B., Xu, Y., Mavrikakis, M. & Adzic, R. R.<br />

Controlling the catalytic activity of platinum-monolayer electrocatalysts<br />

for oxygen reduction with different substrates. Angew. Chem. Int. Ed. 44,<br />

2132–2135 (2005).<br />

23. Muker<strong>je</strong>e, S. & Srinivasan, S. Enhanced electrocatalysis of oxygen reduction on<br />

platinum alloys in proton-exchange membrane fuel-cells. J. Electroanal. Chem.<br />

357, 201–224 (1993).<br />

24. Toda, T., Igarashi, H., Uchida, H. & Watanabe, M. Enhancement of the<br />

electroreduction of oxygen on Pt alloys with Fe, Ni, and Co. J. Electrochem. Soc.<br />

146, 3750–3756 (1999).<br />

25. Mavrikakis, M. Computational methods: A search engine for catalysts.<br />

Nature Mater. 5, 847–848 (2006).<br />

26. Mani, P., Srivastava, R. & Strasser, P. Dealloyed Pt–Cu core–shell nanoparticle<br />

electrocatalysts for use in PEM fuel cell cathodes. J. Phys. Chem. C 112,<br />

2770–2778 (2008).<br />

27. Wang, C. et al. Monodisperse Pt3Co nanoparticles as a catalyst for the<br />

oxygen reduction reaction: Size-dependent activity. J. Phys. Chem. C 113,<br />

19365–19368 (2009).<br />

28. Wang, C. et al. Monodisperse Pt3Co nanoparticles as electrocatalyst: The<br />

effects of particle size and pretreatment on electrocatalytic reduction of oxygen.<br />

Phys. Chem. Chem. Phys. 12, 6933–6939 (2010).<br />

29. Van der Vliet, D. et al. Platinum-alloy nanostructured thin film catalysts for<br />

the oxygen reduction reaction. Electrochim. Acta 56, 8695–8699 (2011).<br />

30. Zeis, R., Mathur, A., Fritz, G., Lee, J. & Erlebacher, J. Platinum-plated<br />

nanoporous gold: An efficient, low Pt loading electrocatalyst for PEM fuel cells.<br />

J. Power Sources 165, 65–72 (2007).<br />

31. Habas, S. E., Lee, H., Radmilovic, V., Somorjai, G. A. & Yang, P. Shaping<br />

binary metal nanocrystals through epitaxial seeded growth. Nature Mater. 6,<br />

692–697 (2007).<br />

32. Tao, F. et al. Reaction-driven restructuring of Rh–Pd and Pt–Pd core–shell<br />

nanoparticles. Science 322, 932–934 (2008).<br />

33. Inaba, M. et al. Controlled growth and shape formation of platinum<br />

nanoparticles and their electrochemical properties. Electrochim. Acta 52,<br />

1632–1638 (2006).<br />

34. Paulus, U. A. et al. Oxygen reduction on carbon-supported Pt–Ni and Pt–Co<br />

alloy catalysts. J. Phys. Chem. B 106, 4181–4191 (2002).<br />

35. Antonietti, M. & Ozin, G. A. Promises and problems of mesoscale materials<br />

chemistry or why meso? Chem. Eur. J. 10, 28–41 (2004).<br />

36. Muker<strong>je</strong>e, S., Srinivasan, S., Soriaga, M. P. & Mcbreen, J. Role of structural<br />

and electronic-properties of Pt and Pt alloys on electrocatalysis of oxygen<br />

reduction—an in-situ XANES and EXAFS investigation. J. Electrochem. Soc.<br />

142, 1409–1422 (1995).<br />

37. Debe, M. K. & <strong>Dr</strong>ube, A. R. Structural characteristics of a uniquely<br />

nanostructured organic thin-film. J. Vacuum Sci. Technol. B 13,<br />

1236–1241 (1995).<br />

38. Debe, M. K. & Poirier, R. J. Postdeposition growth of a uniquely nanostructured<br />

organic film by vacuum annealing. J. Vacuum Sci. Tech. A 12, 2017–2022 (1994).<br />

39. Debe, M. K., Schmoeckel, A. K., Vernstrom, G. D. & Atanasoski, R. High<br />

voltage stability of nanostructured thin film catalysts for PEM fuel cells.<br />

J. Power Sources 161, 1002–1011 (2006).<br />

40. Debe, M. K. in Handbook of Fuel Cells – Fundamentals, Technology and<br />

Applications (eds Vielstich, W., Lamm, A. & Gasteiger, H. A.) Ch. 45 (Wiley,<br />

2003).<br />

41. Debe, M. K. et al. Extraordinary oxygen reduction activity of Pt3Ni7. J.<br />

Electrochem. Soc. 158, B910–B918 (2011).<br />

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials 7<br />

© 2012 Macmillan Publishers Limited. All rights reserved.


ARTICLES<br />

42. Somorjai, G. A. Surface science and catalysis. Science 227, 902–908 (1985).<br />

43. Tao, A. R., Habas, S. & Yang, P. D. Shape control of colloidal metal nanocrystals.<br />

Small 4, 310–325 (2008).<br />

44. Lai, S. C. S. & Koper, M. T. M. The influence of surface structure on selectivity<br />

in the ethanol electro-oxidation reaction on platinum. J. Phys. Chem. Lett. 1,<br />

1122–1125 (2010).<br />

45. Wadayama, T. et al. Oxygen reduction reaction activities of Ni/Pt(111) model<br />

catalysts fabricated by molecular beam epitaxy. Electrochem. Commun. 12,<br />

1112–1115 (2010).<br />

46. Gancs, L., Kobayashi, T., Debe, M. K., Atanasoski, R. & Wieckowski, A.<br />

Crystallographic characteristics of nanostructured thin-film fuel cell<br />

electrocatalysts: A HRTEM study. Chem. Mater. 20, 2444–2454 (2008).<br />

47. Subbaraman, R., Strmcnik, D., Paulikas, A. P., Stamenkovic, V. R. & Markovic,<br />

N. M. Oxygen reduction reaction at three-phase interfaces. Chem. Phys. Chem.<br />

11, 2825–2833 (2010).<br />

48. Debe, M. K. et al. Nanostructured thin film catalysts for PEM fuel cells by<br />

vacuum web coating. (Society of Vacuum Coaters—50th Annual Technical<br />

Conference Proceedings, 2007).<br />

Acknowledgements<br />

The research was conducted at Argonne National Laboratory, which is a US Department<br />

of Energy Office of Science Laboratory operated by UChicago Argonne, LLC under<br />

NATURE MATERIALS DOI: 10.1038/NMAT3457<br />

contract no. DE-AC02-06CH11357. The portion of work related to extended<br />

single-crystalline and thin-film surfaces was supported by the US Department of Energy,<br />

Office of Science, Office of Basic Energy Sciences, Materials Sciences and Engineering<br />

Division. The portion of work exploring practical thin-film-based electrocatalysts<br />

was supported by the Office of Energy Efficiency and Renewable Energy, Fuel Cell<br />

Technologies Program. The authors thank Hitachi High Technologies America for the<br />

access to high-resolution electron microscopy facilities and J. Pearson and A. P. Paulikas<br />

for supporting thin film deposition experiments. V.R.S. is grateful to S. D. Bader and<br />

G.W. Crabtree for productive discussions.<br />

Author contributions<br />

D.F.V., C.W. and V.R.S. designed the experiments. C.W., D.F.V., D.T., D.S., X.F.Z.,<br />

R.T.A., M.K.D. and V.R.S. carried out the experimental work. D.F.V., C.W., M.K.D.,<br />

N.M.M. and V.R.S. discussed the results and V.R.S. wrote the manuscript.<br />

Additional information<br />

Supplementary information is available in the online version of the paper. Reprints and<br />

permissions information is available online at www.nature.com/reprints. Correspondence<br />

and requests for materials should be addressed to V.R.S.<br />

Competing financial interests<br />

The authors declare no competing financial interests.<br />

8 NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials<br />

© 2012 Macmillan Publishers Limited. All rights reserved.


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

LETTERS<br />

PUBLISHED ONLINE: XX MONTH XXXX | DOI: 10.1038/NMAT2883<br />

Selective catalysts for the hydrogen oxidation and<br />

oxygen reduction reactions by patterning of<br />

platinum with calix[4]arene molecules<br />

Bostjan Genorio 1,2 , Dusan Strmcnik 3 , Ram Subbaraman 4 , Dusan Tripkovic 3 , Goran Karapetrov 3 ,<br />

Vojislav R. Stamenkovic 3 , Stane Pejovnik 1 and Nenad M. Marković 3 *<br />

The design of new catalysts for polymer electrolyte membrane<br />

fuel cells must be guided by two equally important fundamental<br />

principles: optimization of their catalytic behaviour as well as<br />

the long-term stability of the metal catalysts and supports<br />

in hostile electrochemical environments1,2 . The methods used<br />

to improve catalytic activity are diverse3–8 , ranging from the<br />

alloying3,4 and de-alloying5 of platinum to the synthesis of<br />

platinum core–shell catalysts6 . However, methods to improve<br />

the stability of the carbon supports and catalyst nanoparticles<br />

are limited9,10 , especially during shutdown (when hydrogen<br />

is purged from the anode by air) and startup (when air is<br />

purged from the anode by hydrogen) conditions when the<br />

cathode potential can be pushed up to 1.5 V (ref. 11). Under the<br />

latter conditions, stability of the cathode materials is strongly<br />

affected (carbon oxidation reaction) by the undesired oxygen<br />

reduction reaction (ORR) on the anode side. This emphasizes<br />

the importance of designing selective anode catalysts that<br />

can efficiently suppress the ORR while fully preserving the<br />

Pt-like activity for the hydrogen oxidation reaction. Here, we<br />

demonstrate that chemically modified platinum with a selfassembled<br />

monolayer of calix[4]arene molecules meets this<br />

challenging requirement.<br />

The problem of cathode degradation during the shutdown/startup<br />

conditions in polymer electrolyte membrane fuel cells<br />

(PEMFCs) was realized two decades ago; however, there are still<br />

few methods to improve the selectivity of anode catalysts. This is<br />

because the required ensemble of active sites to achieve the maximum<br />

reaction rate of the ORR and the hydrogen oxidation reaction<br />

(HOR) is extremely small12 , making the design of selective anode<br />

catalysts extremely challenging. To meet this challenge, significant<br />

improvements in the design of anode materials must be realized.<br />

We present a chemically modified electrode (CME) approach<br />

to the design and improvement of a selective anode catalyst<br />

that can overcome the shutdown and startup limitations that<br />

prevent a full implementation of PEMFC technology. The use<br />

of CME approaches to modify electrocatalytic properties has<br />

been studied previously13,14 . The application of enzyme-mediated<br />

catalysis, specifically using H2-selective hydrogenase enzymes to<br />

carry out selective HOR even under aerobic conditions, has been<br />

reported recently15,16 . So far, the CME approach with a selfassembled<br />

monolayer (SAM) of calix[4]arene or similar organic<br />

molecules has been used as a controllable junction between sizedependant<br />

selectivity for gas separation17,18 and for chloride anion<br />

transport across lipid bilayers 19 . Here, we use such an approach to 44<br />

demonstrate that it is indeed possible to make an ‘oxygen-tolerant’ 45<br />

selective catalyst for the HOR. Although we focus primarily on 46<br />

the selectivity of Pt(111)–calixad systems, we also demonstrate that 47<br />

calix[4]arene-modified Pt(100) and polycrystalline Pt electrodes are 48<br />

highly selective for the ORR and HOR. 49<br />

We start with the analysis of the scanning tunnelling microscopy 50<br />

(STM) images of bare Pt(111) and Pt(111) modified with a 51<br />

calix[4]arene adlayer. In agreement with ref. 19, we found 52<br />

that Pt(111) is not an ideal surface, showing the presence of 53<br />

small adislands of Pt atoms that are separated by well-resolved 54<br />

monoatomic terrace-edge-step sites running roughly parallel to 55<br />

the (111) substrate direction (Fig. 1a). The STM image for a 56<br />

surface modified by the highest coverage of calix[4]arene (Fig. 1b) 57<br />

is characterized by a close-packed, long-range ordered monolayer 58<br />

of parallel arrays of molecules that almost completely cover the 59<br />

(111) terrace sites. Figure 1c is a schematic representation of 60<br />

the Pt(111)/calix interface, in which the wide rim of the cone 61<br />

(representing the anchoring groups) may serve as an electrode 62<br />

surface protector, the narrow rim of the cone as a molecular sieve 63<br />

and the lateral surface of the ‘truncated cone’ as a ‘blocking wall’. 64<br />

The driving force for ordering such large molecules is presum- 65<br />

ably governed by a synergy between the strong chemical interaction 66<br />

of Pt with calix[4]arene molecules and the surface homogeneity of 67<br />

Pt(111) that results in collective interaction of adsorbed molecules 68<br />

on terrace sites. It should be recognized, however, that on the basis 69<br />

of STM data alone it was impossible to deduce either the number 70<br />

of calix[4]arene-free Pt atoms or the nature (coordination) of the 71<br />

remaining bare Pt atoms. Although the structural details of the 72<br />

STM images lie beyond the scope of the present discussion, it is 73<br />

reasonable to suggest that as a result of steric effects most of the 74<br />

step-edges observed in Fig. 1a are not decorated with these large 75<br />

molecules. In contrast, it appears that smaller Pt adislands formed 76<br />

on the terraces of annealed Pt(111) may be ‘buried’ under the 77<br />

large calix[4]arene molecules. The determination of the number of 78<br />

available Pt sites is, however, less challenging given that a reasonable 79<br />

assessment of bare Pt atoms on Pt(111)–calixad can be obtained by 80<br />

monitoring how the fractional coverages (�) of underpotentially 81<br />

deposited hydrogen (Hupd) and hydroxyl species (OHad) are affected 82<br />

by �calix, as we demonstrate further below. 83<br />

Not surprisingly, Fig. 2a shows that pseudo-capacitive features 84<br />

corresponding to Hupd formation, double-layer charging and OHad 85<br />

formation on Pt(111) are almost completely suppressed on the 86<br />

1 Faculty of Chemistry and Chemical Technology, University of Ljubljana, Aškerčeva cesta 5, 1000 Ljubljana, Slovenia, 2 National Institute of Chemistry,<br />

Hajdrihova 19, SI-1001 Ljubljana, Slovenia, 3 Materials Science Division, Argonne National Laboratory, Argonne, Illinois 60439, USA, 4 Nuclear Engineering<br />

Division, Argonne National Laboratory, Argonne, Illinois 60439, USA. *e-mail: markovic@anl.gov.<br />

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials 1


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

LETTERS<br />

a b<br />

c<br />

Lateral surface<br />

Wide rim<br />

40 nm<br />

Narrow rim<br />

20 nm<br />

Figure 1 | Imaging of Pt(111) and Pt(111)–calix electrodes by STM and<br />

corresponding model for adsorbed calix[4]arene molecules on Pt(111).<br />

a, 200×200 nm 2 image of an as-prepared Pt(111) showing large terraces,<br />

divided by mono-atomic steps and covered with a small number of Pt<br />

adislands (the average size of the clusters is 2 nm). b, 100×100 nm 2 image<br />

of the calix[4]arene adlayer (surface coverage of ∼0.98 ML) showing<br />

long-range order of the self-assembled molecules. c, Schematic<br />

representation of calix[4]arene molecules attached to the surface via -SH<br />

groups located on the molecule’s wide rim.<br />

surface covered by organic molecules, confirming a high coverage<br />

of calix[4]arene molecules. Figure 2b also reveals that a systematic<br />

increase in �calix leads to a marked decrease in availability of Pt sites,<br />

that is, for Hupd and OHad from 16% on the least covered surface to<br />

2% on a surface with the most closely packed calix[4]arene adlayer.<br />

When the free Pt sites on the Pt(111)–calix surfaces are transformed<br />

into density of active sites (N), then a noticeable effect of �calix<br />

on the availability of Pt becomes even more obvious, as evidenced<br />

in Table 1. This suggests that the ions/water from the supporting<br />

electrolyte are unable to penetrate the narrow rim of calix[4]arene<br />

molecules. On the basis of these observations, along with the<br />

structural insight derived from the STM images, we propose that<br />

the only sites available for adsorption of electrolyte components are<br />

the relatively small number of Pt unmodified step-edges and/or the<br />

small ensembles of terrace sites between the anchoring groups.<br />

Having illustrated the effect of �calix on the formation of Hupd<br />

and OHad adlayers, a key question arises concerning the extent<br />

to which the �Hupd/OHad versus E curves (Fig. 2b) can be used for<br />

a<br />

b<br />

Current density (µA cm ¬2 )<br />

Charge density (µC cm ¬2 )<br />

NATURE MATERIALS DOI: 10.1038/NMAT2883<br />

0<br />

100<br />

0.2<br />

E (V versus RHE)<br />

0.4 0.6 0.8<br />

50<br />

0<br />

¬50<br />

¬100<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

I II<br />

III<br />

Pt(111)<br />

Pt(111)-A<br />

Pt(111)-B<br />

Pt(111)-C<br />

Pt(111)-D<br />

¬5<br />

0 0.2 0.4 0.6 0.8<br />

E (V versus RHE)<br />

Figure 2 | Relationships between surface coverages by calix[4]arene<br />

molecules and Hupd/OHad on Pt(111) in 0.1 M HClO4. a, The effect of �calix<br />

(curve A with 84% coverage, curve B with 95% coverage, curve C with<br />

96% coverage and curve D with 98% coverage) on cyclic voltammetry of<br />

Pt(111) (black dashed line), including the Hupd in region I, double layer<br />

(region II) and OHad adsorption in region III. b, Corresponding charge<br />

density versus E curves for Hupd and OHad, which are assessed on the basis<br />

of the assumption of one Hupd per Pt on the Pt(111)–(1×1) surface and<br />

∼160 µC cm −2 corresponds to the Hupd charge on Pt(111) in potential<br />

range I (ref. 20). A grey potential region in this and all other figures<br />

represents the potential window of importance to the anode selectivity.<br />

Sweep rates 50 mV s −1 .<br />

analysing the polarization curves for the HOR and the ORR in 19<br />

Fig. 3. As in ref. 21, in this analysis we use the general rate expression 20<br />

in which activity (current i) of the ORR and the HOR is simply 21<br />

Table 1 | Calix[4]arene coverages and TOFs for the ORR and HOR for five samples as described in the Methods section.<br />

