You are on page 1of 6

Journal of Non-Crystalline Solids 576 (2022) 121248

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids


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

Relationship between the elastic properties and structure of


BaO-TiO2-GeO2-SiO2 glasses
Hirokazu Masai a, *, Yasuhiro Fujii b, Naoyuki Kitamura a, Masato Yamawaki c
a
Department of Materials and Chemistry, National Institute of Advanced Industrial Science and Technology, 1-8-31 Midorigaoka, Ikeda, Osaka 563-8577, Japan
b
Department of Physical Sciences, Ritsumeikan University, 1-1-1 Nojihigashi, Kusatsu, Shiga 525-8577, Japan
c
Research Institute for Measurement and Analytical Instrumentation, National Institute of Advanced Industrial Science and Technology, 1-1-1 Umezono, Tsukuba,
Ibaraki 305-8568, Japan

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

keywords: The macroscopic elastic modulus of a solid-state matter is generally affected by the packing density and average
Glass bond strength of the constituent atoms. However, because of the diversity of network connections in oxide
Elastic constants glasses, an understanding based on the network structure is also required for precise structural categorisation.
Raman spectroscopy
Here, we examine the relationship between the elastic modulus and structural parameters, viz., the correlation
Spectroscopy
Oxides
length of the Boson peak and cavity size calculated from the positron annihilation measurement, of BaO-TiO2-
GeO2-SiO2 glasses. The Poisson’s ratio correlates with the ratio of the longitudinal and transversal sound ve­
locities, and thermal stability parameter ΔT. The Boson peak frequencies of these glasses at the GeO2-rich region
exhibit an apparent plateau behavior, indicating that the rate of change in the sound velocity is similar to that in
the correlation length. Although the numerical calculation of the cavity in this glass by positron annihilation
spectroscopy is unavailable because of the D-electron of Ti cations, the τ1 value shows a similar compositional
dependence as the correlation length of the Boson peak. It is expected that a combination of elastic modulus and
inelastic light scattering of glasses will be helpful in understanding glass-forming melts.

1. Introduction The cavities, that is, voids, of glass are open spaces at the angstrom
scale in glass that are unoccupied by cations or anions. It is expected that
Elastic modulus is one of the characteristic mechanical properties of cavities are key factors not only for the elastic modulus, but also for
a matter in the solid state. The elastic modulus generally originates from transparency and the crystallisation of oxide glasses. Although elastic
the average bond strength and packing density of the constituents [1–3]. modulus is described using several parameters such as the Young’s
In contrast to crystalline structures, whose structure is well constructed modulus, shear modulus, and Poisson’s ratio, the relationship between
based on diffraction and simulation studies, the structures of glasses, these parameters and cavities in oxide glasses has not yet been well-
whose chemical compositions and physical properties are much more clarified. The remarkable difference between metallic glasses and
diverse than those of crystalline materials, are not well organised with a oxide glasses is the existence of cavities, that is, free volumes. In contrast
guideline. to metallic glasses, whose atomic packing is sphere-like, network con­
In the case of crystalline materials, whose structure is well studied by nections of oxide glasses are affected by the lone pair of electrons on
diffraction and simulation studies, the elastic properties can be precisely oxygen atoms. For a precise illustration of the cavities in glass, a three-
simulated theoretically [4–6]. On the other hand, because the chemical dimensional depiction of the network connection is necessary [7,8].
and physical properties of oxide glasses are more diverse than those of However, even with the three-dimensional description of the network,
crystalline materials, the categorisation of the elastic properties of oxide the quantitative analysis of the cavities in an oxide glass is not easy
glasses is less developed. We assume that it is not easy to categorise the because the cavities are defined by the characteristic threshold values of
elastic modulus of various kinds of oxide glasses, and that the difficulty each glass.
originates from the variety of networks formed with defects and cavities For the quantitative analysis of cavities in oxide glasses, we have
in the matrix. focused on positron annihilation measurements [9–11]. The

* Corresponding author.
E-mail address: hirokazu.masai@aist.go.jp (H. Masai).

