You are on page 1of 10

Solid State Ionics 111 (1998) 125–134

Mechanism and kinetics of oxygen reduction on porous


La 12x Sr x CoO 32d electrodes

S.B. Adler*
Ceramatec Inc., 2425 S. 900 W., Salt Lake City, UT 84119, USA
Received 7 November 1997; accepted 3 June 1998

Abstract

The linear polarization behavior of La 12x Sr x CoO 32d electrodes has been measured as a function of x, T, and PO 2 using
a.c. impedance. These measurements indicate a reaction mechanism in which O 2 is reduced chemically at the porous mixed
conductor surface, followed by diffusion of oxygen through the mixed conductor to the electrolyte. Analysis of the electrode
kinetics using a recent model for this mechanism yields values for the oxygen vacancy diffusion coefficient (Dv ) and
surface-exchange rate constant (r 0 ) as a function of x, T, and PO 2 . These values agree well with published independent
measurements of oxygen chemical diffusion and isotope surface exchange (D˜ and k). The distance (d ) that the reaction
extends beyond the electrode / electrolyte interface depends on both Dv and r 0 , varying between 0.3 and 10 microns for a
typical electrode surface area.  1998 Elsevier Science B.V. All rights reserved.

Keywords: Perovskite; Mixed conductor; Electrode kinetics; Ion conductor; Chemical diffusion; Surface exchange; Transition metal oxide

1. Introduction conductor may play a significant (if not dominant)


role in determining the electrode polarization: (i)
Much debate surrounds the mechanism of the large faradaic capacitances (up to 1 farad / cm 2 ) [4,5],
oxygen–reduction reaction at a porous mixed-con- (ii) semi-infinite diffusion behavior at intermediate
ducting oxygen electrode. In analogy to a porous frequencies [6], (iii) linear i-V characteristics at
metal catalyst (e.g. Pt on zirconia), many workers moderate overpotential [7], (iv) strong correlation
apply a traditional paradigm in which O 2 reduction is between electrode kinetics and bulk diffusion and
limited by a charge-transfer process occurring locally surface-exchange properties [8,9]. Even a poor
at the one-dimensional ‘three-phase boundary’ inter- mixed conductor such as La 12x Srx MnO 32d (LSM)
face between electrode, electrolyte, and gas [1–3]. shows evidence for involvement of bulk-transport
However, much evidence suggests that oxygen ab- [10,11].
sorption and transport in the bulk of the mixed In order to incorporate these effects, some workers
have applied modified charge-transfer or equivalent
*Corresponding author. Address: Dept. of Chemical Engineer- circuit models that assume non-charge-transfer pro-
ing, Case Western Reserve University, 10900 Euclid Ave., cesses occurring in the electrode can be treated as
Cleevelan, Ohio 44106, U.S. Tel.: 11 216 368 4182; e-mail: series or parallel combinations of charge-transfer
stu@ceramatec.com elements [12]. However, this assumption has been

0167-2738 / 98 / $ – see front matter  1998 Elsevier Science B.V. All rights reserved.
PII: S0167-2738( 98 )00179-9
126 S.B. Adler / Solid State Ionics 111 (1998) 125 – 134

