You are on page 1of 13

Solid State Ionics 177 (2006) 1529 1541 www.elsevier.

com/locate/ssi

Oxide anode materials for solid oxide fuel cells


Jeffrey W. Fergus
Auburn University, Materials Research and Education Center, 275 Wilmore Laboratories, Auburn, AL 36849, United States Received 30 November 2005; received in revised form 28 February 2006; accepted 5 July 2006

Abstract A major advantage of solid oxide fuel cells (SOFCs) over polymer electrolyte membrane (PEM) fuel cells is their tolerance for the type and purity of fuel. This fuel flexibility is due in large part to the high operating temperature of SOFCs, but also relies on the selection and development of appropriate materials particularly for the anode where the fuel reaction occurs. This paper reviews the oxide materials being investigated as alternatives to the most commonly used nickelYSZ cermet anodes for SOFCs. The majority of these oxides form the perovskite structure, which provides good flexibility in doping for control of the transport properties. However, oxides that form other crystal structures, such as the cubic fluorite structure, have also shown promise for use as SOFC anodes. In this paper, oxides are compared primarily in terms of their transport properties, but other properties relative to SOFC anode performance are also discussed. 2006 Elsevier B.V. All rights reserved.
Keywords: Solid oxide fuel cells; Anode; Perovskite; Fluorite

1. Introduction Solid oxide fuel cells (SOFCs) are a promising clean power source for a variety of stationary and mobile applications [14]. One of the advantages of SOFCs over polymer electrolyte membrane (PEM) based fuel cells is their fuel flexibility, which includes tolerance to impurities in the fuel as well as operation with fuels other than hydrogen [26]. As a result of this fuel flexibility, SOFCs are being developed for automotive auxiliary power units (APUs) [7] which can operate using reformed diesel fuel, and thus can be implemented before establishment of a hydrogen infrastructure. The primary reason for the improved fuel flexibility of SOFCs is their high operating temperature, which accelerates the electrode reaction rates. However, this high operating temperature also increases the rates of undesired reactions and creates thermal stresses during cycling, so a major objective in SOFC development is to decrease the operating temperature. As the operating temperature decreases, reduced electrode reaction rates can lead to decreased conversion efficiencies of certain

Tel.: +1 334 844 3405; fax: +1 334 844 3400. E-mail address: jwfergus@eng.auburn.edu. 0167-2738/$ - see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.ssi.2006.07.012

fuels and/or increased inhibition by impurities, both of which can limit fuel flexibility. The anode material must be compatible (chemical and thermal expansion) with the electrolyte and provide means for transport of gas from the fuel, oxygen ions from the electrolyte and electrons to the interconnect. The catalytic properties of the anode to the fuel oxidation reaction are also important, particularly as the operating temperature is decreased. The most widely used anode for SOFCs is a cermet of nickel and yttria-stabilized zirconia (YSZ) [811]. Ni-YSZ cermets have been used successfully in SOFCs, but there are some challenges remaining. One is the deposition of carbon on the electrode when used in carbon-containing fuels [10,12,13]. While a small amount of carbon can be beneficial [14], larger amounts block the anode reaction. One approach to overcoming this limitation is to replace nickel with copper, which does not catalyze CC bond formation, so carbon deposition is reduced [8,11,12]. However, this approach still requires fabrication of a two-phase cermet. Ni-YSZ anodes can also be poisoned by sulfur [12], and the deposition of sulfides can increase with reduced operating temperature [15]. In addition, Ni-YSZ anodes can undergo microstructural changes during oxidation and reduction, which can reduce the three-phase (electrolytenickelgas) area, and thus reduce the electrode activity.

1530

J.W. Fergus / Solid State Ionics 177 (2006) 15291541

Fig. 1. Conductivity of yttrium/lanthanide-doped SrTiO3 in fuel atmospheres (pO2 = 10 1510 14 Pa) at 800 C [1720].

These limitations of Ni-YSZ anodes have led to the development of single-phase oxide anode materials [1113]. A single-phase anode material that covers the entire electrolyte surface must be a mixed (oxygenelectron) conductor, so that oxygen ions can be transported to and electrons transported from the electrode reaction surface. This eliminates the need for a three-phase contact, which simplifies fabrication and reduces problems with sintering of the electrode during use. The target conductivity for such an anode material is 100 Scm 1, but this requirement can be reduced to as low as 1 Scm 1 through careful design of the electrode morphology [11]. Oxide anodes must be mixed conductors and thus typically contain transition metals to provide the electronic conductivity. Nickel and cobalt oxides are not stable in reducing conditions, so more stable oxides, such as those containing titanium, chromium, manganese or iron are typically used for anode materials. For anodes, conduction in reducing conditions is relevant, so n-type

conducting oxides are typically used. However, the conductivity of a p-type oxide will increase when the oxygen electrochemical potential increases at the anode under polarization, so p-type oxides are attractive if the initial conductivity in the fuel atmosphere is sufficiently high [11]. 2. Perovskite-related structures Many of the promising oxides used as SOFC anodes form perovskite-related structures [16]. The anode requires stability in reducing conditions, so the most commonly used perovskites are titanates and chromites [12]. 2.1. Titanates Strontium titanate (SrTiO3) is a good electronic conductor at low oxygen partial pressures, so for application as an SOFC

Fig. 2. Conductivity of niobium-doped SrTiO3 niobium/yttrium-doped SrTiO3 in fuel atmospheres (pO2 = 10 14 Pa) at 800930 C [1719,2426].

