Anda di halaman 1dari 21

J. Chem.

Thermodynamics 113 (2017) 162–182

Contents lists available at ScienceDirect

J. Chem. Thermodynamics
journal homepage: www.elsevier.com/locate/jct

Review

The viscosity of glycerol


Abel G.M. Ferreira a,⇑, Ana P.V. Egas a, Isabel M.A. Fonseca a, Ana C. Costa b, Danielly C. Abreu b,
Lélio Q. Lobo a,1
a
Department of Chemical Engineering, University of Coimbra, Polo II, Rua Silvio Lima, 3030-970 Coimbra, Portugal
b
Departamento de Engenharia Química, Universidade Federal de São João del-Rei, Campus Alto Paraopeba – C.A.P, Rod.: MG 443, KM 7 Ouro Branco, MG 36420-000, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: The dynamic viscosities of glycerol were measured over the temperature range (293–394) K and atmo-
Received 11 February 2017 spheric pressure using a Brookfield thermosel system. The combined expanded uncertainty of reported
Received in revised form 29 May 2017 viscosity is better than 3.0% with a level of confidence 0.95 (k = 2). The presented results are in good
Accepted 30 May 2017
agreement with most values from the literature. An extensive viscosity database for this substance
Available online 1 June 2017
was developed combining the values of this work with those published in literature covering a wide
range of temperatures at atmospheric pressure from the calculated glass transition temperature
Keywords:
(Tg = 188 K), and measurements of viscosity reported over the temperature range (263–398) K at pres-
Glycerol
Viscosity
sures from (104 to 3) GPa. The main purpose of database construction was the development and evalu-
Glass transition ation of reliable correlation models of viscosity valid in wide ranges of temperature and pressure. The
BSCNF equation physically sound equations of Mauro (MYEGA) and the Bond Strength-Coordination Number
VFT equation Fluctuation (BSCNF) model were used to correlate values selected from the database. From BSCNF the
Scaling of viscosity structural effects taking place near the glass transition were discussed in light of results obtained by
recent experimental techniques. Some of the tested equations give overall absolute deviations less than
6% in the range (190–440) K, a value which is close the experimental uncertainty. Stickel derivatives were
calculated for the correlation equations and they were compared with values found from viscosity data
using numerical techniques. High-pressure viscosity data allowed studying the temperature, pressure
and density dependences of this property using the free volume theory and the thermodynamic scaling
of viscosity. The free volume proved to be more accurate in the calculation of viscosity in wide T and p
ranges with average absolute percentual deviations (AADs) ranging from 4% to 14% for data from different
authors. The pressure dependence of glass transition temperature and fragility index of glycerol were cal-
culated from model equations being the results in good agreement with data selected from the literature.
Ó 2017 Elsevier Ltd.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
2. Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
2.1. Chemicals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
2.2. Viscosity measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
3. Viscosity models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
4. Results and discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
4.1. Fixed points and sources of viscosity data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
4.2. Temperature dependence of viscosity at atmospheric pressure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
4.3. Influence of pressure on glass transition parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
4.4. Viscosity dependence on temperature, pressure and density. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
4.4.1. Free volume model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
4.4.2. Scaling of viscosity with density . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177

⇑ Corresponding author.
E-mail address: abel@eq.uc.pt (A.G.M. Ferreira).
1
Deceased 8 June 2016.

http://dx.doi.org/10.1016/j.jct.2017.05.042
0021-9614/Ó 2017 Elsevier Ltd.
A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182 163

4.4.3. Scaling of viscosity with pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179


5. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
Appendix A. Supplementary data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180

1. Introduction the intermolecular repulsive potential. This behaviour was


reported for the relaxation time of several glass-forming organic
It is well-known that among glass-forming liquids (GFL), liquids by Dreyfus et al. [12] who have shown that the viscosity
hydrogen-bonded systems tend to be unusual with regard to their of glycerol could obey to such scaling law. Thermodynamic scaling
supercooling propensity and highly complex behaviour in the of viscosity was performed by Roland et al. [13] for organic and
supercooled state. Glycerol (1,2,3-propanetriol) is one of the most ionic liquids, by Bair and Casalini [14] for liquids used in elastohy-
extensively studied hydrogen-bonded systems. The presence of drodynamic lubrication and Paluch et al. [15] studied organic liq-
three hydroxy groups in the molecule gives rise to particularly rich uids and polymers.
and complex conformational and structural behaviour and it is one The VFT and AM equations will be used in this work to describe
of the archetypal glass-forming liquids. The description of the ther- the temperature dependence of viscosity at isobaric conditions and
mophysical properties of supercooled liquid and the nature of its for this purpose the measurements made in this work were com-
glass transition has been the object of several experimental bined with selected data from the literature. BSCNF was used in
investigations. this context and also in the knowledge of phenomena related to
One of the most intriguing features of supercooled liquids is the the liquid glycerol structure near glass transition. Of particular
dramatic rise in the viscosity as it is cooled toward the glass tran- interest will be the calculation of the glass transition temperature
sition. The viscosity of glycerol has been measured using different and fragility as a function of pressure from correlation equations.
type of viscometers with consequent disagreement in data from The free volume theory and scaling laws were used to study the
the various sources. The differences in viscosity are more notorious dependence of viscosity on (PVT) variables.
at low temperature particularly near the glass transition where
equilibrium viscosity is difficult to obtain owing to the long struc- 2. Experimental
tural relaxation time. Thus it will be important to find out models
that best describes the temperature dependence of viscosity in a 2.1. Chemicals
wide range of temperature by using the fewest possible number
of fitting parameters. Because only two parameters are needed Glycerol was obtained from Sigma-Aldrich (CAS number 56-81-
for a simple Arrhenius description, modeling of super-Arrhenius 5) with stated mass fraction purity (GC)  0.995 (water by Karl
behaviour near glass transition will require a minimum of three Fisher  0.001). Since glycerol is highly hygroscopic, samples were
parameters. Several three-parameter models for temperature degassed ultrasonically, dried over freshly activated molecular
dependence viscosity of glass-forming have been used, being the sieves (Type 3 Å) supplied by Aldrich and further purified by evap-
most popular the Vogel-Fulcher-Tamman (VFT) [1–3], the Adam– oration in a rotary evaporator working at 343 K. The water content
Gibbs (AG) [4] and the Avramov-Milchev (AM) [5–7] equations. was determined with a Metrohm 831 Karl Fisher coulometer indi-
González et al. [8] proposed the use of AM equation with two cating that the purification procedure reduced the water mass frac-
parameters to represent the temperature dependence of viscosity tion from 1.3  103 to 7  104.
from the glass transition temperature, Tg, up to 450 K. However,
the analysis based on the use of these equations is not sufficient
2.2. Viscosity measurements
to fully understand the physics behind the structural relaxation
near glass transition. Aniya [9] proposed the Bond Strength-
Dynamic viscosity measurements were made over the temper-
Coordination Number Fluctuation (BSCNF) model, to describe the
ature (293–394) K at atmospheric pressure using a Brookfield
temperature dependence of the viscosity of GFLs in terms of energy
Thermosel system with a rotational DV-II+(model LVDV-II) vis-
and coordination number parameters of the structural units form-
cometer. Temperature was controlled to better than ±0.1 K. The
ing the liquid to account for the structural effects taking place near
stirring action of the rotating spindle, plus the small sample vol-
the glass transition. The BSCNF model incorporates the VFT
ume helped to keep the temperature gradient across the sample
description under special physical conditions.
to a minimum. Direct readout of sample temperature was provided
For pressure dependence of the viscosity data, the free volume
with the RTD sensor of viscometer (accuracy of ±0.1 K) inserted
(or excluded-volume) theory is often used with quite good results
into the hole of the thermo container. However, for each measure-
over a wide range of pressure. In this theory, the volume depen-
ment, the temperature was also measured by a platinum resistance
dence of viscosity is expressed by three-parameter Doolittle’s
thermometer ERTCO-Eutechnics High Precision Digital Thermome-
equation [10] which usually gives reasonable fit over ca. ten dec-
ter certified in the ITS90 with accuracy of ±0.05 K. The spindles
ades of experimental high-pressure viscosity data for different
SSA/18/13R and SSA/31/13R were used for the viscosity measure-
classes of substances and also provides good estimates of the glass
ments with sample volumes of (8.0 and 10) cm3, respectively,
transition pressures (using extrapolation to viscosity g = 1012 Pas)
and rotational speeds from (5 to 200) RPM were used. To minimize
[11].
contamination of the sample with dust and moisture, we impro-
In the last two decades, several authors have shown that both
vised a system consisting of perspex cylinders perfectly set
relaxation time, s, and viscosity g, of GFLs at different pressure-
between the upper surface of the thermosel unit and the body of
volume-temperature (PVT) conditions can be scaled into a master
the viscometer. For each temperature, the viscosity was measured
curve, s(or g) = F(TV c) where F is the function used to describe
for increasing rotational speeds and after following decreasing
the master curve, V is the volume, and c is a parameter related
speeds and the reading of viscosity was made only when there is
to exponent (3c) of the power law describing the local slope of
no change in the temperature.
164 A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182

According Brookfield manuals, the viscosity measurements are values of m and they are fragile systems. Glycerol has an interme-
guaranteed to be accurate to within ±1% of the full-scale viscosity diate behaviour (values of m are in the range 48 [17] to 54 [18] at
range for a given spindle/speed combination (this percentage, atmospheric pressure).
expressed in mPas, is equal to the spindle factor) and the repro- Typically, the equations for viscosity are of the form ln g = A + F
ducibility is to within ±0.2% [16]. For SSA/18/13R and SSA/31/13R (T), where F(T) is an analytical function of temperature which has
spindle factors are 30/N and 300/N where N is the rotational speed the property that F(T) ? 0 (and thus g ? eA) as T ? 1. The
(RPM). The full-scale range of viscosity corresponding to any spin- Vogel-Fulcher-Tammann (VFT) equation [1–3]:
dle/speed combination is the factor multiplied by 100. As the same
BVFT
accuracy of ±1% applies to the readings in the range, the accuracy ln g ¼ AVFT þ ð2Þ
T  T0
of viscosity measurement will increase as the reading approaches
100% of full-scale. Thus it is possible to measure almost the same is the most popular three-parameter equation used for the analysis
viscosity value with different error levels at different speeds. Con- of the temperature dependence of viscosity. In Eq. (2), AVFT = ln g1
sidering the following example from our measurements: for glyc- is the pre-exponential factor equivalent to the value of the viscosity
erol at 383.29 K, using spindle SSA/18 at N = 100 RPM the at the high temperature limit. The constants AVFT, BVFT and T0 are the
measured viscosity was 11.10 mPas. For this spindle/speed combi- fitting parameters. Furthermore, BVFT = DVFT T0 where DVFT is the
nation the factor is 0.3, the full-range scale viscosity is 30 mPas, strength parameter and T0 is the Vogel temperature, also called
and the possible error will be ±1% of this range, or ±0.3 mPas. Thus ‘‘ideal glass transition temperature” at which all movements of
the readings on the viscosimeter will be g = 11.10 mPas (37% of the constituent elements of a liquid are considered to be totally fro-
full-scale), and the error is ±0.3 mPa.s (relative error of 2.7%). zen. By choosing adequately DVFT and T0, Eq. (2) can reproduce rel-
Changing the speed to N = 200 RPM, the factor changes to 0.15, atively well the viscosity behaviours of strong and fragile systems.
the full-range scale viscosity to 15 mPas, and the possible error However, usually the temperature dependence of viscosity of
will be ±0.15 mPas (relative error of 1.4%). The measured viscosity supercooled liquids cannot be described over the entire range of
at stable conditions was 11.00 mPas which corresponds to 73.3% temperature with a single VFT equation [8,19]. For this reason other
of full-scale. Whenever possible, measurements were made at equations can be preferred for data correlation over large ranges of
speeds corresponding to the higher percentage readings for the temperatures.
spindle. At each temperature the viscosity was measured several The Avramov and Milchev (AM) equation [5–7] is another
times and the mean value of measurements and corresponding three-parameter model (AM) derived based on the temperature
standard deviation (usually less than 0.6% relative to the mean dependence of average jump frequency. It is assumed that mole-
value) were derived providing the uncertainty of measurements. cules in a flowing liquid jump from the holes formed by the nearest
Prior to measurements, the viscometer was calibrated with neighbours to one of the adjoining holes and that the viscosity of
Brookfield viscosity standards at 298.15 K. The used standards the liquid is inversely proportional to the average frequency of
were silicone oil fluids with Newtonian behaviour, certified by these jumps. As a final result the dependence of liquid viscosity
methods traceable to the United States National Institute of Stan- on temperature is:
dards and Technology (NIST). For the standard with given viscosity
 C
49.9 mPa.s we have measured g = (50.9 ± 0.1) mPas at T = 298.21 K BAM AM
and for that with viscosity 975 mPa.s the corresponding measure- ln g ¼ AAM þ ð3Þ
T
ment was g = (992 ± 0.3) mPas at T = 298.20 K. Therefore the accu-
racy of measurements relative to viscosity standards is better than where AAM = ln g1, BAM and CAM are fitting parameters.
2%. Recently, Mauro et al. [20] proposed a new model for represen-
Taking into account the uncertainties of temperature and tation of viscosity of a large variety of glass forming liquids over a
repeatability of measurements, the relative combined expanded wide range of temperature. Using the temperature-dependent con-
uncertainty of viscosity with level of confidence 0.95 (k = 2) is esti- straint model of Gupta and Mauro [21], which relates the configu-
mated to be better than 3.0%. rational entropy with topological degrees of freedom per atom, the
The measurement of the atmospheric pressure was made using resulting equation (MYEGA) is [20]:
a calibrated pressure transducer (AFRISO Euro-Index, DMU03).
 
