Open access peer-reviewed chapter - ONLINE FIRST

Dynamic Specific Heat and Glass Transitions

Written By

Seiji Kojima

Submitted: 14 June 2023 Reviewed: 21 June 2023 Published: 24 September 2023

DOI: 10.5772/intechopen.1002805

Innovative Heat Exchanger Technologies, Developments and Applications IntechOpen
Innovative Heat Exchanger Technologies, Developments and Applicat... Edited by Peixin Dong

From the Edited Volume

Innovative Heat Exchanger Technologies, Developments and Applications [Working Title]

Peixin Dong and Xin Sui

Chapter metrics overview

57 Chapter Downloads

View Full Metrics

Abstract

The dynamical properties such as fragility, and non-Debye behavior of glass-forming materials have been studied by the frequency-dependent dynamic specific heat. Kubo’s formula on the fluctuation-dissipation theorem defines dynamic specific heat using the correlation function of enthalpy fluctuations. Dynamic specific heat is important for analyzing and understanding various relaxation processes. The dielectric relaxation is caused only by polar atomic motions, while the enthalpy relaxation is caused by total degrees of freedom of atomic motions. This chapter introduces two experimental methods to measure dynamic specific heat: (1) temperature-modulated differential scanning calorimetry (MDSC) and (2) photoacoustic spectroscopy. The experimental results of the dynamical properties of glass transitions in oxide glasses with covalent bond network structures and hydrogen-bonded glass-forming materials are reviewed.

Keywords

  • dynamic specific heat
  • modulated DSC
  • photoacoustic spectroscopy
  • relaxation
  • fragility
  • glass transition

1. Introduction

When a simple liquid is cooled from a high temperature, it undergoes a liquid-solid phase transition and crystallizes below a melting temperature, Tm. In contrast, when a complex liquid is cooled down below Tm, it changes into a supercooled liquid, which is a nonequilibrium state. Upon further cooling, a liquid-glass transition occurs at a glass transition temperature, Tg and it changes into a nonequilibrium glassy state. Figure 1 shows the enthalpy of liquid phase (AB), supercooled liquid state (BD), glass state (DE), and crystal phase (CG) as a function of temperature. The enthalpy of a liquid phase and supercooled liquid state crosses to that of a crystal phase at the Kauzmann temperature [1], TK (F), which is an ideal glass transition temperature determined by the static condition. Another ideal glass transition temperature is the Vogel-Fulcher temperature [2], TVF, which is determined by the Vogel-Fulcher law on the main structure relaxation process. In most glasses TK is close to TVF.

Figure 1.

The enthalpy is plotted as a function of temperature. Above Tm only a liquid phase (AB) exists. Below Tm a simple liquid crystallizes and crystal phase (CG) appears. While, a complex liquid keeps a supercooled liquid state (BD) between Tm and Tg. Below Tg a glass state (DE) appears.

Upon cooling a supercooled liquid, the relaxation time of a main structural relaxation process increases and undergoes a liquid-glass transition into a glass state at, Tg, where the relaxation time is 100 s. Both supercooled liquid and glass belong to nonequilibrium states. Understanding the dynamical properties of glass-forming materials is very important to clarify the mechanism of a liquid-glass transition. Their structural relaxation processes have been extensively studied by dielectric and ultrasonic spectroscopies, which are responsible for dielectric and elastic relaxations, respectively [3, 4].

According to the linear-response theory, dynamic specific heat is defined by Kubo’s formula on the fluctuation-dissipation theorem using the correlation function of enthalpy fluctuations [5]. The dynamic-specific heat provides a general understanding on the relaxation phenomena. Because the enthalpy relaxation contains the total degrees of freedom of atomic motions including polar atomic motions. In contrast, dielectric relaxation contains only polar atomic motions, and other nonpolar atomic motions do not contribute to the dielectric relaxation [6]. Therefore, it is important to study the structural relaxation processes of glass-forming materials by the frequency-dependent dynamic specific heat [7].

Advertisement

2. Dynamic susceptibility

Various physical quantities show time-dependent responses when an external field is applied to a system to be studied. Such a time-dependent response in a nonequilibrium state is analyzed by a response function or a relaxation function in linear-response theory, assuming that an external field is too weak to induce a nonlinear response. In observing a time-dependent response of physical quantities in a frequency-domain, dynamical susceptibility shows a frequency dispersion. The frequency dispersions of complex dielectric constants and complex elastic constants have been measured to study dynamical processes in a nonequilibrium state.

2.1 Complex dielectric susceptibility

In nonequilibrium statistical physics, dynamic (or complex) dielectric susceptibility is defined by the fluctuation-dissipation theory on the fluctuation of electric flux density, D(t) [5]. In the classical limit, it holds that

εω=ε0+kBT0DtD¯D0D¯eiωtdt.E1

Here, kB is Boltzmann constant, D(t) is electric flux density as a function of time, is an expected value, and D¯is the average value of D(t).

Dynamic/complex dielectric susceptibility has been discussed by linear-response theory. When the response of electric flux density, D(t), is linearly proportional to the external field, E(t), dielectric flux density is given by,

Dt=εEt+ε0εtϕttEtdt.E2

Here ε, ε0, and ϕt are dielectric constant of high-frequency limit, static dielectric constant, and dielectric response function, respectively. Figure 2 shows the impulse response of D(t).

