Brillouin scattering was performed on an elastomeric polyurea, an important technological polymer. Being widely used for impact modification, of particular interest is its response to extreme pressure conditions. We applied pressures up to 13.5 GPa using a diamond anvil cell and measured the longitudinal and transverse sound velocities via Brillouin light scattering. From these data, the equation of state, the elastic moduli, and Poisson's ratio were obtained. By comparison with previous dilatometry measurements up to 1 GPa, we show how viscoelastic effects can be accounted for in order to obtain an accurate equation of state. Because of the extreme strain-rate hardening of vitrifying polyurea, the property changes associated with its solidification are more subtle in the high frequency Brillouin data than observed in conventional mechanical testing and dilatometry.

The past decade has seen a dramatic increase in the use of pressure as an experimental variable, particularly in the study of amorphous liquids and polymers.1,2 The interpretation of high pressure data usually requires knowledge of the density, ρ, for measurements performed as a function of temperature, T, and pressure, P, that is, the equation of state (EoS), ρ(T,P), is required. A prominent recent example of this is thermodynamic scaling of dynamic quantities and transport properties, in which measurements at various T and P are expressed as a function of Tργs, in which γs is a material constant related to the intermolecular repulsive potential.3–5 Polyurea is a technologically important polymer, which has seen increasing use as a coating for infrastructure protection and military armor.6,7 Because of its application as an impact coating, much work has been done characterizing the properties of polyurea under extreme conditions, such as high strain rates and high pressures.6–19 It is therefore of interest to determine its density and mechanical properties for large compressive forces.

Measuring the EoS of amorphous systems is substantially more difficult than for crystalline materials. The density of the latter can be obtained from x-ray and neutron diffraction, but these methods are not applicable to materials lacking long range order. In high pressure work, Bridgman measured the EoS for many amorphous materials using dilatometry, in which a piston-cylinder apparatus or a sample within a confining fluid was employed.20 Zoller et al. reported the P-V-T results for many polymers by the latter method, but most of the data were limited to 0.2 GPa.21 

Diamond anvil cell (DAC) methods together with an expanding range of measurement techniques are now capable of determining the properties of materials up to several hundred GPa.22 Standard dilatometry techniques are challenging in a DAC, but Brillouin light scattering can be used to obtain density changes from changes in the acoustic sound velocities.23–27 An acknowledged problem with this technique, however, is that Brillouin scattering probes materials at very high (GHz) frequencies, and polymers, which are viscoelastic, exhibit sound velocities (and mechanical properties) that are invariably strongly dependent on frequency.28,29 Neglect of this effect leads to an underestimation of density changes in the typical isothermal compression experiment; nevertheless, no corrections for viscoelasticity have been applied in previous high pressure Brillouin scattering studies of polymers.

In a recent study of polyurea, we measured the glass transition temperature, Tg, up to 6 GPa using a DAC technique based on the disappearance of pressure gradients in a glassy sample upon heating through Tg.30 At 22 °C, the polymer vitrifies at 1.1 GPa. The EoS for the polyurea up to 1 GPa was also obtained using dilatometry.30 In this work, we report the density and moduli of the same polyurea at pressures up to 13.5 GPa from Brillouin measurements using a DAC. While the theory of Brillouin scattering is described in detail elsewhere,31 we briefly discuss pertinent aspects herein. In any material, thermally generated acoustic phonons, having a range of momenta, are continuously created and destroyed. Incident laser light can absorb or create such a phonon, resulting in a shift in frequency ±υ and momentum of the scattered photon. By collecting light along a well-defined scattering angle, conservation of momentum restricts the phonons probed to a specific momentum q. Longitudinal acoustic (LA) phonons at this q have a peak frequency υL and phase velocity vL determined by the elastic properties of the medium. Transverse acoustic (TA) phonons with frequency υT and velocity vT can also propagate if the viscosity is high enough to transiently support a shear wave. Scattering from these phonons gives rise to symmetric peaks in the resulting frequency spectrum. Conservation of energy and momentum yields an equation specific to equal-angle forward scattering32 

υi=2viλ0sin(θ2)
(1)

in which λ0 is the wave length and θ the scattering angle. This equation is used to calculate sound velocities vi from peak positions υi.

