You are on page 1of 7

Pergamon

NanoStructured Materials, Vol. 11, No. 1, pp. 125131, 1999 Elsevier Science Ltd Copyright 1999 Acta Metallurgica Inc. Printed in the USA. All rights reserved. 0965-9773/99/$see front matter

PII S0965-9773(99)00025-2

MECHANOCHEMICAL SYNTHESIS OF GADOLINIUM OXIDE NANOPARTICLES


T. Tsuzuki, E. Pirault and P.G. McCormick
Special Research Centre for Advanced Mineral and Materials Processing, The University of Western Australia, Nedlands, Perth, WA 6907, Australia (Received July 31, 1998) (Accepted December 21, 1998)
AbstractSynthesis of ultrafine Gd2O3 powder by mechanochemical reaction and subsequent calcination was studied using X-ray diffraction measurements, transmission electron microscopy, surface area measurements and thermal analysis. Dried GdCl3 and NaOH powders were mechanically milled together and subsequently calcined at various temperatures. A solid-state displacement reaction GdCl3 3NaOH 3 Gd(OH)3 3NaCl was induced during milling. Calcining of the as-milled powder at 500C led to the formation of Gd2O3 particles of 1 m having a cubic structure. Adding sufficient amount of NaCl diluent in the starting powder resulted in the formation of ultrafine Gd2O3 particles of 20 nm in size. 1999 Acta Metallurgica Inc.

1. Introduction Rare earth oxides have received much attention due to their many uses such as nuclear, electronic, laser, optical, catalyst and phosphor materials. For example, Gd2O3 is used as a catalyst for dimarization of many organic compounds (1), a neutron converter in imaging plate neutron detectors (2), an additive in UO2 fuel rods for nuclear reactors (3). Gd2O3 also has applications as additives in ceramics such as ZrO2, SiC, Si3N4 to enhance sintering properties (4 6). Ultrafine powder is important for many of these applications, because of improved sintering properties achieving high-density and low sintering temperature, as well as catalytic properties due to high-surface area. Many methods have been investigated to produce ultrafine rare earth oxide powders. Thermal decomposition of suitable precursors such as hydroxides is often used for producing highly porous catalyst powders for high surface reactivity (7). However, the minimum particle size obtained with this method is limited by aggregation during thermal decomposition (8). Mechanochemical processing has been recently applied to the synthesis of a wide range of nanocrystalline and nanoparticulate materials (9,10). Ding et al. (1117) have reported the synthesis of nanoparticles of a number of transition metals and ceramics, such as Al2O3, ZrO2, Fe2O3 and ZnS, by a novel method involving the mechanical activation of solid-state displacement reactions. Milling of precursor powders leads to the formation of a nanoscale composite structure of the starting materials which react during milling or subsequent heat treatment to form a mixture of separated nanocrystals of the desired phase within a soluble salt matrix. For example, ultrafine ZrO2 powder was synthesized by milling and subsequent heat treatment of a mixture of ZrCl4 and CaO powder. The displacement reaction of ZrCl4 2CaO 3 ZrO2 2CaCl2 was induced in a steady-state manner, forming ZrO2 nanoparticles within a CaCl2 matrix. Removal of the CaCl2 by-product with a simple washing process resulted in separated ZrO2 particles of 5 nm in size (15).
125

