Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Journal of the American Ceramic Society

Deconvoluting interrelationships between low-energy


vibrational modes and elastic properties in CaO - Al2O3
glasses

Journal: Journal of the American Ceramic Society

Manuscript ID JACERS-49233

Manuscript Type: Research Article


Fo
Date Submitted by the
22-Mar-2022
Author:

Complete List of Authors: Nasikas, Nektarios; Hellenic Army Academy Department of Mathematics
rP

and Engineering, Mathematics and Engineering Sciences


Kalampounias, Angelos; University of Ioannina, Chemistry; Institute of
Chemical Engineering and High-Temperature Processes, University of
Ioannina
ee

Keywords: Raman spectroscopy, Young-s modulus, calcium aluminate

Author-supplied Keyword: If
rR

there is one additional


keyword you would like to Boson Peak
include that was not on the
list, please add it below::
ev
iew

Journal of the American Ceramic Society


Page 1 of 30 Journal of the American Ceramic Society

1
2
3
4
5 Deconvoluting interrelationships between low-energy vibrational modes and elastic
6
7 properties in CaO – Al2O3 glasses
8
9
10
11
12 Nektarios K. Nasikas1,* and Angelos G. Kalampounias2,3,*
13
14
15
16 1Department of Military Studies, Division of Mathematics and Engineering Sciences, Hellenic Army Academy, GR
17
18 16673, Vari, Attica, Greece.
19
Fo
20 2Department of Chemistry, University of Ioannina, Ioannina, GR-45110, Greece
21
3University Research Center of Ioannina (URCI), Institute of Materials Science and Computing, Ioannina, Greece
22
23
rP

24
25
26
ee

27
28
29
Abstract
rR

30
31 By employing a non – conventional glass formation technique, namely the aerodynamic levitation
32
ev

33 (ADL) combined with CO2 laser melting, we have been able to synthesize a series of glasses in
34
35
the binary system xCaO – (1 – x)Al2O3, (x = 0.45, 0.50, 0.55, 0.60, 0.65, 0.70, 0.75 and 0.80).
36
iew

37
38 Utilizing Raman spectroscopy, we have systematically followed the compositionally induced
39
40 spectral changes in the low – frequency regime of the spectra, the so-called Boson Peak (BP)
41
42 region and unraveled interesting interrelationships between the low energy vibrational modes and
43
44
45 the elastic properties of Calcium Aluminate glasses. It was found that the BP frequency is not only
46
47 systematically dependent on altering CaO content by exhibiting a minimum close to the eutectic
48
49 composition. Interestingly, the above-mentioned compositional dependence of the BP, which is
50
51
52
exhibited in the shift of the BP frequency versus the altering CaO content, is also strongly
53
54 correlated with its elastic properties and seems to follow similar patterns for the binary Calcium
55
56 Aluminate glassy system. The above observations point towards the estimation that the eutectic
57
58 1
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 2 of 30

1
2
3 point can serve as a strong indicator for various physicochemical phenomena taking place in the
4
5
6 solid – liquid transition and thus provide new insights in an effort to deconvolute the ambiguous
7
8 interrelationships occurring in the low vibrational frequency regime of these glasses.
9
10
11
12
13
* Corresponding authors:
14 NKN: Department of Military Studies, Division of Mathematics and Engineering Sciences,
15
16 Hellenic Army Academy, GR 16673, Vari, Attica, Greece; email: nasikas@sse.gr
17
18 AGK: Department of Chemistry, University of Ioannina, GR-45110, Ioannina, Greece.
19
Tel.: +30 26510 08439; e-mail: akalamp@uoi.gr
Fo
20
21
22
23
rP

Keywords: calcium aluminates glasses; Raman spectroscopy; low frequency vibrations; elastic
24
25
26 properties; Young’s modulus; sound velocities; Boson peak.
ee

27
28
29
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 2
59
60 Journal of the American Ceramic Society
Page 3 of 30 Journal of the American Ceramic Society

1
2
3 1. Introduction
4
5
6 Amorphous and glassy materials are amongst the most intriguing subjects of materials science
7
8 and engineering, mostly due to their unique physicochemical and structural characteristics
9
10 exhibited in the amorphous state.1 These characteristics have been thoroughly studied in the
11
12
13
previous decades, with a variety of experimental techniques and theoretical approaches.2-5 In this
14
15 context, oxide glasses, have a unique place as they are materials exhibiting a broad spectrum of
16
17 scientific interest, stretching from their geological importance6,7 to even their biocompatibility
18
19
applications for health-related issues.8,9
Fo
20
21
22 Calcium Aluminate (CA) glasses are amongst those possessing both the above-mentioned
23
rP

24 characteristics. They are important geological constituents of the Earth’s mantle and crust10 but
25
26 also important ingredients for biomaterials having a variety of applications in dentistry and
ee

27
28
29 elsewhere.11,12 More specifically, CA and CA based glasses have been extensively studied in the
rR

30
31 past in an effort to elucidate their structural characteristics with a variety of experimental
32
ev

33 techniques, such as Raman spectroscopy, multiple quantum nuclear magnetic resonance (MQ –
34
35
36
NMR) and XANES (X – ray absorption near edge structure) being the most notable.13-15 The vast
iew

37
38 majority of these works have provided valuable insights revealing important structure property
39
40 relations.13-17
41
42
Especially Raman spectroscopy has proven a very powerful tool in this effort and has provided
43
44
45 a very detailed and consistent picture regarding the structural characteristics of the existing Ca –
46
47 O and Al – O polyhedra in these glasses.15,18,19 The vibrational spectra of these corresponding
48
49 polyhedra are depicted as broad overlapping bands in the medium-to-high frequency spectral
50
51
52 regions. These bands constitute some very distinct vibrational signatures and have been attributed
53
54
55
56
57
58 3
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 4 of 30

1
2
3 to various AlOx polyhedra with coordination numbers varying from x = 4, 5 and 6, existing in this
4
5
6 binary CA glassy system.15
7
8 On the contrary, very little attention has been given to the low-frequency region, close to the
9
10 laser’s excitation line of the Raman spectra acquired from CA glasses, where the so-called Boson
11
12
13
peak arises. The low-frequency regime (below 150 cm-1) of the Raman spectra acquired from
14
15 glassy materials is dominated by the presence of a broad asymmetric band, originating by the
16
17 overlapping contributions of two major sources.
18
19
The first source is the vibrational contribution of the so called Boson Peak (BP), whose
Fo
20
21
22 intensity generally follows the temperature dependence arising by the Bose factor, namely n(ω,T)
23
rP

24 = [exp(ℏω/𝑘𝐵T) – 1]-1. The second source is attributed to the quasi – elastic (QE) factor that is
25
26
generally linked to the presence of a very fast relaxation process with a magnitude of picoseconds.
ee

27
28
29 The BP in general is manifested in most of the cases in the spectral region between 10 and 50
rR