J.W. Fergus / Solid State Ionics 177 (2006) 15291541

1531

Fig. 3. Conductivity of (La,Sr)TiO3+ in fuel atmospheres at 800 C [19,25,3034]. The maximum oxygen content for Ti4+, La3+ and Sr2+ are given in the legend.

anode it is doped to increase the ionic conductivity. Fig. 1 shows the increase in the conductivity of SrTiO3 when strontium is replaced with various lanthanides or with yttrium [1720]. Each system has a peak in conductivity with increasing dopant concentration. The highest conductivity among the oxides shown in Fig. 1 is 64 Scm 1 for Sr0.88Y0.08TiO3 [19] and can be increased further to 82 Scm 1 with increases in the deficiency on the strontium site to form Sr0.86Y0.08TiO3 [19,21]. The conductivities reported by different researchers should be compared with caution, because variations with processing parameters, which can affect properties such as porosity, and sluggish equilibration of point defects can lead to variations in measured conductivities [22]. Frozen-in defects from previous heat treatments can dominate electrical behavior [23], particularly at low oxygen partial pressures, since the reduction reaction is considerably slower than the oxidation reaction [21]. Fig. 2 [1719,2426] shows that doping with niobium on the

titanium site can also increase the conductivity. The lower conductivity for Sr(Ti,Nb)O3 that was oxidized prior to measurement of the conductivity in reducing conditions [24] is evidence of the sluggish reduction reaction. The conductivity of yttrium-doped SrTiO3 can be further increased by simultaneous doping with niobium [26]. Yttrium-doped SrTiO3 has also been used in composites with YSZ [27] or ceria [28] electrolytes, where the titanate, rather than a metal, provides the electronic conduction. Lanthanumstrontium titanate is of interest as an anode material due to it resistance to sulfur [29], which, as mentioned above, is one of the limitations of Ni-YSZ cermet anodes. The conductivities of several (La,Sr)TiO3 materials in a fuel atmosphere are shown in Fig. 3 [9,25,3034]. The conductivity generally increases with lanthanum concentration and appears to level off at a lanthanum content of around 0.4. As in the case of yttrium doping, the conductivity is affected by strontium (A-

Fig. 4. Conductivity of (La,Sr)TiO3+ in fuel atmospheres at 800 C [19,25,3034]. The value of is for Ti4+, La3+ and Sr2+.

1532

J.W. Fergus / Solid State Ionics 177 (2006) 15291541

Fig. 5. Conductivity of (La,Sr)TiO3+ and (La,Ca)TiO3+ in fuel atmospheres at 1000 C [33,34,36,37]. The value of is for Ti4+, La3+, Ca2+ and Sr2+.

site according to the general perovskite formula ABO3) deficiency. This difference is reflected in the figure legend, which indicates the maximum oxygen content assuming that all the titanium is Ti4+ (also assuming La3+ and Sr2+). The data from Fig. 3 is plotted according to this maximum oxygen content (represented as in (La,Sr)TiO3+) in Fig. 4, in which case agreement among the different reports is better, indicating that oxygen excess (for the fully oxidized composition) leads to improved conductivity [35]. At low oxygen partial pressures, the titanium is reduced to Ti3+, such that the oxygen content is reduced by the formation of oxygen vacancies. Some of the variation between results from different researchers can be due to differences in the reducing atmosphere. In some cases, the exact oxygen partial pressure is not given (such as dry ArH2 [32]) or different oxygen partial pressures are used for different compositions [33]. Fig. 5 shows that the trend shown in Fig. 4

for (La,Sr)TiO3+ continues at higher temperatures (i.e. 1000 C) and that (La,Ca)TiO3+ shows a similar increase in conductivity with , which is due to increasing lanthanum content, but begins to decrease with oxygen excesses of 0.10 0.15 [33,34,36,37]. Fig. 6 shows that the conductivities for oxygen excess compositions are greater than those for oxygen deficient compositions [25,3741]. Note that A-site deficiency can increase the conductivity for compositions with the same oxygen deficiency. (La,Sr)TiO3 has been doped with several transition metals (Ni, Co, Cu, Cr and Fe) and Ce [42]. The most effective among these dopants is cerium, which significantly decreases the polarization resistance, although iron also produces modest improvements. Other titanates investigated as SOFC anodes include other A-site deficient lanthanides (Nd and Pr) [43] and Fe-doped BaTiO3 [44].

Fig. 6. Conductivity of (La,Sr/Ca)(Ti,Sc/Nb/Cr)O3 in fuel atmospheres at 9001000 C [25,3741]. The value of is for Nb5+, Ti4+, La3+, Sc3+, Cr3+, Ca2+ and Sr2+.

J.W. Fergus / Solid State Ionics 177 (2006) 15291541

1533

Fig. 7. Conversion of methane in water at 800900 C by lanthanum chromite based anode [5461].

2.2. Chromites Fig. 6 shows that chromium has been used as a dopant for titanate anodes. Further increase in the chromium content leads to a transition from n-type to p-type conduction [39,45,46]. LaCrO3 is a stable perovskite [47] and, when doped with calcium or strontium, is the most common ceramic interconnect material for SOFCs [48]. Strontium doping improves the stability in reducing environments [49,50] and reduces the polarization resistance [51]. The polarization resistance can be further decreased with the addition of nickel [52]. Strontium-doped lanthanum chromite does not catalyze carbon deposition and thus is a potential anode for the direct oxidation of methane [53]. Fig. 7 shows the catalytic activity for several lanthanum chromite based anodes as a function of methane-to-water ratio [5461], where the catalytic activity is

represented by the methane conversion (CCH4) according to Eq. (1),   pCH4 f CCH4 100%d 1 pCH4 i 1

where (pCH4)i and (pCH4)f are the methane partial pressures before and after conversion, respectively. Ruthenium is the most effective dopant for increasing the conversion rate, but nickel is also effective. For very reducing conditions, such as pCH4/pH2O = 18.7, carbon deposition, which will increase the apparent conversion rate, has been observed [59]. Fig. 8 shows that the conversion rates are much higher for the reaction of methane with oxygen and that nickel and iron both improve the conversion rates [55,56,59]. The polarization resistances for some ruthenium-doped lanthanum

Fig. 8. Conversion of methane in oxygen at 900 C by lanthanum chromite based anode [55,56,59].