https://doi.org/10.1016/j.jnoncrysol.2021.121248
Received 25 June 2021; Received in revised form 18 October 2021; Accepted 20 October 2021
Available online 13 November 2021
0022-3093/© 2021 Elsevier B.V. All rights reserved.
H. Masai et al. Journal of Non-Crystalline Solids 576 (2022) 121248

measurement of positron annihilation is one of the techniques that can also shown.
numerically evaluate cavities in insulators. For the evaluation of cav­ Fig. 1a shows the transverse sound velocity, VT and longitudinal
ities, the lifetime of the ortho-positronium (o-Ps), which is a tight sound velocity, VL as a function of SiO2 fraction. The shear modulus, G0
binding state of positron and electron, is used. This technique has and Young’s modulus, E0 of the BTGS glasses are calculated using the
already been adopted in the analysis of several glass systems for which following equations:
the information of cavities has not been correlated with other physical
G 0 = ρVT 2 (1)
parameters.
Among several glasses, we focused on BaO-TiO2-GeO2-SiO2 glasses, ( )/( ( 2 ))
ν = VL 2 − 2VT 2 2 VL − VT 2 (2)
wherein fresnoite Ba2Ti(Ge,Si)2O8 crystallites were precipitated by heat
treatment [12–15]. The thermal parameters, crystallisation behavior,
E0 = 2G0 (1 + ν) (3)
and luminescence properties of this glass system have already been re­
ported [12–14]. Although the stoichiometric chemical composition of where ρ and ν are the density and Poisson’s ratio, respectively. The re­
Ba2Ti(Ge,Si)2O8 crystallites are 40BaO-20TiO2-40(Ge, Si)O2 glass, the lationships between these sound velocities and the SiO2 fraction are
nucleation rates of fresnoite from stoichiometric glass are extremely almost linear. The compositional dependence of G0 and E0 are plotted in
higher than those of conventional crystal [15]. Therefore, Fig. 1b. Not only the two sound velocities but also the two elastic moduli
non-stoichiometric (Ge, Si)O2-rich chemical compositions were used to change depending on the SiO2 fraction.
prevent the precipitation of fresnoite nanocrystals in the glass. Based on Because it is not obvious that the Poisson’s ratio increases (or de­
the numerical data obtained by spectroscopic analyzes, we discuss the creases) depending on the chemical composition from Eq. (2), we
relationship between the elastic modulus and the cavities in these plotted the Poisson’s ratio, ν as a function of the SiO2 fraction, as shown
glasses. in Fig. 2a. The molar volumes of these glasses are also shown. With
increasing SiO2 fraction, both the Poisson’s ratio and molar volume
2. Experimental decrease non-linearly depending on the chemical composition. As the
Poisson’s ratio is an indicator of the deformation of the samples against
The 30BaO-15TiO2-(55− x)GeO2-xSiO2 (BTGSx) glasses were ob­ the external force, the result indicates that the deformation of the sample
tained by a conventional melt-quenching method using BaCO3 (purity: is suppressed by the addition of SiO2. The molar volume takes similar
99.95%), TiO2 (purity: 99.99%), GeO2 (purity: 99.99%), and SiO2 (pu­ values in both GeO2-rich and SiO2-rich regions, whereas a remarkable
rity: 99.999%) as precursors. These oxides were mixed and melted at change is observed between x = 27.5 (the specimen comprising of the
1550 ◦ C for 1 h in air. The details may be found in a previous report [12]. same amount of SiO2 and GeO2) and 40. Considering the molar volume
The ultrasonic velocities of the longitudinal (VL) and transverse (VT) of these BTGSx glasses, it is expected that differences in the bond
waves were measured at room temperature by the ultrasonic pulse-echo dissociation energy between Si–O and Ge–O affect the Poisson’s ratio.
method (DPR300, JSR Ultrasonics). The frequencies of the longitudinal Therefore, the Poisson’s ratio can be an indicator of the fragility of glass
transducer and transverse transducer were 10 and 5 MHz, respectively.
The Young’s modulus (E0), instantaneous shear modulus (G0), bulk
modulus (K0), and Poisson’s ratio (ν) were calculated according to
previously reported methods [16]. The errors in E0, G0 and K0 are less
than ±0.1 GPa. The back-scattering micro-Raman measurements were
performed using a single-frequency diode-pumped solid-state laser
oscillating at 532 nm (Oxxius, LCX-532S-300) and a custom-built mi­
croscope with ultra-narrowband notch filters (OptiGrate). The incident
laser was attenuated to 7 mW and focused using a 20x objective lens.
The scattered light collected by the same lens was analyzed using a
single monochromator (Jovin-Yvon, HR320, 1200 grooves/mm)
equipped with a charge-coupled device (CCD) camera (Andor,
DU420-OE). The results given in the latter section show the depolarised
Raman spectra. Positron annihilation lifetimes were measured using a
PSA TypeL-II system (Toyo Seiko Co., Ltd.) with the anti-coincident
system [17]. The 22Na source with a diameter of 15 mm was encapsu­
lated in a Kapton film. The accumulated counts for each sample were
107.

