Secondary Relaxation in Supercooled Liquid Propylene Glycol under

Sep 4, 2018 - This is the second (after glycerol) example of observation of Johari–Goldstein relaxation under ultrahigh pressures P > 2 GPa. View: A...
0 downloads 0 Views 418KB Size
Subscriber access provided by Kaohsiung Medical University

B: Fluid Interfaces, Colloids, Polymers, Soft Matter, Surfactants, and Glassy Materials

Secondary Relaxation in Supercooled Liquid Propylene Glycol under Ultra-High Pressures Revealed by Dielectric Spectroscopy Measurements Mikhail V. Kondrin, Alexei A. Pronin, and Vadim V. Brazhkin J. Phys. Chem. B, Just Accepted Manuscript • DOI: 10.1021/acs.jpcb.8b07328 • Publication Date (Web): 04 Sep 2018 Downloaded from http://pubs.acs.org on September 8, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Secondary Relaxation in Supercooled Liquid Propylene Glycol under Ultra-High Pressures Revealed by Dielectric Spectroscopy Measurements M.V.Kondrin,∗,† A. A. Pronin,‡ and V. V. Brazhkin† †Institute for High Pressure Physics, Russian Academy of Sciences, Troitsk, Moscow, 108840 Russia ‡General Physics Institute, Russian Academy of Sciences, Moscow, 117942 Russia E-mail: [email protected]

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Abstract 1-2-propanediol (propylene glycol) is a well-known glassformer, which easily vitrifies under wide range of cooling rates. An interesting feature of propylene glycol is that, similar to glycerol, it retains one-mode primary relaxation (slow α process) under a wide range of external P − T conditions. It was demonstrated that the emergence of secondary (β) relaxation requires the application of very high pressures P > 4.5 GPa. In this pressure range, the observation of secondary relaxation is partially obfuscated by the presence of strong decoupling of the static (ionic) conductivity and primary relaxation (the fractional Debye–Stokes–Einstein effect). However, secondary relaxation can be unambiguously extracted from experimental data by the correlation procedure of the imaginary and real parts of the dielectric response by means of Cole–Cole plots. This is the second (after glycerol) example of observation of Johari–Goldstein relaxation under ultra-high pressures P > 2 GPa.

Introduction The electric permittivity spectra of supercooled liquids exhibit several relaxation features, which are believed to be the key to achieve a better understanding of the glass transition. The most prominent of them is the structural relaxation (α) characterized by a strong dependence of the relaxation time on temperature and pressure, traced over a wide range of frequency. In addition, all glassformers demonstrate faster secondary or β-relaxation processes. They appear as a pronounced peak in dielectric loss spectra (type B systems) or as an excess wing on the high-frequency side of the structural relaxation (type A). Secondary relaxations named Johari–Goldstein 1 (JG) relaxation similarly to α-peak are caused by the motion of the molecules as a whole. They are believed to be the intrinsic properties of supercooled liquids but their microscopic origin is still questioned. JG relaxations should be distinguished from relaxations caused by intramolecular motions. The application of external pressure can help “separate” JG relaxation from the main relaxation and so to say converts type-A glassformers 2

