Sunteți pe pagina 1din 4

Ionics (2008) 14:3336 DOI 10.

1007/s11581-007-0169-9

ORIGINAL PAPER

Defect thermodynamic and transport properties of nanocrystalline Gd-doped ceria


S. Surble & G. Baldinozzi & M. Doll & D. Gosset & C. Petot & G. Petot-Ervas

Received: 7 September 2007 / Accepted: 2 October 2007 / Published online: 29 October 2007 # Springer-Verlag 2007

Abstract Nanocrystalline CeO2-doped (5, 7.5, 10, and 15 mol%) Gd2O3 powders, with a particle size of about 17 nm, were synthesized through the combustion of glycine/ nitrate gels. Dense nanocrystalline materials were obtained by hot uniaxial sintering. Optical microscopy, scanning electron microscopy and transmission electron microscopy examinations, as well as X-ray diffraction analyses, have allowed us to characterize these polycrystals. The grain sizes, included between 10 and 80 nm, depend on both the sintering temperature and the amount of dopant. A comparison of the transport properties of these nanocrystalline samples to the values obtained with coarsened grained materials of same composition shows that the ionic conductivity passes through a maximum for mean grain sizes included between 300 and 500 nm. Furthermore, an enhancement of the ionic conductivity is observed when the amount of dopant increases. This was attributed to a grain-size-dependent gadolinium segregation at the periphery of the grains confirmed by X-ray photoelectron spectroscopy characterizations. Keywords Yttria-doped Ceria . Nanocrystalline materials . Segregation . Ionic conductivity . SOFC

Introduction Solid oxide fuel cells are regarded to be among the cleanest energy conversion systems. However, to be economically competitive, it is necessary to lower their temperature
Paper presented at the 11th EuroConference on the Science and Technology of Ionics, Batz-sur-Mer, France, 915 Sept. 2007. S. Surble : G. Baldinozzi : M. Doll : D. Gosset : C. Petot : G. Petot-Ervas (*) Research Group CNRS/SPMS- Ecole Centrale Paris, 92295 Chtenay Malabry/SRMA-CEA Saclay, 91191 Gif sur Yvette, France e-mail: georgette.petot@yahoo.fr

operation in the range 500600 C. This will allow not only to minimize the cost of fabrication through the use of less expensive materials for interconnections, heat exchangers, and structural components but also to improve the performance of the cells due to a lower degradation rate of components. To realize these requirements, this implies new system concepts and the use of electrode and electrolyte materials with high oxygen ion conductivity, good mechanical stability, and higher performance at the electrode levels (solid electrolyte/electrode material interface) to make the electrode over potential losses acceptable. In the last decade, nanocrystalline materials are considered to have a high potential to overcome these constraints. They possess a wide range of original properties, including enhanced mechanical properties and chemical reactivity, useful in various technological fields, such as in catalysis and in fuel cells. Nevertheless, to date, their long term behavior is unknown, their physico-chemical stability is not fully understood yet, and the impact of this change of scale may have on the transport properties remains unclear. The aim of this paper was then to gain an understanding on the influence of the nanostructure on the transport properties. Gd-doped dense nanocrystalline ceria was chosen due to its higher ionic conductivity as compared to yttria-stabilized zirconia and to its enhanced electrode reactions due to the catalytic effect of the Ce+3/Ce+4 couple. Its electrical conductivity has been characterized in direct comparison with polycrystalline specimens whose grain sizes were in the micrometer range. The set of results is analyzed, taking into account the segregation of gadolinium at the periphery of the grains.

Sample synthesis and characterization The starting materials were Gd-doped-CeO2 nanocrystalline powders prepared by the combustion technique. A mixture

34 Fig. 1 Microstructural characterization (TEM) of Gd2O3 (7.5 mol%)-doped ceria powder synthesized by the combustion technique. Left panel: bright field image showing the narrow distribution of the particle sizes. Right panel: dark field image showing that each grain is generally a single crystal