3. Results and discussion

The thermal properties of BTGSx glasses are shown in Table 1 along


with their densities. The parameter ΔT = Tx − Tg, which is defined as the
thermal stability parameter for the crystallisation of the glass [12], is

Table 1
Thermal properties of 30BaO-15TiO2-(55-x)GeO2-xSiO2 (BTGSx) glasses [12].
3
Tg/ ◦ C Tx/ ◦ C Tp/ ◦ C ΔT/ ◦ C Density/g cm−

BTGS0 667 742 756 75 4.82


BTGS13.75 684 785 807 101 4.58
BTGS25 698 804 838 106 4.37
BTGS30 702 809 849 107 4.29 Fig. 1. a Transverse sound velocity, VT and longitudinal sound velocity, VL as a
BTGS41.25 719 831 867 112 4.11
function of SiO2 fraction. b Shear modulus, G0 and Young modulus, E0 of BTGSx
BTGS55 734 845 898 111 3.85
glasses as a function of SiO2 fraction.

2
H. Masai et al. Journal of Non-Crystalline Solids 576 (2022) 121248

[18,19]. Duval et al. have proposed the categorisation of strong glasses


(ν < 0.2) and fragile glasses (ν > 0.25) using the Poisson’s ratio; the
former is discussed based on E0 and the latter is discussed based on G0
[18]. If we follow this categorisation, these BTGSx glasses belong to the
class of fragile glasses. As the strong–fragile categorisation could not be
made based on the Poisson’s ratio, the VL/VT value, which is proposed
for the evaluation of the fragility of glass-forming liquids by Duval et al.
[18] is plotted as a function of SiO2 fraction in Fig. 2b. The thermal
stability of the glass, ΔT is also shown for comparison. The average
values of VL/VT are approximately 1.83, which indicate that the glass
can be categorised as strong rather than fragile. [19] We suggest that
both the Poisson’s ratio and the VL/VT value are important to categorise
the nature of glass-forming liquids. There is roughly an inverse rela­
tionship between VL/VT and ΔT, indicating that ΔT can also be used for
the evaluation of the fragility of glass-forming liquids.
Hereby, we assume that some structural parameters of the glass
network also influence the Poisson’s ratio. To examine the numerical
parameters of BTGSx glasses, we performed two spectroscopic mea­
surements, Raman scattering and positron annihilation spectroscopy
(PAS). Using these techniques, numerical structural parameters, such as
Boson peak frequency fBP, correlation length, or cavity sizes, can be
obtained for the discussion of the elastic modulus.
Fig. 3a shows the Raman spectra of BTGSx glasses recorded with HV
polarisation. These spectra consist of five broad bands at 55, 300, 500,
750, and 1000 cm− 1. These five vibration modes are assigned to the
Boson peak that characterise amorphous materials [20–22], the vibra­
tion mode of Ba–O [23], symmetric stretching and bending vibrations of
(Ge,Si)–O–(Ge,Si) [23], combination of the penta-coordinated Ti–O*
and Ge–O–Ge vibrations, [22] and symmetric Si–O stretching vibrations
[24,25], respectively. The intensities of these bands change systemati­
cally depending on the Ge–Si substitution. Fig. 3b shows the normalised
Raman spectra at the Boson region. We confirmed the compositional
Fig. 2. a Poisson’s ratio (ν) and molar volume of BTGSx glasses as a function of dependence at both the lower (10–20 cm− 1) and the higher (70–150
SiO2 fraction. b VL/VT and ΔT of BTGSx glasses as a function of SiO2 fraction.
cm− 1) region of the Boson peak. Here, as the Boson peak is known to
show a log-normal line shape in the reduced spectrum obtained from

Fig. 3. a Raman spectra of BTGSx glasses recorded with the HV polarisation. b Normalised Raman spectra at the Boson region. c Boson peak frequency fBP as a
function of SiO2 fraction. d Correlation length ξT and ξL as a function of SiO2 fraction. ξT and ξL are calculated from the values of VT and VL, respectively.