1534

J.W. Fergus / Solid State Ionics 177 (2006) 15291541

Fig. 9. Polarization resistance of (La,Sr)(Cr,Ru)O3 in H2/H2O and CH4/H2O at 850 C [6264].

strontium chromites are shown in Fig. 9 [6264]. The polarization resistances are higher in methane as compared to hydrogen for all compositions. The amount of ruthenium has little effect on the resistance (i.e. 0.02 and 0.05 ruthenium are about the same), but there is clearly a minimum in resistance when the strontium content is 0.30. The catalytic activity of perovskite oxides has been shown to be highest when the oxide-ion conductivity is high [65], so the B-site of strontium-doped lanthanum chromite has been doped with various transition metals (Mn, Fe, Co, Ni, Cu) to create oxygen vacancies [12]. Of these dopants, Ni provides the lowest polarization resistance, but has poor stability in low oxygen partial pressures. Large manganese additions (e.g. La0.75Sr0.25Cr0.5Mn0.5O3) result in a low polarization resistance and the reformation of methane without excess steam

[6668]. The success of this oxide has led to its use with electrolytes in composite anodes for direct oxidation of methane [6972]. Vanadium additions have also been shown to improve the catalytic properties of lanthanumstrontium chromites for steam reforming [73]. 2.3. Other perovskites The beneficial effect of vanadium has been further exploited in the development of strontium-doped lanthanum vanadate anodes, which have good sulfur tolerance [74,75]. Fig. 10 shows that the conductivity of (La,Sr)VO3 is high and is increased further with the nickel doping [7678]. Although stability in both oxidizing and reducing conditions is a challenge with vanadates [12], they have been used with NiO

Fig. 10. Conductivity of (La,Sr)(Ti,Ni,V)O3 [7678] and Sr(Nb,Mn,Cu)O3 [98,99] in fuel atmospheres at 800 C and 900 C, respectively.

J.W. Fergus / Solid State Ionics 177 (2006) 15291541

1535

Fig. 11. Conductivity of Y/Pr-doped BaCeO3 [95,101103] in fuel atmospheres at 800 C.

[79,80] or nickel [81,82] in composite anodes with reduced polarization resistance. The addition of ruthenium has been shown to improve the steam reforming performance [73]. Ferrites have also been investigated as potential anode materials. One of the most promising alternatives to the most commonly used SOFC cathode, lanthanum strontium manganate (LSM), is lanthanum strontium cobaltite ferrite (LSCF) due to its high conductivity [8385], good oxygen exchange kinetics [86,87] and low polarization resistance [88]. LSCF also has good catalytic activity for methane conversion, but has poor stability at low oxygen partial pressures [89,90]. The stability, however, can be improved by mixing with Gd-doped ceria [91,92]. Strontiumcobaltferrites are also good mixed conductors, but are p-type conductors and are thus more suited for cathodes or for gas separation membranes [9395]. However, chromium-doped lanthanumstrontium ferrite has

shown promise for catalyzing the methane oxidation reaction [96]. Perovskite gallates and niobates have been investigated as potential anode materials [97100]. Although, as shown in Fig. 10 [98,99], their conductivities are relatively low, they show good compatibility with lanthanum strontium gallium magnesium electrolyte materials [100]. The perovskite barium cerate is a solid electrolyte, but the reduction of cerium at low oxygen partial pressures results in an increase in electronic conduction, so fuel cells using doped barium cerate as both the electrolyte and the anode have been investigated [101,102]. Fig. 11 shows that the conductivity is relatively low, but generally increases with yttrium doping [95,101103]. Electronic conduction can also be induced in an electrolyte by doping. For example, the addition of titanium to the perovskite oxide-ion conducting

Fig. 12. Conductivity of yttriumzirconiumtitanium oxides (YZT) in fuel atmospheres at 8001000 C [103,107109,114].

1536

J.W. Fergus / Solid State Ionics 177 (2006) 15291541

LaAlO3 leads to n-type conduction at low oxygen partial pressures [104]. 3. Cubic fluorite structures Solid electrolytes with other structures have been used as anode materials through inherent reduction of one of the ions or through doping. In either case, good contact and compatibility between the anode and the electrolyte is expected. 3.1. Zirconia-based Stabilized zirconia remains a pure ionic conductor in SOFC fuel atmospheres, so doping is required to induce electronic conductivity. Fortunately, titanium, the reduction of which can lead to n-type electronic conduction, has a large solubility in yttria-stabilized zirconia [105107]. Electronic conduction is generally attributed to a Ti4+/Ti3+ small polaron hopping mechanism [108110]. The presence of Ti 3+ has been confirmed with electron spin resonance measurements [110 112], but ion-blocking measurements indicate that association of the titanium defects with oxygen vacancies leads to a large polaron mechanism for large dopant levels [113]. Fig. 12 shows that the total conductivity of yttriumtitaniumzirconium oxides (YTZ) does not vary strongly with titanium content [103,107109,114], so the effect of titanium is to increase the electronic, rather than total, conductivity. The conductivity of YTZ is lower than most other potential anode materials, but YTZ does exhibit good electrode kinetics [115,116]. The impedance spectra contain two contributions; a small highfrequency contribution, which is attributed to charge transfer, and a larger low-frequency contribution, which is attributed to adsorption on the electrode. The high-frequency and total (approximately equal to the low-frequency) resistances for YTZ in CH4/H2O and H2/H2O are shown in Fig. 13 [116,117]. The resistances in methane are higher than those in hydrogen. In