shown by continuum theory to be generally incorrect


for treating oxygen absorption and solid-state / gase-
ous diffusion, which contribute to the impedance in a
convoluted manner [13]. The equivalent circuit ap-
proach also leaves the electrode reaction mechanism
only vaguely or empirically defined, with poor
connection between modeling parameters and the
properties of the materials. Although such a model
may provide a useful set of parameters to ‘fit’ data
accurately, it provides little mechanistic insight.
In order to establish a more direct connection
between performance and properties of mixed-con-
ductors, several workers have used continuum
models to treat explicitly the non-charge-transfer
processes occurring in a membrane [14] or electrode
[13,15]. An impedance model which uses this ap-
proach [13,16,17] has shown preliminary evidence
that the kinetics for mixed conductors with fast ionic
transport (such as La 12x Sr x Co 12y Fe y O 32d ) is domi-
nated by surface chemical exchange of O 2 and solid-
state oxygen diffusion. As shown schematically in
Fig. 1, this model treats the overall electrode re-
action, ]21 O 2 1 2e 2 → O 22 , as a homogeneous chemi-
cal reaction occurring over the internal surface area
of the electrode material. The absorption of neutral
oxygen by the mixed-conductor serves to convert
electronic to ionic current over a finite region of the
electrode thickness. This active region is described Fig. 1. (a) Schematic of the electrode reaction model. (b)
by the characteristic length, d, related to the ex- Schematic of the steady-state electronic and ionic currents in the
change and diffusion properties of the mixed conduc- mixed conductor.
tor. Estimates of d based on SIMS measurements of
D* and k yield values of several microns. This tions quantitatively, and link electrode kinetics di-
model also predicts that gas phase diffusion becomes rectly to material properties. This link not only
dominant below 1% oxygen in N 2 , even at currents allows us to examine the implications of this mecha-
well below limiting. nism, but also to examine and gain insight from
In this paper, we apply this model to understand- where the model breaks down.
ing mixed-conducting La 1-x Sr x CoO 3-d (LSCO) elec-
trodes on ceria. LSCO is an excellent material for
fundamental studies because it is reversibly reduc- 2. Theory
ible, has high rates of surface exchange and diffu-
sion, and forms a clean well-characterized electrode As shown in Fig. 2a, the model discussed above
system with little or no interfacial charge-transfer predicts that the zero-bias impedance of a symmetri-
resistances. LSCO has been well studied in the cal cell (electrode / electrolyte / electrode) can be ex-
literature, and independent measurements of its pressed as a sum of charge-transfer resistances and
thermodynamic, kinetic, and transport properties impedances, plus a ‘chemical’ impedance associated
have been made over a wide range of temperature with non-charge-transfer processes:
and PO 2 conditions. Thus application of the model to
LSCO / ceria allows us evaluate the models assump- Z 5 R electrolyte 1 Zinterfaces 1 Zchem , (1)
S.B. Adler / Solid State Ionics 111 (1998) 125 – 134 127

]]]]]]]
S Dœ
RT
R chem 5 ]]2
2F
]]]]]]]
t
(1 2 ´)c v Dv ar 0 (af 1 ab )
(3)

c v (1 2 ´)
t chem 5 ]]]] (4)
Aar 0 (af 1 ab )
where c v is the vacancy concentration, A 5 2 1 /
(2n) 5 2 ]21 ≠ln(PO 2 ) / ≠ln(c v ) is the thermodynamic
factor, Dv is the vacancy diffusion coefficient, ´, a,
and t are the porosity, surface area, and solid-phase
tortuosity, respectively, of the porous mixed conduc-
tor, r 0 is the exchange neutral flux density (in
analogy to exchange-current density), and af and ab
are kinetic parameters of order unity that depend on
the specific mechanism of the surface chemical
exchange reaction.
The model also predicts that at steady state, the
mixed conductor will be reduced by an amount that
decays exponentially with distance from the elec-
trolyte interface, with a characteristic length describ-
ing the size of the active region [13]:
]]]]
c v Dv (1 2 ´)

Fig. 2. (a) Schematic of the total impedance as predicted by Eq.


œ
d 5 ]]]]
ar 0 (af 1 ab )t
(5)

Finally, the model also predicts that at low O 2


(1). (b) Plot of the ‘chemical’ impedance in the limit of Eq. (2).
concentration in the gas, diffusion in the pores of the
mixed conductor may become important. The
where R electrolyte is the electrolyte resistance,
characteristic resistance of this diffusion process,
Zinterfaces is the impedance of the electron-transfer
which contributes to Zchem , is given by [13]:
and ion-transfer processes occuring at the current
collector / electrode and electrode / electrolyte inter- 2LRT(1 2 xxp )tp
faces, respectively, and Zchem is the convoluted Zp 5 ]]]]] (6)
16F 2 cx´xp DAB
contribution of non-charge-transfer processes includ-
ing oxygen surface exchange, solid-state diffusion, where L is the thickness of the electrode, x is the
and gas-phase diffusion inside and outside the elec- oxygen mole fraction, c is the total gas phase
trode [13]. concentration (P/RT ), DAB is the binary diffusion
As shown in Fig. 2b, in the limit of a semi-infinite coefficient between O 2 and the diluent gas, xp 5DK /
(thick) electrode with no gas-phase diffusion limita- (DK 1DAB ) is a correction factor for Knudsen diffu-
tions, the non-charge-transfer term reduces to: sion, DK is the Knudsen diffusivity of O 2 in the
pores of the solid, and tp is the gas-phase tortuosity.
]]]]
1
œ
Zchem 5 R chem ]]]]
1 1 jv t chem
(2)
3. Experimental

