2010 - Preparation of Nanocrystalline Lithium Niobate Powders at Low Temperature

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Cryst. Res. Technol. 45, No. 9, 977 – 982 (2010) / DOI 10.1002/crat.

201000105

Preparation of nanocrystalline lithium niobate powders at low


temperature

Ting Ting Su1,2, Heng Jiang*1, Hong Gong1, and Yu Chun Zhai2
1
School of Chemistry and Materials Science, Liaoning Shihua University, Fushun 113001, P. R. China
2
School of Materials Science and Metallurgy, Northeastern University, Shenyang 110005, P. R. China

Received 22 February 2010, revised 1 July 2010, accepted 3 July 2010


Published online 15 July 2010

Key words lithium niobate powders, ammonium niobium oxalate, alternative solid-state method.

A facile route to prepare lithium niobate (LiNbO3) powders was proposed by an alternative solid-state
method. Stoichiometric Li2C2O4 and ammonium niobium oxalate were mixed with small amounts of water
and then dried at room temperature. It was demonstrated that Li[NbO(C2O4)2]·nH2O intermediate was
produced by an ion-exchange reaction. Pure LiNbO3 powders were successfully synthesized by heating the
intermediate at 500, 600 and 700 °C for 3 h. X-ray diffraction (XRD), scanning electron microscopy (SEM),
Fourier-transform infrared (FTIR) spectroscopy, UV-Vis diffuse reflectance (UV-Vis) spectroscopy and
thermogravimetric (TG) analysis were used to characterize the precursor compound and as-prepared samples.
XRD results reveal that all the products are identified as hexagonal structure with high relative crystallinity
(>87%). The particle size is found to be about 40 nm for the mixture calcined at 500 °C according to XRD
data, which is in good agreement with SEM data. The as-prepared LiNbO3 powders by this method are high
quality according to FTIR spectra. (Li0.996Nb0.005)Nb0.999O3 phase was formed when the calcination
temperature was raised to 800 °C.
© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1 Introduction
Lithium niobate (LiNbO3) is widely used in ferroelectric, pyroelectric, dielectric and piezoelectric devices due
to its excellent electro-optical, photo-refractive and nonlinear optical properties [1-5]. In recent years, despite it
is potential to use wet chemical synthesis to prepare LiNbO3 powders, there are still some problems existing in
the methods. For example, alkoxide route [6-13] has serious drawbacks: the higher price of raw materials, the
inflammability during heating and the high reactivity with water. Therefore, sol-gel method must be kept away
from atmospheric moisture and carried out under dry inert atmosphere. Alternatively, the use of niobium
chloride (NbCl5) often results in residual chlorine as impurity and easily forms toxic hydrogen chloride (HCl)
gas in the final product. Large amounts of organic compounds are used in preparation of LiNbO3 powders by
homogeneous precipitation [14,15] or polymerized complex [16] methods, which must be calcined above
600 °C with high weigh loss in order to obtain pure LiNbO3. Hydrothermal synthesis is an efficient method in
decreasing the synthetic temperature, however, the reaction time is more than 10 h [17-19], even longer
(3 days) [13].
It has been known that LiNbO3 is usually fabricated by conventional solid-state reaction between niobium
pentoxide (Nb2O5) and lithium carbonate (Li2CO3) above 1000 °C for 24 h [20]. Evaporation of volatile
components during long-term, high-temperature solid-state synthesis may lead to uncontrolled changes in the
Li:Nb ratio in the reaction product, which may yield not only LiNbO3 but also Li3NbO4 and LiNb3O8
components in the whole system [21]. Xue et al. [22,23] have reported the preparation of high-quality LiNbO3
powders with urea as fuel at 550-600 °C. LiNbO3 powders have been rapidly prepared at 650 °C by a molten
salt method, in which a wet ball-milling process has been involved and the contamination has been introduced
inevitably [24].
Nb2O5 shows an extremely high chemical stability and poor water-solubility in the special case of niobium-
based compounds. In this work, the synthesis of LiNbO3 powders is studied using water-soluble ammonium
____________________

* Corresponding author: e-mail: hjiang78@hotmail.com

© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


978 Ting Ting Su et al.: Preparation of nanocrystalline lithium niobate powders

niobium oxalate (ANO hereafter) [25-27] and lithium oxalate (Li2C2O4) as raw materials by an alternative
solid-state method. The stoichiometric mixture of raw materials has been uniformly distributed by adding small
amounts of water, which replaces the ball milling process. The whole chemical reaction can be finished by a
facile route and also reduces the reaction temperature (500 °C). X-ray powder diffraction and FTIR spectrum
are employed to characterize the quality of the synthesized LiNbO3 powders.

