Controlled amorphous crystallization: An easy way to make transparent nanoceramics
Lin Mei,Guang-Hua Liu,Gang He,Li-Li Wang,Jiang-Tao Li
DOI: https://doi.org/10.1016/j.optmat.2011.05.013
IF: 3.754
2012-01-01
Optical Materials
Abstract:Highlights ► New method: amorphous sintering followed by controlled crystallization (ASCC). ► Controlled grain growth was achieved by controlled crystallization. ► Transparent LaAlO 3 /t-ZrO 2 nanoceramic with an average grain size of 40 nm. ► Transmittance at 800 nm 55%, Vickers hardness 19.05 GPa, and fracture toughness 2.64 MPa m 1/2 . Abstract Fabrication of transparent nanoceramics is attracting more and more interests recently. In this study a new method of amorphous sintering followed by controlled crystallization (ASCC) was developed, and transparent LaAlO 3 /t-ZrO 2 nanoceramic was prepared as an example. Based on a eutectic composition of Al 2 O 3 –La 2 O 3 –ZrO 2 , glass powders were synthesized, sintered, and then converted to nanoceramics by post-heat-treatment. The heat-treatment performed at 1200 °C for 2 h produced a transparent LaAlO 3 /t-ZrO 2 nanoceramic with an average grain size of 40 nm. Due to the nanoscale microstructure, the ceramic showed a transparency up to 55% at 800 nm (1 mm thick), Vickers hardness of 19.05 GPa, and fracture toughness of 2.64 MPa m 1/2 , respectively. Keywords Nanoceramics Amorphous sintering Controlled crystallization Optical properties Mechanical properties 1 Introduction Transparent ceramics are receiving wide attentions because of their important applications in high-performance laser hosts instead of single crystals [1–4] . Many works have been done to tailor the microstructure, e.g. the elimination of pores and refinement of grain size, in order to reduce light scattering and increase laser excitation efficiency. It is generally considered that pore-free nanoceramics are promising to achieve excellent combination properties [1,4] . Recently, encouraging advances in optical performance of transparent ceramics in non-cubic system caused by the refinement of grain size have been reported. For example, the in-line transmittance of Al 2 O 3 ceramic greatly increases when the grain size is reduced into submicron scale, which makes translucent sample become transparent [5] . This indicates that the optical properties of the non-cubic ceramics are intensively dependent on the grain size because of their high anisotropic lattice (for α-Al 2 O 3 , c / a = 2.72) [6] . Based on the Rayleigh–Gans–Debye light-scattering theory, Apetz et al. developed a model to describe the light transmission characteristics of fine-grained non-cubic ceramics [7] . This model explains the noteworthy transparency of the submicron-grained α-Al 2 O 3 ceramic, and suggests that the light transmission will remarkably increase with the reduction of grain size. That is, if the grain size can further be reduced to nano-scale, the light transmission of transparent ceramic will be greatly improved. While in conventional sintering of ceramic powders, the high temperature required for densification would promote extensive grain growth, which makes it very difficult to prepare fully-dense nanoceramics [8] . Therefore, three main methods have peen proposed to prevent grain growth during sintering: application of hot-pressing or sintering forging under high pressure low temperature, addition of dopant to modify diffusion process, and sintering powders in metastable crystallographic phases that during sintering act as seeds for the final phases [9] . However, these methods are usually accompanied by rigorous conditions, such as ultra-high pressure, long production cycle and expensive equipments, which limit the wide applications of transparent nanoceramics. Recent researches indicate that coherent, dense, amorphous bulk alloys and glasses can be successfully densified from amorphous powders by viscous flow within the kinetic window Δ T = T g − T s , where T g is the glass transition temperature and T s is the starting crystallization temperature [10,11] . Based on these work, we develop a new method of amorphous sintering followed by controlled crystallization (ASCC), in which the amorphous powders are prepared, sintered via viscous flow, and then converted to transparent nanoceramics by heat-treatment. This approach makes densification and grain growth take place through different mechanisms, which can offer the opportunity to control two processes separately. However, only some transparent nanoglass–ceramics have been obtained from glasses by crystallization. This should