Sie sind auf Seite 1von 7

Solid State Ionics 167 (2004) 99 105

www.elsevier.com/locate/ssi

Hydrogen permeability of SrCe1xMxO3d (x=0.05, M=Eu, Sm)


S.-J. Song a, E.D. Wachsman a,*, J. Rhodes a, S.E. Dorris b, U. Balachandran b
a

Department of Materials Science and Engineering, University of Florida, Gainesville, FL 32611-6400, USA
b
Argonne National Laboratory, Energy Technology Division, Argonne, IL 60439, USA
Received 20 May 2003; received in revised form 8 December 2003; accepted 9 December 2003

Abstract
The hydrogen permeability of SrCe0.95Eu0.05O3  d and SrCe0.95Sm0.05O3  d was studied as a function of temperature, hydrogen partial
pressure ( PH2) gradient, and water vapor partial pressure ( PH2O) gradient. Under a 100% dry hydrogen condition at 1123 K, the hydrogen
permeation rates of dense membranes (1.72 mm thick) are c 3.19  10 9 mol/cm2 s for SrCe0.95Eu0.05O3  d and 2.33  10 9 mol/cm2 s for
SrCe0.95Sm0.05O3  d. Under wet hydrogen conditions at 1123 K, the hydrogen permeation rates are c 2.89  10 9 and 1.21  10 9 mol/cm2
s, respectively, for the same materials. The dopant dependence of hydrogen permeability is explained in terms of the ionization potential of the
dopant. Electronic conductivity was calculated from hydrogen permeation fluxes; activation energies for electron conduction under both dry
and wet conditions were also calculated. The PH2O dependence of electronic conductivity and hydrogen permeability is discussed.
D 2004 Elsevier B.V. All rights reserved.
Keywords: Hydrogen permeation; Electronic conductivity; SrCe0.95Eu0.05O3d; SrCe0.95Sm0.05O3d

1. Introduction
In recent years, the development of dense ceramic
membranes with mixed protonic and electronic conductivity
has received considerable attention due to their possible
application in hydrogen-based energy, petrochemical processes, fuel cells, separation membranes, and other technologies [1 4]. In particular, proton transport properties in
multivalent cation-substituted barium cerate and strontium
cerate (BaCe1  xMxO3  d, SrCe1  xMxO3  d, M = Yb, Tm,
Sm, Eu) have been widely studied in connection with hightemperature hydrogen separation [5 9]. Among these membranes, SrCe1  xEuxO3  d is a promising candidate material
because of the high proton selectivity of strontium cerate
and the electronic conductivity provided by the multivalent
europium dopant [6]. For these reasons, our previous work
aimed not only to elucidate the effect of Eu on the hydrogen
permeation but also to theoretically model the behavior of
SrCe1  xEuxO3  d [10 13].
The overall hydrogen permeation process consists of
three consecutive kinetic steps: gas/solid interfacial reaction,
solid-state diffusion, and solid/gas interfacial reaction. The
relative control of the overall kinetics is often determined by
* Corresponding author. Tel.: +1-352-846-2991; fax: +1-352-8460326.
E-mail address: ewach@mse.ufl.edu (E.D. Wachsman).
0167-2738/$ - see front matter D 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.ssi.2003.12.010

the characteristic length of a specimen, defined as LC = D*/


k, where k is the surface exchange coefficient, and D* is the
tracer diffusion coefficient. The characteristic length determines the transition from bulk diffusion-limited to surface
exchange rate-limited transport [14]. Hamakawa et al. [8]
reported that H2 permeation rates at 950 K were controlled
by bulk diffusion through dense SrCe0.95Yb0.05O3  d membranes even for 2-Am films. Therefore, permeation through
1.72-mm-thick SrCe1  xMxO3  d membranes (the thickness
we used in experiments reported in this paper) should be
bulk diffusion-controlled.
Once a hydrogen chemical potential gradient (the thermodynamic driving force for hydrogen permeation) is applied,
hydrogen will permeate because the diffusion of protons and
electrons is ambipolar. The motion of electrons, the minority
carrier, gives rise to the hydrogen permeation by chargecompensated transport of protons in the same direction [15].
Generally, the hydrogen permeation flux across an oxide
membrane can be calculated from the Wagner equation,
which assumes that bulk diffusion is the rate-limiting step:
"
Z POW
Z PHW
2
2
1 RT
RT
JOHO 
rt tOHO tVO dlnPO2 2
rt tOHO
L 4F 2 POV2
2F PHV2
#
 tVO teVdlnPH2 ;