2 Experimental
Material preparation The raw materials of Li2C2O4, Li2CO3, and Nb2O5 were obtained from commercial
sources and were of analytical grade. In the synthesis of LiNbO3 powders, ANO was prepared firstly as
previously reported process [25]. The molecular formula for the recrystallized ANO was established as
NH4[NbO(C2O4)2(H2O)2](H2O)3 by XRD analysis, which matched well with the standard data of JCPDS 83-
1993. The water content of each reagent was measured by TG analysis considering the hygroscopic property.
The stoichiometric Li2C2O4 (1.53 g) and ANO (NH4[NbO(C2O4)2(H2O)2](H2O)3, 12.38 g, Nb2O5 content is
32.30% by TG analysis) with 1: 1 molar ratio of Li: Nb were mixed by adding 2 mL water to make them as
slurry. The slurry was pestled for about 1-2 h under IR lamp to obtain uniformly distributed mixture and then
dried at room temperature. The mixture was heated in a corundum crucible at 475, 500, 600, 700, and 800 °C
for 3 h, respectively. These obtained samples were named as LN475, LN500, LN600, LN700 and LN800,
respectively.
Two controlled experiments were performed using different raw materials, i.e., Li2C2O4/Nb2O5 and
Li2CO3/ANO. The operation conditions were the same as described above.
Material characterization The crystal structures of the samples were determined by powder X-ray
diffraction (XRD) using a D/MAX-RB X-ray diffractometer (XRD, Rigaku, Japan) with graphite
monochromater and CuKα radiation (40 kV, 100 mA). The scanned range was 2θ=5-70°, with a step width of
0.02°, at a rate of 4° (2θ) per min. The phase identification and calculation of relative crystallinity were
performed by MDI Jade 5.0 software. The degree of relative crystallinity is calculated according to the total
area of the diffraction peaks at 23.7, 32.7, 34.8, 39.0, 40.1, 42.6, 48.5, 53.3, 56.1, 57.0, 61.1, 62.5 and 68.6° in
the XRD pattern of the samples. For LN series samples, the powder crystalline calcined at 800 °C (LN800) was
considered to be 100% in crystallinity for the biggest area [28]. The degree of relative crystallinity was
determined by comparing the total area under crystalline peaks of other samples with that of LN800,
respectively [29]. The morphologies were observed using a scanning electron microscope (SEM) (JSM-7500F-
EDS). Fourier transform infrared (FTIR) spectra were obtained in KBr discs on a Perkin-Elmer Spectrum GX.
Sixteen scans were co-added with a resolution of 4 cm-1, in the range of 2000-400 cm-1. UV-Vis diffuse
reflectance spectra were recorded with a Lamda 900 UV-Vis-NIR spectrophotometer. Thermogravimetric (TG)
analysis was carried out on a Perkin-Elmer Pyris 1 TGA. The atmosphere was air with a flow rate of
20 mL min-1 at 20 K min-1 in the range from 30 to 800 °C.

