Sunteți pe pagina 1din 6

Journal of Non-Crystalline Solids 473 (2017) 108–113

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids


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

Structure and crystallization behavior of 50CuO-xTiO2-(50-x)P2O5 MARK


(x = 5–20) glasses
Liyuan Zhang, Shiquan Liu⁎
School of Materials Science and Engineering, University of Jinan, Jinan 250022, Shandong, China

A R T I C L E I N F O A B S T R A C T

Keywords: Glass samples with molar compositions of 50CuO-xTiO2-(50-x)P2O5 (x = 5, 10, 15, 20) were prepared. The
CuO-TiO2-P2O5 glass crystalline phases, characteristic thermal temperature as well as crystallization kinetic parameters were in-
Depolymerization vestigated using X-ray diffraction (XRD) and differential thermal analysis (DTA) under non-isothermal condi-
Aggregation tions. The Fourier transformed infrared spectroscopy (FTIR) results show that the glass network depolymerizes
Crystallization
with the increase of TiO2. However, TiO2 has an aggregation effect on the glass network. As a result of the
depolymerization and aggregation effects, the glass transition temperature first increases and then decreases
with TiO2. In the meantime, the glass is easier to crystallize. The XRD measurements indicate that when the
content of TiO2 is more than 10 mol%, desired CuTi2(PO4)3 phase appears in the heat-treated samples. The
crystallization kinetics of the samples was studied using the Kissinger and Augis-Bennett models. It is evidenced
that the crystallization activation energy increases with TiO2, while surface and bulk crystallization mechanisms
dominate in the sample with 5 mol and 10–20 mol% TiO2, respectively.

1. Introduction TiO2 is used as a nucleating agent in silicate glasses [12]. However,


there is not much literature on binary TiO2-P2O5 glasses. Binary TiO2-
Phosphate glasses have potential applications in batteries, laser P2O5 glasses could be prepared by quenching method [13]. However, it
technologies and biology, etc. [1,2]. The basic structure unit of phos- was found that glass formation region in the binary TiO2-P2O5 system is
phate glasses is PO4 tetrahedron [3]. Various phosphate anions can be narrow. Alkaline, alkaline earth or transition metal oxides were added
formed by linking through covalent bridging oxygen atoms. Three-di- to improve the glass forming ability of titanium containing phosphate
mensional cross-linked network of phosphate glasses contains three system [14]. More profoundly, especially founctionalized titanium
bonging oxygen atoms and this structure may change into a polymer- phosphate glass materials have been developed. For example, Li2O-,
link metaphosphate which includes two covalent bridging oxygen CaO- and CuO-containing titanium phosphate glasses have been made
atoms. Polymer-link metaphosphate may further transform to pyr- into functional porous glass-ceramics via crystallization and subsequent
ophosphate which has only one bonding oxygen atom. The actual acid leaching procedures. Hosono et al. prepared porous glass-ceramics
structure of a phosphate glass depends on the O/P ratio in its compo- with CaTi4(PO4)6 and β-Ca3(PO4)2 phases from CaO-TiO2-P2O5 glass
sition [4,5]. and with Li1 + xTi2 − xAlx(PO4)3 and β-Ca3(PO4)2 phases from Li2O-
The poor chemical durability associated with the chain-like struc- CaO-TiO2-Al2O3-P2O5 glass [15,16]. The porous glass-ceramics with a
ture phosphate glasses limits the practical application of glasses in the skeleton of LiTi2(PO4)3 made from Li2O-CaO-TiO2-Al2O3-P2O5 glass
fields that conventional silicate glasses have been used [6]. There has showed excellent exchange ability with Ag+ ions [17,18], which led to
been great efforts to improve the chemical stability of phosphate glasses the formation of microporous materials with AgTi2(PO4)3 and Ti
by introducing suitable dopants, such as Al2O3, Fe2O3, MgO, CaO, etc. (HPO4)2·2H2O crystals. The latter materials exhibited bacteriostatic
[7–10]. These additions may not only enhance the chemical durability activity and ammonia gas adsorption capacity [19]. Special attention is
of the phosphate glasses but also can impart special functions to the given in this article to CuO-TiO2-P2O5 glasses, from which porous glass-
glasses and expand the glass application fields. For example, with ad- ceramics with a skeleton of NASICON-type crystal CuTi2(PO4)3 could be
dition of CaO, phosphate glass can act as bioactive glass and phosphate formed [20]. The prepared porous glass-ceramics showed catalytic ac-
glasses containing large amount of TiO2 have shown excellent photo- tivity in the oxidation of propene to acrolein [21]. It could also be used
catalytic properties after heat-treatment [10,11]. as a solid electrolyte to replace zirconia one in oxygen gas sensors [22].