most cases, the resistance decreases with increasing fuel content as would be expected for an n-type conductor. However, the total resistance for YTZ in CH4/H2O initially increases with methane content, which is attributed to direct reaction of methane with oxygen ions in the anode when the methane content is much larger than the water vapor content [117]. Titania-doping of YSZ in cermets (i.e. Ni-YTZ cermets) has been shown to improve the electrode performance [118,119]. Niobium-doped zirconia has also been investigated as an SOFC anode material [120122]. Niobium is not as easily reduced as titanium and is thus not as effective for increasing the electronic conductivity. Partial replacement of yttrium in YTZ with scandium has been used to increase the conductivity [123 125]. The increased conductivity is attributed to the smaller size mismatch between scandia and zirconia as compared to that between yttria and zirconia, which is consistent the higher conductivity of scandia-stabilized zirconia as compared to that of yttria-stabilized zirconia [126]. Ceria has also been added, along with titanium, to zirconia [127,128], which provides an additional ion for reduction and leads to the transition to significant n-type electronic conduction occurring at higher oxygen partial pressures [127]. The increased size of Ce3+, as compared to Ce4+, leads to an increase in the coefficient of thermal expansion for large cerium additions [128]. 3.2. Ceria-based Doped-ceria is an attractive electrolyte for SOFCs, but as discussed above, the reduction of Ce4+ to Ce3+ leads to electronic conduction, so doped ceria has been considered for both the electrolyte and anode in SOFCs [129,130]. The conductivities of some doped ceria materials are shown in Fig. 14 [103,131135]. The high conductivities of gadoliniaand samaria-doped ceria has been attributed to the good match in ionic radii [133]. Although the conductivity of terbium-doped ceria is lower than those for other dopants, the

Fig. 13. High frequency (HT) and total resistances of yttriumtitaniumzirconium oxides (YTZ) in CH4/H2O and H2/H2O at 932 C [116,117].

J.W. Fergus / Solid State Ionics 177 (2006) 15291541

1537

Fig. 14. Conductivity of doped ceria in fuel atmospheres at 800 C [103,131135].

Tb4+/ Tb3+ reaction may improve the electrode kinetics [134]. The conductivity of neodymia-doped ceria is not shown, but its performance is comparable to that of gadolinia-doped ceria in hydrogen [136]. The conductivity of doped ceria is relatively low, so cermets of ceria are preferred to singlephase electrodes for better performance [137]. Gadoliniadoped ceria has shown good catalytic activity for the oxidation of methane [138] with little or no carbon deposition [139]. The composition with the lowest polarization resistance is Ce0.9Gd0.1O2 [140]. The electrocatalytic activity can be improved with the addition of small amounts of ruthenium [141] or nickel [80,131,142]. Like YSZ, ceria is used in cermets with nickel [143,144] and copper [144,145], and as an addition to NiYSZ cermets [146]. (Ce,La)O2 has been mixed with an conducting oxide, (La,Sr)TiO3, which as

shown in Fig. 15, goes through a maximum in conductivity for 20% (Ce,La)O2. 4. Other oxide structures One of the other crystal structures used for oxide anodes is the pyrochlore structure [75,148150]. Fig. 16 shows that the conductivity of the pyrochlore Gd2Ti2O7 increases when Ti is replaced with Mo or Mn [75,151154]. Although the conductivities of these compounds are high, they have poor redox stability and thus do not perform well as SOFC anodes [11,12,154,155]. The spinel, Mg2TiO4, can be reduced at low oxygen partial pressures and becomes an electronic conductor [156]. Although the results in Fig. 17 [157,158] indicate that the magnitude of

Fig. 15. Conductivity of (La,Sr)TiO3(Ce,La)O2 mixtures in fuel atmospheres at 1000 C [147].

1538

J.W. Fergus / Solid State Ionics 177 (2006) 15291541

Fig. 16. Conductivity of the pyrochlore Gd2(Ti, Mn, Mo)2O7 in fuel atmospheres at 800 C [75,151154].

the conductivity does not depend strongly on the dopant ion, Co and Fe lead to n-type conduction, whereas Mn leads to p-type conduction [158]. Another structure investigated for SOFC anodes is the tungsten bronze structure [159162], which, as shown in Fig. 17, has a reasonablygoodconductivity.However,thecoefficientsofthermal expansionofthesecompounds(e.g.6.7 10 6 C 1 forSr0.2Ba0.4Ti0.2Nb0.8, which has a conductivity of 10 Scm 1 [162]) are significantlylowerthanthatofzirconia. Some NbTiO compounds form compounds with good conductivity at low oxygen partial pressure [163]. The highest conductivity occurs for the compound Nb2Ti2O7, which forms the rutile structure. Fig. 17 shows that the conductivities of

some (Ti,Nb,Fe)O2 compounds are greater than 100 Scm 1 [164], but these compounds have low coefficients of thermal expansion (2.36.5 10 6 C 1 [12,163,164]). Bismuth oxide is a potential solid electrolyte, but suffers from electronic conductivity at low oxygen partial pressures. This, however, is desired in an anode, so bismuth-based oxides have been investigated as SOFC anodes [165]. Bi2O3Ta2O5 mixtures performed well as the anode in fuels cells for direct reaction of butane, but the stability of the anode may be an issue due to the relatively high vapor pressure of bismuth, which forms in reducing atmospheres. A fuel cell with a Bi2O3Ta2O5 anode has been shown to operate for 200 h at 700 C with a relatively stable cell voltage [165].

Fig. 17. Conductivity of oxides with the spinel, (Mg,Ni/Co/Fe/Mn)TiO4 [157,158], tungsten bronze, (Ba/Sr,Sr/La/Ca)0.6(Ti,Nb)O3 [159] and rutile structures (Nb,Ti, Fe)O2 [164] in fuel atmospheres at 900930 C.