Figure 2.

(a) Application of impulse electric field with a peak height E(0) at t = 0. (b) Impulse response of D(t), where ϕ(t) is a response function and D(0) = ε0E(0).

By Fourier transformation from time-domain to frequency-domain, it holds that

Dω=0Dteiωtdt=εωEω.E3

Here, the complex dielectric constant, εω, is related to a dielectric response function, ϕt .

εω=ε+ε0ε0ϕteiωtdt,E4

Dielectric relaxation function, Ψt, is defined by

ϕt=ddtΨt.E5

By the substitution of Eq. (5) for Eq. (4), we obtain

εω=ε0+ε0ε0ΨteiωtdtE6

In the Debye model with a relaxation time τ [8], the relaxation function is given by

Ψt=exptτ.E7

The complex dielectric constant is given by

εω=ε+ε0ε1+iωτ.E8

Real and imaginary parts of complex dielectric constants are given by

εω=ε+ε0ε1+ω2τ2,ε"ω=ε0εωτ1+ω2τ2.E9

Figure 3 shows real and imaginary parts of complex dielectric constant with ε = 20, ε0 = 50, and τ = 1.6 × 10−9 s.

Figure 3.

Real and imaginary parts of complex dielectric constant with ε = 20, ε0 = 50, and τ = 1.6 × 10−9 s.

As the frequency increases from a low frequency, the real part gradually decreases from the low-frequency limit, ε0, and in the vicinity of ω = 1/τ remarkable decreases. For further increase in frequency, it approaches the high-frequency limit, ε. The imaginary part shows the maximum at ω = 1/τ and ε = (ε0ε)/2. In a liquid-glass transition, upon cooling from a high-temperature liquid phase, the relaxation time of structural relaxation becomes longer toward a liquid-glass transition temperature, Tg. In a dielectric measurement, the relaxation time is determined by the maximum frequency of the imaginary part of the dielectric constant by the equation ωτ = 1.

Figure 4 shows the temperature dependence of the observed imaginary parts of glass-forming propylene glycol with Tg = 173 K in the large frequency range from 1 mHz to 10 GHz [9]. The broadband dielectric spectra were measured by the combination of two kinds of measurement. The low-frequency range between 1 mHz and 10 MHz was measured by the Impedance/Gain-phase analyzer (Solartron SI 1260). The high-frequency range between 1 MHz and 10 GHz was measured by the time-domain reflectometry (TDR) system, which is the combination of HP54750A digital Oscilloscope and HP54754A Differential TDR module. Upon cooling from a high temperature, the peak frequency at 315 K in a GHz range remarkably decreases down to that in a mHz range. The structural relaxation time obeys the Vogel-Fulcher-Tammann law.

Figure 4.

The temperature and frequency dependences of the imaginary part of dielectric susceptibility of glass-forming propylene glycol.

In the Debye relaxation model, the equation of semi-circle holds for εωand εω.

εωε0+ε22+εω2=ε0ε22E10

The plot of εω versus εω is called a Cole-Cole plot as shown in Figure 5.

Figure 5.

Cole-Cole plot of a complex dielectric constant.

In the non-Deby case, the relaxation time has a distribution, and a complex dielectric constant is given by the summation of relaxation processes with a distribution function of relaxation time, g(τ):

εω=ε+ε0ε0gτ1+iωτ.E11

If the complex dielectric constant over a very large frequency is available, the distribution function can be determined by inverse-integral transformation. However, for the analysis of experimental results with a limited frequency range, the following empirical formulae with few fitting parameters have been used.

  1. Davidson-Cole equation [10]

    εω=ε+ε0ε1iωτγ,0<γ<1E12

  2. Cole-Cole equation [11]

    εω=ε+ε0ε1iωτα,0<α<1E13

  3. Haviliak-Negami equation [12]

    εω=ε+ε0ε1iωταλ,0<α,λ<1E14

In the time-domain, Kohlrausch-Wiliams-Watts (KWW) function has been used in non-Debye relaxation [13, 14]:

Ψt=exptτβ,0β1E15

The numerical study on the relation of parameters between time-domain KWW function and the frequency-domain Haviliak-Negami equation was reported [15].

lnτHNτKWW2.61β0.5exp3β,αγβ1.23.E16

Here, τHNand τKWWare the relaxation times of Haviliak-Negami equation and KWW function, respectively.

2.2 Dynamic specific heat

In the equilibrium statistical mechanics, a specific heat at a constant pressure Cp is defined by

Cp=HT,E17

where H is enthalpy. In this definition, Cp is time-independent and includes the contribution of the total degrees of freedom.

According to the nonequilibrium statistical mechanics, dynamic (or complex) specific heat is Cp(ω) is defined by the fluctuation-dissipation theorem [16, 17] using the correlation function of enthalpy fluctuations. In the classical limit, it holds that

Cpω=Cp0+iωVkBT20hth¯h0h¯eiωtdt,E18

where, h(t) is the time-dependent enthalpy at constant pressure, h¯ is the averaged h(t), and V is the volume.

The specific heat of nonequilibrium materials is generally not a constant on time. The specific heat of nonequilibrium states has been discussed using a response function ϕ(t). The frequency-domain enthalpy is defined by the change of enthalpy h(t) ≡ δH = H(t) − Heq, where Heq is the value at an equilibrium state, is given by

hω=0hteiωtdt=CpωTωE19

The dynamic-specific heat Cpω is related to the enthalpy response function ϕt .