ACS Paragon Plus Environment

Page 2 of 21

Page 3 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

in type-B ones. However, the crossover pressure turns out prohibitively high in many cases (for glycerol, this pressure is P > 2.5 GPa). In this paper, we consider propylene glycol (PG) that is a well-known small molecular glassformer. 2,3 It is characterized by a high dielectric response consisting of the single α process, which does not split by the application of a moderate pressure P < 1 GPa. 4,5 For this reason, we believe that it is a good candidate to check whether application of higher pressures can influence its excess wing and lead to the emergence of pronounced secondary relaxation. The parameters characterizing the vitrification process of PG (glass-transition temperature, fragility index, etc.) at high pressure can be established too. Besides that, interesting information can be obtained regarding correlation between the static conductivity and characteristic frequency of α-relaxation. In the normal liquids, the most general relation between the ionic viscosity ηi and the ionic conductivity σ0 is the Nernst–Einstein law: σ0 1 = 2 ηi ne The substitution of Maxwell’s relation ηi = βG∞ /ν0 into this equation yields the Debye– Stokes–Einstein (DSE) equation for the conductivity σ0 /ν0 ≈ const , which is actually observed in experiments with simple molecular glassformers. 6 Deviations from this proportionality, where ν0 is associated with some charge relaxation process, are occasionally observed in supercooled liquids. Such deviations are commonly described by the phenomenological relation σ0 ∼ ν0s 0 < s < 1, which was named fractional DSE (fDSE) relation. 7–18 However, this effect is observed rather sporadically and the origin of the fDSE relation is unknown. It is interesting to compare results obtained for PG and for its two derivatives: glycerol and propylene carbonate (PC). Glycerol is a weakly hydrogen-bonded glassformer but its bonds are stronger than that of PG. Propylene carbonate is a van-der-Waals-bonded liquid. Influence of high pressure on hydrogene bonding is poorly studied subject. 19 It was shown previously that the application of pressures above 2.5 GPa leads to onset of distinct secondary

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

relaxation in glycerol 20,21 and no transformation of the relaxation process was observed in PC 22 at pressures up to 4.2 GPa. So, one might expect an intermediate behavior for PG, which ultimately will be demonstrated in our paper.

Methods The samples of 99.5%-purity 1-2-propanediol were obtained from Acros Organics and were used as delivered. The high-pressure dielectric spectroscopy experiments were carried out in Toroid type anvils 23 in the P −T region P < 5.2 GPa and 150 K < T < 373 K. As a sensor, we used a plane capacitor made of copper plates separated by Teflon spacers (with a thickness of 0.1 − 0.2 mm) and 1.5-mm-diameter holes punched in them. Its typical empty capacitance was about 10 pF. The measurement capacitor was immersed in the liquid contained in a Teflon cell. Thus, the studied liquid also served as a pressure-transmitting medium. The signal from the capacitor was recorded by a Novocontrol-Alpha analyzer in the frequency range of 1 Hz – 2 MHz. The pressure and temperature sensors consisting of a manganin wire and a chromel–alumel thermocouple were also placed in the Teflon cell. The overall accuracies of determination of the pressure and temperature were better than 0.05 GPa and 0.1 K, respectively. In every experimental run, two or three isobars and isotherms were swept in a “stepwise” manner. That is, we first increased the pressure until the α-relaxation frequency became lower than the range of our measurements. After that, we started to heat our setup until this frequency became higher than our frequency range, etc. In the process of experiment, we took special care to prevent the transition of the sample to the glass state (which could be easily checked by the breakage of a manganin sensor). The rate of pressure and temperature variation was not strictly controlled but the typical rates can be estimated as 0.02 GPa/min and 0.5 K/min, respectively. Dielectric spectroscopy data presented in this paper were obtained in several runs of this type. The frequency-dependent complex dielectric response was obtained from the raw ca-

4

ACS Paragon Plus Environment

Page 4 of 21

Page 5 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

pacitance C(ν) and conductivity σ(ν) (converted to unit length by multiplication with the appropriate geometric factor l/S, where l and S are the length and surface area of the capacitor, respectively): ε(ν) = ε′ + iε′′ =

σ(ν) C(ν) +i ε0 2πε0 ν

This response can be described by the general formula

ε(ν) = i

σ0 X + ∆εi Ri (ν/νi ) + ε∞ ν

(1)

All coefficients in this formula are real (and physically meaningful) values: σ0 is the static conductivity (in reduced values of the permittivity of free space ε0 ), ∆εi are the intensities of several relaxation processes Ri with characteristic frequencies νi contributing to the dielectric response and ε∞ is the dielectric constant at infinite frequency (in practice, this means the frequency at which the influence of all considered relaxation processes is negligible). Note that, due to the Kramers–Kronig relation, ∆εi are real constants that allows one to fit the real and imaginary parts of experimental data simultaneously. Throughout this paper, we adopt the Cole–Davidson function (due to its simplicity) for the description of the dielectric relaxation response: RCD (ν) =

1 (1 − iν/ν0 )β

(2)

As an example, the fit of this formula to the experimental data collected at P = 3.2 GPa is shown in Fig. 1.