where R chem and t chem are a characteristic resistance LSCO powders of composition x50.2, 0.3, and
and time constant, respectively, related to the 0.4 were manufactured at Ceramatec, and verified as
thermodynamic, surface kinetic, and transport prop- phase-pure and impurity-free by XRD, XRF, and
erties of the mixed conductor [13]: SEM / EDS. The powders were made into an ink, and
128 S.B. Adler / Solid State Ionics 111 (1998) 125 – 134

used to fabricate symmetrical LSCO / ceria / LSCO


cells with 2.0 cm 2 active area. The electrodes were
fired in air onto the doped ceria electrolyte (of
proprietary composition) between 900 and 12008C
for between 1 and 8 h. Following firing, the surface
areas of the electrode materials were measured by
BET, and the porosity determined from the electrode
weight and thickness (SEM). The solid-phase tor-
tuosity was estimated from SEM, based on quali-
tative comparison to pore geometries typical in the
catalysis literature [18]. Table 1 shows a summary of
these electrode geometry measurements.
As explained in the Section 5 below, minor
impurities and certain electrolyte dopants can react
with the mixed-conductor, altering its diffusion and
surface exchange properties, or producing insulating
phases at the electrode / electrolyte interface. These
phenomena can cause a dependence of the electrode
kinetics on electrolyte dopant, and non-zero interfa-
cial impedances (Zinterfaces ±0). Much effort was
taken in this study to choose electrolyte materials
that minimize these effects, and to carefully examine
the electrolyte surface and electrode / electrolyte in-
terface by SEM to insure no obvious second-phase
formation had occurred.
Measurements of a.c. and d.c. polarization were
made under controlled atmosphere and temperature Fig. 3. (a) Schematic of the test apparatus used to measure the cell
using the experimental apparatus shown schematical- impedance. (b) Deconvolution of ohmic and non-ohmic contribu-
ly in Fig. 3a. The electrochemical measurements tions to the cell impedance.
were made using a Solartron 1287 potentiostat and
1260 frequency response analyzer, and the oxygen impedance (Z) was typically ohmic or purely induc-
concentration was verified using an in-situ zirconia tive, with a real value R ohmic . Measurements were
oxygen sensor. Measurements were made between made on electrolytes of several different thicknesses
550 and 8008C, and between 1% and 100% O 2 in in order to confirm that R ohmic vs. thickness fits a
blended gases of O 2 / He and O 2 / N 2 . In most cases straight line with zero intercept, and has a slope
three cells were studied and the results averaged. equal to the electrolyte resistivity. The remaining
The electrolyte and electrode contributions to the impedance, Z2R ohmic , typically consisted of a single
cell impedance were deconvoluted as shown in Fig. resonance with a shape similar to Fig. 2b, and was
3b. In the limit of high frequency (.5 kHz), the total taken to be the impedance of the electrode reactions
(Zchem ). Although previous work with this family of
Table 1 perovskites has shown a second, high frequency
Geometric parameters of the electrode as a function of com- interfacial impedance [6,19] (especially at low tem-
position measured or estimated from BET, SEM, and weight
perature), none were observed under the conditions
measurements
investigated here. Where appropriate, Zchem was fit to
x L (mm) a (mm 21 ) ´ t Eq. (2) using non-linear least-squares in order to
0.2 15 1.8 0.5 2.5 extract values for R chem and t chem .
0.3 15 2.3 0.5 2.5 At low gas-phase concentrations of oxygen, the
0.4 15 1.6 0.5 2.5
electrodes become limited by gas-phase diffusion,
S.B. Adler / Solid State Ionics 111 (1998) 125 – 134 129