Corresponding author at: School of Materials Science and Engineering, University of Jinan, West Campus, No. 336, West Road of Nan Xinzhuang, Jinan 250022, China.
E-mail addresses: liusq_ujn@hotmail.com (L. Zhang), mse_liusq@ujn.edu.cn (S. Liu).

http://dx.doi.org/10.1016/j.jnoncrysol.2017.08.003
Received 15 June 2017; Received in revised form 31 July 2017; Accepted 2 August 2017
Available online 09 August 2017
0022-3093/ © 2017 Published by Elsevier B.V.
L. Zhang, S. Liu Journal of Non-Crystalline Solids 473 (2017) 108–113

For the above phosphate glasses, their structure and crystallization indicates that sample Ti25 was devitrified. It will not be further studied.
behavior are vital to the special functions of the final derived materials. The FTIR spectra of the glass samples Ti5 to Ti20 are shown in
However, there still lacks a systematical study on the crystallization of Fig. 2. From the FTIR spectrum of sample Ti5, it can be seen that there
copper titanium phosphate glasses. In this work, the structure and are mainly five absorption bands at 1200 1080, 940, 760 and
crystallization behavior of 50CuO-xTiO2-(50-x)P2O5 (x = 5, 10, 15, 20) 525 cm− 1, respectively. The band at 1200 cm− 1 is associated with the
glasses have been studied. stretching vibration of PO2 groups [2,28,29]. The band shifts to a lower
wave number from Ti5 to T10, and disappears in the cases of Ti15 and
2. Experimental procedures Ti20. The band at 1080 cm− 1 is ascribed to the asymmetric stretching
vibrations of PeO bonds in non-bridging oxygen groups [2]. And this
Glasses with a composition (in mol%) of 50CuO-xTiO2-(50-x)P2O5 band shifts to lower wave numbers with increase of TiO2. The absorp-
(where x = 5, 10, 15, 20) were prepared using analytic grade reagents tion band at around 940 cm− 1 is related to the asymmetric stretching
of CuO, TiO2 and H3PO4 (Conc. 85%). These raw materials were put vibration of PeOeP linkages [30]. The absorption band at 760 cm− 1 is
into a Teflon beaker and mixed with 30 ml water to make a slurry [23]. attributed to the symmetric vibrations of PeOeP bonds [31]. The ab-
The slurry was evaporated and dried at 200 °C for 12 h. Then, the dried sorption band at around 525 cm− 1 can be ascribed to the deformation
mixture was melted in an electric furnace at 1250 °C for 1 h. Finally, the modes of PeO−(PO43 −) groups [32,33] and its intensity decreases with
melts were poured out and quenched into water to avoid the crystal- the increase of TiO2. The above results are indicative of the depoly-
lization of the melts during cooling process. The collected particles merization of the phosphate chains [34], suggesting that the integrity of
were finally dried at 100 °C in a furnace. The as-synthesized samples glass network decreases with the increase of TiO2. On the other hand,
were denoted respectively as Ti5, Ti10, Ti15 and Ti20 corresponding to there appears a new weak absorption band at 642 cm− 1 (see the arrow
the contents of TiO2. pointed band) in the cases of Ti15 and Ti20. This new band is attributed
The glass samples were thermally treated in a furnace with a to the TieO vibration [12]. Due to the large electric strength of tita-
ramping rate of 5 °C/min at the glass transition points for 1 h, then at nium ions, the aggregation effect of TiO2 on the glass network cannot
the crystallization peak temperatures for 2 h. The heat-treated samples be neglected [29].
were denoted as Tix-HT-T, where x stands for the content of TiO2 in the
parent glass and T represents the heat treatment temperature. 3.2. DTA analysis of the glasses
The Fourier transform infrared spectroscopy (FTIR) spectra of
samples were recorded on a Nicolet iS10 Spectrometer (Thermo, USA) The DTA curves of the glasses are shown in Fig. 3. The curves dis-