J.W. Fergus / Solid State Ionics 177 (2006) 15291541

1539

5. Conclusions Oxide anodes can overcome some of the limitations of Ni YSZ cermet anodes. In addition to general requirements, such as stability and compatibility with other fuel cell components (e.g. chemical stability and thermal expansion match), oxide anodes must have the good transport and catalytic properties. While the ideal oxide anode material has not been discovered, progress is being made in the development of mixed-conducting single-phase oxides for use as anodes in solid oxide fuel cells. References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] B.C.H. Steele, A. Heinzel, Nature 414 (6861) (2001) 345. P. Singh, N.Q. Minh, Int. J. Appl. Ceram. Tech. 1 (1) (2004) 5. S.C. Singhal, Solid State Ionics 152153 (2003) 405. N.Q. Minh, Solid State Ionics 174 (2004) 271. H. Tu, U. Stimming, J. Power Sources 127 (12) (2004) 284. K. Sasaki, K. Watanabe, K. Shiosaki, K. Susuki, Y. Teraoka, J. Electroceram. 13 (2004) 669. P. Lamp, J. Tachtler, O. Finkenwirth, S. Mukerjee, S. Shaffer, Fuel Cells 3 (3) (2003) 146. S.P. Jiang, S.H. Chan, J. Mater. Sci. 39 (2004) 4405. W.Z. Zhu, S.C. Deevi, Mater. Sci. Eng., A A362 (2003) 228. A. McEvoy, in: S.C. Singhal, K. Kendall (Eds.), High Temperature Solid Oxide Fuel Cells, Elsevier Ltd, Oxford, UK, 2003, p. 149. A. Atkinson, S. Barnett, R.J. Gorta, J.T.S. Irvine, A. McEvoy, M. Mogensen, S.C. Singhal, J. Vohs, Nat. Mater. 3 (1) (2004) 17. S. Tao, J.T.S. Irvine, Chem. Rec. 4 (2004) 83. J.T.S. Irvine, A. Sauvet, Fuel Cells 1 (34) (2001) 205. S. McIntosh, M.J. Vohs, R.J. Gorte, J. Electrochem. Soc. 150 (4) (2003) A470. Y. Matsuzaki, I. Yasuda, Solid State Ionics 132 (2000) 261. B.A. Boukamp, Nat. Mater. 2 (5) (2003) 294. N.H. Chan, R.K. Sharma, D.M. Smyth, J. Electrochem. Soc. 128 (8) (1981) 1762. S. Hui, A. Petric, Mater. Res. Bull. 37 (2002) 1215. S. Hui, A. Petric, J. Electrochem. Soc. 149 (1) (2002) J1. Y. Yoo, S. Koutcheiko, A. Petric, in: J. Huijsmans (Ed.), Fifth European Solid Oxide Fuel Cell Forum, European Fuel Cell Forum, Oberrohfdorf, Switzerland, 2002, p. 507. S. Hui, A. Petric, J. Eur. Ceram. Soc. 22 (2002) 1673. Q. Fu, F. Tietz, D. Stver, Electrochem. Soc. Proc. 200507 (SOFC IX) (2005) 1417. R. Moos, K.H. Hrdtl, J. Am. Ceram. Soc. 80 (10) (1997) 2549. T. Kolodiazhnhyi, A. Petric, J. Electroceram. 15 (2005) 5. P.R. Slater, D.P. Fagg, J.T.S. Irvine, J. Mater. Chem. 7 (12) (1997) 2495. S. Koutcheiko, Y. Yoo, A. Petric, in: J. Huijsmans (Ed.), Fifth European Solid Oxide Fuel Cell Forum, European Fuel Cell Forum, Oberrohfdorf, Switzerland, 2002, p. 655. H. He, Y. Huang, J.M. Vohs, R.J. Gorte, Solid State Ionics 175 (2004) 171. S. Koutcheiko, Y. Yoo, A. Petric, I. Davidson, Ceram. Int. 32 (2005) 67. R. Mukundan, E.L. Brosha, F.H. Garzon, Electrochem. Solid-State Lett. 7 (1) (2004) A5. G. Pudmich, W. Jungen, F. Tietz, Electrochem. Soc. Proc. 9919 (SOFC VI) (1999) 577. P. Pudmich, B.A. Boukamp, M. Gonzalez-Cuenca, W. Jungen, W. Zipprich, F. Tietz, Solid State Ionics 135 (2000) 433. J. Canales-Vzquez, S.W. Tao, J.T.S. Irvine, Solid State Ionics 159 (2003) 159. R. Moos, K.H. Schllhammer, Appl. Phys. A 65 (1997) 291. O.A. Marina, N.L. Canfield, J.W. Stevenson, Solid State Ionics 149 (2002) 21. J. Canales-Vzquez, W. Zhou, J.T.S. Irvine, Ionics 8 (34) (2002) 252.

[21] [22] [23] [24] [25] [26]

[27] [28] [29] [30] [31] [32] [33] [34] [35]

