You are on page 1of 5

Solid State Sciences 11 (2009) 671675

Contents lists available at ScienceDirect

Solid State Sciences


journal homepage: www.elsevier.com/locate/ssscie

Hydrothermal processing and characterization of Ce1xPbxO2d solid solutions


Guofang Zhang a, b, Liping Li a, *, Guangshe Li a, Xiaoqing Qiu a, Guofeng Yan a, c
a

State Key Structural Chemistry Laboratory, Fujian Institute of Research on the Structure of Matter, 155 Yangqiao West Road, Fuzhou 350002, PR China Rare Earth school, Inner Mongolia University of Science and Technology, Baotou 014010, PR China c Research Institute of Photocatalysis, Fuzhou University, Fuzhou 350002, PR China
b

a r t i c l e i n f o
Article history: Received 25 April 2008 Received in revised form 1 October 2008 Accepted 28 October 2008 Available online 5 November 2008 Keywords: Hydrothermal Buffer solution Solid solubility Oxygen storage capacity

a b s t r a c t
Nanocrystals of Ce1xPbxO2d (x 0.00, 0.05, 0.10, 0.15, 0.20, 0.25, 0.30, and 0.35) were prepared by a hydrothermal reaction route. During the formation reaction, buffer solutions were explored as an effective additive to retain the initial molar ratio. With increasing the Pb2 content, the average crystallite size was slightly retarded. Morphologies observed by transmission electron microscope indicated that the particles were spherical-like and highly uniformed. Pb2 ions are homogenously distributed in the solid solutions. Analyses using X-ray diffraction, Raman and UV spectroscopies showed that the solid solubility limit of Pb2 in CeO2 was about x 0.20. For x < 0.20, with increasing the Pb2 content, the bulk conductivity increased, and the oxygen storage capacity was enhanced as followed by a decrease in reduction temperature. 2008 Elsevier Masson SAS. All rights reserved.

1. Introduction Cerium dioxide (CeO2) is one of the most important materials that present a great deal of applications such as electrolyte materials in solid oxide fuel cells [1], gas sensors [2], ultraviolet blocking materials [3], and high temperature oxidation resistant coatings [4]. Recently, very interesting phenomena have been reported on nanoceria for catalytic uses due to the unique valence characteristics and the presence of intrinsic oxygen vacancies in the CeO2 lattice. It is well known that chemical doping can greatly inuence the number of vacancies in lattice and further alter the properties [5]. It has been proved that the lattice of CeO2 can be modied by incorporating metal ions with smaller radius and lower valence such as Fe, Cu, Mn [68], or larger lanthanide elements with similar properties to Ce [911]. However, the main group elements are barely successfully doped into CeO2 to form solid solutions, because most of these elements like Pb belong to the fourth main group (IVA) and have bigger ionic radius than Ce and mixed valence states, though such doping is expected to improve remarkably the performance of CeO2. The preparation of solid solutions Ce1xPbxO2d is a challenging issue to chemists and engineers of materials. Owing to a large lattice mismatch between Pb oxides and CeO2, the solubility limit of Pb in CeO2 is still unclear. It is reported that the solid solubility of PbO in CeO2 is about 9 mol% by ring two

compounds at 800  C [12]. However, high temperature calcinations could result in severe environmental pollution of lead or non-stoichiometry in the nal products, which further limits the materials design and property tailoring for technological applications. Hydrothermal processing has attracted a lot of attention since the desired particle size and shape can be precisely modied when experimental parameters like pH of the reaction solution, temperature and time, concentration and the type of solvents, are carefully controlled [13]. Thus, in this paper, we reported on the preparation of nanocrystalline Ce1xPbxO2d solid solutions using a low temperature hydrothermal process. The structures of the asprepared Ce1xPbxO2d solid solutions were systematically analyzed by X-ray diffraction, scanning electron microscope, transmission electron microscope, ultraviolet-visible and Raman spectroscopy. Pb2 ions were found to be distributed in the samples by substituting Ce4. We also studied the properties including lattice dimension, solid solubility limit, band-gap energy, and electrical properties as a function of Pb2 content.