3 Results and discussion


For the sake of comparison, Li2C2O4 and ANO powders were mixed thoroughly with a molar ratio of [Li] :
[Nb]=1 : 1 under solvent-free condition by conventional solid-state reaction. The components of Li2C2O4/ANO
mixed without water were investigated by FTIR and TG analysis. FTIR spectra of Li2C2O4, ANO and
Li2C2O4/ANO mixed without water are shown in figure 1. In figure 1c, peaks of Li2C2O4/ANO mixture are
nearly the superposition of Li2C2O4 (Fig. 1a) and ANO (Fig. 1b). Five shoulder peaks at 1644, 1326, 773, 512
and 444 cm-1 in figure 1c are derived from the characteristic peaks of Li2C2O4 (assigned with #1-#5 dot line in
figure 1a). It can be inferred that the Li2C2O4/ANO powders are the mechanical mixture of Li2C2O4 and ANO.
In the present work, the raw materials have been uniformly mixed by adding small amounts of water and then
dried. Figure 1d and e show the FTIR spectra of (NH4)2C2O4 and Li2C2O4/ANO mixed with water. It can be
found from figure 1e, the disappearance of the characteristic peaks of Li2C2O4 and the presence of new bands
(assigned with #6 and #7 dot line) indicate that ion-exchange reaction was occurred betweent Li2C2O4 and
ANO (Equation 1). Li[NbO(C2O4)2]·nH2O precursor and (NH4)2C2O4 were formed when Li2C2O4 and ANO are
ground in the presence of water. It can be confirmed by the following FTIR and TG analysis.
2NH 4 [NbO(C2 O4 )2 (H2 O)2 ](H2 O)3 +Li 2 C2 O4 → 2Li[NbO(C2 O4 )2 ] ⋅ nH2 O+(NH4 )2 C2 O4 +2(5-n)H 2 O (1)

© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.crt-journal.org


Cryst. Res. Technol. 45, No. 9 (2010) 979

IR bands for the stretch vibration of C=O at 1712, 1687 cm-1 and C-O at 1396, 1260 cm-1 can be observed.
Peak at 965 cm-1 is due to the stretch vibration of Nb=O. Bands at 910 and 800 cm-1 are assigned to COO-. 542
and 505 cm-1 bands are the result of M-O stretch vibration. The new bands at 720 and 635 cm-1 indicate the
presence of (NH4)2C2O4 product.

Fig. 1 FTIR spectra of (a) Li2C2O4, (b) ANO, (c) Fig. 2 TG curves of (a) Li2C2O4, (b) ANO, (c)
Li2C2O4/ANO mixed without water (d) (NH4)2C2O4 and (e) Li2C2O4/ANO mixed without water (d) (NH4)2C2O4 and (e)
Li2C2O4/ANO mixed with water. Li2C2O4/ANO mixed with water.

Fig. 3 XRD powder diffraction pattern of Li2C2O4/ANO Fig. 4 XRD patterns of the as-synthesized LN475,
mixed without water and calcined at 500 °C for 3 h. LN500, LN600, LN700, and LN800.

The thermal behaviors of Li2C2O4, ANO, Li2C2O4/ANO mixed without water, (NH4)2C2O4 and Li2C2O4/ANO
mixed with water are investigated by TG analysis as shown in figure 2. In figure 2b and c, the TG curve of
Li2C2O4/ANO mixed without water is almost superposed with that of ANO below 300 °C. The subsequent two
main weight loss steps (450-550 °C and 600-650 °C) in figure 2c are coincided with those in figure 2a and b,
respectively. It can be inferred from above analysis the decomposition process of the Li2C2O4/ANO mixed
without water is an incorporation of Li2C2O4 and ANO. As shown in figure 2b, ANO decomposes to niobic
acid (Nb2O5·xH2O) at about 330 °C and then yields Nb2O5 at 620 °C [30]. In figure 2e, the weight loss between
50 and 500 °C is due to the decomposition of Li[NbO(C2O4)2]·nH2O and (NH4)2C2O4. The decomposition rate
of precursor is faster than that of ANO due to the weigh loss process of (NH4)2C2O4 below 240 °C. There is no
weight loss in figure 2e above 500°C, which is different from figure 2c. TG analysis of the
Li[NbO(C2O4)2]·nH2O and (NH4)2C2O4 precursor shows a 64.4 % weight loss from 50 to 500 °C. Pure LiNbO3
was also prepared using ANO and LiNO3 as raw material in the presence of sugar/PVA, however, the weight
loss reached 95 % to remove the organic compound [31].
According to above FTIR and TG analysis, only a mechanical mixture of Li2C2O4 and ANO can be formed
when they are mixed without water. Figure 3 shows the XRD result of the mechanical mixture calcined at
500 °C for 3 h. As shown in figure 3, large amounts of unreacted Li2CO3 and Nb2O5 have been detected, which