[36] V. Vashook, L. Vasylechko, M. Knapp, H. Ullmann, U. Guth, J. Alloys Compd. 354 (2003) 13. [37] V. Vashook, R. Mller, J. Zosel, K. Ahlborn, H. Ullmann, U. Guth, Mat.Wiss. U. Werkstofftech. 33 (2002) 335. [38] V. Vashook, L. Vasylechko, H. Ullmann, U. Guth, Solid State Ionics 158 (2003) 317. [39] V. Vashook, L. Vasylechko, J. Zosel, R. Mller, E. Ahlborn, U. Guth, Solid State Ionics 175 (2004) 151. [40] V. Vashook, J. Zosel, W. Preis, W. Sitte, Guth, Solid State Ionics 175 (2004) 441. [41] J. Canales-Vzquez, J.C. Ruiz-Morales, J.T.S. Irvine, W. Zhou, J. Electrochem. Soc. 152 (7) (2005) A1458. [42] O.A. Marina, L.R. Pederson, in: J. Huijsmans (Ed.), Fifth European Solid Oxide Fuel Cell Forum, European Fuel Cell Forum, Oberrohfdorf, Switzerland, 2002, p. 481. [43] F.J. Lepe, J. Fernndez-Urbn, L. Mestres, M.L. Martnez-Sarrin, J. Power Sources 151 (2005) 74. [44] E.A. Mashkina, L.A. Dunuyshkina, A.K. Demin, M. Gbbels, R. Hock, A. Magerl, in: J. Huijsmans (Ed.), Fifth European Solid Oxide Fuel Cell Forum, European Fuel Cell Forum, Oberrohfdorf, Switzerland, 2002, p. 695. [45] V. Vashook, L. Vasylechko, J. Zosel, W. Gruner, H. Ullman, U. Guth, J. Solid State Chem. 177 (2004) 3784. [46] M. Gonzlez-Cuenca, W. Zipprich, B.A. Boukamp, G. Pudmich, F. Tietz, Fuel Cells 1 (34) (2001) 256. [47] T. Nakamura, G. Petzow, L.J. Gauckler, Mater. Res. Bull. 14 (5) (1979) 649. [48] J.W. Fergus, Solid State Ionics 171 (12) (2004) 115. [49] J. Sfeir, J. Power Sources 118 (2003) 276. [50] J. Sfeir, J. Van herle, A.J. McEvoy, J. Eur. Ceram. Soc. 19 (1999) 897. [51] P. Vermoux, Ionics 3 (34) (1997) 270. [52] J. Sfeir, J. Van herle, R. Vasquez, in: J. Huijsmans (Ed.), Fifth European Solid Oxide Fuel Cell Forum, European Fuel Cell Forum, Oberrohfdorf, Switzerland, 2002, p. 570. [53] J. Vulliet, B. Morel, J. Laurencin, G. Gauthier, L. Bianchi, S. Giraud, H.Y. Henry, F. Lefebvre-Joud, Electrochem. Soc. Proc. 200307 (SOFC VIII) (2003) 803. [54] A.-L. Sauvet, J. Fouletier, J. Power Sources 101 (2001) 259. [55] S. Tao, J.T.S. Irvine, Electrochem. Soc. Proc. 200307 (SOFC VIII) (2003) 793. [56] S. Tao, J.T.S. Irvine, Chem. Mater. 16 (2002) 4116. [57] A.L. Sauvet, J.T.S. Irvine, Solid State Ionics 167 (2004) 1. [58] A.-L. Sauver, J.T.S. Irvine, in: J. Huijsmans (Ed.), Fifth European Solid Oxide Fuel Cell Forum, European Fuel Cell Forum, Oberrohfdorf, Switzerland, 2002, p. 490. [59] J. Sfeir, P.A. Buffat, P. Mckli, N. Xanthopoulos, R. Vasquez, H.J. Mathieu, J. Van herle, K.R. Thampi, J. Catal. 202 (2001) 229. [60] A.J. McEvoy, Mat.-Wiss. U. Werkstofftech 33 (2002) 331. [61] P. Vernoux, J. Guindet, M. Kleitz, J. Electrochem. Soc. 145 (10) (1998) 3487. [62] P. Vernoux, E. Djurado, M. Guillodo, J. Am. Ceram. Soc. 84 (10) (2001) 2289. [63] A.-L. Sauvet, J. Guindet, J. Fouletier, Ionics 5 (1-2) (1999) 150. [64] A.-L. Sauvet, J. Fouletier, Electrochim. Acta 47 (2001) 987. [65] R. Doshi, C.B. Alcock, N. Gunasekaran, J.J. Carberry, J. Catal. 140 (1993) 557. [66] S. Tao, J.T.S. Irvine, Nat. Mater. 2 (5) (2003) 320. [67] S. Tao, J.T.S. Irvine, J. Electrochem. Soc. 151 (2) (2004) A252. [68] S. Tao, J.T.S. Irvine, J.A. Kilner, Adv. Mater. 17 (2005) 1734. [69] S.P. Jiang, X.J. Chen, S.H. Chan, J.T. Kwok, K.A. Khor, Solid State Ionics 177 (12) (2006) 149. [70] X.J. Chen, K.A. Khor, S.H. Chan, S.P. Jiang, Electrochem. Soc. Proc. 200507 (SOFC IX) (2005) 1352. [71] S.A. Barnett, J. Liu, B.D. Madsen, Z. Zhan, Electrochem. Soc. Proc. 200226 (Solid-State Ionic Devices III) (2002) 53. [72] J. Liu, B.D. Madsen, Z. Ji, S.A. Barnett, Electrochem. Solid-State Lett. 5 (6) (2002) A122.