100

S.-J. Song et al. / Solid State Ionics 167 (2004) 99105

where rt is the total conductivity; ti is the transference number


S SS
of the charged species (i = OHO, VO , eV), F is the Faraday
constant, and d ln PO2 and d ln PH2 are the chemical potential
gradients across an oxide membrane.
In this work, we studied the hydrogen permeability of
SrCe1  xMxO3  d (x = 0.05, M = Eu, Sm) as a function of
temperature, PO2, PH2, and PH2O. Partial conductivities at a
given pressure were also investigated.

gas flow was 50 sccm of H2/Ar mixtures. The hydrogen


content of the permeate stream was measured with Q100MS
Dycor Quadlink mass spectrometer. Leakage of neutral gas
through pores in the sample or through an incomplete seal
was checked by measuring the argon tracer content of the
permeate stream. No discernible leak was detected. Comprehensive experimental details will be discussed in a future
publication [16].

2. Experimental

3. Results and discussion

Polycrystalline SrCe 0.95 Eu 0.05 O 3  d and SrCe 0.95


Sm0.05O3  d samples were prepared by conventional solidstate reaction methods. High-purity oxide powders of SrCO3
(99.9%; Alfa Aesar) and CeO2 (99.9%; Alfa Aesar) were
mixed with Eu 2 O 3 (99.99%; Alfa Aesar) or Sm 2 O 3
(99.99%; Alfa Aesar), ground in a ball mill with stabilized
zirconia balls in ethanol, and calcined at 1573 K for 10 h in
air. The calcined oxides were then crushed, sieved to < 45
Am, ground in a ball mill again, pressed into pellets, cold
isostatic-pressed, and sintered at 1773 K for 10 h in air. The
densities of the resultant disks were 96% of theoretical. Xray diffraction spectra confirmed that all specimens obtained
in this way exhibit a single phase of the orthorhombic
perovskite structure.
Hydrogen permeation were measured on dense disks 24
mm in diameter  1.72 mm thick. The planar surfaces of
each disk were polished with 1200-grit SiC paper. Polished
disks were sealed with ceramic sealant to alumina tubes. In
order to minimize the edge contribution to the permeation,
the membrane edge was sealed with ceramic sealant. The
sweep side flow was a constant 20 sccm of helium; the feed

3.1. Hydrogen permeation fluxes


From the hydrogen content measured in the helium
sweep side of the permeation assembly and the helium flow
rate, the total hydrogen permeation rate (mol/s) was calculated, while assuming the ideal gas law. Then, the permeation fluxes (mol/cm2 s) were calculated by dividing the
permeation rates by the effective surface area of the disk
membranes.
The variation in hydrogen flux with temperature for Euand Sm-doped SrCeO3 is shown in Fig. 1. The hydrogen
fluxes increase with temperature for both systems, and the
SrCe0.95Eu0.05O3  d exhibits higher permeability when compared with SrCe0.95Sm0.05O3  d over the entire investigated
temperature range under both dry and wet conditions.
The influence of applied hydrogen chemical potential
gradient on the hydrogen permeability of SrCe 0.95
Eu 0.05 O 3  d at various values of P H 2 O and SrCe 0.95
Sm0.05O3  d in dry conditions is shown in Fig. 2. The
hydrogen permeation flux of the Eu-doped specimen shows
higher hydrogen permeability and a greater DPH2 depen-

Fig. 1. Hydrogen flux as a function of temperature for Eu- and Sm-doped SrCeO3 d: closed symbols = 100% H2; dry and open symbols: PH2 = 0.972 atm,
PH2O = 0.028 atm.