www.crt-journal.org © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


980 Ting Ting Su et al.: Preparation of nanocrystalline lithium niobate powders

indicates that pure LiNbO3 phase can not be formed at 500 °C when Li2C2O4 and ANO powers are mixed
without any solvent.
It can be seen from above TG analysis that there is no weight loss above 500 °C when Li2C2O4 and ANO
are mixed with water. Based on TG data, the dried Li2C2O4/ANO mixture (mixed with water) was calcined at
temperatures ranging from 475 to 800 °C for 3 h. Typical XRD patterns of LN475, LN500, LN600, LN700,
and LN800 powders are shown in figure 4. At lower temperature (475 °C), only the diffraction peaks of Nb2O5
and Li2CO3 powders are detected. It can be seen that single-phase LiNbO3 powders can be prepared at 500 °C.
All diffraction peaks of LN500-700 samples can be assigned to the hexagonal LiNbO3 structure (JCPDS 88-
0289). Each peak is assigned corresponding crystallographic plane by using its d-spacing value. The
broadening of XRD peaks at low calcination temperature (500 °C) may be due to smaller crystallite size. The
average crystalline size of these particles synthesized at 500 °C is estimated to be about 40 nm using the
Scherrer’s equation. The calculated lattice parameters is a=5.1466 Å and c=13.8546 Å, which are well
consistent with the values of the standard card a=5.148 Å and c=13.863 Å (JCPDS 88-0289). A few
differences exist in the calculated and standard lattice parameters, which may be attributed to the change in
preparative conditions. The intensity of the peaks is more intense with the increase of heat-treatment
temperature, which indicates that the size of the powders become larger. The results can be confirmed through
the SEM studies further. While in the case of LN800 powders, the crystal is identified as hexagonal
(Li0.996Nb0.005)Nb0.999O3 phase and all peaks in the diffractogram match well with reported values (JCPDS 85-
2453). It is well known that when temperature is high enough, lithium cation diffuses easily and even it can
evaporate from the raw materials. As a result, it may cause a scarcity of lithium in the obtained samples.
Therefore, it can be concluded that such high temperature is not benefit for the synthesis of pure stoichiometric
LiNbO3. The relative crystallinities of LN500-800 samples are estimated as about 87.1%, 91.4%, 98.4% and
100%. It can also be found that the crystallinity is clearly improved with the increase of temperature. It should
be noted that 500 °C is really a low temperature for the synthesis of pure nanocrystalline LiNbO3 powders by
an alternative solid-state method.
The controlled experiments have been performed to investigate the effect of different raw materials on
product compositions. Table 1 lists the XRD results of experimental products. As shown in table 1, except for
LiNbO3, small amounts of Nb2O5 crystal can be observed when Li2C2O4 and Nb2O5 were calcined at 500 °C.
Trace amount of LiNb3O8 crystal appears with the increase of temperature. Pure LiNbO3 powders could not be
obtained until 800 °C. It can be concluded that the reactive activity of niobic acid is much higher than that of
Nb2O5. Therefore, niobic acid plays an important role in the synthesis process. Similarly, pure LiNbO3 can not
be obtained when Li2CO3 reacts with ANO in temperature range from 500 to 800 °C. It is because that Li2C2O4
degrades to Li2CO3 in temperature range from 350 to 550 °C. The activity of Li2CO3 obtained by in situ
degradation of Li2C2O4 is higher than that of Li2CO3 reagent. It can be concluded that the single-phase LiNbO3
is produced in situ along with the degradation of Li2C2O4 and ANO.

Table 1 The XRD results of controlled experimental products. (a: small; b: trace)
Starting Materials Calcination Temperature Calcination Product Matching JCPDS no.
500 °C LiNbO3 + Nb2O5a 88-0289 and 27-1003
Li2C2O4 + Nb2O5
600 °C LiNbO3 + LiNb3O8b 88-0289 and 36-0307
(mixed with water
700 °C LiNbO3 + LiNb3O8b 88-0289 and 36-0307
and then dried)
800 °C LiNbO3 88-0289
500 °C LiNbO3 + LiNb3O8a 88-0289 and 36-0307
Li2CO3 + ANO
600 °C LiNbO3 + LiNb3O8a 88-0289 and 36-0307
(mixed with water
700 °C LiNbO3 + LiNb3O8b 88-0289 and 36-0307
and then dried)
800 °C (Li0.996Nb0.005)Nb0.999O3 85-2453