Fig. 4. Impedance of a symmetrical LSCO / ceria / LSCO cell


(x50.3 and T57508C) as a function of PO 2 with He and N 2 as the
balance gas.

severely limiting the range of PO 2 we were able to


study. For example, Fig. 4 shows the impedance of
an LSCO / ceria / LSCO cell as a function of PO 2 in
He vs. N 2 . At 21% O 2 , results in He and N 2 are
nearly identical, but at 4%, a repeatable difference of
0.20 V cm 2 appears. A calculation of the gas-phase
diffusion resistance [Eq. (6)] based on the electrode
morphology and known published values for binary
Fig. 5. (a) Impedance of a symmetrical LSCO / ceria / LSCO cell at
and Knudsen diffusivities [18] yields values con-
x50.4, T57508C, P hO h2jj50.21 atm. (b, c) Values of R chem and
sistent with this difference: 0.04 and 0.20 V cm 2 , ] ]
t chem as a function of x, T, and PO 2 , extracted from the impedance
respectively, for 4% O 2 in He and N 2 . For the values data by fitting to Eq. (2). Circles, squares, and triangles indicate
of R chem studied here, this effect can be ignored if x50.2, 0.3, and 0.4, respectively.
helium is used. However, below |1% O 2 (especially
in N 2 or Ar), the impedance becomes dominated by
the gas phase, showing a semicircular response with function of x and T in air, and as a function of x and
a resistance proportional to PO 2 . Similar effects were PO 2 at 7508C. The results show that R chem depends
reported in Ref. [6]. more strongly on temperature than t chem (|140 vs.
|100 kJ / mol, respectively), while t chem depends
21 20.3
more strongly than R chem on PO 2 (P O 2 vs. P O 2 ).
Both R chem and t chem depend strongly on x. As shown
4. Results below, these dependencies can be explained quantita-
tively in terms of the properties of LSCO.
4.1. Electrode kinetics

Fig. 5a shows the impedance response of a 4.2. Extraction of Dv and r0


LSCO / ceria / LSCO cell (x50.4) in air at 7508C.
Above |5 kHz the response becomes purely induc- Using Eqs. (3) and (4), the values of Dv and
tive, with a constant real component, R ohmic (sub- r 0 (af 1 ab ) were calculated as a function of T, PO 2 ,
tracted from the data). The response fits well to Eq. and x based on the values of R chem and t chem reported
(2), yielding values for R chem and t chem . The behavior in Fig. 5. No adjustable parameters were used in this
was similar at all values of x, T, and PO 2 , and Fig. 5b calculation. The electrode geometry parameters ´, t,
and c show a summary of R chem and t chem as a and a were taken from Table 1, while the thermo-
130 S.B. Adler / Solid State Ionics 111 (1998) 125 – 134

dynamic properties of LSCO were taken from the


literature [20,21] and incorporated using the Lan-
khorst rigid-band formalism [22], which succinctly
expresses c v and the thermodynamic factor A in
terms of T, PO 2 , and x:

4
S
m O0 2 (T ) 1 RT lnPO 2 5 Eox (x) 2 ]] 6] 2 x
cv
g(´F ) c mc D
S
2 TSox (x) 2 2RT ln ]]]
cv
c mc 2 c v D
(7)

12 cv
A 5 1 1 ]]] ] (8)
g(´F )RT c mc

where c mc is the concentration of lattice sites in


LSCO (0.091 mol / cm 3 ), Eox and Sox are a standard
Helmholtz energy and entropy, and g(´F ) is the
density of states at the Fermi level, all assumed
independent of T and PO 2 [21]. The results are
shown in Figs. 6 and 7.
Fig. 6a shows the calculated values of Dv as a
function of temperature and x in air. For comparison,
values of Dv 5 D˜ /A calculated from permeation data
[21,23] are also shown. The agreement at high
temperature is good, showing a similar activation
energy and dependence on Sr content. There is some Fig. 6. Values of Dv and r 0 (af 1 ab ) extracted from the electrode
deviation from Arrhenius behavior at x50.2 and low kinetics as a function of x and T (in air). Circles, squares, and
temperature, corresponding to a less quantitative fit triangles indicate x50.2, 0.3, and 0.4, respectively. Closed
of the impedance data to Eq. (2). As discussed symbols indicate impedance data. Open symbols indicate values
based on independent chemical diffusion and isotopic exchange
below, this deviation suggests that the model may
measurements.
begin to break down at low vacancy concentration.
Fig. 6b shows the calculated values of r 0 (af 1 ab )
as a function of temperature and x in air. For content, also supported by chemical diffusion and
comparison, the values of r 0 (af 1 ab )5kc mc calcu- permeation measurements [21,23].
lated from surface exchange measurements at x50.2 Fig. 7b shows the calculated values of r 0 (af 1 ab )
are shown (open symbols) [24]. For this comparison, as a function of PO 2 and x at 7508C. In contrast to
the values of k (measured at PO 2 50.7 atm) were Dv , the surface exchange constant depends strongly
corrected slightly (to 0.21 atm) based on the PO 2 on oxygen concentration, varying as P 11 O 2 , and there
dependence shown in Fig. 7b, below. Agreement is is little dependence on Sr content. This strong
reasonable in both absolute value and activation dependence on PO 2 , and commensurate weak depen-
energy. dence on x (electron occupancy of the Co3d–O2p
Fig. 7a shows the calculated values of Dv as a band) suggests the exchange reaction is limited by
function of PO 2 and x at 7508C. This data shows a O 2 dissociation rather than availability of electronic
weak dependence of Dv on PO 2 , in good agreement or ionic species. The nonlinear behavior at x50.2
with chemical diffusion measurements (coulometric may indicate that other steps limit the exchange
titration) [21]. There is a strong dependence on Sr reaction at lower vacancy concentration.
S.B. Adler / Solid State Ionics 111 (1998) 125 – 134 131

material properties is extremely valuable because it


allows us to examine quantitatively some of the
implications of this mechanism, and to examine and
gain insight from the limitations and deficiencies of
the model.

5.2. Size of the extension region

One question we can address is how far the


reaction extends from the electrolyte / electrode inter-
face. This distance is characterized by D, calculated
from Eq. (5) using the values of Dv and r 0 (af 1 ab )
reported above. Fig. 8 show the results of this
calculation as a function of x, T and PO 2 . These
results show that at under realistic operating con-

Fig. 7. Values of Dv and r 0 (af 1 ab ) extracted from the electrode


kinetics as a function of x and PO 2 (at 7508C). Symbols are as in
Fig. 6.

5. Discussion

5.1. Reaction mechanism

The results presented above suggest that O 2


reduction on a porous La x Sr 12x CoO 32d electrode is
limited primarily by surface chemical exchange and
solid state diffusion. This conclusion stands in
contrast to the traditional view that the electrode
reaction is limited by a charge-transfer process,
presumably occurring at the three phase boundary.
Indeed, this analysis shows that the electrode kinetics
can be explained quantitatively, with no adjustable
parameters, in terms of independently measured
Fig. 8. Values of the extension region, D, as a function of x, T, and
transport and chemical-kinetic properties of the PO 2 , calculated based on the values of Dv and r 0 (af 1 ab ) in Figs.
mixed conductor. 6 and 7. Circles, squares, and triangles indicate x50.2, 0.3, and
This link between the electrode kinetics and the 0.4, respectively.
132 S.B. Adler / Solid State Ionics 111 (1998) 125 – 134