S.-J. Song et al. / Solid State Ionics 167 (2004) 99105

101

Fig. 2. Hydrogen flux as a function of applied hydrogen chemical potential gradients under various conditions of PH2O at 850 jC.

dence than did the hydrogen permeation flux of the Smdoped specimen.
The dependence of hydrogen permeability on the
dopant can be explained by an electronic conduction
mechanism in reducing atmospheres and/or by a proton
transport mechanism. Kreuer et al. [17] suggest that an
observed increase in activation energy for the diffusivity
of protonic defects with dopants may be related to the
general increase in the oxygen basicity and to the proton
transfer barrier. Muon spin relaxation measurements [18]
on Sr-doped SrZrO3 also suggest the existence of trapping
centers adjacent to the dopant ions. Furthermore, several
simulations have been performed [19 21] to predict the
effect of dopant ions on proton conduction. However, the
effect of dopant ions on hydrogen permeation is most
likely due to the effect of the dopant on electronic
conductivity.
Our previous work on electronic conduction [12]
describes it as a small polaron model because the electron
transport in dopants was understood in terms of a charge
transfer reaction, wherein an electron is transferred from a
dopant ion in a low-oxidation state to a neighboring ion in a
high-oxidation state. In addition, our earlier work [11]
showed that the ionized dopant concentration depends on
the ionization potential of each dopant under given thermodynamic conditions. The third ionization potential of Eu is
24.8 eV, which is larger than that of Sm (23.3 eV) [22].
Therefore, the reaction Eu2 + ! Eu3 + requires more energy
than Sm2 + ! Sm3 +, and Eu2 + is more thermodynamically
stable than Sm2 +. This greater thermodynamic stability of
Eu2 + means that there is a greater concentration of EuCe
U in
SrCe1  xEuxO3  d than SmCe
U in SrCe1  xSmxO3  d. Therefore, the n-type electronic conduction due to Eu Ce
U !
EuCe
V + eVis greater than the n-type conduction in Sm-doped
SrCeO3  d.

Fig. 3 shows the hydrogen flux of Eu-doped SrCeO3  d


as a function of temperature under various PH2and PH2O
conditions. For both dry and wet circumstances, hydrogen
permeability increases with increasing PH2 because of the
contribution of proton and electron concentrations. As PH2O
increases, hydrogen permeability decreases because of an
increase in PO2, which leads to a decrease in the concentration of electrons.
Hydrogen fluxes through 1.72-mm-thick membranes are
not high enough for practical systems, but the absolute
values of hydrogen permeation can be improved by more
effective experimental design. Hamakawa et al. [8] reported
that hydrogen permeation rates are inversely proportional to
the thickness of dense membranes down to micrometer
ranges. Their results showed that hydrogen permeation rates
were more than 100 times greater through 2-Am SrCe0.95
Yb0.05O3  d films than through 1-mm-dense membranes of
an identical composition. The permeation flux can be further
increased by improving the surface exchange kinetics by
coating the surface of the membrane with a porous layer,
thus increasing the effective surface area, or by coating the
membrane surface with materials that exhibit superior
hydrogen exchange properties. Also, by optimizing dopant
concentrations in a specimen, hydrogen permeation can be
enhanced, as shown in previous work on Eu-doped
BaCeO3  d [6]. Moreover, recent investigations, to be
reported in a future publication, confirm that hydrogen
permeation can be enhanced by manipulating the microstructure.
3.2. Partial conductivities
The ambipolar conductivities of 1.72-mm-thick SrCe0.95
Eu0.05O3  d at various PH2, PH2O, and temperatures are
shown in Fig. 4. When both PH2 and PH2O differences are

102
S.-J. Song et al. / Solid State Ionics 167 (2004) 99105

Fig. 3. Hydrogen flux of SrCe0.95Eu0.05O3d as a function of temperature: (a) dry; (b) PH2O = 0.028 atm; (c) PH2O = 0.051 atm; and (d) PH2O = 0.086 atm.