30
31 wavenumbers, although some binary oxide glasses, such as Magnesium Silicates, occurring
32
ev

33 between the metasilicate and the orthosilicate compositions corresponding to Enstantite (MgSiO3)
34
35
36 and Forsterite (Mg2SiO4) respectively, have demonstrated BP frequency values in the spectral
iew

37
38 region between 90 and 100 wavenumbers.20
39
40 One of the first works aiming to elucidate the nature of the BP21 proposed that it originates
41
42
43 from light scattering in the low – frequency regime of the Raman spectra and more specifically
44
45 from acoustic modes induced by disorder exhibited in the glassy state. Similar efforts proposed a
46
47 model aiming to unify the experimental data produced by spectroscopic and diffraction
48
49
50
experiments of amorphous solids22,23 However, despite systematical efforts to investigate the
51
52 origin of the BP, it remains debatable and especially for CA has just started to be systematically
53
54 investigated.24
55
56
57
58 4
59
60 Journal of the American Ceramic Society
Page 5 of 30 Journal of the American Ceramic Society

1
2
3 Most studies, so far, that have dealt with the elucidation of structure property relations for
4
5
6 oxide glasses have looked systematically into the compositional dependence of the above-
7
8 mentioned linkage. Even less attention has been given to the linkage between the compositional
9
10 dependence of the Boson peak and the elastic properties of the corresponding glasses. CA glasses
11
12
13
are known to possess very interesting elastic properties.25,26
14
15 In this work, we have synthesized via aerodynamic levitation combined with CO2 laser
16
17 melting, a series of glasses in the binary xCaO – (1-x)Al2O3, (x = 0.45, 0.50, 0.55, 0.60, 0.65, 0.70,
18
19
0.75 and 0.80) system with x denoting the mole fraction. The aerodynamic levitation (ADL)
Fo
20
21
22 technique combined with CO2 laser melting has been proven a very versatile and quite reliable
23
rP

24 technique that can provide homogeneous glasses, while at the same time expanding the established
25
26 glass formation region due to the extremely high cooling rates achieved during synthesis.24,27,28
ee

27
28
29 This expansion in the established glass forming region, is extremely important as the explosion in
rR

30
31 the interest shown in glasses and amorphous materials due to their extensive use in technology
32
ev

33 related devices and applications, calls for an ever-evolving effort for novel glasses, with novel
34
35
36
compositions and properties.29,30
iew

37
38 Here we report evidence that shed light in the intriguing linkages between the low-frequency
39
40 vibrational modes in CA glasses, namely the Boson Peak behavior and how this behavior is
41
42
connected to the elastic properties of CA glasses. The above linkages will be discussed in the
43
44
45 framework of a systematic and very thorough change in the CaO content, of just 0.05 mole, in an
46
47 effort to deconvolute a ternary interrelationship between, the Boson Peak, the elastic properties
48
49 and composition changes in CA glasses.
50
51
52
53
54
55
56
57
58 5
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 6 of 30

1
2
3 2. Experimental
4
5
6 2.1 Previtrification treatment of the starting reagents and glass composition preparations
7
8 It is essential, before going into the vitrification process of each of the desired glass
9
10 compositions, to perform a series of “previtrification” processes in order to ensure that the prepared
11
12
13
batches are homogeneous and relieved of any kind of impurities. The first step is to dry the reagent
14
15 grade chemicals, namely CaO (Alfa Aesar, 99.95% purity) and Al2O3 (Alfa Aesar, 99.99% purity)
16
17 under vacuum at 150 0C for 24 h, removing this way any possible absorbed H2O from the
18
19
atmosphere’s existing moisture.
Fo
20
21
22 The second step is to carefully weigh and extensively mix the desired stoichiometric masses
23
rP

24 corresponding to CaO and Al2O3 respectively in order to prepare completely homogeneous


25
26 mixtures that will be used for glass synthesis. The as prepared batches with stoichiometries within
ee

27
28
29 the compositional range 0.45 ≤ x ≤ 0.80 corresponding to the binary system xCaO – (1-x)Al2O3,
rR

30
31 with x denoting the mole fraction, and compositional step of 0.05 mole fraction, can provide a
32
ev

33 very detailed picture about the changes induced by altering composition.


34
35
36
iew

37
38 2.2 Vitrification process of Calcium Aluminate glasses achieved with aerodynamic levitation
39
40 combined with CO2 laser melting
41
42
The vitrification process utilized in order to synthesize the glasses with compositions shown
43
44
45 in Table 1, was through a non – conventional technique, namely the aerodynamic levitation (ADL)
46
47 combined with CO2 laser. This technique has proven to be the “technique of choice” when it comes
48
49 to vitrify glassy compositions exhibiting either high melting temperatures30,31 or to expand the
50
51
52 glass formation region achieved so far with conventional melt quenching techniques.24,28 The ADL
53
54 technique is executed in two distinct steps for the “home made” levitator setup.32,33
55
56
57
58 6
59
60 Journal of the American Ceramic Society
Page 7 of 30 Journal of the American Ceramic Society

1
2
3 Firstly, small amounts, usually a few milligrams (<10 mg) of the glass compositions intended
4
5
6 for vitrification, still in fine grain powder form, are placed on a water-cooled copper hearth for
7
8 initial melting. The heating source used to melt the desired glass compositions is the 10.6 μm laser
9
10 line of a 240 W, water – cooled, continuous wave CO2 laser (Synrad Evolution series).
11
12
13
Subsequently, the CO2 laser was turned on and its intensity was gradually increased with a step of
14
15 about 10 % of its maximum power per minute. The maximum laser intensity reached at this stage
16
17 was limited to 60% of its maximum output. The laser beam was irradiating the desired glass
18
19
composition from above covering the full size of the sample, ensuring this way homogeneity of
Fo
20
21
22 the irradiated surface and excluding any unwanted reflections that could create temperature
23
rP

24 gradients. As the temperature is carefully increased, ensuring homogeneous melting of the


25
26 powdered sample, a liquid droplet (due to surface tension) starts to form that is “self-supported”
ee

27
28
29 by its own material. This guarantees the absence of any factor that could possibly contaminate the
rR

30
31 sample since there is complete absence of any kind of container. This characteristic is the main
32
ev

33 reason why the ADL technique can also be found in the literature as “containerless” technique.34
34
35
36
After the formation of the droplet, the CO2 laser beam is blocked by the laser’s shutter leading to
iew

37
38 rapid cooling of the melt. At this stage the solidified droplet is still opaque and polycrystalline, as
39
40 the presence of the powder material in the base of the droplet provides crystallization nucleus
41
42
throughout this procedure.
43
44
45 The second stage, involved in the ADL setup, is as follows. Initially, we exchange the water-
46
47 cooled coper hearth with a conical nozzle (CN) through which Ar gas flows. The CN has seven
48
49 engraved air flow channels in its body, namely the center one, with the rest laying in a honeycomb
50
51
52 shape around it, creating an “air-cage” that can “trap” and levitate stably the sample. The
53
54 polycrystalline, almost spherical, sample is placed in the center of the CN and the Ar gas flow is
55
56
57
58 7
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 8 of 30