ditions, the reaction is not confined to the three- A breakdown of the model for small d is consistent
phase boundary, but may extend a distance ap- with its formal assumption that the active region is
proaching the electrode thickness (15 mm). As one much larger than the average particle size d p 53 /A
might expect, d has a strong dependence on x and [13]. However, it is unclear if this breakdown is
PO 2 since the active region grows with increasing caused by the two-dimensionality of the electrode
vacancy concentration or lower r 0 . There is a much geometry under these conditions, or the presence of a
weaker or inverted dependence on T since Dv and second, parallel reaction mechanism. For example,
r 0 (af 1 ab ) have similar activation energies, and thus surface diffusion might become important at low
tend to cancel each other in Eq. (5). vacancy concentration, explaining the apparent over-
In doing this analysis, we found evidence that Eq. prediction of Dv for x50.2 at low temperature (Fig.
(2) begins to break down at exceptionally small and 6a).
large values of d. This effect is illustrated in Fig. 9a, At large values of d (.6 mm), the impedance also
which shows the shape of the impedance as a appears slightly distorted from Eq. (2), which we
function of d, normalized to its value at d 52 mm. expect to occur as d approaches L / 3 (|5 mm). This
When d is moderate (2 mm), the impedance has a effect is more clearly illustrated in Fig. 9b, which
shape closely matching Eq. (2). However at small compares the response of a low surface area elec-
d (,0.7 mm), the shape of the impedance becomes trode (a,0.1 mm 21 ), made by firing the electrode to
less well-defined, more like a depressed semicircle. higher temperature, normalized to the response of an
electrode of high surface area (a52.0 mm 21 ). The
response of the low-surface area electrode is nearly
semicircular, with only slight evidence of diffusive
behavior at high frequency.
This distortion can be understood by re-examining
the reaction / diffusion limit of the model, but relax-
ing the assumption of semi-infinite electrode thick-
ness. With a finite electrode thickness L, the non-
charge-transfer term in equation 12 of Ref. [13]
becomes:
]]]]

œ
1 L ]]]
S
Zchem 5 R chem ]]]]uTanh ]œ1 1 jv t chem .
1 1 jv t chem d
D
(9)

As L approaches infinity, the Tanh(L /d ) term ap-


proaches unity, yielding Eq. (2). As illustrated in
Fig. 1b, the relevant length parameter in this limit
becomes d, which governs the rates of both solid-
state diffusion and surface exchange. On the other
hand, as d becomes larger than L, the Tanh(L /d )
term in Eq. (9) becomes linear, yielding:

RT 1
Zchem 5 ]]2 ]]]]]]]] (10)
2F aLr 0 (af 1 ab )(1 1 jv t chem )
Fig. 9. (a) Shape of the impedance as a function of D for a high
surface area electrode (x50.3, T57508C, varying PO 2 ). (b) which is a semicircle with a characteristic resistance
Comparison of the impedance at high and low surface area
(x50.4, T57508C, PO 2 50.21 atm). Schematics of the corre- related only to surface exchange (1 /aLr 0 ). The
sponding steady-state reduction profiles in the electrode are relevant length parameter in this limit becomes L,
illustrated. corresponding to the case where solid-state diffusion
S.B. Adler / Solid State Ionics 111 (1998) 125 – 134 133