126

GADOLINIUM OXIDE NANOPARTICLES

Vol. 11, No. 1

The synthesis of Gd2O3 by mechanical milling and heat treatment via the reaction of 2GdCl3 3CaO 3 Gd2O3 3CaCl2 has been reported (18). Only large particles of 100 nm were obtained due to the complicated reaction route. GdOCl was formed during milling and the reaction 6GdOCl 2Ca4OCl6 CaO 3 3Gd2O3 9CaCl2 occurred only during heat treatment at high temperature. Furthermore, this type of reaction between a rare earth chloride and an alkali-earth oxide can not yield a rare earth oxide of La, Ce, Pr, Nd and Sm, because the free energy change in the reaction 2ROCl CaO 3 R2O3 CaCl2 (R rare earth) is positive for light-weight lanthanides such as La, Nd and Sm, while it is negative for Gd and the heavier lanthanides. Therefore, an alternative route for synthesizing rare earth oxide nanoparticles is needed. In this paper, we report the synthesis of Gd2O3 nanoparticles via mechanochemical synthesis of Gd(OH)3 and subsequent calcination. 2. Experimental Procedures The starting materials were anhydrous GdCl3 powder (Cerac, 99.9%, 20 mesh), NaOH (Aldrich, 99.99%, pellets), and NaCl (Sigma, 99.8%, beads). GdCl3 and NaCl were dried under vacuum at 200C prior to use. A stoichiometric mixture of starting powders was sealed in a hardened steel vial with steel balls of 6.4 mm in diameter, under a high-purity argon atmosphere. Milling was performed with a Spex 8000 mixer/mill for 24 hours using a ball to powder mass ratio of 10: 1. In order to detect a possible combustion event (19,20), the surface temperature of the vial was measured during milling using a thermocouple attached to the outside surface of the vial. Removal of the NaCl by-product was carried out by washing the powder with de-ionized water, using an ultrasonic bath and a centrifuge. The washed powder was dried in an oven (60C). Calcining of the as-milled powder was carried out under an air atmosphere in alumina crucibles for 1 hour. The structure of the powder was examined under an Ar-gas atmosphere at room temperature with a Siemens D5000 X-ray diffractometer with Cu-K radiation. The mean crystallite size of the powder was estimated from diffraction peak widths using the Scherrer formula (21). The microstructure of the powder was studied using a Philips 430 transmission electron microscope (TEM) and a JEOL 2000 FX TEM equipped with a Link energy-dispersive-spectroscopy (EDS) system. For TEM studies, the washed powder was dispersed in methanol using an ultrasonic bath and a drop of solution was placed on a copper grid coated with holey carbon film. Simultaneous differential thermal analysis (DTA) and thermogravimetric analysis (TGA) were carried out using a Rigaku Thermoflex thermal analysis system under constant Ar-gas flow of 2 cc/min with a heating rate of 20C/min. BET surface area was measured using a Micromeritics Gemini 2360 Surface Area Analyser.

3. Results and Discussion Two milling stoichiometries were investigated, corresponding to the following reaction equations: GdCl3 GdCl3 3NaOH 3 Gd(OH)3 3NaCl 14NaCl (1) (2)

3NaOH

11NaCl 3 Gd(OH)3

In the equation (1), the volume ratio of NaCl to Gd(OH)3 in the product phase is 2.2: 1. The second equation was chosen so that the volume ratio of NaCl to Gd(OH)3 in the product phase is 10: 1. Since no thermodynamic data for Gd(OH)3 is available, the free energy change of the reactions was not calculated.

Vol. 11, No. 1

GADOLINIUM OXIDE NANOPARTICLES

127

Figure 1. X-ray diffraction patterns of the GdCl3 and (c) calcined at 500C.

3NaOH mixture powder; (a) the as-milled powder, (b) calcined at 350C,

3.1. GdCl3

3 NaOH

Figure 1(a) shows the XRD patterns of the GdCl3 3NaOH powder mixture after milling for 24 hours. The diffraction pattern consisted primarily of peaks associated with NaCl, Gd(OH)3 and minor peaks associated with the -Gd2O3 phase were also present. Since no combustion was detected during milling, the solid-state displacement reaction (1) occurred in a steady state manner during milling, forming Gd(OH)3 and NaCl. The cubic -Gd2O3 phase should have derived from dehydration of Gd(OH)3, suggesting occurrence of mechanical decomposition. A dehydration process during high-energy ball milling has also been reported for -AlOOH by Tonejc et al. (22). The mean crystallite sizes estimated from diffraction peak width were 16 nm, 40 nm and 100 nm for Gd(OH)3, NaCl and Gd2O3, respectively. In order to determine the temperature required to dehydrate the synthesized Gd(OH)3, simultaneous TG/DTA measurements were carried out. Figure 2 shows the DTA and TGA curves for the as-milled powder. Two steps of weight loss around 300C and 450C were evident in the TGA curve. The DTA curve has two endothermic peaks around the same temperatures. Hence, the as-milled powder was calcined at 350C and 500C. Figures 1(b) and (c) show the XRD patterns of the calcined powders. After calcining the as-milled powder at 350C, the diffraction pattern comprised peaks associated with NaCl, -Gd2O3 and GdOOH, whereas after calcining at 500C only NaCl and -Gd2O3 phases were evident in the XRD pattern. This suggests that the thermal decomposition occurred in two stages:

Figure 2. TG/DTA curves of the as-milled GdCl3

3NaOH mixture powder.