be attributed to the different crystallization abilities of the multi-component mother glasses, which make it difficult to achieve a high crystallization degree up to 95% by heat-treatment. In the ASCC method, eutectic compositions are used to prepare transparent nanoceramics because of their medium glass-forming abilities [12] . In the present work, transparent LaAlO 3 -based nanoceramic was prepared by the ASCC method as an example. The glass powders with a eutectic composition were prepared, sintered and converted into transparent LaAlO 3 /t-ZrO 2 nanoceramics via a controlled crystallization process. The feasibility of this method for other systems had also been demonstrated. 2 Experimental 2.1 Preparation Raw materials used in this study included: Al 2 O 3 (purity > 99.99%, D V50 = 200 nm, Dalian Luming Nanometer Material Co., Ltd.), La 2 O 3 (purity ⩾ 99.99%, D V50 = 509 nm, China Minmetals (Beijing) Research Institute) and m-ZrO 2 (purity ⩾ 99.8, D V50 = 46 nm, Dongguan SG Ceramics Technology Co., Ltd). The powder mixtures with a molar composition of 53%Al 2 O 3 –20%La 2 O 3 –27%m–ZrO 2 [13] (named hereafter as ALZ) were dispersed as an aqueous suspension with 1.5 wt% polyvinyl alcohol (PVA, molecular weight 1700, purity ⩾ 99.0%, Beijing Chemical Works) as binder, and spray dried into spherical particles. Spray dried agglomerates were feed into a C 2 H 2 /O 2 flame at a rate of 20 g/min, and then sprayed into water to prepare the ALZ glass microspheres, where the C 2 H 2 flow rate was 20 l/min and the O 2 flow rate was 15 l/min. The flame was vertical to the water surface, and produced a combustion temperature up to 3200 °C. The particles were heated, melted and quenched into water to obtain a cooling rate of ∼10 3 °C/s. The glass microspheres deposited at the bottom of a steel container were collected, dried and sieved. For hot-pressing, 5.0 g glass powder passed through 500 mesh was loaded into a graphite die ( Φ 20 mm), the internal surface of which was covered with a graphite sheet to avoid direct contact between the powder compact and the graphite die. The sintering temperature was elevated from room temperature to the final sintering temperature (about 900 °C) at 10 °C/min, and dwelled for several minutes to achieve densification with the aid of a pressure exerted before the dwell. After cooled down, the obtained glasses were heat-treated at different temperatures ranging from 900 °C to 1300 °C to prepare the transparent nanoceramics. 2.2 Characterization Glass transition temperature ( T g ) and starting crystallization temperature ( T s ) of the glass microspheres were determined by a differential thermal analysis (DTA, NETZSCH STA 449C, Germany). Phase assemblage was identified by X-ray diffraction (XRD, D8 Focus, Bruker, Germany) using Cu Kα radiation. Microstructure was characterized by scanning electron microscopy (SEM, Hitachi S-4300, Japan), and energy dispersive spectroscopy (EDS, Shimazu 6853-H, Japan) was used to analyze the chemical composition in selected area. The in-line transmittance spectras were obtained using an UV–VIS–NIR Spectrophotometer (Cary 5000, Varian, America) for 175–2000 nm and a Fourier Transform Infrared Spectra (FTIR; Excalibur 3100, Varian, America) for 2000 nm–25 μm. All the samples were polished down to 1 mm thick. According to the mixture rule, the refractive index ( n g ) of the ALZ glass is given by [14] : (1) n = Φ 1 n 1 + Φ 2 n 2 + Φ 3 n 3 where n 1 , n 2 and n 3 are the refractive indexes of Al 2 O 3 , La 2 O 3 and ZrO 2 , respectively, Φ 1 , Φ 2 and Φ 3 are the volume fractions of the respective components described above. Therefore, the n g is 2.00. Based on these results, the theoretical transmittance ( T th ) of the ALZ glass was calculated as 80% using the relations given in the literature [7] . The sample density was measured by Archimedes method, and the theoretical density of the ALZ glass can be calculated to be 4.56 g/cm 3 by an empirical equation [15] (2) ρ th = 0.53 · ∑ ( M i · x i ) ∑ ( V i · x i ) where M i is molar weight (kg/mol), x i is molar fraction (mol%), and V i is packing density parameter (m 3 /mol) for an oxide M X O Y . Based on the density variation, the degree of crystallization ( α ) of the heated sample was evaluated as reported in the literature [16] . Vickers hardness ( H v ) and fracture toughness ( K IC ) were measured using a microhardness machine (HXD-1000TM, Shanghai), and by application of 500 gf (4.9 N) load on polished surface for 15 s. 3 Results and discussion 3.1 Preparation of ALZ glass powder In this study, the