Anisotropic 18O tracer diffusion in epitaxial films of GdBaCo2O5+δ cathode material with different...

7
Anisotropic 18 O tracer diusion in epitaxial lms of GdBaCo 2 O 5+d cathode material with dierent orientations James Zapata, a onica Burriel, * b Pablo Garc´ ıa, a John A. Kilner bc and Jos´ e Santiso * a The layered structure of the orthorhombic GdBaCo 2 O 5+d (GBCO) double perovskite compound, currently considered as a promising cathode material in Solid Oxide Fuel Cells (SOFCs), is believed to induce a high degree of anisotropy in the oxygen diusion coecient, being maximum along the ab plane in comparison to the diusion along the c-axis direction. In this study we have deposited lms with dierent orientation: pure c-axis and a-axis orientation on SrTiO 3 (001) and NdGaO 3 (110) single crystals, respectively. The oxygen diusion was analysed by isotopic 18 O exchange depth proling (IEDP) and Time-of-ight Secondary Ion Mass Spectrometry (ToF-SIMS) in the lms along the longitudinal and transverse directions at dierent exchange temperatures and exposure times. The magnitude of longitudinal D* at low temperatures shows a clear anisotropy. The oxygen diusion along the a-axis shows comparable values to the bulk polycrystalline GBCO, while it is about one order of magnitude lower along the c-axis of the structure. The corresponding oxygen surface exchange rates k* do not show any anisotropy having comparable values for c-axis and a-axis orientation. These k* values are slightly larger than those reported for bulk material showing that thin lm textured cathodes may have enhanced activity for oxygen reduction at low temperatures. Introduction It is generally believed that one of the limiting steps to reduce the working temperatures of current Solid Oxide Fuel Cell (SOFC) devices is the overpotential caused by the low activity of the cathode reactions, which decreases the performance of the whole device. In this sense there is an intense activity in the quest for active cathode materials which present at the same time large electronic conductivities, high oxygen diusion and large surface oxygen exchange rates at moderate temperatures. Among the possible compound candidates are perovskite oxides based on substituted (A,A 0 )(B,B 0 )O 3d with A,A 0 ¼ La, Ba, Sr and B,B 0 ¼ Co, Fe, as well as oxides with perovskite-related structure members of the RuddlesdenPopper A n+1 B n O 3n+1 series, such as La 2 NiO 4+d (n ¼ 1) and others. 1,2 Recent work has attracted attention to one of these compounds, namely the double perovskite GdBaCo 2 O 5+d (GBCO), which has shown very high oxygen diusion coecients D* (similar to those of pure ionic conducting materials such as YSZ) and large oxygen surface exchange rates k*, 3 along with superior chemical stability against the typical CeO 2 electrolytes. 4 In this compound Gd and Ba cations alternate in sequential A-site layers along the c-axis of the structure, whereas Co ions occupy the B-sites. The stoichiometric material under standard conditions of pO 2 and T holds a substantial oxygen deciency with oxygen vacancies located mostly on the GdO layers of the structure, parallel to the ab plane, and arranged forming channels along the crystallographic a-axis 5,6 and d changing from 0 to 1. This gives rise to a superstructure that doubles along the c- and b-axis and forms a double perovskite a 2a 2a cell. The structure of the GBCO material is reported to be orthorhombic Pmmm with a ¼ 3.862 ˚ A, b/2 ¼ 3.934 ˚ A and c/2 ¼ 3.786 ˚ A. 6 The mechanism for the fast oxygen diusion in this material is reported to involve mainly those oxygen vacancy sites arranged forming parallel channels along the a-axis, and is therefore expected to be highly anisotropic, as it has been inferred from molecular dynamics calculations, 7 the oxygen diusivity being much larger in the ab planes of the structure when compared to the diusivity along the c-axis. One of the current techniques to probe the oxygen surface exchange and diusion kinetics in solids is isotope exchange depth proling by secondary ion mass spectrometry (IEDP- SIMS) 8,9 This has been successfully used for the characterisation of thin epitaxial lms of dierent cathode materials: La 2 NiO 4+d , 10 Ba 0.5 Sr 0.5 Co 0.8 Fe 0.2 O 3d , 11 La 0.7 Sr 0.3 MnO 3 , 12 as well as La 1x Sr x Co 3d . 13 a Centro de Investigaci´ on en Nanociencia y Nanotecnolog´ ıa, CIN2 (CSIC-ICN), Campus UAB, 08913 Bellaterra, Barcelona, Spain. E-mail: [email protected]; Tel: +34 93 7373634 b Department of Materials, Imperial College London, London SW7 2AZ, UK. E-mail: m. [email protected] c International Institute for Carbon Neutral Energy Research, 744 Motooka, Nishi-ku, Fukuoka 819-0395, Japan Cite this: J. Mater. Chem. A, 2013, 1, 7408 Received 20th February 2013 Accepted 15th April 2013 DOI: 10.1039/c3ta10749c www.rsc.org/MaterialsA 7408 | J. Mater. Chem. A, 2013, 1, 74087414 This journal is ª The Royal Society of Chemistry 2013 Journal of Materials Chemistry A PAPER Published on 18 April 2013. Downloaded by UNIVERSITAT GIESSEN on 28/10/2014 16:14:06. View Article Online View Journal | View Issue

Transcript of Anisotropic 18O tracer diffusion in epitaxial films of GdBaCo2O5+δ cathode material with different...