128

GADOLINIUM OXIDE NANOPARTICLES

Vol. 11, No. 1

Figure 3. TEM images of the Gd2O3 powder calcined at 500C and subsequently washed; (a) bright field image, and (b) dark field image.

Gd(OH)3 3 GdOOH GdOOH 3 0.5Gd2O3

H2O 0.5H2O

(3) (4)

at 300C and 450C, respectively. In fact, the weight loss of the reaction at 300C was about twice as large as at 450C, consistent with the two-step decomposition equations [3] and [4]. The Gd2O3 crystallite size estimated from diffraction peak widths was 100 nm. The powder calcined at 500C and subsequently washed to remove NaCl was examined by TEM. It was found that the powder consisted of particles with sizes between 0.1 and 1 m. Figures 3(a) and (b) show the TEM bright field and dark field images, respectively, of a typical Gd2O3 particle. It is evident that the particle had a porous structure composed of large crystallites. This morphology is due to the low volume ratio of NaCl to Gd(OH)3. The volume of NaCl was not sufficient enough to form separate Gd(OH)3 particles during milling, resulting in an interconnected network of Gd(OH)3 after thermal decomposition. It is well known that calcined powder of hydrates such as -Al2O3 made from Al(OH)3 can have very high surface areas ( 300 m2/g) regardless of micrometer size particles, due to the formation of intercrystalline pores between small crystallites ( 10 nm) and intracrystalline pores during dehydration (23,24). However, BET specific surface area of the Gd2O3 powder obtained in the present work was only 17 m2/g which is significantly lower for a powder made from the calcination of a hydroxide. This surface area gives a BET particle size of 40 nm, nearly the same dimension as the Gd(OH)3 crystallite size before dehydration. Therefore, this low BET surface area was due to the large crystalline sizes and lack of intracrystalline pores. 3.2. GdCl3 3NaOH 11NaCl

Figure 4(a) shows the XRD pattern of the as-milled powder. The pattern consisted of peaks associated with NaCl and Gd(OH)3. Contrary to the case of the milling without a NaCl diluent, Gd2O3 phase was not evident in the pattern. Therefore, mechanical dehydration of Gd(OH)3 did not occur during milling, possibly because the milling energy was reduced by NaCl diluent. The crystallite size of Gd(OH)3 estimated from peak width was 20 nm. After calcining the as-milled powder at 500C, the XRD pattern consisted of peaks corresponding to NaCl and -Gd2O3 phases (Fig. 4(b)), suggesting that the thermal decomposition of Gd(OH)3 was completed. After washing the calcined powder, only the Gd2O3 phase was seen in the XRD pattern (Fig. 4(c)). The mean crystallite size estimated from peak width was 30

Vol. 11, No. 1

GADOLINIUM OXIDE NANOPARTICLES

129

Figure 4. X-ray diffraction patterns of the GdCl3 500C, and (c) washed after calcining.

3NaOH

11NaCl mixture powder; (a) As-milled powder, (b) calcined at

nm, which is nearly the same order as that of Gd(OH)3 before calcining. Therefore, the Gd(OH)3 particles were well isolated during thermal decomposition due to the sufficient amount of NaCl diluent in the starting powder. This is consistent with the previous study of the synthesis of ZnS (17) in that separated single crystal nanoparticles were formed only in sample milled with sufficient volume fraction of diluent. Figure 5 shows a TEM image of the powder calcined and subsequently washed. Particles with sizes between 5 to 50 nm are evident in the figure. Using dark field imaging, each particle was found to be a single crystal. Only weak agglomeration of the particles was evident in the powder, indicating that the individual particles are separable from each other. The particles have rather irregular shapes due to the decomposition process. No trace of Na or Cl was detected by EDS. The size distribution of Gd2O3 particles estimated from TEM micrographs is shown in Fig. 6. It is evident that the mean particle size is 20 nm. Mazdiyasni et al. (25) have developed a dynamic calcination technique to prevent the aggregation during thermal decomposition of Gd(OH)3, and have obtained Gd2O3 particles of 28 nm in size with a size distribution of 1290 nm. The particles obtained in this study have a smaller mean particle size and a narrower size distribution than those obtained the above dynamic calcination technique.

Figure 5. TEM image of the Gd2O3 powder obtained from GdCl3

3NaOH

11NaCl mixture powder.