Taking the observed values covering January to March 2017 the BMYEGA ðCMYEGA =TÞ
lng ¼ AMYEGA þ e ð4Þ
mean value was p = (101.86 ± 0.23) kPa in close agreement with a T
previous 2016 value p = (102.24 ± 0.39) kPa.
with fitting parameters AMYEGA = ln g1, BMYEGA and CMYEGA.
Eqs. (2)–(4) can be rewritten as equivalent equations in terms of
3. Viscosity models
the same set of physically meaningful quantities namely the glass
transition temperature, Tg, the fragility index, m, the high temper-
Viscosity is often represented on an Angell plot of logarithm of
ature asymptotic limiting viscosity, ln g1, and the viscosity at Tg, ln
the viscosity versus the ratio (x = Tg/T), which is a Tg-scaled version
gg. From the definition of m, taking into account x (=Tg/T) the Eqs.
of an Arrhenius plot. For fixed pressure, Angell plots give curves
(2)–(4) are arranged as:
with different degrees of non-Arrhenius behaviour. The degree of
½ln ðgg =g1 Þ
2
deviation from the Arrhenius behaviour is called fragility and it
VFT : ln g ¼ ln g1 þ ð5Þ
is widely accepted that the viscous flow behaviours of any kind ½lnð10Þ mðx  1Þ þ ln ðgg =g1 Þ
of GFLs are characterized by the fragility index defined as [17]:
  where
d log g
m¼ ð1Þ ½ln ðgg =g1 Þ
2
dðT g =TÞ T¼T g BVFT ¼ T g ð6Þ
lnð10Þ m
Small values of m are assigned to glass formers materials with
 
high polymerized networks such as SiO2 (m = 20) and they are lnðgg =g1 Þ
called strong materials. On the other hand, the systems with T0 ¼ Tg 1  ð7Þ
lnð10Þ m
non-directional interatomic or intermolecular bonds exhibit high
A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182 165

 
lnð10Þ m the only parameter to be found will be B (=B⁄) provided that ln
AM : ln g ¼ ln g1 þ ln ðgg =g1 Þ x ln ðgg =g1 Þ
ð8Þ gg and ln g1 are fixed. The equation (BSCNF1) obtained under
those conditions is:
where
ln g ¼ ln g1
BAM ¼ T g ln ðgg =g1 Þ ln ðgg =g1 Þ = lnð10Þ m ð9Þ h   i
1=2 g
ð1  B Þ ðx þ B x2 Þ ln g g þ 0:5 lnð1  B Þ
1
þ
ð1 þ B Þð1  B x2 Þ
1=2
lnð10Þ m
C AM ¼ ð10Þ
ln ðgg =g1 Þ  0:5 lnð1  B x2 Þ ð17Þ

" ! # Ikeda and Aniya [22] proved that when c = 1, BSCNF1 is equiv-
lnð10Þm alent to VFT equation and then the following relation between
MYEGA : ln g ¼ ln g1 þ xlnðgg =g1 Þ exp  1 ðx  1Þ
lnðgg =g1 Þ ideal glass transition temperature, T0, and Tg is [22]:
 pffiffiffiffi
ð11Þ 1þ B
C  0:5 lnð1  B Þ
T0 1B
¼1 ð18Þ
where Tg lnð10Þ m
" #
lnð10Þ m or
BMYEGA ¼ T g ln ðgg =g1 Þ exp 1  ð12Þ  
ln ðgg =g1 Þ g
ln g g
T0 1
¼1 ð19Þ
" # Tg lnð10Þ m
lnð10Þ m
C MYEGA ¼ T g 1 ð13Þ An important and interesting quantity obtained from BSCNF1 is
ln ðgg =g1 Þ
the number of structural units broken defined as NB = Eg / (E0Z0)
In the course of lowering the temperature, or increasing pres- where Eg is the activation energy for the viscous flow and (E0Z0)
sure the value of the viscosity increases drastically reaching is the average total binding energy per one structural unit. From
approximately 1012 Pas at the glass transition temperature. The BSCNF1 model [23]:
glass-forming liquid is formed by an agglomeration of clusters or h   i
g
B  C  þ 2 ln g g þ 0:5 lnð1  B Þ
structural units which are bound to others by a certain bond NB ¼
1
ð20Þ
strength retaining its spatial random connectivity. Thermally acti- ð1  B ÞC 
vated viscous flow occurs due to bond-breaking and bond-
The physical meaning of NB is the indication of the cooperativity
switching and it is not necessary to break all the bonds connecting
among the structural units that compose the liquid near the glass
to the nearest neighbour components when the thermally acti-
transition. Following Ikeda and Aniya [23] NB gives the number
vated viscous flow occurs. For example, bond twisting may also
of the structural units involved in the thermally activated viscous
result in the viscous flow by enrolling the movement of second
flow, and it is expected to provide the degree of fluidity within
or more distant components of the liquid near its glass transition.
the glass-forming liquids. Relating Eq. (20) to the fragility index
Aniya [9] developed a new viscosity model (BSCNF) which
described by Eq. (15) it is obtained:
describes analytically strong and fragile systems. In this model,
the GFL is characterized by Gaussian distributions of the binding mðB ; C  Þ
NB ¼ lnð10Þ ð21Þ
energy, E, and the coordination number, Z, between structural C
units. The binding energy and coordination number are This result provides the connection between the fragility and
E = E0 + DE and Z = Z0 + DZ respectively, where E0 and Z0 are aver- the cooperativity near glass transition.
age values and DE and DZ are fluctuations. The BSCNF model is [9]: For pressure (and density) dependence of the viscosity the free
nh i o volume (or excluded-volume) (FV) model is often used with quite
g
Cx þ Cx2 lnðg g Þ þ 0:5 lnð1  BÞ 1B
C
1 good results over a wide range of viscosity. The physical basis for
ln g ¼ ln g1 þ
1

1  Bx2 this model is that viscosity in liquids should depend on the inter-
 0:5 lnð1  Bx Þ 2
ð14Þ action of a molecule with its closest neighbors and on the space
that is available for motion, which is the free volume. Free volume,
h i2 (VV1), is the difference between the total volume and the volume
ðD EÞðDZÞ E0 Z 0
where B ¼ RT g
;C¼ RT g
. occupied by the liquid. Doolittle proposed [10]:
The adjustable parameters in Eq. (14) are ln gg, ln g1, B and C.
V1
However, fixed values of gg and g1 are commonly used in calcula- ln g ¼ A þ B ð22Þ
tions and thus the number of coefficients could be only two. From V  V1
Eq. (14) the fragility index is [9]: being V1 the volume corresponding to infinite viscosity and B is a
n h   io parameter which is usually considered a function of pressure
g
1 B  C þ 2 ln g g þ 0:5 lnð1  BÞ describing the viscosity divergence as V?V1. Eq. (22) is normalized
1
m¼ ð15Þ using the density and viscosity at p = 0.1 MPa because the density
lnð10Þ 1B
and viscosity measured at this pressure are more accurate than at
According BSCNF theory, parameters B and C obey the relation high pressure. The normalized form of Eq. (22) is:
[22]:    
    g V1 V1
2cð1  BÞ gg jDEj =E0 ln ¼B  ð23Þ
C¼ pffiffiffi ln þ 0:5 lnð1  BÞ ; c ¼ ð16Þ g0 V  V1 V0  V1
2c þ Bð1 þ c2 Þ g1 jDZj =Z 0
where g0 and V0 are the viscosity and volume at atmospheric pres-
When c = 1, that is when the ratio between the fluctuations of sure. Eq. (23) was used by Cook et al. [11] to represent their viscos-
energy and coordination number becomes equal, the parameter C ity data in the ranges T = (273.15–398.15) K and p = (104 to 3) GPa.
(=C⁄) calculated from Eq. (16) can be substituted in Eq. (14) and Eq. (23) is written in terms of density as:
166 A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182

   
g q q0 4. Results and discussion
ln ¼B  ð24Þ
g0 q1  q q1  q0
The existing data in the literature for viscosity of glycerol were
According to Cook et al. [11], the free volume equations usually
collected and discussed in this section. In view of the considerable
give reasonable fit over ca. ten decades of high pressure viscosity
importance that the glass transition temperature plays as fixed
data for different classes of substances and also provides good esti-
point in modeling behaviour of GFL viscosity we have searched
mates of the glass transition pressure (using extrapolation to
the literature to obtain reliable values of this property as a function
gg = 1012 Pas).
of pressure and temperature. The collected data of viscosity and
The results of nonequilibrium molecular dynamics simulation
glass transition temperature is used latter for the modeling of this
show that the repulsive intermolecular potential controls the
property using equations presented in Section 3.
transport of momentum. If the power law behaviour U(r) (r/
r)3c is assumed valid for a repulsive potencial, then the viscosity
of molecular liquids scale as g = f(C) where C = T1Vc and c is
4.1. Fixed points and sources of viscosity data
the scaling parameter [13]. Viscosities measured under different
conditions of temperature and pressure and varying over several
Several properties are characteristic of GFLs, being of particular
orders of magnitude colapse onto a single master curve which
relevancy the melting temperature, Tm, the glass transition tem-
can be represented through the equation given by Bair and Casalini
perature, Tg, and the crossover temperature Tc located well above
[14]:
the glass transition temperature. Other quantities related with
    the step variation of viscosity at glass transition are the fragility
g C
index, m, and the viscosity itself at the glass transition, gg. The
ln ¼B ð25Þ
g1 C1  C high-temperature asymptotic value of viscosity, g1, is also consid-
ered in several studies. All these properties have been used in
where g1, B and C1 are adjustable parameters. Historically, the model equations applied to viscosity correlation.
scaling parameter was the value of 4 according to the exponent For the normal melting temperature of glycerol Wilhoit et al.
3c = 12 of Lennard-Jones repulsion. However, in the most recent [26] selected the value Tm = (291.8 ± 0.2) K reported by Volmer
studies it is assumed that is adjustable and a broad range of varia- and Marder [27] who made the measurements by carefully pro-
tion of this parameter has been found for the scaling of chemical tecting a dry sample from atmospheric moisture. This value has
different substances [13,24]. For molecular liquids the values of been given as the triple point temperature by NIST [28] and was
parameter c decrease with increasing size, being larger for liquids considered as such in the present work.
composed of small molecules interacting only through van der The knowledge of the glass transition temperature, Tg, is impor-
Waals forces and c is lower for ionic liquids and associated molec- tant from the theoretical and physical point of view [29,30] and for
ular compounds [24]. Henceforth, the scaling of viscosity with den- the development of equations used to describe the behaviour of
sity will be referred as (DS) model. the supercooled liquid or the glassy state [8,31–33]. Glycerol is
A model that is still little explored in modeling the influence of T one of the most classical low molecular glass forming liquids for
and p on viscosity was provided by Avramov [25]. In this model, an which the most extensive data on Tg is available at atmospheric
analytical form is used to eliminate the temperature dependence of pressure and pressures up to 12 GPa [34]. Several methods have
viscosity by formulating a pressure dependent dimensionless func- been proposed to evaluate Tg and the resulting values could differ
tion Ug(p) as [25]: in some degrees from each other. Selected values are given in
Table 1 and a brief discussion about them follows.
ln gðp; TÞ  ln g1  p b Calorimetry has been widely used for determination of Tg and
Ug ðpÞ ¼ ¼ 1þ ð26Þ
ln g0 ðTÞ  ln g1 P following Angell: Tg, per si, is usually defined as the temperature
of onset heat capacity increase, during heating, usually at
where ln g0(T) can be described by a chosen model at atmospheric 10 Kmin1 [29]. The definition of Tg is always arbitrary and there
pressure. The parameters G and b of organic substances including is no international convention on the subject. Gibson and Giauque
glycerol and geoscientifically relevant systems were calculated by [49] observed a sudden change of glass-liquid heat capacity start-
Avramov [25] from Eq. (26) using the AM model to describe the vis- ing at ca. 179.8 K and ending ca. 191.6 K. From their data, the value
cosity at atmospheric pressure. The dimensionless function Ug(p) Tg
190 K was assigned [39]. At atmospheric pressure, values close
leads to the colapse of viscosity distributed for different isotherms to this one were obtained: using differential scanning calorimetry
data into a single master curve which is function of pressure only. (DSC) Wang et al. [40] reported Tg = 188.9 K, Hempel et al. [38]
This behaviour reminds the scaling with density (DS) model dis- reported Tg = 189.15 K from DSC and Tg = 193.15 K from tempera-
cussed before, and for this reason the calculation of viscosity from ture modulated calorimetry (TMDSC). Sou et al. [46] used DSC
function Ug will be abbreviated as pressure scaling (PS) model. techniques and reported Tg in the range 178.3 K to 185.5 K using
To evaluate the predictive ability of the models aforementioned, heating scanning rates between 0.05 and 4.0 mKs1. In their study,
the relative deviations (RDs) between calculated and experimental Tg was taken as the temperature corresponding to the maximum of
values of viscosities were calculated using Eq. (27) and the average relaxation peaks of the thermograms for each scanning rate.
absolute percentage deviation (AAD) was calculated through Eq. Recently Swenson et al. [47] reported Tg = 190 K by DSC and this
(28), where N is the number of data points, gi is the experimental technique was used by Trofymluk et al. [41] who studied the vitri-
value of viscosity of this study or reported measurement in the lit- fication of glycerol in mesoporous silica in a wide range of pore
erature, and gi,cal is the calculated viscosity with models. sizes. They obtained the minimum value Tg = (189.0 ± 1) K for the
bulk liquid and a maximum Tg = (196 ± 1) K for the minimum pore
gi;cal  gi size (2.6 mm). Elsaesser et al. [43] developed a methodology where
RD ¼ 100  ð27Þ
gi the reversible glass-liquid transition was observed through a kink
in the relative volume change on isobaric heating of the sample.
N
100 X g  gi
i;cal
This kink was observed in several heating cycles where the result-
AADg ¼  ð28Þ ing isobaric thermal expansivity shows a sharp rise in the glass to
N i¼1
gi
liquid transition in a very narrow temperature range and this rise
A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182 167

Table 1
Selected values of glass transition temperature Tg of glycerol. The uncertainties, rTg, and range of pressure, Dp, are given.