3
H. Masai et al. Journal of Non-Crystalline Solids 576 (2022) 121248

dividing Raman intensity by f{n(f, T) + 1}, we fitted the Boson peak components. The first, second, and third components are attributed to
with a log-normal function, [26] the lifetimes of para-positronium (p-Ps), positron annihilated without
( ) the formation of Ps, and ortho-positronium (o-Ps), respectively [31].
A (lnf − μ)2 The decay curves for the BTGS glass from PAS are depicted in Fig. 4a.
Ired (f ) = exp − , (4)
f 2σ 2 The intensities are normalised for comparison. The inset shows the
enlarged decay curves of BTGSx glasses (x = 0, 30, and 55); they suggest
where f, A, eμ , and σ are the frequency shift, amplitude, median of the
that small deviations are observable for these glasses. For the calculation
log-normal distribution, and standard deviation of lnf, respectively. The
2
of the cavity sizes, we initially used a conventional decay model with o-
values of the fBP were calculated as eμ− σ using fitting results. The region Ps for insulators [32]. The results are shown in Table 2. The cavity sizes
selected for the fitting of fBP is from the lower frequency at half of SiO2-containing glasses, which are calculated using the τ3 values,
maximum to the higher one where the intensity is 0.85 times the become larger with increasing SiO2 fraction. However, the results are
maximum (Fig. 3b). Although the energy shift of fBP in the ternary unsuitable for quantification of cavity sizes, because these relative in­
K2O–Nb2O5-GeO-SiO2 glass was not clearly observed, [27] the fre­ tensities of the third components, I3, are much smaller than those of SiO2
quency shift of fBP of the BTGS glass is confirmed by the good glass, and the relative intensities of I2 and I3 are different independent of
signal-noise ratio of the Boson peaks. Fig. 3c shows fBP as a function of the chemical composition. Considering the cavity size of the standard
the SiO2 fraction. It seems that fBP is almost constant within the error SiO2 glass, and the previous report on the porosity of oxide glasses,
margin below the SiO2 fraction of 30 mol%. In contrast, a remarkable wherein the generation of o-Ps in glass is prevented by the D-electron of
frequency shift is observed in the BTGS glasses containing more than 30 impurities [33], it is expected that the size estimation using the con­
mol% of SiO2. It is assumed that the main network-forming oxide ventional decay model with o-Ps for insulators is not adoptable. It was
changes from GeO2 to SiO2 at ~27.5 mol%, which is half of the fraction reported that a part of Ti4+ in BaO-TiO2-GeO2 glasses was reduced to
of SiO2 and GeO2 (55 mol%). The decrease in fBP with increasing SiO2 Ti3+ during the melting of the material at 1550 ◦ C [34]. Although there
fraction was previously observed for the SiO2-rich xGeO-(1 − x)SiO2 is no clear evidence at present, Ti3+ species are expected to prevent the
glass (x ≤ 0.37) [28]. From the values of fBP, we calculated the corre­ formation of o-Ps. A previous report on the PAS of oxide glasses suggests
lation length, ξ, which is conventionally given by fBP ≃ V/ξ, where V is that two-dimensional angular correlation of the annihilation radiation
the sound velocity [26]. For the evaluation of ξ, the values of VT have (ACAR) measurement is useful for estimating the porosity of oxide
been used in the analysis of soft materials. However, it is natural that the glasses [33]. However, we believe that the superiority of ACAR data is
present glasses, which have E0 values of more than 75 GPa, should be not established due to following reasons. (1) There is an inflection point
treated as elastic materials, based on the VL, although the Poisson’s ratio of radius in GeO2-SiO2 glasses determined by ACAR, although such a
is more than 0.28. Therefore, we calculated both ξT and ξL, which are structural anomalous composition has not been clarified. (2) The volume
calculated using the VT and VL values, respectively. Fig. 3d shows ξT and change (shrinkage) of SiO2 glass induced by high pressure is much larger
ξL as functions of the SiO2 fraction. Both the values of ξ increase with than that of GeO2 glass, indicating that there are many cavities in SiO2
increasing SiO2 fraction, indicating the predominance of an averaged glass [35]. (3) The periodicity, which is calculated by the Q value of the
intermediate structure comprising various SiO4 connections. It should first sharp diffraction peak in neutron diffraction, of SiO2 glass is larger