1
2
3 gradually increased until stable levitation of the sample is achieved. The Ar gas flow is carefully
4
5
6 monitored through a mass-flow meter with the ability to control the flow in 0.1% intervals. The
7
8 next step involves the gradual increase of the CO2 laser’s intensity in order to melt the levitated
9
10 sample. This step is very crucial as the increase in temperature, if not properly monitored, might
11
12
13
lead to temperatures far above the liquidus line and thus inflict compositional alterations due to
14
15 evaporation. In order to avoid the above-described situation, the temperature of the levitated and
16
17 concomitantly melted sample was carefully monitored through an optical pyrometer.
18
19
When complete melt of the levitated sample occurred, the CO2 laser beam was suddenly
Fo
20
21
22 blocked by the laser’s shutter, attaining this way extremely high-cooling rates for the levitated
23
rP

24 sample and subsequently causing its vitrification. The synthesized glasses were spherical beads of
25
26 a few mm in diameter, completely transparent, colorless and homogeneous.
ee

27
28
29
rR

30
31 2.3 Acquisition of the Raman spectra for the corresponding Calcium Aluminate glasses.
32
ev

33 For all glasses synthesized via the above-mentioned procedure and had corresponding
34
35
36
compositions shown in Table 1, we acquired Stokes – side polarized (parallel polarization) and
iew

37
38 depolarized (crossed polarization) Raman spectra. The 514.5 nm laser line originating from an Ar+
39
40 laser (Spectra Physics. Stabilite 2017) was used as the excitation source and the collection of the
41
42
Raman scattered light was performed in a 900 geometry by a triple monochromator (Jobin – Yvon,
43
44
45 T – 64000) that had the ability to operate in a double subtractive mode, permitting this way the
46
47 recording of the Raman spectra very close to the excitation line. In order to receive a very high-
48
49 quality Raman spectra, in terms of resolution, we employed holographic gratings of 1800
50
51
52 grooves/mm and their optimized diffraction efficiency was set at 2 cm-1. A two-dimensional
53
54 charged coupled device (CCD) camera was used in order to analyze and detect the scattered light,
55
56
57
58 8
59
60 Journal of the American Ceramic Society
Page 9 of 30 Journal of the American Ceramic Society

1
2
3 which had a 1024×256 pixels array and was cooled down to 140 K by filing a proper container
4
5
6 with N2 in order to minimize thermal noise. In order to collect the Raman spectra, every bead was
7
8 placed on a proper stage with the ability to move in all three axes. The laser line was carefully
9
10 passed along the equatorial plane of each bead in order to minimize any stray light reflections and
11
12
13
at the same time passing through the full diameter of the bead enhances the collected signal.
14
15 In order to ensure that the whole procedure is systematic and reproducible, we performed
16
17 periodic calibrations with the aid of CCl4 which was sealed inside a quartz cell. This procedure
18
19
guarantees that the polarization ratio for both scattering geometries (VV and VH) is correct and
Fo
20
21
22 accounts for possible shifts of the monochromator’s gratings. The acquisition times for all spectra
23
rP

24 were several minutes and were chosen as to acquire a very high signal-to-noise ratio spectra
25
26 leading to the acquisition of high-quality Raman spectra.
ee

27
28
29 Similar procedures for acquiring Raman spectra from molten oxides and salts at high
rR

30
31 temperatures have been described in detail elsewhere35,36 and can provide a very detailed picture
32
ev

33 of the versatility of the ADL technique and its many uses.


34
35
36
iew

37
38 3. Results and discussion
39
40 The reduced low-frequency Stokes-side Raman spectra acquired by all synthesized glasses are
41
42
shown in Figure 1. The reduced low-frequency VV (RLFVV) representation of the Raman spectra,
43
44
45 shows the evolution of the BP with altering CaO content in the binary system xCaO – (1-x)Al2O3,
46
47 (x = 0.45, 0.50, 0.55, 0.60, 0.65, 0.70, 0.75 and 0.80) with x denoting the mol fraction. We restricted
48
49 our analysis in the 0-300 cm-1 spectral region, where BP dominates the spectrum.
50
51
52
53
54
55
56
57
58 9
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 10 of 30

1
2
3
4
5
6
7
8 BP
9 xCaO-(1-x)Al2O3
10
11
12
13
14
15
16 x
17
0.80
g(ω) C(ω) / ω2

18
19
Fo
20
21 0.75
22
23
rP

24
0.70
25
26
ee

27
28 0.65
29
rR

30
31
0.60
32
0.55
ev

33
34
0.50
35 Exprt VV
0.45
36
iew

fit
37
38
39 0 100 200 300 400
40
41
42 Raman shift [cm-1]
43
44
45 Figure 1. Reduced Raman spectra versus frequency in the Boson peak region for the VV
46
47 polarization geometry. Open circles and thick solid lines represent experimental points and fittings,
48
49
50
respectively.
51
52
53
54
55
56
57
58 10
59
60 Journal of the American Ceramic Society
Page 11 of 30 Journal of the American Ceramic Society

1
2
3 The BP dominates over low-lying vibrational bending modes the low-frequency regime of the
4
5
6 Raman spectra and its characteristic frequency is strongly depended on altering CaO content.24
7
8 If one considers the harmonic approximation for the Stokes-side Raman spectra that is
9
10 collected by amorphous materials, it was shown by Shuker and Gammon37 that the Raman intensity
11
12
13
can be written as follows:
14
15
16
17 𝐼𝛼𝛽 𝛼𝛽
𝑒𝑥𝑝(𝜔, 𝛵) = 𝐶 (𝜔)𝑔(𝜔)𝜔
―1
[𝑛(𝜔,𝛵) +1] (1)
18
19
Fo
20
21
22 In the above equation, the 𝐶𝛼𝛽(𝜔) factor denotes the photon – phonon (Raman) coupling
23
rP

24 coefficient, which represents the activity of the vibrational density of states (VDoS) and
25
26 ―1
ee

27 𝑛(𝜔,𝛵) = [𝑒𝑥𝑝 (ħ𝜔


𝑘𝑇 ) ― 1] is the Bose – Einstein factor.37 The superscript αβ denotes the two
28
29
rR

30 different polarization geometries VV and VH. Considering all the above, we have represented the
31
32 spectral characteristics exhibited by the BP shown in Figure 1 in the RLFVV form in an effort to
ev

33
34 quantitatively follow its behavior. Therefore, the vertical axis is calculated in the g(ω)C(ω)/ω2
35
36
iew