Results The temperature evolution of the characteristic frequency and parameter β of the main relaxation process along several isobars in the pressure range P < 3.2 GPa is shown in Fig. 2. It should be noted that dielectric relaxation in this pressure range still retains its unimodal character and an increase in pressure results in the broadening of the relaxation 5

ACS Paragon Plus Environment

The Journal of Physical Chemistry

log10(ν Hz) 0

1

2

3

4

5

6

5

6

2

log10(ε ″)

1.5

1

0.5

0 300

T (K) 320 340

360

1.8 1.6 log10(ε ′)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 21

1.4 1.2 1 0.8 0

1

2

3

4

log10(ν Hz)

Figure 1: Fit of Eq. 1-2 to the dielectric spectroscopy data obtained at P = 3.2 GPa. The color of each curve specifies the temperature at which it was collected.

6

ACS Paragon Plus Environment

Page 7 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

process, which can be deduced from a decrease in the parameter β from ≈ 0.6 at ambient pressure to ≈ 0.2 at P = 3.2 GPa. The temperature dependence of the characteristic frequency obtained for the main dielectric relaxation process can be well interpolated by the Waterton–Mauro 24,25 relation (see Fig. 2 b): log10 (ν0 ) − log10 (ν∞ ) =

E A exp( ) T T

(3)

Previous studies 26,27 demonstrated that the Waterton–Mauro relation has many benefits in comparison to the more known Vogel–Fulcher–Tammann equation. Relation (3) catches a non-divergent character of the relaxation times, which is likely present in glasses, the more or less realistic values of fitting parameters (ν∞ and E), and, the last but not the least, the more robust fitting of experimental data collected in the limited temperature range. Both the Waterton–Mauro and Vogel–Fulcher–Tammann laws yield approximately similar results for the composite parameters of a vitrification process such as the fragility index   mp ≡ d log10 ν10 /d TTg |T =Tg (measure of deviation of the relaxation time from the Arrhenius behavior) and the dielectric glass-transition temperature Tg (by convention, the temperature at which ν0 = 10−3 Hz). The “dielectric” glass-transition temperatures (obtained by the extrapolation of dielectric relaxation times to ν0 = 10−3 Hz) and fitted parameters in Eq. 3 are presented in Table 1, where the fragility index is related to the parameter A in Eq. 3 as mp = A/Tg exp(E/Tg )(1 + E/Tg ). Table 1: Glass-transition temperature Tg , fragility index mp , and the parameters log10 (ν∞ ) and E of the Waterton–Mauro equation (3) in propylene glycol at high pressures P , GPa 0.0 0.8 1.6 2.2 3.2

Tg , K 164.43 195.37 236.97 251.35 279.27

∞ log10 ( νHz ) 15.25 11.78 15.47 13.47 14.67

mp 42.47 42.77 47.94 47.32 51.27

E, K 382.6 565.2 615.1 722.14 810.1

The comparison of glass-transition temperatures Tg and fragility indices mp in PG, PC, 7

ACS Paragon Plus Environment

The Journal of Physical Chemistry

0.003

0.0035

1 , K−1 T 0.004 0.0045

0.005

0.0055

0.006

0.6

β

0.5 0.4

a)

0.3 6

b)

4

log10(ν0 Hz)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 21

2

0

−2

0.003

0.0035

0.004

0.0045

0.005

0.0055

0.006

1 , K−1 T

Figure 2: Plots of the relaxation parameter β (see Eq. 2, panel a) and characteristic frequency ν0 (panel b) along several isobars P = (+)0, (◦) 0.8, () 1.6, (♦) 2.2, and (△) 3.2 GPa. Thin lines are fits by Eq. 3.

8

ACS Paragon Plus Environment

1

3

4

250

300

0

P, GPa 2

200

Tg , K

90

a)

70

80

b)

50

60

mp

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

350

Page 9 of 21

0

1

2

3

4

P, GPa Figure 3: Parameters Tg (panel a) and mp (panel b) for () propylene glycol, (♦) PC, and (◦) glycerol. The data for PC and glycerol are taken from Refs. 21,22 Solid curves are guides to eye.