2. Experimental procedure Pure and Pb doped CeO2 were synthesized by a hydrothermal process, in which chemicals Ce(NO3)3$6H2O (A.R.), Pb(NO3)2 (A.R.), NH4OH (25%, A.R.), NH4NO3 (A.R.) were used as the starting materials. The synthetic process can be described as follows: rstly, solutions of 0.3 mol/L Ce(NO3)3 and 0.3 mol/L Pb(NO3)2 with Ce:Pb molar ratios from 95:5 to 65:35 were fully mixed. Then, NH4OH solution was dropped into the mixed solution to adjust the pH

* Corresponding author. Tel.: 86 591 83792846; fax: 86 591 83714946. E-mail address: lipingli@fjirsm.ac.cn (L. Li). 1293-2558/$ see front matter 2008 Elsevier Masson SAS. All rights reserved. doi:10.1016/j.solidstatesciences.2008.10.013

672

G. Zhang et al. / Solid State Sciences 11 (2009) 671675

value to 9.5. In order to retain the Pb content in the nal products as to the initial ratio, NH4NO3 was added to the suspension to form a buffer solution. The pH value was kept constant during the chemical reactions. The as-obtained suspension was transferred to Teon-lined stainless steel autoclaves, sealed and thermally treated at 200  C for 24 h. When the formation reaction was completed, autoclaves were cooled naturally to room temperature. The products were harvested by pressure ltration, washed fully with distilled water, and dried at 80  C for 12 h. No Pb ions were detected in the supernatant solution by chemical method with AgCl after the formation reactions. Phases of the samples were identied by X-ray diffraction (Rigaku, DAMX2500, Japan) using Cu Ka radiation at room temperature. The average particle size, D, was calculated from the diffraction peak (111) using Scherrer formula: D 0.9l/(b cos q), where l is the X-ray wavelength, b is the half-width of the (111) peak, and q is the diffraction angle of the (111) peak. The lattice parameters were rened by least-squares methods using the Rietica program with nickel as the internal standard for peak position calibrations. Morphologies and particle sizes were observed by scanning electron microscopy (SEM) in a JSM6700 electron microscope equipped with an energy-dispersive spectroscopy system (EDS) and transmission electron microscopy (TEM) with a JEM-2010 electron microscope. The samples for SEM and TEM observation were dispersed in ethanol followed by ultrasonic vibration and then placed a drop of the dispersion onto a copper grid coated with a layer of amorphous carbon. Raman spectra of the samples were recorded at room temperature using a Renishaw1000 unpolarized Raman spectrometer with a He ions laser in the region of 2001000 cm1. The excitation wavelength is 514.5 nm and output powder is 6.0 mW. Diffuse reectance UV visible (UVvis) spectra of the samples were recorded on lambda 900 UVvisNIR spectrophotometer (Perkin Elmer). BaSO4 was used as a reference material. Ac-impedance spectra of the samples were measured using Agilent 4284A LCR analyzer. The samples were extruded to 4 7 mm 2 mm tablet with Ag as the electrodes. TPR experiments were evaluated with a Micromeritics TPD/TPR 2900 Chemisorption Analyzer. The experiments were performed as follows: a weight amount of the sample (w10 mg) was placed in a quartz reactor, pretreated in a ow of Ar (99%) gas at 773 K for 1 h (ramp rate of 10 K min1), and cooled to room temperature. A gas mixture of H2 (10%)Ar (90%) was then passed (25 mL min1) through the reactor. The temperature was raised up to 973 K at a heating rate of 10 K min1. A thermal conductivity detector was employed at the outlet of the reactor to measure the volume of hydrogen consumed during the reduction of the samples.

Relative intensity (arb.units.)

x=1.00

x=0.35 x=0.30 x=0.25 x=0.20 x=0.15 x=0.10 x=0.05 x=0.00


20 30

2 (degree)

40

50

60

Fig. 1. XRD patterns of Ce1xPbxO2d (x 0.00, 0.05, 0.10, 0.15, 0.20, 0.25, 0.30, 0.35, and 1.00). The diffraction peaks denoted by * represent the internal standard of Ni.