Journal ofMaterials Chemistry A

PAPER

Publ

ishe

d on

18

Apr

il 20

13. D

ownl

oade

d by

UN

IVE

RSI

TA

T G

IESS

EN

on

28/1

0/20

14 1

6:14

:06.

View Article OnlineView Journal | View Issue

aCentro de Investigacion en Nanociencia y N

UAB, 08913 Bellaterra, Barcelona, Spain.

7373634bDepartment of Materials, Imperial College L

[email protected] Institute for Carbon Neutral

Fukuoka 819-0395, Japan

Cite this: J. Mater. Chem. A, 2013, 1,7408

Received 20th February 2013Accepted 15th April 2013

DOI: 10.1039/c3ta10749c

www.rsc.org/MaterialsA

7408 | J. Mater. Chem. A, 2013, 1, 74

Anisotropic 18O tracer diffusion in epitaxial films ofGdBaCo2O5+d cathode material with differentorientations

James Zapata,a Monica Burriel,*b Pablo Garcıa,a John A. Kilnerbc and Jose Santiso*a

The layered structure of the orthorhombic GdBaCo2O5+d (GBCO) double perovskite compound, currently

considered as a promising cathode material in Solid Oxide Fuel Cells (SOFCs), is believed to induce a

high degree of anisotropy in the oxygen diffusion coefficient, being maximum along the a–b plane in

comparison to the diffusion along the c-axis direction. In this study we have deposited films with

different orientation: pure c-axis and a-axis orientation on SrTiO3(001) and NdGaO3(110) single crystals,

respectively. The oxygen diffusion was analysed by isotopic 18O exchange depth profiling (IEDP) and

Time-of-flight Secondary Ion Mass Spectrometry (ToF-SIMS) in the films along the longitudinal and

transverse directions at different exchange temperatures and exposure times. The magnitude of

longitudinal D* at low temperatures shows a clear anisotropy. The oxygen diffusion along the a-axis

shows comparable values to the bulk polycrystalline GBCO, while it is about one order of magnitude

lower along the c-axis of the structure. The corresponding oxygen surface exchange rates k* do not

show any anisotropy having comparable values for c-axis and a-axis orientation. These k* values are

slightly larger than those reported for bulk material showing that thin film textured cathodes may have

enhanced activity for oxygen reduction at low temperatures.

Introduction

It is generally believed that one of the limiting steps to reducethe working temperatures of current Solid Oxide Fuel Cell(SOFC) devices is the overpotential caused by the low activity ofthe cathode reactions, which decreases the performance of thewhole device. In this sense there is an intense activity in thequest for active cathode materials which present at the sametime large electronic conductivities, high oxygen diffusion andlarge surface oxygen exchange rates at moderate temperatures.Among the possible compound candidates are perovskiteoxides based on substituted (A,A0)(B,B0)O3�d with A,A0 ¼ La, Ba,Sr and B,B0 ¼ Co, Fe, as well as oxides with perovskite-relatedstructure members of the Ruddlesden–Popper An+1BnO3n+1

series, such as La2NiO4+d (n ¼ 1) and others.1,2

Recent work has attracted attention to one of thesecompounds, namely the double perovskite GdBaCo2O5+d (GBCO),which has shown very high oxygen diffusion coefficients D*(similar to those of pure ionic conducting materials such as YSZ)

anotecnologıa, CIN2 (CSIC-ICN), Campus

E-mail: [email protected]; Tel: +34 93

ondon, London SW7 2AZ, UK. E-mail: m.

Energy Research, 744 Motooka, Nishi-ku,

08–7414

and large oxygen surface exchange rates k*,3 along with superiorchemical stability against the typical CeO2 electrolytes.4 In thiscompound Gd and Ba cations alternate in sequential A-site layersalong the c-axis of the structure, whereas Co ions occupy theB-sites. The stoichiometricmaterial under standard conditions ofpO2 and T holds a substantial oxygen deciency with oxygenvacancies located mostly on the GdO layers of the structure,parallel to the a–b plane, and arranged forming channels alongthe crystallographic a-axis5,6 and d changing from 0 to 1. Thisgives rise to a superstructure that doubles along the c- and b-axisand forms a double perovskite a � 2a � 2a cell. The structure ofthe GBCO material is reported to be orthorhombic Pmmm witha ¼ 3.862 A, b/2 ¼ 3.934 A and c/2 ¼ 3.786 A.6

The mechanism for the fast oxygen diffusion in this materialis reported to involve mainly those oxygen vacancy sitesarranged forming parallel channels along the a-axis, and istherefore expected to be highly anisotropic, as it has beeninferred from molecular dynamics calculations,7 the oxygendiffusivity being much larger in the a–b planes of the structurewhen compared to the diffusivity along the c-axis.