Sample Hupd charge<br />

(µC cm −2 )<br />

Calix<br />

coverage (%)*<br />

Available<br />

surface (%)*<br />

Number of<br />

available<br />

sites (N cm −2 ) †<br />

Min. TOF for<br />

ORR @0.8 V ‡<br />

Pt(111) 161 0 100 10 15 9 8<br />

Pt(111)-A 26 84 16 1.6×10 14 9 49<br />

Pt(111)-B 9 94 6 5.5×10 13 9 129<br />

Pt(111)-C 7 96 4 4.2×10 13 6 194<br />

Pt(111)-D 4 98 2 2.4×10 13 6 388<br />

*Calix coverages and available surface were calculated by comparing Hupd charge on the bare Pt(111) and modified Pt(111).<br />

† Number of available sites was calculated assuming one H per Pt adsorption and number of total sites on Pt(111) surface 1.5×10 15 .<br />

Min. TOF for<br />

HOR @0.1 V ‡<br />

‡ Values are calculated by taking the measured current density @0.8 and @0.1, using the equation TOF = iE1/nFN. As current is under diffusion control for the HOR and in some cases for the ORR, the<br />

values are presented as minimum TOFs for the reactions.<br />

2 NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

NATURE MATERIALS DOI: 10.1038/NMAT2883 LETTERS<br />

a<br />

b<br />

c<br />

Current density (mA cm ¬2 )<br />

Current (mA)<br />

Current (mA)<br />

5<br />

0<br />

¬5<br />

¬10<br />

¬15<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0<br />

¬0.5<br />

¬1.0<br />

¬1.5<br />

Ring<br />

Disc<br />

0.1 M HClO 4<br />

E (V versus RHE)<br />

0 0.2 0.4 0.6 0.8<br />

Pt (1111)<br />

Pt (1111)-A<br />

Pt (1111)-B<br />

Pt (1111)-C<br />

Pt (1111)-D<br />

¬2.0<br />

0 0.2 0.4 0.6 0.8 1.0<br />

E (V versus RHE)<br />

Figure 3 | Design of ‘O2-tolerant’ selective anode catalysts for the HOR<br />

by controlling �calix on Pt(111). a, Polarization curves for the HOR in 0.1 M<br />

HClO4 on Pt(111) are the same as for all Pt(111)–calix modified surfaces.<br />

The curves for Pt(111)-B (grey) and -C (blue) are the same but they are<br />

omitted for clarity. Small variations may be observed in the hydrogen<br />

evolution region. These arise from variability in the IR correction for the<br />

electrode. For electrodes with high �calix values, the variability in IR values<br />

for the HOR can alter the overall hydrogen evolution reaction currents<br />

because of the large values of these currents. b, Peroxide oxidation currents<br />

recorded at 1.1 V versus RHE on the ring electrode during the ORR on the<br />

Pt(111)–calix disc electrodes. c, Corresponding polarization curves for the<br />

ORR. The rotation rate for all measurements was 1,600 r.p.m.; a sweep rate<br />

of 50 mV s −1 was used in all experiments.<br />

dependent on the number of available Pt sites:<br />

iE1 = nFK1creac.(1−�ad) (1)<br />

where n is the number of electrons, K1 is a constant, creac. is the<br />

concentration of H2 or O2 in the solution and �ad = �Hupd +<br />

�OHad + �calix is the fraction of the surface masked by the siteblocking<br />

species, that is, reaction intermediates are adsorbed at low<br />

coverages. Closely following our analysis in ref. 21, equation (1) is<br />

developed on the basis of a simple assumption that the Pt–O2 and 8<br />

Pt–H2 energetics as well as their reaction intermediates on bare Pt 9<br />

atoms is not affected by the surrounding calix[4]arene molecules. 10<br />

We also assume that if ions/water from the supporting electrolytes 11<br />

are unable to penetrate through the narrow end (see Fig. 1c), then 12<br />

the same should be valid for H2 and O2; that is, the adsorption of 13<br />

H2 and O2 occurs on a small number of calix[4]arene-free Pt sites. 14<br />

To demonstrate that the number of active sites required for the 15<br />

maximum rates of the HOR and ORR is extremely low, the reactivity 16<br />

of Pt(111)–calix surface was converted into a turnover frequency, 17<br />

TOF=iE1/nFN , and is summarized in Table 1. We notice in passing 18<br />

that any rigorous kinetic analyses of the ORR and HOR lie beyond 19<br />

the scope of the present discussion, because this analysis is not 20<br />

necessary to demonstrate the main aspect of our Letter, which is the 21<br />

high selectivity of the Pt(111)–calix surfaces for HOR. 22<br />

We start first with the analysis of polarization curves for the HOR 23<br />

on Pt(111) and Pt(111) modified with calix[4]arene (Fig. 3a). As 24<br />

found previously22 , the HOR on Pt(111) is an extremely fast process 25<br />

that, below 0.1 V, is determined predominantly by the surface 26<br />

coverage of spectator species Hupd (ref. 23). Above this potential, the 27<br />

reaction rate is always under pure diffusion control. An important 28<br />

observation from Fig. 3 is that for various surface coverages of 29<br />

calix[4]arene molecules the HOR is essentially the same. This 30<br />

suggests that the required number of Pt sites for the maximum rates 31<br />

of the HOR is rather small, that is, calculated on the basis of 2% 32<br />

surface site availability, the minimum TOF for HOR on Pt sites 33<br />

could be as high as 388 molecules/(site*s). Notice that this result 34<br />

has fulfilled the first requirement for designing highly active anode 35<br />

catalysts under operating PEMFC conditions—a Pt-like activity for 36<br />

the HOR. The second requirement, to make an O2-tolerant selective 37<br />

catalyst for the anode, is much more challenging and, to the best of 38<br />

our knowledge, a method suitable for the design of such catalysts 39<br />

has not been developed thus far. 40<br />

The ORR is a more complex multi-electron reaction in which 41<br />

O2 is reduced either to water without peroxide formation (4e− 42<br />

reduction) and/or to water plus peroxide formation (a mixed 4e− + 43<br />

2e− reduction) and/or completely to peroxide via a 2e− reduction 44<br />

process24 . As in our previous experiments25 , to analyse possible 45<br />

reaction pathways (neglecting any rigorous kinetic analyses) of the 46<br />

ORR on Pt(111) and Pt(111)–calix surfaces we use the rotating ring- 47<br />

disc electrode (RRDE) method. This method provides information 48<br />

on both the total currents for the ORR on the disc electrode (Fig. 3c) 49<br />

as well as the concomitant production of peroxide on the ring 50<br />

electrode (Fig. 3b). For Pt(111), starting at ∼0.95 V and sweeping 51<br />

the Pt(111) disc potential in the negative direction to 0.45 V, the 52<br />

ring currents were essentially zero, implying that in this potential 53<br />

region the ORR proceeds entirely through the direct 4e− pathway. 54<br />

Figure 3b shows that the appearance of peroxide oxidation currents 55<br />

on the ring electrode begins at potentials below 0.45 V and the 56<br />

limiting current corresponding to an exactly two-electron reduction 57<br />

of O2 is reached at the negative potential limit. 58<br />

There are two general observations concerning the ORR on 59<br />

Pt(111)–calix systems. First, the disc currents show that the 60<br />

ORR is inhibited on the calix[4]arene-modified surfaces and 61<br />

the deactivation increases with an increase of �calix (Fig. 3c and 62<br />

Table 1). The inhibition of O2 adsorption is so strong that the 63<br />

theoretical diffusion-limited currents corresponding to the disc 64<br />

geometric area are never reached. On the basis of our previous 65<br />

discussion in ref. 26, we propose that the currents observed on 66<br />

Pt(111)–calixad surfaces are related to the ORR on uncovered or 67<br />

partially covered active Pt patches. Given that the relative number of 68<br />

these patches is too small to allow a full overlap between the adjacent 69<br />

diffusion zones (spherical diffusion regions), it is reasonable to 70<br />

suggest that the currents below 0.6 V are diffusion-limited currents; 71<br />

but for conditions of highly blocked surfaces. Second, peroxide 72<br />

formation is hardly observed in the potential region of significance 73<br />

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials 3


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

LETTERS<br />

a<br />

c<br />

e<br />

Current density (mA cm ¬2 )<br />

Current density (mA cm ¬2 )<br />

Current density (µA cm ¬2 )<br />

5<br />

0<br />

¬5<br />

¬10<br />

¬15<br />

0<br />

¬2<br />

¬4<br />

¬6<br />

60<br />

40<br />

20<br />

0<br />

¬20<br />

¬40<br />

0.1 M HClO 4<br />

0 0.2 0.4 0.6 0.8<br />

E (V versus RHE)<br />

NATURE MATERIALS DOI: 10.1038/NMAT2883<br />

Pt(100)<br />

0<br />

Pt(Poly)<br />

Pt(100)-D<br />

Pt(Poly)-D<br />

0 0.2 0.4 0.6 0.8 1.0<br />

0 0.2 0.4 0.6 0.8 1.0<br />

E (V versus RHE) E (V versus RHE)<br />

¬60<br />

0 0.2 0.4<br />

E (V versus RHE)<br />

0.6 0.8<br />

b<br />

d<br />

f<br />

Current density (mA cm ¬2 )<br />

Current density (µA cm ¬2 )<br />

Current density (mA cm ¬2 )<br />

5<br />

¬5<br />

¬10<br />

¬15<br />

0<br />

¬2<br />

¬4<br />

¬6<br />

100<br />

50<br />

0<br />

¬50<br />

0 0.2 0.4 0.6 0.8<br />

E (V versus RHE)<br />

¬100<br />

0 0.2 0.4 0.6 0.8 1.0<br />

E (V versus RHE)<br />

Figure 4 | High selectivity of the ORR and HOR is also observed on the calix[4]arene-covered Pt(100) and polycrystalline Pt electrodes, suggesting that<br />

the Pt–calix systems are of broad fundamental and technological importance. a,b, Polarization curves for the HOR on bare and covered Pt(100) (a) and<br />

Pt(Poly) (b). c,d ORR on bare and calix[4]arene-covered Pt(100) (c) and Pt(Poly) (d) electrodes. e,f, Cyclic voltammograms for these surfaces in 0.1 M<br />

HClO4. Slight kinetic inhibition in HOR between 0.0 and 0.15 V observed for Pt(100) and Pt(Poly) surfaces is most likely the result of the high coverage of<br />

surface species 26 (Hupd) as well as calix for these electrodes. Both Pt(100) and Pt(Poly) electrodes exhibit the same diffusion-limiting currents for HOR as<br />

the bare surfaces while showing significant deactivation for the oxygen reduction reaction at potentials >0.6 V.<br />

for startup/shutdown conditions (E > 0.6 V). This is an important<br />

result as H2O2 may affect degradation of the Nafion membrane.<br />

As revealed in Fig. 3b, below 0.6 V a monotonic increase in the<br />

peroxide production is observed on Pt(111)–calixad. Furthermore,<br />

on this surface the onset potential for peroxide formation is shifted<br />

∼300 mV positive relative to Pt(111), indicating a change in the<br />

reaction pathway on these two surfaces. In short, below 0.6 V the<br />

2e− reduction process is predominant on calix[4]arene-modified<br />

Pt(111) surfaces. This is consistent with our previous proposition<br />

that larger ensembles of Pt sites are required for efficient cleavage of<br />

the O–O bond than for the adsorption of O2 and the concomitant<br />

formation of H2O2 (ref. 27).<br />

Importantly, Fig. 3 shows that on the 98% covered electrode the<br />

ORR is completely inhibited between 0.6 and 0.85 V; however, on<br />

the same surface and within the same potential window the HOR is<br />

under pure diffusion control. This unique selectivity of such CMEs<br />

may be attributed, by inference, to very strong ensemble effects in 17<br />

which the critical number of bare Pt atoms required for adsorption 18<br />

of O2 (that is, the ORR) is much higher than that required for 19<br />

the adsorption of H2 molecules and a subsequent HOR. It is also 20<br />

important to point out that the established selectivity was possible 21<br />

only because the required number of active sites for maximal rates 22<br />

of the HOR is, in fact, extremely small and just 2% of the available 23<br />

active surface sites are sufficient to reach the diffusion limiting 24<br />

currents. Finally, it is important to emphasize that the observed 25<br />

O2 selectivity is not unique to the Pt(111)–calix system and, as 26<br />

summarized in Fig. 4 and detailed in the caption, an exceptional 27<br />

anode selectivity for the ORR and HOR is also observed on the 28<br />

Pt(100) and polycrystalline Pt electrodes. This suggests that Pt–calix 29<br />

systems are of broad fundamental and technological importance. 30<br />

The results presented here indicate some important new 31<br />

directions in the quest to design selective anode catalysts for the 32<br />

4 NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

15<br />

16<br />

17<br />

18<br />

19<br />

20<br />

21<br />

22<br />

23<br />

24<br />

25<br />

26<br />

27<br />

28<br />

29<br />

30<br />

31<br />

32<br />

33<br />

34<br />

35<br />

36<br />

37<br />

38<br />

39<br />

40<br />

41<br />

42<br />

43<br />

44<br />

45<br />

46<br />

47<br />

48<br />

49<br />

50<br />

51<br />

52<br />

53<br />

54<br />

55<br />

56<br />

57<br />

58<br />

59<br />

60<br />

61<br />

NATURE MATERIALS DOI: 10.1038/NMAT2883 LETTERS<br />

Figure 5 |<br />

HOR that could improve the stability of cathode catalysts in<br />

PEMFC. We have demonstrated that a chemically modified Pt<br />

electrode with a SAM of calix[4]arene molecules can selectively<br />

block the ORR, but in such a way that the HOR proceeds with<br />

Pt-like activity. The optimum selectivity has been achieved by fine<br />

tuning the surface coverage of calix[4]arene molecules, leading to<br />

the formation of a critical ensemble of O2-tolerant Pt sites that<br />

are very active for the adsorption of H2 and consequent H–H<br />

bond breaking. The CME approach we have outlined here is not<br />

restricted to the Pt–calix systems, and may have many applications<br />

in analytical, synthetic and materials chemistry as well as in chemical<br />

energy conversion, selective fuel production and energy storage.<br />

Methods<br />

Synthesis of calix[4]arene. Several functional groups are known to bind to<br />

metal surfaces28 . Thiols are the most widely used compounds in chemisorption<br />

studies because of their well-known chemistry and the favourable self-assembly<br />

characteristics. Our previous studies showed that quadropod thio derivates of<br />

calix[4]arene ensure a strong bond with Pt substrates, resulting in relatively<br />

stable SAMs (ref. 29). For the present study we used protected thio derivatives of<br />

calix[4]arene. Quadropod anchoring compound was synthesized using a three-step<br />

reaction from the corresponding calix[4]arene. The alkyl-protected calix[4]arene<br />

1 was brominated to yield the bromo derivate 2. Lithiation with t-butyllithium,<br />

followed by introduction of sulphur and protection of the thiol group in the form<br />

of thiolacetate, gave a reasonable yield of the final compound 3 (Fig. 5). For the<br />

detailed procedure and data on the compounds see ref. 29.<br />

Preparation of Pt(111), Pt(100) and Pt(Poly) and self-assembly procedure. Pt<br />

electrodes were prepared by inductive heating for 10 min at ≈1,100 K in an argon<br />

hydrogen flow (3% hydrogen). The annealed specimen was cooled slowly to room<br />

temperature in this flow stream and immediately covered by a droplet of water.<br />

The electrode was then immersed in a tetrahydrofuran solution of calix[4]arene<br />

for 24 h, allowing the formation of a calix[4]arene SAM. The concentrations of<br />

calix[4]arene in tetrahydrofuran were 90, 130, 250 and 400 µM to obtain samples<br />

A, B, C and D, respectively, each with different calix coverages for the Pt(111)<br />

electrode (see Table 1). For Pt(100) and Pt(Poly) electrodes, 600 µM solutions were<br />

used to demonstrate the validity of the CME approach even at very high coverages.<br />