Fig. 2 shows high resolution-transmission electron microscopy (HRTEM) images of a typical sample at x 0.20. The sample consisted of aggregated nanoparticles. All particles are single crystals of a nearly spherical morphology with an average diameter of 5 nm, which is in good agreement with the estimated values from XRD. For the formation of solid solution Ce1xPbxO2d in pure phase, pH control is crucial. As well known, the pH value of solutions would be altered during hydrothermal process. The variation of pH value might result in Pb loss because PbOx oxides can be dissolved in either strong acidic solution or alkaline solution. In our experiment, we use buffer solution to keep the pH at a stable range, by which Pb ions could be precipitated completely for the formation reactions. Our elemental analysis showed that the atomic ratio of Pb to Ce is much closer to the initial one. As a result, buffer solution used in the preparation conditions plays a key role in formation

3. Results and discussion 3.1. X-ray diffraction analysis Fig. 1 shows XRD patterns of Ce1xPbxO2d (x 0.000.35) samples. All samples adopted a standard uorite cubic structure and no other impurity phases were detected. XRD pattern for x 1.00 prepared under the same hydrothermal condition is also shown in Fig. 1 for comparison. It is found that the product for x 1.00 is mainly composed of Pb3O4 and PbO phases with Pb ions in 2 and 4 oxidation states. With increasing the Pb content, all diffraction peaks became broadened, which is associated with poor crystallinity and ne nature of the samples. The average crystallite size calculation using Scherrer formula for (111) peak indicates that the crystallinte size decreased from 12 to 5 nm when Pb content increased from x 0.00 to 0.35. Therefore, the incorporation of Pb ions kinetically prevented the grain growth of the particles.

Fig. 2. HRTEM image of typical sample Ce1xPbxO2d at x 0.20.

G. Zhang et al. / Solid State Sciences 11 (2009) 671675

673

reaction of the solid solution Ce1xPbxO2d. Furthermore, in order to investigate the distribution of Pb ions in the as-obtained samples, both line scan and elemental mappings for x 0.2 were measured. As indicated in Fig. 3, the proles of Ce and Pb composition are similar. The ratio of Ce/Pb is almost a constant. Therefore, both elements Ce and Pb were homogeneously distributed over the whole particles. Considering that lattice parameters of the solids are extremely sensitive in the chemical composition [14], the solution limit of Pb in CeO2 nanoparticles could be determined by structural renement. Fig. 4 shows the variations of lattice parameters of Ce1xPbxO2d solid solutions as a function of Pb content. The dopant content dependence of lattice parameter for Ce1xLaxO2d solid solutions [11] is also shown for comparison. It is seen that a-axis increases linearly with Pb content to x 0.20, while beyond x 0.20, a-axis keeps almost constant within the experimental errors, which suggested that the solubility limit of Pb in CeO2 is about x 0.20. The variation of lattice parameter as a function of Pb content seems to be consistent with the relative ionic size of Pb2 to Ce4, because ionic radius of Pb2 in 8-coordination is 0.129 nm which is much larger than that of 0.097 nm for Ce4 in the same coordination [16]. It should be noted that for Pb doped samples, the slope (Da/Dx) obtained by linearly tting the lattice parameter is only 0.011(0.001), which is much smaller than 0.030(0.002) for La doped solution. The ionic radius of La3 ions (0.116 nm in 8coordiantion [16]) is smaller than Pb2. Therefore, in addition to the doping of Pb2, there should be other possibilities that account for the variation of lattice parameter of Ce1xPbxO2d. Firstly, signicant size effect is expected to be a major cause since the particle sizes of the samples are only of several nanometers which may give rise to an apparent lattice expansion [15]. Taking the size effect into consideration, the lattice parameter for the doped samples is still larger than the pure CeO2 with the same particle size. We may have to consider the mixed valence characteristic of Pb ions in the samples. In oxides, Pb ions can appear in 2 and/or 4. Because Pb4 is smaller with an ionic radius of 0.094 nm, incorporation of Pb4 into CeO2 would lead to a slight lattice contraction. Therefore, the slightly increased lattice parameter with Pb content below the solubility limit should be related to the mixed valences of Pb2 and Pb4.