One of the current techniques to probe the oxygen surfaceexchange and diffusion kinetics in solids is isotope exchangedepth proling by secondary ion mass spectrometry (IEDP-SIMS)8,9 This has been successfully used for the characterisationof thin epitaxial lms of different cathode materials:La2NiO4+d,10 Ba0.5Sr0.5Co0.8Fe0.2O3�d,11 La0.7Sr0.3MnO3,12 as wellas La1�xSrxCo3�d.13

This journal is ª The Royal Society of Chemistry 2013

Paper Journal of Materials Chemistry A

Publ

ishe

d on

18

Apr

il 20

13. D

ownl

oade

d by

UN

IVE

RSI

TA

T G

IESS

EN

on

28/1

0/20

14 1

6:14

:06.

View Article Online

Therefore, in order to experimentally explore the anisotropicdiffusion along different crystallographic directions it isnecessary to study either single crystals14 or highly textured(epitaxial) thin lms. Recent studies on polycrystalline ceramicsamples of the praseodymium double perovskite compoundPrBaCo2O5.5 have been able to experimentally probe theanisotropy in the oxygen diffusion which is at least four timeslarger in crystallites with [100] or [010] orientation than in thosewith [001] orientation.15 The intrinsic anisotropy might be evenlarger, as the measurements were performed on polycrystallinesamples with small grain size, and therefore there were exper-imental limitations associated to inter-grain diffusion and tothe random orientation of the grains. Previous attempts to growepitaxial GBCO lms by Pulsed Laser Deposition (PLD) fromstoichiometric targets showed a tendency to form Co-decientlms. This induced the generation of planar defects consistingof supplementary GdO layers, which had an important impacton the charge transport in the lms.16 The purpose of this studyis to analyse the oxygen surface exchange and diffusion processon epitaxially grown stoichiometric GBCO lms with differentorientations. With this aim the composition of the PLD targetwas varied to obtain stoichiometric lms. We make use of the18O tracer diffusion experiments in both longitudinal andtransverse directions, using the same methodology as previ-ously described for La2NiO4 epitaxial lms.17

Experimental

Thin lms of GBCO were grown by the Pulsed Laser Deposition(PLD) technique using a Compex Pro 201 KrF excimer laser(248 nm) at 10 Hz pulse repetition at a laser uence of around3 J cm�2 (100 mJ in 4.35 mm2 spot size). The target used wassynthesized by conventional solid state reaction using highpurity commercial (Alfa Aesar) oxides and carbonates asprecursor materials, that is, Gd2O3 (99.99%), BaCO3 (99.997%),and Co3O4 (99.9985%) oxides (with an excess of 5% cobaltadded to compensate the deviations in the cation compositionpresent in layers deposited using a stoichiometric target7). Aera de-carbonation process at 900 �C for 20 h the powder waspressed into a circular pellet and annealed over 20 h at 1000 �Cwith a nal heat treatment at 1250 �C for 3 h to densify it.

Films about 160 nm thick (which correspond to 4000 abla-tion pulses) were deposited at a substrate temperature of 850 �Cin a 60 mTorr atmosphere of pure O2 and with the target–substrate distance xed at about 55 mm. Substrates werecommercial single crystals of SrTiO3(001) (STO) andNdGaO3(110) (NGO) (from CrysTec GmbH). Aer deposition thelms were cooled down under the same pO2 and subsequentlyinvestigated by X-Ray Diffraction (XRD) in q/2q congurationusing a PANalytical X'Pert Pro-MRD instrument (Cu Ka radia-tion and PIXel detector). The thickness of the GBCO layers wasestimated as proportional to the number of pulses aer aprevious calibration on thinner GBCO lms measured by X-rayreectometry and scanning electron microscopy (SEM). For thecomposition analysis of the lms on the STO substrate, we useda Cameca SX-50 Microprobe (with four crystal analysers forWDS). The typical voltages used were 12, 15 and 20 keV. A thin

This journal is ª The Royal Society of Chemistry 2013

lm analysis program (STRATAGem) was used to determine theexact composition and thickness of the thin lms withoutinterference of the substrate composition.

Film texture along with in- and out-of-plane lattice parame-ters were accurately determined performing X-ray reciprocalspace maps (RSMs) of both lm and substrate asymmetricreections.

The oxygen isotope exchange depth prole (IEDP) techniquewas used to measure the oxygen surface exchange and diffusionproperties of GBCO lms in transverse and longitudinalcongurations (perpendicular and parallel to the lm surface,respectively). The isotopic exchange was carried out at fourdifferent temperatures ranging from 300 to 600 �C.

The 18O isotopic exchange experiment was performed usingthe same equipment and methodology as described in aprevious study. In this case the 18O-enriched gas compositionwas 27.8%. The effective 18O exchange time was calculatedtaking into account the nite increase/decrease in the sampletemperature.17 In the case of the transverse proles, the repor-ted surface exchange values were used in the exchanged timecorrection, since this is the rate-limiting step for oxygen trans-port in the thin lms, while in the case of longitudinal proles,the corresponding oxygen diffusion values were used.