To well study the microstructure of the as-synthesized LiNbO3 powders, representative SEM images of
LiNbO3 powders are shown in figure 5. It can be observed that the LiNbO3 powders calcined at 500 °C are
composed of nanocrystals and the sizes vary between 20 and 70 nm (Fig. 5a). These fine particles are
coarsening into relatively big aggregates and display polyhedral morphology with the increase of temperature
(see figure 5b and c). The particles are about 100-300 nm in diameter when the calcination temperature is
increased to 700 °C. The grain sizes have been enlarged to about 400-800 nm for LN800 sample (Fig. 5d).
To check out the quality of the synthesized samples by this method, FTIR technology was employed to
characterize the products and the results were shown in figure 6. As shown in figure 6, the carboxylate bands

© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.crt-journal.org


Cryst. Res. Technol. 45, No. 9 (2010) 981

disappear when the precursor is calcined above 500 °C. Although peak centered at 1460 cm-1 is attributed to
the vibration of CO32–, it is very weak when LiNbO3 precursor is heated at 500 and 600 °C. Hence, it may be
concluded that the dwell time of 3 h is long enough to eliminate the CO32– impurity from the obtained samples.
Only when the calcination temperature reaches 700 °C, the CO32– is completely eliminated from the as-
synthesized powders. It had been reported that four absorption modes at 783, 670, 625 and 440 cm-1 were
observed for LiNbO3 crystal [11,32]. Liu et al. [17] reported three absorption modes at 693, 646 and 437 cm-1
for LiNbO3 were obtained by hydrothermal method. In our work, there were three absorption peaks at 666, 635
and 439 cm-1, which were in agreement with literature values [10,12]. It can be concluded that the LiNbO3
samples synthesized between 500 and 700 °C are high-quality without organic and CO32- impurities.

Fig. 5 SEM images of (a) LN500, (b) LN600, (c) LN700 and (d) LN800.

Fig. 6 FTIR spectra of the LN500-800 and Fig. 7 UV-Vis diffused reflectance spectra of LN500,
Li2CO3 powders. LN600 and LN700 samples.

The UV-Vis diffuse reflectance spectra of LN500-700 samples are shown in figure 7. The valence band and
conduction band of LiNbO3 are composed of O2p and Nb4d orbital, respectively. The band gap absorption edge
of LN500-700 samples is estimated as 322, 313 and 307 nm, respectively. The result shows that a blue-shift of
the absorption edge has been observed with the increase of calcination temperature. The shift may be due to the
particle size effect. The average band-gap absorption edge of them is determined to be 314.0 nm,
corresponding to the average band-gap energy of 3.95 eV. The absorption curves have a steep edge for all the
samples, which indicate that the UV light absorption is due to the band-to-band transition from O2p to Nb4d
directly (as shown in inset of figure 7).

4 Conclusion
In the alternative solid-state method, an ion-exchange reaction between NH4+ and Li+ occurs when Li2C2O4 and
ANO are ground in water. Pure LiNbO3 powders were obtained by heating Li[NbO(C2O4)2]·nH2O precursor at
500, 600, and 700 °C for 3 h, respectively. The raw materials of ANO and Li2C2O4 are excellent water-soluble
complex. The raw materials were adjusted to uniformly distribute by adding small amounts of water instead of
the ball milling process, which avoided the introduction of contamination into LiNbO3 powders. LiNbO3
powders obtained by this method are high quality by FTIR and XRD analysis. The average band gap
absorption edge of as-prepared LiNbO3 powders is determined to be 314.0 nm. The band gap energy of

www.crt-journal.org © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


982 Ting Ting Su et al.: Preparation of nanocrystalline lithium niobate powders

LiNbO3 at different calcination temperatures is estimated to be average of 3.95 eV from the onset of UV-Vis
diffuse reflectance spectra. It is found that typical environmental problems, as the use of toxic reagent,
presence of chlorine ion, or pollutant secondary products, have been successfully bypassed.