Authors Year Tg/K rTg /K Dp/GPa Methoda Purity/mass fraction


Johari and Walley [35] 1972 217–258 (0.6–2.2) DD Non-available (na)
Sandberg et al. [36] 1977 188–248 (104–1.8) DSC na
Herbst et al. [37] 1993 178–270 (104–3) VIS 0.995
Cook et al. [11] 1994 192–280 (0.3–3.7) VIS na
Hempel et al. [38] 2000 189.15 104 DSC na
193.15 104 MTDSC na
Bartoš et al. [39] 2001 189 104 PALS na
Wang et al. [40] 2002 188.9 104 DSC 0.995
Paluch et al. [18] 2002 190–356 (104–7.4) RT na
Trofymluk et al. [41] 2005 189–196 ±1 104 DSC >0.996
Reiser and Kasper [42] 2006 191–212 (104–0.7) RT na
Elsaesser et al. [43] 2007 189–227 ±2 (0.05–1) VOL na
Pawlus et. al. [44] 2009 189,239.2 (104,1.8) RT
Pronin et al. [45] 2010 184–305 (104–4.5)
Sou et al. [46] 2011 178.3–185.5 104 DSC 0.995
Swenson et al. [47] 2013 190 104 DSC 0.999
Roussenova et al. [48] 2014 185b ±2 104 PALS na
190c ±2 104 PALS na
a
VIS: from viscosity equations using gg = 1012 Pas; DD: dielectric data; DSC: differential scanning calorimetry; MTDSC: temperature modulated DSC; VOL: volumetric
piezometer; RT: relaxation times; PALS: positron annihilation lifetime spectroscopy.
b
Measurements in bulk phase.
c
Confined in porous Vycor glass.

Fig. 1. Pressure dependence of the glass transition temperature, Tg. , Johari and
Walley [35]; , Sandberg et al. [36]; +, Herbst et al. [37]; h, Cook et al. [11]
calculated from viscosity; D, Paluch et al. [18]; r, Reiser and Kasper [42]; s,
Elsaesser et al. [43]; , Pawlus et al. [44]; , Pronin et al. [45], }, AM Eq. (8); , Fig. 2. The temperature dependence of viscosity for glycerol near Tg = (188 ± 5) K.
MYEGA Eq. (11); , BSCNF1 Eq. (17); - - - -, Eq. (29). The lines are tangents to the experimental curves near Tg/T = 1. Symbols: D,
Schröter and Donth [19]; , Huck et al. [53]; , Möbius et al. [55]. Solid line
corresponds to m = 45 and was determinated with data of Schröter and Donth and
of Huck et al. at Tg/T > 0.95. Dashed line corresponds to m = 41 obtained from data of
ended with an overshoot effect corresponding to a micro inhomo- Möbius et al. at Tg/T > 0.95.
geneous structure of glycerol [47,50]. Elsaesser et al. [43] deter-
mined Tg with reproducibility of ±2 K in the pressure range p =
(0.05–1.0) GPa. to the convention gg = 1012 Pas. Their (Tg, p) results and those
For viscosity calculations of glasses, it is common practice to reported by other authors mentioned in Table 1 were displayed
accept the value gg = 1012 Pas as the liquid viscosity at the glass in Fig. 1.
transition temperature. At atmospheric pressure, viscometric tech- The (Tg, p) data calculated by Cook et al. [11] are higher than
niques are feasible in the measuring range 0.1–1010 Pas and thus it those presented by other authors for increasing pressures from
is needed to extrapolate to 3 orders of magnitude to locate the 0.6 GPa and with the exception of data of Pawlus et al. [44] and
glass-liquid transition. High-pressure accurate viscosity measure- Pronin et al. [45], the values reported in the literature from the sev-
ments are scarce and they are usually available up to 106 Pas. Cook enties until now show good agreement. Particular mention could
et al. [11] measured the viscosity of glycerol in the ranges p = (104 be made to the values given by Paluch et al. [18] available for pres-
to 3) GPa corresponding to T = (273.15–398.15) K and they calcu- sures up to 7.4 GPa. The overwhelming majority of experimental
lated Tg as a function of pressure using model equations according (Tg, p) data indicate an increasing of Tg with pressure coupled with
168 A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182

Table 2
Sources of data for the viscosity of glycerol. The uncertainties in viscosity, ug, range of pressure, Dp, and of viscosity, Dg are given.

Authors Year Np DT/K Dp/GPa (Dg/mPa.s) (ug)a Methodb Purity/mass fraction


4 4 6
Tamman and Hesse [2] 1926 14 231.15–268.95 10 (1.5  10 –6.7  10 ) FBV
Parks and Gilkey [63] 1929 7 186.15–221.15 104 (108–1015) CC
Sheely [64] 1932 5 293.15–303.15 104 (583–1385) BV 0.9966
Vand [65] 1947 10 353–440 104 (0.8–32) CC
Segur and Oberstar [66] 1951 11 273.15–373.15 104 (1.207  104–14.8) (1–3%) OV 0.9997
Piccirelli and Litovitz [67] 1957 17 267.15–313.15 104 (1.1  107–46) HV
McDuffie and Kelly [52] 1964 21 263.15–283.15 104-0.4 (3.3  103–2.3  105) RBV 0.995
Harrison and Gosser [68] 1965 8 348.15 104-0.5 (40.7–413) (r < 2%)e RBV
Slie and Madigosky [69]d 1968 273–523 104 (0.9–9825) (±3%) OV 0.996
Berdyev et al. [70] 1974 15 243.15–453.15 104 (10–106) OV,HV
Kiyachenco and Litvinov [71] 1985 10 243.15–323.15 104 (190–7  105) CV >0.995
Huck et al. [53] 1988 13 192–262 104 (5104–1014) PV
Shankar and Kumar [72]c 1994 9 283.15–323.15 104 (114–2950) ( ± 4%) UV >0.99
Cook et al. [11] 1994 100 267.15–433.15 104–3 (8–2  109) RBD
Kinart and Kinart [73] 1996 4 288.15–303.15 104 (624–2336) HV
Schröter and Donth [19] 2000 71 192.35–320.85 104 (168–1013) PPR 0.999
Möbius et al. [55] 2010 14 185–297.9 104 (850-7  1013) CPPR
Ge et al. [74] 2010 9 298.15–338.15 104 (62–944) (±1%) RV >0.997
Klotz et al. [75] 2012 8 295.95 104–2 (580-3  106) (±5%) RB >0.995
Kijevcanin et al. [76] 2013 3 298.15,303.15 104 (600–912) ( ± 3  103) Stabinger 0.995
Bandarkar et al. [77] 2015 16 293.15–323.15 104 (141–1411) (r < 0.7%)e OV 0.994
Lynam et al. [78] 2016 14 298.15–413.15 104 (7–873) (±1%) CPR >0.990
a
The uncertainty in viscosity is given in mPas or percentage. Upper and lower limits of viscosity were rounded whenever possible.
b
FBV: Falling-ball viscometer; CC: Couette cell; BG: Bingham; OV: Ostwald viscometer; RBV: rolling-ball viscometer; HV: Hoppler viscometer; PV: penetro-viscometer;
UV: Ubbelohde viscometer; RBD: rolling-ball diamond anvil cell; PPR: parallel plate rheology; CPPR: Couette and plate-plate rheometer; CPR: cone and plate rheometer.
c
Kinematic viscosity in mm2s1.
d
Results given in equation form.
e
Standard deviation r from mean value.

a decrease of dTg/dp. From DSC measurements of the glass transi- or as a fitting parameter. The direct measurement of g1 is not pos-
tion temperature given in Table 1 at p = 0.1 MPa, the calculated sible but it could be obtained by extrapolation of measured data or
mean is <T0g> = (188 ± 5) K (where superscript ° refers to atmo- by some model including this quantity. The discussion in the liter-
spheric pressure). The value of < T0g> is in good agreement with ature about the calculation of g1 either from the theoretical point
those from positron annihilation lifetime spectroscopy [39,48], at of view [60,61] or using model equations refer to oxide glasses.
its upper limit is close the value (Tg = 195 K) which is obtained Nascimento and Aparício [62] reported log (g1/Pas) = 5.1 ± 2.0
using the rule Tg
2/3 Tm [51], and can be compared favourably from the application of VFT equation to 38 silicate and borate sys-
with selected values used in literature studies: 187 K [52], 186 K tems with fragilities ranging between 18 and 97. The high uncer-
[53,54], 190 K [55] and 191 K [33]. The (Tg, p) data of Sandberg tainty attached to log g1 and, on another hand, the fact that the
et al. [36], Herbst et al. [37], Paluch et al. [18] (for pressures up value was obtained for inorganic substances prevents its use as a
to 4 GPa), Reiser and Kasper [42] and Elsaesser et al. [43] were used fixed value in viscosity equations.
to fit the equation proposed by Drozd-Rzoska et al. [34]: The information on viscosity from the literature is given in
Table 2. The characteristics of data such as the temperature and
Dp 1=b ðDp=cÞ pressure range, measurement method and purity of samples are
T g ¼ T 0g ð1 þ Þ e ð29Þ
P provided whenever possible. At atmospheric pressure, the viscos-
where Dp = p-p0g, P, b, and c are parameters and T0g and p0g are the ity measurements cover a wide range of temperature, between Tg
reference temperature and pressure. In this work, we have used and 523 K. For pressures up to 3.0 GPa the covered temperature
T0g = <Tg°>, and the resulting parameters of Eq. (29) were range is T = (263–433) K. It is somewhat surprising that since the
P = 4.1554 GPa, b = 0.5657 and c = 5.0574 GPa, being the standard first viscosity measurements under pressure made in 1964 by
deviation of fit r = ± 2 K. Reiser et al. [56] gives T0g = (188.3 ± 0.5) K McDuffie and Kelly [52] and after more than fifty years only three
and their (Tg, p) equation follows closely Eq. (29) (maximum devia- pressure data sets have been reported, the last measurements
tion of 0.8% at 0.7 GPa). being made by Klotz et al. [75].
Since fragility gives the slope of viscosity at Tg, and it plays a At atmospheric pressure, the measurements of various authors
central role in understanding the behaviour of GFLs and its classi- could not be considered either due to the inconsistency of data,
fication, it would be convenient to calculate glycerol fragility values not reported or for other reasons. Slie and Madigosky [69]
directly, from experimental viscosity data. Following Senkov and made measurements in extensive temperature range but data is
Miracle [57], the value of m can be determined from the tangent given in equation form. Kiyachenko and Litvinov [71] measured
(maximum slope) of log g vs. Tg/T in the range 0.95 < Tg/T  1. For viscosity by light techniques and capillary viscosimetry but the
glycerol, the viscosity data available in this range of temperature values of viscosity were not reported. The measurements of Huck
are from Schröter and Donth [19], Huck et al. [53] and Möbius et al. [53] were given in graphical form.
et al. [55]. The two first sets of measurements are in good agree- Figs. S1 and S2 (Supplementary materials) show the viscosity
ment near glass transition as it can be seen in Fig. 2. From them, data distribution in temperature and pressure coordinates. The vis-
the slope of log g vs. Tg/T gives m = (45 ± 1) a value which is in close cosity measurements found in literature, made at atmospheric
agreement with m = 48 given by Rössler et al. [17]. Values found in pressure are presented in Fig. 3 as (ln g, T) plots.
literature are usually higher: 51 [40], 53 [54,58,59], and 54 [18]. From Fig. 3 it is obvious that data reported by Berdyev et al. [70]
The model equations given in Section 3 include the high- present anomalous behaviour and they will not be used in this
temperature viscosity limit (ln g1) which can be considered fixed work. For temperatures lower than 230 K the measurements made
A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182 169

Fig. 3. Experimental viscosity of glycerol as a function of temperature from literature at low and medium (a) and high (b) temperature ranges. e, Tamman and Hesse [2]; ,
Cook et al. [11], p = (104 to 0.97) GPa; , Cook et al. [11], p = (104 to 2.9) GPa; , Cook et al. [11] at p = 0.7 GPa; , Cook et al. [11] at p = 1.3 GPa; , Cook et al. [11], p = (104
to 2.2) GPa; , Cook et al. [11], p = (104 to 2.6) GPa; , Cook et al. [11], p = (104 to 3.8) GPa; , Vand [65]; s, Segur and Oberstar [66]; D, Schröter and Donth [19]; , Ge
et al. [74]; , Kijevcanin et al. [76]; 5, Bandarkar et al. [77]; , Slie and Madigosky [69]; h, Piccirelli and Litovitz [67]; , Lynam et al. [78]; ., Berdyev et al. [70]; ✰, McDuffie
and Kelly [52]; , Huck et al. [53]; ( , ), Möbius et al. [55]; ▲, Parks and Gilkey [63]; , Harrison and Gosser [68], p = (104 to 0.5) GPa; , Shankar and Kumar [72]; + , Klotz
et al. [75], p = (104 to 1.8) GPa; x, this work. Solid line corresponds to fit AM model, Eq. (3). Vertical line corresponds to glass transition temperature, Tg = 188 K.