Pb doped 0.548 La doped

0.546

(a/x)La=0.030(4)

a (nm)
0.544

(a/x)Pb=0.011(1) 0.542

0.0

0.1

0.2

0.3

0.4

Pb content x
Fig. 4. Dopant content dependence of lattice parameter for the solid solution Ce1xPbxO2d (C). The data for Ce1xLaxO2d (>) from Ref [11] is shown for comparison.

materials. Any changes in lattice parameters and chemical environment can give signicant shifts in the band frequencies [11]. Here, Raman spectroscopy was used to study the microstructure of Ce1xPbxO2d solid solutions. Fig. 5 shows the room temperature Raman spectra of pure CeO2 and doped samples at given Pb concentration. An intense Raman peak located at 462 cm1 was observed for pure CeO2, which is assigned to the triply degenerated Raman-active mode (F2g) [11]. With increasing the Pb content, this Raman mode showed a downshift toward 452 cm1 for x 0.20, and then kept at the same value when Pb content is larger than x 0.20. This result further conrms the solubility limit of x 0.20 for Ce1xPbxO2d. Meanwhile, the full width at half maximum (FWHM) of this Raman mode was increased with x, which might be related to the lattice distortion produced by incorporation of Pb ions and the reduction of particle size. In addition to the F2g mode, a weak broad Raman peak was also observed in the 500600 cm1 region, which is ascribed to the oxygen vacancies [17]. The increase of its intensity

3.2. Raman analysis Raman scattering is an excellent, nondestructive, and rapid analysis technique for investigating the phonon spectra of

x=0.30

Intensity (arb. unit)

x=0.20

X=0.10

x=0.0

200

400

600

800

1000

Raman shift (cm-1)


Fig. 3. Line scan of elements Ce and Pb in Ce1xPbxO2d at x 0.20 by EDS analysis. Fig. 5. Raman spectra of the samples at the given Pb contents.

674

G. Zhang et al. / Solid State Sciences 11 (2009) 671675

with dopant content is consistent with the Pb2 doping. All these observations provide direct evidence that the Pb ions were incorporated in the CeO2 lattice. 3.3. UVvis spectroscopy Lattice modication of CeO2 by Pb incorporation also affects the band structure. CeO2 is a wide-band semiconductor [18] with a band-gap energy of about 3.3 eV for direct transition [19]. The band gaps of solid solutions Ce1xPbxO2d were investigated by measuring UV absorption edge. Fig. 6(a) shows the UVvis diffuse spectra of Ce1xPbxO2d samples. All samples showed the similar UV absorption spectra as the pure CeO2. Peaks located at 260 and 320 nm are governed with partially resolved O2 / Ce4 charge transfer process. The former one is a reectance band arising from the surface sites, while the latter one is a diffuse reectance band that originated from the bulk CeO2 [8]. It is noted that no absorption associated with Pb ions was observed, which can be understood by the main group element behavior (with no d electron transition). With increasing the Pb content, the absorption edge gradually shifted toward higher wavelength. When the Pb content is beyond the solution limit of x 0.20, the absorption edge did not show any apparent change. Fig. 6(b) exhibits the dopant content dependence of the bandgap energies. The band-gap energy for CeO2 nanocrystals is 3.34 eV.

Incorporation of Pb ions into CeO2 resulted in a red shift of band gap. This observation can be understood in terms of the lattice modication of CeO2 from the Pb doping. As indicated in Fig. 4, the incorporation of Pb2 ions led to a lattice expansion, which gave rise to a negative pressure and further a band-gap narrowing. Comparatively, the contribution of size effect can be ignored, since small particle size would result in a blue shift in the ultraviolet absorption [2022]. 3.4. Transport conductivity properties Room-temperature conductivity of the samples was determined by ac complex impedance technique. Fig. 7(a) illustrates a typical impedance plot for x 0.10 in the frequency range of 201 106 Hz. Other doped samples showed similar plots. The impedance plot for x 0.10 consisted of three parts: a small semi-circle in the high frequencies range is attributed to the bulk conduction, the middle part is related to the grain boundary conduction, while the linear part at low frequency is ascribed to the electrode exchange process. With increasing the Pb content, bulk conductivity was increased (Fig. 7b). It should be mentioned that the conductivities of solid solutions Ce1xPbxO2d are governed by many factors, such as concentration of electron/vacancy, the mobility of oxygen ions, and surface absorbed water molecules. Among these factors, the conduction of oxygen ions should play a dominant role as it is usually dependent on the concentration of oxygen vacancies. The incorporation of Pb ions in CeO2 lattice resulted in an increase in oxygen vacancies and therefore enhanced the bulk conductivity. When Pb content is larger than x 0.20, part of Pb ions may locate in the interstitial sites or on the surfaces. In this case, the concentration of the trap centers and barriers could be increased to retard the carrier conduction, which results in a decrease in conductivity. On the other hand, for heavy doping, oxygen vacancies would bound to cation defects to form defect associations, such as  fPb00 VO g, which could also reduce the concentration of free Ce oxygen vacancies for conduction [23].