9

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

and glycerol reveals an unambiguous trend depending on the strength of intermolecular forces (Fig. 3). All three molecules have the same hydrocarbon backbone consisting of three carbon atoms. As was already mentioned, PC is a purely van-der-Waals bonded liquid, and PG and glycerol are weakly hydrogen bonded ones, although hydrogen bonds in glycerol are stronger because it has three hydrogen bonds per molecule while PG has only two. Correspondingly, the weaker the bonds, the steeper the pressure dependence of the glasstransition temperature and the weaker the pressure dependence of the fragility index (in PC, it is negative). At higher pressures, qualitative changes in dielectric relaxation in PG took place and secondary relaxation emerged at pressures P > 4.5 GPa. This secondary relaxation manifests itself as a small hump in the high-frequency wing of the imaginary part of dielectric relaxation data collected along the 350-K isotherm (see Fig. 4). Due to the decoupling of the static conductivity and characteristic frequency of primary relaxation, the peak corresponding to the α-process is submerged under the low-frequency wing corresponding to ionic conductivity. This effect makes the emergence of secondary relaxation less evident but it can be detected by means of the Cole–Cole plot, i.e., the plot of the imaginary vs. real part of the dielectric response (see Fig. 5). Because amplitudes of primary and secondary relaxations are oder of magnitude different it is convenient to draw the Cole-Cole plots in slightly modified manner that is in double logarithmic scale which makes changes in high-frequency flank of dielectric response (corresponding to the lower left corner of the Cole-Cole plot) more spectacular. The comparison of isobaric data collected at P = 3.2 GPa (Fig. 5 a) and along the 350-K isotherm in the pressure range 2.2 GPa < P < 5.2 GPa (Fig. 5 b) clearly demonstrates a qualitative difference between them and obviously indicates that the secondary relaxation attains a significant amplitude at pressures P > 4.5 GPa. This pressure is significantly higher than the secondary relaxation onset pressure in glycerol (P = 2.5 GPa 21 ), which until now was a record-breaking value. We remind that similarly to PG glycerol has distinct secondary relaxation at high pressure (see Fig. 6 a) while in PC only one relaxation is observed upto

10

ACS Paragon Plus Environment

Page 10 of 21

Page 11 of 21

log10(ν Hz) 0

1

2

3

5

4

6

3 2.5 2.5 3 3.5 4 4.5 5

log10(ε ″)

2

P (GPa) 1.5 1 0.5 0

2 1.8 1.6 log10(ε ′)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1.4 1.2 1

0

1

2

3

4

5

6

log10(ν Hz)

Figure 4: Relaxation in propylene glycol along the 350-K isotherm at P = 2.2 − −5.2 GPa. The color of each curve specifies the pressure at which it was collected.

11

ACS Paragon Plus Environment

The Journal of Physical Chemistry

a)

300

320

T (K) 340

360

log10(ε ″)

1.5

1

0.5

0

1

1.2

1.4

1.6

1.8

log10(ε ′)

3

b) 2.5 3 3.5 4 4.5 5

2.5

P (GPa)

2 log10(ε ″)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 21

1.5

1

0.5

0

1

1.2

1.4

1.6

1.8

2

log10(ε ′)

Figure 5: The Cole–Cole plots of dielectric relaxation (panel a) along the 3.2-GPa isobar at T = 295−375 K and (panel b) along the 350-K isotherm at P = 2.2−−5.2 GPa. Qualitative difference between the plots demonstrates the emergence of secondary β relaxation at T = 350 K and P > 4.5 GPa. The color of each curve specifies the pressure or temperature at which it was collected. 12

ACS Paragon Plus Environment

Page 13 of 21

a)

2

log10(ε ″)

1.5

1

0.5

300

310

0

320

330

T, K

1.2

1.4

1.6

1.8

log10(ε ′)

1

b)

− 0.5

log10(ε ″) 0 0.5

P, GPa 3.4 3.6 3.8

4

4.2

−1

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1

1.2

1.4

1.6

1.8

log10(ε ′)

Figure 6: Correlation between the real and imaginary parts of the permittivity (the Cole– Cole plot) in glycerol (P = 4 GPa, T = 290 − 340 K, panel a) and PC (T = 410 K, P = 3.3 − 4.3 GPa, panel b). The color of each curve specifies the pressure or temperature at which it was collected. The data from Refs. 21,22 were used.