37 form. Advantages on the use of the reduced representation of the Raman spectra can be found
38
39 elsewhere.37,38
40
41 The fitting procedure for the RLFVV spectra shown in Figure 1 involved the use of a log-
42
43
44
normal function, in order to systematically follow the compositional dependence of the BP for
45
46 these glasses. The evolution of the BP frequency versus the altering CaO content can be seen in
47
48 Figure 2a. It is readily observed that the BP frequency for both scattering geometries, VV and VH,
49
50
follows almost identical paths exhibiting a monotonical decrease from ~100 cm-1 corresponding
51
52
53 to the x = 0.45 glass to ~80 cm-1 corresponding to the x = 0.65 glass. From that point on, an almost
54
55 monotonical increase in the BP frequency is observed reaching almost 85 cm-1 for the x = 0.80
56
57
58 11
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 12 of 30

1
2
3 glass. This crossover point is observed at a mole fraction for CaO close to the eutectic points
4
5
6 observed in the CaO – Al2O3 binary phase diagram.39 This composition, namely around the 65
7
8 mol% in CaO which lays close to a eutectic point in the binary Calcium Aluminate phase diagram,
9
10 seems to play a key role in deconvoluting the interrelationships between the low-frequency
11
12
13
vibrations, but also for the various physicochemical phenomena occurring in these glasses. A
14
15 somewhat similar trend with varying composition is exhibited in the enthalpy of fusion for the
16
17 CaO – Al2O3 binary system.39 This observation is quite interesting as it could reveal a missing link
18
19
between these two very important quantities in CA glasses.
Fo
20
21
22
23
rP

24
25
26
ee

27
28
29
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 12
59
60 Journal of the American Ceramic Society
Page 13 of 30 Journal of the American Ceramic Society

1
2
3
4
5
6
7 100
8 (a)
9 96

[cm-1]
10
11 92
12 VH
13 ΒPmax
ΩVV,
14
88
15 VV
16 84 VH
17
18 80
19 0.4 0.5 0.6 0.7 0.8
Fo
20
21
22
Sound velocities [m/s]

23 7500 (b)
rP

24
25
7000 longitudinal
26
ee

27
28
29 4000
rR

30
31
32 3500 transversal
ev

33
34 0.4 0.5 0.6 0.7 0.8
35
36 CaO content [mole fraction]
iew

37
38
39
40
41
Figure 2. (a) Composition dependence of the BP frequency for both polarization geometries VV
42
43 (red squares) and VH (black circles) (b) Composition dependence of the longitudinal (orange
44
45 rhombus) and transversal (blue stars) sound velocities in CA glasses.
46
47
48
49
50 The observed non-coincidence for the BP value in the VV and VH representations of the
51
52 spectra is quite significant. Analogous results have been observed in other glassy systems, such as
53
54 Sr – Mg Fluorophosphates.40 The significant non – coincidence between the polarized and
55
56
57
58 13
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 14 of 30

1
2
3 depolarized values corresponding to the BP frequency is attributed to the different function of C(ω)
4
5
6 corresponding to the VV and VH geometries.
7
8 Figure 2b shows the compositional dependence of the longitudinal and transversal sound
9
10 velocities in CA glasses reported in the literature.41,42 Both velocities decrease with increasing CaO
11
12
13
content following similar behavior in their rate of decrease. No abrupt changes are observed but
14
15 rather smooth and systematic. It is also readily observed that the transversal sound velocities and
16
17 the BP frequency shown in Figure 2 exhibit similar compositional dependence at least up to
18
19
x=0.70. The above finding indicates that the low-frequency vibrations observed in the Raman
Fo
20
21
22 spectra acquired from CA glasses originate from transversal phonon vibrations. The above link
23
rP

24 has also been shown to hold valid also in the case of fluorophosphate glasses40 and in other systems
25
26 both experimentally43 and theoretically.44
ee

27
28
29 It is known that by using the BP frequency, one can derive the Debye temperature, through the
rR

30
31 following equation:44
32
ev

33
34
35
36 𝜃𝐷 = ħ𝜔𝐷/2𝜋𝑘𝐵 (2)
iew

37
38
39
40 In the above equation, ħ and 𝑘𝐵 are the Plank’s and Boltzmann’s constants, respectively. Also,
41
42
43 𝜔𝐷 is the Debye frequency corresponding to the depolarized (VH) BP frequency. The Debye
44
45 temperature, 𝜃𝐷, calculated through equation (2) and the elastic properties are shown in Figure 3.
46
47
48 For the latter case, namely the derivation of the Debye temperature through the elastic
49
50 properties of these glasses, we made use of the following equation:
51
52
53
54
ℎ 3𝑃𝑁 1/3
55 𝜃𝐷 = ( (3)
56 𝑘𝐵)( 4𝑉𝜋 ) 𝑢𝑚
57
58 14
59
60 Journal of the American Ceramic Society
Page 15 of 30 Journal of the American Ceramic Society

1
2
3
4
5
6 In equation (3), V is the molar volume derived from the ratio between the effective molar
7
8 weight and density. Finally, N is the Avogardo’s number and P is the number of atoms in the
9
10 molecular formula. The rest of the symbols have their usual meanings.
11
12
13
The quantity 𝑢𝑚 in equation (3) is the mean sound velocity, which can be calculated using the
14
15 longitudinal 𝑢𝐿and transversal 𝑢𝑇 velocities, respectively, as shown in equation (4).
16
17
18
19 1
Fo
20 ―3
21
22
𝑢𝑚 = [3 (
1 2
𝑢3𝑇
1
+ 𝑢3 ]
𝐿
) (4)
23
rP

24
25
26
ee

27 It is readily observed that the Debye temperature derived from the BP frequency is
28
29 monotonically decreasing, while the Debye temperature derived from the elastic properties, seems
rR

30
31 to be initially decreasing in an almost linear manner, reaching a minimum around 65 mol% in CaO
32
ev

33
34
and then increases slightly in a linear manner again. The latter trend is almost identical to the trend
35
36 exhibited by the BP frequency with altering CaO content (see Figure 2a).
iew

37
38 The compositional dependence of the Debye temperature resembles that of the compositional
39
40
evolution of the corresponding longitudinal and transversal sound velocities as shown in Figure
41
42
43 2b. The latter can also be used as an indicator for the cross-linking density in these glasses since
44
45 Debye temperature is also strongly related to the heat capacity of glasses and thus reflecting their
46
47 cross-linking density.40
48
49
50
51
52
53
54
55
56
57
58 15
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 16 of 30

1
2
3
4
5
6
7 600
8
9
10 Debye temperature [K]
11
560
12
13
14
15
520
16
17
18 LFR
19 Elastic properties
480
Fo
20
21
22 0.4 0.5 0.6 0.7 0.8
23
rP

24 CaO content [mole fraction]


25
26
ee

27 Figure 3. Composition dependence of the Debye temperature. Orange half circles correspond to
28
29
the Debye temperature derived from the elastic properties, while blue squares correspond to the
rR