15

12

FR(%)

x=0.20 x=0.10 x=0.05 x=0.00

a
-Z'' ( M)

0 300 400 500 600

Wavelength (nm)

grain

b
3.3

0 0

bulk 3 6 9 12

Z' ( M)

b
Egap (eV)
3.2

10 8 6 4 2 0 -2

3.1

bulk(10-6-1cm-1)

0.00

0.05

0.10

0.15

0.20

0.25

0.00

0.05

0.10

0.15

0.20

Pb content x
Fig. 6. (a) UVvis spectra of the solid solution Ce1xPbxO2d at given Pb content, and (b) dopant content dependence of band-gap energies.

Pb content x
Fig. 7. (a) Complex impedance spectrum of Ce1xPbxO2d at x 0.10 and (b) the variation of bulk conductivity with Pb content.

G. Zhang et al. / Solid State Sciences 11 (2009) 671675

675

4. Conclusions
area (a.u.)
0.8 0.6 0.4 0.2 0.0 0.0 0.1 0.2 0.3

x=0.35 x=0.20 x=0.15 x=0.10 x=0.05 x=0.00

100

200

300

400

500

600

Temperature ( C)
Fig. 8. TPR proles of solid solution Ce1xPbxO2d at given Pb content. Inset shows the relationship between the peak area and Pb content.

Nanosized Ce1xPbxO2d solid solutions were prepared by a hydrothermal method, in which buffer solution was used to promote the formation reaction and maintain the molar ratio of Pb to Ce. The solid solutions had a cubic uorite structure with high sample uniformity. Pb ions were homogenously distributed in the lattice of CeO2. The solid solubility of Pb in CeO2 was determined to be x 0.20, which was much higher than other methods reported previously. Substitution of Ce by Pb slightly retarded the grain growth of the nanoparticles, and gave rise to a lattice expansion. Within the solid solubility, the expansion slope of lattice parameter was only 0.011(1), which is related to the mixed valence state of Pb2 and Pb4. With increasing the Pb content, the Raman peak of solid solution gradually shifted toward lower energies, which is accompanied by a broadened width of the phonon modes. The incorporation of Pb in CeO2 lattice also shifted the band-gap energy toward longer wavelength. Because of the presence of oxygen vacancies, the bulk conductivity gradually increased and showed a maximum at x 0.15. The reducibility of the solid solutions is widely modied by Pb doping. Cerium lead mixed oxides are much more reducible than pure CeO2. Acknowledgments This work was nancially supported by NSFC under the contract (No. 20831004, 20671092, 20771101), National Basic Research Program of China (973 program, No. 2009CB939801), FJIRSM (No. SZD-08002-3). References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] K. Eguchi, T. Inoue, H. Arai, Solid State Ionics 52 (1992) 165. P. Jasinski, T. Suzuki, H.U. Anderson, Sens. Actuators, B 95 (2003) 73. T. Masui, K. Fujiwara, K. Machida, G. Adachi, Chem. Mater. 9 (1997) 2197. S. Patil, S.C. Kuiry, S. Seal, Proc. R. Soc. A 460 (2004) 3569. G. Sanon, R. Rup, A. Mansingh, Phys. Rev. B 44 (1991) 5672. G.S. Li, R.L. Smith, H. Inomata, J. Am. Chem. Soc. 123 (2001) 11091. W.J. Shan, W.J. Shen, C. Li, Chem. Mater. 15 (2003) 4761. B. Murugan, A.V. Ramaswamy, Chem. Mater. 17 (2005) 3983. W. Huang, P. Shuk, M. Greenblatt, Chem. Mater. 9 (1997) 2240. M.J. Godinho, R.F. Gonalves, L.P. S Santos, J.A. Varela, E. Longo, E.R. Leite, Mater. Lett. 61 (2007) 1904. S. Patil, S. Seal, Y. Guo, Appl. Phys. Lett. 88 (2006) 243110. M. Hrova, A. Benc, J. Holca, M. Koseca, Mater. Res. Bull. 36 (2001) 767. K. Yamashita, K.V. Ramanuhachary, M. Greenblatt, Solid State Ionics 81 (1995) 53. R.X. Li, S. Yabe, M. Yamashita, S. Momome, S. Yoshida, S. Yin, T. Sato, Solid State Ionics 151 (2002) 235. S. Deshpande, S. Patil, S.V.N.T. Kuchibhatla, S. Seal, Appl. Phys. Lett. 87 (2005) 133113. A. Trovareli, Catal. Rev. 38 (1996) 439. J.R. Mcbride, K.C. Hass, B.D. Poindexter, W.H. Weber, J. Appl. Physiol. 76 (1994) 2435. P. Patsalas, S. Logothetidis, L. Sygellou, S. Kennou, Phys. Rev. B 68 (2003) 035104. S. Tsunekawa, T. Fukuda, A. Kasuya, J. Appl. Physiol. 87 (2000) 1318. M.G.B. Nunes, L.S. Cavalcante, V. Santos, J.C. Sczancoski, M.R.M.C. Santos, L.S. Santos-Junior, E. Longo, J. Sol-Gel. Technol. 47 (2008) 38. Y. Nosaka, J. Phys. Chem. 95 (1991) 5054. S. Tsunekawa, J.T. Wang, Y. Kawazoe, J. Appl. Physiol. 94 (2003) 3654. V. Butler, C.R.A. Catlow, B.E.F. Fender, J.H. Harding, Solid State Ionics 8 (1983) 109. Y. Zhang, S. Andersson, M. Muhammed, Appl. Catal. B-Environ. 6 (1995) 325. D. Terribile, A. Trovarelli, C. Leitenburg, et al., Catal. Today 47 (1999) 133. H.C. Yao, Y.F. Yao, J. Catal. 86 (1984) 254. D.C. Ji, X.H. Qian, S.L. Liu, Chin. J. Catal. 25 (2004) 475.