Cpω=Cp+Cp0Cp0ϕteiωtdt,E20

where Cp is the value that equilibrates instantaneously, and Cp0 is the static specific heat in an equilibrium state.

The specific-heat spectroscopy was developed by Birge [6], using an alternating current calorimetry technique, is not generally in use in laboratories and requires expert skill to detect weak third harmonic signals. In fact, this method measures the product of specific heat and thermal conductivity, i.e., the thermal effusivity [18]. Photoacoustic spectroscopy was also applied to study the dynamic specific heat of glass-forming materials [7]. Recently, temperature-modulated DSC (MDSC) is commercially available, and measurement can be performed in routine DSC scan [19]. In this chapter, at first, MDSC and the application to liquid-glass transitions of oxide network glasses are described. Secondly, photoacoustic spectroscopy is described, and the application to liquid-glass transitions of alcohols is described.

Advertisement

3. Temperature-modulated differential scanning calorimetry

In the early 1990s, a new type of differential scanning calorimetry (DSC), modulated temperature DSC (MDSC), was developed by Reading et al. [20, 21]. It has been shown that MDSC is a powerful tool to investigate the dynamical properties of glass transitions [22, 23]. MDSC is the extended equipment of a standard DSC. A small sinusoidal temperature perturbation is superimposed on the linear temperature ramping of the standard DSC, as shown in Figure 6. The MDSC was analyzed by Schawe based on the linear-response theory [24, 25].

Figure 6.

The temperature-modulated DSC (MDSC) applies a small temperature fluctuation, generally sinusoidal temperature oscillation, to linear temperature ramping of conventional differential scanning calorimetry.

The calorimeter response is given by

dQdt=CpdTdt+ftT.E21

Here, dQ/dt, dT/dt, Cp, and ftT are heat flow, heating rate, specific heat, and kinetic component, respectively. This heating rate-dependent component is called reversing, and the other independent one is nonreversing. MDSC can simultaneously measure these two components and the total DSC heat flow.

In DSC, the differential heat flow (HF) was determined by the temperature difference between the sample and the reference. The resultant HF is analyzed by the deconvolution of the sample’s response into the underlying heating rate [26]. The raw resultant HF is averaged over the period of more than one modulation. After the averaged HF was subtracted from the resultant HF, the modulated contribution to the resultant HF is calculated by the discrete Fourier transformation (DFT) to obtain the amplitude of modulated contribution AHF, and the phase lag (ϕ) between the modulated contribution and modulation in the heating rate. The phase angle is corrected according to the calculation reported in Ref. [27]. The absolute value of Cp* (ω) is calculated by the amplitudes of modulated contribution AHF and the heating rate Aq as shown in Eq. (22). The real part Cp′ and imaginary part Cp″ are calculated by the phase lag(ϕ) from Cp*:

Cp=AHFMAq,E22

where M is the mass of the sample.

Modulated temperature differential scanning calorimetry (MDSC) measurements were performed by using TA instruments DSC2920 to obtain a thermal relaxation process in the vicinity of a glass transition temperature [26]. Figure 7 shows the absolute value of Cp* and phase lag(ϕ) observed in the vicinity of a glass transition temperature of glucose. It has been shown that MDSC can provide the frequency-dependence of the specific heat in the order of 0.01–0.1 Hz. The information obtained by MDSC determines the frequency-dependent specific heat Cp*(ω). Then, Cpω and Cpω are calculated using a phase lag ϕ,

Figure 7.

The curves of the absolute value of complex specific heat |Cp| and phase lag of glucose obtained by MDSC [26].

Cpω=Cpωeϕ=CpωiCp"ωE23
Cpω=Cpωcosϕ,Cp"ω=Cpωsinϕ,E24

where Cpω relates to an in-phase component of the modulation and Cp"ω relates to an out-of-phase component. The temperature dependence of real Cp and imaginary Cp" parts of glass-forming glucose is shown in Figure 8.

Figure 8.

The real and imaginary parts of dynamic specific heat in the vicinity of a glass transition temperature of glucose.

The dynamic heat capacity of sodium borate glasses was measured by DSC 2920 and DSC T-zero Q200 (TA Instruments, Tokyo, Japan) [28]. A glass sample of about 10.0 mg was put into an aluminum pan. The sample pan was heated over Tg with a linear heating rate of 0.5 or 1.0°C/min. The temperature modulation conditions are amplitude ±1 °C and the period between 20 and 200 s. As a purge gas, dry nitrogen gas flowed in a sample chamber with a flow rate of 20 ml/min. The frequency dependences of the real and imaginary parts of the dynamic heat capacity of xNa2O-(1 − x)B2O3 glass (x = 0.41) in the vicinity of glass transition temperature are shown in Figure 9 [28].

Figure 9.

Frequency dependences of the real and imaginary parts of the dynamic specific heat of xNa2O-(1 − x)B2O3 glass (x = 0.41) observed by T-zero DSCQ200.