We show herein that the effect of viscoelasticity on the bulk modulus is significant, causing the Brillouin values to be substantially larger than obtained from low frequency dilatometry experiments. We provide a correction for this effect to obtain accurate densities in agreement with the lower pressure dilatometry data.30 As expected, the longitudinal modulus increases strongly with pressure, with a change in pressure-sensitivity upon vitrification.

Polyurea was prepared by a reaction of Versalink P1000 polyamine (Air Products) with Isonate 143 L isocyanate (Dow Chemical) in a 4:1 mass ratio. Films 100–200 μm thick were prepared by pressing the reactants between Teflon sheets, followed by curing for 8 h at 75 °C. The sample was then allowed to equilibrate under ambient conditions, with a consequent small water uptake (<1% by weight). Small discs were cut from the molded sheet and loaded into a DAC using a stainless steel gasket. One or two ruby chips were also positioned in the sample cell (dimensions approximately 300 μm in diameter and 60 μm thick) for in-situ pressure determination.33 Ruby fluorescence spectra were collected using a Princeton Instruments Acton SP2300 spectrometer with an 1800BLZ grating, giving a pressure uncertainty of ±0.1 GPa. The experiments were repeated thrice, each with a separate sample loading.

Brillouin spectra were collected using a single-mode, Ar-ion laser (λ0=514.5nm) as an excitation source. The laser power was kept below 100 mW; nevertheless, some minor photodegradation of the polyurea was observed. Light from the DAC was collected in a symmetric, equal angle forward scattering geometry (external scattering angle = 61.61°). The scattered light was focused onto a 100 μm pinhole of a 3 + 3 tandem Fabry-Perot interferometer. A schematic of the experimental setup is shown in Fig. 1. A typical spectrum was collected for 20–60 min at each pressure, with longer collection times at higher pressures due to the weaker signal. Measurements herein were at room temperature (22 °C), at pressures from 0.1 MPa to 13.5 GPa. Typically, pressures were systematically increased although a cyclic compression/decompression measurement gave no indication of hysteresis.

FIG. 1.

Experimental setup for Brillouin light scattering in a DAC.

FIG. 1.

Experimental setup for Brillouin light scattering in a DAC.

Close modal

Typical Brillouin spectra are shown in Fig. 2, in which prominent LA mode peaks are evident; these shift to higher frequencies with pressure. Also present are the much weaker TA mode peaks. Being too low in frequency to be observed in the ambient pressure spectrum, they emerge from the central line at ∼0.4 GPa. This appearance of the TA modes at high pressure is discussed in Ref. 23. The shift of the TA modes with pressure is much weaker than for the LA modes. The amplitudes of both peaks decrease at higher pressures. When the polyurea is deep in the glassy state (>5 GPa), the TA modes became too weak to resolve; however, the LA modes can be observed up to the highest pressures herein (>13 GPa).

FIG. 2.

Selected Brillouin spectra from polyurea showing the pressure evolution of longitudinal (LA) and transverse (TA) acoustic mode peaks. Solid lines are fits to Eq. (2).

FIG. 2.

Selected Brillouin spectra from polyurea showing the pressure evolution of longitudinal (LA) and transverse (TA) acoustic mode peaks. Solid lines are fits to Eq. (2).

Close modal

To extract the peak frequencies υL and υT, spectra were fit with each mode modeled by a damped harmonic oscillator

I(υ)=I0(2Γυ)2+(υ2υi2)2,
(2)

where I0 is the intensity and Γ the half width. Half widths decreased with pressure up to vitrification due to the increasing phonon lifetimes on compression. In the glassy state, there is an apparent broadening of the Brillouin doublets due to non-hydrostatic conditions (non-uniform local pressures). From the peak shift frequencies, the LA (vL) and TA (vT) sound velocities were calculated via Eq. (1) (Fig. 3). While both vL and vT increase monotonically with pressure, the effect is markedly stronger at low pressures (below 1 GPa). The pressure at which the response becomes glassy, Pg = 1.1 GPa at 22 °C for low frequencies,30 is indicated in the figure.

FIG. 3.

