From amorphous phase separations to nanostructured materials in sol–gel derived ZrO2:Eu3+/SiO2 and ZnO/SiO2 composites

Share Embed


Descrição do Produto

Journal of Non-Crystalline Solids 352 (2006) 2152–2158 www.elsevier.com/locate/jnoncrysol

From amorphous phase separations to nanostructured materials in sol–gel derived ZrO2:Eu3+/SiO2 and ZnO/SiO2 composites A. Gaudon, F. Lallet, A. Boulle, A. Lecomte, B. Soulestin, R. Guinebretie`re *, A. Dauger Science des Proce´de´s Ce´ramiques et de Traitements de Surface, UMR CNRS no. 6638, ENSCI, 47 Avenue Albert Thomas, 87065 Limoges, France Received 21 July 2005; received in revised form 30 January 2006 Available online 4 May 2006

Abstract ZrO2:Eu3+–SiO2 and ZnO–SiO2 composites have been synthesized by a sol–gel method by using a specific gelation and drying procedure. In the two cases we were able to produce large and transparent monolithic samples. Microstructural properties of these materials were investigated by thermo-differential and thermo-gravimetric analysis, X-ray diffraction, transmission electron microscopy and small angle X-ray scattering. The existence of a miscibility gap in both systems results in the formation of nanocomposites where crystallized zirconia or amorphous zinc oxide nanoparticles are dispersed in a silica glass matrix. These two kinds of nanocomposites are potential high efficiency luminescent materials because the nanoparticles size is easily controlled by the annealing conditions. Ó 2006 Elsevier B.V. All rights reserved. PACS: 81.07.b; 81.20.Fw; 61.46.+w Keywords: Nanocrystals; X-ray diffraction; Nanocomposites; Silica; Silicates; Sol–gels (xerogels)

1. Introduction In the recent years there has been intense interest in the development of new luminescent materials with a high efficiency. Nanocrystalline materials, and especially ceramics built with nanocrystallites [1,2], have recently attracted special attention because a significant enhancement of the optical properties occurs when the crystallites size is reduced to a few nanometers. Due to size effects such materials behave differently, in various spectroscopic processes, than their large-grain counterparts [3,4]. A special form of nanocomposites attract more and more attention: silica glass doped with various light-emitting nanocrystallites [5,6]. Preparation of such transparent nanostructured ceramics is a difficult technological task. Sol–gel routes can result *

Corresponding author. Tel.: +33 5 55 45 22 07; fax: +33 5 55 79 09 98. E-mail address: [email protected] (R. Guinebretie`re).

0022-3093/$ - see front matter Ó 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.jnoncrysol.2006.02.054

in a significant simplification for the elaboration of such materials [6–8] with a high degree of compositional and homogeneity control. In particular, we have recently shown [9] that using sol–gel processing in a silica-based binary oxide system including a miscibility gap could conduct to bulk nanocomposites with a good control of the size and the spatial distribution of the crystalline particles. In this work we focus on two kinds of nanocomposites [10,11] as potential luminescent materials. In the first one, we introduce small quantities of a well known emissive dopant: Eu3+ in host zirconia nanoparticles distributed in a silica glass matrix. In the second one we disperse zinc oxide nanoparticles in a silica glass matrix. ZnO is well known [12,13] to be a large gap semiconductor with promising optical properties. In both cases the corresponding phase diagram exhibits a miscibility gap. In this paper we present recent results on the investigation of sol–gel derived ZrO2:Eu3+/SiO2 and ZnO/SiO2 bulk