In order to determine the oxygen transport along the lmplane, i.e. longitudinal conguration, a dense and uniform Authin lm of approximately 160 nm was sputtered to cover theGBCO epitaxial lm and prevent oxygen exchange from the topsurface. A trench was opened at the lateral edge of the lm,allowing the exchange with the 18O enriched exchange gasphase and thus ensuring the diffusion of the 18O species onlyalong the direction parallel to the lm plane.18

All the exchanged samples were measured by time-of-ightsecondary ion mass spectrometry (ToF-SIMS) on a Ion ToF-SIMS5 machine (ION-TOF GmbH, Munster, Germany) equipped witha bismuth LMIS pulsed gun incident at 45�. A 25 kV Bi+ primaryion beam was used to generate the secondary ions using the lowmass resolution or burst alignment mode (eight pulses) foranalysis and a Cs+ beam (2 kV) incident at 45� for sputtering. Forthe longitudinal conguration secondary ion images forselected areas of the GBCO lm, including the edge of theopened trench, were acquired. For the transverse conguration(perpendicular to the lm plane) tracer oxygen depth proleswere measured from the exchanged surface penetrating down tothe substrate by sputter depth proling.

Results and discussionFilm orientation

Fig. 1(a and b) show typical X-ray diffraction patterns obtainedfor GBCO lms grown at a xed oxygen partial pressure pO2 of60 mTorr and 850 �C on STO(100) and NGO(110) substrates,respectively. On STO the patterns show one set of 00l GBCOreections with an out-of-plane parameter of about 3.775 A,consistent with the c/2 value of the bulk GBCO material. Thepresence of intense reections in the intermediate positionsbetween the main substrate reection (indexed as l ¼ odd) isalso characteristic of the GBCO double perovskite

J. Mater. Chem. A, 2013, 1, 7408–7414 | 7409

Fig. 1 XRD patterns of GBCO films deposited at 850 �C by PLD on STO (001) (a)and NGO (110) (b) substrates. Intense substrate peaks are labelled with asterisks(STO) and triangles (NGO). Reciprocal space maps of those films on STO (c) andNGO (d) in the area of the pseudocubic primitive �303 reflection (3�36 for the NGO)reveal the orientation of GBCO film being c-axis oriented on STO and a-axisoriented on NGO. The schemes in (e) and (f) depict the corresponding epitaxialorientation between the film and the substrate on STO and NGO, respectively(models have been generated using Vesta software.25 Coloured spheres in GBCOcorrespond to Gd, purple, and Ba, green, while Sr is green and Nd is golden in STOand NGO, respectively).

Journal of Materials Chemistry A Paper

Publ

ishe

d on

18

Apr

il 20

13. D

ownl

oade

d by

UN

IVE

RSI

TA

T G

IESS

EN

on

28/1

0/20

14 1

6:14

:06.

View Article Online

superstructure arising from the Gd/Ba sequential A-sitearrangement in the [001] direction. No additional peaks corre-sponding to other phases or orientations appear under thesedeposition conditions. Therefore, the lms on STO correspondto purely c-axis oriented GBCO. This texture was also conrmedby the reciprocal space maps around the �303 STO reection(Fig. 1c). The map shows a lm reection at Qy ¼ 0.6122 r.l.u.(reciprocal lattice units) of GBCO, corresponding to an out-of-plane parameter c/2 ¼ 3.77(1) A, in good agreement with thevalue obtained from the q/2q scans. The elongated shape of the

7410 | J. Mater. Chem. A, 2013, 1, 7408–7414

peak in the horizontal direction denes two distinct maxima(labeled A and B in the map), which were attributed to theincipient a/b differentiation of the orthorhombic phase. Theircorresponding in-plane parameter values are 3.89(1) A for A and3.92(1) A for B. These values are slightly larger and shorter thanthe corresponding a ¼ 3.862 A and b/2 ¼ 3.934 A, respectively,reported for the bulk phase,6 and are therefore assigned to thesecrystallographic directions indicating a certain degree of biaxialin-plane strain, tensile along the a-axis and compressive alongthe b-axis.

Standard q/2q XRD patterns of the lms deposited on NGO(110), Fig. 1b, do not show clearly differentiated lm peaks, onlya weak shoulder at the larger angle side of the 110 substratepeak (equivalent to 001 reection for the primitive cubic cell)was observed.