30
31
32 Debye temperature derived from the BP frequency.
ev

33
34
35
36
iew

37
The BP can also be correlated to another very important parameter in these glasses, namely the
38
39 mean atomic volume 𝑉𝑚, through the following equation:45
40
41
42
43 𝑀
44 𝑉𝑚 = 𝜌𝑛 (5)
45
46
47
48
49 where, M, ρ and n are the atomic weight, the density and the number of atoms that are involved in
50
51
52 the stoichiometric formula for a specific glass composition, respectively. The evolution of the BP
53
54
55
56
57
58 16
59
60 Journal of the American Ceramic Society
Page 17 of 30 Journal of the American Ceramic Society

1
2
3 maximum frequency for both scattering geometries VV and VH (𝛺𝑉𝑉,𝑉𝐻
𝐵𝑃𝑚𝑎𝑥 ) versus the mean atomic
4
5
6 volume 𝑉𝑚 are shown in Figure 4.
7
8
9
10 102
11
12 100 VV
13 VH
14 98
[cm-1]

15 96
16
17 94
18
VH

92
19
ΒPmax
ΩVV,
Fo
20 90
21
22 88
23
rP

86
24
25 84
26
ee

27 82
28 7.2 7.4 7.6 7.8 8.0 8.2 8.4 8.6 8.8 9.0
29
rR

30 mean atomic volume Vm [cm3/mol]


31
32
ev

33 Figure 4. Evolution of the BP maximum frequency for both scattering geometries VV (red circles)
34
35
36
and VH (black triangles) versus the mean atomic volume for CA glasses.
iew

37
38
39
40 The results show an astonishing and almost identical dependence of the maximum BP
41
42 frequency for both scattering geometries VV and VH from the mean atomic volume to the
43
44
45 analogous dependence from composition (see Figure 2a).
46
47 Another interesting issue is to investigate the compositional dependence of the bulk and the
48
49 Young’s modulus as this relation can reveal important information about the cross-linking density
50
51
52 in the CA glasses studied here. The above is shown in Figure 5 a and b, respectively.
53
54
55
56
57
58 17
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 18 of 30

1
2
3
4
5
6 100
7

Bulk modulus [GPa]


8 96 (a)
9
10
11 92
12
13 88
14
15
16
84
17
18 80
19
Fo
20 0.4 0.5 0.6 0.7 0.8
21
22 140
23
rP
Young's modulus [GPa]

24
25 (b)
26 120
ee

27
28
29
rR

30 100
31
32
ev

33
34 80
35
36
iew

37
0.4 0.5 0.6 0.7 0.8
38 CaO content [mole fraction]
39
40
41
42
43 Figure 5. a) Bulk modulus in relation to CaO content in CA glasses (Orange rhombus) b) Young’s
44
45 modulus in relation to CaO content in CA glasses (Black squares).
46
47
48
49
50 It is readily observed that both the bulk and the Young’s modulus follow a similar decreasing
51
52 trend with increasing CaO content. This decrease, especially for the bulk modulus with increasing
53
54 CaO content could be associated with the structural evolution of the various AlOy (y = 4, 5, 6)
55
56
57
58 18
59
60 Journal of the American Ceramic Society
Page 19 of 30 Journal of the American Ceramic Society

1
2
3 polyhedra existing in these glasses15 and their contribution towards reducing the rigidity of the
4
5
6 glass network upon addition of CaO, which presumably acts as a glass modifier in this system.
7
8 In this direction, it is also interesting to investigate the relation between the Poisson’s ratio
9
10 and the Debye temperature with composition, namely the CaO content. Figure 6 a and b shows
11
12
13
the above-mentioned dependence for Poisson’s ratio and Debye temperature, respectively.
14
15
16
17
18
19 0.36
Fo
20 (a)
Poisson's ratio

21
22 0.32
23
rP

24
25 0.28
26
ee

27
28
0.24
29
rR

30 0.20
31 0.4 0.5 0.6 0.7 0.8
32
ev

33 620
34
Debye temperature [K]

35 600
36
iew

580 (b)
37
38 560
39
40 540
41
520
42
43 500
44
45 480
46 460
47 0.4 0.5 0.6 0.7 0.8
48
49 CaO content [mole fraction]
50
51
52
53
Figure 6. Poisson’s ratio (a) and Debye temperature (b) in relation to CaO content in CA glasses
54
55
56 (crossed open squares).
57
58 19
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 20 of 30

1
2
3
4
5
6 The Poisson’s ratio seems to be following a somewhat increasing trend with increasing
7
8 CaO content, while the Debye temperature was found to decrease with increasing CaO content. It
9
10 has been shown that the Debye temperature and the Poisson’s ration can be proven very useful
11
12
13
tools in order to estimate the cross-linking density of the glass network.46 Values for the Poisson
14
15 ratio between 0.1 and 0.2 are indicative of high cross-linking, while values between 0.3 and 0.5
16
17 are indicative of the opposite, namely low cross-linking density.46Here, the reported values for the
18
19
CA glasses under investigation show a variation between 0.27 and 0.34. This can be considered as
Fo
20
21
22 indicative of a small cross-linkage transformation occurring in these glasses and thus the
23
rP

24 corresponding glass networks are characterized from low cross-linking density, respectively.
25
26 In previous works regarding binary alkali and alkaline earth silicate glasses, but also
ee

27
28
29 aluminosilicates, the above-mentioned dependence of the 𝛺𝑉𝑉,𝑉𝐻
𝐵𝑃𝑚𝑎𝑥 versus 𝑉𝑚 (see Figure 4) was
rR

30
31 found to be linear.47,48 We argue that for CA glasses or for glasses where there is no archetypal
32
ev

33
glass former involved and one of the two components plays the role of the glass former (here nor
34
35
36 CaO nor Al2O3 can from glasses independently, hence are not considered to be archetypal glass
iew

37
38 formers, like SiO2 for example) knowledge of the mean atomic volume can provide a very close
39
40 estimation of how the BP maximum is behaving in these glasses. This will lift significant
41
42
43 experimental difficulties in acquiring a high-quality BP in glasses, as its proximity to the elastic
44
45 scattering hinders its recording.
46
47 It is also worth exploring the dependence of the 𝛺𝑉𝑉,𝑉𝐻
𝐵𝑃𝑚𝑎𝑥 versus the elastic properties of these
48
49
50 glasses and most importantly the shear modulus. The sear modulus or rigidity G, that generally
51
52 shows a system’s resistance to shear can be calculated by equation (6):
53
54
55
56
57
58 20
59
60 Journal of the American Ceramic Society
Page 21 of 30 Journal of the American Ceramic Society