S.-J. Song et al. / Solid State Ionics 167 (2004) 99105

103

Fig. 4. Ambipolar conductivity of SrCe0.95Eu0.05O3  d as a function of temperature: (a) dry; and (b) wet (solid symbols: PH2O = 0.028 atm; open symbols:
PH2O = 0.051 atm; open symbols with bar, PH2O = 0.086 atm).

applied to an oxide membrane, the thermodynamic driving


force for hydrogen permeation is not only the PH2 gradient
but also the PO2 gradient, as shown in Eq. (1). Under our
experimental conditions, no discernible oxygen permeation
was detected by the mass spectrometer so tVO  can be
assumed to be zero. Therefore, Eq. (1) can be simplified to:

JOHO 

RT
2F 2 L

PHW

PHV2

rt tOHO teVdln PH2 :

Using Eq. (2), we can extract the ambipolar conductivity


ramb from the hydrogen permeation flux as:

ramb



rOHO reV
4F 2 L BJH2

:
RT BPHW2 Pref
rOHO reV

H2

The ambipolar conductivity can be calculated from the


slope of the hydrogen permeation flux versus the logarithm of the hydrogen partial pressure. In this calcula-

104

S.-J. Song et al. / Solid State Ionics 167 (2004) 99105

Fig. 5. Electronic conductivity of SrCe0.95Eu0.05O3  d as a function of temperature.

tion, the mean slope at a given PHU 2 was determined by


using the datum at the PHU 2 and interpolation between its
nearest neighbors. As shown in Fig. 4, the ambipolar
conductivity varies with both PH2 and PO2. The hydrogen
exponent m, ramb~PHm2 , takes a value of c 1/2, as shown
in Fig. 4a. This result was expected from our previous
work [12] on defect structure and n-type electrical properties of SrCe0.95Eu0.05O3  d. The 1/2 slope corresponds to
the dependence of electron concentration on hydrogen
partial pressure in region IV (Fig. 3 in Ref. 12), charac-

terized by the charge neutrality condition 2[EuCe


U ]=[OHO]
( p < KA):
1=2

n~AVPH2
and by
KH2 O

PH2 O
1=2

PO2 PH2

1=4 1=2

n~APO2 PH2 O ;

where A and AV are constants that depend on temperature.

Fig. 6. Electronic conductivity of SrCe0.95Eu0.05O3  d as a function of PH2O at 850 jC.

S.-J. Song et al. / Solid State Ionics 167 (2004) 99105

The oxygen exponent m, ramb~POm2, is c  1/4, as


shown in Fig. 4b. The PO2 dependence of ambipolar conductivity within the investigated PO2 range agrees with the PO2
dependence of electrons [12]. To satisfy the PH2 and PO2
dependence of ambipolar conductivity, the proton conductivity should still be larger than the electron conductivity so
the ambipolar conductivity can be considered equivalent to
the electron conductivity. If rOHSO >reV, then ramb c reV.
Therefore, it is fair to say that, in the wet reducing atmosphere
explored in this experiment, SrCe0.95Eu0.05O3  d is in the
proton-dominant region, where protons are charge compensated by ionized europium ions.
If we assume that ramb c reV, the electronic conductivity
as a function of temperature is as shown in Fig. 5. From the
temperature dependence of the electronic conductivity, the
activation energies calculated from the dry and wet condition are c 0.63 and 0.72 eV, respectively. The addition of
water vapor to hydrogen reduces the electronic conductivity
and increases the activation energy of electron conduction in
the investigated temperature range.
Fig. 6 illustrates the electron conductivity dependence of
PH2O at various PH2 values. As expected from Fig. 5,
electronic conductivity slightly decreases with PH2O. Because the oxygen chemical potential increases with PH2O, an
increase in proton concentration with PH2O is followed by a
decrease in electrons due to the dominance of the redox
reaction in the oxide membrane:
1
H2 g O2 g X H2 Og
2



H2 Og VO OxO X 2OHO
OxO

1
X V
O 2e O2 g:
2

terms of the ionization potential of the dopant. The


ambipolar conductivity calculated from hydrogen permeation fluxes shows the same PO2 and PH2 dependence as
the electronic conductivity for the experimental conditions.
Electronic conductivity and hydrogen permeation flux
decrease with PH2O. Further effort is required to enhance
the electron conductivity of Eu-doped strontium cerate by
adjusting the doping level, increasing the effective surface
area, and/or improving hydrogen exchange properties at the
surface.