Calix coverages were calculated by Hupd comparison between the clean Pt(111)<br />

surface and chemically modified surfaces.<br />

STM method. For the as-prepared surface, the STM images were acquired<br />

with a Digital Instruments Multi-Mode Dimension STM controlled by a<br />

Nanoscope III control station. During the measurement, the microscope with<br />

the sample was enclosed in a pressurized cylinder with a CO atmosphere. For<br />

further details, see ref. 12.<br />

For modified surfaces, STM measurements were carried out on a home-built<br />

low-temperature STM similar to ref. 30 equipped with an RHK SPM1000<br />

controller. The samples were prepared according to the method described above<br />

and transferred to a helium glove box. Any water drops remaining on the sample<br />

were removed by blowing the surface with helium gas. The sample was then<br />

mounted on the STM stage and the STM head was sealed and transferred to the<br />

cryostat. The STM was cooled down to 4.2 K and the surface was scanned at a<br />

bias voltage of 500 mV and tunnelling current of 20 pA. Measurements at 4.2 K<br />

provided minimum drift during scanning (less than a few ångströms per hour).<br />

A high tunnelling resistance was necessary to ensure that the tip did not touch<br />

calix[4]arene molecules on the surface.<br />

RRDE method electrolytes and electrochemical set-up. After extensive rinsing,<br />

the electrode was embedded into the RRDE and transferred into a standard<br />

three-compartment electrochemical cell containing 0.1 M HClO4 (Sigma-Aldrich).<br />

In each experiment, the electrode was immersed at 0.07 V in a solution saturated<br />

with Ar. After obtaining a stable cycle between 0.07 and 0.7 V the polarization<br />

curve for the ORR was recorded on the disc whereas the peroxide oxidation signal<br />

was measured at the ring, which was held at 1.1 V versus the reversible hydrogen<br />

electrode (RHE). Peroxide currents presented are already corrected for the<br />

collection efficiency of 0.24. Subsequently, oxygen was purged from the solution,<br />

1 2 3<br />

replaced with hydrogen and HOR polarization curves were measured. Finally, the 62<br />

voltammetric response was again recorded in an argon-purged solution to confirm 63<br />

that the calix coverage had not changed significantly. 64<br />

All gases were 5N5 quality purchased from Airgas. The sweep rate for all 65<br />

measurements was 50 mV s −1 ; for the ORR measurements the electrode was rotated 66<br />

at 1,600 r.p.m. Electrode potentials are given versus the RHE. 67<br />

Received 25 June 2010; accepted 17 September 2010; 68<br />

published online XX Month XXXX 69<br />

References 70<br />

1. Vielstich, W., Lamm, A. & Gasteiger, H. (eds) Handbook of Fuel Cells 71<br />

(Fundamentals and Survey of Systems, Vol. 1, Wiley, 2003). 72<br />

2. Gasteiger, H. A. & Markovic, N. M. Just a dream—or future reality. Science 73<br />

324, 47–48 (2009). 74<br />

3. Stamenkovic, V. R. et al. Improved oxygen reduction activity on Pt3Ni(111) 75<br />

via increased surface site availability. Science 315, 493–497 (2007). 76<br />

4. Greeley, J. et al. Alloys of platinum and early transition metals as oxygen 77<br />

reduction electrocatalysts. Nature Chem. 1, 552–556 (2009). 78<br />

5. Srivastava, R., Mani, P., Hahn, N. & Strasser, P. Efficient oxygen reduction fuel 79<br />

cell electrocatalysis on voltammetrically dealloyed Pt–Cu–Co nanoparticles. 80<br />

Angew. Chem. Int. Ed. 46, 8988–8991 (2007). 81<br />

6. Nilekar, A. U. et al. Bimetallic and ternary alloys for improved oxygen reduction 82<br />

catalysis. Top. Catal. 46, 276–284 (2007). 83<br />

7. Debe, M. K., Schmoeckel, A. K., Vernstrom, G. D. & Atanasoski, R. High 84<br />

voltage stability of nanostructured thin film catalysts for PEM fuel cells. 85<br />

J. Power Sources 161, 1002–1011 (2006). 86<br />

8. Zhang, J., Sasaki, K., Sutter, E. & Adzic, R. R. Stabilization of platinum 87<br />

oxygen-reduction electrocatalysts using gold clusters. Science 315, 88<br />

220–222 (2007). 89<br />

9. Kinoshita, K. Carbon Electrochemical and Physicochemical Properties 90<br />

(JWS, 1998). 91<br />

10. Wang, J. & Swain, G. M. Fabrication and evaluation of platinum/diamond 92<br />

composite electrodes for electrocatalysis—Preliminary studies of the 93<br />

oxygen-reduction reaction. J. Electrochem. Soc. 150, E24–E32 (2003). 94<br />

11. Borup, R. et al. Scientific aspects of polymer electrolyte fuel cell durability and 95<br />

degradation. Chem. Rev. 107, 3904–3951 (2007). 96<br />

12. Strmcnik, D. S. et al. Unique activity of platinum adislands in the CO 97<br />

electrooxidation reaction. J. Am. Chem. Soc. 130, 15332–15339 (2008). 98<br />

13. Bard, A. J. & Stratmann, M. (eds) Modified Electrodes (Encyclopedia of 99<br />

Electrochemistry, Vol. 10, Wiley, 2007). 100<br />

14. Strmcnik, D. S. et al. Nature Chem.; Published online Aug 15 2010. 101<br />

15. Cracknell, J. A. et al. Enzymatic oxidation of H2 in atmospheric O2: 102<br />

The electrochemistry of energy generation from trace H2 by aerobic 103<br />

microorganisms. J. Am. Chem. Soc. 130, 424–425 (2008). 104<br />

16. Vincent, K. A., Cracknell, J. A., Parkin, A. & Armstrong, F. A. Hydrogen cycling 105<br />

by enzymes: Electrocatalysis and implications for future energy technology. 106<br />

Dalton Trans. 3397–3403 (2005). 107<br />

17. Markowitz, M. A., Janout, V., Castner, D. G. & Regen, S. L. Perforated 108<br />

monolayers—design and synthesis of porous and cohesive monolayers from 109<br />

mercurated calix[n]arenes. J. Am. Chem. Soc. 111, 8192–8200 (1989). 110<br />

18. Zhang, L., Hendel, R. A., Cozzi, P. G. & Regen, S. L. A single Langmuir–Blodgett 111<br />

monolayer for gas separations. J. Am. Chem. Soc. 121, 1621–1622 (1999). 112<br />

19. Sidorov, V., Kotch, F. W., Kuebler, J. L., Lam, Y-F. & Davis, J. T. Chloride 113<br />

transport across lipid bilayers and transmembrane potential induction by an 114<br />

oligophenoxyacetamide. J. Am. Chem. Soc. 125, 2840–2841 (2003). 115<br />

20. Clavilier, J. et al. Study of the charge displacement at constant potential during 116<br />

CO adsorption on Pt(110) and Pt(111) electrodes in contact with a perchloric 117<br />

acid solution. J. Electroanal. Chem. 330, 489–497 (1992). 118<br />

21. Strmcnik, D. et al. The role of non-covalent interactions in electrocatalytic 119<br />

fuel-cell reactions on platinum. Nature Chem. 1, 466–472 (2009). 120<br />

22. Marković, N. M., Grgur, B. N. & Ross, P. N. Temperature-dependent hydrogen 121<br />

electrochemistry on platinum low-index single-crystal surfaces in acid 122<br />

solutions. J. Phys. Chem. B 101, 5405–5413 (1997). 123<br />

23. Strmcnik, D. et al. Adsorption of hydrogen on Pt(1 1 1) and Pt(1 0 0) surfaces 124<br />

and its role in the HOR. Electrochem. Commun. 10, 1602–1605 (2008). 125<br />

24. Wroblowa, H. S., Pan, Y-C. & Razumney, G. Electroreduction of oxygen: A 126<br />

new mechanistic criterion. J. Electroanal. Chem. 69, 195–201 (1976). 127<br />

NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials 5


1<br />

2<br />

3<br />

4<br />

5<br />

6<br />

7<br />

8<br />

9<br />

10<br />

11<br />

12<br />

13<br />

14<br />

LETTERS<br />

25. Marković, N. M., Gasteiger, H. A., Grgur, B. N. & Ross, P. N. Oxygen<br />

reduction reaction on Pt(111): Effects of bromide. J. Electroanal. Chem. 467,<br />

157–163 (1999).<br />

26. Strmcnik, D. S. et al. Relationship between the surface coverage of spectator<br />

species and the rate of electrocatalytic reactions. J. Phys. Chem. C 111,<br />

18672–18678 (2007).<br />

27. Markovic, N. M., Gasteiger, H. A. & Ross, P. N. H2 and CO electrooxidation<br />

on well-characterized Pt, Ru, and Pt–Ru. 1. Rotating disk electrode<br />

studies of the pure gases including temperature effects. J. Phys. Chem. 99,<br />

8290–8301 (1995).<br />

28. Love, J. C. et al. Self-assembled monolayers of thiolates on metals as a form of<br />

nanotechnology. Chem. Rev. 105, 1103–1169 (2005).<br />

29. Genorio, B. et al. Synthesis and self-assembly of thio derivatives of calix[4]arene<br />

on noble metal surfaces. Langmuir 24, 11523–11532 (2008).<br />

NATURE MATERIALS DOI: 10.1038/NMAT2883<br />

30. Renner, Ch. et al. A versatile low-temperature scanning tunneling microscope. 15<br />

J. Vac. Sci. Technol. A 8, 330–332 (1990). 16<br />

Acknowledgements 17<br />

This work was supported by the Director, Office of Science, Office of Basic Energy 18<br />

Sciences, Division of Materials Sciences, US Department of Energy under Contract No. 19<br />

DE-AC03-76SF00098. B.G. acknowledges support from the Center of Excellence Low 20<br />

Carbon Technologies (CO NOT) and the Ministry of Higher Education, Science and 21<br />

Technology of Slovenia (ARRS-3311-04-831034). 22<br />

Additional information 23<br />

The authors declare no competing financial interests. Reprints and permissions 24<br />

information is available online at http://npg.nature.com/reprintsandpermissions. 25<br />

Correspondence and requests for materials should be addressed to N.M.M. 26<br />

6 NATURE MATERIALS | ADVANCE ONLINE PUBLICATION | www.nature.com/naturematerials


Page 1<br />

Query 1: Line no. 1<br />

‘novel’ changed to ‘new’ in the first sentence of<br />

the first paragraph. OK?<br />

Query 2: Line no. 1<br />

The title has been edited according to style. OK?<br />

Page 2<br />

Query 3: Line no. 7<br />

Please check ‘transformed into’ here. Can it be<br />

changed to, for example, ‘expressed as’?<br />

Query 4: Line no. 19<br />

Can the text in figure 2b’s caption be changed to<br />

‘which are assessed on the basis of the assumptions<br />

of one Hupd per Pt on the Pt(111)–(1×1) surface<br />

and that the Hupd charge on Pt(111) in potential<br />

range I is ∼160 µC cm −2 (ref. 20).?’<br />

Page 3<br />

Query 5: Line no. 1<br />

In figure 3’s caption should ‘IR’ be ‘iR (2<br />

instances), and can it be defined?<br />

Query 6: Line no. 1<br />

In figure 3’s caption, ‘HER’ written in full as<br />

‘hydrogen evolution reaction’. OK?<br />

Query 7: Line no. 33<br />

Can the text here be changed to ‘388 molecules<br />

per site per second’?<br />

Query 8: Line no. 41<br />

Can the sentence ‘The ORR is ... reduction<br />

process 24 .’ be changed to, for example, ‘The ORR is<br />

a more complex multi-electron reaction in which<br />

O2 is reduced to water and/or peroxide. The former<br />

process can occur directly (4e − reduction) and/or<br />

indirectly (via peroxide formation, which is a 2e −<br />

reduction process) 24 .’<br />

Page 4<br />

Query 9: Line no. 1<br />

Please check that the intended meaning of figure<br />

4’s caption has been retained after editing for style,<br />

and see the query below.<br />

Query 10: Line no. 1<br />

Can the text ‘for these surfaces in’ in figure 4e,f’s<br />

caption be changed to ‘for bare and covered Pt(100)<br />

(e) and Pt(Poly) (f)’?<br />

Query 11: Line no. 1<br />

‘Pt-Poly’ changed to ‘Pt(Poly)’ throughout.<br />

OK?<br />

Page 5<br />

Query 12: Line no. 1<br />

Please provide a short title, and a caption defining<br />

any acronyms not already defined in the main text,<br />

for figure 5.<br />

Query 13: Line no. 36<br />

‘H-UPD’ changed to ‘Hupd’ here, and in table 1.<br />

OK?<br />

Query 14: Line no. 60<br />

Text reworded to ‘at 1.1 V versus the reversible<br />

hydrogen electrode (RHE)’ here. OK?<br />

Query 15: Line no. 67<br />

It is now compulsory for authors to include<br />

an ‘Author contributions’ statement to specify<br />

the contributions of each co-author, such as<br />

experimental work, pro<strong>je</strong>ct planning, data analysis<br />

etc. This statement should be short, and refer to<br />

authors by their initials.<br />

Query 16: Line no. 101<br />

Please provide title and other details for ref. 14.<br />

Query 17: Line no. 107<br />

Please provide volume number for ref. 16.<br />

Page 6<br />

Query 18: Line no. 22<br />

Please cite the Supplementary Information<br />

somewhere relevant in the text.


ISSN 0036-0244, Russian Journal of Physical Chemistry A, 2011, Vol. 85, No. 13, pp. 2299–2304. © Pleiades Publishing, Ltd., 2011.<br />

Enhanced Activity in Ethanol Oxidation<br />

of Pt 3Sn Electrocatalysts Synthesized by Microwave Irradiation 1<br />

S. Stevanovi a , D. Tripkovi b , J. Roganc , D. Mini d , A. Gavrilovi e ,<br />

A. Tripkovi a , and V. M. Jovanovi a<br />

c � c � c � c �<br />

c �<br />

c �<br />

a ICTM Department of Electrochemistry, University of Belgrade, N<strong>je</strong>goševa 12, Belgrade, Serbia<br />

b Materials Science Division, Argonne National Laboratory, Argonne, IL 60439, USA<br />

c Faculty of Technology and Metallurgy, University of Belgrade, Karnegi<strong>je</strong>va 4, Belgrade, Serbia<br />

d Faculty of Physical Chemistry, University of Belgrade, Studentski trg 12, Belgrade, Serbia<br />

e CEST Centre of Electrochemical Surface Technology, Viktor–Kaplan–Strasse 2, A–2700 Wiener Neustadt, Austria<br />

e-mail: vlad@tmf.bg.ac.rs<br />

Received November 2, 2010<br />

Abstract—High surface area carbon supported Pt and Pt 3Sn catalysts were synthesized by microwave irradiation<br />

and investigated in the ethanol electro-oxidation reaction. The catalysts were obtained using a modified<br />

polyol method in an ethylene glycol solution and were characterized in terms of structure, morphology and<br />

composition by employing XRD, STM and EDX techniques. The diffraction peaks of Pt 3Sn/C catalyst in<br />

XRD patterns are shifted to lower 2θ values with respect to the corresponding peaks at Pt/C catalyst as a consequence<br />

of alloy formation between Pt and Sn. Particle size analysis from STM and XRD shows that Pt and<br />

Pt 3Sn clusters are of a small diameter (~2 nm) with a narrow size distribution. Pt 3Sn/C catalyst is highly<br />

active in ethanol oxidation with the onset potential shifted for ~150 mV to more negative values and with<br />

~2 times higher currents in comparison to Pt/C.<br />

Keywords: enhanced activity, Pt 3Sn electocatalyst, microwave irradiation, ethanol oxidation.<br />

DOI: 10.1134/S0036024411130309<br />

INTRODUCTION<br />

In various applications, fuel cells are widely recognized<br />

as very attractive devices to obtain electric<br />

energy directly from the combustion chemical product.<br />

Alcohols, mainly methanol, are widely proposed<br />

as possible fuels for mobile applications such as electric<br />

vehicles. Methanol is the simplest alcohol (only<br />

one carbon) and the electrocatalysis of its oxidation<br />

should be also simplest, but methanol is a toxic compound.<br />

However, ethanol offers an attractive alternative<br />

as a fuel in low temperature fuel cells because it<br />

can be produced in large quantities from agricultural<br />

products and it is the major renewable biofuel from the<br />

fermentation of biomass. Ethanol has also lower toxicity<br />

and higher mass energy density than methanol (6.1<br />

and 8.0 kWh/kg for methanol and ethanol, respectively).<br />

However, in the case of ethanol as a fuel, the<br />

problem is more complicated because ethanol contains<br />

two carbon atoms and for the complete oxidation<br />

of ethanol to CO 2 a good electrocatalyst must<br />

activate the C–C bond breaking (while avoiding the<br />

poisoning of the catalytic surface by CO species).<br />

The oxidation mechanism of ethanol in acid solution<br />

1 The article is published in the original.<br />

CHEMICAL KINETICS<br />

AND CATALYSIS<br />

2299<br />

may be summarized in the following schema of parallel<br />

reactions:<br />

CH3CH2OH [ CH3CH2OH] ad<br />

C1ad, C2ad CO2( total oxidation),<br />

(1)<br />

CH3CH2OH [ CH3CH2OH] ad<br />

(2)<br />

CH3CHO CH3COOH( partial oxidation).<br />

Carbon supported platinum is commonly used as<br />

anode catalyst in low temperature fuel cells. Pure Pt,<br />

however, is not the most efficient anodic catalyst for<br />

the direct ethanol fuel cell. On the other side platinum<br />

it self is known to be rapidly poisoned by strongly<br />

adsorbed species originating from the dissociative<br />

adsorption of ethanol [1]. Efforts to minimize the poisoning<br />

of Pt have been focused on the addition, of cocatalysts,<br />

such as Ru, Mo, Rh, Bi, Pd or Sn, to platinum<br />

to promote CO oxidation [1–4]. In this sense<br />

platinum is mostly alloyed with Sn to form binary catalysts.<br />

The activity of binary electrocatalyst was attributed<br />

to a bi-functional effect (Pt adsorbs alcohol and<br />

oxidizes H, while Sn oxides supply oxygen providing<br />

oxidation of the blocking intermediate CO) as well as<br />

the electronic interaction between Pt and alloyed metals<br />

[5, 6].