be noted that the correlation length evaluated from the Boson peak than that of pure GeO2 glass [8]. This indicates that the average char­
originates from the glass network. Since the molar volume decreases acteristic length of the SiO4 network is longer than that of the GeO4
with increasing SiO2 fraction, the shift in the Boson peak of the SiO2-rich network, suggesting that the porosity size of SiO2 glass is larger than that
glass agrees with a denser packed structure, which is also observed for of GeO2 glass. Based on these reasons, a quantitative discussion by PAS
densified glasses [29]. From the viewpoint of the correlation length, the was performed herein. We assume that the τ1 values reflect the electron
longer the correlation length in SiO2-rich glasses is, the more ordered density in the cavities, and the single-component analysis of the τ1
structure in random network formed is [30]. As fBP exhibits a bending values is effective for evaluating the porosity of the BTGS glasses.
point, it can be stated that the change rate of the correlation length with Therefore, we used the fitting, in which the parameters except for τ1 of
respect to the SiO2 fraction differs for the GeO2-rich and SiO2-rich re­ the BTGS55 glass were used (the values of τ2 and τ3 are 0.369 and 2.532
gions. After the SiO2:GeO2 ratio exceeds 1:1, the SiO2 network appears ns, and the values of I1, I2, and I3 are 25.2, 74.3, and 0.52%, respec­
more strongly to sharpen the change in the slope of the correlation tively). The results are shown in Table 3. The τ1 values of the BTGSx
length. It seems peculiar that the values of fBP are almost constant below glasses in Table 3 are shown in Fig. 4b as a function of the SiO2 fraction.
30 mol% of SiO2, although the sound velocity and elastic modulus The increase in τ1 values along with the SiO2 fraction suggests that larger
changed linearly in the same region with relative changes of 7%. This rings were formed in the SiO2-rich glasses.
means that the correlation length, which is an indicator of the structural We speculate that there is no direct relationship between them
ordering of glass, increases in this compositional range (Fig. 3d). We because the origin of the cavity by PAS and that of ξ by Raman spec­
assume that the rate of change in the sound velocity is similar to that in troscopy are different. Because of the uncertainty of numerical values by
the correlation length, and it induces an apparent plateau behavior of fBP PAS, the analysis of glass has not been considered useful. However, it is
below 30 mol% of SiO2. expected that a comparison of these data with those of other glass sys­
On the other hand, we can obtain the characteristic length of glass by tems will provide further details for understanding this relationship. For
PAS. In contrast with the correlation length based on the Boson peak, the example, for the case of SiO2 melted at higher temperatures, the esti­
length evaluated by PAS correlates with the cavity in the materials. mation of cavity diameter from the lifetime of the positronium becomes
Therefore, it seems natural that the correlation length is larger than that possible [11]. We have recently confirmed that this approach can be
of the cavity volume. Remarkably, the detected size is mainly due to the applied to silicate- [36] and borate- [37] based glasses. In contrast, we
cavities preferred by positrons, and a clear relationship between these have found that the formation of o-Ps is also prevented in the ZnO-P2O5
parameters has not been established so far. However, undoubtedly, the glass [38]. According to these results, it is expected that there is a
cavity size obtained by PAS is one of the indicators of the structure of the threshold of defects for the generation of o-Ps. We assume that the
glasses and their compositional dependence related to a particular threshold for positronium generation might serve as another means of
structural characteristic on the nanometre scale. In the case of in­ categorising oxide glasses. However, the number of reports on the life­
sulators, such as conventional organic polymers [31], the observed times of the positronium in different kinds of oxide glasses is not very
decay curves originate from the lifetime of the positronium (Ps), the large, and there are currently no guidelines for determining the
bound state of a positron and an electron. It has been reported that threshold for all oxide glasses. It should be noted that the calculated
positronium decays can generally be deconvoluted into three diameter changed depending on the homogeneity of the samples.