Longitudinal and shear sound velocities for polyurea calculated from Eq. (1). Dashed line is the fit of Eq. (3). Vertical line denotes the glass transition pressure, Pg = 1.1 GPa at room temperature.30 The error bars are no larger than the symbol size.

FIG. 3.

Longitudinal and shear sound velocities for polyurea calculated from Eq. (1). Dashed line is the fit of Eq. (3). Vertical line denotes the glass transition pressure, Pg = 1.1 GPa at room temperature.30 The error bars are no larger than the symbol size.

Close modal

Even though the TA modes are not visible in all spectra, the values of vT for all pressures are required to calculate the density ρ and elastic parameters. Therefore, the available vT measurements can be fit with a phenomenological model

vT(P)=vT0+CTln(1+PbT),
(3)

where bT and CT are constants and vT0(T) is the sound velocity at zero pressure. To verify Eq. (3), we applied it to the data in Ref. 23 and found it to describe accurately the isothermal pressure dependence of vL and vT for three polymers over similar pressure ranges. The fit to vT for polyurea shown in Fig. 3 yields vT0 = 0.866 ± 0.15 km/s, CT = 0.855 ± 0.076 km/s, and bT = 0.7 ± 0.2 GPa. We then extrapolated to estimate vT in the low (<0.4 GPa) and high (>5 GPa) pressure regimes.

In order to calculate ρ at high pressures using Brillouin sound velocity data, most studies use the equation

ρρ0=P0PγvB2dP,
(4)

where ρ0 is the ambient density, γ the adiabatic index, and vB the bulk sound velocity introduced for mathematical convenience23,27,34

vB2=vL243vT2.
(5)

However, Eq. (4) neglects the fact that Brillouin scattering measures sound velocities at very high (GHz) frequencies, which due to viscoelastic effects yields values higher than would be measured using a low frequency probe.35 In polymers, the principal viscoelastic time constant is the local segmental relaxation time τ, defining the onset of glassy behavior.1,2,29 When the time scale of a dynamic measurement is much longer than τ, the material response is liquid-like (“relaxed”), whereas solid-like behavior is observed for measurements occurring faster than the material response. In the case of the polyurea at room temperature, previous dielectric measurements reported τ ∼ 1 μs at 0.1 MPa.36,37 Acoustic phonons measured by Brillouin scattering are in the GHz frequency range and thus probe the polymer on times scales of a few ns, i.e., much shorter than τ at ambient conditions. Since compression further increases τ, our Brillouin measurements are all in the unrelaxed, solid-like regime. In a study of liquid methanol for which τ∼1 ns at ambient pressure, correction for viscoelastic effects was shown to change the response from that of a relaxed to an unrelaxed state with increasing pressure approaching vitrification.38 Equation (4) is derived based upon the assumption of a relaxed response and cannot yield accurate densities if applied directly to our sound velocities herein. While Ref. 23 alluded to this problem, no attempt was made to estimate the error.

We derive a correction to Eq. (4) for the viscoelastic contribution, relying on the density measurements from dilatometry on this same polyurea to ∼1 GPa.30 These data were described by the Tait EoS21 

ρ(P)=ρ0(T)[1Cln(1+PB(T))],
(6)

where ρ0(T) was modeled with a 2nd degree polynomial, B(T)=b0exp(b1T), and C is a constant. From this, we calculated the pressure dependence of the isothermal bulk modulus

KT=ρdPdρ|T,
(7)

and using the Maxwell relation, the low frequency adiabatic bulk modulus KS0

1KS0=1KTα2TρcP,
(8)

where α is the thermal expansivity and cP the isobaric specific heat capacity. We calculated α from the EoS and measured cP = 1.97 J/g · K using differential scanning calorimetry (cP was approximated as constant).