Ionics (2008) 14:3336

[1] of glycine and gadolinium nitrate (Gd(NO3)3.6H2O) was added to an aqueous solution containing 0.1 mol of Ce (NO3)3.6H2O to obtain the desired Ce/Gd molar ratio and a molar ratio of 2 between nitrate ions and glycine. This ratio allows to control the combustion temperature near 1,100 C and to achieve a uniform doping of the obtained powders while controlling their size. Highly homogeneous and nonaggregated powders were obtained after calcinations at 500 C in air. X-ray diffraction shows that the powders are a single phase and that the grain sizes is below 17 nm. Transmission electron microscopy (TEM) observations confirm this result (Fig. 1). They show that the grains consist of wellcrystallized single crystals having regular shapes and sizes between 5 and 17 nm. Calcined powders were sintered into pellets under 250 MPa uniaxial pressure at 750 C in WC dies. A series of samples (10 mol% Gd2O3) was also sintered at 850 C to study the influence of the sintering temperature. The polycrystals have densities >95% of the theoretical density. Scanning electron microscopy (SEM) shows that all samples have an average grain size less than 100 nm, which depends on the amount of dopant. As an example, Fig. 2 displays the microstructures of two samples sintered at the same temperature but doped with a different amount of Gd (7.5 and 15 mol% Gd2O3). This figure clearly shows that the more doped sample exhibits the smaller grain size, a tendency which is observed throughout all the compositions. Moreover, for a given composition, the grain size increases when both the sintering temperature and the annealing time are increased. In Fig. 3, we have reported the microstructure of a reference ceria sample doped with Gd (10 mol% Gd2O3) and obtained by natural sintering of nanocrystalline powders at 1,400 C during 2 h [2]. These sintering conditions lead to a final average grain size of about 500 nm (Fig. 3a). Furthermore, the TEM image shows the absence of second phase along grain boundaries.

Fig. 2 Microstructural characterization (SEM) of the Gd-doped ceria polycrystals containing 7.5 (a) and 15 mol% Gd2O3 (b) sintered at 750 C under 250 MPa

Ionics (2008) 14:3336

35

Fig. 4 Influence of the grain size and of the amount of gadolinium on the ionic conductivity of nanocrystalline ceria. Comparison with the values obtained with a polycrystal R prepared by conventional sintering

of a reference sample (R) doped with 10 mol% Gd2O3 (Fig. 3) and obtained by natural sintering at 1,400 C for 2 h. These results clearly show that the ionic conductivity of the nanocrystalline samples, with an average grain size lower than 150 nm, is smaller than the ionic conductivity of the reference material R. They also show that the ionic conductivity depends on both the grain size and the Gd composition. As shown in Fig. 5, the ionic conductivity of polycrystalline specimens, whose grain sizes are in the micrometer range, depends not only on the cooling rate at the end of sintering (P8 cooled at a higher cooling rate than that of P6) but also on both the grain size and the Gd amount, as the

Fig. 3 a SEM image of a fracture surface of a Gd (10 mol% Gd2O3)doped ceria reference polycrystal (R) sintered at.1,400 C for 2 h showing the distribution of grain sizes and b TEM image showing the absence of precipitates at grain boundaries

Results The electrical conductivity was characterized by complex impedance spectroscopy [2, 6] using silver as an electrode material (=g/R, where g=l/s). In Fig. 4, we have reported the results obtained with nanocrystalline samples doped with different Gd amounts and sintered according to the conditions detailed above. These results are compared to those

Fig. 5 Influence of the grain size and of the amount of gadolinium on the ionic conductivity of microcrystalline Gd-doped ceria

36

Ionics (2008) 14:3336

Table 1 Influence of the grain size, of the amount of Gd and of the cooling rate at the end of sintering on the amount of Gd segregated in the first monolayers at the periphery of the grains Gd/Ce initial F P9 P8 P6 R 0.22 0.11 0.22 0.22 0.22 Gd/Ce <6 monolayers 2.6 1.3 1.5 1.7 0.9 Gd/Ce <11 monolayers 0.6 0.3 0.4 0.4 0.3 grain (m) 1020 0.52 0.53 0.73 0.20.8 Sintering conditions 1,600 1,300 1,300 1,300 1,400 C, 30 h C, 1 h C, 1 h C, 1 h C2 h

All P samples were sintered at 1,300 C under a pressure of 0.24 MPa, but the cooling rate at the end of sintering was faster for sample P8 than for sample P6.

nanocrystalline materials. However, it should be noted that the ionic conductivity of these samples increases as the grain size decreases. Therefore, according to the results reported in Fig. 4, the ionic conductivity goes through a maximum that occurs at about 300500 nm for the CeO2doped (10 mol%) Gd2O3 polycrystals.

fraction of grain boundaries in the sample but large enough to provide a significant enrichment of gadolinium (close to Gd/Ce 0.9) near the grain boundaries.