eutectic point of the ALZ system is 1665 °C [13] . Although the flame spray process provided a high temperature up to 3200 °C, only smaller spray-dried particles could be heated to a temperature above the melting point during the supersonic spray process. It was found that the ALZ glass powders with particle size below 38.5 μm (−500 mesh) were fully spherical and transparent under an optical microscope because of the higher degree of melting and quenching. Differential Thermal Analysis indicated that the kinetic window (Δ T ) of the ALZ amorphous microspheres was 816–895 °C. According to the empirical rule, the sintering temperature is generally chosen to be 20 °C lower than T s so as to avoid devitrification [17] . Therefore, the optimal sintering temperature for the ALZ system was 875 °C, which is well located in the kinetic window for sintering. 3.2 Sintering of ALZ transparent glass It is generally considered that the sintering of glass frit into bulk forms is realized by viscous flow within the kinetic window, while the densification rate is positively proportional to the applied pressure P according to the d ρ /d t = 3 P (1− ρ )/4 η , where η is the viscosity [18] . Thus, during the sintering process conducted at 875 °C, a pressure ranging from 30 MPa to 90 MPa was applied to achieve full densification as soon as possible. The shrinkage behaviors are shown in Fig. 1 . It is observed that the shrinkage process consisted of three stages: accelerating shrinkage, rapid shrinkage and shrinkage stagnation. The rapid shrinkage stage made a largest contribution up to 90% to the densification. Meanwhile, the applied pressure accelerated the shrinkage rate dramatically. When the pressure reached 90 MPa, the time for full densification decreased from 1350 s to 525 s, as seen in the inset of Fig. 1 . The shorter dwell time is well beneficial for avoiding the surface devitrification of glass microspheres, so as to aid the elimination of residual pores. SEM images in Fig. 2 also verify that the higher pressure promoted the densification degree notably. Comparing with the sample sintered at 30 MPa, the glass obtained at 90 MPa had no obvious pores on the polished surface. Meanwhile, the sample seemed to be a whole bulk, indicating that the interfaces of glass microspheres had disappeared by viscous flow. This is very different from polycrystalline ceramics sintered at high temperatures. These results suggest that the higher pressure instead of temperature is essential to sinter transparent glass. The density of the ALZ glass sintered at 90 MPa was measured as 4.55 g/cm 3 , which reached 99.8% of the theoretical value. Fig. 3 shows that the glass had a transparency in a wide wavelength range of 265 nm–6.54 μm, covering visible, near infrared and middle infrared regions. The transmittance of this sample reached 68% at 800 nm, and reached the maximum value of 75% in the range of 2.4–4.4 μm. Compared with the theoretical transmittance (80%), the loss should be attributed to the light scattering because of residual pores, minor translucent and opaque glass microspheres. X-ray diffraction analysis (XRD) in Fig. 4 indicates that the sintered sample remained full amorphous, which is beneficial for the followed crystallization treatment to achieve desirable nano-sized grains. 3.3 Preparation of ALZ transparent nanoceramics by crystallization Controlling the crystallization kinetics of amorphous bulk is of critical importance to obtain nano-sized ceramics, by optimizing the post heat-treatment conditions, such as heating temperature and time [11] . Therefore, the heat-treatments for the ALZ bulk glass were conducted in the temperature range of 900–1300 °C for 2 h, and the influences of the heat-treatment temperature on crystallization behavior were thoroughly investigated. Fig. 4 presents the XRD patterns of ALZ specimens after heat-treatment. It is noted that the sample treated at 900 °C precipitated a crystalline phase La 2 Zr 2 O 7 . This should be attributed to the nucleation effect of ZrO 2 [19] . The amount of ZrO 2 in this composition was considerably more than what is commonly used in glass–ceramic technology as a nucleation agent. The precipitation of plenty of fine compound La 2 Zr 2 O 7 renders heterogeneous nucleation sites and improves nano-crystallization. At 1000 °C, the diffraction peak of La 2 Zr 2 O 7 almost disappeared, and the following crystalline phases precipitated: LaAlO 3 and t-ZrO 2 . As for the sample treated at 1100 °C and 1200 °C, except LaAlO 3 and t-ZrO 2 , no new crystalline phase was detected. When