13

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

4.2 GPa (see Fig. 6 b). In our opinion this difference is connected with the strength of the intermolecular forces present in the liquids: stronger the bonds the lower is the pressure where onset of the secondary relaxation is observed. It is conjectured that for PC even higher pressures should be applied (about 8 GPa or higher) to reveal the secondary relaxation. It is also interesting to note an enhancement of the fDSE effect in PG under high pressure, which leads to the decoupling of the static conductivity and frequency of primary relaxation mentioned above (see Fig. 1). The relation between the static conductivity and characteristic frequency in PG at ambient pressure is almost linear (typical Debye–Stokes–Einstein relation). However, nonlinear deviation occurs at higher pressures and a power law σ0 ∼ ν00.67 is observed at P = 3.2 GPa (fDSE, 7–18 Fig. 7). It should be noted that a similar behavior has already been detected in PG and its oligomers at high pressure. 4,5 However, the origin of these deviations is not well understood. Recently, it was found that an increase in pressure above a threshold value of 0.5 GPa results in a fractional relation between the static conductivity and viscosity in 2E1H 28 and PC 29 too, although the normal DSE is observed at ambient pressure in these glassformers. Recently, 29 it was proposed that a similar deviation from the DSE relation in PC at high pressure can be caused by the interference of some high-frequency process present in liquids 30–32 in the terahertz range. This additional relaxation process provides so to say a “shunting” mechanism of charge transfer, which appears at high pressures and low temperatures where the static conductivity ensured by the main relaxation process is too low.

Conclusions To summarize, we have established that the application of high pressure leads to the splitting of main α-relaxation in PG and appearance of secondary fast relaxation as a separate peak on the high-frequency side of the imaginary part of the α-process. Pressures at which this splitting takes place are higher than 4.5 GPa and are significantly higher than that for

14

ACS Paragon Plus Environment

Page 14 of 21

Page 15 of 21

50

40

σ0 (kHz)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

30

20

10

0

200

400

600

800

1000

1200

ν0 (kHz) Figure 7: Correlation between the static conductivity σ0 and characteristic frequency ν0 at P = 3.2 GPa. The nonlinear dependence is a result of the fDSE effect.

15

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

glycerol (P > 2.5 GPa 21 ). There is a definite dependence of the secondary relaxation onset pressure on the strength of intermolecular forces in the sequence of glycerol, PG, and PC. The weaker the forces, the higher this pressure. For weakly van-der-Waals bonded PC, there are no traces of α-relaxation splitting up to the pressure P = 4.2 GPa. It is remarkable that the pressure-dependent glass-transition temperature and fragility index also significantly correlate with the type of intermolecular forces and their strength. The application of high pressure also leads to decoupling of the static conductivity and the characteristic relaxation frequency, which takes place at pressures above 3 GPa. As a result, we have observed a nonlinear relation between the static conductivity and characteristic frequency at high pressures, which can be described by the fDSE relation with an exponent of 0.67. The strong difference of this exponent at ambient (where it equals 1) and high pressures makes PG a convenient experimental object where microscopic mechanism of ionic transport responsible for the fDSE relation can be studied.

Acknowledgement This work was supported by the Russian Science Foundation (grant no. 14-22-00093). The work of A.A.P. was supported by the Russian Foundation for Basic Research (grant no. 16-02-01120).

References (1) Johari, G. P.; Goldstein, M. Viscous Liquids and the Glass Transition. II. Secondary Relaxations in Glasses of Rigid Molecules. The Journal of Chemical Physics 1970, 53, 2372–2388. (2) Davidson, D. W.; Cole, R. H. Dielectric Relaxation in Glycerol, Propylene Glycol, and n-Propanol. The Journal of Chemical Physics 1951, 19, 1484–1490.