Reciprocal space maps around the 3�36 NGO reection(equivalent to the �303 reection for the primitive cubic cell),shown in Fig. 1d, evidence the presence of two clear contribu-tions, labeled C and D, at common Qy position. This corre-sponds to an out-of-plane parameter of 3.86(1) A, very close tothe a-axis parameter of the bulk phase. The coincidence of thisvalue with that of the NGO substrate explains why the GBCOpeaks are not observed in the standard patterns since theyperfectly overlap with the NGO peaks. The corresponding in-plane parameter values for C and D are 3.815 A and 3.925 A,respectively, these values being slightly longer (and shorter)than those corresponding to c/2 (and b/2) of the bulk GBCOphase. This indicates pure a-axis orientation perpendicular tothe lm, and the coexisting c- and b-axis parallel to the main in-plane [100]/[010] crystallographic orientations of the substrateprimitive cell. Fig. 1e and f show the schemes of the relativeorientations of lm and STO and NGO substrates, respectively.As in the case of c-axis oriented lms on STO, pure a-axisoriented lms on NGO show a certain degree of biaxial strain.Table 1 collects the calculated mismatch between lm andsubstrate cells for the different orientations as well as themeasured cell parameters on both the STO and NGO substrates.Mismatch calculations involving orthorhombic structures maybecome confusing if one compares the linear differencesbetween cell parameters for each crystallographic orientation(two cell parameters for each pseudocubic h100i orientation).Instead, a simpler approach consists of comparing the areamismatch between the corresponding rectangular sides of eachprimitive cell with the plane cut of the substrate. As an example,the mismatch for the GBCO a-axis oriented domains on NGO iscalculated by comparing the areas of the rectangular plane cutof the substrate: a0 � b0 ¼ 3.864 � 3.855 ¼ 14.896 A2, and therectangular plane formed by GBCO b/2 � c/2 ¼ 3.934 � 3.786 ¼14.894 A2. Their difference (Arealm � Areasubs)/Areasubs is�0.02% (slightly compressive), as depicted in Table 1. This alsohelps to realize that even though the direct comparison of theseparate cell parameters may induce a uniaxial strain of oppo-site sign along the two in-plane directions (a0 < c/2, b0 > b/2) thearea mismatch is very small. Therefore it is not surprising thatthe corresponding GBCO out-of-plane value is not very differentto the bulk value (less than +0.1% difference). By comparing thedifferent lm-substrate orientations the lower mismatch area

This journal is ª The Royal Society of Chemistry 2013

Table 1 Cell parameters of GBCO, STO and NGO structures along with calculated mismatch for the different orientations. The table also contains the values of the filmcell parameters measured from reciprocal space maps

Substrate cellparameters (A)

Plane cutparameters (A)

GBCO areamismatcha (%)

Measured out-of-planeparameter (A)

Measuredin-plane-parameters (A)

GBCOorientation

STO Pm-3m (001) ab: +0.37 3.775 c-axisCubic a0 ¼ b0 ¼ 3.905 bc: +2.3 3.8965 (A)a ¼ 3.905 ac: +4.1 3.925 (B)NGO Pbnm (110) ab: �1.99 3.867 a-axisOrthorhombic a0 ¼ 3.864 bc: �0.02 3.815 (C)a ¼ 5.43 b0 ¼ 3.855 ac: +1.85 3.925 (D)b ¼ 5.50c ¼ 7.71

a Calculated from the difference between the area of the corresponding plane of the primitive GBCO cell (orthorhombic Pmmm, a ¼ 3.862 A; b/2 ¼3.934 A; c/2 ¼ 3.786 A) and the corresponding substrate.

Paper Journal of Materials Chemistry A

Publ

ishe

d on

18

Apr

il 20

13. D

ownl

oade

d by

UN

IVE

RSI

TA

T G

IESS

EN

on

28/1

0/20

14 1

6:14

:06.

View Article Online

was clearly obtained for c-axis orientation on STO and a-axisorientation on NGO, in good agreement with the experimentalobservation.

Occasionally, in some of the lms deposited on NGOsubstrates, a certain amount of c-axis oriented domains (veryweak peaks) were observed depending on the depositionconditions. This indicates that growth of GBCO on thissubstrate is much more sensitive to subtle changes in thedeposition parameters.

Despite some subtle differences in their strain, we mayconclude that depending on the substrate type the lms arepurely c-axis and predominantly a-axis oriented on STO andNGO substrates, respectively. Since in the GBCO Pmmm struc-ture the fast oxygen migration path would follow the crystallo-graphic a-axis, it would be expected that depending on the lmorientation the lms may behave in a very distinct way. In lmswith c-axis orientation their corresponding crystallographica-axis lays parallel to both the main in-plane directions, while ina-axis oriented lms it is perpendicular to the lm.

The average cationic composition measured by WDSmicroprobe analysis of the lms deposited on STO (001) fromthe Co-enriched (5%) target corresponded to 0.96 : 1.02 : 2.02.This corrected the observed Co-decient lms observed in lmsdeposited by PLD from stoichiometric 1 : 1 : 2 (([Gd] + [Ba])/[Co] ¼ 1) targets which induced the formation of a large densityof planar defects.16

Fig. 2 Normalised 18O isotope fraction along the transverse direction (depthprofile) of GBCO films of about 160 nm thick deposited on STO (a) and NGO (b) atexchange temperatures of 328, 384 and 463 �C, and times of 544, 883 and 630 s,respectively. The scheme in (c) depicts the experimental arrangement for the 18Oexchange directly through the top film surface.