the temperature increased to 1300 °C, minor LaAl 11 O 18 (β) appeared. However, the diffraction intensities of LaAlO 3 and t-ZrO 2 only increased slightly. This indicates that the crystallization process is almost complete. Fig. 5 shows that both the density ( ρ ) and crystallization degree ( α ) of the samples increased dramatically with an increase of heat-treatment temperature. A complete crystallization was achieved at 1300 °C for 2 h, accompanied by a density increase from 4.55 g/cm 3 to 5.34 g/cm 3 . Notably, when the treatment was performed at 1200 °C for 2 h, the crystallization degree had reached 95.2%, which suggests that the transition from glass to crystalline ceramic had been realized. XRD patterns in Fig. 4 also show that the sample heated at 1200 °C for 2 h mainly consisted of LaAlO 3 and t-ZrO 2 crystallines, and no amorphous characteristic was observed. Thus, LaAlO 3 /t-ZrO 2 nanoceramic was successfully prepared in this study. Scanning electron microscopy (SEM) observations on fracture surface also provide good insight into the processes occurring during crystallization, as shown in Fig. 6 . The LaAlO 3 /t-ZrO 2 ceramic obtained at 1200 °C exhibited a compact surface, although a great volume shrinkage took place during the crystallization process. When the heating temperature was elevated to 1300 °C, the fracture surface of the sample appeared to be intergranular, which directly indicates that the sample had been completely crystallized. The larger acicular grains should be designated as the LaAl 11 O 18 phase (β) based on the EDS analysis. These are well consistent with the results described above. Controlled grain growth was achieved by controlled crystallization via modifying the heating temperature from 900 °C to 1300 °C. The grain sizes were obtained from the XRD and SEM analysis, and summarized in Table 1 . It is observed that the average grain size ( D ) remained below 40 nm until the heating temperature was elevated to 1200 °C, accompanied by a slight loss in transparency due to the neglectable light scattering from grain boundaries. So the LaAlO 3 /t-ZrO 2 ceramic obtained at 1200 °C showed a transparency up to 70% in the infrared region, which is almost the same as that of the parent glass, while in the visible region the transmittance at 800 nm decreased from 68% to 55%. The refractive index of LaAlO 3 /t-ZrO 2 nanoceramic was estimated by a mixture rule described as Eq. (1) , and the theoretical transmittance is plotted in Fig. 3 ( T th = 78%, dotted line). By comparison with the T th , the obtained LaAlO 3 /t-ZrO 2 nanoceramic showed a relative high transparency in the infrared region, and some absorption and scattering losses in the visible region because of the presence of nanocrystals ( Fig. 6 ). When the heating temperature was 1300 °C, the grains grew quickly to ∼300 nm, and the sample became opaque. According to the Rayleigh–Gans–Debye light-scattering theory, the optical properties of non-cubic systems are highly dependent on the grain size because of the light scattering from birefringence [7] . To illustrate the discovery, let us take again the example of Al 2 O 3 ceramic ( c / a = 2.72). If the grain size can be reduced to nano-scale, the transmittance will be improved to the theoretical value. Contrastly, if the grain size is in micrometer scale, the sample is translucent. In this study, the transmittance of LaAlO 3 /t-ZrO 2 nanoceramic exhibited a similar dependence on the grain size because of the high birefringence of this extremely anisotropic lattice of LaAlO 3 ( c / a = 2.44), while actually t-ZrO 2 produced a much smaller deviation in transparency from the cubic symmetry ( c / a = 1.01). Moreover, the shrinkage-induced pores and cracks are liable to form during crystallization, although these defects were not detected by the SEM analysis. This could be one more reason for the opacity of the completely crystallized samples obtained at 1300 °C. In addition, the heat-treated ALZ samples exhibited enhanced Vickers hardness ( H V ) and fracture toughness ( K IC ) (see Table 1 ). Compared with the sintered basic glass, the transparent LaAlO 3 /t-ZrO 2 nanoceramic obtained at 1200 °C had an almost doubled Vickers hardness of 19.05 GPa and fracture toughness of 2.64 MPa m 1/2 , respectively. This remarkable improvement should be attributed to the increased density and crystallization degree, and the nanoscale microstructure. It is well known that optical sapphire (Al 2 O 3 ), spinel (MgAl 2 O 4 ) and AlON (Al 23−1/3 x O 27+ x N 5− x , 0.429 < x < 2) have widely been used as window materials because of their excellent combined properties [20] . Of these, AlON transparent polycrystalline ceramic is an extremely durable material with mechanical properties similar to sapphire, for example, a Vickers hardness of 16.00 GPa and fracture toughness of 2.40 MPa m 1/2 , respectively [21] . In this work, the prepared transparent LaAlO 3 /t-ZrO 2 nanocomposite exhibited more attractive mechanical properties, which suggests a potential application in window materials. However, more efforts should be made to improve the optical properties of LaAlO 3 /t-ZrO 2 nanocomposite by optimizing the new developed process. 