A. Gaudon et al. / Journal of Non-Crystalline Solids 352 (2006) 2152–2158

nanocomposites. These material are synthesized by a specific gelation and drying procedure that allows to produce large samples of bulk xerogel [9]. For each system the effects of thermal treatment on the microstructure have been investigated using thermal-analysis (thermo-differential DTA and thermo-gravimetric TGA), X-ray diffraction (XRD), transmission electron microscopy (TEM) and small angle X-ray scattering (SAXS). 2. Experimental 2.1. Material synthesis 2.1.1. ZrO2:Eu2O3–SiO2 system For this study undoped (ZrO2–SiO2) and doped (ZrO2:Eu2O3–SiO2) materials were prepared. According previous work [9] and in order to be in the middle of the silica–zirconia miscibility gap [14] the mixture has been fixed to 30 mol% ZrO2–70 mol% SiO2, doping was achieved by adding 2 mol% Eu2O3 in relation to the Zr content. All the sols were synthesized in a dry atmosphere to avoid uncontrolled hydrolysis of precursors. We used tetraethyl-orthosilicate (TEOS: Si(OEt)4, Aldrich Chemicals), europium chloride (EuCl3, Merk), zirconium n-propoxide (Zr(OPr)4, Alfa Products) as SiO2, Eu2O3 and ZrO2 precursors and acetylacetone (acac: C5H8O2, Merk) as chelating agent for the zirconium n-propoxide [15]. The reaction rate of silica and zirconia precursors toward water is quite different; silicon alkoxide without any catalyst is ‘stable’ over a long period of time before gelation begins, while zirconium alkoxide reacts immediately. To match the precursors reactivity [16], the silica solution was pre-hydrolyzed in acidic conditions [17] at room temperature under mechanical stirring with a mixed solution of propanol, H2O and HCl. Then, the chelated zirconia precursor solution was slowly added while continuously stirring. Finally the final amount of H2O is slowly mixed. The sol parameters are the concentration in alkoxides, C = [Zr(OPr)4] + [Si(OEt)4] = 0.5 mol L1, the hydrolysis ratio W = [water]/[alkoxides] = 10, the complexing ratio R = [acac]/[Zr(OPr)4] = 0.7 and the doping ratio x = [Eu2O3]/ [ZrO2] = 0 or 2 mol%. 2.1.2. ZnO–SiO2 system For this study, ZnO–SiO2 gels were prepared with 10 mol% ZnO. The starting materials were tetraethylorthosilicate (TEOS:Si(OEt)4, Aldrich Chemicals) and zinc nitrate (Zn(NO3)2,6H2O, Aldrich Chemicals). First, zinc nitrate is dissolved at room temperature in ethanol under mechanical stirring. Then this solution is mixed with the silica precursor. The resulting mixture is then slowly hydrolyzed with the required quantity of water under mechanical stirring. The sol parameters are the hydrolysis ratio W = [water]/[alkoxides] = 10 and the concentration in metallic cations, C = [Si(OEt)4] + [Zn(NO3)2, 6H2O] = 1 mol L1.

2153

2.1.3. Drying procedure All the sols gelled after a period between few hours and few days at 60 °C. Gelling times for silica–zirconia and silica–zinc oxide sols are noticeably different; the gelling point is reached in only 2 h for SiO2–ZrO2 samples whereas one week is required for SiO2–ZnO samples. The gels were slowly dried at 60 °C under a 100% humidity atmosphere in order to form massive xerogels, as first proposed by Nogami [18] and as we have reported recently [9]. Resulting samples were then slowly fired at 400 °C for 5 h with a heating rate of 1 °C/min, what allows removing the major part of residual organics compounds and preventing the apparition of cracks into the materials. Finally the samples are cut out and polished so as to obtain slices of definite thickness. Subsequently the samples are fired with a heating rate of 5 °C/min at a temperature ranging between 1000 and 1300 °C for the zirconia-based specimens and between 600 and 900 °C for the ZnO–SiO2 samples. SAXS experiments were carried out on optically polished small slices about 100 lm in thickness. 2.2. Characterization SAXS measurements were performed using a home made set up [19,20] with a point like collimation geometry. The monochromatic X-ray incident beam was provided by a double channel cut (2 2 0) germanium monochromator adapted to a 12 kW rotating anode with a Cu target. The scattered intensities were recorded using a linear position sensitive detector (Elphyse). The sample to detector distance, fixed to 0.5 m, allows to cover a q-range from 0.1 to 4 nm1 where q is the length of the scattering vector, q = 4pk1sin(h), k is the Cu Ka1 wavelength and 2h is the scattering angle. In order to determine the particles size distribution the SAXS curves are fitted with a model in which the interaction between particles is described by a hard sphere potential [21–24]. In the present work we assume, according to TEM observation (see Fig. 4), that particles are spherical and that their size distribution p(D) can be described by a lognormal distribution. For a polydisperse system with inter-particle interference effect, the scattering intensities I(q) are given by [23,24] Z 1 2 IðqÞ ¼ N Dp pð2RÞF ðq; RÞ Sðq; 2Rhs ; ghs ÞdR; ð1Þ 0