16

ACS Paragon Plus Environment

Page 16 of 21

Page 17 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(3) Leo´ n, C.; Ngai, K. L.; Roland, C. M. Relationship between the Primary and Secondary Dielectric Relaxation Processes in Propylene Glycol and Its Oligomers. The Journal of Chemical Physics 1999, 110, 11585–11591. (4) Casalini, R.; Roland, C. M. Dielectric α-Relaxation and Ionic Conductivity in Propylene Glycol and Its Oligomers Measured at Elevated Pressure. The Journal of Chemical Physics 2003, 119, 11951–11956. (5) Casalini, R.; Roland, C. M. Excess Wing in the Dielectric Loss Spectra of Propylene Glycol Oligomers at Elevated Pressure. Phys. Rev. B 2004, 69, 094202. (6) Stickel, F.; Fischer, E. W.; Richert, R. Dynamics of Glass-Forming Liquids. II. Detailed Comparison of Dielectric Relaxation, dc-Conductivity, and Viscosity Data. The Journal of Chemical Physics 1996, 104, 2043–2055. (7) Koike, T. Study of Dielectric Relaxation of an Epoxide Oligomer by a Direct Current Transient Method. Polymer Engineering & Science 1993, 33, 1301–1307. (8) Corezzi, S.; Capaccioli, S.; Gallone, G.; Livi, A.; Rolla, P. A. Dielectric Behaviour Versus Temperature of a Monoepoxide. Journal of Physics: Condensed Matter 1997, 9, 6199. (9) Corezzi, S.; Capaccioli, S.; Gallone, G.; Lucchesi, M.; Rolla, P. A. Dynamics of a GlassForming Triepoxide Studied by Dielectric Spectroscopy. Journal of Physics: Condensed Matter 1999, 11, 10297. (10) Corezzi, S.; Lucchesi, M.; Rolla, P. A.; Capaccioli, S.; Gallone, G.; Paluch, M. Temperature and Pressure Dependences of the Relaxation Dynamics of Supercooled Systems Explored by Dielectric Spectroscopy. Philosophical Magazine Part B 1999, 79, 1953– 1963.

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(11) Bielowka, S. H.; Psurek, T.; Ziolo, J.; Paluch, M. Test of the Fractional Debye-StokesEinstein Equation in Low-Molecular-Weight Glass-Forming Liquids Under Condition of High Compression. Phys. Rev. E 2001, 63, 062301. (12) Psurek, T.; Hensel-Bielowka, S.; Ziolo, J.; Paluch, M. Decoupling of the DC Conductivity and α-Structural Relaxation Time in a Fragile Glass-Forming Liquid under High Pressure. The Journal of Chemical Physics 2002, 116, 9882–9888. (13) Psurek, T.; Ziolo, J.; Paluch, M. Analysis of Decoupling of DC Conductivity and Structural Relaxation Time in Epoxies with Different Molecular Topology. Physica A: Statistical Mechanics and its Applications 2004, 331, 353–364. (14) Paluch, M.; Wojnarowska, Z.; Hensel-Bielowka, S. Heterogeneous Dynamics of Prototypical Ionic Glass CKN Monitored by Physical Aging. Phys. Rev. Lett. 2013, 110, 015702. (15) Drozd-Rzoska, A.; Rzoska, S. In Metastable Systems under Pressure; Rzoska, S., DrozdRzoska, A., Mazur, V., Eds.; NATO Science for Peace and Security Series A: Chemistry and Biology; Springer Netherlands, 2010; pp 141–149. (16) Wojnarowska, Z.; Wang, Y.; Pionteck, J.; Grzybowska, K.; Sokolov, A. P.; Paluch, M. High Pressure as a Key Factor to Identify the Conductivity Mechanism in Protic Ionic Liquids. Phys. Rev. Lett. 2013, 111, 225703. (17) Rodrigues, A. C.; Viciosa, M. T.; Dan`ede, F.; Affouard, F.; Correia, N. T. Molecular Mobility of Amorphous S-Flurbiprofen: A Dielectric Relaxation Spectroscopy Approach. Molecular Pharmaceutics 2014, 11, 112–130. (18) Abou Elfadl, A.; El-Sayed, S.; Hassen, A.; Abd El-Kader, F.; Said, G. Charge Transport and Glassy Dynamics in Polyisoprene. Polymer Bulletin 2014, 71, 2039–2052.