3.5. Redox property of Ce1xPbxO2d samples Generally, doped ceria exhibits excellent oxygen storage capacity. For example, oxygen storage capacity of Ce1xPbxO2d prepared by coprecipitation oxalate precursors was reported to be remarkably enhanced, and the temperature of maximum reactivity was signicantly lowered compared with pure CeO2 and other doped CeO2 such as Ce1xZnxO2d and Ce1xMxO2d (M Ca, Nd) [24]. The redox behavior of our samples was investigated by TPR. As indicated by TPR proles in Fig. 8, reduction of CeO2 by H2 occurs from 250  C and reaches two maximum H2 consumption at 462 and 620  C, which correspond to the surface and bulk reductions, respectively [25,26]. For the doped samples, only one single reduction feature was detected and the bulk reduction was inhibited. Moreover, the reduction capability of Pb doped solid solution strongly depends on the Pb content. As shown in Fig. 8, all Pb doped samples exhibited much lower redox temperature than that of pure CeO2 and PbO which shows two reduction peaks at 450 and 640  C [27]. The redox temperature of Pb doped solid solution gradually decreased with Pb content. For example, for x 0.05, the redox temperature was 364  C, which was reduced to 296  C when Pb content was increased to x 0.20. Accompanying with the decrease of redox temperature, the intensity of TPR peak increased with Pb content. Inset of Fig. 8 gives the variation of peak area with Pb content. The increase in peak area suggests that the amount of H2 consumption was increased to activate a deeper reduction. By incorporation of Pb ions into CeO2 lattice, the reducibility of the solid solutions is widely modied. Cerium lead mixed oxides are much more reducible than pure CeO2. Moreover, the deeper reduction was gradually activated with increase of Pb content. In Pb doped solid solutions, both Ce4 and Pb4 are reduced simultaneously, while only Ce4 is reduced in pure CeO2 nanocrystals. As a result, the reduction temperature of the former becomes much lower than that of the latter. Since introduction of Pb2 into CeO2 lattice decreased the particle size of the samples, the reduction reaction could easily take place in the bulk.

H2 consumption ( arb.units.)

You might also like