by Möbius et al. [55] obtained with plate-plate rheology are always 4.2. Temperature dependence of viscosity at atmospheric pressure
lower than those reported by Schröter and Donth [19] measured
with the same technique. However, the measurements made by The viscosity measurements made in this work were combined
Möbius et al. [55] with Couette rheology are in reasonable agree- with those reported in literature namely those performed by
ment with those reported by Schröter and Donth [19] at overlap- Schröter and Donth [19], McDuffie and Kelly [52], Vand [65], Segur
ping temperatures (RDs ca. 10%). Data reported by Piccirelli and and Oberstar [66], Kinart and Kinart [73], Ge et al. [74], and
Litovitz [67] deviate ca. 20% from the values of Schröter and Donth Kijevcanin et al. [76], giving a working viscosity set of 154 data
[19]. McDuffie and Kelly [52] reported values which are in good points covering the range T = (192.4–440) K suitable for study of
agreement with data from other authors as Schröter and Donth the correlation equations at atmospheric pressure. The viscosity
[19] and Segur and Oberstar [66] (RDs less than 2%). The values data dispersed by other sets from other authors
reported by Cook et al. [11] at atmospheric pressure present vari- [2,53,55,63,67,69,72,77,78] were used for comparison purposes.
able deviations between zero and 6% compared with data from Two decades ago, Stickel et al. [79,80] indicated that the validity
other authors. Over the range of temperature (230–300) K the vis- of VFT description in supercooled glass-forming liquids is associ-
cosity data from Tammann and Hesse [2] are ca. 18% lower than ated with the linearity of plotting /T = [dlng/d(1/T)]1/2 vs. 1/T
the values of most authors. (Stickel T-plot) while Arrhenius behaviour corresponds to a con-
The viscosity measured with the rotational Brookfield DVII+ vis- stant line plot. In subsequent years, this representation became a
cosimeter at atmospheric pressure in the range T = (292.9–394.1) K key tool for estimating the so-called dynamic crossover tempera-
is presented in Table 3. At each temperature, the mean value of vis- ture, Tc, between two dynamic domains at which the a-
cosity and the corresponding standard deviation are presented. The relaxation dynamics starts to change. It was shown that at least
standard deviations are less than 1% of the mean values, most often two VFT equations are necessary for describing pre-vitrification
less than or equal to 0.6%, which presumes a good accuracy of data. slowing down in a broader range of temperatures. The temperature
In Fig. 3 the experimental data of this work are compared with val- Tc has been also recognized as a milestone in the way of approach-
ues from the literature. The viscosity values from Vand [65], Segur ing Tg since a lot of exceptional phenomena disclose there [81].
and Oberstar [66], Schröter and Donth [19], Ge et al. [74], Kijev- In Fig. 4 the Stickel plot (/T, T1) is presented. It was constructed
canin et al. [76] and Bandarkar et al. [77] follow closely our data using our data and selected values from other authors. For each
with deviations less than 5% and after T = 394 K, the data by Vand author data set, numerical derivative values were calculated by
show the trend of our own data. Systematic RDs of ca. 10% between using quadratic interpolating functions over narrow temperature
our values and those from Slie and Madigosky [69] are observed. ranges with neighbor ranges overlapping in a way that it was pos-
Values measured by Piccirelli and Litovitz [67] deviates from our sible to have a mean value of the Stickel derivative at each temper-
determinations by ca. 12% up to 310 K and deviations decrease ature. It can be observed the good agreement between the selected
for higher temperatures. Data measured by Lynam et al. [78] devi- sources of data in different ranges of temperature displaying differ-
ates appreciably after 383 K from our determinations. ent slopes. This characteristic of data will be used to find the
170 A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182

Table 3
Experimental viscosity of glycerol g at temperature T and p = 101.86 kPa.a

T/K N/rpm g/(mPas) <g>/(mPas) r/(mPas) T/K N/ rpm g/(mPas) <g>/(mPas)d r/(mPas)d
292.89 10 1455 353.67 100 32.25
292.89 15 1468 353.67 120 32.10
b
292.89 20 1487 1470 13 353.67 150 31.60
298.10 15 978.8 353.67 200 31.25 31.80b 0.40
298.10 20 965.8 364.12 30 21.30
298.10 30 967.1 970.6b 5.9 364.12 60 21.10
302.59 15 644.9 364.12 100 20.90
302.59 20 644.2 364.12 120 20.90 21.05c 0.17
302.59 30 648.4 645.8b 1.9 374.38 20 14.90
312.81 20 303.6 374.38 60 14.80
312.81 30 308.4 374.38 80 14.60
312.81 40 303.7 374.38 100 14.70
312.81 80 305.6 305.3b 1.9 374.38 120 14.70 14.74b 0.10
323.88 80 146.4 383.29 100 11.10
323.88 100 147.3 383.29 120 10.90
323.88 120 148.5 147.4b 0.9 383.29 150 11.00
333.15 80 86.40 383.29 200 11.00 11.00c 0.07
333.15 100 84.75 394.08 120 8.30
333.15 120 84.20 394.08 150 8.32
333.15 160 84.20 394.08 200 8.35 8.32b 0.02
333.15 200 84.00 84.71b 0.79
343.80 120 50.00
343.80 150 50.00
343.80 200 49.50 49.83c 0.24
a
Standard uncertainties: u(p) < 0.23 kPa, u(T) = 0.1 K, and the combined expanded viscosity uncertainty is ca. Uc(g) = 3% with 0.95 level of confidence (k = 2).
b
Measurements with SSA/31 (viscosity accuracy of 300/N).
c
Measurements with SSA/18 (viscosity accuracy of 30/N).
d
The mean value of viscosity <g> and the standard deviation from the mean r at each temperature were calculated over the values measured at different rotational speeds, N.

and, thus, the representation of (/T, T1) will be a straight line from
which the parameters DVFT and T0 can be calculated. From Fig. 4 in
can be seen that VFT equations can describe reasonably the temper-
ature dependence of /T in the supercooled range using different sets
of AVFT, DVFT and T0.
The results in Fig. 4 indicate three changes of regime at (291 ± 4,
345 ± 30 and 401 ± 14) K corresponding to the change of slope of
the linear representations. The first change of regime occurs at
Tc = (291 ± 4) K close to the crossover temperature reported by
Schröter and Donth [19] (T = 283.5 ± 6) K and it is consistent with
Tc = 285 K found from dielectric techniques [82]. González et al.
[8] found crossovers at (263, 343 and 393) K using data from Tam-
man and Hesse [2], Parks and Gilkey [63], Sheely [64], Piccirelli and
Litovitz [67] and Huck et al. [53]. As mentioned in Section 4.1, the
viscosity data from Tamman and Hesse and from Parks and Gilkey
present important disagreements with more plausible data from
other authors and their use by González et al. [8] could be respon-
sible for the differences with crossover temperatures of this work.
According to Eq. (30) the slope and intersection with the ordinate
give the parameters DVFT, T0 (and BVFT) and from them, the values
of m and ln (gg/g1) can be derived via Eqs. (6) and (7). An extra
Fig. 4. Temperature-derivative analysis (Stickel plot) of viscosity from viscosity fit of Eq. (2) to the viscosity data using the values of BVFT and T0
measurements of this work and of literature. , Vand [65]; s, Segur and Oberstar
[66]; D, Schröter and Donth [19]; h, Piccirelli and Litovitz [67];  , this work. Also
found from Stickel plot will provide the remaining parameter
shown are the VFT regressions for the data (black solid lines) determining the AVFT = ln g1. The parameters calculated in this way for the temper-
crossing temperatures, and the lines corresponding to AM, Eq. (8) (red), MYEGA, Eq. ature range T = (192–291) K (hereafter named as VFTS) are given in
(11) (blue) and constrained BSCNF, Eq. (14) (green). Dashed line corresponds to AM Table 4. Following another route, we have fitted Eqs. (2) and (5) to
equation fitted by González et al. [8]. The crossover temperatures found in this
viscosity data in the same range of temperature using the data
work are indicated by dashed vertical lines.
reported by McDuffie and Kelly [52], Kinart and Kinart [73] and
Schröter and Donth [19]. Table 4 shows that parameters AVFT, BVFT,
and T0 obtained with VFTS and VFT Eqs. (2) and (5) are comparable
changes of dynamic behaviour (crossover temperature) and to and that the set (ln g1, ln gg, m and Tg) obtained by fitting Eq. (5) to
compare the performance of viscosity equations in term of the data combine to give the same set (A, B and T0) as does Eq. (2). Eq.
derivative /T. The Stickel derivative of VFT Eq. (2) is: (5) predicts a glass transition temperature (Tg = 189 K) consistent
 0:5 with < Tg° > found in Section 4.1. The value log (g1/Pas)
7 Pas
0:5
d ln g 1 T0 1 found in all VFT methods is very low, without any physical mean-
ð 1
Þ ¼ 0:5
 ð30Þ
dT ðDVFT T 0 Þ DVFT T ing for glycerol. It can be explained by the lack of high-temperature
A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182 171

Table 4
Parameters of model equations and constrains applied to selected viscosity data of glycerol at p = 0.1 MPa.

Eqs./Constrains DT/Kc A = ln(g1/mPa.s) T0 /K B C ln (gg/mPa.s) me Tg/Ke AADg (%)f r ln g R2


VFT
VFTSa R1 -8.8360 121.5 2759.3 32.6572 51 ± 4 188 d
6.6 0.0896
Eq. (2)b R1 -9.3998 118.2 2911.2 32.3082 49 188 d
2.0 0.034 0.9999
Eq. (5)b R1 -9.3998 118.2 2911.2 31.7188 48 ± 4 189 2.0 0.034 0.9999
AM
Eq. (3) R2 - 1.6029 608.62 2.987 31.7987 43 188 d 3.4 0.049 0.9999
Eq. (8) R2 - 1.6029 608.62 2.987 34.1773 46 ± 3 184 ± 6 3.4 0.049 0.9999
MYEGA
d
Eq. (4) R2 - 4.1459 931.45 375.56 32.3785 47 188 5.9 0.073 0.9998
Eq. (11) R2 - 4.1459 931.45 375.56 33.7846 50 186 5.9 0.073 0.9998
BSCNF
Eq. (14) R2 1.7836 0.2225 -20.727 31.2589 46 ± 1 186 ± 1 3.3 0.045 0.9999
Eq. (14)g R2 -5.0017 0.4840 6.2414 34.5388 61 ± 2 188 12 0.182 0.9985
a
Parameters B and T0 calculated from Stickel plot.
b
VFT fitted to selected viscosity data.
c
Temperature ranges of the fittings: R1 and R2 correspond to T = (192–291) K and T = (192–440) K.
d
Not used in the fitting, considered as fixed parameter for the calculation of ln gg and m.
e
Values of m and Tg were rounded to the nearest integer and their calculated uncertainties were not given unless they are higher than ±1.
P
Average absolute deviation, AADg ¼ ð100=NÞ N i¼1 jgcal  gji =g where gcal and g are calculated and experimental viscosity for data point i.
f

g
Constrained fit by simultaneously fixing Tg = <T0g> and log(gg/Pas) = 12.

Fig. 5. Relative deviations as a function of temperature between the calculated values based on Eq. (2) (gVFT) and the experimental values (g) in the range T = (192 to 291) K.
(a) VFTS calculated via Stickel plot and (b) Eq. (2) fitted to selected experimental data of Segur and Oberstar [66], Schröter and Donth [19] and Kinart and Kinart [73]. e,
Tamman and Hesse [2]; s, Segur and Oberstar [66]; D, Schröter and Donth [19]; h, Piccirelli and Litovitz [67]; ✰, McDuffie and Kelly [52]; , Huck et al. [53]; ▲, Parks and
Gilkey [63]; , Möbius et al. [55]; , Kinart and Kinart [73]. Also shown are comparisions of VFTS and VFT, Eq. (2) with Chen and Pearlstein [84] (solid line), and AM equations
(dashed line) fitted by González et al. [8].

viscosities for the fittings. The values of fragility calculated with all (AAD = 6.6%) and that a very good representation of data is
VFT are in good agreement with m = 45 ± 1, obtained in Section 4.1 obtained with Eqs. (2) or Eq. (5) since the AAD = 2% is close to
using experimental data. the smallest uncertainty found in experimental determinations.
About the ability of VFT equations to representation observable The RDs of the calculated viscosities from the literature data are
viscosity data, Table 4 shows that VFTS gives reasonable accuracy plotted as a function of temperature in Fig. 5(a) and (b) for VFTS
172 A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182

and Eq. (2), respectively. The results from Eq. (5) are almost the 12
same as those obtained from Eq. (2). -2.0