is fast, and the entire internal surface area of the properties. We have observed strong effects of cer-
electrode (aL) is utilized for O 2 reduction. tain types of ceria dopants on the surface-exchange
A fit of the low-surface-area data in Fig. 9b to Eq. properties of LSCO. This effect is seldom homoge-
(10) (assuming the same material properties as for neous, and thus alters the shape as well as the
the high surface area electrode) yields a value of magnitude of the impedance. Data from Liu and Wu
a50.05 mm 21 , which is consistent with estimates [3] show large effects on the electrode impedance
based on BET data (0.01|0.1 mm 21 ). The calculated with Ba and Bi as electrolyte dopants. These ele-
value of D based on this area is 30 mm (twice the ments are likely to be highly reactive with LSCO.
electrode thickness), indicating that the entire elec- We have also observed that certain impurities intro-
trode is indeed utilized. duced during bonding may influence the diffusion
properties of LSCO. Indeed, Bae and Steele (and
5.3. Electrolyte effects others) have seen strong effects on electrode prop-
erties of Si contamination from the electrolyte or
The electrode model embodied in Eq. (2) predicts from processing [26,27]. Finally, for materials hav-
that the electrode kinetics are independent of the ing small d, doping of the electrolyte by elements in
electrolyte material. However, ongoing observations the electrode may create mixed-conduction in the
at Ceramatec and elsewhere show various dependen- electrolyte, altering the effective active area [28].
cies on electrolyte composition and cell processing.
These effects have been interpreted by some authors
as ‘proof’ that the mechanism of the reaction must be 6. Conclusions
limited by a charge-transfer process at the three-
phase boundary [3]. We believe this interpretation is The measurements presented in this paper suggest
oversimplified, and offer two alternative explana- that O 2 reduction on a porous La x Sr 12x CoO 32d
tions, neither of which require invoking a localized electrode is limited primarily by surface chemical
three-phase boundary mechanism. exchange and solid state diffusion. This conclusion
First, we have observed that during firing, stands in contrast to the traditional view that the
ionically-insulating phases can form at the interface electrode reaction is limited by a charge-transfer
between electrode and electrolyte, limiting oxide-ion process, presumably occurring at the three phase
transfer and creating a secondary high frequency boundary. Indeed, this analysis shows that the elec-
charge-transfer impedance (Zinterface ). It has been trode kinetics can be explained solely in terms of the
known in the literature for many years that perov- transport and chemical-kinetic properties of the
skites in the family (La,Sr)(Mn, Co, Fe)O 32d can mixed conductor, with all charge-transfer steps as-
react with yttria-stabilized zirconia (YSZ) to form sumed to be equilibrated.
insulating La-Zr-O and Sr-Zr-O phases that degrade For typical electrode geometries, the utilization
electrode performance. Recent data from Liu and Wu region (d ) over which the reduction reaction occurs
[3], appear to show this effect for extends several microns from the electrode / elec-
La 12x Sr x Co 12y Fe y O 32d (LSCF) electrodes on YSZ, trolyte interface, approaching the electrode thickness
as well as for LSCF on Sr-doped ceria. We have also under some conditions. For electrodes of low surface
observed this effect with LSCO on ceria when the area (,0.1 mm 21 ), the reaction extends over the
ceria contains certain types of dopants (details pro- entire surface of the electrode, and is limited almost
prietary). Small amounts of second phase that are not entirely by surface chemical exchange. These dimen-
detectable by SEM or XRD are sufficient to cause sions show that the material and geometry of the
this effect. Interfacial resistance can also be caused electrode far from the three-phase boundary are
by impurities at the interface, and is pronounced at important in determining overall performance.
low temperature [25]. Below O 2 concentrations of |0.01 atm, the elec-
Second, reaction and interdiffusion of electrode trode kinetics become dominated by gas-phase /
and electrolyte occur during bonding, and this may Knudsen diffusion in the pores of the electrode. This
have an impact on both electrolyte and electrode observation can be explained quantitatively based on
134 S.B. Adler / Solid State Ionics 111 (1998) 125 – 134