References
[1] D. Xue and K. Kitamura, Solid State Commun. 122, 537 (2002).
[2] R. S. Weis and T. K. Gaylord, Appl. Phys. A 37, 191 (1985).
[3] G. A. Magel, M. M. Fejer, and R. L. Byer, Appl. Phys. Lett. 56, 108 (1990).
[4] P. Baldi, M. P. Micheli, K. E. Hadi, S. Nouh, A. C. Cino, P. Aschieri, and D. B. Ostrowsky, Opt. Eng. 37, 1193
(1998).
[5] A. F. Benner, H. F. Jordan, and V. P. Heuring, Opt. Eng. 30, 1936 (1991).
[6] M. Niederberger, N. Pinna, J. Polleux, and M. Antonietti, Angew. Chem. 116, 2320 (2004).
[7] E. R. Camargo and M. Kakihana, Chem. Mater. 13, 1905 (2001).
[8] M. D. Aguas and I. P. Parkin, J. Mater. Sci. Lett. 20, 57 (2001).
[9] S. C. Navalea, V. Samuelb, and V. Ravi, Mater. Lett. 59, 2476 (2005).
[10] L. H. Wang, D. R. Yuan, X. L. Duan, X. Q. Wang, and F. P. Yu, Cryst. Res. Technol. 42, 321 (2007).
[11] H. C. Zeng and S. K. Tung, Chem. Mater. 8, 2667 (1996).
[12] M. N. Liu, D. F. Xue, and C. Luo, J. Alloys Compd. 426, 118 (2006).
[13] M. N. Liu, D. F. Xue, and K. Li, J. Alloys Compd. 449, 28 (2008).
[14] V. Samuel, A. B. Gaikwad, A. D. Jadhav, S. A. Mirji, and V. Ravi, Mater. Lett. 61, 765 (2007).
[15] E. R. Camargo and M. Kakihana, Solid State Ionics 151, 413 (2002).
[16] A. D. Li, J. B. Cheng, R. L. Tang, Q. Y. Shao, Y. F. Tang, D. Wu, and N. B. Ming, Mater. Res. Soc. Symp. Proc.
942, 0942-W04-03 (2006).
[17] J. Yu and X. Liu, Mater. Lett. 61, 355 (2007).
[18] C. An, K. Tang, C. Wang, G. Shen, Y. Jin, and Y. Qian, Mater. Res. Bull. 37, 1791 (2002).
[19] M. Liu and D. Xue, Mater. Lett. 59, 2908 (2005).
[20] G. Bhagavannarayana, R. V. Ananthamurthy, G. C. Budakoti, B. Kumarb, and K. S. Bartwalc, J. Appl. Cryst. 38,
768 (2005).
[21] V. T. Kalinnikov, O. G. Gromov, G. B. Kunshina, A. P. Kuz’min, E. P. Lokshin, and V. I. Ivanenko, Inorg. Mater.
40, 482 (2004).
[22] M. Liu, D. Xue, S. Zhang, H. Zhu, J. Wang, and K. Kitamura, Mater. Lett. 59, 1095 (2005).
[23] M. Liu and D. Xue, Solid State Ionics 177, 275 (2006).
[24] Y. Lin, H. Yang, J. Zhu, F. Wang, and H. Luo, Mater. Manuf. Processes 23, 791 (2008).
[25] K. Beck, H. Seyeda, U. Sulkowski, and A. Rosenkranz, US patent 7 241 911, 10 July 2007.
[26] R. C. Emerson and K. Masato, Chem. Mater. 13, 1905 (2001).
[27] T. Asai, E. R. Camargo, M. Kakihana, and M. Osada, J. Alloys. Compd. 309, 113 (2000).
[28] M. A. L. Nobre, E. Longo, E. R. Leite, and J. A. Varela, Mater. Lett. 28, 215 (1996).
[29] M. S. Kamat, T. Osawa, R. J. DeAngelis, Y. Koyama, and P. P. DeLuca, Pharm. Res. 5, 426 (1988).
[30] T. T. Su, Y. C. Zhai, H. Jiang, and H. Gong, J. Therm. Anal. Calorim. 98, 449 (2009).
[31] S. Wohlrab, M. Weiss, H. Du, and S. Kaskel, Chem. Mater. 18, 4227 (2006).
[32] E. J. Baran and I. L. Botto, J. Mater. Sci. Lett. 5, 671 (1986).

© 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.crt-journal.org

You might also like