In contrast to the relaxed modulus KS0 obtained from dilatometry, Brillouin scattering determines (at least when τ is large) the high frequency adiabatic bulk modulus, KS=ρvB2, which is the modulus in the unrelaxed limit. KS can be modeled by

KS=KS0+ΔKS,
(9)

where ΔKS is the relaxing part of the modulus describing the increase at very high frequencies.28 [The assumption of Eq. (9) that stresses are additive is open to question.39] Fig. 4 shows the values of KS and KS0 from ambient pressure up to 1.1 GPa. It is apparent that KS is significantly larger than KS0; however, the curves are essentially parallel over the entire pressure range, indicating that the pressure-dependence is consistent for the two methods. Thus, the relaxing part is constant to within the uncertainty, ΔKS=1.56±0.23GPa. This value is used in subsequent calculations for P < Pg.

FIG. 4.

Measurements of KS∞ from Brillouin data and KS0 calculated from the dilatometry measurements of Ref. 30 using Eq. (8). The two data points at P = 0.6 GPa are measurements from two separate DAC loadings, and thus a measure of the reproducibility. Inset shows the longitudinal modulus (open symbols; left ordinate) and shear modulus (solid symbols; right ordinate) measured at 0.1 MPa using dynamic mechanical,19 ultrasonic,40 and Brillouin scattering. Dashed lines represent the KWW function with βKWW = 0.3.

FIG. 4.

Measurements of KS∞ from Brillouin data and KS0 calculated from the dilatometry measurements of Ref. 30 using Eq. (8). The two data points at P = 0.6 GPa are measurements from two separate DAC loadings, and thus a measure of the reproducibility. Inset shows the longitudinal modulus (open symbols; left ordinate) and shear modulus (solid symbols; right ordinate) measured at 0.1 MPa using dynamic mechanical,19 ultrasonic,40 and Brillouin scattering. Dashed lines represent the KWW function with βKWW = 0.3.

Close modal

The viscoelastic effect is also evident by comparing the longitudinal modulus M and shear modulus G measured at different frequencies. In the inset of Fig. 4 are M and G at ambient conditions from ultrasonic spectroscopy at 1 MHz,40 dynamic mechanical spectroscopy at 1 Hz,19 and the current Brillouin data at several GHz (calculated as M=ρvL2 and G=ρvT2). The moduli increase significantly with increasing measurement frequency. The dashed lines show data for a Kohlrausch-Williams-Watts (KWW) relaxation function29 with a representative value of the stretch exponent, βKWW= 0.3.

The adiabatic index can be obtained from the low pressure dilatometry data γ=KS0/KT. While γ is typically ∼1.2 at ambient pressure, it decreases towards unity at high pressure.23 Previous Brillouin studies often assumed γ=1,23,41 but from our values of KT and KS0, we calculate an ambient value γ0=1.12, decreasing with pressure to asymptotically approach γP=1.08. For pressures beyond the range of the EoS, we held γ constant at γP. The modeling of this parameter gives a more accurate calculation of density via Eq. (4) than the usual approximation γ=1.

In order to incorporate ΔKS in the calculation of ρ, we express KT in terms of ΔKS

KT=KS0γ=1γ(KSΔKS).
(10)

Inserting into Eq. (7) along with KS=ρvB2 yields

dρ=ργρvB2ΔKSdP,
(11)

which upon integration gives

ρρ0=P0PγvB2(1ΔKSρvB2)1dP.
(12)

This is Eq. (4) corrected for viscoelasticity. Since the integrand contains ρ, formally the integration is not allowed. However, we proceed by first obtaining an approximate ρ(P) from Eq. (4) and then inserting those values into the integrand of Eq. (12). By iteration, we arrive at the most accurate values of ρ(P).

The value determined for ΔKS in Eq. (12) is strictly applicable only to the polyurea above Tg. In the glassy state, τ is so large that the density increases very slowly as the polymer equilibrates, a process known as physical ageing.42 This means that our experimental time scale (i.e., time for a pressure increment and spectral measurement, tens of minutes) is much shorter than τ for the glassy polyurea; thus, a pressure change induces a density response according to K rather than K0. Accordingly, we set ΔKS=0 for pressures higher than Pg=1.1GPa.