Oxygen surface exchange in transverse geometry

The oxygen transport along the transverse direction (perpen-dicular to the lm plane) for the GBCO thin lms was deter-mined from the depth proles obtained with the IEDPtechnique in combination with ToF-SIMS analysis. Fig. 2a and bshow the corresponding normalized 18O isotopic oxygen frac-tion for the different exchange temperatures, on STO and NGOsubstrates, respectively. Fig. 2c shows a scheme of themeasuring geometry.

The calculation of the oxygen transport parameters is ach-ieved by tting the experimental data to the appropriate solu-tion of the diffusion equation. We assume for this congurationthe model of an innite solid slab extending over the region ofthickness 2l in which the initial isotope fraction is Cbg and the

This journal is ª The Royal Society of Chemistry 2013 J. Mater. Chem. A, 2013, 1, 7408–7414 | 7411

Journal of Materials Chemistry A Paper

Publ

ishe

d on

18

Apr

il 20

13. D

ownl

oade

d by

UN

IVE

RSI

TA

T G

IESS

EN

on

28/1

0/20

14 1

6:14

:06.

View Article Online

diffusion annealing takes place in a large volume of gas ofconstant isotope fraction Cg. The symmetrical solutions givenby this plane sheet model (x being the distance from the lm–

substrate interface, and l being the lm thickness) occupyingthe region �l < x < l, apply also to the sheet 0 < x < l when theface x ¼ 0 is non-permeable. The solution to the diffusionequation given by Crank19 is as follows:

crðx; tÞ ¼ Cðx; tÞ � Cbg

Cg � Cbg

¼ 1�XN

n¼1

2Lcosðbnx=lÞexp�� bn

2D*t=l2�

cosbn

�bn

2 þ L2 þ L�

(1)

where C(x, t) is the 18O isotope fraction obtained by SIMS, i.e.18O/(18O + 16O), t is the time of the isotope exchange, D* is theoxygen tracer diffusion coefficient, k* is the surface tracerexchange coefficient and bn (n ¼ 1, 2,.) are the nonnegativeroots of b tan b ¼ L, and L ¼ l$k*/D*. The limitations of thetechnique are examined in detail elsewhere.8,20–22

The proles are almost at across the lm thickness down tothe interface with the substrate. This indicates that oxygendiffusion is too fast to be measurable for the current lmthickness under the exchange conditions (T and t), so the rate-limiting step in this geometry is the oxygen surface exchange.No information about diffusion was extracted for this geometryand thus only tracer exchange rates k* were obtained. Fig. 3depicts an Arrhenius plot of the corresponding k* values,obtained for the best t of the experimental data to the diffusionequation solution, eqn (1), along with those values reported inthe literature at similar T. For the low temperature region, theoxygen surface exchange rates for both lm orientations areabout half an order of magnitude larger than those reported forthe bulk material, with similar activation energies, while thebulk and thin lm values are comparable at the high tempera-tures. The larger surface activities at low temperatures could bedue to the textured lms the surface of which predominantlyexpose particular crystallographic planes, (001) or (100) for purec- or a-axes oriented lms, respectively. This indicates greaterpotential for thin lms with controlled surface orientation ascathodes for SOFC application at intermediate operation

Fig. 3 The oxygen surface exchange rates (k*) plotted as a function of reciprocaltemperature for the GBCO films with c- and a-axes orientation (deposited on STOand NGO, respectively). The graph also includes values for the bulk GBCO ceramicfrom the literature.3 The dashed lines are only a guide to the eye.

7412 | J. Mater. Chem. A, 2013, 1, 7408–7414

temperatures. The comparison between the lms indicates thatregardless their texture, lms with c- or a-axes orientation showcomparable surface activity for oxygen exchange. Either thereare little differences in the surface chemical termination, or thepossible differences in the surface termination plane (presum-ably of pure AO or A0O planes for pure c-axis oriented lms ormixed AO/A0O in the case of pure a-axis oriented lms) do notplay an important role in oxygen exchange. However, the slightreduction observed at higher temperatures might be due to asurface atomic rearrangement induced by the long annealingtimes in a pure O2 atmosphere needed for the annealing stepprior to the 18O exchange, which might cause surface segrega-tion, as observed in Sr-substituted LaCoO3

23 as well as inLa0.8Sr0.2MnO3 perovskite.24

Oxygen surface exchange and diffusion in longitudinalgeometry

Fig. 4a and b show longitudinal concentration proles obtainedfor the lms deposited on STO and NGO substrates at threedifferent temperatures. For tting the longitudinal concentra-tion proles the semi-innite medium solution of the diffusionequation was used.19 For this particular geometry, the analysisprovides both the oxygen exchange, k*, and diffusion coeffi-cients, D*, along the plane of the lm, which correspond todifferent crystallographic orientations according to the lmtexture, i.e. along the a/b direction for c-axis oriented lms onSTO, and along the b/c direction for a-axis oriented lms onNGO. The solution to the diffusion equation can be written interms of dimensionless parameters and as x0 ¼ x=

ffiffiffiffiffiffiffiffiffiffi4D*t

pand

h0 ¼ ðk*=D*Þ= ffiffiffiffiffiffiffiffiffiffi4D*t

pas

crðx; tÞ ¼ Cðx; tÞ � Cbg

Cg � Cbg

¼ erfc�x0�� exp

�2h0x0 � h02

�erfc

�x0 þ h0

(2)