Fig. 5(a) shows that with the exception of temperatures -2.5


10
approaching glass transition, the RDs of VFTS relative to data from
-3.0
Schröter and Donth [19] and Segur and Oberstar [66] are in the

log(η /mPa.s)
range ±5% which is not far from the usual uncertainty of viscosity 8 -3.5
measurements. Near glass transition viscosities of Schröter and
-4.0
Donth [19] are reproduced with RDs between 15% and 25%. The
values of Piccirelli and Litovitz [67] show RDs between 13% and 6 -4.5

log[η /(Pa.s)]
20%, giving corresponding to AAD of 17%. The data sets not used -5.0
in VFTS fitting behave as follows. The values of Huck et al. [53] 4
-5.5
show variable RDs, with a maximum value of 47% at 192 K, being 0.0 0.1 0.2 0.3 0.4 0.5
AAD = 16%. The values from Möbius et al. [55] measured with Cou- Tg/T
ette methodology deviates by more than 10% as temperature 2
decreases. The values due to Tamman and Hesse [2] deviate largely
from calculated values (AAD of 51%). Its contemporary values of
0
viscosity reported by Parks and Gilkey [63], which even now are
part of the few viscosity data available at low temperature, have
high negative RDs (-38%) near the glass transition, becoming high -2
positive (54% at 221 K). For Eq. (2), Fig 5(b) shows that the calcu-
lated viscosities are in very good agreement with values reported
by Schröter and Donth [19] (AAD of 1.7%), RDs distributing in the -4
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
range ±3% with 5% near the glass transition. The rejected data
Tg/T
in the fitting show RDs with a similar behaviour to that in Fig. 5
(a) for VFTS. However, data by Huck et al. [53] show smaller devi- Fig. 6. Tg-scalled Arrhenius representation (Angell plot) for glycerol at atmospheric
ations than for the calculations with VFTS at temperatures near pressure. The inset shows the high temperature behaviour of models giving
glass transition. different values of ln g1. e, Tamman and Hesse [2]; , Vand [65]; s, Segur and
The main conclusions of this part can be summarized as follows. Oberstar [66]; D, Schröter and Donth [19]; , Ge et al. [74]; , Kijevcanin et al. [76];
, Slie and Madigosky [69]; , Kinart and Kinart [73]; , this work. The results from
The Stickel plot gives an indication of a crossover temperature
AM (solid line), MYEGA (dashed line) and constrained BSCNF (dash-point line) are
located near the melting temperature. It is possible to find a VFT shown.
equation from Stickel analysis of data which describes viscosity
data with reasonable accuracy (AAD = 6.6%) from glass transition
to crossover temperature. In this range, very good accuracy in vis-
cosity calculation (AAD = 2%) and reliable derived parameters (Tg values of Tg and m are in good agreement with the values resulting
and m) are obtained by fitting VFT Eq. (5) to viscosity data. In the from other models and from literature.
whole temperature domain covered by Stickel analysis no Arrhe- The obtained correlation coefficient, R2, indicate a very good
nius behaviour was detected. correlation of viscosity for all the models. The same good perfor-
To overcome the limitations of applying VFT equation over lim- mance is revealed by inspection of the Angell plot presented in
ited ranges of temperature, AM, MYEGA and BSCNF equations were Fig. 6. In this figure, the calculated viscosities from AM, MYEGA
studied over the temperature range between 192.4 K and 440 K and constrained BSCNF were compared with the selected experi-
(154 data points), corresponding to viscosity data sets selected mental data.
from literature as mentioned in Section 4.1. The fitted parameters Different log g1 values are predicted from the model equations.
for standard AM and MYEGA from Eqs. (3) and (4), their equivalent The majority of values of the viscosity at high temperature
Eqs. (8) and (11), and BSCNF Eq. (14) are presented in Table 4. As limit are predicted from the model equations as close to log
explained in Section 3, all the BSCNF parameters have a clear phys- (gg/Pas) = 11. Data of Slie and Madigovsky [69] at higher temper-
ical meaning related with cooperativity phenomena at low temper- atures follow the results from AM but deviates considerably from
ature and only imposing constraints one obtains suitable values for the MYEGA and BSCNF. This is more clearly observed in Fig. 7
the parameters. From Table 4 it can be seen that when BSCNF is where the RDs were represented as a function of temperature for
used without constraints, it gives a good fit (AAD = 3.3% and the model equations.
R2 = 0.9999). However, the value of C must be positive [22] and fur- From Fig. 7, it can be seen that MYEGA and constrained BSCNF
thermore, the obtained high positive ln g1 value is unacceptable. represent data of Slie and Madigovsky [69] with RDs of the order of
Fixed values of gg and g1 are commonly used in the fitting BSCNF 20% at medium temperature ranges but they exceed this value at
model [22]. Fixing g1 does not make sense as discussed before in temperatures higher than 450 K. On the contrary, AM model repre-
Section 4.1. Consequently, for this model, we have constrained sent these data with RDs less than 10% after 350 K reaching 5% at
Eq. (14) by simultaneously fixing Tg at <T0g>and log(gg/Pas) = 12. the highest temperatures. This observation could play favorably
With this procedure, one obtains suitable values for the B and C as an indication that log g1 is not far from the value log (g1 /
parameters even though some loss of accuracy of viscosity calcula- Pas)=- 3.7 (see Table 4).
tion can result. Observing the RDs in Fig. 7(a) one can conclude that AM equa-
With regard to the prediction ability of the properties charac- tions provide a good representation of all selected data for the fit-
teristic of glass-forming liquids, Table 4 shows values of glass tran- ting in the entire the range of temperature with AAD of 3.4% a value
sition temperature predicted with AM and MYEGA which are in close to the reliable viscosity measurements. Some scattering is
good agreement with <T0g>. Fragility index is consistent with the observed in deviations near the glass transition. Schröter and
value found in this work and with other values from literature, Donth viscosities [19] making up about 45% of selected values
characterizing glycerol as a moderate fragile liquid. Despite uncon- deviate with AAD = 4%. Our viscosities (28% of data points) are rep-
strained BSCNF model gives unrealistic C parameter, the predicted resented with AAD = 2%. This value was also found as AAD for the
A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182 173

Fig. 7. Relative deviations as function of temperature between the calculated values with model equations (gEq(AM,MYEGA,BSCNF)) and the experimental values (g). (a) AM, Eq
(3); b) MYEGA, Eq. (4); (c) Constrained BSCNF, Eq. (14). Symbols: e, Tamman and Hesse [2]; , Vand [65]; s, Segur and Oberstar [66]; D, Schröter and Donth [19]; , Ge et al.
[74]; , Kijevcanin et al. [76]; 5, Bandarkar et al. [77]; , Slie and Madigosky [69]; h, Piccirelli and Litovitz [67]; , Lynam et al. [78]; ✰, McDuffie and Kelly [52]; , Huck
et al. [53]; ▲, Parks and Gilkey [63]; , Möbius et al. [55]; , Shankar and Kumar [72]; , Kinart and Kinart [73]; , this work. Also shown are comparisions of AM, MYEGA
and BSCNF equations with Chen and Pearlstein [84] (solid line), and AM equations (dashed line) fitted by González et al. [8].

high temperature data reported by Vand [65]. Data reported by 30% at 413 K. The determinations made recently by Bandarkar
Segur and Oberstar [66] extending from 273.15 K to 373.15 K have et al. [77] in the range 293 K to 323 K give AAD = 2%. The kinematic
AAD = 2.7%. The excluded data behave as follows. The plate-plate viscosities measured by Shankar and Kumar [72] between 283 K
rheology data of Möbius et al. [55] deviates by more than 200%. and 323 K were changed into dynamic values using the density cal-
Viscosities of Tamman and Hesse [2] show high RDs reaching val- culated from GMA equation of state [83] and they deviate with
ues of ca. 100% at 230 K. The measurements made by Piccirelli and AAD = 4% from AM.
Litovitz [67] extending from 230 K to 343 K exhibit AAD of 17%. The viscosity RDs corresponding to MYEGA and constrained
The recent determinations at high temperature made by Lynam BSCNF are presented in Fig. 7(b) and (c) respectively. Compared
et al. [78] show increasing deviations with temperature, reaching with AM equations, somewhat higher RDs are observed: the
174 A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182

Table 5
Parameters for the fit with BSCNF1 equation to viscosity at p = 0.1 MPa. Derived properties are the number of structural units broken (NB) fragility index (m) CRR diameter (nB) and
(T0/Tg) ratio.

Eqs./Constrains A = ln(g1/mPa.s) B* C* ln(gg/mPas) mc Tg/Kc NB nB/nm T0/Tg AADg (%) e


rln g R2
a
R1
NCb 7.0346 0.4890 12.638 35.329 61 185 ± 3 11 1.34 0.699 11.9 0.148 0.9990
Cb 6.5116 0.5123 11.567 34.5388c 63 188d 12 1.40 0.715 12.9 0.185 0.9985
R2a
NCb 9.3636 0.4162 15.763 35.3249 55 183 8 1.21 0.645 5.7 0.034 0.9999
Cb 6.2417 0.5159 11.387 34.5388c 63 188d 13 1.41 0.718 18.4 0.263 0.9968
a
R1 and R2 refer to ranges of temperature T = (192–440) K and T = (192–291) K, respectively.
b
NC refers to unconstrained fit and C to constrained fit by simultaneously fixing Tg = <T0g> and log(gg/Pas) = 12.
c
Values of m and Tg were rounded to the nearest integer and their calculated uncertainties were not given unless they are higher than ±1.
d
Not used in the fitting, considered as fixed parameter for the calculation of ln gg and m.
P
Average absolute deviation, AADg ¼ ð100=NÞ N i¼1 jgcal  gji =g where gcal and g are calculated and experimental viscosity for point i.
e

average absolute deviations are 6% and 12%, for MYEGA and BSCNF, ting (R2 = 0.9999 and AADg = 6%) is obtained and predicted values
respectively. The main points discussed before for AM relative to of m and Tg are reliable. The limiting viscosity log (g1/Pas)
7,
different authors data sets can be applied also to MYEGA and is the same as for VFT equation (cf. Table 4). The ratio T0/Tg calcu-
BSCNF with particular differences observable at low temperature lated from Eq. (18) is given in Table 5. Taking the value correspond-
near the glass transition and at the high temperature. ing to unconstrained fit made in R2 one obtains T0 = 118 K which is
By fixing log (gg/Pas) = 9.5, Tg = 195.4 K and considering (ln g1) the same value found before from VFT Eqs. (2) and (5).
and (CAM) as adjustable parameters, González et al. [8] fitted AM Parameter NB calculated from Eq. (20) of BSCNF1 gives the num-
Eq. (8) using viscosities of Tamman and Hesse [2], Parks and Gilkey ber of the structural units involved in the thermally activated vis-
[63], Piccirrelli and Litovitz [67] and Schröter and Donth [19]. They cous flow and therefore it is related to the cooperativity
fitted also the same viscosity data sets with the four-parameter rearranging region (CRR). The concept of CRR was introduced by
equation proposed by Chen and Pearlstein [84]. These representa- Adam and Gibbs [4] and was defined as the smallest subsystem
tions of viscosity are compared with model equations of this work which, upon a sufficient thermal fluctuation, can undergo a confor-
in Fig. 7. Large RDs, especially at mid and higher temperature mational rearrangement independently of its environment. Calori-
ranges are observed which possibly can be explained due to the metric experiments give the size of the cooperativity [38,85] and it
wrong selection of data: as we have shown in Fig. 3 and discussed, was reported that CRR increases with the decrease of temperature.
the viscosities reported by Schröter and Donth [19] are inconsis- According to BSCNF model the larger size of CRR for more fragile
tent with data from the other data sets especially with the deter- systems results from the preferential break of weaker parts of
minations of Tamman and Hesse. On the other hand, the value the bonds between structural units and thus the degree of cooper-
Tg = 195.4 K selected for the fittings was probably too high. ativity is evaluated from NB by using Eq. (20). By assuming a spher-
In Fig. 4 the Stickel plots for AM, MYEGA and BSCNF models are ical shape for the region of cooperativity, the cooperative volume is
shown. The Stickel derivatives from AM equation are in close [86], VB = NBV1 = (4p/3) (nB/2)3, where nB is the diameter of the
agreement with the numerical values found from experimental spherical region and V1 is the volume of one structural unit, which
data, over wide temperature range (up to 400 K). MYEGA shows is evaluated by V1 = M/(NAq) where M is the molar mass, NA the
good agreement up to 330 K and BSCNF performs better in the T Avogadro’s number and q is the density. In Table 5, nB was calcu-
range where VFTS was developed, i.e., the supercooled region. lated from NB and density q (Tg = 188 K) = 1324 kgm3 [50]. By
The Stickel derivative calculated from AM equation fitted by Gon- using NB = 11 corresponding to unconstrained fit made in R1,
zález et al. [8] was calculated and represented in Fig. 4. It is nB = 1.34 nm is obtained, which is in good agreement with the
observed that calculated values don’t follow the expected trend cooperativity length reported by Fragiadakis et al. [87] using 4-D
and they display a strange behaviour for temperatures higher than NMR (n4-D NMR = 1.30 nm). Similarly, taking NB = 12 and NB = 13
370 K. corresponding to the constrained fittings made in R1 and R2
The BSCNF1 model given by Eq. (17) was studied in some detail respectively, nB = 1.40 nm and 1.41 nm were calculated in excellent
because some important parameters for GFLs can be derived from agreement with the correlation (or cooperativity) length evaluated
it. In fitting BSCNF1 to viscosity, two situations were considered: by Hong et al. [33] from boson peak analysis technique (nBP = 1.42 -
the use of data available over temperatures ranges (R1) corre- nm). These results indicate that BSCNF equations can be a reliable
sponding to T = (192.4–440) K, and over the reduced range (R2) tool to explore glass forming liquids properties and structure with
for T = (192–291) K, where VFT equation was applied. The correla- respect to their relations with viscosity.
tion using these ranges was decided because the ‘‘perfect” equiva-
lence between BSCNF1 and VFT equations claimed by Ikeda and 4.3. Influence of pressure on glass transition parameters
Aniya [22] and, therefore, it will be interesting to study the differ-
ences in performance of the model in different temperature ranges. In addition to the atmospheric pressure, some viscosity data at
The constrain Tg = <T0g> and log(gg/Pas) = 12, applied before to other isobaric conditions are available. At pressures of 0.7 GPa and
BSCNF was used. The results were compiled in Table 5. The fittings 1.3 GPa and corresponding T = (267–398) K and T = (284–433) K
made in R1 constrains show acceptable quality (R2 = 0.999 and respectively, extensive viscosity data were given by Cook et al.
AAD = 12% and 13%). Now the unconstrained fit gives positive C, [11]. Restricted data were obtained by these authors from interpo-
contrary to what was observed for BSCNF. Without constrain lation of isothermal experiments at p = 2.2 GPa and p = 3.0 GPa.
imposed, the fitted glass transition temperature (Tg = 185 ± 3 K) is This information was used to fit AM, MYEGA and BSCNF1 with
in good agreement with <T0g> and the predicted viscosity at glass the main purpose of collecting information on Tg, m, and NB as a
transition is also in agreement with the standard value from liter- function of pressure. Unconstrained fittings were applied to
ature. The obtained fragility index is a little higher than the values BSCNF1 model. The results for fitting equations to isobaric (g, T)
generally accepted. Working in R2 without constraint, a good fit- data are given as Supplementary Materials in Table S1. The results
A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182 175