130

GADOLINIUM OXIDE NANOPARTICLES

Vol. 11, No. 1

Figure 6. Size distribution of Gd2O3 particles.

The BET surface area of the powder calcined and subsequently washed was 43 m2/g. This value gives a BET particle size of 18 nm, which is in reasonable agreement with the results of XRD and TEM measurements. The BET measurements support the results of XRD and TEM measurements that the particles are single crystallites and separable from each other. 4. Conclusions A solid-state displacement reaction between GdCl3 and NaOH was induced by mechanochemical processing, forming Gd(OH)3 in a NaCl matrix. After calcining the as-milled powder, ultrafine Gd2O3 powder was obtained. NaCl by-product/diluent was removed by a simple washing process. Without addition of diluent, calcining of the mechanochemically synthesized Gd(OH)3 resulted in the formation of porous Gd2O3 particles of 0.11 m in size. Milling with sufficient amount of diluent led to the formation of Gd2O3 nanoparticles with sizes smaller than 50 nm. Therefore, mechanochemically formed hydroxide precursors were well isolated in the NaCl matrix which prevented the particles from agglomerating during thermal decomposition. This synthesis method is suitable for producing a large quantity of ultrafine rare earth oxide powders. Moreover, this method is also applicable to the production of ternary rare earth oxides ultrafine powders such as Gd2O3 doped CeO2. References
1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. G. A. M. Hussein, J. Phys. Chem. 98, 9657 (1994). K. Takahashi, S. Tazaki, J. Miyahara, Y. Karasawa, and N. Niimura, Nucl. Instrum. Methods Phys. Res. A377, 119 (1996). G. Gunduz and I. Uslu, J. Nucl. Mater. 231, 113 (1996). S. Bhattacharyya and D. C. Agrawal, J. Mater. Sci. 30, 1495 (1995). Z. Chen, J. Am. Ceram. Soc. 79, 530 (1996). C. M. Wang, X. Pan, M. J. Hoffmann, R. M. Cannon, and M. Ruehle, J. Am. Ceram. Soc. 79, 788 (1996). G. A. M. Hussein, J. Anal. Appl. Pyrol. 37, 111 (1996). N. B. Kirk and J. V. Wood, J. Mater. Sci. 30, 2171 (1995). P. G. McCormick, J. Ding, H. Yang, and T. Tsuzuki, Mater. Res. 96, IMMA 1, 85 (1996). G. B. Schaffer and P. G. McCormick, Mater. Forum. 16, 91 (1992). J. Ding, W. F. Miao, P. G. McCormick, and R. Street, Appl. Phys. Lett. 67, 3804 (1995). J. Ding, T. Tsuzuki, P. G. McCormick, and R. Street, J. Alloys Compounds. 234, L1 (1996). J. Ding, T. Tsuzuki, P. G. McCormick, and R. Street, J. Phys. D Appl. Phys. 29, 2365 (1996). J. Ding, T. Tsuzuki, and P. G. McCormick, J. Am. Ceram. Soc. 79, 2956 (1996). J. Ding, T. Tsuzuki, and P. G. McCormick, Nanostruct. Mater. 8, 75 (1997). J. Ding, T. Tsuzuki, and P. G. McCormick, Nanostruct. Mater. 8, 739 (1997). T. Tsuzuki, J. Ding, and P. G. McCormick, Phys. B. 239, 378 (1997). T. Tsuzuki, W. T. A. Harrison, and P. G. McCormick, J. Alloys Compounds, submitted for publication. G. B. Schaffer and P. G. McCormick, Metallurg. Trans. A22, 3019 (1991). H. Yang and P. G. McCormick, J. Solid State Chem. 107, 258 (1993).

Vol. 11, No. 1

GADOLINIUM OXIDE NANOPARTICLES

131

21. B. Cullity, Elements of X-ray Diffraction, 2nd edn., Addison-Wesley, Redding, MA (1978). 22. A. Tonejc, C. Kosanovic, M. Stubicar, A. M. Tonejc, B. Subotic, and I. Smit, J. Alloys Compounds. 204, L1 (1994). 23. Z. C. Kang and L. Eyring, J. Solid State Chem. 88, 303 (1990). 24. B. C. Lippens and J. J. Steggerds, Physical and Chemical Aspects of Adsorbents and Catalysts, p. 171 (1970). 25. K. S. Mazdiyasni and L. M. Brown, J. Am. Ceram. Soc. 54, 479 (1971).

You might also like