in the wave number range of 400–1350 cm− 1. The measurements were play endothermic glass transition points (Tg) at 461 °C, 488 °C, 504 °C
made on glass powder dispersed in KBr pellets. The crystalline phases in and 493 °C, respectively. The results show that the glass transition
the heat treated specimens were identified by X-ray diffraction analysis points of the samples first increases with the increase of TiO2 from 5 to
(XRD) carried out on a D8 ADVANCED diffractometer (Bruke, 15%mol. This is due to the increasing aggregation effect of TiO2 on the
Germany). The relative contents of the crystalline phases were esti- glass network and slow mobility of the large titanium ions, which lead
mated using the Matrix-flushing method [24]. to the increase in glass viscosity and consequently the rise in the
Differential thermal analysis (DTA) was performed on an HCT-1 transition point temperature [26]. With further increasing the content
(Henven, China) thermal analyzer with a heating rate of 20 °C/min in of TiO2 from 15 to 20 mol%, the Tg decreases, suggesting that the de-
air with an empty alumina crucible as reference. Powder samples with polymerization of the glass network became dominate.
particles sized 106–125 μm were applied in the DTA analysis. The glass The DTA curves show a general trend with obvious crystallization
transition points were determined using the tangent extrapolation of peaks before 1000 °C, above which the crystals become to melt. The
the onset of first endothermic peak on the DTA curves. To determine the first exothermic peaks of samples Ti5, Ti10 and Ti15 are located around
crystallization activation energy (E) of the glasses, heating rates of 5, 10 685 °C. However, the first exothermic peak of Ti20 shifts to a lower
and 15 °C/min were also applied in the DTA analysis. E is calculated temperature of 646 °C. In addition, from sample Ti5 to Ti10, the first
based on the following Kissinger eq. [25,26] crystallization peak becomes significantly sharper and the intensity
greatly increases with the content of TiO2. Further increasing TiO2 to 15
ln (Tp2 α) = (E R)(1 Tp) + C (1) and 20 mol%, the intensity of this peak does not change. The second
exothermic peaks of all the samples are very weak. The temperatures
where Tp is the crystallization peak temperature in the DTA curve; α,
for this peak of samples Ti5 and Ti10 are 800 °C and 813 °C, respec-
the heating rate; R, the universal gas constant and C, a calculation
tively. However, it shifts to a lower temperature of around 750 °C for
constant. As can be seen from Eq. (1), a plot of ln(Tp2/α) versus 1/Tp
Ti15 and Ti20. The third exothermic peak appears at around 985 °C for
will give a straight line, indicating that the data of ln(Tp2/α) and 1/Tp
samples Ti5, Ti10 and Ti15. However, there exists an obvious en-
have a linear relationship. E can then be obtained from the slope (E/R)
dothermic peak at 948 °C for sample Ti20, suggesting the melting of
of the straight line. In addition, the crystallization mechanism of a glass
crystals formed at the lower crystallization temperatures. The following
can be correlated with the Avrami parameter, n, which is obtained from
sharp and intense exothermic peak at 961 °C indicates a recrystalliza-
the eq. [27]
tion of the melted phase. The above results indicate that the crystal-
n = (2.5 ΔT )RTp2 E (2) lization of the glass samples generally became easier with the increase
of TiO2. The ease of crystallization of the glass is a result of both the
where ΔT is the width of the crystallization peak at half height in the depolymerization of phosphate glass network and the aggregation ef-
DTA curves. fect of TiO2.