where N is the precipitates density number and Dp is the scattering length difference between matrix and precipitates. Rhs is the hard-sphere radius and ghs is the volume fraction of hard-sphere. The volume fraction (g) of the precipitates can be calculated as  3 R g ¼ ghs . ð2Þ Rhs F(q, R) is the particle form factor, which for spherical precipitates of radius R can be expressed as

2154

A. Gaudon et al. / Journal of Non-Crystalline Solids 352 (2006) 2152–2158

F ðq; RÞ ¼ 4pR3

sinðqRÞ  qR cosðqRÞ

. ð3Þ 3 ðqRÞ S(q, 2Rhs, ghs) is the interparticle structure factor for hard-sphere model describing the interparticle interference effect. It can be calculated analytically in the Perkus–Yevick approximation [25,26]. XRD experiments were carried out in the Debye–Scherrer geometry with a home-made diffractometer previously described [27] and based on a sealed tube operating at 37.5 kV/28 mA, a quartz monochromator (Cu Ka1 radiation) and a curved position sensitive detector (Inel CPS 120). In order to determine the particles size distribution the XRD curves are fitted with a model based on physical description of the sample taking into account crystallite size (and size distribution) effects, lattice disorder effects as well as the contribution of diffractometer to the XRD signal. The intensity can be written as a Fourier series X IðqÞ ¼ AS ðLÞAD ðLÞAI ðLÞeiqL ; ð4Þ L

where L is a length in real space and AS, AD and AI are the Fourier coefficients describing the effects of size, lattice disorder and the instrument. In the present case the effect of lattice disorder has been found to be negligible, so that AD(L)  1. In consistency with SAXS we assume the crystallites to be spherical and lognormally distributed. With these assumptions an expression of AS can be found in [28]. A detailed description of AI has been given elsewhere [29]. The microstructure of the sample was also investigated using a Jeol 2010 transmission electron microscope operating at 200 kV. Thermal analysis data were recorded on a Setaram Setsys thermoanalyzer. All the experiments were carried out with a heating rate of 10 °C/min under a dry air atmosphere. 3. Results 3.1. Silica–zirconia nanocomposites The DTA and TGA curves for the doped and undoped gels are shown in Fig. 1(a) and (b). Even if it is more obvi-

ous for undoped samples, all the gels show an endothermic peak around 250 °C in the DTA curve which correspond to a weight loss in TGA. According to previous results on zirconia xerogels [15] and on silica–zirconia xerogels [7] that can be unambiguously attributed to desorption of physically adsorbed water in that temperature range. The continuous weight loss below 600 °C and corresponding exothermic peaks around 450 °C can be associated to the pyrolysis of the residual organic compounds. At low temperature doped and undoped gels have fairly the same thermal behavior with only different quantities of physically adsorbed water. TGA (Fig. 1(b)) and X-ray diffraction (Fig. 2(a) and (b)) results prove that the xerogels (doped and undoped) obtained after 1 h at 600 °C are amorphous with a low residual content of organic compounds. SAXS investigations show that these materials are not homogeneous at a nanometric scale. The corresponding intensity distributions (Fig. 3(a)) clearly show a peak at qm = 0.3–0.4 nm1 corresponding to a correlation length of about 18 nm in real space. TEM observations on the same sample (Fig. 3(b)) show the existence of compositional fluctuations characterized by a length scale in accordance with this value. Further inspection of the diffraction patterns reported in Fig. 2(a) and (b) reveals that between 800 °C and 1000 °C distinct peaks appear which can be unambiguously attributed to tetragonal zirconia. Both DTA (Fig. 1(a)) curves show the same significant exothermic peak at 947 °C, according to XRD data this peak can be assigned to the crystallization of zirconia crystals. The addition of a small amount of europium chloride to silica–zirconia xerogels does not change their crystallization behavior. According to the affinity of Eu2O3 for zirconia, which form a wellknown solid-solution [30,31], we can assume, although it remains to be proved, that zirconia nanocrystals are effectively doped with Eu3+ ions. TEM micrograph (Fig. 4(a)) and the corresponding SAXS curve (Fig. 4(b)) for a sample fired 5 h at 1100 °C illustrate the resulting microstructure of this heterogeneous crystallization process. For such high temperature heat treatments (higher than 1000 °C) the microstructure of 0