glycerol and the same conclusion was reported by Root and Berne
[90] from molecular dynamics. Root and Berne [90] explained the
increase of m with pressure from the broader distribution of hydro-
gen bond angles and thus bond strengths. This situation leads up to
the increase in the range of hydrogen bond energies resulting in
the increasing of fragility.
From Table S1 it can be concluded that according to BSCNF1
equation after a steep ascent of NB starting at atmospheric pressure
where NB = (8–11) there is a stabilization of NB in the form of a pla-
teau and thus the number of structural units that are breaking on
weaker parts of the bonds becomes almost constant (NB = 34). This
behaviour is closely related to what happens with parameters m
and C⁄ through Eq. (21).

4.4. Viscosity dependence on temperature, pressure and density

In this section viscosity data of glycerol available in the litera-


ture covering the ranges T = (263.15–398.15) K and p = (104 to
3.0) GPa [11,52,68] were used to study the free volume and scaling
Fig. 8. Pressure dependence of fragility índex, m, calculated herein and from the equations.
literature. D, Cook et al. [11] from viscosity isobaric calculations; ▲, Cook et al. [11]
from viscosity isochoric calculations; d, Paluch et al. [18]; h, Reiser and Kasper 4.4.1. Free volume model
[42]; , Pronin et al. [45], }, AM Eq. (8); , MYEGA Eq. (11); , BSCNF1 Eq. (17).
The use of Eq. (24) requires the conversion of pressure to den-
sity. This was accomplished using the GMA equation of state given
of calculations of glass transition temperature as a function of pres- by us in a previous work [83] where we have shown that GMA
sure are displayed in Fig. 1 for all the models. The main conclusion enables the calculation of density at pressures of 50 GPa with devi-
is that the calculated Tg are in reasonable agreement with litera- ations of only ±2%. Eq. (24) was fitted to isothermal data sets using
ture values and the agreement is better for BSCNF1 equation. The least squares with resulting parameters giving the temperature
fragility index is represented in Fig. 8. dependence. Values of g0 were calculated by standard AM model
From Fig. 8 it can be seen that at increasing pressures the iso- with Eq. (3) because of the very good quality of calculations.
choric and in particular the isobaric values derived from viscosity Parameters B and q1, the density at p = 0.1 MPa from GMA, q0,
data by Cook et al. [11] are notoriously higher than the values the average absolute percentage deviation in viscosity AADg, the
reported by Paluch et al. [18] and Reiser and Kasper [42]. The val- standard deviation rln g, the correlation coefficient and the num-
ues of m from AM and MYEGA are consistent with values of Paluch ber of data points for each isotherm are presented in Table 6. In
et al. [18] showing a kind of plateau after an initial increase. The Fig. 9 the dependencies on temperature of parameters B and q1
values of m calculated from BSCNF1 are higher than those derived are represented.
from AM and MYEGA, showing, however identical behaviour as Sigmoidal shapes were obtained with a sharp jump beginning at
pressure increases. Cook et al. [11] pointed out that the strong 286 K and ending at 306 K. It can be verified that the crossing point
effect of pressure in fragility índex in glycerol was understandable temperature obtained from the Stickel plot obtained at atmo-
if hydrogen bonding decreases with the increase of pressure: the spheric pressure, 291 K, is close to the midpoint of the observed
loss of associations would progress towards a hard-sphere-like jump. Similar results were obtained by Cook et al. [11] from the
fragile liquid. The decrease in hydrogen bonding in glycerol with application of free volume model to their high-pressure viscosity
increasing pressure seems reasonable since it is known [88] that data. Parameters B and q1 were fitted to sigmoidal functions:
in water this phenomenon occurs. Compression of water leads to b3
significant distortion and/or disruption of the hydrogen bond net- B ¼ b1 þ b2 T þ ð31Þ
1 þ eðTT B Þ=b4
work with the important consequence that the dynamic behaviour
of water under high compression resembles more that of a ‘‘nor- where b1 = 10.4808, b2 = 0.0310 K1, b3 = 8.1929, TB = 295.8 K,
mal” molecular liquid of comparable molecular size [88]. However, b4 = 3.0733, standard deviation rB = ±0.283 and correlation coeffi-
NMR studies [89] indicate the increase of hydrogen bonding in cient R2 = 0.9958, and:

Table 6
Parameters of free volume equation, Eq. (24), fitted to isothermal (q, g) data.

Authors q0/gcm3 B q1/gcm3 AADg (%) r ln g R2


T/K (Dp/GPa)
McDuffy and Kelly [52]
263.15 (0–0.21) 1.2753 2.1167 1.5934 3.8 0.022 0.9992
273.15 (0–0.27) 1.2712 2.3957 1.6392 5.6 0.031 0.9986
283.15 (0–0.38) 1.2666 1.8099 1.6136 3.6 0.057 0.9974
Cook et al. [11]
273.15 (0–0.92) 1.2712 1.8206 1.6198 32.3 0.154 0.9962
295.65 (0–1.1) 1.2602 5.3207 1.9355 5.9 0.073 0.9998
323.15 (0–2.2) 1.2435 8.7253 2.2345 8.6 0.193 0.9956
348.15 (0–2.6) 1.2268 7.7715 2.2368 16.3 0.191 0.9953
398.15 (0–3.0) 1.1952 6.3184 2.2431 10.1 0.120 0.9973
Harrison and Gosser [68]
348.15 (0–0.49) 1.2268 1.6228 1.6479 5.5 0.027 0.9990
176 A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182

Fig. 9. Temperature dependence of parameters B (a) and q1 (b) of free volume model, Eq. (24). ▲, McDuffie and Kelly [52]; ., Harrison and Gosser [68]; , Cook et al. [11].
Full lines correspond to Eqs. (31) and (32) for (a) and (b), respectively.

ficient R2 = 0.9984. A 3D plot obtained from FV theory showing the


logarithm of viscosity as a function of temperature and density was
displayed in Fig. 10. Over the surface, the lines corresponding to the
standard pressure and temperature, i. e. at p = 0.1 MPa and at
T = 298.15 K and the results corresponding to experimental data
for isothermal paths are shown.
In Fig. 11(a) the surface projection in (ln g, q) plane is depicted.
Isobaric data plotted at p = 0.1 MPa were calculated with AM, and
at the pressures 0.7 GPa and 1.3 GPa for whose values Cook et al.
[11] reported ‘‘pseudoisobaric” experimental data, the calculations
were made with Eq. (24). In Fig.11(b) the RDs between calculated
and experimental values of viscosity are presented. A reasonable to
a good agreement is obtained between calculated and experimen-
tal data with global average absolute deviation AAD = 12%. The
highest contribution for AAD comes from Cook et al. viscosity data
(14%), being the contributions from McDuffie and Kelly of 4% and
from Harrison and Gosser 6%. It should be mentioned that Cook
et al. reported an overall uncertainty of approximately 10% for
their viscosity measurements [11]. From Fig. 11 the experimental
scattering of data in viscosity isotherms given by Cook et al. can
be concluded.
It is well recognized that density plays a key role in the under-
Fig. 10. Surface plot of the function ln (g/mPas) versus temperature, (T/K), and standing of viscosity behaviour. Further information on this subject
density, (q/ gcm3), calculated with FV, Eq. (24). A color gradient was used on the can be obtained by inspection of the isokom (i. e. constant viscos-
surface in order to enhance the sudden increase in viscosity in the low temperature
and high density regions. s, Cook et al. [11], , McDuffie and Kelly [52], , Harrison
ity) lines at the (q, T) plane represented in Fig. 12 for equally
and Gosser [68]; , p = 0.1 MPa; white line corresponds to the calculations for spaced values of ln g. In this figure, the experimental data of the
T = 298.15 K. authors have been plotted.
As shown before in Fig. 11, a pronounced rise in viscosity is
q1 ¼ q1 þ q2 eFðTÞ ð32Þ observable for values of ln g are higher than 15, at temperatures
lower than 300 K. The curvature of the isokom lines changes
with FðTÞ ¼ eðTT q Þ=q3 , being q1 = 1.6164, q2 = 0.6225, q3 = 5.2552, markedly in this region especially near 290 K, close the crossing
Tq = 293.5 K, standard deviation rq1 = ± 0.017 and correlation coef- temperature found at atmospheric pressure. The (Tg, qg)
A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182 177

Fig. 12. Isokom lines (i.e. at constant viscosity) at the (q,T) plane calculated with
free-volume FV model, Eq. (24). The symbols correspond to the data of authors used
in FV: s, Cook et al. [11], , McDuffie and Kelly [52], , Harrison and Gosser [68].
Glass transition: - - -, results from FV; , results from Eq. (29); , data from
Cook et al. [11].

Fig. 11. The free-volume (FV) and density scaling (DS) models in terms of density (q)
and temperature compared with experimental viscosity. (a) Isothermal (q, ln g)
representations; (b) relative deviations between the FV calculated values (gcalc) and
the experimental data (g). 4, Cook et al. [11], T = 273.15, and p = (104 to 0.97) GPa;
5, Cook et al. [11], T = 295.65 K, p = (104 to 2.9) GPa; s, Cook et al. [11], T = 323.15 K,
p = (104 to 2.2) GPa; h, Cook et al. [11], T = 348.15 K, p = (104 to 2.6) GPa; e, Cook
et al. [11], T = 398.15 K, p = (104 to 3.8) GPa; , Cook et al. [11] at p = 0.7 GPa, T =
(267.15 to 398.15) K; +, Cook et al. [11] at p = 1.3 GPa, T = (284.05 to 433.15) K; ,
McDuffie and Kelly [52], T = 263.15, and p = (104 to 0.21) GPa; , McDuffie and Kelly
[52], T = 273.35 K, p = (104 to 0.27) GPa; , McDuffie and Kelly [52], T = 283.35 K,
p = (104 to 0.38) GPa; , Harrison and Gosser [68], T = 348.15 K, p = (104 to 0.49)
GPa; (━, , ), FV; , FV at Tc = 191 K; - - -, DS; , AM (p = 0.1 MPa).

coordinates corresponding to glass transition are displayed and


compared with the results obtained from Eq. (29) and those
reported by Cook et al. [11]. The glass transition temperature
increase always with the density of the liquid, and FV theory pre-
dicts a very rapid increase of Tg with density up to ca. 1.52 gcm3
and becoming very smooth afterward.
Taking the ‘‘pseudoisobaric” viscosity paths given by Cook et al.
[11] represented in Fig. 11, one can see that increasing the pres-
Fig. 13. Density scaling (DS) of the viscosities for glycerol. The viscosities
sure, glycerol shows increasingly isochoric behaviour along the iso- encompass temperatures from 191 K up to 440 K and pressures up to 3.8 GPa.
baric lines and that at isochoric conditions, the viscosity is The scaling exponent, c = 1.8, was determined from the power-law fit of the
markedly temperature dependent as density increases. This effect temperature and density at constant viscosity (4, g = 50 mPa.s; 5, g = 103 mPa.s; s,
can differ in magnitude for different fluids [11] and change from g = 7103 mPa.s; h, g = 3.4104 mPa.s; e, g = 3105 mPa.s) shown in the inset, from
data of Cook et al. [11]. Symbols for DS: 4, 5, s, h, e, Cook et al. [11]; , , ,
one density range to another and thus it will be difficult to scale McDuffie and Kelly [52]; , Harrison and Gosser [68]; , This work; , Vand [65];
the effects by simple superimposing the curves through transla- , Schröter and Donth [19]; , Eq. (25).
tions along density axis. However, for glycerol, the influence of
temperature on viscosity at different isochoric paths is less than of viscosity, the scaling relation is g = f(C) with C = T1qc where
for other substances as for example dibutylphthalate [11]. This fact c is a liquid constant and f a unique function for which the form
justifies the next step which is the scaling attempt of viscosity in is usually unspecified, although Bair and Casalini [14] derived Eq.
terms of liquid density. (25) which describes experimental data very well. Dreyfus et al.
[12] shown that viscosity data of glycerol reported by Cook et al.
4.4.2. Scaling of viscosity with density [11] scale with c = 1.8. For each isokom, the scaling parameter C
Density scaling of dynamic variables and transport properties of will be a constant and thus a double linear logarithmic plot of T
GFL substances have been studied recently [12,14,15]. In the case versus q has slope c. This power-law plot is shown in the inset
178 A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182