1
2
3 𝐺 = 𝜌𝑢2𝑠 (6)
4
5
6
7
8 The evolution of the 𝛺𝑉𝑉,𝑉𝐻
𝐵𝑃𝑚𝑎𝑥 in relation to G is shown in Figure 7. The relation between these
9
10
11 two quantities is very similar to the one observed in Figure 4, where a characteristic minimum for
12
13 both polarization geometries VV and VH is distinctly observed. This similarity is somewhat
14
15 expected, as both the mean atomic volume 𝑉𝑚 and shear modulus are depended on the anionic
16
17
18 interactions occurring between the structural units existing in these glasses.
19
In this direction, it also interesting to address the intensity of the BP as it is derived from the
Fo
20
21
22 fitting of the Raman spectra shown in Figure 1 and how it is dependent on the inverse shear
23
rP

24
25
modulus for both polarization geometries VV and VH. The results are presented in Figure 8.
26
ee

27
28
29
rR

30
31 100 VV
32 VH
ev

33
[cm-1]

34
95
35
36
iew
VH

37
ΒPmax
ΩVV,

38 90
39
40
41
42 85
43
44
45
46 80
25 30 35 40 45 50
47
48 Shear modulus G [GPa]
49
50
51
52
Figure 7. Evolution of the BP maximum frequency for both polarization geometries VV (red
53
54 circles) and VH (black triangles) versus the shear modulus for CA glasses.
55
56
57
58 21
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 22 of 30

1
2
3
4
5
6 In Figure 8, it is directly observed that the BP intensity is linearly correlated with the inverse sear
7
8 modulus (1/G).
9
10
11
12
13 24
14
15 20
[arb. units]

16
16 Pearson's r=0.993
17
18 VV
12
19
VH
Fo
20 8
21 linear fit
VH

22
23
rP
IVV,
ΒP

24 3
25
26
ee

2 Pearson's r=0.995
27
28
29
rR

30 1
0.020 0.025 0.030 0.035 0.040
31
32 1/G [GPa-1]
ev

33
34
35
36
Figure 8. BP intensity in relation to the inverse shear modulus for both polarization geometries
iew

37
38 VV (black squares) and VH (red circles).
39
40
41
42
The strong linear relation between the BP intensity and the inverse shear modulus (1/G) is
43
44
45 reflected in the respective Pearson’s coefficients, which have values equal to 0.993 for the VV and
46
47 0.995 for the VH polarization geometry. It is worth noting that there is a significant difference in
48
49 slopes and intercepts for the lines corresponding to VV and VH polarizations. This is another
50
51
52 indication of the transverse vibrational origin of the BP.
53
54
55
56
57
58 22
59
60 Journal of the American Ceramic Society
Page 23 of 30 Journal of the American Ceramic Society

1
2
3 Tanaka and Shintani,44 proposed a very simple relation between the BP intensity and the
4
5
6 elastic moduli as shown in the following equation:
7
8
9
10 1 1 1
11
𝐼𝐵𝑃~𝐾 + 𝐺 ≈ 𝐺 (7)
12
13
14
15
The quantities K and G in equation (7), represent the bulk and the shear modulus, respectively.
16
17
18 The bulk modulus, K, compared to the shear modulus, G, is larger so the inversed quantity 1/K is
19
Fo
20 very small compared to 1/G. Hence, the BP intensity in equation (7) is considered being only
21
22 proportional to the inverse shear modulus, 1/G. Analogous interrelationships between the BP and
23
rP

24
25 the elastic properties have been shown recently for polymer glasses.49
26
ee

27
28
29 4. Conclusions
rR

30
31
32
In this work we have systematically tried to deconvolute the interesting and overlapping
ev

33
34 interrelationships between the low-frequency vibrational modes and the elastic properties of CA
35
36
iew

glasses. For this purpose, we have successfully synthesized a large variety of glasses occurring in
37
38
the binary system xCaO – (1-x)Al2O3, (x = 0.45, 0.50, 0.55, 0.60, 0.65, 0.70, 0.75 and 0.80). The
39
40
41 above wide compositional range combined with a very narrow compositional step of just 0.05
42
43 mole provided a very systematic and thorough investigation of the above-mentioned
44
45 interrelationships.
46
47
48 By utilizing Raman spectroscopy, we have been able to systematically follow a well resolved
49
50 BP, which exhibits a strong dependence on altering CaO content. This trend, when taken together
51
52 with the sound velocities in CA glasses, provided strong evidence for the transversal origin of the
53
54
55
BP. The determination of the elastic properties combined with altering CaO provided a very clear
56
57
58 23
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 24 of 30

1
2
3 picture about the dependence owed to altering CaO content. The results suggested that the cross-
4
5
6 linking density as depicted in the Poisson’s ratio and the Debye temperature is somewhat low and
7
8 these glasses, who are also characterized by reduced rigidity of the glass network as shown by the
9
10 Young’s and bulk modulus, respectively.
11
12
13
Finally, we argue that the eutectic points are very important points, not only for glass
14
15 formation, but also for elucidating the various physicochemical phenomena occurring in
16
17 amorphous solids and possibly reveal interesting links between the low-frequency vibrations, and
18
19
the elastic properties of glasses.
Fo
20
21
22
23
rP

24 5. Acknowledgements
25
26 Professor Emeritus G. N. Papatheodorou and Dr. G. A. Voyiatzis are greatly acknowledged
ee

27
28
29 for providing access to experimental facilities.
rR

30
31
32
ev

33
34
35
36
iew

37
38
39
40
41
42
References
43
44
45 1. Varshneya AK, Mauro JC. Fundamentals of Inorganic Glasses. 3rd Ed. Elsevier; 2019.
46
47 2. Angell CA. Perspective on the Glass Transition. J. Phys. Chem. Solids. 1988;49(8):863 – 871.
48
49 https://doi.org/10.1016/0022-3697(88)90002-9
50
51
52 3. Stillinger FH. A topographic view of supercooled liquids and glass formation. Science.
53
54 1995;267(5206):1935 – 1939. https://doi.org/10.1126/science.267.5206.1935
55
56
57
58 24
59
60 Journal of the American Ceramic Society
Page 25 of 30 Journal of the American Ceramic Society

1
2
3 4. Greaves GN, Sen S. Inorganic glasses, glass – forming liquids and amorphizing solids. Adv.
4
5
6 Phys. 2007;56 (1):1 – 166. https://doi.org/10.1080/00018730601147426
7
8 5. Hemley RJ, Mao HK, Bell PM, Mysen BO. Raman spectroscopy of SiO2 glass at high pressure.
9
10 Phys. Rev. Lett. 1986;57(6):747 – 750. https://doi.org/10.1103/PhysRevLett.57.747
11
12
13
6. Navrotsky A. Progress and new direction in high temperature calorimetry. Phys. Chem.
14
15 Minerals. 1977;2:89 – 104. https://doi.org/10.1007/BF00307526
16
17 7. Shim S-H, Duffy TS, Shen G. Stability and structure of MgSiO3 Perovskite to 2300 –
18
19
Kilometer depth in Earth’s Mantle. Science. 2001;293(5539):2437 – 2440.
Fo
20
21
22 https://doi.org/10.1126/science.1061235
23
rP