3. Results and discussion 3.3. Crystallization of the glasses at different temperatures

3.1. XRD and FTIR analysis of the as-synthesized samples The XRD patterns of the heat-treated samples are shown in Fig. 4. It
is can be seen that the diffraction peaks become stronger with the in-
The XRD patterns of the as-synthesized samples are presented in crease of the heat treatment temperature for all the samples. For sample
Fig. 1. The patterns in Fig. 1(a) consist of broad halo peaks at low Ti5, the types of the crystalline phases do not change with the heat
diffraction angles, which confirm the amorphous nature of the as-syn- treatment temperature. There exist Cu2P2O7(JCPDF card No.71-2177),
thesized samples Ti5 to Ti20. However, the pattern in Fig. 1(b) Ti7O13 (JCPDF card No.71-0629), Cu3(PO4)2(JCPDF card No.80-0991),

109
L. Zhang, S. Liu Journal of Non-Crystalline Solids 473 (2017) 108–113

Fig. 1. XRD patterns of the as-synthesized samples with different


amounts of TiO2.

increase of TiO2 promoted the crystallization of CuTi2(PO4)3. It is also


noted from the XRD results that heat-treatments performed at lower
temperatures help to reduce the types of crystals formed. In addition, it
can be seen from Table 3 and Table 4 that the content of CuTi2(PO4)3
phase decreases with the increase of the heat-treatment temperature.
Therefore, the heat-treatment of samples containing 15–20 TiO2 at re-
latively low temperatures facilitated the crystallization of the Cu-
Ti2(PO4)3 phase.

3.4. Crystallization kinetics of the glasses

Considering the importance of the formation of desired crystalline


phases in the functionalization of CuO-TiO2-P2O5 glasses, discussed
below are the crystallization kinetics of the glass samples. The discus-
sion is only based on the first crystallization peaks in the DTA curves
obtained with different ramping rates. Fig. 5 presents the ln (Tp2/β)
against 1/Tp curves of the samples. The correlation coefficients of linear
Fig. 2. FTIR spectra of the glass samples with different amounts of TiO2.
fitting the data are 0.9950, 0.9915, 0.9894 and 0.8990 for samples from
Ti5 to Ti20, respectively. The calculated activation energy for the
crystallization and the Avrami parameter (n) are summarized in
Table 5.
It can be seen that the crystallization activation energy increases
with the content of TiO2. Since the XRD results demonstrated that all
crystallized samples exhibited multi-phase crystals, the crystallization
activation energy reflects the energy barrier that the overall crystal-
lization needs to overcome. However, Ti20 has the largest E value,
lowest crystallization temperature and highest amount of CuTi2(PO4)3
phase. The XRD results (see Tables 3 and 4) also suggest the crystal-
lization of CuTi2(PO4)3 phases in samples Ti15 and Ti20 are highly
sensitive to the heat-treatment temperature, while the other main phase
(Cu2P2O7 and Cu3(PO4)2 in the cases of Ti15 and Ti20, respectively)
varies much less with the increase of the heat-treatment temperature.
These results may suggest that the crystallization of CuTi2(PO4)3 phase
needs a relative low activation energy.
It is well known that the crystallization of glasses involves nuclea-
tion and crystal growth stages [35]. Nucleation can occur on the surface
Fig. 3. DTA curves of the glass samples with different amounts of TiO2.
or in the bulk of the glass, resulting in surface or bulk crystallization.
The Avrami parameter, n, based on the Eq. (2) can be an indication of
CuO(JCPDF card No.80-0076), Ti(PO3)3(JCPDF card No.82-1178), the crystallization mechanism of a glass. An n value close to 1 denotes a
TiO2(JCPDF card No.71-1167) phases. The relative contents of the surface crystallization while a value of 3 implies a bulk crystallization
identified crystalline phases are listed in Table 1. The data indicate that [26,36]. The Avrami parameters presented in Table 5 show that n first
Cu2P2O7 is the major phase and the content of Cu3(PO4)2, Ti(PO3)3 and increases significantly from 1.79 to 2.77 from Ti5 to Ti10, then it de-
TiO2 phases increases with the heat treatment temperature, while the creases slowly with the increase of TiO2. The results suggest that sample
Ti7O13 phase greatly decreases in Ti5-HT-985. As can be seen from Ti5 may have dominant surface crystallization along with bulk crys-
Fig. 4(b) the phase of Ti7O13 disappears in sample Ti10-HT (Table 2). tallization, while other samples may mainly have bulk crystallization.
Instead, a new phase which is indexed to CuTi2(PO4)3 (JCPDF card This variation in the crystallization mechanism with the increase of
No.84-1355) appears in sample Ti10-HT-982 with a content of 8.2%. TiO2 is a result of the depolymerization of the phosphate glass network
This phase significantly increases with the increase of TiO2 and be- and the aggregation effect of TiO2.
comes the second main phase in samples Ti15-HT-686 and Ti15-HT-760
and the main phase in sample Ti20-HT-646. These results imply that the

110
L. Zhang, S. Liu Journal of Non-Crystalline Solids 473 (2017) 108–113

Fig. 4. XRD patterns of the heat-treated samples.