2300<br />

Electrocatalytic performance of the Pt-based catalysts<br />

is highly dependent not only on the nature of the<br />

metals added, but also on a variety of surface conditions<br />

of the synthesized materials (i.e. surface composition,<br />

morphology, impurity, contamination). Hence<br />

an ideal synthesis method of Pt-based electrocatalysts<br />

should be facile, cost effective, controllable and reproducible.<br />

It is also well known that the metal catalytic<br />

activity is strongly dependent on the particle shape,<br />

size and particle size distribution [7]. Conventional<br />

preparation techniques based on wet impregnation or<br />

chemical reduction of metal precursors does not provide<br />

satisfying control of particle size and shape [7].<br />

There has been continuing effort to develop an alternative<br />

synthesis method such as microwave irradiation<br />

[8, 9]. Microwave synthesis method has many advantages<br />

compared with conventional heating synthesis<br />

such as prompt start up and very short heating time<br />

thus enables homogeneous nucleation and shorter<br />

crystallization time leading to formation of small uniform<br />

metal particles.<br />

In this work, Pt and Pt 3Sn nanoparticles, synthesized<br />

by microwave assisted polyol method, were physically<br />

and electrochemically characterized for the ethanol<br />

electrooxidation reaction.<br />

EXPERIMENTAL<br />

Preparation of Pt/C and Pt 3Sn/C Electocatalysts<br />

In this procedure, in a 100 ml beaker mixture of<br />

0.5 ml of 0.05 M H 2PtCl 6 and 0.5 ml of 0.017 M SnCl 2<br />

solution was mixed with 25 ml of ethylene glycol under<br />

magnetic stirring. Then 0.8 M NaOH was added drop<br />

wise to adjust pH ~12. The beaker was placed in the<br />

center of an ordinary microwave oven and heated 60 s<br />

for the Pt and 90 s for Pt 3Sn catalyst at 700 W. After<br />

microwave heating, mixture was uniformly mixed with<br />

20 ml water suspension of 20 mg for Pt and 31.4 mg for<br />

Pt 3Sn of Vulcan XC-72 carbon and 150 ml 2 M H 2SO 4<br />

solution for 3 h with magnetic stirring. The resulting<br />

suspension was filtered and the residue was washed<br />

with high purity water. The solid product was dried at<br />

160°C for 3 h in N 2 atmosphere. The metal loading for<br />

both catalysts should be ~20 wt %. Thermogravimetric<br />

analysis (TGA) confirmed 20 wt % for Pt/C, while for<br />

Pt 3Sn/C a slightly lower portion of metal (~18 wt %)<br />

was found.<br />

Characterization of Pt/C and Pt 3Sn/C Electrocatalysts<br />

The thermogravimetric (TGA) and differential thermal<br />

(DTA) analyses were performed simultaneously<br />

(30–800°C range) on a SDT Q600 TGA/DSC<br />

instrument (TA Instruments). The heating rates were<br />

20 K min –1 and the sample mass was less than 10 mg.<br />

The furnace atmosphere consisted of air at a flow rate<br />

of 100 cm 3 min –1 .<br />

STEVANOVI C� et al.<br />

Unsupported Pt and Pt 3Sn nanoparticles were<br />

characterized by scanning tunnelling microscopy<br />

(STM). STM characterizations were performed using<br />

a NanoScope III A (Veeco, USA) microscope. The<br />

images were obtained in the height mode using a<br />

Pt–Ir tip (set-point current, i t, from 1 to 2 nA, bias<br />

voltage, V b = –300 mV). The mean particle size and<br />

distribution were acquired from a few randomly<br />

chosen areas in the STM images containing about<br />

100 particles.<br />

The X-ray diffraction (XRD) patterns of the powder<br />

catalysts were recorded with an Ital Structure<br />

APD2000 X–ray diffractometer in a Bragg–Brentano<br />

geometry using CuK α radiation (λ = 1.5418 Å) and<br />

step-scan mode (range: 2θ = 15°–85° step-time: 2.50 s,<br />

step-width: 0.02°). The program Powder Cell [10],<br />

was used for phase analysis and calculation of unit cell<br />

parameters.<br />

Microstructural examination was performed by<br />

scanning electron microscopy (SEM). An XL 30<br />

ESEM-FEG (environmental scanning microscope<br />

with field emission gun, manufactured by FEI, Netherlands)<br />

device equipped with an energy dispersive<br />

X-ray spectrometer from EDAX was used. The samples<br />

were inspected using 5, 10 and 20 kV acceleration<br />

voltages at magnifications of 20000× and 10000×<br />

respectively.<br />

Electrocatalytic activity of the catalysts was investigated<br />

by potentiodynamic and chronoamperometric<br />

tests using a PINE-RDE4 model potentiostat/galvanostat<br />

and a three-electrode compartment cell at<br />

room temperature. The working electrode was a thin<br />

layer of Nafion-impregnated Pt/C or Pt 3Sn/C catalysts<br />

applied on a glassy carbon disk electrode with the<br />

loading of 20 μg/cm 2 of the catalyst. The thin layer was<br />

obtained from a suspension of 2 mg of the respective<br />

catalyst in a mixture of 1 ml water and 50 μ1 of 5%<br />

aqueous Nafion solution, prepared in an ultrasonic<br />

bath, placed onto the substrate and dried at room temperature.<br />

Pt wire and a saturated calomel electrode<br />

(SCE) were used as the counter and reference electrode,<br />

respectively. The electrocatalytic activity of as<br />

prepared Pt/C and Pt 3Sn/C was studied in 0.1 M<br />

HClO 4 + 0.5 M C 2H 5OH solution. The electrolyte was<br />

prepared with high purity water and deaerated by N 2.<br />

Ethanol was added to the supporting electrolyte solution<br />

while holding the electrode potential at –0.2 V.<br />

The potential was then cycled up to 0.3 V i.e. the<br />

potential range of technical interest (E < 0.6 V (RHE))<br />

at sweep rate of 20 mV/s. Current–time transient<br />

curves were recorded after immersion of the freshly<br />

prepared electrode in the solution at –0.2 V for 2 s followed<br />

by stepping the potential to 0.2 V and holding<br />

the electrode at that potential for 30 min. A commercially<br />

available Pt/C catalyst from E-TEK with a nominal<br />

Pt loading of 20 wt % and particle size of ~2 nm was<br />

used as a benchmark.<br />

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A Vol. 85 No. 13 2011


RESULTS AND DISCUSSION<br />

Catalysts Characterizations<br />

The particle size and surface morphology of unsupported<br />

Pt and Pt 3Sn catalysts were characterized by<br />

STM. As observed from the top view of STM images<br />

(Fig. 1), both catalysts have rather uniform particles of<br />

a small diameter. Most of particles are in spherical<br />

shape. Cross section analysis (Fig. 1) confirmed particle<br />

size of


2302<br />

Intensity, a.u.<br />

Pt/C<br />

Pt 3Sn/C<br />

STEVANOVI C� et al.<br />

20 40 60 80<br />

2θ, deg<br />

Fig. 2. XRD patterns of Pt/C and Pt 3 Sn/C catalysts.<br />

j, mA/mg Pt 1<br />

120<br />

80<br />

40<br />

0<br />

−0.2 0 0.2 0.4<br />

E, V(vs SCE)<br />

Fig. 3. Potentiodynamic curves for the oxidation of 0.5 M<br />

C 2H 5OH at (1) Pt/C, (2) Pt 3Sn/C and (3) Pt/C–Tanaka<br />

in 0.1 M HClO 4, v = 20 mV/s.<br />

2<br />

3<br />

alyst (table). Because Sn has a bigger atomic radius<br />

than Pt (R Pt = 1.39 Å, R Sn = 1.61 Å) the addition of<br />

some amount of Sn to Pt could induce the extension of<br />

Pt unit cell parameter. Mean particle sizes were calculated<br />

by Scherrer formula [11] and for each catalyst are<br />

larger than that obtained by STM (table) possibly<br />

because unsupported catalysts were used for STM<br />

analysis. Still, the agreement can be described as very<br />

good because the values calculated by Scherrer formula<br />

account for a very probable lattice stress too.<br />

The small particle sizes and homogeneous size distributions<br />

of both catalysts with metal loading of<br />

20 wt % should be attributed to the advantages of<br />

microwave assisted modified polyol process in ethylene<br />

glycol solution. It is generally agreed that the size<br />

of metal nanoparticles is determined by the rate of<br />

reduction of the metal precursor. The dielectric constant<br />

(41.4 at 298 K) and the dielectric loss of ethylene<br />

glycol are high and hence rapid heating occurs easily<br />

under microwave irradiation [12]. In ethylene glycol<br />

mediated reactions (the “polyol” process), ethylene<br />

glycol also acts as a reducing agent to reduce the metal<br />

ions to metal powders [13]. The fast and uniform<br />

microwave heating accelerated the reduction of the<br />

metal ions and the formation of metallic nuclei, thus<br />

greatly facilitated small and uniform particle formation.<br />

Ethylene glycol has also big viscosity and it<br />

could prevent the agglomerations of the obtained<br />

nanoparticles.<br />

EDX analysis of Pt 3Sn/C catalyst revealed that the<br />

elemental composition agrees rather well with the<br />

nominal composition in the initial mixture (table).<br />

Electrochemical Performances<br />

The electrocatalytic activities of as prepared Pt/C,<br />

Pt/C-Tanaka and Pt 3Sn/C catalysts were studied in<br />

0.1 M HClO 4 + 0.5 M C 2H 5OH solution and positive<br />

scan of voltammetric curves are presented in Fig. 3.<br />

Pt 3Sn/C catalyst is highly active in ethanol oxidation<br />

with the onset potential at approximately –0.15 V<br />

(shifted for >0.1 V towards more negative potentials<br />

compared to Pt/C) and rapid kinetics, which is in<br />

accordance with the relevant results in the literature in<br />

the potential range of technical interest, i.e. at E < 0.6 V<br />

(RHE) [14]. Hydrogen adsorption/desorption peaks<br />

are clearly suppressed because the ethanol adsorption<br />

replaces the adsorbed hydrogen from the interface.<br />

The current densities in the whole potential region are<br />

two times higher for Pt 3Sn/C catalyst in comparison<br />

with Pt/C catalyst. Pt/C catalyst exhibits a lower<br />

activity than Pt 3Sn/C catalyst but higher than commercially<br />

Pt/C catalyst—Tanaka with the same metal<br />

loading probably due to its better dispersion and<br />

smaller particles as the results of microwave assisted<br />

synthesis.<br />

The stability of catalyst was studied in the chronoamperometric<br />

experiments and presented in Fig. 4.<br />

The higher initial current density at 0.2 V on Pt 3Sn/C<br />

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A Vol. 85 No. 13 2011


catalyst in comparison to Pt/C catalyst is in accordance<br />

with potentiodynamic measurments. The current<br />

decreases rapidly at Pt/C catalyst reaching study<br />

state values in a few minutes. On the other hand, the<br />

initial current decreases slightly at Pt 3Sn/C and stabilizes<br />

in the experimental period of time at the value<br />

which is about two times higher than for Pt/C catalyst.<br />

The Pt 3Sn/C catalyst is evidently less poisoned than<br />

Pt/C catalyst.<br />

Although Pt/C synthesized by microwave assisted<br />

polyol method is more active for ethanol oxidation<br />

reaction than commercial Pt/C-Tanaka catalyst, the<br />

activity of Pt 3Sn/C synthesized by polyol method is<br />

lower in comparison with activity of commercial<br />

Pt 3Sn/C–Tanaka catalyst [15]. However, in comparison<br />

with other Pt 3Sn/C catalysts prepared from colloid<br />

solutions but by usual heating, not microwave<br />

irradiation, our Pt 3Sn/C catalyst exhibits larger shift of<br />

onset potential to negative values [1, 6] as well as lower<br />

poisoning [16].<br />

The lower activity of our Pt 3Sn/C catalyst in comparison<br />

with commercial one according to Colmenares<br />

et al. [16] can be tentatively explained by<br />

lower alloying degree in polyol synthetized catalysts.<br />

Beneficial role of Pt–Sn alloy phase has been pointed<br />

out by other authors too [17, 18]. On the other hand,<br />

there are reports in the literature [1, 6] on remarkable<br />

promotion of ethanol oxidation by PtSn/C catalysts<br />

with main part of Sn in non-alloyed oxidized state<br />

what is probably the reason for better performance of<br />

our Pt 3Sn/C in comparison with similar other catalysts<br />

described in the literature.<br />

Thus, the presence of Sn in the catalyst can promote<br />

ethanol oxidation by an electronic effect in the<br />

Pt-based electrode material affecting on adsorption<br />

properties of the surface making this system less<br />

prompt to poisoning by organic species than pure Pt.<br />

Sn or its oxides can supply surface oxygen-containing<br />

species at lower potentials by activation of the interfacial<br />

water molecule necessary to complete the oxidation<br />

of adsorbed reaction intermediates leading to carbon<br />

dioxide, in the situation that the C–C bond was<br />

broken, or the formation of acetic acid [19, 20]. This<br />

oxidative removal of CO-like species strongly<br />

adsorbed on adjacent Pt active sites proceeds through<br />

so-called bi-functional mechanism.<br />

As displayed by the XRD results, the addition of Sn<br />

in the case of Pt 3Sn/C catalyst synthesized by microwave<br />

assisted polyol method induced slight extension<br />

of Pt-Pt distances resulting in rather low alloying<br />

degree. TG and EDX analysis revealed negligible loss<br />

of catalysts components what leads to reasonable<br />

assumption that significant quantity of non-alloyed Sn<br />

is present in the catalyst and on its surface. Thus, the<br />

increased activity of Pt 3Sn/C in comparison to Pt/C<br />

catalyst is most probably promoted by bi-functional<br />

mechanism although the electronic effect of alloyed<br />

Sn and extended Pt–Pt distance could play some role<br />

as well.<br />

ENHANCED ACTIVITY IN ETHANOL OXIDATION 2303<br />

j, mA/mgPt 16<br />

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A Vol. 85 No. 13 2011<br />

12<br />

8<br />

4<br />

0<br />

500<br />

1000<br />

1500<br />

Fig. 4. Chronoamperometric curves for the oxidation of<br />

0.5 M C 2 H 5 OH at 0.2 V on (1) Pt/C and (2) Pt 3 Sn/C catalysts<br />

in 0.1 M HClO 4 .<br />

2000<br />

t, s<br />

CONCLUSIONS<br />

A microwave assisted rapid heating polyol method<br />

was used to prepare carbon supported Pt and Pt3Sn nanoparticles with high electrocatalytic activities for<br />