Conclusion In conclusion, a comparison between the ionic conductivity of dense nanocrystalline Gd-doped ceria samples (grain size <150 nm) and specimens whose grain sizes were in the micrometer range shows first that the ionic conductivity of Gd2O3 (10 mol %)CeO2 polycrystals goes through a maximum for an average grain size close to 300500 nm. Furthermore, these results also show that the conductivity values depend on both the nominal Gd composition and the sintering conditions. XPS analyses suggest that this behavior is related to the segregation of Gd in the region near the grain boundaries, enrichments which are less pronounced when the grain size decreases. Consequently, the ionic conductivity values obtained with the samples doped with 10 mol% Gd2O3 suggest that the maximum of conductivity corresponds to a mean enrichment at the periphery of the grains close to Gd/ Ce 0.9. The decrease of the ionic conductivity with the amount of gadolinium in nanocrystalline samples of same grain sizes supports this conclusion [4, 5].

Discussion To explain these results, it is important to take into account the occurrence of kinetic demixing phenomena during the sintering process [2, 6]. The X-ray photoelectron spectroscopy (XPS) analyses results reported in Table 1 show the influence of the grain size, of the amount of Gd, and of the cooling rate at the end of sintering on the amount of Gd segregated at the periphery of the grains (depth close to 0.2 nm). As already observed in previous studies [36], the kinetic demixing effects are less pronounced when the grain size decreases. It is also interesting to notice (Table 1) that whatever the nominal amount of dopant in the sample, for a given grain size, a close Gd enrichment in the first monolayers is observed (P6, P8, P9). Furthermore, a slight reduction of the kinetic demixing effect occurs when the cooling rate at the end of sintering is faster (P6<P8). When one considers the results in Table 1 and the ionic conductivity values reported in Figs. 4 and 5, it seems reasonable to conclude that the ionic conductivity of these samples, mainly controlled by the grain boundaries in the temperature range investigated, is affected by the segregation of Gd at the periphery of the grains. In the nanocrystalline samples, the effects of kinetic demixing are strongly reduced as shown previously in Mg-doped alumina [5]. In fact, the ratio between surface area and volume is much larger in these samples, characterized by very small grain sizes. In this case, it is not possible to achieve the strong Gd segregation observed in the microcrystalline samples simply because of the large areas of the nano-grain surfaces. Therefore, the present results obtained with the ceria polycrystals doped with 10 mol% Gd2O3 show that an optimum for the ionic conductivity is obtained for a grain size small enough (300500 nm) to provide a significant

References
1. Chick LA, Pederson LR, Mopan GD, Bates JL, Thomas LE, Exarhos GJ (1990) J Mater Lett 10:6 2. Petot-Ervas G, Petot C, Raulot JM, Kusinski J, Sproule I, Graham M (2003) Role of the microstructure on the transport properties of Ydoped zirconia and Gd-doped ceria. Ionics 9:195201 3. Rizea A, Petot C, Petot-Ervas G, Graham MJ, Sproule GI (2001) Kinetic demixing and grain boundary conductivity of yttria-doped zirconia. Ionics 7:7284 4. Aoki M, Chiang YM, Kosaki I, Lee LJ, Tuller H, Liu Y (1996) Solute segregation and grain boundary impedance in high-purity stabilized zirconia. J Am Ceram Soc 79:11691180 5. Monceau D, Petot C, Petot-Ervas G, Fraser JW, Graham MJ, Sproule JI (1995) J Eur Ceram Soc 15:851858 6. Rizea A, Petot C, Petot-Ervas G, Abrudeanu M, Graham MJ, Sproule GI (2007) Transport properties of yttrium-doped zirconia influence of kinetic demixing. Solid State Ion 117:34173424

S-ar putea să vă placă și