Table 1
Relative contents (%) crystalline phases in Ti5-HT samples.

Sample Cu2P2O7 Ti7O13 Cu3(PO4)2 CuO Ti(PO3)3 TiO2

Ti5-HT-686 47.19 ± 0.05 28.40 ± 0.11 9.37 ± 0.08 7.47 ± 0.02 5.77 ± 0.04 1.80 ± 0.02
Ti5-HT-800 40.02 ± 0.03 33.37 ± 0.09 10.27 ± 0.06 5.88 ± 0.02 7.87 ± 0.04 2.59 ± 0.02
Ti5-HT-985 53.96 ± 0.04 6.08 ± 0.10 16.37 ± 0.07 6.63 ± 0.01 12.23 ± 0.04 4.37 ± 0.01

Table 2
Relative contents (%) crystalline phases in Ti10-HT samples.

Sample Cu2P2O7 Cu3(PO4)2 Ti(PO3) CuO TiO2 CuTi2(PO4)3

Ti10-HT-684 56.14 ± 0.07 17.60 ± 0.13 12.62 ± 0.08 8.12 ± 0.03 5.52 ± 0.03 –
Ti10-HT-813 50.35 ± 0.03 21.40 ± 0.06 16.91 ± 0.03 5.76 ± 0.02 5.58 ± 0.02 –
Ti10-HT-982 54.72 ± 0.03 15.52 ± 0.07 12.09 ± 0.03 5.41 ± 0.01 4.06 ± 0.02 8.20 ± 0.03

Table 3
Relative contents (%) crystalline phases in Ti15-HT samples.

Sample Cu2P2O7 CuTi2(PO4)3 CuO Cu3(PO4)2 Ti5O4(PO4)4

Ti15-HT-686 52.59 ± 0.08 39.96 ± 0.08 7.45 ± 0.03 – –


Ti15-HT-760 58.06 ± 0.07 33.90 ± 0.07 8.04 ± 0.03 – –
Ti15-HT-985 54.55 ± 0.03 7.23 ± 0.04 5.96 ± 0.02 20.42 ± 0.06 11.84 ± 0.04

Table 4
Relative contents (%) crystalline phases in Ti20-HT samples.

Sample CuTi2(PO4)3 Cu3(PO4)2 Cu2P2O7 CuO TiO2

Ti20-HT-646 43.15 ± 0.10 40.52 ± 0.18 9.76 ± 0.11 6.57 ± 0.04 –


Ti20-HT-748 37.25 ± 0.07 48.35 ± 0.14 10.41 ± 0.07 4.01 ± 0.04 –
Ti20-HT-961 10.35 ± 0.03 41.61 ± 0.05 32.19 ± 0.04 5.63 ± 0.02 10.22 ± 0.02

4. Conclusions that the glass network depolymerized with the increase of TiO2.
However, TiO2 has an aggregation effect on the glass network. It is
Glasses with nominal compositions of 50CuO-xTiO2-(50-x)P2O5 evidenced that the glass is easy to crystallize and the content of desired
(x = 5, 10, 15 and 20) have been successfully prepared by melting and CuTi2(PO4)3 phase in crystallized samples increases with the increase of
water quenching technique. The FTIR measurement results indicate TiO2. In the meantime, the dominant crystallization mechanism turns

111
L. Zhang, S. Liu Journal of Non-Crystalline Solids 473 (2017) 108–113

Fig. 5. ln (Tp2/β) versus − 1/Tp curves correspondent to the crystallization peak 1 of Ti5 (a) Ti10 (b) Ti15 (c) and Ti20 (d) glasses.