4
H. Masai et al. Journal of Non-Crystalline Solids 576 (2022) 121248

Fig. 4. a Positron decay curves of BTGSx glasses. Inset shows the extended decay curves of BTGSx glasses (x = 0, 30, and 55). b Values of τ1 as a function of
SiO2 fraction.

used for the quantification of the characteristic length not only in simple
Table 2
glasses but also in multicomponent glasses.
Relative intensities (I1, I2, and I3) and their respective lifetimes (τ1, τ2, and τ3) of
BTGS glasses from three-component analysis of positron lifetime spectra without
restrictions.
CRediT authorship contribution statement

τ1 (ns) τ2 (ns) τ3 (ns) I1 (%) I2 (%) I3 (%)


Hirokazu Masai: Formal analysis, Investigation, Writing – original
BTGS0 0.242 ± 0.353 ± 2.621 ± 30.2 69.4 0.39 ± draft, Writing – review & editing. Yasuhiro Fujii: Investigation.
0.007 0.003 0.083 ± 3.7 ± 3.7 0.01 Naoyuki Kitamura: Investigation. Masato Yamawaki: Investigation.
BTGS13.75 0.251 ± 0.355 ± 2.226 ± 27.2 72.3 0.55 ±
0.010 0.004 0.058 ± 5.2 ± 5.2 0.02
BTGS25 0.211 ± 0.348 ± 2.491 ± 15.6 84.0 0.41 ± Declaration of Competing Interest
0.019 0.002 0.074 ± 1.7 ± 1.7 0.01
BTGS30 0.216 ± 0.351 ± 2.345 16.1 83.5 0.48 ±
±
The authors declare no competing financial interest.
0.008 0.002 0.061 ± 2.0 ± 2.0 0.01
BTGS41.25 0.299 ± 0.396 ± 2.361 ± 61 ± 38 ± 0.45 ±
0.008 0.002 0.100 11 11 0.03 Acknowledgments
BTGS55 0.250 ± 0.369 ± 2.532 ± 25.2 74.3 0.52 ±
0.008 0.003 0.065 ± 3.5 ± 3.5 0.02 This work was partially supported by the Japan Society for the
Promotion of Science Grant-in-Aid for Scientific Research (B) Number
18H01714. The author (H.M.) greatly acknowledges the fruitful dis­
Table 3 cussion with Prof. H. Mizuno (University of Tokyo).
Lifetime τ1 of BTGS glasses of positron lifetime spectra with the fixed parameters
(τ2 (0.369 ns), τ3 (2.532 ns), I1 (25.2%), I2 (74.3%), and I3 (0.52%)). The error References
bars of τ1 is ± 0.001 ns.
τ1 (ns) τ2 (ns) τ3 (ns) I1 (%) I2 (%) I3 (%) [1] A. Makishima, J.D. Mackenzie, Calculation of bulk modulus, shear modulus and
Poisson’s ratio of glass, J. Non Cryst. Solids 17 (1975) 147–157.
BTGS0 0.193 0.369 2.532 25.2 74.3 0.52 [2] T. Rouxel, Elastic properties and short-to medium-range order in glasses, J. Am.
BTGS13.75 0.212 0.369 2.532 25.2 74.3 0.52 Ceram. Soc. 90 (2007) 3019–3039.
BTGS25 0.213 0.369 2.532 25.2 74.3 0.52 [3] S. Inaba, S. Fujino, K. Morinaga, Young’s modulus and compositional parameters of
BTGS30 0.219 0.369 2.532 25.2 74.3 0.52 oxide glasses, J. Am. Ceram. Soc. 82 (1999) 3501–3507.
BTGS41.25 0.239 0.369 2.532 25.2 74.3 0.52 [4] P. Ravindran, L. Fast, P.A. Korzhavyi, B. Johansson, J. Wills, O. Eriksson, Density
BTGS55 0.250 0.369 2.532 25.2 74.3 0.52 functional theory for calculation of elastic properties of orthorhombic crystals:
application to TiSi2, J. Appl. Phys. 84 (1998) 4891–4904.
[5] Z. Wu, E. Zhao, H. Xiang, X. Hao, X. Liu, J. Meng, Crystal structures and elastic