70

(b)

(a)

60

-5

50 Doped

30

TGA (%)

DTA (µV)

40

20 10

-10 Doped

-15

Undoped

0

Undoped

-20

-10 -20

-25 0

300

600

900

Temperature (°C)

1200

1500

0

300

600

900

1200

1500

Temperature (°C)

Fig. 1. DTA (a) and TGA (b) curves or the doped and undoped silica–zirconia xerogels. Heating rate is 10 °C/min. The exothermic peak at 947 °C is associated to crystallization of tetragonal zirconia.

A. Gaudon et al. / Journal of Non-Crystalline Solids 352 (2006) 2152–2158

(a)

1300°C 1h

1200°C 1h

Tetragonal zirconia

Intensity (a.u.)

Intensity (a.u.)

Tetragonal zirconia

40

60

80

2 Theta (°)

(b)

1300°C 1h

1200°C 1h

1100°C 1h

1100°C 1h 1000°C 1h 800°C 1h 600°C 1h

20

2155

1000°C 1h 800°C 1h 600°C 1h

20

40

60

2 Theta (°)

80

Fig. 2. XRD patterns of (a) undoped and (b) dopped ZrO2–SiO2 xerogels calcined at different temperatures. For both samples, characteristics peaks of the tetragonal zirconia phase are observed above 1000 °C.

Fig. 3. SAXS intensity distribution (a) of undoped and doped ZrO2–SiO2 xerogels calcined at 600 °C for 1 h and corresponding TEM micrograph (b) for pure ZrO2–SiO2 xerogel.

Fig. 4. TEM micrograph (a) and corresponding SAXS intensity distribution (b) of a sample fired for 5 h at 1100 °C. The sample exhibits a correlation length (18 nm) related to the mean distance between crystalline zirconia particles.

these materials can be described as follows: tetragonal zirconia nanocrystals are not randomly dispersed in a dense silica glass matrix. Annealing at high temperature (respectively 1 h at 1000, 1100, 1200 and 1300 °C), promotes the growth of the crystalline particles as attested by SAXS (Fig. 5(a)). The evolution of size and size distribution of the zirconia

particles can be extracted by fitting the SAXS and the XRD patterns. The resulting distributions (Fig. 6(a) and (b) respectively for SAXS and XRD) have the same behavior and the average dimensions are in good agreement with TEM observations. However, it can be observed that dimensions obtained by XRD are systematically smaller than these obtained by SAXS. This can be understood

2156

A. Gaudon et al. / Journal of Non-Crystalline Solids 352 (2006) 2152–2158 180000

1000

1000

1300°C 1h

(b)

(a)

3

100

500

120000

I(q/qm(t)).qm(t)

Intensity (a.u.)

1000°C 1h

1200°C 1h

0 0.0

60000

0.5

1.0

1.5

10

1

0.1 1100°C 1h

0 0.00

0.25

0.50

0.75

0.01 0.1

1.00

1

10

q/qm

-1

q (nm )

Fig. 5. SAXS intensity distribution (a) for europium doped silica zirconia nanocomposites calcined at different temperatures (respectively 1000 °C 1 h, 1100 °C 1 h, 1200 °C 1 h and 1300 °C 1 h) and the corresponding dynamical scaling behavior of the structure factors (b).