Table 5 [10] J. Fu, Photocatalytic properties of TiO2–P2O5 glass ceramics, Mater. Res. Bull. 46
(2011) 2523–2526.
The crystallization activation energy (E) and the Avrami parameter (n) of the glass
[11] R.C. Lucacel, O. Ponta, E. Licarete, T. Radu, V. Simon, Synthesis, structure, bioac-
samples.
tivity and biocompatibility of melt-derived P2O5-CaO-B2O3-K2O-MoO3 glasses, J.
Non-Cryst. Solids 439 (2016) 67–73.
Samples E (kJ/mol) n [12] D.P. Mukherjee, S.K. Das, The influence of TiO2 content on the properties of glass
ceramics: crystallization, microstructure and hardness, Ceram. Int. 40 (2014)
Peak 1 Peak 1 4127–4134.
[13] D.E. Harrison, F.A. Hummel, Reactions in the system TiO2-P2O5, J. Am. Ceram. Soc.
Ti5 364 ± 0.3 1.79 ± 0.07 42 (2006) 487–490.
Ti10 435 ± 0.1 2.77 ± 0.06 [14] A. Kishioka, Glass formation in the Li2O-TiO2-P2O5, MgO-TiO2-P2O5, and CaO-TiO2-
Ti15 442 ± 0.2 2.56 ± 0.05 P2O5 systems, Bull. Chem. Soc. Jpn. 51 (1978) 2559–2561.
Ti20 500 ± 0.4 2.39 ± 0.05 [15] H. Hosono, Z. Zhang, Y. Abe, Porous Glass-ceramic in the CaO-TiO2-P2O5 System,
72 (2005), pp. 1587–1590.
[16] H. Hosono, Y. Abe, Porous glass-ceramics with a skeleton of the fast-lithium-con-
ducting crystal Li1 + xTi2 − xAlx (PO4)3, J. Am. Ceram. Soc. 75 (1992) 2862–2864.
from surface crystallization to bulk crystallization. [17] H. Hosono, F. Tsuchitani, K. Imai, Y. Abe, M. Maeda, Porous glass-ceramics cation
exchangers: cation exchange properties of porous glass-ceramics with skeleton of
References fast Li ion-conducting LiTi2(PO4)3 crystal, J. Mater. Res. 9 (1994) 755–761.
[18] H. Hosono, Y. Abe, Silver ion selective porous lithium titanium phosphate glass-
ceramics cation exchanger and its application to bacteriostatic materials, Mater.
[1] P. Hejda, J. Holubová, Z. Černošek, E. Černošková, The structure and properties of Res. Bull. 29 (1994) 1157–1162.
vanadium zinc phosphate glasses, J. Non-Cryst. Solids 462 (2017) 65–71. [19] Kasuga Toshihiro, H. Nakamura, K. Yamamoto, A. Masayuki Nogami, Y. Abe,
[2] I. Jlassi, N. Sdiri, H. Elhouichet, Electrical conductivity and dielectric properties of Microporous materials with an integrated skeleton of AgTi2(PO4)3 and Ti
MgO doped lithium phosphate glasses, J. Non-Cryst. Solids 466-467 (2017) 45–51. (HPO4)2·2H2O crystals, Chem. Mater. (1998) 3562–3567.
[3] D.A. Magdas, N.S. Vedeanu, D. Toloman, Study on the effect of vanadium oxide in [20] K. Yamamoto, T. Kasuga, M. Nogami, Y. Abe, Preparation of porous glass-ceramics
calcium phosphate glasses by Raman, IR and UV–vis spectroscopy, J. Non-Cryst. with a skeleton of NASICON-type copper titanium phosphate crystal, Phosphor. Res.
Solids 428 (2015) 151–155. Bull. 6 (1996) 341–344.
[4] U. Hoppe, G. Walter, R. Kranold, D. Stachel, Structural specifics of phosphate [21] K. Yamamoto, Y. Abe, Enhanced catalytic activity of microporous glass-ceramics
glasses probed by diffraction methods: a review, J. Non-Cryst. Solids s 263–264 with a skeleton of NASICON-type copper(I) titanium phosphate crystal, Mater. Res.
(2000) 29–47. Bull. 35 (2000) 211–216.
[5] R.K. Brow, Review: the structure of simple phosphate glasses, J. Non-Cryst. Solids s [22] K. Yamamoto, T. Kasuga, M. Nogami, An oxygen sensor based on copper(I)-con-
263–264 (2000) 1–28. ducting CuTi2(PO4)3 glass ceramics, Appl. Phys. Lett. 73 (1998) 3297–3299.
[6] Y. Attafi, S. Liu, Conductivity and dielectric properties of Na2O-K2O-Nb2O5-P2O5 [23] K. Yamamoto, T. Kasuga, Y. Abe, Preparation of porous glass-ceramics with a
glasses with varying amounts of Nb2O5, J. Non-Cryst. Solids 447 (2016) 74–79. skeleton of NASICON-type crystal CuTi2 (PO4)3, J. Am. Ceram. Soc. 80 (1997)
[7] E. Metwalli, R.K. Brow, Modifier effects on the properties and structures of alu- 822–824.
minophosphate glasses, J. Non-Cryst. Solids 289 (2001) 113–122. [24] F.H. Chung, Quantitative interpretation of X-ray diffraction patterns of mixtures. I.
[8] S. Reis, M. Karabulut, D. Day, Chemical durability and structure of zinc–iron Matrix-flushing method for quantitative multicomponent analysis, J. Appl.
phosphate glasses, J. Non-Cryst. Solids 292 (2001) 150–157. Crystallogr. 7 (1974) 519–525.
[9] X. Li, H. Yang, X. Song, Y. Wu, Glass forming region, structure and properties of zinc [25] H.E. Kissinger, Reaction kinetics in differential thermal analysis, Anal. Chem. 29
iron phosphate glasses, J. Non-Cryst. Solids 379 (2013) 208–213. (1957) 1702–1706.