properties of superhard IrN2 and IrN3 from first principles, Phys. Rev. B 76 (2007),
Therefore, if the diversity of the calculated diameters is observed, the 054115.
atomistic heterogeneity of the samples exists in the matrix. There is no [6] G.V. Sin’ko, N.A. Smirnov, Ab initio calculations of elastic constants and
doubt that PAS is a probing method that fails to clarify the glass struc­ thermodynamic properties of bcc, fcc, and hcp Al crystals under pressure, J. Phys.
Condens. Matter 14 (2002) 6989–7005.
ture when used individually. Therefore, we believe that our compre­ [7] Y. Onodera, S. Kohara, H. Masai, A. Koreeda, S. Okamura, T. Ohkubo, Formation of
hensive examination will inspire other studies on oxide glasses using a metallic cation-oxygen network for anomalous thermal expansion coefficients in
combination of other experimental methods. binary phosphate glass, Nat. Commun. 8 (2017) 15449.
[8] Y. Onodera, S. Kohara, S. Tahara, A. Masuno, H. Inoue, M. Shiga, A. Hirata,
K. Tsuchiya, Y. Hiraoka, I. Obayashi, K. Ohara, A. Mizuno, O. Sakata,
4. Conclusion Understanding diffraction patterns of glassy, liquid and amorphous materials via
persistent homology analyzes, J. Ceram. Soc. Jpn. 127 (2019) 853–863.
We evaluated several elastic properties of BaO-TiO2-(Ge,Si)O2 [9] R.N. West, Positron studies of condensed matter, Adv. Phys. 22 (1973) 263–383.
[10] R.A. Pethrick, Positron annihilation—a probe for nanoscale voids and free volume?
glasses to examine the correlation length of these glasses. The Boson Prog. Polym. Sci. 22 (1997) 1–47.
peak frequency and molar volume of these glasses exhibit similar ten­ [11] M. Ono, K. Hara, M. Fujinami, S. Ito, Void structure in silica glass with different
dencies, while the VL/VT and ΔT show the opposite tendency. The in­ fictive temperatures observed with positron annihilation lifetime spectroscopy,
Appl. Phys. Lett. 101 (2012), 164103.
flection of the Boson peak frequency at the intermediate composition [12] H. Masai, N. Iwafuchi, Y. Takahashi, T. Fujiwara, S. Ohara, Y. Kondo, N. Sugimoto,
suggests a change in the main network-forming oxides in these glasses. Preparation of crystallized glass for application in fiber-type devices, J. Mater. Res.
The correlation length estimated from the Boson peak shows a compo­ 24 (2009) 288–294.
[13] S. Ohara, H. Masai, Y. Takahashi, T. Fujiwara, Y. Kondo, N. Sugimoto, Space-
sitional dependence similar to the τ1 values calculated by PAS. Although selectively crystallized fiber with second-order optical nonlinearity for variable
we cannot correlate the ordered network suggested by the Boson peak optical attenuation, Opt. Lett. 34 (2009) 1027–1029.
with a cavity size estimated by PAS, structural analysis using PAS will be [14] T. Kato, G. Okada, N. Kawaguchi, H. Masai, Y. Yanagida, Scintillation properties of
BaO-TiO2-GeO2-SiO2 glass-ceramics, J. Non Cryst. Solids 501 (2018) 116–120.