18

ACS Paragon Plus Environment

Page 18 of 21

Page 19 of 21 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(19) Fanetti, S.; Citroni, M.; Dziubek, K.; Nobrega, M. M.; Bini, R. The Role of H-bond in the High-Pressure Chemistry of Model Molecules. Journal of Physics: Condensed Matter 2018, 30, 094001. (20) Pronin, A. A.; Kondrin, M. V.; Lyapin, A. G.; Brazhkin, V. V.; Volkov, A. A.; Lunkenheimer, P.; Loidl, A. Glassy Dynamics under Superhigh Pressure. Phys. Rev. E 2010, 81, 041503. (21) Pronin, A. A.; Kondrin, M. V.; Lyapin, A. G.; Brazhkin, V. V.; Volkov, A. A.; Lunkenheimer, P.; Loidl, A. Pressure-Induced Change in the Relaxation Dynamics of Glycerol. JETP Letters 2010, 92, 479–483. (22) Kondrin, M. V.; Gromnitskaya, E. L.; Pronin, A. A.; Lyapin, A. G.; Brazhkin, V. V.; Volkov, A. A. Dielectric Spectroscopy and Ultrasonic Study of Propylene Carbonate under Ultra-High Pressures. The Journal of Chemical Physics 2012, 137, 084502. (23) Khvostantsev, L. G.; Slesarev, V. N.; Brazhkin, V. V. Toroid Type High-Pressure Device: History and Prospects. High Pressure Research: An International Journal 2004, 24, 371–383. (24) Waterton, S. C. The Viscosity-Temperature Relationship and Some Inferences on the Nature of Molten and of Plastic Glass. J. Soc. Glass Technol. 1932, 16, 244–249. (25) Mauro, J. C.; Yue, Y.; Ellison, A. J.; Gupta, P. K.; Allan, D. C. Viscosity of GlassForming Liquids. Proceedings of the National Academy of Sciences 2009, 106, 19780– 19784. (26) Lunkenheimer, P.; Kastner, S.; K¨ohler, M.; Loidl, A. Temperature Development of Glassy α -Relaxation Dynamics Determined by Broadband Dielectric Spectroscopy. Phys. Rev. E 2010, 81, 051504.

19

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(27) Kondrin, M. V.; Pronin, A. A.; Brazhkin, V. V. Crystallization and Vitrification of Ethanol at High Pressures. The Journal of Chemical Physics 2014, 141, 194504. (28) Pawlus, S.; Klotz, S.; Paluch, M. Effect of Compression on the Relationship between Viscosity and Dielectric Relaxation Time in Hydrogen-Bonded Primary Alcohols. Phys. Rev. Lett. 2013, 110, 173004. (29) Kondrin, M. V.; Brazhkin, V. V.; Lebed, Y. B. Fluctuation-Dissipation Theorem and the Dielectric Response in Supercooled Liquids. The Journal of Chemical Physics 2015, 142, 104505. (30) Trachenko, K. Heat Capacity of Liquids: An Approach from the Solid Phase. Phys. Rev. B 2008, 78, 104201. (31) Bolmatov, D.; Trachenko, K.; Brazhkin, V. V. The Phonon Theory of Liquid Thermodynamics. Sci. Rep. 2012, 2, 00421. (32) Kondrin, M. V. High Frequency Asymptotics of Dielectric Permittivity in Supercooled Liquids: Experimental Data and Conclusions of Mode Coupling Theory. Journal of Experimental and Theoretical Physics 2014, 119, 707–713.

20

ACS Paragon Plus Environment

Page 20 of 21

Page 21 of 21

0.5

log10(ε ″) 1

1.5

2

Graphical TOC Entry

T=350 K P=2.2−5.2 GPa 0

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1

1.2

1.4

1.6

1.8

log10(ε ′)

21

ACS Paragon Plus Environment