The integration of Eq. (12) yields ρ(P) up to 13.5 GPa, as displayed in Fig. 5. Note that the values obtained with Eq. (12) are much more consistent with the low pressure dilatometry data than using Eq. (4), which does not correct for viscoelasticity. In Ref. 23, Brillouin measurements were used to calculate ρ(P) using Eq. (4) for three other polymers up to about 14 GPa. When compared with lower pressure dilatometry data, these ρ(P) values were found to be significantly too low. However, we find that these differences vanish when the data are analyzed via Eq. (12) with estimates of ΔKS. The parameter ΔKS is applicable to any material exhibiting a viscoelastic response although of course its value would be material dependent. The aforementioned finding that ΔKS is constant is important because it suggests that it is sufficient to have a single determination of KS0 at ambient pressure in order to apply Eq. (12). This would be useful for a Brillouin study of a material for which dilatometric characterization is lacking.

FIG. 5.

Pressure dependence of ρ for polyurea calculated with viscoelastic corrections [Eq. (12)] and without corrections [Eq. (4)]. Dilatometry values are from Ref. 30; solid and dashed lines are fits of Eq. (6) to data below and above Pg, respectively. Inset shows Poisson's ratio determined from: Brillouin (circles), dynamic mechanical at 1 Hz19 (star), and ultrasonic at 1 MHz40 (diamond).

FIG. 5.

Pressure dependence of ρ for polyurea calculated with viscoelastic corrections [Eq. (12)] and without corrections [Eq. (4)]. Dilatometry values are from Ref. 30; solid and dashed lines are fits of Eq. (6) to data below and above Pg, respectively. Inset shows Poisson's ratio determined from: Brillouin (circles), dynamic mechanical at 1 Hz19 (star), and ultrasonic at 1 MHz40 (diamond).

Close modal

Many different EoS models have been used to describe the pressure dependence of ρ; for example, Ref. 23 compared five EoS for their ability to fit data. Herein, we report results for the Tait EoS, Eq. (6), used previously for dilatometry data.30 For an isothermal pressurization at T =22 °C, only the parameters C and B were adjusted in fitting the data, with the ambient density fixed at ρ=1.098 g/ml. We fit separately the regimes above and below Pg, using for the latter the parameters from Ref. 30; results are given in Table I. As seen in Fig. 5, the Tait equation accurately describes the densities for both rubbery and glassy polyurea.

TABLE I.

Parameters from fits of Eq. (6).

CB (GPa)
P < 1.1 GPa 0.0889±0.001 0.239±0.007 
P > 1.1 GPa 0.1058±0.001 0.358±0.007 
CB (GPa)
P < 1.1 GPa 0.0889±0.001 0.239±0.007 
P > 1.1 GPa 0.1058±0.001 0.358±0.007 

Using the EoS for the entire pressure range, we calculated the pressure dependence of the high frequency longitudinal and shear moduli via the relations M=ρvL2 and G=ρvT2; these are shown in Fig. 6. There is a marked decrease in the sensitivity to pressure above Pg, the weaker dependence due to the reduced compressibility of the glass.

FIG. 6.

High frequency longitudinal M and shear G moduli for polyurea. Vertical line denotes Pg = 1.1 GPa. Inset shows the pressure derivative of M.

FIG. 6.

High frequency longitudinal M and shear G moduli for polyurea. Vertical line denotes Pg = 1.1 GPa. Inset shows the pressure derivative of M.

Close modal

From the ratio of transverse to longitudinal strain, we calculate Poisson's ratio

ν̃=vL22vT22vL22vT2,
(13)

which is plotted versus pressure in the inset of Fig. 5. From a value of 0.41 at ambient, ν̃ decreases to 0.37 over the first few GPa, remaining constant for higher pressures. The neglect of viscoelasticity also reduces the magnitude of the apparent ν̃. From ultrasonic measurements at 1 MHz,40 ν̃ = 0.44.

Using Brillouin scattering, the longitudinal and shear sound velocities, density, elastic moduli, and Poisson's ratio were determined for polyurea at pressures to 13.5 GPa. By utilizing lower pressure dilatometry and lower frequency dynamic measurements, we corrected the high frequency Brillouin data for viscoelastic effects. Accordingly, the densities reported herein are more accurate than previous studies on other polymers; moreover, we show that the viscoelastic correction is generally applicable to Brillouin data. The glass transition at ambient temperature occurs at a relatively low Pg ∼ 1.1 GPa. Because of the extreme strain-rate hardening of the polyurea when vitrification is imminent, the changes in properties upon solidification are more subtle in the high frequency Brillouin results than observed in conventional mechanical testing and dilatometry. The mechanical properties reported herein are of value in modeling polyurea under the high pressure, high loading rate conditions found in many applications of the material.