In Figure 9, the peak temperature of Cp”(ω), Tg(ω), increases as the temperature modulation frequency increases. The α-relaxation time τα is related to the modulation period, and it was determined by

τα=1ω=P2π,E25

where P is the temperature modulation period. Figure 10 shows the Angell plot [29] on the observed values of τα, where the horizontal axis is normalized by static Tg, which was determined as the temperature when τα becomes 100 s. By the Angell plot, fragility index m is determined by

Figure 10.

Angell plot of xNa2O-(1 − x)B2O3 glasses with x = 0.11 and 0.41 Mol%. The horizontal axis is normalized by Tg. The straight lines are the fits by a least square method.

m=dlogταdTgTT=Tg.E26

The range of m changes from strong glass-forming liquids with m ≈ 16 to fragile glass-forming liquids with m ≈ 200 [30]. From the Angell plots of x = 11 and 41 mol% in Figure 10, the fragility index m was calculated by Eq. (26), corresponding to the steepness of the linear curve. Such a determination of fragility index using MDSC was also reported in lithium borate glasses [31], glucose, fructose [26], ethylene glycol-glucose aqueous solutions [32], and poly (propylene glycol)s [33].

The Na2O content dependence of the fragility index m of xNa2O-(1 − x)B2O3 glasses is shown in Figure 11 with the reference values summarized by Chryssikos et al. [34]. The values observed by the MDSC are in good agreement with the experimental result of viscosity. The agreement indicates the reliability of the MDSC in studying the fragility of glass-forming materials. It is found that the steepness increases as the Na2O content increases. This result indicates the change of the fragility from “strong” to “fragile” as the Na2O content increases below x = 0.45. According to the Vilgis model, the degree of fragility relates to the distribution of the coordination number of a key element. In the sodium borate glasses, the composition dependence of m and the variation of the coordination number of boron atoms have similarities. It indicates that the origin of the fragility can be related to the distribution of the coordination number of boron. A similar result was also reported in lithium borate glass using MDSC [31].

Figure 11.

Composition dependence of the fragility index m of xNa2O-(1 − x)B2O3 glasses. The values observed by the MDSC are plotted by solid circles. The reference values summarized by Chryssikos et al. [34] are plotted by open circles.

For the study on the non-Debye nature of structural relaxation, the Cole-Col plot of xLi2O-(1 − x)B2O3 glass (x = 0.40) is shown in Figure 12. The imaginary part, Cp″ is plotted against the real part Cp. Experimental conditions: heating rate 1°C/min, amplitude of modulation 1.0°C, period of modulation 100 s. This plot shows the temperature averaged non-Debye nature in the vicinity of a dynamic glass transition temperature. Circles denote the experimental results. The solid line denotes the fitted values by the HN equation. The values of parameters are α = 0.94 and γ = 0.67, and both parameters are less than 1.0, and non-Debye nature is observed [17].

Figure 12.

Cole-Cole plot of dynamic specific heat of 40 Li2O·60 B2O3. Circles denote the experimental results. The solid line denotes the fitted values by Haviliak-Negami equation (α = 0.94 and γ = 0.67).

The origin of non-Debye nature relates to the coexistence of the different types of intermediate structural units with the three- and four-coordinated boron atoms. In alkali borate glasses, the addition of Li2O leads to the formation of tetrahedral B∅4, where ∅ is bridging oxygen. Therefore, the heterogeneity of the intermediate structure occurs under a low composition range of Li2O. The distribution of intermediate structures causes the distribution of relaxation times and the non-Debye nature. Such a non-Debye nature was also observed by the Cole-Cole plot of complex heat capacity in sodium borate glass [28], fructose, and glucose [26].

Advertisement

4. Photoacoustic method

In 1880, the photophonic phenomena were discovered by Alexander Graham Bell. When a solid in an enclosed cell is illuminated by a rapidly modulated beam of sunlight shines, it is possible to hear an audible sound by a hearing tube attached to the cell [35]. Tyndall and Rontgen reported that an audible sound could also be emitted when a gas in an enclosed cell is illuminated with intermitted Drummond’s limelight [36, 37]. Bell subsequently experimented with the production of sound by radiant energy with a variety of solids, liquids, and gases [38]. After 100 years had passed, the interest in photoacoustic effects and related experimental techniques increased. The first conference on photoacoustic effect took place in February 1981 in Bad Honnef in Germany [39]. Spectroscopy and detection of minute concentrations, monitoring chemical reactions, energy conversion processes, and calorimetric applications to phase transitions were extensively discussed. By using a finely focused laser beam, photoacoustic imaging (PAI) is also possible, and a liquid-solid interface was observed in the vicinity of a melting temperature [40]. Recently, PAI is an emerging noninvasive imaging modality. PAI provides better resolution than pure ultrasonic imaging and deeper penetration than pure optical imaging [41].

Rosencwaig and Gersho reported a quantitative derivation for the acoustic signal of solid and semisolid matter in a photoacoustic cell in terms of the optical, thermal, and geometric parameters of the system [42]. A one-dimensional model of the heat flow in the cell resulting from the absorbed light energy was calculated.

A standard cylindrical photoacoustic cell is shown in Figure 13 by the cross-sectional view. A sample is located between backing material and gas. Here, the thermal diffusion length of a sample, μs, the thermal diffusion length of gas, μg, the thermal conductivity (cal/cm s °C), ki, the density (g/cm3), ρi, the specific heat (cal/g°C), Cpi, the thermal diffusivity i (cm2/s) αi = ki/ρiCi, the thermal diffusion coefficient (cm), ai = (ω/2αi)1/2; the thermal diffusion length (cm) μi = 1/ai, where subscript i denotes s, g, and b for the solid, gas, and backing material, respectively. The modulated frequency of the incident light to a cell is ω in radians per second.