24 8. Hoppe A, Güldal NS, Boccaccini AR. A review of the biological response to ionic dissolution
25
26
products from bioactive glasses and glass – ceramics. Biomaterials. 2011;32(11):2757 – 2774.
ee

27
28
29 https://doi.org/10.1016/j.biomaterials.2011.01.004
rR

30
31 9. Hench LL. The story of Bioglass®. J. Mater. Sci: Mater. Med. 2006;17(11):967 – 978.
32
ev

33 https://doi.org/10.1007/s10856-006-0432-z
34
35
36 10. Kojitani H, Akaogi M. Melting enthalpies of mantle periotite: calorimetric determinations in
iew

37
38 the system CaO – MgO – Al2O3 – SiO2 and application to magma generation. Earth Planet Sc
39
40 Lett. 1997;153(3-4):209 – 222. https://doi.org/10.1016/S0012-821X(97)00186-6
41
42
43 11. Dimitriadis K, Moschovas D, Tulyaganov DU, Agathopoulos S. Glass – ceramics in the CaO
44
45 – MgO – Al2O3 – SiO2 system as potential dental restorative materials. Int. J. Appl. Ceram. Tec.
46
47 2021;18(6):1938 – 1949. https://doi.org/10.1111/ijac.13836
48
49
12. Kaga M, Kobayashi M, Ishikawa K, Minamikawa H, Yawaka Y, Watari F. Cytotoxicity of
50
51
52 strengthened glass – ionomer cement by compounding short fibers with CaO – P2O5 – SiO2 –
53
54 Al2O3 glass. Nano Biomedicine. 2010;2(1):23 – 30. https://doi.org/10.11344/nano.2.23
55
56
57
58 25
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 26 of 30

1
2
3 13. Neuville DR, Cormier L, Massiot D. Al environment in tectosilicate and peraluminus glasses:
4
5 27Al
6 A MQ – MAS NMR, Raman, and XANES investigation. Geochim. Cosmochim. Ac.
7
8 2004;68(24):5071 – 5079. https://doi.org/10.1016/j.gca.2004.05.048
9
10 14. Neuville DR, Cormier L, Massiot D. Al coordination and speciation in calcium
11
12
13
aluminosilicate glasses: Effects of composition determined by 27Al MQ – MAS NMR and
14
15 Raman spectroscopy. Chem. Geol. 2006;229(1-3):173 – 185.
16
17 https://doi.org/10.1016/j.chemgeo.2006.01.019
18
19
15. McMillan P, Piriou B. Raman spectroscopy of calcium aluminate glasses and crystals. J. Non
Fo
20
21
22 – Cryst. Solids. 1983;55(2):221 – 242. https://doi.org/10.1016/0022-3093(83)90672-5
23
rP

24 16. Sebdani MM, Mauro JC, Jensen LR, Smedskjaer MM. Structure – property relations in calcium
25
26 aluminate glasses containing different divalent cations and SiO2. J. Non – Cryst. Solids.
ee

27
28
29 2015;427:160 – 165. https://doi.org/10.1016/j.jnoncrysol.2015.07.047
rR

30
31 17. Huang C, Behrman EC. Structure and properties of calcium aluminosilicate glasses. J. Non –
32
ev

33 Cryst. Solids. 1991;128(3):310 – 321. https://doi.org/10.1016/0022-3093(91)90468-L


34
35
36
18. McMillan P, Piriou B, Navrotsky A. A Raman spectroscopic study of glasses along the joins
iew

37
38 silica-calcium aluminate, silica-sodium aluminate, and silica-potassium aluminate. Geochim.
39
40 Cosmochim. Acta. 1982;46(11):2021 – 2037. https://doi.org/10.1016/0016-7037(82)90182-X
41
42
19. Daniel I, McMillan PF, Gillet P, Poe BT. Raman spectroscopic study of structural changes in
43
44
45 calcium aluminate (CaAl2O4) glass at high pressure and high temperature. Chem. Geol.
46
47 1996;128(1-4):5 – 15. https://doi.org/10.1016/0009-2541(95)00159-X
48
49 20. Kalampounias AG, Nasikas NK, Papatheodorou GN. Glass formation and structure in the
50
51
52 MgSiO3–Mg2SiO4 pseudobinary system: From degraded networks to ioniclike glasses. J.
53
54 Chem. Phys. 2009;131:114513. https://doi.org/10.1063/1.3225431
55
56
57
58 26
59
60 Journal of the American Ceramic Society
Page 27 of 30 Journal of the American Ceramic Society

1
2
3 21. Martin AJ, Brening W. Model for Brillouin Scattering in Amorphous Solids. Phys. Status
4
5
6 Solidi B. 1974;64(1):163 - 172. https://doi.org/10.1002/pssb.2220640120
7
8 22. Elliott SR. Medium-range structural order in covalent amorphous solids. Nature.
9
10 1991;354(6353):445 – 452. https://doi.org/10.1038/354445a0
11
12
13
23. Elliott SR. A unified model for the low-energy vibrational behaviour of amorphous solids.
14
15 Europhys. Lett. 1992;19(3):201 - 206. https://doi.org/10.1209/0295-5075/19/3/009
16
17 24. Nasikas NK, Kalampounias AG. Compositional dependence of the Boson peak and fractal
18
19
dimensionality in Calcium Aluminate glasses prepared by aerodynamic levitation and CO2-
Fo
20
21
22 laser melting. J. Phys. Chem. Solids. 2022;164:110621.
23
rP

24 https://doi.org/10.1016/j.jpcs.2022.110621
25
26 25. Yeganeh-Haeri A, Ho CT, Weber R, Diefenbacher J, McMillan PF. Elastic properties of
ee

27
28
29 aluminate glasses via Brillouin spectroscopy. J. Non-Cryst. Solids. 1998;241(2-3):200-203.
rR

30
31 https://doi.org/10.1016/S0022-3093(98)00804-7
32
ev

33 26. Kang E-T, Lee S-J, Hannon AC. Molecular dynamics simulations of calcium aluminate
34
35
36
glasses. J. Non-Cryst. Solids. 2006;352(8):725 – 736.
iew

37
38 https://doi.org/10.1016/j.jnoncrysol.2006.02.013
39
40 27. Kohara S, Suzuya K, Takeuchi K, Loong C – K, Grimsditch M, Weber J. K. R., et al. Glass
41
42
formation at the limit of insufficient network formers. Science. 2004;303(5664):1649 –
43
44
45 1652. https://doi.org/10.1126/science.1095047
46
47 28. Nasikas NK, Chrissantopoulos A, Bouropoulos N, Sen S, Papatheodorou GN. Silicate glasses
48
49 at the ionic limit: Alkaline – earth sub – orthosilicates. Chem. Mater. 2011;23(16):3692 –
50
51
52 3697. https://doi.org/10.1021/cm2012582
53
54
55
56
57
58 27
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 28 of 30