the ethanol electrooxidation reaction. Structural<br />

characterization (XRD) and surface characterization<br />

(STM) of the catalysts revealed that by this method the<br />

catalysts with the small particles and with rather uniform<br />

size distribution were synthesized.<br />

Electrochemical measurements revealed the higher<br />

activity for the ethanol oxidation of the Pt3Sn/C. This<br />

catalyst has two times higher oxidation currents and<br />

significantly lower reaction onset potential than Pt/C<br />

catalyst. Chronoamperometric measurements revealed<br />

that Pt3Sn/C catalyst is notably less poisoned than<br />

Pt/C catalyst. Also, in comparison to other similar<br />

catalysts prepared by usual heating, not microwave<br />

irradiation, our catalyst exhibits larger shift of onset<br />

potential to negative values as well as lower poisoning.<br />

The increased activity of Pt3Sn/C in comparison to<br />

Pt/C catalyst is predominantly due to bi-functional<br />

mechanism although the electronic effect of alloyed<br />

Sn and extended Pt–Pt distances should also contribute<br />

the increase in activity.<br />

ACKNOWLEDGMENTS<br />

This work was financially supported by the Ministry<br />

of Science and Technological Development,<br />

Republic of Serbia, contract no. H–142056.<br />

REFERENCES<br />

1. C. Lamy, S. Rousseau, E. M. Belgsir, C. Coutanceau,<br />

and J. M. Leger, Electrochim. Acta 49, 3901 (2004).<br />

2. E. Antolini, J. Power Sources 170, 1 (2007).<br />

2<br />

1


2304<br />

3. E. Antolini, F. Colmati, and E. R. Gonzalez, J. Power<br />

Sources 193, 555 (2009).<br />

4. M. Li, A. Kowal, K. Sasaki, N. Marinkovic, D. Su,<br />

E. Korach, P. Liu, and R. R. Adzic, Electrochem. Acta<br />

55, 4331 (2010).<br />

5. W. J. Zhou, S. Q. Song, W. Z. Li, Z. H. Zhou, G. Q. Sun,<br />

Q. Xin, S. Douvartzides, and P. Tsiakaras, J. Power Sorces<br />

140, 50 (2005).<br />

6. M. Zhu, G. Sun, and Q. Xin, Electrochim. Acta 54,<br />

1511 (2009).<br />

7. I. S. Armadi, Z. L. Wang, T. C. Green, A. Henglein, and<br />

M. A. El-Sayed, Science 272, 1924 (1996).<br />

8. W. X. Chen, J. Y. Lee, and Z. L. Liu, Chem. Commun.,<br />

2588 (2002).<br />

9. Z. Liu, B. Guo, L. Hong, and T. H. Lim, Electrochem.<br />

Commun. 8, 83 (2006).<br />

10. W. Kraus and G. Nolze, Powder Cell for Windows, V. 2.4<br />

(Federal Inst. for Materials Research and Testing, Berlin,<br />

Germany, 2000).<br />

11. H. P. Klug and L. E. Alexander, X-Ray Diffraction Procedures,<br />

2nd ed. (Wiley, New York, 1974), p. 687.<br />

12. R. C. Weast, Handbook of Chemistry and Physics, 47th<br />

ed. (The Chemical Rubber, Cleveland, OH, 1966).<br />

STEVANOVI C� et al.<br />

13. F. Fivet, J. P. Lagier, B. Blin, B. Beaudoin, and<br />

M. Figlarz, Solid State Ionics 32–33, 198 (1989).<br />

14. Q. Wang, G. Q. Sun, L. H. Jiang, Q. Xin, S. G. Sun,<br />

Y. X. Jiang, S. P. Chen, Z. Jusys, and R. J. Behm, Phys.<br />

Chem. 9, 2686 (2007).<br />

15. A. V. Tripkovi c�, K. Dj. Popovi c�, J. D. Lovi c�,<br />

V. M. Jovanovi c�, S. I. Stevanovi c�, D. V. Tripkovi c�,<br />

A. Kowal, Electrochem. Commun. 11, 1030 (2009).<br />

16. L. Colmenares, H. Wang, Z. Jusys, L. Jiang, S. Yan,<br />

G. Q. Sun, and R. J. Behm, Electrochim. Acta 52, 221<br />

(2006).<br />

17. R. Alcala, J. W. Shabaker, G. W. Huber, M. A. Sanchez-<br />

Castillo, and J. A. Dumesic, J. Phys. Chem. B 109,<br />

2074 (2005).<br />

18. X. Tang, B. Zhang, Y. Li, Y. Xu, Q. Xin, and W. Shen,<br />

J. Mol. Catal. A: Chem., 235 (2005)<br />

19. S. Rousseau, C. Coutanceau, C. Lamy, and J. M. Leger,<br />

J. Power Sources 18, 158 (2006).<br />

20. F. C. Simoes, D. M. Dos Anjos, F. Vigier, J. M. Leger,<br />

F. Hahn, C. Coutanaceau, E. R. Gonyaley, G. Tremiliosi-Filho,<br />

A. R. De Andrade, P. Olivi, and K. B. Kokoh,<br />

J. Power Sources 11, 1567 (2007).<br />

RUSSIAN JOURNAL OF PHYSICAL CHEMISTRY A Vol. 85 No. 13 2011


Abstract #1510, 221st ECS Meeting, © 2012 The Electrochemical Society<br />

Electrocatalysts for the Oxygen Evolution Reaction<br />

N. Danilovic 1 , R. Subbaraman 2,3 , D. Strmcnik 2 , D.<br />

Tripkovic 2 , A.P. Paulikas 2 , K.-C. Chang 2 , V.R.<br />

Stamenkovic 2 , D. Myers 1 , N.M. Markovic 2<br />

1 Chemical Sciences and Engineering Division,<br />

2 Material Science Division<br />

3 Nuclear Engineering Division<br />

Argonne National Laboratory, Lemont, IL 60439<br />

Electrocatalysis of the oxygen evolution reaction (OER)<br />

is critical to the operation of electrolyzers. Currently,<br />

large-scale electrochemical production of hydrogen from<br />

water splitting is constrained by two limitations: 1) high<br />

overpotentials for the OER and 2) lack of stability of the<br />

electrode materials. While a large number of materials<br />

have been tested for the OER in alkaline and acid<br />

environments, the efforts were guided by a trial-and-error<br />

and/or a combinatorial approach with a resulting lack of<br />

studies focusing on systematic understanding of<br />

fundamental catalytic properties of the OER on wellcharacterized<br />

materials.<br />

In this presentation, using well-defined extended surfaces<br />

we demonstrate that, for proton exchange membrane<br />

(PEM) electrolyzers, activity and stability depend on the<br />

nature of the metal-oxide interface. Extended surfaces of<br />

Ir and Ru were used to investigate the OER on<br />

electrochemically or thermally formed oxides, while the<br />

traditional approach uses high surface area catalysts. By<br />

tuning specific electrochemical or thermal treatments on<br />

the Ir and Ru we selectively grow either amorphous<br />

electrochemical oxides or crystalline oxides with welldefined<br />

surfaces.<br />

Figure 1, illustrates the cyclic voltammograms (CVs) of<br />

polycrystalline Ir, IrO2 electrochemical oxide and IrO2<br />

thermal oxide in 0.1M HClO4.<br />

Figure 1. CVs of different Ir surfaces<br />

The electrochemical and thermal treatments have a drastic<br />

effect on the surface of Ir. Ir metal, prepared by radio<br />

frequency (RF) heat treatment in a reducing atmosphere,<br />

has a well-defined hydrogen underpotenial deposition<br />

(Hupd) and electrochemical double layer regions, with<br />

mixed oxide formation and OER region starting above 1.2<br />

V. Both thermally and electrochemically grown oxides<br />

block the Ir metal surface and prevent Hupd from<br />

forming. The effect of these different surface structures<br />

on the OER is shown in Figure 2 where the OER activity<br />

is shown.<br />

Figure 2. OER on different Ir surfaces<br />

A dramatic increase in activity is seen when the OER is<br />

performed on the electrochemically-grown oxide surface,<br />

while it is suppressed when the reaction is on a thermal<br />

oxide. The OER on the thin oxide on the metal surface,<br />

formed prior to the OER scan has an intermediate<br />

activity. This data suggests that the amorphous<br />

electrochemical oxide, allows promotes breakage of the<br />

O-H bond required for oxygen recombination and<br />

evolution.<br />

Similar data is observed for Ru and its electrochemically-<br />

and thermally-grown extended oxide surfaces.<br />

In conclusion, there is a strong structure-property<br />

relationship between the nature of the surface oxide and<br />

the rate oxygen evolution reaction.


Abstract #1532, 221st ECS Meeting, © 2012 The Electrochemical Society<br />

Fundamental investigations of precious metal<br />

stability in energy conversion systems<br />

D. Strmcnik, D. Tripkovic, R. Subbaraman, N. Danilovic, D. van der<br />

Vliet, A.P. Paulikas, V. R. Stamenkovic and N. M. Markovic<br />

Materials Science Division, Argonne National Laboratory, Argonne,<br />

Illinois 60439 USA<br />

Fundamental understanding of metal electrode-electrolyte<br />

interfaces has long been recognized as one of the<br />

cornerstones of successful commercialization of energy<br />

conversion systems such as fuel cells and electrolyzers.<br />

One of the major challenges faced by these systems today<br />

is the catalyst durability.<br />

State-of-the-art precious metal catalysts and their alloys<br />

are exposed to various harsh conditions under normal<br />

operating cycles, causing degradation of the catalyst and<br />

in their parent devices. In broad terms the degradation of<br />

the catalyst can be understood as any change in catalyst’s<br />

properties, including its size, shape, morphology, activity<br />

etc. While it is well known that precious metal catalysts<br />

degrade in the course of fuel cell/electrolyzer operation,<br />

to date, mostly due to the complexity of real catalyst<br />

systems, there is very little systematic research done into<br />

the fundamental aspects of catalyst degradation.<br />

In this lecture, a systematic scanning tunneling<br />

microscope (STM) and electrochemical study of the<br />

electrochemical interfaces of well-defined single crystal<br />

surfaces during conditions relevant for energy conversion<br />

systems is presented. The results give new insights into<br />

the degradation of electrocatalysts.<br />

Figure 1: An STM image of a) as prepared Pt(111)<br />

surface and b) Pt(111) surface cycled to 1.3 V.<br />

Acknowledgement.<br />

This work was supported by the contract (DE-AC02-<br />

06CH11357) between the University of Chicago and Argonne,<br />

LLC, and the US Department of Energy.


Abstract #1560, 221st ECS Meeting, © 2012 The Electrochemical Society<br />

Controlling Reactivity of Electrochemical Interfaces by Tuning Noncovalent<br />

Interactions<br />

D. Strmcnik, R. Subbaraman, D. Tripkovic, K. Chang, N. Danilovic, D. van<br />

der Vliet, P. Lopes, A. Paulikas, V. Stamenkovic and N. M. Markovic<br />

Argonne National Laboratory, Materials Science Division<br />

Argonne Il 60439<br />

In aqueous electrolytes, depending on the nature of the reacting species, the supporting<br />

electrolyte, and the electrode materials, two types of interactions have traditionally been considered: (i) direct –<br />

covalent - bond formation between adsorbates and electrodes, involving chemisorption, electron transfer, and<br />

release of the ion hydration shell; and (ii) relatively weak non-covalent forces that may affect the concentration<br />

of ions in the vicinity of the electrode but do not involve direct surface-adsorbate bonding. The range of<br />

physical phenomena associated with these two classes of bonds is unusually broad, and are of paramount<br />

importance to understand activity of classical two phase interfaces and Nafion-based three phase interfaces.<br />

In this lecture, we address the importance of both covalent and non-covalent interactions in<br />

controlling catalytic activity of the oxygen reduction reaction (ORR), hydrogen evolution reaction (HER)<br />

and oxidation of organic molecules in alkaline environments. Using surface x-ray scattering measurements<br />

we demonstrate that hydrated cations are located 3.4 Å away from the surface, suggesting that they are<br />

partially hydrated, though not adsorbed at the surface. The effect of these quasi-specifically adsorbed cations<br />

on electrochemical reactions ranges from highly inhibiting (ORR) to highly promoting (HER) to a modest<br />

effects during the oxidation of organic molecules. Although the field is still in its infancy, a great deal has<br />

already been learned and trends are beginning to emerge that give new insight into the relationship<br />

between the nature of bonding interactions and catalytic activity/stability of electrochemical interfaces.


Abstract #1570, 221st ECS Meeting, © 2012 The Electrochemical Society<br />

In-situ Infrared Spectroscopy at Solid-Liquid Interfaces as a<br />

Tool for Evaluation of Nanoscale Surface Morphology<br />

Dusan Tripkovic, Dusan Strmcnik, Dennis F. van der Vliet, Chao Wang, Nenad<br />

M. Markovic and Vojislav R. Stamenkovic<br />

Argonne National Lab, Argonne, IL, USA<br />

Single crustal surfaces of platinum have been characterized electrochemically,<br />

spectroelectrochemically and by scanning tunneling microscopy (STM) in order to reveal<br />

structure-function relationships between surface morphology and electrochemical activity for CO<br />

oxidation. Our findings strongly suggest that the surface morphology along with surface defects<br />

do have dominant role in determining intrinsic catalytic activity for CO electrooxidation. The<br />

results acquired on different single crystal surfaces in rapid-scan mode [1] suggest that the νCO - E<br />

curve, (where E is applied electrode potential), can be used as a spectral fingerprint for recognition<br />

of surface morphologies, including defects and low coordinated sites. FTIR combined with STM<br />

have brought out invaluable details about the true active sites and in combination with<br />

electrochemical methods enabled us to draw a true correlation between stability and morphology of the<br />

surfaces. Moreover, an interesting yet largely unexplored issue in characterizing the morphology of<br />

nanoparticles on the atomic scale is the lack of a surface specific method capable of detecting the<br />

presence of defects - low coordinated sites on nanoscale surfaces. The presence of such undercoordinated<br />

sites is challenging to detect in-situ and to monitor their transformation under reaction<br />

conditions. We demonstrated that a powerful method for characterizing nanoscale surfaces is<br />

utilization of in-situ FTIR in rapid-scan mode due to its sensitivity to the vibrational properties of<br />

adsorbed CO (νCO) which are dependent on surface morphology. Preliminary data, demonstrate that<br />

the adsorbate-induced changes provide a unique possibility for characterizing in-situ structural<br />

changes of nanoscale surfaces that are essential for understanding of structure-function relationships.<br />

1. Stamenkovic et al. Vibrational-properties of CO at the Pt(111) solution interface:the anomalous starktuning<br />

slope. J. Phys. Chem. B 2005, 109, 678-680.


Abstract #1650, Honolulu PRiME 2012, © 2012 The Electrochemical Society<br />

Advanced Electrocatalysts for PEM Fuel Cells<br />

C. Wang, D. van der Vliet, D. Tripkovic, D. Strmcnik, D.<br />

Li, N. M. Markovic and V. R. Stamenkovic<br />

Materials Science Division<br />

Argonne National Laboratory<br />

Argonne, IL 60439, USA<br />

The major barriers to commercial fuel cells are the cost,<br />

performance and durability of the current state-of-the-art<br />

Pt-based electrodes. Advancement in electrocatalysts for<br />

PEM fuel cells relies on the capability to alter their<br />

fundamental properties, employed materials and available<br />

technologies for rational synthesis. In this report we will<br />

summarize the most recent efforts from our group that<br />

have led to substantial improvements in durability and<br />

activity of electrocatalysts. Nanoscale materials in the<br />

form of multimetallic particles dispersed in high surface<br />

area carbon as well as nanostructured thin film catalysts<br />

will be presented. In all cases we relied on the knowledge<br />

obtained from well-defined systems in order to develop<br />

advanced catalysts with tailored properties. The reported<br />

synergy between well-defined systems and corresponding<br />

and nanoparticles emphasizes has been proven as an<br />

efficient approach in design and synthesis of advanced<br />

catalyst for PEM fuel cells.<br />

Such research effort involves the following<br />

steps: 1) characterization of well-defined solid materials by<br />

varying their surface structure, composition profile and<br />

electronic properties [1]; 2) atomic/molecular level<br />

characterization of electrochemical interfaces [2]; 3)<br />

theoretical modeling of structure/function relationship; 4)<br />

identification of the most active and stable sites at atomic<br />

scale during reaction conditions; 5) modification of<br />

electrified solid-liquid interfaces; 6) altering of the surface<br />

and subsurface composition by other constituents aimed to<br />

improving functionality [3]; 7) design and synthesis of<br />

nanomaterials with tailored structures.<br />

�<br />

[1] Stamenkovic et al. Science 315 (2007) 493.<br />

[2] Stamenkovic et al. Nature Materials 6 (2007) 241.<br />

[3] Genorio et al. Nature Materials 9 (2010) 998.<br />

[4] Wang et al. Nano Letters 11 (2011)919.


217th ECS Meeting, Abstract #1755, © The Electrochemical Society<br />

Design catalytic properties of electrochemical interfaces<br />

D. Strmcnik, K. Kodama, D. Tripkovic, D. van der Vliet, C. Wang,, V. Stamenkovic, and N. M.<br />

Marković<br />

Material Sciences Division Argonne National Laboratory, USA<br />

The successful deployment of advanced energy conversion and storage systems depends<br />

critically on our understanding of the fundamental bonding interactions at electrochemical<br />

interfaces []. In aqueous electrolytes, depending on the nature of the reacting species, the<br />

supporting electrolyte, and the electrode material, two types of interactions have traditionally<br />

been considered: (i) direct – covalent - bond formation and (ii) relatively weak non-covalent<br />

metal-ion forces that may affect the concentration of ions in the vicinity of the electrode but do<br />

not involve direct electrode - reactant/intermediate bonding. The range of physical phenomena<br />

associated with these two classes of bonds is unusually broad, ranging from charge transfer,<br />

release of the ion hydration shell, chemisorption, and catalysis.<br />

In this presentation, we address the importance of both covalent and non-covalent interactions<br />

in controlling catalytic activity of electrochemical interfaces. In order to give a brief<br />

overview of the field, this presentation will provide a carefully balanced selection of results<br />

for the oxygen reduction reaction, the hydrogen oxidation reaction and electrooxidation of<br />

small organic molecules first on metal single crystal surfaces and then on corresponding real<br />

nanoparticles. By focusing on the mechanism of action, we demonstrate that the ability to<br />

make a controlled arrangement of surface atoms presages a new era of advances in our<br />

knowledge of the electrochemical reactions.<br />

1


Abstract #1924, 219th ECS Meeting, © 2011 The Electrochemical Society<br />

Electrocatalysis on Well-Defined Solid-Liquid Interfaces<br />

Chao Wang, Dusan Strmcnik, Dusan Tripkovic, Dennis van der Vliet,<br />

Nenad M. Markovic and Vojislav R. Stamenkovic<br />

Argonne National Laboratory<br />

Materials Science Division<br />

Argonne, IL<br />

Novel nanomaterials with uniquely reactive surfaces can solve challenging problems in the areas<br />

as diverse as clean energy production, energy conversion, energy storage, corrosion,<br />

bioengineering, sensors, electronic devices etc. The ultimate goal in the heterogeneous/electrocatalysis<br />

would be to tune the electronic and structural properties of nanoparticles in order to<br />

achieve superior catalytic and stability enhancements [1].<br />

The material-by-design-approach, would be used to demonstrate how the knowledge obtained<br />

from the well-defined extended surfaces can be employed to create tailor-made nanocluster<br />

surfaces with advanced catalytic properties. The surfaces of polycrystalline bimetallic PtM alloys<br />

(M=Ni,Co,Fe,V,Ti,Re) as well as Pt3Ni(hkl) and Pt(hkl) single crystals were characterizaed in<br />

ultra-high vacuum chamber by AES, LEIS and UPS before transfer into electrochemical<br />

environment.<br />

Enhanced catalytic properties are induced by the second metal, and the mechanism of<br />

enhancement may occur through several effects: (1) Electronic effect, due to changes in the<br />

metallic d-band center position vs. Fermi level; and (2) Structural effect, which refers correlation<br />

between surface atom arrangements, and/or corrosion-induced dissolution – surface roughening.<br />

It has been found that each alloy, depending on prehistory, could form surfaces with two different<br />

compositions. The annealed alloys form the outermost Pt-skin surface layer, which consists only<br />

platinum atoms, while the sputtered surfaces have the bulk ratio of alloying components [2,3].<br />

Post-electrochemical UHV surface characterizations revealed that Pt-skin surfaces are stable after<br />

immersion to an electrolyte, while sputtered surfaces formed Pt-skeleton outermost layers as a<br />

result of dissolution of transition metal atoms. Modification in Pt electronic properties induced<br />

by alloying elements alters adsorption/catalytic properties of corresponding bimetallic alloy.<br />

The very same levels of catalytic enhancement have been established between extended surfaces<br />

and corresponding materials in the form of nanoparticles. In addition to the d-band center<br />

position, it has been found how catalytic activity could be affected by the arrangement of surface<br />

atoms, presence of defects and substrate. In order to engineer catalysts with the desirable<br />

physicochemical properties one should consider all relevant factors that could influence catalytic<br />

activity and stability of tailored surfaces.<br />

[1] V. Stamenkovic, B. Fowler, B.S. Mun, G. Wang, P.N. Ross, C.A. Lucas, N.M. Markovic, Science 315<br />

(2007) 493-497.<br />

[2] V. Stamenkovic, B.S. Mun, K.J.J. Mayrhofer, P.N. Ross, N.M. Markovic, J. Rossmeisl, J. Greeley, J.K.<br />

Nørskov, Angewandte Chemie International Edition 45 (2006) 2897.<br />

[3] V. Stamenkovic, B.S. Mun, M. Arenz, K.J.J. Mayrhofer, C. A. Lucas, G. Wang, P.N. Ross, N.M.<br />

Markovic, Nature Materials 6 (2007) 241.