D* and k* values obtained for the measured regions of theexchanged samples are depicted in the Arrhenius plots in Fig. 5a and b. Despite the inherent variability in the measurement ofk*, the values for this longitudinal geometry follow a similartrend to the one observed in the transverse geometry. Thesurface exchange coefficients for lms on STO and NGO(Fig. 5a) show values from half to one order of magnitude largerthan the bulk ones, and are slightly larger than those reportedfor transverse geometry. This could be related to a higheractivity of the freshly cut trenches opened in this experiment incomparison with the top lm surface more exposed to surfacecontamination and degradation in the case of transversegeometry. Besides, this longitudinal conguration may be lessaffected by surface segregation during the low temperatureannealing because of the particular geometry and the surface-to-volume ratio and, therefore, remain more active for oxygenreduction. Again, there are no signicant differences related tolm orientation, as it would be expected in this case since theexposed surfaces very likely consist of a combination ofdifferent crystallographic planes, i.e.: (100)/(010) for the c-axisoriented lms on STO, and (001)/(010) for the a-axis orientedlms on NGO.

This journal is ª The Royal Society of Chemistry 2013

Fig. 4 Normalised isotopic fraction profile along the longitudinal direction ofthe c-axis (a) and a-axis (b) GBCO oriented films deposited on STO and NGOsubstrates, respectively, and exchanged at 338, 391 and 473 �C, during 17 655,628 and 4941 s, respectively. The scheme in (c) shows the experimentalarrangement for this longitudinal geometry.

Fig. 5 Arrhenius plots of the tracer oxygen surface exchange rate k* (a) andoxygen tracer diffusivities D* (b) measured in the longitudinal geometry for thedifferent GBCO film orientations. The graphs include also the values reported forthe bulk material for comparison.3 Dashed lines are only a guide to the eye.

Paper Journal of Materials Chemistry A

Publ

ishe

d on

18

Apr

il 20

13. D

ownl

oade

d by

UN

IVE

RSI

TA

T G

IESS

EN

on

28/1

0/20

14 1

6:14

:06.

View Article Online

The most remarkable differences are observed in the D*tracer diffusion coefficient, in Fig. 5b. While the D* valuesobtained on c-axis oriented GBCO/STO are slightly higher thanthe reported bulk material in the full range of temperatures,those obtained on a-axis oriented GBCO/NGO deviate substan-tially, being almost one order of magnitude smaller in the lowtemperature range between 300 and 400 �C. For pure c-axisoriented lms the diffusion path in the longitudinal geometry isparallel to the a/b crystallographic planes, and therefore themeasured D* values are taken as the maximum intrinsic values.On the other hand, in a pure a-axis oriented lm there would bea/b planes perpendicular to the diffusion path, but also paralleldue to the mixed b and c orientation within the lm plane.Depending on the grain connectivity, there is little chance for a

This journal is ª The Royal Society of Chemistry 2013

direct diffusion path along a–b planes, as in the case of c-axisoriented lms. Therefore, the observed reduction of the tracerdiffusion coefficient D* is evidence of the diffusion anisotropy,as has also been proven in polycrystalline samples of PrBa-Co2O5.5. For this material the diffusion coefficient was found tobe larger in the [100]/[010] direction compared to [001].15 At thehighest temperature the a-axis oriented lm shows a compa-rable D* value to the c-axis oriented and bulk values. It cannotbe ruled out that the presence of a small amount of c-axisorientation in predominantly a-axis oriented samples, as couldbe the case for this particular sample, might provide a fastdiffusion path and therefore a substantial increase in thediffusion coefficient. If that was the case the anisotropy inGBCO may be even of a larger magnitude and therefore thelargest anisotropy value measured (one order of magnitude)constitutes the lower bound of the measurements taking intoaccount the possible fast diffusion paths through grainboundaries or regions with different orientation. The fact thatD* values along the fastest diffusion direction are similar tothose reported for polycrystalline GBCO material (with arandom distribution of domain orientations) indicates that thebulk material presents a substantial connectivity between a/b-planes so the fastest diffusion path dominates, which reinforcesthe suitability of these ordered perovskites for MIEC cathodes.

J. Mater. Chem. A, 2013, 1, 7408–7414 | 7413

Journal of Materials Chemistry A Paper

Publ

ishe

d on

18

Apr

il 20

13. D

ownl

oade

d by

UN

IVE

RSI

TA

T G

IESS

EN

on

28/1

0/20

14 1

6:14

:06.