The work at NRL was supported by the Office of Naval Research, in part by Code 332 (R.G.S. Barsoum). The Brillouin scattering measurements were performed at the Carnegie Institution of Washington, using facilities supported by the U.S. Department of Energy/National Security Administration (No. DE-NA-0002006, CDAC). T.C.R. acknowledges an American Society for Engineering Education postdoctoral fellowship. We thank D. Fragiadakis for a careful reading of the manuscript and insightful suggestions.

1.
C. M.
Roland
,
S.
Hensel-Bielowka
,
M.
Paluch
, and
R.
Casalini
,
Rep. Prog. Phys.
68
,
1405
1478
(
2005
).
2.
C. M.
Roland
,
Macromolecules
43
,
7875
(
2010
).
3.
C. M.
Roland
,
S.
Bair
, and
R.
Casalini
,
J. Chem. Phys.
125
,
124508
(
2006
).
4.
D.
Coslovich
and
C. M.
Roland
,
J. Phys. Chem. B
112
,
1329
(
2008
).
5.
C. M.
Roland
,
J. L.
Feldman
, and
R.
Casalini
,
J. Non-Cryst. Sol.
352
,
4895
(
2006
).
6.
R. G. S.
Barsoum
,
Elastomeric Polymers with High Rate Sensitivity: Applications in Blast, Shockwave, and Penetration Mechanics
(
William Andrew
,
2015
).
7.
N.
Iqbal
,
M.
Tripathi
,
S.
Parthasarathy
,
D.
Kumar
, and
P.
Roy
,
RSC Adv.
6
,
109706
(
2016
).
8.
A. J.
Hsieh
,
T. L.
Chantawansri
,
W.
Hu
,
K. E.
Strawhecker
,
D. T.
Casem
,
J. K.
Eliason
,
K. A.
Nelson
, and
E. M.
Parsons
,
Polymer
55
,
1883
1892
(
2014
).
9.
T.
Jiao
,
R. J.
Clifton
, and
S. E.
Grunschel
,
AIP Conf. Proc.
845
,
809
812
(
2006
).
10.
M. R.
Amini
,
J.
Isaacs
, and
S.
Nemat-Nasser
,
Int. J. Impact Eng.
37
,
82
89
(
2010
).
11.
M. R.
Amini
and
S.
Nemat-Nasser
,
Int. J. Fract.
162
,
205
217
(
2010
).
12.
L.
Xue
,
W.
Mock
, and
T.
Belytschko
,
Mech. Mater.
42
,
981
1003
(
2010
).
13.
V.
Chakkarapani
,
K.
Ravi-Chandar
, and
K. M.
Liechti
,
J. Eng. Mater. Technol.
128
,
489
494
(
2006
).
14.
S. A.
Tekalur
,
A.
Shukla
, and
K.
Shivakumar
,
Compos. Struct.
84
,
271
281
(
2008
).
15.
J.
Yi
,
M. C.
Boyce
,
G. F.
Lee
, and
E.
Balizer
,
Polymer
47
,
319
329
(
2006
).
16.
C. M.
Roland
,
J.
Twigg
,
Y.
Vu
, and
P. H.
Mott
,
Polymer
48
,
574
578
(
2007
).
17.
J. A.
Pathak
,
J. N.
Twigg
,
K. E.
Nugent
,
D. L.
Ho
,
E. K.
Lin
,
P. H.
Mott
,
C. G.
Robertson
,
M. K.
Vukmir
,
T. H.
Epps
, and
C. M.
Roland
,
Macromolecules
41
,
7543
7548
(
2008
).
18.
T.
Choi
,
D.
Fragiadakis
,
C. M.
Roland
, and
J.
Runt
,
Macromolecules
45
,
3581
3589
(
2012
).
19.
P. H.
Mott
,