5
H. Masai et al. Journal of Non-Crystalline Solids 576 (2022) 121248

[15] A.A. Cabral, V.M. Fokin, E.D. Zanotto, C.R. Chinaglia, Nanocrystallization of [28] B.E. Hubbard, J.J. BE, N.I. Agladze, A.J. Sievers, Optical activity of the boson peak
fresnoite glass. I. Nucleation and growth kinetics, J. Non Cryst. Solids 330 (2003) and two-level systems in silica-germania glasses, Phys. Rev. B 67 (2003), 144201.
174–186. [29] Y. Inamura, M. Arai, M. Nakamura, T. Otomo, N. Kitamura, S.M. Bennington, A.
[16] A. Kato, M. Nagao, A. Sakuda, A. Hayashi, M. Tatsumisago, Evaluation of young’s C. Hannon, U. Buchenau, Intermediate range structure and low-energy dynamics of
modulus of Li2S–P2S5–P2O5 oxysulfide glass solid electrolytes, J. Ceram. Soc. Jpn. densified vitreous silica, J. Non Cryst. Solids 293-295 (2001) 389–393.
122 (2014) 552–555. [30] Y. Onodera, S. Kohara, P.S. Salmon, A. Hirata, N. Nishiyama, S. Kitani, A. Zeidler,
[17] M. Yamawaki, Y. Kobayashi, K. Hattori, Y. Watanabe, Novel system for potential M. Shiga, A. Masuno, H. Inoue, S. Tahara, A. Polidori, H.E. Fischer, T. Mori,
nondestructive material inspection using positron annihilation lifetime S. Kojima, H. Kawaji, A.I. Kolesnikov, M.B. Stone, M.G. Tucker, M.T. McDonnell, A.
spectroscopy, Jpn. J. Appl. Phys. 50 (2011), 086301. C. Hannon, Y. Hiraoka, I. Obayashi, T. Nakamura, J. Akola, Y. Fujii, K. Ohara,
[18] E. Duval, T. Deschamps, L. Saviot, Poisson ratio and excess low-frequency T. Taniguchi, O. Sakata, Structure and properties of densified silica glass:
vibrational states in glasses, J. Chem. Phys. 139 (2013), 064506. characterizing the order within disorder, NPG Asia Mater 12 (2020) 85.
[19] V.N. Novikov, A.P. Sokolov, Poisson’s ratio and the fragility of glass-forming [31] R.A. Pethrick, Positron annihilation—a probe for nanoscale voids and free volume?
liquids, Nature 431 (2004) 961–963. Prog. Polym. Sci. 22 (1997) 1–47.
[20] V.K. Malinovsky, A.P. Sokolov, The nature of boson peak in Raman scattering in [32] R.N. West, Positron studies of condensed matter, Adv. Phys. 22 (1973) 263–383.
glasses, Solid State Commun. 57 (1986) 757–761. [33] K. Inoue, H. Kataoka, Y. Nagai, M. Hasegawa, Y. Kobayashi, Short and medium
[21] E. Duval, A. Boukenter, T. Achibat, Vibrational dynamics and the structure of range order in two-component silica glasses by positron annihilation spectroscopy,
glasses, J. Phys. Condens. Matter 2 (1990) 10227. J. Appl. Phys. 115 (2014), 204903.
[22] H. Shintani, H. Tanaka, Universal link between the boson peak and transverse [34] H. Masai, K. Hamaguchi, K. Iwasaki, Y. Takahashi, R. Ihara, T. Fujiwara, Effect of
phonons in glass, Nat. Mater. 7 (2008) 870–877. melt temperature on the structure of BaO–TiO2–GeO2 glass, Mater. Res. Bull. 47
[23] S.A. Markgraf, S.K. Sharma, A.S. Bhalla, Raman study of fresnoite-type materials: (2012) 4065–4070.
polarized single crystal, crystalline powders, and glasses, J. Mater. Res. 8 (1993) [35] P.S. Salmon, A. Zeidler A, Networks under pressure: the development of in situ
635–648. high-pressure neutron diffraction for glassy and liquid materials, J. Phys. Condens.
[24] Y. Su, M.L. Balmer, B.C. Bunker, Raman spectroscopic studies of silicotitanates, Matter 27 (2015), 133201.
J. Phys. Chem. B 104 (2000) 8160–8169. [36] H. Masai, H. Kimura, M. Akatsuka, T. Kato, N. Kitamura, T. Yanagida, 252Cf-
[25] H. Masai, Structure studies of BaO-TiO2-SiO2 Glass-Ceramics using 29Si MAS NMR induced luminescence of cerium-doped lithium silicate glasses, J. Lumin. 241
and Raman spectroscopy, Bull. Chem. Soc. Jpn. 91 (2018) 950–956. (2022), 118481.
[26] V.K. Malinovsky, V.N. Novikov, A.P. Sokolov, Log-normal spectrum of low-energy [37] H. Masai, T. Ohkubo, Y. Fujii, A. Koreeda, T. Yanagida, T. Ina, K. Kintaka,
vibrational excitations in glasses, Phys. Lett. A 153 (1991) 63–66. Examination of structure and optical properties of Ce3+‑doped strontium borate
[27] L.F. Santos, L. Wondraczek, J. Deubener, R.M. Almeida, Vibrational spectroscopy glass by regression analysis, Sci. Rep. 11 (2021) 3811.
study of niobium germanosilicate glasses, J. Non Cryst. Solids 353 (2007) [38] H. Masai, Y. Onodera, S. Kohara, T. Ohkubo, A. Koreeda, Y. Fujii, M. Koshimizu,
1875–1881. M. Yamawaki, Correlation between structures and physical properties of binary
ZnO–P2O5 glasses, Phys. Status Solidi B 257 (2020), 2000186.

You might also like