Figure 13.

One-dimensional model of photoacoustic effect in gas-microphone method on the optically thick sample (Ls > μβ > μs), where μs and μg are thermal diffusion lengths of a sample and gas, respectively. μβ = 1/β is the optical absorption length.

The thermal diffusion equation in the solid, taking into account the distributed heat source, can be written by

2ϕx2=1αsϕtAexpβx1+expjωt,forLsx0.E27

Here, A=βI0η/2ks, ϕ and η are temperature and efficiency, respectively.

The full expression for the incremental pressure,

δPt=Qexpjωtπ/4,whereQ=γP0θ/2lgagT0.E28

The full expression for δP(t) is complicated for interpretation. However, it is understandable by considering typical conditions. For optically opaque (μβ = 1/β ≪ 1) and thermally thick (μs ≪ l; μβ < μs) solids, it holds that

Q1jμs2agμsksY,whereY=γP0I022lgT0.E29

This case is shown in Figure 14(a) [42, 43].

Figure 14.

(a) Optically thin and thermally thick (μβ < μs < Ls), (b) optically very thin and thermal thick (μβ ≪ μs < Ls). The modulated light is absorbed in an opaque film, and thermal waves are generated at interfaces with adjacent gas and a sample.

The photoacoustic signals of glass-forming liquids were measured in the condition of Figure 14(b) [7, 44]. The modulated light is absorbed in an opaque film, and thermal waves are generated at interfaces with adjacent gas (0 < x < Lg) and a sample (−Ls < x < 0). During a cycle of modulation, thermal waves propagate μs and μg to minus and plus directions along the x-axis, respectively.

According to their theory on the case of an optically very thin and thermal thick sample, the photoacoustic signal is related to squared complex thermal effusivity, Cpνκν,which is the product of complex specific heat and complex thermal conductivity, as below. Linear-response theory defines dynamic thermal conductivity by the fluctuation of heat flux density [5]:

Cpνκν=Aptν2.E30
tan1Cpνκν"Cpνκν=2ϕptν.E31

Here, Aptν and ϕptν are the frequency-dependent photoacoustic amplitude and the phase lag, respectively. The Cpν and κν are dynamic specific heat and dynamic thermal conductivity, respectively. From Eqs. (30) and (31), the frequency-dependence of the product of a dynamic specific heat and a dynamic thermal conductivity can be determined by the frequency-dependence of observed Aptν and ϕptν. Figure 15 shows the temperature dependence of Apt and −2 ϕpt of propylene glycol at 80 Hz [44]. Upon cooling from a high temperature, the relaxation frequency (the inverse of the relaxation time) decreases and approaches the modulation frequency of an incident light. When the relaxation frequency becomes the same as the modulation frequency, the maximum of −2 ϕpt appears and Apt decreases.

Figure 15.

Temperature dependence of amplitude and phase of photoacoustic signals in propylene glycol [44].

Cole-Cole plot of dynamic Cpκ of propylene glycol is shown in Figure 16 [44]. These values of the Davidson-Cole exponent of photoacoustic method were compared with the values determined by the dielectric and the ultrasonic measurements, as shown in Table 1. It is found that the Davidson-Cole exponent of photoacoustic method in the two liquids is smaller than those of the dielectric and ultrasonic measurements [45, 46]. The dielectric relaxation is related to only the fluctuation of polarization. In contrast, a photoacoustic response is related to thermal fluctuations, which contain all degrees of freedom including polarization and strain. Therefore, the distribution of relaxation time of photoacoustic method can be broader than those of the dielectric and ultrasonic measurements. Therefore, it is reasonable that Davidson-Cole exponents of photoacoustic method can be smaller than those of the dielectric and elastic measurements. This fact is seen in Table 1. There are two and three hydroxyl groups in a molecule for propylene glycol and glycerol, respectively. Since the dominant interaction among molecules is the intermolecular hydrogen bonding, the nearest interaction of molecules of glycerol is more complex than that of propylene glycol, and the distribution of relaxation time of glycerol can be broader than that of propylene glycol. Therefore, the exponent of glycerol is smaller than that of propylene glycol, which is smaller than those of n- and i-propanol [47].

Figure 16.

Cole-Cole plot of dynamic Cpκ of propylene glycol. The solid line shows calculated values by the Davidson-Cole equation at βCD=0.52, where (Cpκ)′ and (Cpκ)” are real and imaginary parts of dynamic Cpκ [44].

MethodResponseGlycerolPropylene glycol
Dielectric [45]Polarization0.58 ± 0.030.66 ± 0.01
Ultrasonic [46]Strain0.42 ± 0.05
Photoacoustic [44]Enthalpy0.35 ± 0.030.52 ± 0.05

Table 1.

Comparison of Davidson-Cole exponents of dielectric, ultrasonic, and photoacoustic measurements in glass-forming glycerol and propylene glycol.