1540

J.W. Fergus / Solid State Ionics 177 (2006) 15291541 [116] A. Kelaidopoulou, A. Siddle, A.L. Dicks, A. Kaiser, J.T.S. Irvine, Fuel Cells 1 (34) (2001) 226. [117] A. Kelaidopoulou, A. Siddle, A.L. Dicks, A. Kaiser, J.T.S. Irvine, Fuel Cells 1 (34) (2001) 219. [118] D. Skarmoutsos, P. Nikolopoulos, F. Tietz, I.C. Vinke, Solid State Ionics 170 (2004) 153. [119] S. Skarmoutsos, P. Nikolopoulos, A. Tsoga, Ionics 5 (56) (1999) 455. [120] J.T.S. Irvine, D.P. Fagg, J. Labrincha, F.M.B. Marques, Catal. Today 38 (1997) 467. [121] D.P. Fagg, J.T.S. Irvine, Ionics 4 (12) (1998) 61. [122] D.P. Fagg, A.J. Feighery, J.T.S. Irvine, J. Solid State Chem. 172 (2003) 277. [123] J.T.S. Irvine, S. Tao, A.J. Feighery, T.D. McColm, Electrochem. Soc. Proc. 200116 (SOFC VII) (2001) 738. [124] S. Tao, J.T.S. Irvine, J. Solid State Chem. 165 (2002) 12. [125] S. Tao, J.T.S. Irvine, J. Electrochem. Soc. 151 (4) (2004) A497. [126] J.W. Fergus, J. Mater. Sci. 38 (2003) 4259. [127] F. Capel, C. Moure, P. Durn, Ceram. Int. 28 (2002) 627. [128] R.A. Buchanan, J.T.S. Irvine, in: J. Huijsmans (Ed.), Fifth European Solid Oxide Fuel Cell Forum, European Fuel Cell Forum, Oberrohfdorf, Switzerland, 2002, p. 417. [129] M. Mogensen, D. Lybye, K. Kammer, N. Bonanos, Electrochem. Soc. Proc. 200507 (SOFC IX) (2005) 1068. [130] V.V. Kharton, F.M. Figueiredo, L. Havarro, E.N. Naumovich, A.V. Kovalevsky, A.A. Yaremchenko, A.P. Viskup, A. Carneiro, F.M.B. Marques, J.R. Frade, J. Mater. Sci. 36 (2001) 1105. [131] H. Ushida, M. Sugimoto, M. Watanabe, Electrochem. Soc. Proc. 2001 16 (SOFC VII) (2001) 653. [132] S. Wang, T. Kobayashi, M. Dokiya, T. Hashimoto, J. Electrochem. Soc. 147 (10) (2000) 3606. [133] H. Yahiro, K. Eguchi, H. Arai, Solid State Ionics 36 (1989) 71. [134] A. Martnez-Arias, A.B. Hungra, M. Fernndez-Garcia, A. Iglesias-Juez, J.C. Conesa, G.C. Mather, G. Munuera, J. Power Sources 151 (2005) 43. [135] I.K. Naik, T.Y. Tien, J. Electrochem. Soc. 126 (4) (1979) 562. [136] K. Yashiro, T. Nakamura, A. Kaimai, T. Otake, T. Kawada, J. Mizusaki, Electrochem. Soc. Proc. 200507 (SOFC IX) (2005) 1369. [137] A.O. Strmer, P. Holtappels, H.Y. Tu, U. Stimming, Mat.-Wiss. U. Werkstofftech. 33 (2002) 339. [138] P. Aguiar, W. Ramrez-Cabrera, A. Atkinson, L.S. Kershenbaum, D. Chadwick, Electrochem. Soc. Proc. 200116 (SOFC VII) (2001) 703. [139] O.A. Marina, C. Bagger, S. Primdahl, M. Mogensen, Solid State Ionics 123 (1999) 199. [140] K. Kammer, M. Mogensen, Electrochem. Solid-State Lett. 8 (2) (2005) A108. [141] H. Ushida, S. Suzuki, M. Watanabe, Electrochem. Soc. Proc. 200307 (SOFC VIII) (2003) 728. [142] H. Uchida, N. Osada, S. Suzuki, M. Watanabe, Electrochem. Soc. Proc. 200507 (SOFC IX) (2005) 1410. [143] B. Rsch, H. Tu, A.O. Strmer, A.C. Mller, U. Stimming, Electrochem. Soc. Proc. 200307 (SOFC VIII) (2003) 737. [144] E.V. Tsipis, V.V. Kharton, J.R. Frade, J. Eur. Ceram. Soc. 25 (2005) 2623. [145] M.L. Faro, G. Monforte, V. Antonucci, A.S. Aric, Electrochem. Soc. Proc. 200507 (SOFC IX) (2005) 1445. [146] P.E. Murray, T. Tsai, S.A. Barnett, Nature 400 (6745) (1999) 649. [147] O.A. Marina, J.W. Stevenson, Electrochem. Soc. Proc. 200226 (SolidState Ionic Devices III) (2002) 91. [148] M. Smith, A.J. McEvoy, Electrochem. Soc. Proc. 200507 (SOFC IX) (2005) 1437. [149] S. Zha, C. Zhe, M. Liu, Electrochem. Solid-State Lett. 8 (8) (2005) A406. [150] S. Kramer, M. Spears, H.L. Tuller, Solid State Ionics 72 (1994) 59. [151] O. Porat, C. Heremans, H.L. Tuller, J. Am. Ceram. Soc. 80 (9) (1997) 2278. [152] O. Porat, C. Heremans, H.L. Tuller, Solid State Ionics 94 (1997) 75. [153] O. Porat, M.A. Spears, C. Heremans, I. Kosacki, H.L. Tuller, Solid State Ionics 8688 (1996) 285. [154] J.J. Sprague, H.L. Tuller, J. Eur. Ceram. Soc. 19 (1999) 803. [155] H. Holtappels, F.W. Poulsen, M. Mogensen, Solid State Ionics 135 (2000) 675.