Abstract #549, 221st ECS Meeting, © 2012 The Electrochemical Society<br />

Electrochemical Interfaces for Energy Conversion and Storage<br />

Ram Subbaraman, Dusan Tripkovic, Dusan Strmcnik, Gustav Wiberg, Jakub<br />

Jirkovsky, Chao Wang, Vojislav Stamenkovic and Nenad M. Markovic<br />

Argonne National Laboratory, Materials Science Division<br />

Argonne Il 60439<br />

Design and synthesis of energy efficient electrochemical interfaces (materials and double<br />

layer components) with tailored properties for accelerating and directing chemical transformations is one<br />

key to developing new alternative energy systems – fuel cells, electrolyzers and batteries. In the past,<br />

researchers working in the field of fuel cells (converting hydrogen and oxygen into water) and<br />

electrolyzers (splitting water back to H2 and O2) seldom focused on understanding the electrochemical<br />

compliments of these reactions in battery systems, e.g., the lithium-air system.<br />

In this lecture, we bridge the gap between the “water battery” (fuel cell ↔ electrolyzer) and<br />

the Li-air battery systems. We demonstrate that this would require fundamentally new knowledge in<br />

several critical areas. In water-based environments, we need to identify a common descriptor (the bond<br />

strength) that can be used to establish the guiding principles for designing reactivity of electrochemical<br />

interfaces. We demonstrate that the bonding energy of hydroxyl species to the varieties of wellcharacterized<br />

materials (OHad-M) spanning from metals, metal/metal-oxides and oxides can be used as a<br />

single descriptor to tune the activity of catalysts and double layer components. In turn, successful<br />

identification of such descriptor, together with associated electrocatalytic trends, provides the foundation<br />

for rational design of practical electrocatalysts. In organic-based environments, we need to establish the<br />

missing structure-function relationships for discharging processes in which chemical energy of Li and<br />

oxygen is transformed to LiO 2- /Li2O2/LiO2 and charging processes in which deposited<br />

intermediates/products are transformed back to Li ions and O2. We also need to understand at atomic<br />

and molecular levels a possible interaction of products/intermediates with organic solvents. We conclude<br />

that understanding the complexity (simplicity) of electrochemical interfaces would open new avenues for<br />

design and deployment of alternative energy systems.


Abstract #947, 220th ECS Meeting, © 2011 The Electrochemical Society<br />

Advanced Electrocatalysts: - From Extended to Nanoscale Surfaces<br />

Chao Wang, Dongguo Li, Dennis van der Vliet, Dusan Strmcnik, Dusan Tripkovic, Nenad M. Markovic,<br />

Vojislav R. Stamenkovic<br />

Materials Science Division, Argonne National Laboratory, Argonne, IL 60439, USA<br />

Highly efficient catalysts for energy conversion have become the primary target for the practical<br />

application of renewable energy technologies. The development of novel electrocatalysts relies<br />

on fundamental principles to improving high catalytic activity, durability, and achieving<br />

competitive cost effectiveness. Conventional platinum (Pt) catalysts that have been widely<br />

employed in fuel cells are not only expensive, but also have limited resources considering largescale<br />

applications. Moreover, Pt that is generally considered to be chemically inert becomes<br />

unstable when exposed to hostile electrochemical environments. All of that has led us to develop<br />

a systematic approach toward the rational design and synthesis of advanced Pt-bimetallic and -<br />

multimetallic electrocatalysts for fuel cells. Via synergistic studies on extended surfaces and<br />

high-surface-area catalysts, we managed to control and tune the critical parameters of the<br />

catalysts such as particle size, shape, composition and elemental distribution. These studies were<br />

supported also with theoretical simulation and they provided comprehensive understanding of the<br />

structure-function correlation (Fig. 1). As a result, we developed highly active and durable<br />

nanoscale materials, which can be employed in fuel cells.<br />

Figure 1. Schematic illustration of the stabilization mechanisms in highly durable Au/FePt3 catalysts for<br />

the oxygen reduction.


216th ECS Meeting, Abstract #868, © The Electrochemical Society<br />

Active sites for PEM fuel cell reactions<br />

Dusan Strmcnik 1 , Dusan Tripkovic 1 , Dennis van der Vliet 1 , Jeff Greely 2 ,<br />

Alexander Brownrigg 3 , Chris Lucas 3 , Goran Karapetrov 1 , Vojislav<br />

Stamenkovic 1 and Nenad M. Markovic 1<br />

1 Materials Science Division, 2 Center for Nanoscale Materials,<br />

Argonne National Laboratory, Argonne, Illinois 60439 USA<br />

3 Oliver Lodge Laboratory, Department of Physics, University of<br />

Liverpool, Liverpool, L69 7ZE, UK<br />

.<br />

The development of electrocatalytic materials of<br />

enhanced activity and efficiency through careful<br />

manipulation, at the atomic scale, of the catalyst surface<br />

structure, has long been a goal of electrochemists. To<br />

accomplish this ambitious ob<strong>je</strong>ctive, it would be<br />

necessary to obtain a thorough understanding of the<br />

relationship between the atomic-level surface structure<br />

and the catalytic properties.<br />

In the past, the most common approach for establishing<br />

structure function relationships was to correlate<br />

electrochemical behavior of different single crystal<br />

surfaces, assuming perfectly ordered surfaces.<br />

In our work, however, we focus on the sub-structures<br />

which exist on single crystal surfaces. Using sample<br />

preparation and characterization (STM, FT-IRAS and<br />

SXS) methods of metal single crystal surfaces, we were<br />

able to monitor and control adsorbate-/ potential-induced<br />

changes in the coordination of surface atoms. This, in<br />

turn, allowed us to find correlations between the nature as<br />

well as the number of active sites and the rate of PEM-FC<br />

reactions.<br />

Special focus was awarded to oxygen reduction reaction<br />

and CO oxidation reaction on Pt (100), Pt (111) and Pt<br />

(1099) surfaces to understand, how to combine different<br />

morphological features in one catalyst exhibiting high<br />

activities for both reactions.<br />

The knowledge about the active sites for a particular<br />

reaction is a prerequisite and the first step to<br />

understanding how to tune a catalyst performance.<br />

In our recent work, however, we have already made the<br />

second step, producing a high surface area Pt catalyst that<br />

shows unprecedented activities for CO oxidation as well<br />

as oxygen reduction reaction.<br />

Acknowledgement.<br />

This work was supported by the contract (DE-AC02-<br />

06CH11357) between the University of Chicago and Argonne,<br />

LLC, and the US Department of Energy. Use of the Center for<br />

Nanoscale Materials was supported by the U.S. Department of<br />

Energy, Office of Science, Office of Basic Energy Sciences,<br />

under contract No. DE-AC02-06CH11357. We acknowledge<br />

computer time at the Laboratory Computing Resource Center<br />

(LCRC) at Argonne National Laboratory, The National Energy<br />

Research Scientific Computing Center (NERSC), and the<br />

Molecular Science Computing Facility (MCSF) at Pacific<br />

Northwest National Laboratory.<br />

Wave number [cm -1 ]<br />

Current density [mA/cm 2 ]<br />

4<br />

3<br />

2<br />

1<br />

0<br />

2090<br />

2080<br />

2070<br />

2060<br />

2050<br />

EC<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

E [V vs. RHE]<br />

STM<br />

DFT<br />

FTIR<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

E [V vs. RHE]<br />

Figure 1. (top) Cyclic voltammograms (50 mv/s) for<br />

Pt(111) in 0.1 M HClO4 in the presence and absence of<br />

CO. (as prepared surface) (middle) STM image of the as<br />

prepared surface. Inset – DFT model of CO and OH<br />

coverage at the as prepared surface. (bottom) in-situ FT-<br />

IRAS experiment : dependence of νCO and ICO2 on<br />

potential.<br />

CO 2 band intensity [arb. u.]


N M. Markovic<br />

H. Gasteiger and N. M. Markovic


Fuel Cells and Electrocatalysis<br />

�More than just devise for energy conversion - the development of fuel<br />

cell catalysts in the last three decades has transformed electrocatalysis<br />

from art to science<br />

Cattech 4 (2000) 110<br />

H 2 /CO<br />

ORR/OER: a key for the<br />

development of alternative<br />

energies<br />

CO: a key “test molecule”<br />

HOR/HER : the mother of all<br />

electrochemical reactions


Surface Science Approach<br />

Single<br />

Crystals<br />

UHV<br />

Experiment<br />

MATERIALS-BY-DESIGN<br />

DESIGN<br />

Model<br />

Nano<br />

–– SYNTHESIZE<br />

CHARACTIZATION METHODS<br />

SXS<br />

FTIR<br />

STM/AFM RAMAN<br />

(111)<br />

METALS<br />

M -OXIDES<br />

OXIDES<br />

(100)<br />

–– CHARACTERIZE<br />

HRTEM<br />

ACTIVITY, SELECTIVITY AND STABILITY MAPPING<br />

CsOH<br />

LiOH<br />

Au<br />

Pt<br />

Ru<br />

Real<br />

Nano<br />

KOH<br />

Theory/Modeling<br />

ANIONS<br />

WATER<br />

CATIONS<br />

a)<br />

CHEMICAL<br />

Metal-Air Batteries<br />

SYNTHESIS METHODS<br />

Pt(111)<br />

0.85V<br />

�OH = 0.7<br />

Reactants: O 2 , CH 3 OH, H 2<br />

DOUBLE-LAYER-BY-DESIGN<br />

M + M +<br />

M +<br />

–– UNDERSTAND<br />

REAL SYSTEM<br />

Pt/C<br />

–– APPLY<br />

PHYSICAL<br />

Fuel Cell � Electrolyzer<br />

O 2<br />

H 2<br />

3


ORR- The reaction pathway<br />

O 2<br />

(O 2) ad<br />

2e-<br />

2H +<br />

4e-<br />

(H 2O 2) ad<br />

(H 2O 2) solution<br />

2e-<br />

2H +<br />

H 2O


Tailoring Electronic Properties<br />

dN(E)/dE (arb. units)<br />

dN(E)/dE (arb. units)<br />

Intensity [arb. units]<br />

Intensity [arb. units]<br />

Pt Pt 237 237<br />

Pt Pt 158 158<br />

Pt Pt 251 251<br />

AES (a) LEED<br />

200 200 400 400 600 600 800 800<br />

Auger Electron Energy [eV]<br />

LEIS<br />

Pt Pt 390 390<br />

Pt 3 Ni(hkl)<br />

Ni Ni 716 716Ni<br />

Ni 783 783<br />

Ni Ni 848 848<br />

Pt 3Ni(hkl)<br />

Pt 0.71<br />

Surface composition = 100% Pt<br />

0.2 0.4 0.6 0.8<br />

E/E0 UPS<br />

Pt3Ni (111)<br />

(100)<br />

(110)<br />

-8 -6 -4 -2 0<br />

Binding Energy [eV]<br />

Science, 315(2007)493<br />

(b)<br />

(c)<br />

AES 3 keV<br />

[01-1]<br />

[001]<br />

[011]<br />

[1-10]<br />

(d)<br />

p(1x1)<br />

(e) Pt 3 Ni(100)<br />

(f)<br />

Pt 3 Ni(111)<br />

c(5x1)<br />

Pt 3 Ni(110)<br />

(110)-(1x1)<br />

i [�A/cm 2 ]<br />

i [�A/cm 2 ]<br />

i [�A/cm 2 ]<br />

20<br />

0<br />

-20<br />

20<br />

0<br />

-20<br />

20<br />

0<br />

-20<br />

CV<br />

Metal Alloys(Pt 3M)<br />

Pt 3 Ni(111)<br />

Pt(111)<br />

E [V] vs. RHE<br />

Relative Expansion [%]<br />

i [�A/cm 2 ]<br />

i [�A/cm 2 �Hupd [%] ]<br />

i [�A/cm 2 ]<br />

�Hupd H O [%]<br />

2 2<br />

� Hupd [%]<br />

Hi [mA/cm O [%]<br />

2 2 2 ]<br />

i [mA/cm 2 ]<br />

(g) 0.0<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

Pt 3 Ni(100)<br />

Pt(100)<br />

0.1M HClO 4<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

E [V] vs. RHE<br />

Pt 3 Ni(110)<br />

Pt(110)<br />

-0.5<br />

60<br />

-1.0<br />

30<br />

-1.5<br />

0<br />

60<br />

-30<br />

60<br />

(h)<br />

-60<br />

30<br />

30<br />

0<br />

0<br />

-30 60<br />

-30<br />

-60 30<br />

-60<br />

0<br />

60<br />

50 60<br />

(i)<br />

30<br />

0<br />

30<br />

0.0 0.2 0.4 0.6 0.8 1.0 1.2<br />

E [V] vs. RHE<br />

-2<br />

0<br />

-4<br />

50<br />

-60<br />

0<br />

-2<br />

-4<br />

-6<br />

Pt[111]-Skin<br />

Pt[111]<br />

i = n F K(1-� s) exp(-γ�G i / RT)<br />

E [V] vs. RHE<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

E [V] vs. RHE<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

Pt [at. %]<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

a’)<br />

1 2 3 4 5 6 7<br />

Atomic layer<br />

20 E [V] vs. RHE<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

60<br />

80<br />

100<br />

Pt<br />

Pt3Ni(111) 3Ni(111) 0<br />

40<br />

Pt(111)<br />

Pt(111)<br />

0.0 0.2 0.4 0.6 0.8 0.1 1.0 HClO4 E [V] vs. RHE<br />

Pt poly<br />

60 o C<br />

1600 rpm<br />

III<br />

III<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

E [V] vs. RHE<br />

II<br />

Ni [at. %]<br />

Pt 3 Ni(111)<br />

Pt(111)<br />

II<br />

SXS<br />

0.1 HClO4 60<br />

Pt poly<br />

o 0.0 0.2 0.4 0.6 0.8 1.0<br />

E [V] vs. RHE<br />

C<br />

1600 rpm<br />

Potential/V<br />

I<br />

I<br />

a)<br />

b)<br />

c)<br />

60<br />

30<br />

0<br />

60<br />

30<br />

0<br />

� Oxd [%]<br />

� Oxd [%]<br />

� Oxd [%]