Upon cooling a liquid from a high temperature, several kinds of relaxation processes, an α-relaxation, slow and fast β-processes and the low-energy excitation, boson peaks were observed in the large frequency range between the μHz and THz ranges as shown in Figure 17 [9, 48, 49]. The relaxation time of the main α-structural relaxation process in a fragile liquid diverges at TVF, and it obeys the following Vogel-Tammann-Fulcher law reflecting the cooperative motion in a cage [2].

Figure 17.

Various relaxation processes and boson peaks of propylene glycol. The solid line shows the Vogel-Tammann-Fulcher (VTF) law observed by various measurements.

τ=τ0expBTTVF.E32

Here τ0and B are constants. However, in a strong liquid, such as a silica glass, TVF is far below Tg, and the relaxation time approximately obeys the Arrhenius law. Therefore, no remarkable change in relaxation time is observed in the vicinity of Tg.

Propylene glycol undergoes a liquid-glass transition at about Tg = 168 K. It is one of the typical glass-forming materials with the strong glass-forming tendency [48, 49]. It belongs to an intermediate liquid with a fragility index of m = 48 in the strong-fragile classification [33]. The temperature dependences of α relaxation determined by dielectric and photoacoustic measurements together with the boson peak and fast relaxations determined by the Raman and Brillouin scattering measurements are shown in Figure 7. The boson peak frequency is nearly constant. The fast relaxation determined from the polarized Brillouin scattering spectra is also temperature-independent. In the supercooled liquid, this process is faster than other relaxation processes determined by dielectric and photoacoustic measurements [9, 44]. This process may be attributed to the fast β-process predicted by the mode coupling theory on glass transitions, which is faster than the α-relaxation process [50]. Another relaxation process was determined from the Rayleigh wing in a depolarized Brillouin scattering spectrum, and it was attributed to the α-relaxation at a GHz range [49]. The γ-relaxation was also determined from the width of a longitudinal acoustic (LA) peak, assuming the coupling between the relaxation process and LA mode [51]. Below a GHz range, the relaxation time of the α-relaxation process was observed by dielectric spectroscopy, which detects polar motions [9]. While dynamic heat capacity detects both polar and nonpolar motions. The mean relaxation time of the α-relaxation process obeys the Vogel-Tammann-Fulcher law. For the detailed study on the relaxation processes and vibrational properties in a liquid-glass transition, a complete picture of relaxation map is required in the very large frequency range, in which various kinds of relaxation processes and low-energy excitation are included.

Advertisement

5. Conclusions

In a linear-response theory, the dynamic specific heat is defined by Kubo’s formula on the fluctuation-dissipation theorem using the correlation function of enthalpy fluctuations. It is important that the dynamic specific heat can provide us with a more general interpretation of various relaxation and dynamical processes. In fact, the origin of the enthalpy relaxation is attributed to the total degrees of freedom of atomic motions. While, the dielectric relaxation is related only to polar atomic motions, and ultrasonic spectroscopy is related only to strain. This chapter introduces two experimental methods to measure physical properties related to dynamic specific heat, (1) temperature-modulated differential calorimetry (MDSC) and (2) photoacoustic spectroscopy. Then, the dynamical properties of glass transitions in oxide network glasses and hydrogen-bonded alcohols are reviewed on dynamic glass transition temperature, fragility index, and the distribution of relaxation time.

In contrast, in broadband dielectric spectroscopy, which covers from micro- to terahertz frequency range, the frequency range of observable dynamic heat capacity is limited in the low-frequency range. MDSC detects dynamic specific heat, while specific-heat spectroscopy and photoacoustic methods are based on thermal effusion [18] and detect dynamic thermal effusion, which relates to the product of dynamic specific heat and dynamic thermal conductivity. Therefore, further development of experimental methods is necessary to extend the frequency range of measurements of dynamic specific heat or to detect dynamic specific heat and dynamic thermal conductivity independently.