C. B.
Giller
,
D.
Fragiadakis
,
D.
Rosenberg
, and
C. M.
Roland
,
Polymer
105
,
227
233
(
2016
).
20.
P. W.
Bridgman
,
Rev. Mod. Phys.
18
,
1
(
1946
).
21.
D.
Walsh
and
P.
Zoller
,
Standard Pressure Volume Temperature Data for Polymers
(
CRC Press
,
1995
).
22.
R. J.
Hemley
,
High Pressure Res.
30
,
581
619
(
2010
).
23.
L. L.
Stevens
,
E. B.
Orler
,
D. M.
Dattelbaum
,
M.
Ahart
, and
R. J.
Hemley
,
J. Chem. Phys.
127
,
104906
(
2007
).
24.
L. L.
Stevens
,
D. M.
Dattelbaum
,
M.
Ahart
, and
R. J.
Hemley
,
J. Appl. Phys.
112
,
023523
(
2012
).
25.
A. S.
Benjamin
,
M.
Ahart
,
S. A.
Gramsch
,
L. L.
Stevens
,
E. B.
Orler
,
D. M.
Dattelbaum
, and
R. J.
Hemley
,
J. Chem. Phys.
137
,
014514
(
2012
).
26.
M.-S.
Jeong
,
J.-H.
Ko
,
Y. H.
Ko
, and
K. J.
Kim
,
Curr. Appl. Phys.
15
,
943
946
(
2015
).
27.
Y.-H.
Ko
,
J.
Min
, and
J.
Song
,
Curr. Appl. Phys.
16
,
311
317
(
2016
).
28.
R.
Meister
,
C. J.
Marhoeffer
,
R.
Sciamanda
,
L.
Cotter
, and
T.
Litovitz
,
J. Appl. Phys.
31
,
854
870
(
1960
).
29.
C. M.
Roland
,
Viscoelastic Behavior of Rubbery Materials
(
Oxford
,
2011
).
30.
T. C.
Ransom
,
M.
Ahart
,
R. J.
Hemley
, and
C. M.
Roland
,
Macromolecules
50
,
8274
8278
(
2017
).
31.
I. L.
Fabelinskii
,
Molecular Scattering of Light
(
Springer Science & Business Media
,
2012
).
32.
A.
Polian
,
J. Raman Spectrosc.
34
,
633
637
(
2003
).
33.
G. J.
Piermarini
,
S.
Block
,
J. D.
Barnett
, and
R. A.
Forman
,
J. Appl. Phys.
46
,
2774
2780
(
1975
).
34.
B. W.
Lee
,
M.-S.
Jeong
,
J. S.
Choi
,
J.
Park
,
Y. H.
Ko
,
K. J.
Kim
, and
J.-H.
Ko
,
Curr. Appl. Phys.
17
,
1396
1400
(
2017
).
35.
L.
Comez
,
C.
Masciovecchio
,
G.
Monaco
, and
D.
Fioretto
,
Solid State Phys.
63
,
1
77
(
2012
).
36.
R. B.
Bogoslovov
,
C. M.
Roland
, and
R. M.
Gamache
,
Appl. Phys. Lett.
90
,
221910
(
2007
).
37.
C. M.
Roland
and
R.
Casalini
,
Polymer
48
,
5747
5752
(
2007
).
38.
L.
Loubeyre
,
M.
Ahart
,
S. A.
Gramsch
, and
R. J.
Hemley
,
J. Chem. Phys.
138
,
174507
(
2013
).
39.
P. H.
Mott
and
C. M.
Roland
,
Macromolecules
31
,
7095
7098
(
1998
).
40.
J.
Qiao
,
A. V.
Amirkhizi
,
K.
Schaaf
,
S.
Nemat-Nasser
, and
G.
Wu
,
Mech. Mater.
43
,
598
607
(
2011
).
41.
W. F.
Oliver
,
C. A.
Herbst
,
S. M.
Lindsay
, and
G. H.
Wolf
,
Phys. Rev. Lett.
67
,
2795
(
1991
).
42.
M.
Henkel
,
M.
Pleimling
, and
R.
Sanctuary
,
Ageing and the Glass Transition
(
Springer
,
2007
), Vol.
716
.