0.10

0.10

(a) SAXS

(b) XRD

0.05

p(D)

p(D)

1000°C 1h

1000°C 1h

0.05

1100°C 1h

1100°C 1h

1200°C 1h 1300°C 1h

1200°C 1h 1300°C 1h 0.00

0.00

0

5

10

15

20

25

30

35

40

D (nm)

0

5

10

15

20

25

30

35

40

D (nm)

Fig. 6. Calculated size distributions, (a) SAXS (b) XRD, of the zirconia nanocrystals for europium doped silica zirconia nanocomposites calcined at different temperatures (respectively 1000 °C 1 h, 1100 °C 1 h, 1200 °C 1 h and 1300 °C 1 h).

considering that SAXS is sensitive to compositional fluctuations whereas XRD is sensitive to crystal lattice. For instance polycrystalline grains will appear as different crystals with smaller dimensions in XRD whereas they will appear as a single particles for SAXS. Both methods are hence complementary, depending whether one is interested in the crystalline properties or not. The rather good agreement between the distributions obtained by SAXS and XRD indicates that the proportion of polycrystalline grains must be rather low which is certainly a consequence of phase separation prior to crystallization. 3.2. SiO2–ZnO nanocomposites The behavior of ZnO–SiO2 nanocomposites is very similar to ZrO2–SiO2 nanocomposites we therefore provide only the main results without going into details. DTA and TGA curves for ZnO–SiO2 samples containing 10 mol% ZnO are shown in Fig. 7(a). The endothermic peak around 180 °C in the DTA curve and subsequent weight loss can be attributed to desorption of physically adsorbed water. The continuous weight loss below 600 °C

and corresponding exothermic peaks can be associated to the pyrolysis of the residual organic compounds. The monolithic xerogels obtained after 1 h at 600 °C are totally transparent and according to XRD data (Fig. 7(b)) they are amorphous. SAXS investigations show that these materials are not homogeneous at a nanometric scale. The corresponding intensity distributions (Fig. 8(a)) clearly show a peak at qm = 0.3–0.4 nm1 corresponding to a correlation length of about 18 nm in direct space. TEM observations (Fig. 8(b)) prove that compositional fluctuations are the cause of scattered intensity. The material is amorphous and its microstructure can be described as a distribution of amorphous zinc oxide-rich nanoparticles in a silica-rich amorphous matrix, similarly to that is occurring in the ZrO2–SiO2 system. Fig. 7(a) shows the XRD pattern of samples fired in the 600–900 °C temperature range. Between 600 °C and 900 °C distinct peaks appear which can be unambiguously attributed to zinc silicate. According to the XRD data the exothermic peak found at about 820 °C in the DTA curve (Fig. 7(b)) can be attributed to the crystallization of zinc silicate.

A. Gaudon et al. / Journal of Non-Crystalline Solids 352 (2006) 2152–2158 15

Intensity (a.u.)

(b)

0

10

-5

5

-10

0

-15

TGA (%)

900°C 1h

DTA (µV)

(a)

Zn2SiO4

2157

800°C 1h 700°C 1h 600°C 1h

-5 10

20

30

40

50

60

70

80

90

2 Theta (°)

0

200

400

600

800

-20 1000

Temperature (°C)

Fig. 7. XRD patterns (a) of the ZnO–SiO2 xerogels calcined at different temperatures and corresponding DTA ant TGA curves (b) at a heating rate of 10 °C/min.

Fig. 8. SAXS intensity distribution (a) and corresponding TEM micrograph (b) of a sample fired for 1 h at 600 °C. The sample exhibits a correlation length (18 nm) related to the mean distance between crystalline ZnO-rich particles.