References

  1. 1. Kauzmann W. The nature of the glassy state and the behavior of liquids at low temperatures. Chemical Reviews. 1948;43:219-256. DOI: 10.1021/cr60135a002
  2. 2. Fulcher GS. Analysis of recent measurements of the viscosity of glasses. Journal of the American Ceramic Society. 1925;8:339-355. DOI: 10.1111/j.1151-2916.1925.tb16731.x
  3. 3. Lunkenheimer P, Loidl A. Dielectric spectroscopy of glass-forming materials: α-relaxation and excess wing. Chemical Physics. 2002;284:205-219. DOI: 10.1016/S0301-0104(02)00549-9
  4. 4. Kojima S. Anniversary of Brillouin scattering: Impact on materials science. Materials. 2022;15:3518. DOI: 10.3390/ma15103518
  5. 5. Kubo R. Statistical-mechanical theory of irreversible processes. I. General theory and simple applications to magnetic and conduction problems. Journal of Physical Society of Japan. 1957;12(6):570-586. DOI: 10.1143/JPSJ.12.570
  6. 6. Birge NO, Nagel SR. Specific heat spectroscopy of the glass transition. Physical Review Letters. 1985;54(25):2674-2677. DOI: 10.1103/physrevlett.54.2674
  7. 7. Kojima S, Kashiwada T, Nakahira S. Photoacoustic measurements of thermal dispersion in glass-transitions of glycerol and propylene glycol. High Temperature-High Pressures. 1991;23:701-704
  8. 8. Debye P. Zur Theorie der anomalen Dispersion im Gebiete der langwelligen elektrischen Strahlung. Verh Deut Phys Gesell. 1913;15:777-793
  9. 9. Park IS, Saruta K, Kojima S. Dielectric relaxation and calorimetric measurements of glass transition in the glass-forming dihydroxy alcohols. Journal of Physical Society of Japan. 1998;67(12):4131-4138. DOI: 10.1143/JPSJ.67.4131
  10. 10. Davidson DW, Cole RH. Dielectric relaxation in glycerol, propylene glycol, and n-propanol. Journal of Chemical Physics. 1951;19(12):1484-1490. DOI: 10.1063/1.1748105
  11. 11. Cole KS, Cole RH. Dispersion and absorption in dielectrics, I. Alternating current characteristics. Journal of Chemical Physics. 1941;9(4):341-351. DOI: 10.1063/1.1750906
  12. 12. Havriliak S, Negami S. A complex plane analysis of α-dispersions in some polymer systems. Journal of Polymer Science Part. 1996;14:99-117
  13. 13. Kohlrausch R. Theorie des elektrischen rückstandes in der Leidener flasche. Annalen der Physik. 1854;167:161-320. DOI: 10.1002/andp.18541670203
  14. 14. Willams G, Watts DC. Non-symmetrical dielectric relaxation behavior arising from a simple empirical decay function. Transaction of the Faraday Society. 1970;66:80-85. DOI: 10.1039/TF9706600080
  15. 15. Alvarez F, Alegria A, Colmenero J. Relationship between the time-domain Kohlrausch-Williams-Watts and frequency-domain Havriliak-Negami relaxation functions. Physical Review B. 1991;44(14):7306-7312. DOI: 10.1103/PhysRevB.44.7306
  16. 16. Birge NO. Specific-heat spectroscopy of glycerol and propylene glycol near the glass transition. Physical Review B. 1986;34(3):1631-1642. DOI: 10.1103/PhysRevB.34.1631
  17. 17. Matsuda Y, Fukawa Y, Ike Y, Kodama M, Kojima S. Dynamic specific heat, glass transition and non-Debye nature of thermal relaxation in Lithium borate glasses. Journal of the Physical Society of Japan. 2008;77(8):084602. DOI: 10.1143/JPSJ.77.084602
  18. 18. Christensen T, Olsen NB, Dyre JC. Conventional methods fail to measure cp(ω) of glass-forming liquids. Physical Review E. 2007;75(4):041502. DOI: 10.1103/PhysRevE.75.041502
  19. 19. Matsuda Y, Matsui C, Ike Y, Kodama M, Kojima S. Non-Debye nature in thermal relaxation and thermal properties of Lithium borate glasses studied by modulated DSC. Journal of Thermal Analysis and Calorimetry. 2006;85(3):725-730. DOI: 10.1007/s10973-006-7652-9
  20. 20. Reading M, Elliott D, Hill VL. A new approach to the calorimetric investigation of physical and chemical transitions. Journal of Thermal Analysis. 1993;40:949-955. DOI: 10.1007/BF02546854
  21. 21. Reading M, Luget A, Wilson R. Modulated differential scanning calorimetry. Thermochimica Acta. 1994;238:295-307. DOI: 10.1016/S0040-6031(94)85215-4
  22. 22. Park I, Kojima S. Glass transition and structural relaxation of intermediate liquids by MDSC and dielectric spectroscopy. Thermochimica Acta. 2000;352-353:147-152. DOI: 10.1016/S0040-6031(99)00459-1
  23. 23. Carpentier L, Bustin O, Decamps K. Temperature-modulated differential scanning calorimetry as a specific heat spectroscopy. Journal of Physics D. 2002;35(4):402-408. DOI: 10.1088/0022-3727/35/4/317
  24. 24. Schawe JEK. Principles for the interpretation of modulated temperature DSC measurements. Part 1. Glass transition. Thermochimica Acta. 1995;261(1):183-194. DOI: 10.1016/0040-6031(95)02315-S
  25. 25. Schawe JEK, Höhne GWH. The analysis of temperature modulated DSC measurements by means of the linear response theory. Thermochimica Acta. 1996;287(2):213-223. DOI: 10.1016/0040-6031(96)88984-2
  26. 26. Ike Y, Seshimo Y, Kojima S. Complex heat capacity of non-Debye process in glassy glucose and fructose. Fluid Phase Equilibria. 2007;256(1–2):123-126. DOI: 10.1016/j.fluid.2007.04.017