View Article Online

Conclusions

We have deposited high quality epitaxial lms of the mixedionic-electronic conducting material GdBaCo2O5.5+d ondifferent substrates, and demonstrated that the appropriatechoice of substrate mismatch allows growing lms with eitherpure c-axis or a-axis orientation. This has allowed exploration ofthe potential anisotropy in the oxygen transport and has proventhat indeed the oxygen diffusion at low temperatures is almostone order of magnitude larger along the a-axis compared to thec-axis. This has been related to the arrangement of oxygenvacancies preferentially in the GdO planes forming channelsalong the a-axis, and therefore providing a path for oxygenmigration. However, no inuence of the structure anisotropywas observed in the oxygen surface exchange rates, which wereof similar values regardless of the lm orientation or measuringgeometry. The method described here provides a very powerfultool to explore intrinsic oxygen transport properties in complexanisotropic oxidematerials otherwise dominated by themixtureof all orientations and grain boundary contributions in poly-crystalline samples.

Acknowledgements

The authors would like to acknowledge the Spanish Ministry ofEducation and Culture for nancial support (MAT2011-29081-C02-01 and CONSOLIDER-INGENIO CSD2008-0023 projects).This research was also supported by a Marie Curie Intra Euro-pean Fellowship within the seventh European CommunityFramework Programme (PIEF-GA-2009-252711) (for M.B). J. Z.thanks the Spanish National Research Council (CSIC) for theJAE-PreDoc Scholarship. The authors would like to thank DrS. Fearn and R. Charter for their help with ToF-SIMS analysis.

Notes and references

1 A. Tarancon, M. Burriel, J. Santiso, S. J. Skinner andJ. A. Kilner, J. Mater. Chem., 2010, 20, 3799–3813.

2 J. Santiso and M. Burriel, J. Solid State Electrochem., 2011, 15,985–1006.

3 A. Tarancon, S. J. Skinner, R. J. Chater, F. Hernandez-Ramırez and J. A. Kilner, J. Mater. Chem., 2007, 17, 3175–3181.

7414 | J. Mater. Chem. A, 2013, 1, 7408–7414

4 A. Tarancon, J. Pena-Martinez, D. Marrero-Lopez, A. Morata,J. C. Ruiz-Morales and P. Nunez, Solid State Ionics, 2008, 179,2372–2378.

5 A. Maignan, C. Martin, D. Pelloquin, N. Nguyen andB. Raveau, J. Solid State Electrochem., 1999, 142, 247–260.

6 C. Frontera, J. L. Garcia-Munoz, A. Llobet andM. A. G. Aranda, Phys. Rev. B: Condens. Matter Mater. Phys.,2002, 65, 180405(R).

7 D. Partt, A. Chroneos, A. Tarancon and J. A. Kilner, J. Mater.Chem., 2011, 21, 2183–2186.

8 R. A. De Souza and M. Martin, MRS Bull., 2009, 34, 907.9 J. A. Kilner, S. J. Skinner and H. Brongersmaa, J. Solid StateElectrochem., 2011, 15, 861.

10 M. Burriel, G. Garcia, J. Santiso, J. A. Kilner, J. C. C. Richardand S. J. Skinner, J. Mater. Chem., 2008, 18, 416–422.

11 L. Wang, R. Merkle, J. Maier, T. Acarturk and U. Starke, Appl.Phys. Lett., 2009, 94, 071908.

12 L. Yan and P. A. Salvador, ACS Appl. Mater. Interfaces, 2012, 4,2541.

13 M. Kubicek, Z. Cai, W. Ma, B. Yildiz, H. Hutter and J. Fleig,ACS Nano, 2013, 7, 3276–3286.

14 K. Kerman, C. Ko and S. Ramanathan, Phys. Chem. Chem.Phys., 2012, 14, 11953.

15 M. Burriel, J. Pena-Martinez, R. J. Chater, S. Fearn,A. V. Berenov, S. J. Skinner and J. A. Kilner, Chem. Mater.,2012, 24, 613–621.

16 M. Burriel, M. Casas-Cabanas, J. Zapata, H. Tan, J. Verbeeck,C. Solıs, J. Roqueta, S. J. Skinner, J. A. Kilner, G. VanTendeloo and J. Santiso, Chem. Mater., 2010, 22, 5512–5520.

17 J. A. Kilner, B. C. H. Steele and L. Ilkov, Solid State Ionics,1984, 12, 89.

18 D. R. Killoran, J. Electrochem. Soc., 1962, 190, 170–171.19 J. Crank, The Mathematics of Diffusion, Oxford University

Press, Oxford, 1975.20 R. A. De Souza and R. J. Chater, Solid State Ionics, 2005, 176,

1915.21 E. Fischer and J. L. Hertz, Solid State Ionics, 2012, 218, 18–24.22 F. Ciucci, Solid State Ionics, 2013, 232, 97–105.23 Z. Cai, M. Kubicek, J. Fleig and B. Yildiz, Chem. Mater., 2012,

24, 1116–1127.24 S. Fearn, J. C. H. Rossiny, J. A. Kilner and J. R. G. Evans, Solid

State Ionics, 2012, 211, 51–57.25 K. Momma and F. Izumi, J. Appl. Crystallogr., 2011, 44, 1272–

1276.

This journal is ª The Royal Society of Chemistry 2013