of Fig. 13 where some iso-viscosity (T,p) data sets corresponding to the selected values for scaling exponent. The results of scaling vis-
viscosities in the range 50 mPas–3 105 mPas where collected from cosity are presented in Fig. 13 where ln g is represented as a func-
the data reported by Cook et al. [11] and they were used in a dou- tion of (C = T1q1.8). The isothermal experimental data of Cook et.
ble logarithmic plot. For each (T,p) conditions, the density was cal- al. [11], McDuffy and Kelly [52] and Harrison and Gosser [68] were
culated from GMA Eos reported by us [83]. It can be seen that good represented. Some selected experimental data at p = 0.1 MPa were
linear relationships hold for the tested viscosities. The pairs (g, c) also plotted for T = (192–440) K.
corresponding to the fittings were: (50 mPas, 1.78), (103 mPas, One can see that ln g shows a tendency to collapse on a single
1.85 ± 0.16), (7  103 mPas, 1.86 ± 0.13), (3.4  104 mPas, master curve although some scattering of data is observed at
1.75 ± 0.08) and (3  10 mPas, 1.85 ± 0.15). Therefore, c = 1.8
5
higher values of C. The parameters resulting of fitting Eq. (25) to
was considered for scaling calculations. This value is slightly higher the isothermal viscosity data mentioned above were log (g1/
than the value c = 1.4 found from scaling the a-relaxation frequen- Pas) = 4.8488, B = 17.5423 and C1 = 0.01314, with correlation
cies [91]. The value of scaling exponent can be obtained from the coefficient R2 = 0.9943, and standard deviation rln g = 0.309. From
FV model since the properties involved are viscosity, density, and Fig. 13 one can see that at atmospheric pressure the master curve
temperature. For example, taking the isokom line at ln(g/mPas) is in good agreement with experimental data up to C = 5.2  103
= 5, from (q, T) plane represented in Fig. 12, the double logarithmic K1g1.8cm5.4. This value corresponds approximately to the cross-
relation T and q gives the slope c = (1.79 ± 0.04) and considering ing temperature, Tc = 291 K. For the viscosities at pressures differ-
values from the isokom line at glass transition, ln(g/mPas) ent from atmospheric, all the data follow the master curve with
= 34.54, the slope is (1.55 ± 0.31). Both values are consistent with low scattering up to Tc and afterward increasing deviations appear

Fig. 14. Pressure dependence of function Ug following AM (a), MYEGA (b) and constrained BSCNF (c). 4, 5, s, h, e, Cook et al. [11] isothermal data; , Cook et al. [11]
isobaric data at p = (0.7 and 1.3) GPa; , , , McDuffie and Kelly [52] isothermal data; , Harrison and Gosser [68] isothermal data; —, Eq. (26), with parameters G = 0.9 GPa
and b = 0.7 reported by Avramov [25]; h, Eq. (26) with G = 1.0285 GPa and b = 0.5844; -  -, Eq. (26) with G = 1.0833 GPa and b = 0.5707.
A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182 179

in particular for data of McDuffie and Kelly [52]. The isothermal Ug(p) will result. The parameters G and b of organic substances
values (q, ln g) calculated from Eq. (25) are compared with those including glycerol and geoscientifically relevant systems were
resulting from FV in Fig. 11. The main conclusion is that scaling calculated by Avramov [25] by fitting Eq. (26) to the isothermal
the viscosity using density, gives poor results of predicted viscosity viscosity data of Cook et al. [11] with viscosity at atmospheric
at temperatures lower than 273 K but they are acceptable at pressure calculated from AM model. In the present work Ug(p),
increasing temperatures up to 398 K. The (T, q, ln g) surface has data points were calculated from the isothermal experimental
a soft and low slope at low temperatures compared to what is results of Cook et. al. [11], McDuffy and Kelly [52], and Harrison
observed for the free volume model. Some considerations can be and Gosser [68]. Due to analytical and accuracy differences relative
made about the small value of the scaling exponent obtained for to the model equations used previously to represent the viscosity
glycerol. It is usually accepted that large values of c reflect strong at atmospheric pressure, they were all used in Eq. (26) to search
volume contributions to the cooperative dynamics in glass forming for the best representation of Ug(p). The results are plotted in
liquids as is the case of octane which has c = 8 [13]. At the limit Fig. 14 where the values of Ug(p) were represented as a function
c = 0 the temperature will be the dominant control variable and of pressure.
in this scenario viscosity is a function of T alone. This is the case Although AM equation provides the better representation of
of water for which the collapse of viscosity data onto a master viscosity at atmospheric pressure, the temperature dependence
curve is not been observed. This behaviour is due to the hetero- of Ug was not removed. The other models are suitable for the
geneities in water liquid structure where directional interactions representation of Ug(p) because they show the same range of the
among water molecules give rise to an open H-bonded network variation of Ug, and they give a unique function of the pressure
with lower density than unassociated molecules. However, scaling with the almost complete removal of the temperature dependence.
of viscosity with density can be achieved for H- bonded liquids and Namely temperature removal is achieved for MYEGA Eq. (4), and in
using very low values of scaling parameter as in the case of sorbitol this case, the parameters of Ug(p) obtained by fitting are
[C6H8(OH)6] a multilaterally H-bonded glass former. For this sub- G = (1.0285 ± 0.0598) GPa and b = (0.5844 ± 0.0208), with correla-
stance c = 0.13, obtained from dielectric relaxation times, and thus tion coefficient R2 = 0.9946 and standard deviation rln U = 0.0278.
this value cannot be accepted as meaningful because 3c = 0.39 is Similar results could be obtained from constrained BSCNF and
an unimaginable value for the exponent of repulsive part of the BSCNF1 equations. The logarithm of viscosity calculated from the
potential. It is probable that scaling exponent can provide not only expression:
a measure of the repulsive part of molecular potential but also for  p b
the range order resulting from the formation of clusters or hetero- ln g ðp; TÞ ¼ 1 þ ½ln g0 ðTÞ  ln g1  þ ln g1 ð33Þ
P
geneities within the liquid [56]. For glycerol, there are indications
that H-bonding increase with increasing pressure [90] which is
contrary to what happens with water. From molecular dynamics,
Root and Berne [90] verified that single H bonds that make up a
strong hydrogen-bonded network increase with density in the
range 1280 kgm3 to 1380 kgm3 which translated in terms of
the master curve involves all the tested temperatures (263.15 K
to 398.15 K), pressures (0.1 MPa to 3 GPa) and corresponds to vari-
ations of (T1q1.8) between 3.9  103 K1g1.8cm5.4 and
6.5  103 K1g1.8cm5.4. The single H-bonds are about ten times
more present than the multiple bonds which are fairly strained and
weak [90]. This behaviour should maintain a stable liquid structure
that allows scaling. It is interesting to mention that some aliphatic
alcohols present c close to the value found for glycerol [24]:
methanol, c = 1.76; ethanol, c = 1.94; 1-propanol, c = 1.89. As the
scaling exponent reflects the degree to which volume governs
the temperature and pressure dependence of the dynamic process
there have been attempts to relate scaling exponent with dynamic
quantities. From relaxation times of several different classes of
substances including H-bonded ones, Casalini and Roland [91]
found a relationship between c and the ratio EV/EP of activation
enthalpy at constant volume, EV, and at constant pressure EP. This
ratio assumes the value of 1 or 0, for temperature or volume dom-
inated dynamics, respectively. For glycerol EV/EP = 0.75 and this
value should mean some importance of volume in the dynamics
of glycerol as GFL.
Beyond the utility of scaling the viscosity and the relaxation
time of glass formers as a matter of organization and extrapolation
of data, other important reasons make scaling procedures appeal-
ing. The magnitude of c could be a direct link to the way as mole-
cules use the complex potential energy landscape in which the
density-dependent local barriers exceed the available thermal
energy. Thus, molecular motion follows a many-body cooperative
dynamics evolving with changes in temperature and volume [13]. Fig. 15. The scaling with pressure (PS) in terms of pressure, p, and temperature, T,
compared with experimental viscosity. (a) Isothermal (p, ln g) plot; (b) relative
deviations between PS calculated values (gcalc) and the experimental data (g). 4, 5,
4.4.3. Scaling of viscosity with pressure s, h, e, Cook et al. [11]; , , , McDuffie and Kelly [52]; , Harrison and Gosser
According to Eq. (26) the temperature dependence of viscosity [68]; +, Klotz et al. [75]. The results for PS model reported by Avramov [25] are
can be removed and a pressure dependent dimensionless function displayed in (a) by dashed curves.
180 A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182

were compared with isothermal experimental data from literature is possible. However, the scatter existing for different sets of exper-
in Fig. 15. The results can be considered as reliable, with an overall imental data prevents the establishment of a clear and well-
AADg of 17% with individual AADg sets of 22%, 4% and 12% for data of defined master curve in density scaling. More reliable and consis-
Cook et al., McDuffy and Kelly, and Harrison and Gosser, respec- tent high-pressure viscosity data for glycerol are needed.
tively. The overall AADg is close to that found from the calculations
with free volume, but the individual deviations relative to each
Acknowledgements
author are higher with the exception of McDuffy and Kelly.
Similarly, to the FV and DS models, the higher AADg observed in
The authors are thankful to Prof. Klaus Schröter (Institute of
PS for the viscosities of Cook et. al. is mainly due to their viscosity
Physics, University Halle, Germany), to Prof. Matthias Möbius
data scattering observable at all temperatures. In Fig. 15 the results
(School of Physics, Trinity College, Dublin, Ireland) and to Prof. Ste-
of the calculations with Eq. (33) with the parameters reported by
fan Klotz (CNRS, Université Pierre et Marie Currie, Paris, France) for
Avramov [25] are displayed. It can be seen that agreement between
kindly providing us with their measured viscosities. We acknowl-
experimental and calculated values is getting worse with the
edge Prof. Charles Angell (School of Molecular Sciences, University
increase of temperature and it is very poor at temperatures higher
of London) and Prof. Masaru Aniya (Department of Physics, Gradu-
than 295 K.
ate School of Science and Technology, Kumamoto University,
Japan) for clarifying some doubts in the course of his work.
5. Conclusions
Appendix A. Supplementary data
A major problem is that even today experimental values of vis-
cosity of glycerol in good agreement between the various authors Supplementary data associated with this article can be found, in
in wide temperature ranges are not available. This fact motivated the online version, at http://dx.doi.org/10.1016/j.jct.2017.05.042.
us to make new measurements in the region between 293 K and
394 K.
Using our determinations combined with selected values from References
the literature, we applied well known and recent models to corre- [1] D.H. Vogel, The temperature dependence law of the viscosity of fluids, Phys. Z.
late dynamic viscosity of liquid glycerol over the temperature 22 (1921) 645–646.
range from glass transition to 440 K. [2] G. Tammann, W. Hesse, Die Abhängigkeit der Viscosität von der Temperatur
bie unterkühlten Flüssigkeiten, Z. Anorg. Allg. Chem. 156 (1926) 245–257.
The viscosity at atmospheric pressure is described with low
[3] G.S. Fulcher, Analysis of recent measurements of the viscosity of glasses, J. Am.
average absolute percentage deviations, within the experimental Ceram. Soc. 8 (1925) 339–355.
error of viscosity measurements. The AM, MYEGA and BSCNF [4] G. Adam, J. Gibbs, On the temperature dependence of cooperative relaxation
describe data with AADg of 3%, 6%, and 3%, respectively. Imposing properties in glass-forming liquids, J. Chem. Phys. 43 (1965) 139–146.
[5] I. Avramov, A. Milchev, Effect of disorder on diffusion and viscosity in
constraints is needed to have physical meaning for the parameters condensed systems, J. Non-Cryst. Solids 104 (1988) 253–260.
for BSCNF but constraint is not needed for BSCNF1. With simulta- [6] A. Milchev, I. Avramov, On the influence of amorphization on atomic diffusion
neous use of <T0g> and log(gg/Pas) = 12 as constraints, the AADg is in condensed systems, Phys. Status Solidi B 120 (1983) 123–130.
[7] I. Avramov, Influence of disorder on viscosity of undercooled melts, J. Chem.
maintained at reasonable levels (12–13%). The use of AM Eq. (8) Phys. 95 (1991) 4439–4443.
is recommended for the representation of viscosity of glycerol. [8] J.A.T. González, M.P. Longinotti, H.R. Corti, The viscosity of glycerol-water
The stickel derivative calculated from AM and MYEGA equa- mixtures including the supercooled region, J. Chem. Eng. Data 56 (2011) 1397–
1406.
tions are in good agreement with the numerical calculation from [9] M. Aniya, A model for fragility of the melts, J. Therm. Anal. Calorim. 69 (2002)
experimental data over wide ranges of temperature. The BSCNF 971–978.
equation performs better in the T range where VFT has been devel- [10] A.K. Dolittle, Studies in newtonian flow. II. The dependence of the viscosity of
liquids on free-space, J. Appl. Phys. 22 (1951) 1471–1475.
oped, i.e., the supercooled region. The comparison between numer- [11] R.L. Cook, H.E. King, C.A. Herbst, D.R. Herschbach, Pressure and temperature
ical and model Stickel derivatives is a good test about how close dependent viscosity of two glass forming liquids: glycerol and dibutyl
the models follow the data because the derivative of ln g is phthalate, J. Chem. Phys. 100 (1994) 5178–5189.
[12] C. Dreyfus, A. Le Grand, J. Gapinski, W. Steffen, A. Patkowski, Scaling the a-
involved which is more sensitive to temperature variations.
relaxation time of supercooled fragile organic liquids, Eur. Phys. J. B 42 (2004)
Reliable values of glass transition temperatures and fragility 309–319.
indexes (within the calculated uncertainty given in the literature) [13] C.M. Roland, S. Bair, R. Casalini, Thermodynamic scaling of the viscosity of van
and their pressure dependencies were obtained by fitting AM, der Waals, H-bonded, and ionic liquids, J. Chem. Phys. 125 (2006), 124508-1-8.
[14] S. Bair, R. Casalini, A scaling parameter and function for the accurate
MYEGA and BSCNF1 equations to data. The fragility index obtained correlation of viscosity with temperature and pressure across eight orders of
at atmospheric pressure (m ranging from 43 to 51) is consistent magnitude of viscosity, ASME J. Tribol. 130 (2008) 041802–41807.
with values from literature and characterizes glycerol as a moder- [15] M. Paluch, K. Grzybowska, A. Grzybowski, Effect of high pressure on the
relaxation dynamics of glass-forming liquids, J. Phys.: Condens. Matter 19
ate fragile liquid. (2007), 205117-12.
Several advancements at the light of the physical picture of [16] Brookfield Engineering Labs Inc, More Solutions to Sticky Problems. A guide to
glass melts were possible with the application of BSCNF theory. getting more from your Brookfield viscometer & Rheometer, Technical
Bulletin, Stoughton, MA.
This aspect could be more important than the intrinsic precision [17] E. Rössler, K.-U. Hess, V.N. Novikov, Universal representation of viscosity in
of fitting with this model. BSCNF1 equations suggest that about glass forming liquids, J. Non-Cryst. Solids 223 (1998) 207–222.
(7–11) structural units must be broken in order to observe the vis- [18] M. Paluch, R. Casalini, S. Hensel-Bielowka, C.M. Roland, Effect of pressure on
the a relaxation in glycerol and xylitol, J. Chem. Phys. 116 (2002) 9839–9844.
cous flow near the glass transition at atmospheric pressure. This [19] K. Schröter, E. Donth, Viscosity and shear response at the dynamic glass
number should increase to 34 at 3 GPa. This situation could be transition of glycerol, J. Chem. Phys. 113 (2000) 9101–9108.
the result of the increasing range of hydrogen bond energies with [20] J. Mauro, Y. Yue, A. Ellison, P. Gupta, D. Allan, Viscosity of glass-forming liquids,
Proc. Natl. Acad. Sci. U.S.A. 106 (2009) 19780–19784.
pressure resulting also in the increase of fragility.
[21] P.K. Gupta, J.C. Mauro, Composition dependence of glass transition
High-pressure viscosities are reasonably represented using free temperature and fragility. I. A topological model incorporating temperature
volume and scaling with pressure methods with AADg of 14% and dependent constraints, J. Chem. Phys. 130 (2009) 94503–94510.
17%, respectively. Taking into account the available experimental [22] M. Ikeda, M. Aniya, Bond strength-coordination number fluctuation model of
viscosity: an alternative model for the Vogel-Fulcher-Tammann equation and
viscosity data of glycerol in extended temperature and pressure an application to bulk metallic glass forming liquids, Materials 3 (2010) 5246–
ranges it can be concluded that scaling of viscosity with density 5262.
A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182 181