  27. 27. Weyer S, Hensel A, Schick C. Phase angle correction for TMDSC in the glass-transition region. Thermochimica Acta. 1997;304(305):267-275. DOI: 10.1016/S0040-6031(97)00180-9
  28. 28. Fukawa Y, Matsuda Y, Kawashima M, Kojima S. Determination of complex-specific heat and fragility of sodium borate glasses by temperature-modulated DSC. Journal of Thermal Analysis and Calorimetry. 2010;99:39-44. DOI: 10.1007/s10973-009-0522-5
  29. 29. Angell CA, MacFarlane DR, Oguni M. The Kauzmann paradox, metastable liquids, and ideal glasses: A summary. Annals of the New York Academy of Sciences. 1986;484(1):241-247. DOI: 10.1111/j.1749-6632.1986.tb49574.x
  30. 30. Vashist P, Patial BS, Bhardwaj S, Tripathi SK, Thakur N. On the non-isothermal crystallization kinetics, glass forming ability and thermal stability of Bi additive Se–Te–Ge alloys. Journal of Thermal Analysis and Calorimetry. 2023;148(15):7717-7726. DOI: 10.1007/s10973-023-12271-5
  31. 31. Matsuda Y, Fukawa Y, Kawashima M, Mamiya S, Kojima S. Dynamic glass transition and fragility of lithium borate binary glasses. Solid State Ionics. 2008;179(40):2424-2427. DOI: 10.1016/j.ssi.2008.09.011
  32. 32. Hashimoto E, Seshimo Y, Sasanuma K, Aoki Y, Kanazawa H, Ike Y, et al. Dynamical properties of ethylene glycol and glucose studied by temperature-modulated differential scanning calorimetry and Brillouin scattering. Journal of Thermal Analysis Calorimetry. 2010;99(1):45-50. DOI: 10.1007/s10973-009-0543-0
  33. 33. Koda S, Shibata T, Park IS, Kojima S. Relaxation dynamics and fragility during liquid-glass transitions of poly(propylene glycol)s. Current Applied Physics. 2015;15(7):805-810. DOI: 10.1016/j.cap.2015.04.041
  34. 34. Chryssikos GD, Kamitsos EI, Yiannopoulos YD. Towards a structural interpretation of fragility and decoupling trends in borate systems. Journal of Non-Crystalline Solids. 1996;196:244-248. DOI: 10.1016/0022-3093(95)00594-3
  35. 35. Bell AG. On the production and reproduction of sound by light. American Journal of Science. 1880;s3-20(118):305-324. DOI: 10.2475/ajs.s3-20.118.305
  36. 36. Tyndall J. Action of an intermittent beam of radiation heat upon gaseous matter. Proceedings of Royal Society of London. 1881;31:307-317. DOI: 10.1098/rspl.1880.0037
  37. 37. Rontgen WC. On tones produced by the intermittent irradiation of a gas. Philosophical Magazine Series 5. 1881;11(68):308-311. DOI: 10.1080/14786448108627021
  38. 38. Bell AG, LXVIII. Upon the production of sound by radiant energy. Philosophical Magazine Series 5. 1881;11(71):510-528. DOI: 10.1080/14786448108627053
  39. 39. Luscher E, Coufal HJ, Korpiun P, Tilgner R, editors. Photoacoustic Effect. Braunschweig: Friedr. Vieweg & Sohn; 1984
  40. 40. Kojima S. Photoacoutic imaging for the melting of gallium. Japanese Journal of Applied Physics. 1984;24(12R):1571-1572. DOI: 10.1143/JJAP.24.1726
  41. 41. Ghaliev R. Photoacoustic Imaging: Principles. IntechOpen, London, UK: Advances and Applications; 2020
  42. 42. Rosencwaig A, Gersho A. Theory of the photoacoustic effect with solids. Journal of Applied Physics. 1976;47:64-69. DOI: 10.1063/1.322296
  43. 43. Kojima S. Study of phase transitions by frequency dependent photoacoustic measurements. In: Leroy O, Breazeal MA, editors. Physical Acoustics. New York: Plenum Press; 1991. pp. 399-404
  44. 44. Kojima S, Koketsu T, Takahashi E, Kanayasu M. Complex thermal effusivity of glassy polyhydric alcohols. Fluid Phase Equilibria. 1993;88:209-218. DOI: 10.1016/0378-3812(93)87112-E
  45. 45. Angle CA, Smith DL. Test of the entropy basis of the Vogel-Tammann-Fulcher equation. Dielectric relaxation of polyalcohols near Tg. Journal of Physical Chemistry. 1982;86(19):3845-3852. DOI: 10.1021/j100216a028
  46. 46. Jeong YH, Nagel SR, Bhattachya S. Ultrasonic investigation of the glass transition in glycerol. Physical Review A. 1986;34(1):602-608. DOI: 10.1103/PhysRevA.34.602
  47. 47. Park IS, Mikami M, Kojima S, Takagi Y. Dynamics of liquid-glass transition in propanol. AIP Conference Proceedings. 1999;469:525-526. DOI: 10.1063/1.58534
  48. 48. Kojima S, Novikov V. Correlation of temperature dependence of Quasielastic light scattering intensity and alpha-relaxation time. Physical Review B. 1996;54(1):222-227. DOI: 10.1103/PhysRevB.54.222
  49. 49. Kojima S, Sato H, Yoshihara A. Light scattering of supercooled propylene glycol. Journal of Physics: Condensed Matter. 1997;9(46):10079-10086. DOI: 10.1088/0953-8984/9/46/005
  50. 50. Gotze W, Sjogren L. Relaxation processes in supercooled liquids. Reports on Progress in Physics. 1992;55(3):241-376. DOI: 10.1088/0034-4885/55/3/001
  51. 51. Takanashi K, Yoshihara A, Kojima S. Brillouin scattering study on dynamical properties of a viscoelastic liquid propylene glycol using a tandem Fabry-Perot interferometer. Japanese Journal of Applied Physics. 1994;33(5S):2867-2873. DOI: 10.1143/JJAP.33.2867

Written By

Seiji Kojima

Submitted: 14 June 2023 Reviewed: 21 June 2023 Published: 24 September 2023