Tailoring Activity by Electronic Properties<br />

Specific Activity: ik @ 0.9V [mA/cm 2<br />

real ]<br />

Activity for ORR<br />

19<br />

18<br />

17<br />

Activity mapping<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

3.4<br />

Pt 3 Ti<br />

d-band center [eV]<br />

Pt 3 V<br />

3.0<br />

Pt 3 Fe<br />

Pt-skin surfaces<br />

Target Activity<br />

Pt 3 Ni(111)<br />

Pt 3 Co<br />

Pt-skeleton surfaces<br />

2.6<br />

Pt 3 Ni<br />

- 3.4 - 3.0 - 2.6<br />

d-band d-band center center shift [eV] (eV)<br />

Guiding Principles:<br />

(a)<br />

Pt-poly<br />

Pt/C<br />

3<br />

2<br />

1<br />

Activity improvement factor vs. Pt-poly<br />

Pt 3 Co<br />

Pt 3Fe<br />

Pt 3Ti<br />

1 st Layer<br />

2nd (b)<br />

Pt3Ni Layer<br />

3 rd Layer<br />

4th Layer<br />

Pt<br />

3.0 2.8 2.6<br />

-3.0 -2.8 -2.6<br />

Calculated d-band center [eV]<br />

Calculated d-band center [eV]<br />

6<br />

4<br />

2<br />

0<br />

Nanosegregated Profile<br />

6<br />

4<br />

2<br />

0<br />

Pt[111]-Skin surface<br />

Calculated activity [ 10 -2 eV]<br />

Calculated activity [10 -2 eV]<br />

Pt=100 at.%<br />

Pt=48 at.%<br />

Ni=52 at.%<br />

Pt=87 at.%<br />

Ni=13 at.%<br />

Pt=75 at.%<br />

Ni=25 at.%<br />

Nature Materials, 6(2007)241<br />

�Maximize activity by minimizing surface coverage of spectators<br />

�Without compromising activity protect 3d elements by additional Pt layer


Tailoring Electronic Properties<br />

Colloidal solvo - thermal approach for<br />

monodispersed PtM NPs with controlled<br />

size and composition<br />

Pt Lα<br />

Pt Co<br />

Co Kα<br />

Efficient surfactant removal<br />

1 o Particle size effect applies to Pt-bimetallic NPs<br />

Specific Activity increases with<br />

particle size: 3 < 4.5 < 6 < 9nm<br />

Mass Activity decreases with particle size<br />

Optimal size particle size ~5nm<br />

2o Temperature induced segregation in Pt-bimetallic NPs<br />

Agglomeration prevented<br />

(b)<br />

Optimized annealing temperature 400-500 o C<br />

3 o Surface chemistry of homogeneous Pt-bimetallic NPs<br />

Pt xM (1-x) NPs<br />

Dissolution of non Pt surface atoms leads to Pt-skeleton formation<br />

4o Composition effect in Pt-bimetallic NPs<br />

Pt3M 1 nm<br />

PtM PtM 2 PtM 3<br />

0.22 nm<br />

1 nm<br />

Optimal composition of Pt-bimetallic NPs is PtM<br />

1 nm<br />

9


a<br />

b<br />

From Model to Real Catalysts<br />

c<br />

b<br />

c<br />

d<br />

ed<br />

e<br />

Intensity (a. u.)<br />

counts<br />

Intensity (a. u.)<br />

counts<br />

1 nm<br />

0<br />

Intensity Counts<br />

Calculated Profile<br />

0<br />

0<br />

as-prepared acid leached acid leached/annealed<br />

Intensity Counts<br />

Calculated Profile<br />

1 nm 0 1 2 3 4 5<br />

Position (nm)<br />

Pt<br />

Ni<br />

Pt<br />

Ni<br />

1 2 3 4 5<br />

1 Position 2 3 (nm) 4<br />

Position (nm)<br />

5<br />

1<br />

2 3 4 5<br />

Position (nm)<br />

Intensity (a. u.)<br />

counts<br />

Intensity (a. u.)<br />

counts<br />

1 nm<br />

1 nm 0 1 2 3 4 5<br />

Position (nm)<br />

0<br />

0<br />

0<br />

1 2 3 4 5<br />

1 Position 2 3(nm) 4<br />

Position (nm)<br />

5<br />

1 2 3 4 5<br />

Position (nm)<br />

Au core<br />

PtFe shell<br />

counts<br />

counts<br />

0<br />

0<br />

1 2 3 4 5<br />

Position (nm)<br />

1 2 3 4 5<br />

Position (nm)<br />

Nano Letters, 11(2011)919-928<br />

5o 10<br />

0.2<br />

Pt-bimetallic catalysts with mutilayered Pt-skin surfaces<br />

b<br />

400<br />

RDE after 4K cycles @60o c<br />

before<br />

C (0.6-1.05V vs. RHE):<br />

Mass Activity (A/g)<br />

Specific Surface A<br />

Specific Activity (mA/cm 2 )<br />

30<br />

20<br />

0<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

8-fold 300 after<br />

6<br />

specific and 10-fold mass activity improvements over Pt/C<br />

100<br />

0<br />

before<br />

after<br />

Improvement Factor<br />

(vs. Pt/C)<br />

6<br />

4<br />

2<br />

0<br />

before<br />

Improvement Factor<br />

(vs. Pt/C)<br />

200Advanced<br />

Functional 4 Materials, 21(2011)147<br />

2<br />

0<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

after<br />

Pt/C acid leached acid leached/annealed<br />

PtNi/C PtNi/C<br />

Pt/C Au/PtFe<br />

c<br />

Mass Activity (A/g)<br />

Specific Activity (<br />

0.6<br />

0.4<br />

0.0<br />

400<br />

300<br />

200<br />

100<br />

0<br />

before<br />

after<br />

Pt/C acid leached<br />

PtNi/C<br />

Improvement<br />

(vs. Pt/C<br />

4<br />

2<br />

0<br />

Improvement Factor<br />

(vs. Pt/C)<br />

6<br />

4<br />

2<br />

0<br />

before<br />

14<br />

10<br />

0<br />

after<br />

acid leached/annealed<br />

PtNi/C<br />

6 o Multimetallic NPs can further improve activity and durability<br />

before<br />

After 60K cycles<br />

before<br />

After 60K cycles<br />

8<br />

4<br />

8


Pt (111)<br />

c)<br />

Beyond electronic effects<br />

- CN- - CN- - CN- - CN- - O 2<br />

Inorganic<br />

2- 3- 3- 3-<br />

- SO 4 / PO 4<br />

Keep electronic properties constant<br />

Maximize activity and selectivity by inert self assembled monolayer<br />

b)<br />

Inert Pt-CN ad adlayer<br />

Nature Chemistry 2 (2010) 880<br />

Molecular patterning<br />

Organic<br />

Inert Pt-Calix[4]arene adlayer<br />

Nature Materials, 9, 2010,881<br />

9


Selectivity controlled activity<br />

•ORR is strongly inhibited by adsorbed phosphoric acid anions<br />

E E [V [V vs. vs. RHE]<br />

RHE]<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

c)<br />

a)<br />

I II III<br />

Pt (111) I II III<br />

Pt (111)<br />

0.05 M H 3 PO 4<br />

Pt (111) - O – CN 2<br />

d)<br />

2- 3-<br />

- SO 4 / PO 4<br />

Pt (111)<br />

- CN- - CN- - CN- - CN- - M + - M + - M + - M +<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

E E [V [V vs. vs. RHE]<br />

RHE] - O 2<br />

- M + - M (H2O) x-OHad + - M (H2O) x-OHad + - M (H2O) x-OHad + (H2O) x-OHad 100<br />

50<br />

0<br />

-50<br />

-100<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

-6<br />

Current density [mA/cm2 ] Current density [mA/cm2 Current density [mA/cm ]<br />

2 ] Current density [mA/cm2 ]<br />

Current density [�A/cm2 Current density [�A/cm ] 2 ]<br />

Current density [�A/cm2 Current density [�A/cm ] 2 ]<br />

Pt (111)<br />

c)<br />

“Big vs. Small”<br />

- CN- - CN- - CN- - CN- - O 2<br />

2- 3-<br />

- SO 4 / PO 4<br />

•CN adlayer selectively blocks adsorption of anions<br />

b)<br />

Nature Chemistry 2 (2010) 880<br />

i = n F K (1-� s) exp(-γ�G i / RT)<br />

Current density [mA/cm2 ] Current density [mA/cm2 Current density [mA/cm ]<br />

2 ] Current density [mA/cm2 ]<br />

Current density [�A/cm2 Current density [�A/cm ] 2 ]<br />

Current density [�A/cm2 Current density [�A/cm ] 2 ]<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

E [V vs. RHE]<br />

E [V vs. RHE]<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

100<br />

50<br />

0<br />

-50<br />

-100<br />

a)<br />

b)<br />

I II III<br />

I II III<br />

Pt (111)<br />

Pt (111) – CN<br />

0.1 M HClO 4<br />

-6<br />

0.0 0.2 0.4 0.6 0.8 1.0<br />

E E [V [V vs. vs. RHE] RHE]<br />

0.0<br />

0.0<br />

18 10


Selectivity controlled stability<br />

a) b)<br />

c)<br />

40 nm<br />

2% of active sites<br />

Nature Materials, 9, 2010,881<br />

Improving stability of cathode catalysts<br />

20 nm<br />

Current density [�A/cm2 Current density [�A/cm ] 2 ]<br />

Chargedensity [�C/cm2 Chargedensity [�C/cm ] 2 ]<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

E [V vs. RHE]<br />

0.0 0.2 0.4 0.6 0.8<br />

100<br />

50<br />

0<br />

-50<br />

-100<br />

a)<br />

b)<br />

I II III<br />

Pt (111)<br />

Pt (111)-A<br />

Pt (111)-B<br />

Pt (111)-C<br />

Pt (111)-D<br />

-5<br />

0.0 0.2 0.4 0.6 0.8<br />

E [V vs. RHE]<br />

11


Tailoring Activity/Stability by Molecular Patterning<br />

�Maintain electronic properties and optimize stability by molecular patterning<br />

Extended<br />

O 2<br />

H 2<br />

Activity<br />

� Calix adlayer prevents adsorption of O 2 but not H 2<br />

Pt(111)-Calix[4]arene<br />

H 2 = 2H + + 2e -<br />

O 2 + 4H + +4e _ = H 2O<br />

Reversible potential (E)<br />

Nano<br />

�Selective anode catalysts for the ORR and HOR are of grate importance for stability of<br />

cathode materials in Fuel Cells<br />

14<br />

17


Three Phase (Nafion) Interface<br />

•What type of interactions are possible at three phase interfaces ?<br />

18


Interplay of Covalent, Electrostatic and Non-covalent Interactions<br />

Sulfonate anions<br />

are “specifically” adsorbed<br />

on Pt (covalent interactions)<br />

Pt-beckbone<br />

interaction is very<br />

week (electrostatic<br />

forces)<br />

Non-covalent<br />

cation-sulfonate interaction<br />

J. Phys. Chem., C. (2010) .<br />

26 19


The Spring Model<br />

J. Phys. Chem., C. (2010) .<br />

i = n F k (1-� anions.) exp(-�G i/RT)<br />

27 20


Perspectives<br />

� Understanding complexity (simplicity) of electrochemical interfaces all<br />

the way down to molecular and atomic levels will pave the way to:<br />

� Greatly improve our predictions for<br />

accelerating and directing chemical<br />

transformations<br />

� <strong>Dr</strong>amatic new energy production<br />

and storage technologies<br />

� Enable new mitigation strategies<br />

for environmental damage<br />

CARBON<br />

CYCLE<br />

WATER<br />

CYCLE<br />

Energy<br />

Conversion<br />

&<br />

Storage Group<br />

ENERGY<br />

Real<br />

16<br />

LITHIUM<br />

CYCLE<br />

16


Dusan Strmcnik,<br />

Chao Wang<br />

Dennis van der Vliet<br />

Ken Kodama<br />

Masa Uchimura<br />

Ram Subbaraman<br />

Nemanja Danilovic<br />

Donguo Li<br />

Pietro Lopez<br />

Voya Stamenkovic<br />

K-C Chang<br />

Gustav Wiberg<br />

Dusan Tripkovic<br />

Chris Lucas<br />

Jeff Greeley


Activity and Stability of platinum and platinum bi metal<br />

alloys for the oxygen reduction reaction<br />

Dusan Tripkovic 1 , Dusan Strmcnik 2 , Vojislav Stamenkovic 2 and<br />

Nenad M. Markovic 2<br />

1 Institute of Chemistry, Technology and Metallurgy – Department of Electrochemistry,<br />

University ofBelgrade, Belgrade,Serbia<br />

2 Materials Science Division, Argonne National Laboratory, Lemont, IL -60559<br />

Tripkovic_dusan@yahoo.com<br />

Much of the research activities in the last two decades have focused on the<br />

development of the polymer electrolyte membrane fuel cells (PEMFCs), a device that<br />

directly converts the chemical energy of hydrogen and oxygen into electrical energy.<br />

Many problems still need to be resolved in order for the PEMFCs to fully reach its<br />

commercial implementation – one of the major issues is the development of cathode<br />

materials that can simultaneously reduce the overpotential for the oxygen reduction<br />

reaction (ORR) and provide long-term stability in hostile electrochemical<br />

environments.<br />

A step toward solution of this foremost hurdle has recently been made by<br />

demonstrating a ninety-fold increase in specific activity (activity per surface Pt atom)<br />

on Pt3Ni(111) vs. the state-of-the-art Pt/C. It was proposed that this exceptional<br />

activity is governed by unique electronic (d-band position) and geometric (111-like<br />

symmetry) properties of nanosegregated (3 atomic layers) structures. Although high<br />

energy conversion efficiencies were achieved on the Pt (111)-skin surfaces, a<br />

fundamental knowledge of how to minimize the degradation of such nanosegregated<br />

structures is still lacking.<br />

Given that the stability of catalysts is the integral part of the materials design<br />

process, establishing the potential window of stability of the these nanosegregated<br />

structures is a major challenge that requires a deeper understanding of the oxideinduced<br />

morphological and activity changes<br />

Here, we present data for the ORR on Pt(111), Pt3Ni(111),<br />

Pt3Ni(1099)=20(111)x(100),Pt3Ni(1077)=8(111)x(100), and Pt3Ni(110) surfaces in<br />

0.1 M HClO4 to address the importance of surface morphology on activity and stability<br />

of Pt-skin atoms using electrochemical (cyclic voltammetry, analytical(ICPMS) and<br />

microscopic(STM) techniques.<br />

Learning from the differences and similarities between these systems could be<br />

a step forward in understanding whether the high activity of Pt3Ni(111) is primarily<br />

influenced by the subsurface concentration of Ni or by the geometry of Pt surface<br />

atoms.


High Activity and Stability of Pt2Bi Catalyst in Formic<br />

Acid Oxidation<br />

J.D. Lović a , M.D. Obradović a , D.V. <strong>Tripković</strong> a , K.Dj. Popović a ,<br />

V.M. Jovanović a , S.Lj. Gojković b and A.V. <strong>Tripković</strong> a,*<br />

a ICTM-Institute of Electrochemistry, University of Belgrade,<br />

N<strong>je</strong>goševa 12, P.O.Box 473, 11000 Belgrade, Serbia<br />

b Faculty of Technology and Metallurgy, University of Belgrade,<br />

Karnegi<strong>je</strong>va 4, P.O.Box 3503, 11000 Belgrade, Serbia<br />

*E-mail: amalija@tmf.bg.ac.rs<br />

Formic acid oxidation was studied on Pt2Bi catalyst characterized by<br />

XRD spectroscopy (phase composition), STM (surface morphology) and COads<br />

stripping voltammetry (surface composition). Bulk composition of Pt2Bi<br />

characterized by XRD revealed two phases: 55% PtBi alloy and 45% Pt.<br />

Estimated contribution of pure Pt on the Pt2Bi surface (43.5 %), determined by<br />

COads stripping voltammetry, corresponds closely to bulk composition and<br />

indicates clearly that adsorbed CO prevents leaching of Bi.<br />

-2<br />

j / mA cm R<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Pt 2 Bi<br />

Pt poly<br />

-300 -200 -100 0 100 200 300 400 500 600 700 800 900<br />

E / mV (SCE)<br />

Fig. 1. Cyclic voltammograms for the<br />

oxidation of 0.125 M HCOOH in 0.1 M<br />

H2SO4 solution on Pt2Bi and Pt<br />

catalysts. Scan rate of 50 mV s –1 .<br />

-2<br />

j / mA cm R<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

Pt poly<br />

Pt 2 Bi<br />

0<br />

0 400 800 1200 1600 2000<br />

Fig. 2. Chronoamperometric curves for<br />

the oxidation of 0.125 M HCOOH at 0.2<br />

V in 0.1 M H2SO4 solution on Pt2Bi and<br />

Pt catalysts.<br />

Pt2Bi exhibits high activity and stability in formic acid oxidation. High<br />

activity is caused by the fact that formic acid oxidation proceeds completely<br />

through dehydrogenation path induced by an ensemble effect. Stability of Pt2Bi<br />

surface in formic acid oxidation is based on preventing of Bi<br />

leaching/dissolution during formic acid oxidation as it was recognized by<br />

insignificant change of surface morphology and roughness demonstrated by<br />

STM images before and after electrochemical treatment in formic acid<br />

containing solution, as well as by the absence of surface poisoning by COads<br />

species.<br />

Pt2Bi is powerful catalyst for formic acid oxidation exhibiting up two<br />

order of magnitude larger current densities at 0.0 V and onset potential shifted<br />

for ~0.2 V to more negative value relative to Pt.<br />

t / s


Insight of Sn influence on formic acid oxidation at<br />

Pt based catalysts<br />

S. Stevanović 1 , D. <strong>Tripković</strong> 1 , V. <strong>Tripković</strong> 2 , D. Minić 3 , A.Gavrilović 4 , K. Popović 1 ,<br />