3.4 Preparation of other transparent nanoceramics by ASCC Transparent LaAlO 3 /t-ZrO 2 nanoceramic ( Fig. 7 a) was successfully prepared by the ASCC method in this study, and had attractive optical and mechanical properties. This should be attributed to three reasons: medium glass-forming ability of the eutectic system, wide enough kinetic window for densification, and controlled crystallization process to produce nanostructure. Particularly, the first one is the basic condition for the achievements of glass powders, amorphous densification by hot-pressing and complete crystallization via a simple heat-treatment. That is also the reason why ceramic materials are generally fabricated by a solid-state reaction rather than the crystallization process. As can be seen in Fig. 7 , several transparent nanoceramics have been prepared by the developed ASCC method, including 1.5%Er:LaAlO 3 /t-ZrO 2 , 1%Nd:YAG/HfO 2 and 3%Nd:YAG/HfO 2 . The fluorescent properties of these transparent nanoceramics will be a subject of our further studies. 4 Conclusions In this study, a new method of amorphous sintering followed by controlled crystallization (ASCC) was developed to fabricate transparent nanoceramics. Glass powders from Al 2 O 3 –La 2 O 3 –ZrO 2 were synthesized, sintered by viscous flow, and then converted to transparent LaAlO 3 /t-ZrO 2 nanoceramics by post-heat-treatment. In the densification of glass powders, the higher pressure instead of temperature was essential to sinter transparent glass. The heat-treatment performed at 1200 °C for 2 h produced transparent LaAlO 3 /t-ZrO 2 nanoceramic with an average grain size of 40 nm. Due to the reduced light scattering from birefringence, the transmittance of LaAlO 3 /t-ZrO 2 nanoceramic (1.0 mm thick) was up to 55% at 800 nm. Meanwhile, the nanoceramic exhibited a Vickers hardness of 19.05 GPa, and fracture toughness of 2.64 MPa m 1/2 , respectively. Compared with conventional sintering process, this technique provides a new route to prepare transparent nanocermics by separate the densification stage from grain growth, in which the grain size is easily controlled by modifying the crystallization temperature. The method is also applicable to other systems which have a medium glass-forming ability. Acknowledgment This work was supported by the National Natural Science Foundation of China under Grant Nos. 50772116 and 50932006 (Key Project). References [1] A. Ikesue Y.L. Aung Nature 2 2008 721 [2] A. Ikesue Y.L. Aung J. Am. Ceram. Soc. 89 2006 1936 [3] V. Lupei A. Lupei A. Ikesue Opt. Mater. 30 2008 1781 [4] G.L. Messing A.J. Stevenson Science 322 2008 383 [5] K. Rayashi O. Kobayashi S. Toyoda K. Morinaga Mater. Trans., JIM 32 1991 1024 [6] M. Stuer Z. Zhao U. Aschauer P. Bowen J. Euro. Ceram. Soc. 30 2010 1335 [7] R. Apetz M.P.B. van Bruggen J. Am. Ceram. Soc. 86 2003 480 [8] W.D. Kingery H.K. Bowen D.R. Uhlmann Introduction to Ceramics second ed. 1976 Wiley & Sons New York [9] R. Fedyk D. Hreniak W. Łojkowski W. Stręk H. Matysiak E. Grzanka S. Gierlotka P. Mazur Opt. Mater. 128 2002 248 [10] A. Rosenflanz M. Frey B. Enderson E. Richards C. Schardt Nature 430 2004 761 [11] K. Lu Adv. Mater. 11 1999 1127 [12] N. Sakamoto S. Araki M. Yoshimura J. Am. Ceram. Soc. 92 2009 S157 [13] S.M. Lakiza L.M. Lopato J. Euro. Ceram. Soc. 25 2005 1373 [14] W. Heller J. Phys. Chem. 69 1965 1123 [15] S. Inaba S. Fujino J. Am. Ceram. Soc. 93 2010 217 [16] A. Karamanov M. Pelino J. Euro. Ceram. Soc. 19 1999 649 [17] J.S. Cheng H. Li L.Y. Tang F. He Glass Ceramics first ed. 2006 Chemical Industry Press Beijing [18] T. Vasilos J. Am. Ceram. Soc. 43 1960 517 [19] S. Banijamali B. EftekhariYekta H.R. Rezaie V.K. Marghussian Thermochim. Acta 488 2009 60 [20] J.M. Sands C.G. Fountzoulas G.A. Gilde P.J. Patel J. Euro. Ceram. Soc. 29 2009 261 [21] J.W. McCauley P. Patel M.W. Chen G. Gilde E. Strassburger B. Paliwal K.T. Ramesh D.P. Dandekar J. Euro. Ceram. Soc. 29 2009 223