[73] P. Vernoux, M. Guillodo, J. Fouletier, A. Hammou, Solid State Ionics 135 (2000) 425. [74] L. Aguilar, S. Zha, S. Li, J. Winnick, M. Liu, Electrochem. Solid-State Lett. 7 (10) (2004) A324. [75] S. Li, W. Rauch, M. Liu, A. Burke, J. Winnick, Electrochem. Soc. Proc. 200226 (Solid-State Ionic Devices III) (2002) 113. [76] S. Hui, A. Petric, W. Gong, Electrochem. Soc. Proc. 9919 (SOFC VI) (1999) 632. [77] S. Hui, A. Petric, Solid State Ionics 143 (2001) 275. [78] Z. Cheng, S. Zha, L. Aguilar, M. Liu, Solid State Ionics 176 (2005) 1921. [79] B.D. Madsen, S.A. Barnett, Electrochem. Soc. Proc. 200507 (SOFC IX) (2005) 1185. [80] B.D. Madsen, S.A. Barnett, Solid State Ionics 176 (2005) 2545. [81] S. Primdahl, M. Mogensen, Solid State Ionics 152153 (2002) 597. [82] S. Primdahl, J.R. Hansen, L. Grahl-Madsen, P.H. Larsen, J. Electrochem. Soc. 148 (1) (2001) A74. [83] M.A.J. Cropper, S. Geiger, D.M. Jollie, J. Power Sources 131 (2004) 57. [84] K.C. Wincewicz, J.S. Cooper, J. Power Sources 140 (2005) 280. [85] L.J. Gauckler, D. Beckel, B.E. Buergler, E. Jud, U.P. Muecke, M. Prestat, J.L. Rupp, J. Richter, Chimia 58 (12) (2004) 837. [86] D. Stver, H.P. Buchkremer, S. Uhlenbruck, Ceram. Int. 3 (7) (2004) 1107. [87] N.P. Brandon, S. Skinner, B.C.H. Steele, Annu. Rev. Mater. Res. 33 (2003) 183. [88] W. Gong, S. Gopalan, U.B. Pal, J. Electroceram. 13 (13) (2004) 653. [89] A. Hartley, M. Sahibzada, M. Weston, I.S. Metcalfe, D. Mantzavinos, Catal. Today 55 (2000) 197. [90] M. Weston, I.S. Metcalfe, Solid State Ionics 113115 (1998) 247. [91] A.S. Aric, L.R. Gullo, D. La Rosa, V. Antonucci, Electrochem. Soc. Proc. 200507 (SOFC IX) (2005) 1396. [92] A. Sin, E. Kopnin, Y. Dubitsky, A. Zaopo, A.S. Aric, L.R. Gullo, D. La Rosa, V. Antonucci, J. Power Sources 145 (2005) 68. [93] H. Ullmann, N. Trofimenko, A. Naoumidis, D. Stver, J. Eur. Ceram. Soc. 19 (1999) 791. [94] B. Ma, U. Balachandran, Mater. Res. Bull. 33 (2) (1998) 223. [95] U. Balachandran, B. Ma, P.S. Maiya, R.L. Mieville, J.T. Dusek, J.J. Picciolo, J. Guan, S.E. Dorris, M. Liu, Solid State Ionics 108 (1998) 363. [96] E. Ramirez-Cabrera, A. Atkinson, R. Rudkin, D. Chadwick, in: J. Huijsmans (Ed.), Fifth European Solid Oxide Fuel Cell Forum, European Fuel Cell Forum, Oberrohfdorf, Switzerland, 2002, p. 546. [97] T.D. McColm, J.T.S. Irvine, Solid State Ionics 152153 (2002) 615. [98] S. Tao, J.T.S. Irvine, Solid State Ionics 154155 (2002) 659. [99] S. Tao, J.T.S. Irvine, J. Mater. Chem. 12 (2002) 2356. [100] Q. Fu, X. Xu, D. Peng, X. Liu, G. Meng, J. Mater. Sci. 38 (2003) 2901. [101] A. Tomita, T. Hibino, M. Sano, Electrochem. Solid-State Lett. 8 (7) (2005) A333. [102] D. Hirabahashi, A. Tomita, M.E. Brito, T. Hibino, U. Harada, M. Nagao, M. Sano, Solid State Ionics 168 (2004) 23. [103] Y. Hibino, A. Hashimoto, M. Suzuki, M. Sano, J. Electrochem. Soc. 149 (11) (2002) A1503. [104] J.Y. Park, G.M. Choi, Solid State Ionics 176 (2005) 2807. [105] M.T. Colomer, P. Durn, A. Caballero, J.R. Jurado, Mater. Sci. Eng., A A229 (1997) 114. [106] A.J. Feighery, J.T.S. Irvine, D.P. Fagg, A. Kaiser, J. Solid State Chem. 143 (1999) 273. [107] A. Kaiser, A.J. Feighery, J.T.S. Irvine, Electrochem. Soc. Proc. 9919 (SOFC VI) (1999) 541. [108] P. Durn, F. Capel, C. Moure, A.R. Gonzlez-Elipe, A. Caballero, M.A. Baares, J. Electrochem. Soc. 146 (7) (1999) 2425. [109] H. Arashi, H. Naito, Solid State Ionics 5354 (1992) 436. [110] K.E. Swider, W.L. Worrell, J. Electrochem. Soc. 143 (11) (1996) 3706. [111] K.E. Swider, W.L. Worrell, J. Am. Ceram. Soc. 78 (4) (1995) 961. [112] F. Tietz, W. Jungen, P. Lersch, M. Figaj, K.D. Becker, D. Skarmoutsos, Chem. Mater. 14 (2002) 2252. [113] M.T. Colomer, J.R. Jurado, J. Solid State Chem. 165 (2002) 79. [114] F. Tietz, I.A. Raj, D. Stver, Br. Ceram. Trans. 103 (5) (2004). [115] P. Holtappels, J. Bradley, J.T.S. Irvine, A. Kaiser, M. Mogensen, J. Electrochem. Soc. 148 (8) (2001) A923.

J.W. Fergus / Solid State Ionics 177 (2006) 15291541 [156] [157] [158] [159] [160] [161] D.P. Fagg, S.M. Fray, J.T.S. Irvine, Solid State Ionics 72 (1994) 235. D.M. Flot, J.T.S. Irvine, Ionics 4 (34) (1998) 175. D.M. Flot, J.T.S. Irvine, Solid State Ionics 135 (2000) 513. P.R. Slater, J.T.S. Irvine, Solid State Ionics 120 (1999) 125. P.R. Slater, J.T.S. Irvine, Solid State Ionics 124 (1999) 61. E. Kendrick, M.S. Islam, P.R. Slater, Solid State Ionics 176 (3940) (2005) 2975.

1541

[162] A. Kaiser, J.L. Bradley, P.R. Slater, J.T.S. Irvine, Solid State Ionics 135 (2000) 519. [163] C.M. Reich, A. Kaiser, J.T.S. Irvine, Fuel Cells 1 (34) (2001) 249. [164] A. Lashtabeg, J.T.S. Irvine, A. Feighery, Ionics 9 (34) (2003) 220. [165] D. Hirabahashi, A. Hashimoto, T. Hibino, U. Harada, M. Sano, Electrochem. Solid-State Lett. 7 (5) (2004) A108.

You might also like