A. <strong>Tripković</strong> 1 and V.M. Jovanović 1<br />

1 ICTM, Department of Electrochemistry,University of Belgrade, N<strong>je</strong>goševa 12, Belgrade,<br />

2 Center for Atomic-scale Materials Design (CAMD), Department of Physics, Nano-DTU,<br />

Technical University of Denmark, Building 311, Kgs. Lyngby 2800, Denmark<br />

3 Faculty of Physical Chemistry, University of Belgrade, Studentski trg 12, Belgrade,Serbia<br />

4 CEST Centre of Electrochemical Surface Technology, Viktor-Kaplan-Str. 2,A-2700 Wiener<br />

Neustadt, Austria<br />

One of the extensively studied bimetallic catalysts is PtSn. It appears that<br />

among other catalysts PtSn is the most active for ethanol oxidation. The increased<br />

activity of Pt based catalyst with added Sn in comparison to pure Pt, is explained by an<br />

electronic effect in the Pt-based electrode material affecting on adsorption properties of<br />

the surface, making this system less prompt to poisoning by organic species than pure<br />

Pt. More importantly, Sn or its oxides can supply surface oxygen-containing species at<br />

lower potentials by activation of the interfacial water molecule necessary to complete<br />

the oxidation of adsorbed reaction intermediates through bi-functional mechanism.<br />

Thus, the addition of Sn is mainly related to its influence on CO. CO is the main poison<br />

of polycrystalline Pt in formic acid oxidation, yet this reaction has been rarely studied<br />

on Pt-Sn catalyst.<br />

The aim of our work is to overview the influence of Sn on the oxidation of<br />

formic acid by studding the reaction when Sn is present only on the surface of Pt (Sn<br />

UPD), when Sn is added to Pt in PtSn catalyst and when Sn is leached out from<br />

surface layers of this catalyst – skeleton structure of PtSn. Both Pt and PtSn catalysts<br />

were prepared by microwave assisted pollyol method and characterized by STM, XRD<br />

and EDX techniques. Oxidation of formic acid was studied in 0.1 M HClO4 + 0.5 M<br />

HCOOH solution by cyclic voltammetry (sweep rate 50 mV/s) and chromoamperometry<br />

at 0.2 V vs SCE.<br />

STM analysis of unsupported catalysts showed small spherical particles of<br />

almost equal size below 2 nm. XRD confirmed small particle sizes for both catalysts<br />

and indicated low alloying degree of Pt and Sn, while EDX revealed almost nominal<br />

Pt:Sn ratio.<br />

If Sn is present on the surface either as ad-atom on Pt or as less noble metal in<br />

PtSn catalyst, the onset potential of formic acid oxidation is shifted to the lower values<br />

and currents are significantly higher in comparison to Pt, thus the reaction is shifted<br />

towards direct path. However, on skeleton surface of PtSn catalyst, when Sn should be<br />

only in subsurface layers, the reaction proceeds as on pure Pt upto 0.3 V, but then<br />

contrary to pure Pt the current continues to increase to the maximum at potential which<br />

is more negative than the second peak at pure Pt. The results are discussed and<br />

explained by Sn influence through bi-functional mechanism, morphology changes and<br />

electronic effect.


Ehanol oxidation on carbon supported platinum based<br />

bimetallic catalysts synthesized by microwave assisted<br />

polyol procedure<br />

S. Stevanović 1 , D. <strong>Tripković</strong> 1 , J. Rogan 2 , J. Lović 1 , K. Popović 1 , A. <strong>Tripković</strong> 1 and<br />

V.M. Jovanović 1<br />

1 ICTM- Department of Electrochemistry, University of Belgrade,<br />

N<strong>je</strong>goševa 12, Belgrade,<br />

2 Faculty of Technology and Metallurgy, University of Belgrade,<br />

Karnegi<strong>je</strong>va 4, Belgrade,<br />

Ethanol oxidation was studied in perchloric acid at high surface area<br />

carbon supported Pt/C, PtRh/C and PtSn/C catalysts prepared by microwave<br />

assisted polyol procedure. The catalysts were characterized by XRD, STM and<br />

EDX techniques. The diffraction peaks of the bimetallic catalysts in X-ray<br />

diffraction patterns are slightly shifted to lower (PtSn/C) or higher (PtRh/C) 2θ<br />

values with respect to the corresponding peaks at Pt/C catalyst as a<br />

consequence of alloy formation. Thus, low alloyed PtSn and PtRh catalysts<br />

were obtained. STM analysis of unsupported catalysts shows that the particles<br />

are small (diameter is ~ 2 nm) with a narrow size distribution. This should be<br />

attributed to the advantages of microwave assisted polyol process in ethylene<br />

glycol solution.<br />

j (mA/mg Pt )<br />

210<br />

180<br />

150<br />

120<br />

90<br />

60<br />

30<br />

0<br />

Pt/C<br />

PtRh/C<br />

PtSn/C<br />

-30<br />

-0.2 -0.1 0.0 0.1 0.2 0.3<br />

E (V vs SCE)<br />

Fig. 1: Potentiodynamic curves for the<br />

oxidation of 0.5 M C2H5OH at asprepared<br />

Pt/C, PtRh/C and PtSn/C<br />

catalysts in 0.1 M HClO4, � = 20 mV/s.<br />

It is found that the activity of Pt-based<br />

bimetallic catalyst for ethanol oxidation<br />

greatly depends on secondary metal<br />

and the electrode potential. While<br />

addition of Sn to Pt leads to the<br />

enhancement of ethanol oxidation and<br />

lower poisoning of electrode surface,<br />

addition of Rh does not change overall<br />

electrochemical reaction. The onset<br />

potential for ethanol oxidation at our<br />

PtSn/C catalyst is shifted for ~ 150 mV<br />

to more negative values and the<br />

increase of activity for ~ 3 times in<br />

comparison to Pt/C catalyst is<br />

obtained. The onset potential for this<br />

reaction is remarkably shifted to lower<br />

values in comparison to other PtSn<br />

catalysts with similar composition.<br />

Chronoamperometric measurements<br />

revealed that PtSn/C is notably less poisoned than Pt/C catalyst. The increased<br />

activity of PtSn/C catalyst is mainly due to bifunctional mechanism enabled<br />

most probably by non-alloyed Sn (SnO2) although electronic effect of low<br />

alloyed PtSn could play some role as well.


Advanced Multimetallic Electrocatalysts<br />

Vojislav R. Stamenkovic, Dennis van der Vliet, Dusan Strmcnik, Dusan Tripkovic,<br />

Chao Wang, Nenad M. Markovic<br />

Materials Science Division, Argonne National Laboratory<br />

9700 S. Cass Ave. Argonne, IL 60439, USA<br />

vrstamenkovic@anl.gov<br />

Recently a lot of effort has been placed on the synthesis of alloy nanoparticles<br />

(NPs) for applications in heterogeneous catalysis. 1-3 While it is usually challenging to<br />

reach the optimal catalytic performance for a given chemical reaction by singlecomponent<br />

materials, alloys offer an additional dimension of tailoring electronic and<br />

surface structures of catalysts via ensemble, strain, and/or electronic (ligand) effects.<br />

Fine tuning of these parameters provides the opportunity for catalyst design to reach a<br />

fine balance among activity, selectivity and durability. Bimetallic Pt3M (M = Fe, Co,<br />

Ni, etc.) alloys have been extensively studied and shown to be better catalysts than pure<br />

Pt for the ORR. 2 4 It was first established on well-defined extended surfaces that these<br />

alloys have lower d-band center versus Pt, resulting in reduced binding strength and<br />

adsorption of oxygenated spectator species (e.g., OH − ) on the surface and thus more<br />

active sites accessible for molecular oxygen. A volcano-type dependence of the<br />

electrocatalytic activity on the electronic structures has been revealed, which was<br />

further validated in monodisperse and homogeneous high-surface-area nanocatalysts<br />

prepared by organic solution synthesis. Our successful experience in such<br />

combinational studies of extended surfaces and nanocatalysts have urged us to explore<br />

novel and more complicated systems in the quest for better electrocatalysts.<br />

In this work we have carried out a comprehensive investigation of Pt-ternary<br />

(Pt3(MN)1 with M, N = Fe, Ni or Co) alloy electrocatalysts for the ORR. An organic<br />

solvothermal approach was developed for the synthesis of homogeneous Pt-ternary<br />

alloy NPs. Element mapping based on high-resolution scanning transmission electron<br />

microscopy (STEM) was used to analyze the alloy homogeneity in the NPs. The<br />

obtained NPs were supported on carbon black and then sub<strong>je</strong>ct to electrocatalytic<br />

studies. We also did fundamental studies of extended surfaces for these ternary alloys, 5<br />

in order to confirm the dependence of ORR activity on the alloying elements in the<br />

NPs. The obtained composition-function correlation was further compared to the<br />

results by theoretical simulations, and the previously established relationship for Ptbimetallic<br />

catalysts.<br />

[1]. Greeley, J.; Mavrikakis, M. Nat. Mater. 2004, 3, 810-815.<br />

[2]. Strasser, P.; Koh, S.; Anniyev, T.; Greeley, J.; More, K.; Yu, C. F.; Liu, Z. C.;<br />

Kaya, S.; Nordlund, D.; Ogasawara, H.; Toney, M. F.; Nilsson, A. Nature Chem. 2010,<br />

2, 454-460<br />

[3]. Stamenkovic, V. R.; Mun, B. S.; Arenz, M.; Mayrhofer, K. J. J.; Lucas, C. A.;<br />

Wang, G. F.; Ross, P. N.; Markovic, N. M. Nat. Mater. 2007, 6, 241-247.<br />

[4] Stamenkovic, V. R.; Fowler, B.; Mun, B. S.; Wang, G. F.; Ross, P. N.; Lucas, C.<br />

A.; Markovic, N. M. Science 2007, 315, 493-497


Electrocatalysts with Multifunctional Properties<br />

Dennis van der Vliet, Dusan Strmcnik, Dusan Tripkovic, Chao Wang, Nenad M.<br />

Markovic, Vojislav R. Stamenkovic,<br />

Materials Science Division, Argonne National Laboratory<br />

9700 S. Cass Ave. Argonne, IL 60439, USA<br />

vrstamenkovic@anl.gov<br />

The ability to alter the electronic and structural properties of electrocatalysts can<br />

lead towards superior catalytic activity. This can be achieved by controlling the size<br />

and shape of nanoparticles, the introduction of another constituent into the catalyst<br />

and/or tailoring the surface morphology. The later has been in focus of this report and<br />

can be controlled either by preparation protocol or electrochemical conditions. The<br />

concept of structure function relationships has been used to investigate the influence of<br />

surface atomic arrangement on the electrocatalytic properties of the surface 1 . By<br />

comparing the rates of an electrocatalytic reaction one is able to draw conclusions<br />

about the reaction mechanism as well as the properties of the material needed to boost<br />

the reaction rate 2-5 . However, in this contribution, we focus on the different atomic<br />

arrangements within the same surface, i.e. on the local surface structure of the catalyst.<br />

We utilized ex- and in-situ tools to characterize and tune the morphology of the catalyst<br />

surfaces. That produced substantial differences in catalytic properties. Moreover, we<br />

demonstrate that surface adatoms can control the overall electrochemical behavior of a<br />

catalyst. For that reason, the knowledge about the active sites for a particular reaction<br />

is a prerequisite and the first step to understanding how to tune a catalyst performance.<br />

In order to elucidate the real active sites we used Pt single crystal surafces to establish<br />

the reaction rates for CO bulk electrooxidation and oxygen reduction reaction as a<br />

function of the applied positive potential limits and surface morphology. We managed<br />

to control the presence of different active sites for the same catalysts and thereofre, we<br />

induced multifunctional behavior. The same analogy was used for nanoscale Pt<br />

catalysts and we proved that multifunctionality of electrocatalyst could be tailored.<br />

References:<br />

[1]. N.M.Markovic, P.N.Ross, Surf. Sci. reports, 45, 117 (2002).<br />

[2]. M. Arenz, K.J.J. Mayrhofer, V. Stamenkovic, B.B. Blizanac, T. Tomoyuki, P.N.<br />

Ross, N.M. Markovic, J.Am.Chem.Soc. 127 6819 (2005)<br />

[3]. N.P. Lebedeva, M. Koper, E. Herrero, J.M. Feliu, R.A. van Santen,<br />

J.Electroanal.Chem., 487, 37-44 (2000).<br />

[4]. N.P. Lebedeva, A. Rodes, J.M. Feliu, M.T.M. Koper, R.A van Santen,<br />

J.Phys.Chem.B, 106, 9863-9872 (2002)<br />

[5] Stamenkovic, V. R.; Fowler, B.; Mun, B. S.; Wang, G. F.; Ross, P. N.; Lucas, C.<br />

A.; Markovic, N. M. Science 2007, 315, 493-497


_n~::;::;D::;n"""':::'>-;J"'"<br />

p:J~""~,p:J>«1)~r::-,-:=<br />

nO:;:j.~<br />

Ox-'::;<br />

'JJ"O-;:Jnn;:J(1)OIOU~<br />

('):::..r::~'Q..(1)(1)<br />

0,..';::;:;3' r. if, Q.. _. -,<br />

-"";::'.'"'\<br />

::::~CTn~:JO (1) \0.1 (1) (1) en r:: 0..<br />

;:;_.::J",<br />

-'0..'---'-<br />

';?~;::C3<br />

:: _. '. _.<br />

o:f7<br />

~ :;.<br />

Nr+:J-<br />

(1) _. - ,< 0<br />

::::(1)::;,...<br />

en Q.. 1';:;<br />

--c.o;::;en<br />

is::;:::r 0<br />

:::..5(1)-'J;r-CT~~CT~<br />

;, ~ reo :!. (1) &; 0 ;) c::<br />

- ~ Q.. 'J; x - ~ ::;..<br />

;::: ~ o'<br />

to °:=<br />

-,~~ ~~.:<br />

°:::r -'<br />

s<br />

-'<br />

VJo9:~~';:';::a-1~o..<br />

u; 3<br />

""0,...,.:::;2<br />

;::'. a ~ ~ (6 Q.. ::; ;:::. c:: ::; 3 u<br />

-Oo;-.::J..-,..,- (1);:;::; nO,u<br />

-'-:J'J;(1)0°-t;,- '" ,..,


~<br />

Physical ('hem istry 20 I 0<br />

u-<br />

be mentioned also that Pt/C catalyst synthesized by microwave assisted r<br />

method is ITlOre active then commercial Pt/C-Tanaka catalyst probably due<br />

belter dispersion, However, the activity of PtjSn/C catalyst synthesized by<br />

method is lower in comparison with activity of commercial PtjSn/C-T<br />

catalyst [5], The higher activity of Pt]Sn/C-Tanaka catalyst most likely<br />

consecvcnceof higher alloying degree of this catalyst.<br />

140 -------.---..----<br />

i<br />

-N-<br />

- ---<br />

Pt/C-Tan<br />

Pt/C-MW.aka<br />

I'<br />

120 i -- PI Sn/C-MW<br />

, " 1<br />

~ 100<br />

ci<br />

~ 80<br />

~ "<br />

Fig.I.."<br />

th~ c1ectrooxidation of m<br />

C2H5OH in 0,\ M HCI04,<br />

60 :'<br />

, ,<br />

" ,.<br />

40 : /<br />

, ,<br />

: ,<br />

, ,-<br />

W ,~'<br />

"<br />

,,-<br />

: 0 , / ~,.~-_.":>< /'<br />

-20<br />

"<br />

;:<br />

-0-2 E pJi»s SCE) 0,2 04<br />

Conclusion<br />

The presented results point out the enhanced activity of bimetalic catalyst for<br />

electrooxidation of ethanol which may be mainly ascribed to changes in el<br />

structure due to the al1oy formation.<br />

References<br />

[1] J. Luo, N. Kariuki, L. Han, L. Wang, C. J. Zhong, T. He Electrochil1l.<br />

2006,51,4821. -<br />

L. Colmenares,H. Wang, Z.Jusys, L. Jiang, S. Van, G. Q. Sun, R. .I.<br />

Elcctrochil1l. Acta, 2006, 52, 221.<br />

F. Colmati, E. Antolini, E. R. Gonzalez, J. Alloy and Compounds, 2008,<br />

7(14<br />

Z. Liu, B. Guo, L. Hong, T. H. Lim, Electrochem. Communicatior<br />

83.<br />

A. V. Tripkovic, K. OJ. Popovic, .I. O. Lovic, V. M. .Iovan!<br />

Stevanovic, O. V. Tripkovic, A. Kowal, Electrochem. Comm,.<br />

lmo<br />

270<br />

MANGANESE DIOXII<br />

OL Y ANILINE NANOS<br />

APPLICA TION<br />

B. SUuki~, l. Stojkovic. i'<br />

iversity of' Be/grade, F'ocu/t,/ of<br />

B<br />

stract<br />

nganese dioxide modified carb<br />

,pared via a novel hydrothen<br />

aracterised and examined 1'01' pc<br />

owing a high e\ectrocawlytic a<br />

ueous media,<br />

troduction<br />

recent times, special attention<br />

ide-based catalysts as an altern,<br />

n02) in particular is interesti<br />

rnbination of beneficial ph)<br />

!though having excellent activi<br />

cry low surface areas. Loading<br />

direct synthesis or post graftil<br />

irect synthesis of the oxide onte<br />

ctive phase and the support.<br />

if anoPAN!) were used as a Ii<br />

", .'<br />

,(communICatiOn.<br />

f<br />

~i(J!:xperimental<br />

:~,Carb-nanoPANI was prepared b<br />

!\f)11odification of Carb-nanoPAN<br />

',procedure. First, a reaetion mix!<br />

';in 25 m! of aqueous solution<br />

manganese sulphate (0.06 M) '1<br />

. lined reactor for 15 h at 135"C<br />

". separated from the solution<br />

hom!';

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!