Acknowledgements
This work was supported by the US Department of
Energy, Office of Fossil Energy, National Energy Technology Laboratorys Gasification Technologies Program, under
contract W-31-109-Eng-38 and NASA NAG3-2750.

References
[1]
[2]
[3]
[4]
[5]
[6]

6
7
8

Consequently, the hydrogen permeability activated by


the PH2 and PH2O differences across the SrCe0.95Eu0.05O3  d
membrane is controlled by the transport of electrons.

105

[7]
[8]
[9]
[10]
[11]
[12]
[13]

4. Conclusions

[14]

Hydrogen permeability of SrCe0.95Eu 0.05O 3  d and


SrCe0.95Sm0.05O3  d membranes was studied by measuring
gas permeation as a function of temperature and PH2, and
PH2O gradient. The effect of the dopant ion on hydrogen
permeability through a 1.72-mm-thick membrane was investigated. The SrCe 0.95Eu0.05O3  d membrane shows
higher hydrogen permeability when compared with
SrCe0.95Sm0.05O3  d over the entire investigated range of
temperature and PH2 and PH2O gradients. The dopant
dependence of hydrogen permeability was explained in

[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]

H. Iwahara, Solid State Ionics 77 (1995) 289.


H. Iwahara, Solid State Ionics 125 (1999) 271.
K.D. Kreuer, Solid State Ionics 97 (1997) 1.
F.D. Schutter, J. Vangrunderbeek, J. Luyten, I. Kosacki, R.V.
Landschoot, J. Schran, J. Schoonman, Solid State Ionics 57 (1992) 77.
N. Bonanos, Solid State Ionics 53 56 (1992) 967.
J. Rhodes, E.D. Wachsman, in: E.D. Wachsman, W. Weppner, E.
Traversa, M. Liu, P. Vanysek, N. Yamazoe (Eds.), Solid State Ionic
Devices II- Ceramic Sensors, vols. 2000 2032, The Electrochem.
Soc, Pennington, NJ, 2001, p. 137.
H. Iwahara, T. Yajima, H. Uchida, Solid State Ionics 70 71 (1994)
267.
S. Hamakawa, L. Anwu, E. Iglesia, Solid State Ionics 148 (2002) 71.
X. Qi, Y.S. Lim, Solid State Ionics 130 (2000) 149.
S.-J. Song, E.D. Wachsman, S.E. Dorris, U. Balachandran, Solid
State Ionics 149 (2002) 1.
S.-J. Song, E.D. Wachsman, S.E. Dorris, U. Balachandran, J. Electrochem. Soc. 150 (2003) A790.
S.-J. Song, E.D. Wachsman, S.E. Dorris, U. Balachandran, J. Electrochem. Soc. 150 (2003) A1484.
S.-J. Song, E.D. Wachsman, J. Rhodes, S.E. Dorris, U. Balachandran,
Solid State Ionics 164 (2003) 107.
J. Crank, The Mathematics of Diffusion, 2nd ed., Claren Press, Oxford, UK, 1975, p. 326.
T. Norby, Y. Laring, Solid State Ionics 136 137 (2000) 139.
J. Rhodes, E.D. Wachsman, S.-J. Song, in preparation.
K.D. Kreuer, W. Munch, M. Ise, T. He, A. Fuchs, U. Traub, J. Maier,
Ber. Bunsenges. Phys. Chem. 101 (1997) 1334.
R. Hempelmann, M. Soetratmo, O. Hartmann, R. Wapping, Solid
State Ionics 107 (1998) 267.
M.S. Islam, J. Mater. Chem. 10 (2000) 1027.
M.S. Islam, M. Cherry, Solid State Ionics 97 (1997) 33.
W. Munch, G. Seifert, K.D. Kreuer, J. Maier, Solid State Ionics 86 88
(1996) 647.
J.A. Dean (Ed.), Langs Handbook of Chemistry, McGraw-Hill, New
York, 1991, pp. 3 8.

Das könnte Ihnen auch gefallen