1
2
3 29. Durán A. Press Release: United Nations Approves 2022 as the International Year of Glass.
4
5
6 Glass Phys. Chem. 2021;47(731):731. https://doi.org/10.1134/S1087659621060353
7
8 30. Ballato J, Dragic PD. The uniqueness of glass for passive thermal management for optical
9
10 fibers. Int. J. Appl. Glass Sci. 2021;1-14. https://doi.org/10.1111/ijag.16543
11
12
13
31. Nasikas NK, Sen S, Papatheodorou GN. Structural nature of polyamorphism in Y2O3 – Al2O3
14
15 glasses. Chem. Mater. 2011;23(11):2860 – 2868. https://doi.org/10.1021/cm200241c
16
17 32. Kalampounias AG, Papatheodorou GN. Raman spectroscopic measurements of molten
18
19
ceramic materials at high temperature. In: P. C. Trulove, H. C. De Long, R. A. Mantz, G. R.
Fo
20
21
22 Stafford, M. Matsunago. eds. Proceedings of the 13th International Symposium on Molten
23
rP

24 Salts. Pennington, NJ: The Electrochemical Society; 2002:PV02–19:p.485.


25
26 33. Papatheodorou GN. The High Temperature Raman Spectroscopy. In: P. C. Trulove, H. C. De
ee

27
28
29 Long, R. A. Mantz, G. R. Stafford, M. Matsunago. eds: Proceedings of the 13th International
rR

30
31 Symposium on Molten Salts, Pennington, NJ: The Electrochemical Society; 2002: PV02–
32
ev

33 19:p. 2.
34
35
36
34. Weber JKR. The containerless synthesis of glass. Int. J. Appl. Glass Sci. 2010;1(3):248 – 256.
iew

37
38 https://doi.org/10.1111/j.2041-1294.2010.00026.x
39
40 35. Papatheodorou GN, Kalampounias AG, Yannopoulos SN. Raman Spectroscopy of High
41
42
Temperature Melts. In: M. Gaune- Escard, K. R. Seddon, eds: Molten Salts and Ionic Liquids:
43
44
45 Never the Twain?, John Wiley & Sons, NY; 2009 pp. 301–340.
46
47 36. Papatheodorou GN, Yannopoulos SN. Light Scattering from Molten Salts: Structure and
48
49 Dynamics. In: M. Gaune – Escard, eds: Molten Salts: From Fundamentals to Applications,
50
51
52 Kluwer Academic, Dordrecht; 2002:47 – 106.
53
54
55
56
57
58 28
59
60 Journal of the American Ceramic Society
Page 29 of 30 Journal of the American Ceramic Society

1
2
3 37. Shuker R, Gammon R. Raman-Scattering Selection-Rule Breaking and the Density of States
4
5
6 in Amorphous Materials. Phys. Rev. Lett. 1970;25:222.
7
8 https://doi.org/10.1103/PhysRevLett.25.222
9
10 38. Kalampounias AG. Measurements of fractal dimensionality in 0.1Cs2O–0.9TeO2 glass-
11
12
13
forming tellurite system by Raman spectroscopy. Physica B. 2011;406(4):921 – 925.
14
15 https://doi.org/10.1016/j.physb.2010.12.028
16
17 39. Jerebtsov DA, Mikhailov GG. Phase diagram of CaO – Al2O3 system. Ceram. Int.
18
19
2001;27(1):25 – 28. https://doi.org/10.1016/S0272-8842(00)00037-7
Fo
20
21
22 40. Mpourazanis P, Stogiannidis G, Tsigoias S, Kalampounias AG. Transverse phonons and
23
rP

24 intermediate-range order in Sr-Mg fluorophosphate glasses, Spectrochim. Acta A.


25
26 2019;212:363 – 370. https://doi.org/10.1016/j.saa.2019.01.024
ee

27
28
29 41. Yeganeh-Haeri A, Ho CT, Weber R, Diefenbacher J, McMillan PF. Elastic properties of
rR

30
31 aluminate glasses via Brillouin spectroscopy. J. Non-Cryst. Solids.1998;241(2-3):200 – 203.
32
ev

33 https://doi.org/10.1016/S0022-3093(98)00804-7
34
35
36
42. Kang E-T, Lee S-J, Hannon AC, Molecular dynamics simulations of calcium aluminate
iew

37
38 glasses. J. Non-Cryst. Solids. 2006;352(8):725 – 736.
39
40 https://doi.org/10.1016/j.jnoncrysol.2006.02.013
41
42
43. Pilla O, Caponi S, Fontana A, Goncalves JR, Montana M, Rossi F, et al. The low energy excess
43
44
45 of vibrational states in v-SiO2: the role of transverse dynamics, J. Phys. Condens. Matter.
46
47 2004;16:8519. https://doi.org/10.1088/0953-8984/16/47/006
48
49 44. Shintani H, Tanaka H. Universal link between the boson peak and transverse phonons in glass.
50
51
52 Nat. Mater. 2008;7:870 – 877. https://doi.org/10.1038/nmat2293
53
54
55
56
57
58 29
59
60 Journal of the American Ceramic Society
Journal of the American Ceramic Society Page 30 of 30

1
2
3 45. Soga N, Yamanaka H, Hisamoto C, Kunguri M. Elastic properties and structure of alkaline-
4
5
6 earth silicate glasses. J. Non-Cryst. Solids. 1976;22(1):67 – 76. https://doi.org/10.1016/0022-
7
8 3093(76)90008-9
9
10 46. Mpourazanis P, Stogiannidis G, Tsigoias S, Papatheodorou GN, Kalampounias AG. Ionic to
11
12
13
covalent glass network transition: Effects on elastic and vibrational properties according to
14
15 ultrasonic echography and Raman spectroscopy. J. Phys. Chem. Solids. 2019;125:43 – 50.
16
17 https://doi.org/10.1016/j.jpcs.2018.10.010
18
19
47. Nakamura K, Takahashi Y, Osada M, Fujiwara T. Low-frequency Raman scattering in binary
Fo
20
21
22 silicate glass: Boson peak frequency and its general expression. J. Ceram. Soc. Jpn.
23
rP

24 2013;121(1420):1012 – 1014. https://doi.org/10.2109/jcersj2.121.1012


25
26 48. Ando MF, Benzine O, Pan Z, Garden J-L, Wodraczek K, et al. Boson peak, heterogeneity and
ee

27
28
29 intermediate-range order in binary SiO2 – Al2O3 glasses. Sci. Rep. 2018;8:5394.
rR

30
31 https://doi.org/10.1038/s41598-018-23574-1
32
ev

33 49. Tomoshige N, Mizuno H, Mori T, Kim K, Matubayashi N. Boson peak, elasticity and glass
34
35
36
transition temperature in polymer glasses: Effects of the rigidity of chain bending. Sci. Rep.
iew

37
38 2019;9:19514. https://doi.org/10.1038/s41598-019-55564-2
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58 30
59
60 Journal of the American Ceramic Society

You might also like