112
L. Zhang, S. Liu Journal of Non-Crystalline Solids 473 (2017) 108–113

[26] S. Liu, H. Liu, T. Wang, H. Huang, Structure and crystallization of (1 − x) the structure and properties of (50 − x/2)Na2O-xCuO-(50 − x/2)P2O5 glasses,
(33.33Li2O-33.33Fe2O3-33.33P2O5)−xB2O3 glasses, J. Non-Cryst. Solids 450 Mater. Chem. Phys. 84 (2004) 341–347.
(2016) 38–41. [33] S. Babu, A. Balakrishna, D. Rajesh, Y.C. Ratnakaram, Investigations on lumines-
[27] J.A. Augis, J.E. Bennett, Calculation of the Avrami parameters for heterogeneous cence performance of Sm3 + ions activated in multi-component fluoro-phosphate
solid state reactions using a modification of the Kissinger method, J. Therm. Anal. glasses, Spectrochim. Acta A Mol. Biomol. Spectrosc. 122 (2014) 639–648.
Calorim. 13 (1978) 283–292. [34] H. Maeda, T. Tamura, T. Kasuga, Experimental and theoretical Investigation of the
[28] K. Meyer, Characterization of the structure of binary zinc ultraphosphate glasses by structural role of titanium oxide in CaO-P2O5-TiO2 invert glass, J. Phys. Chem. B
infrared and Raman spectroscopy, J. Non-Cryst. Solids 209 (1997) 227–239. 121 (2017) 5433–5438.
[29] B. Tiwari, A. Dixit, G. Kothiyal, M. Pandey, S. Deb, Preparation and characterization [35] I. Dyamant, A.S. Abyzov, V.M. Fokin, E.D. Zanotto, J. Lumeau, L.N. Glebova,
of phosphate glasses containing titanium, Barc Newslett. 285 (2007) 167. L.B. Glebov, Crystal nucleation and growth kinetics of NaF in photo-thermo-re-
[30] B.H. Jung, D.N. Kim, H.-S. Kim, Properties and structure of (50 − x) fractive glass, J. Non-Cryst. Solids 378 (2013) 115–120.
BaO–xZnO–50P2O5 glasses, J. Non-Cryst. Solids 351 (2005) 3356–3360. [36] H. Yinnon, D.R. Uhlmann, Application of thermoanalytical techniques to the study
[31] P. Jozwiak, DTA, FTIR and impedance spectroscopy studies on lithium–iron–pho- of crystallization kinetics in glass-forming liquids, part I: theory, J. Non-Cryst.
sphate glasses with olivine-like local structure, Solid State Ionics 179 (2008) 46–50. Solids 54 (1983) 253–275.
[32] A. Chahine, M. Et-tabirou, M. Elbenaissi, M. Haddad, J.L. Pascal, Effect of CuO on

113

S-ar putea să vă placă și