the geometry of the electrode and the properties of [6] J.A. Lane, S. Adler, P.H. Middleton, B.C.H. Steele, in: M.
Dokiya, O. Yamamoto, H. Tagawa, S.C. Singhal (Eds.),
the gas. For further details, including the theoretical
Proceedings of the Fourth International Symposium on Solid
effects of combined solid-state and gas-phase diffu- Oxide Fuel Cells, The Electrochemical Soc., New Jersey,
sion, and diffusion in a boundary layer outside the 1995.
electrode, the reader is referred elsewhere [6,13]. [7] Y. Takeda, R. Kanno, M. Noda, Y. Tomida, O. Yamamoto, J.
Finally, we point out observations by ourselves Electrochem. Soc. 134(11) (1987) 2656.
[8] R.H.E. van Doorn, PhD Thesis, Twente University, 1996.
(and others) of a dependence of the electrode kinetics
[9] J.E. ten Elshof, H.J.M. Bouwmeester, H. Verweij, Solid State
on the type of electrolyte and conditions of process- Ionics 81 (1995) 97.
ing. We do not consider these effects to be evidence [10] E. Siebert, A. Hammouche, M. Kleitz, Electrochem. Acta
for a charge-transfer-limited reaction mechanism. 40(11) (1995) 1741.
Rather, we believe these effects are caused by [11] H. Lauret, A. Hammou, J. European Ceram. Soc. 16(4)
(1996) 447.
reactions between the electrode and electrolyte ma-
[12] M. Liu, J. Electrochem. Soc. 145 (1998) 142.
terials, forming ionically insulating phases or [13] S.B. Adler, J.A. Lane, B.C.H. Steele, J. Electrochem. Soc.
modifying the properties of the materials. Although 143(11) (1996) 3554.
parallel, more localized or charge-transfer-limited [14] H. Deng, M. Zhou, B. Abeles, Solid State Ionics 74 (1994)
mechanisms may be present (as is likely with LSM / 75.
[15] A.M. Svensson, S. Sunde, K. Nisanciuglu, J. Electrochem.
ceria [13]), the data and analysis presented here
Soc. 145 (1998) 1390.
suggest they are not rate-limiting for mixed-conduc- [16] S.B. Adler, J.A. Lane, B.C.H. Steele, J. Electrochem. Soc.
tors with high rates of solid-state oxygen transport. 144(5) (1997) 1884.
[17] S. Adler, J.A. Lane, B.C.H. Steele, in: H.U. Anderson, A.C.
Khandkar, M. Liu (Eds.), Proceedings of the Second Interna-
tional Symposium on Ceramic Membranes, PV 94-12, p.
Acknowledgements
159, The Electrochemical Soc., New Jersey, 1995.
[18] C.N. Satterfield, T.K. Sherwood, The Role of Diffusion in
This material is based upon work supported by the Catalysis, Addison-Wesley Pub. Co., 1963.
North Atlantic Treaty Organization under a Grant [19] H.U. Anderson, L.-W. Tai, C.C. Chen, M.M. NAsrallah, W.
awarded in 1993 (NSF-NATO). Samples and ex- Huebner, in: M. Dokiya, O. Yamamoto, H. Tagawa, S.C.
Singhal, (Eds.), Solid Oxide Fuel Cells IV, PV 95-1, p. 375,
perimental facilities were provided by Ceramatec,
The Electrochemical Society Proceedings Series, Pennington,
Special thanks go to Marc Flinders and Dave Padgen NJ, 1995.
for aiding in materials characterization and electro- [20] J. Mizusaki, Y. Mima, S. Yamauchi, K. Fueki, J. Solid State
chemical measurements, and Jon Lane and John Chem. 80 (1989) 102.
Newman for helpful discussions. Thanks also go to [21] M. Lankhorst, Ph.D. Thesis, University of Twente, En-
schede, The Netherlands, 1997.
the University of Twente for copies of the Ph.D.
[22] M.H.R. Lankhorst, H.J.M. Bouwmeester, H. Verweij, Phys.
theses of Martijn Lankhorst and Rene´ van Doorn. Rev. Lett. 77(14) (1996) 2989.
[23] R.H.E. van Doorn, Ph.D. Thesis, University of Twente,
Enschede, The Netherlands, 1996.
References [24] H.J.M. Bouwmeester, H. Druidhof, A.J. Burgraaf, Solid
State Ionics 72 (1994) 185.
[25] J.A. Lane, private communication, 1997.
[1] F.H. van Heuveln, H.J.M. Bouwmeester, J. Electrochem. [26] J. Bae, B.C.H. Steele, Solid State Ionics 106 (1998) 247.
Soc. 144(1) (1997) 134. [27] B.C.H. Steele, private communication, 1998.
¨
[2] M. Godickemeier, K. Sasaki, L.J. Gauckler, I. Reiss, J. [28] J.D. Sirman, J.A. Kilner, J. Electrochem. Soc. 143 (1996)
Electrochem. Soc. 144(5) (1997) 1635. L229.
[3] M. Liu, Z. Wu, Solid State Ionics 107 (1998) 105.
[4] M. Kleitz, F. Petitbon, Solid State Ionics 92 (1996) 65.
[5] M. Kleitz, L. Dessemond, T. Kloidt, M.C. Steil, in: M.
Dokiya, O. Yamamoto, H. Tagawa, S.C. Singhal (Eds.),
Proceedings of the Fourth International Symposium on Solid
Oxide Fuel Cells, The Electrochemical Soc., New Jersey,
1995.

You might also like