4. Discussion 4.1. Silica–zirconia nanocomposites The scattered intensity distribution is proportional to the square modulus of the Fourier transform of the electron density distribution of the sample. SAXS is not sensitive to crystal lattice and the contrast arises from the presence of inhomogeneities in electron density with colloidal size [32] (i.e. few nanometers), such as particles, pores or compositional fluctuations. The main difficulty is to extract information about the nature, the shape and the spatial distribution of these inhomogeneities. In the present case SAXS results (Fig. 3(a)) and TEM observations (Fig. 3(b)) proves that, after 1 h at 600 °C, nano-scaled compositional fluctuation are present in doped and undoped silica – zirconia xerogels. They are amorphous and their microstructure is that of a bi-continuous phase separated system consisting in interconnected ZrO2-rich and SiO2-rich domains. It appears that the existence of this amorphous phase separation before crystallization strongly modifies the crystallization temperature of tetragonal zirconia in such silica–zirconia composites. This temperature is indeed sig-

nificantly different (about 500 °C higher) than for pure zirconia precursor xerogel [14]. Such a result has been also observed for various binary systems including a miscibility gap, for example TiO2–SiO2 [33], HfO2–SiO2 [34,35] and La2O3–SiO2 [34,36]. As we have recently reported [9] the previously formed interconnected amorphous texture constitutes the matrix from which the crystallization appears. Consequently the size distribution of zirconia nanocrystals are fairly monodisperse with a non random special distribution (highlighted by a well defined peak into SAXS intensity distributions). The curves corresponding to high temperature heat treatments (respectively 1 h at 1000, 1100, 1200 and 1300 °C cf. Fig. 5(a)) collapse into a single master curve 3 (Fig. 5(b)) when they are plotted as Iðq=qm2 ðtÞÞ  ðqm2 ðtÞÞ versus ðq=qm2 ðtÞÞ. It shows that the experimental structure factors are time independent and grow with dynamical self-similarity. This scaling behavior is known to occur in the late stages of phase separation after quenching into a miscibility gap [37,38], it applies to two-densities systems with constant compositions and volume fractions and proves that the growth of zirconia nanocrystals follows a pure coalescence mechanism.

2158

A. Gaudon et al. / Journal of Non-Crystalline Solids 352 (2006) 2152–2158

By using different annealing conditions, we were able to generate nanostructures with a definite size of zirconia nanocrystals (from 4 to 23 nm) depending on the annealing temperature. It is striking that the zirconia particles formed subsequently to crystallization are highly monodisperse: the standard deviation of the diameter distribution is only r = 0.98 nm for an annealing duration of 1 h at 1000 °C. This is believed to be a consequence of the bicontinous texture formed in the amorphous state prior to crystallization. The broadening of the distribution for higher temperatures (r = 1.90 nm, 3.75 nm and 6.00 nm for 1 h at 1100, 1200 and 1300 °C respectively) is a consequence of the coalescence mechanism mentioned earlier. 4.2. SiO2–ZnO nanocomposites The behavior of ZnO–SiO2 nanocomposites appears markedly different from the ZrO2–SiO2 nanocomposites. In particular the crystallization of zinc oxide is never observed whereas, and on the contrary to the SiO2–ZrO2 case, the silicate compound formation is observed at low temperature. This point is still under investigation. 5. Conclusion Large monolithic samples of ZrO2:Eu3+–SiO2 and ZnO– SiO2 have been produced by sol gel processing using a specific gelation and drying procedure. These samples were studied by thermal-analysis (DTA and TGA), XRD experiments, SAXS intensities measurements and TEM observations. The addition of few quantities of europium oxide into silica–zirconia xerogels does not change their behavior with respect to their undoped counterparts. Interconnected ZrO2-rich and SiO2-rich amorphous domains appear in the monolithic xerogel fired at low temperature (typically 1 h at 600 °C). The crystallization of zirconia particles occurs from the zirconia rich phase previously formed by phase separation and for high temperature heat treatments (1000 °C) highly monodisperse tetragonal zirconia nanocrystals are distributed throughout a dense silica glass matrix. The size and the size distribution of zirconia nanocrystals can be controlled by further thermal treatments at different temperatures for different durations. By using different annealing conditions we were able to produce significantly different nanostructures with a large range of possible size for zirconia nanocrystals (from 4 to 23 nm). Photoluminescence measurements are now required to confirm the insertion of europium ions on the zirconia structure and to investigate the dependence of the luminescence properties on the microstructure. Transparent silica–zinc oxide nanocomposites have been also produced via a sol–gel process and their behavior is similar to silica–zirconia nanocomposites. At low tempera-

ture (typically 600 °C) an amorphous phase separation has been found, amorphous zinc oxide-rich nanoparticles (about 10 nm) are distributed in a silica-rich amorphous matrix. However through further annealing the crystallization of ZnO zincite phase was never observed whereas the crystallization of zinc silicate occurs rapidly and at relatively low temperature (around 800 °C). Photoluminescence measurements should now be carry out to evaluate the efficiency of these nanocomposites. Acknowledgements The authors would like to express their gratitude towards the European Community (the European Social Funds) and the Limousin Region for their financial support of the present work. 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] [28] [29]

[30] [31] [32] [33] [34] [35] [36] [37] [38]

B.N. Bhargava, J. Lumin. 70 (1996) 501. L. Sun et al., J. Lumin. 87–89 (2000) 447. E.T. Goldburt et al., J. Lumin. 72–74 (1997) 190. H.S. Yang et al., J. Lumin. 83&84 (1999) 139. Z. Fu et al., J. Phys. Cond. Mat. 15 (2003) 2867. D. Hreniak et al., J. Phys. Chem. Solids 64 (2003) 111. D. Li et al., J. Non-Cryst. Solids 261 (2000) 273. J. Teyssier et al., J. Non-Cryst. Solids 341 (2004) 152. A. Gaudon et al., J. Eur. Ceram. Soc. 25 (2004) 283. H. He, Y. Wang, Y. Zou, J. Phys. D: Appl. Phys. 36 (2003) 2972. S.F. Wang et al., Opt. Mater. 27 (2004) 269. M. Abdullah, S. Shibamoto, K. Okuyama, Opt. Mater. 26 (2004) 95. S. Chakrabarti et al., Thin Solids Films 441 (2003) 228. K. Kim, P.C. McIntyre, J. Appl. Phys. 92 (2002) 5094. R. Guinebretiere et al., J. Non-Cryst. Solids 147&148 (1992) 542. O. Stachs, V. Petrov, T. Gerber, J. Appl. Crystallogr. 30 (1997) 670. J. Cihlar, Colloid Surf. 70 (1993) 239. M. Nogami, J. Non-Cryst. Solids 69 (1985) 415. A. Dauger et al., J. Appl. Crystallogr. 24 (1991) 765. A. Lecomte et al., J. Non-Cryst. Solids 225 (1998) 120. R. Triolo, E. Caponetti, S. Spooner, Phys. Rev. B 39 (1989) 4588. M. Hennion, D. Ronzaud, P. Guyot, Acta Metall. 30 (1982) 599. J.P. Pedersen, Phys. Rev. B 47 (1993) 657. J.P. Pedersen, J. Appl. Crystallogr. 27 (1994) 595. N.W. Ashcroft, J. Lekner, Phys. Rev. 145 (1966) 83. R. Triolo et al., Philos. Mag. A 60 (1989) 401. O. Masson, R. Guinebretiere, A. Dauger, J. Appl. Crystallogr. 29 (1996) 540. P. Scardi, M. Leoni, Acta Crystall. A 57 (2001) 604. A. Boulle, Thesis, Diffraction des rayons X sur couches d’oxydes e´pitaxie´es. Elaboration et analyse microstructural. 2002: Limoges, France. A. Gedanken et al., J. phys. chem. B 104 (2000) 7057. R. Reisfeld et al., Chem. Phys. Lett. 349 (2001) 266. A. Guiner, F. Fournet, Small-angle scattering of X-rays, John Wiley & Sons, Inc., New-York, 1955. R.N. Viswanath, S.R. Ramasamy, Colloid. Surf. 133 (1997) 49. J. Robertson, Eur. Phys. J. Appli. Phys. 28 (2004) 265. S. Stemmer et al., Jpn. J. Appl. Phys. 42 (2003) 3593. J.P. Maria et al., J. Appl. Phys. 90 (2001) 3476. G.E.S. Brito et al., J. Non-Cryst. Solids 217 (1997) 41. H. Furukawa, J. Appl. Crystallogr. 21 (1988) 805.

Lihat lebih banyak...

Comentários

Copyright © 2017 DADOSPDF Inc.