[23] M. Ikeda, M. Aniya, Correlation between fragility and cooperativity in bulk longitudinal and shear acoustic dynamics in glycerol and DC704 studied by
metallic glass-forming liquids, Intermetallics 18 (2010) 1796–1799. time-domain Brillouin scattering, J. Chem. Phys. 138 (2013), 12A544-12.
[24] E.R. López, A.S. Pensado, M.J.P. Comuñas, A.A.H. Pádua, J. Fernández, K.R. [55] M.E. Möbius, T. Xia, W. van Saarloos, M. Orrit, M. van Hecke, Aging and
Harris, Density scaling of the transport properties of molecular and ionic solidification of supercooled glycerol, J. Phys. Chem. B 114 (2010) 7439–7444.
liquids, J. Chem. Phys. 134 (2011) 144507–144511. [56] A. Reiser, G. Kasper, S. Hunklinger, Pressure-induced isothermal glass
[25] I. Avramov, Pressure and temperature dependence of viscosity of glassforming transition of small organic molecules, Phys. Rev. B 72 (2005) 094204–94207.
and of geoscientifically relevant systems, J. Volcanol. Geoth. Res. 160 (2007) [57] O.N. Senkov, D.B. Miracle, Description of the fragile behaviour of glass-forming
165–174. liquids with the use of experimentally accessible parameters, J. Non-Cryst.
[26] R.C. Wilhoit, J. Chao, K.R. Hall, Thermodynamic properties of key organic Solids 355 (2009) 2596–2603.
oxygen compounds in the carbon range C-1 to C-4. Part 1. Properties of [58] L.M. Wang, C.A. Angell, R. Richert, Fragility and thermodynamics in
condensed phases, J. Phys. Chem. Ref. Data 14 (1985) 1–175. nonpolymeric glass-forming liquids, J. Phys. Chem. 125 (2006) 074505–74508.
[27] M. Volmer, M. Marder, Zur theorie der linearen kristallisations- [59] R. Böhmer, K.L. Ngai, C.A. Angell, D.J. Plazek, Nonexponential relaxations in
geschwindigkeit unterkuhlter schmeltzen und unterkuhlter fester strong and fragile glass formers, J. Chem. Phys. 99 (1993) 4201–4209.
modifikationen, The theory of the linear crystalisation rate of under cooled [60] V.N. Filipovich, The viscosity of SiO2 and GeO2 glasses added R2O, Glass Phys.
melts and under cooled firm modifications, Z. Physik. Chem. 154 (1931) 97– Chem. 1 (1975) 426.
112. [61] V.M. Fokin, M.L.F. Nascimento, E.D. Zanoto, Correlation between maximum
[28] NIST: National Institute of Standards and Technology, Chemistry WebBook, crystal growth rate and glass transition temperature of silicate glasses, J. Non-
http://webbook.nist.gov/chemistry/fluid/ (accessed 26.05.17). Cryst. Solids 351 (2005) 789–794.
[29] C.A. Angell, Formation of glasses from liquids and biopolymers, Science 267 [62] M.L.F. Nascimento, C. Aparicio, Viscosity of strong and fragile glass-forming
(1995) 1924–1935. liquids investigated by means of principal component analysis, J. Phys. Chem.
[30] J. Li, A. Kushima, J. Eapen, X. Lin, X. Qian, J.C. Mauro, P. Diep, S. Yip, Computing Sol. 68 (2007) 104–110.
the viscosity of supercooled liquids: Markov network model, PLoS ONE 6 (3) [63] G.S. Parks, W.A. Gilkey, Some viscosity data on liquid glucose and glucose-
(2011), e17909-7. glycerol solutions, J. Phys. Chem. 33 (1929) 1428–1437.
[31] S. Magazù, F. Migliardo, N.P. Malomuzh, I.V. Blazhnov, Theoretical and [64] M.L. Sheely, Glycerol viscosity tables, Ind. Eng. Chem. 24 (1932) 1060–1064.
experimental models on viscosity: I, Glycerol, J. Phys. Chem. B 111 (2007) [65] V. Vand, The measurement of the viscosity of glycerol at high temperatures,
9563–9570. Research 1 (1947) 44–45.
[32] A. Takeuchi, H. Kato, A. Inoue, Vogel-Fulcher-Tammann plot for viscosity [66] J.B. Segur, H.E. Oberstar, Viscosity of glycerol and its aqueous solutions, Ind.
scaled with temperature interval between actual and ideal glass transitions for Eng. Chem. 43 (1951) 2117–2120.
metallic glasses in liquid and supercooled liquid states, Intermetallics 18 [67] R. Piccirelli, T.A. Litovitz, Ultrasonic shear and compressional relaxation in
(2010) 406–411. liquid glycerol, J. Acoust. Soc. Am. 29 (1957) 1009–1020.
[33] L. Hong, V.N. Novikov, A.P. Sokolov, Is there a connection between fragility of [68] D.E. Harrison, R.B. Gosser, Rolling ball viscometer for use at temperature to
glass forming systems and dynamic heterogeneity/cooperativity?, J Non-Cryst. 400 °C under pressures to 5 kilobar, Rev. Sci. Instru. 36 (1965) 1840–1843.
Solids 357 (2011) 351–356. [69] W.M. Slie, W.M. Madigosky, Pressure dependence of the elastic moduli of
[34] A. Drozd-Rzoska, S.J. Rzoska, M. Paluch, A.R. Imre, C.M. Roland, On the glass liquid glycerol, J. Chem. Phys. 48 (1968) 2810–2817.
temperature under extreme pressures, J. Chem. Phys. 126 (2007) 164504– [70] A.A. Berdyev, V.A. Lysenko, B. Khemraev, Absortion and dispersion of
164507. ultrasound and hypersound in glycerine, Sov. Phys. JETP 38 (1974) 515–516.
[35] G.P. Johari, E. Whalley, Dielectric properties of glycerol in the range 0.1– [71] Y.F. Kiyachenko, Y.I. Litvinov, Increase in scale length in a liquid as the glass
105 Hz, 218–357 K, 0–53 kb, Farad. Sym. Chem. Soc. 6 (1972) 23–41. transition temperature is approached, JETP Lett. 42 (1985) 266–269.
[36] O. Sandberg, P. Anderson, G. Bäckström, Proc. 7th Symp. Thermophys. Prop. [72] P.N. Shankar, M. Kumar, Experimental determination of the kinematic
(ASME, New York, 1977) 181–184. viscosity of glycerol-water mixtures, Proc. R. Soc. Lond. A 444 (1994) 573–581.
[37] C.A. Herbst, R.L. Cook, H.E. King Jr., High-pressure of glycerol measured by [73] C.M. Kinart, W.J. Kinart, Physicochemical properties of glycerol-formamide
centrifugal-force viscometry, Nature 361 (1993) 518–520. liquid mixtures and their assumed internal structures, Phys. Chem. Liq. 33
[38] E. Hempel, G. Hempel, A. Hensel, C. Schick, E. Donth, Characteristic lengh of (1996) 159–170.
dynamic glass transition near Tg for a wide assortment of glass-forming [74] M.-L. Ge, J.-L. Ma, B. Chu, Densities and viscosities of propane-1,2,3-triol
substances, J. Phys. Chem. B 104 (2000) 2460–2466. +ethane-1,2-diol at T= (298.15 to 338.15) K, J. Chem. Eng. Data 55 (2010)
[39] J. Bartoš, O. Šanša, J. Krištiak, T. Blodrowicz, E. Rössler, Free-volume 2649–2651.
microstructure of glycerol and its supercooled liquid-state dynamics, J. [75] S. Klotz, K. Takemura, Th. Strässle, Th. Hansen, Freezing of glycerol-water
Phys.: Condens. Matter 13 (2001) 11473–11484. mixtures under pressure, J. Phys.: Condens. Matter 24 (2012) 325103–325106.
[40] L.M. Wang, V. Velikov, C.A. Angell, Direct determination of kinetic fragility [76] M.Lj. Kijevcanin, E.M. Zivkovic, B.D. Djordjevic, I.R. Radovic, J. Jovanovic, S.P.
índices of glassforming liquids by differential scanning calorimetry: kinetic Šerbanovic, Experimental determination and modeling of excess molar
versus thermodynamic fragilities, J. Chem. Phys. 117 (2002) 10184–10192. volumes, viscosities and refractive indices of the binary systems (pyridine +
[41] O. Trofymluk, A.A. Levchenko, A. Navrotsky, Interfacial effects on vitrification 1-propanol, +1,2-propanediol, +1,3-propanediol, and +glycerol). New UNIFAC-
of confined glass-forming liquids, J. Chem. Phys. 123 (2005), 194509 -7. VISCO parameters determination, J. Chem. Thermodynamics 56 (2013) 49–56.
[42] A. Reiser, G. Kasper, On the pressure dependence of fragility, Europhys. Lett. 76 [77] F. Bandarkar, I.S. Khattab, F. Martinez, M. Khoubnasabjafaric, S. Vahdati, A.
(2006) 1137–1143. Jouyban, Viscosity and surface tensiono f glycerol+N-methyl-2-pyrrolidone
[43] M.S. Elsaesser, I. Kohl, E. Mayer, T. Loerting, Novel method to detect the mixtures from 293 to 323 K, Phys. Chem. Liq. 53 (2015) 104–116.
volumetric glass?liquid transition at high pressures: glycerol as a test case, J. [78] J.G. Lynam, G.I. Chow, P.L. Hyland, C.J. Coronella, Corn stover pretreatment by
Phys. Chem. B 111 (2007) 8038–8044. ionic liquid and glycerol mixtures with their density, viscosity and
[44] S. Pawlus, M. Paluch, J. Ziolo, C.M. Roland, On the pressure dependence of the thermogravimetric properties, ACS Sustain. Chem. Eng. 4 (2016) 1765–1774.
fragility of glycerol, J. Phys.: Condens. Matter 21 (2009) 332101–332103. [79] F. Stickel, E.W. Fischer, R. Richert, 104, Dynamics of glass-forming liquids. II.
[45] A.A. Pronin, M.V. Kondrin, A.G. Lyapin, V.V. Brazhkin, A.A. Volkov, P. Detailed comparison of dielectric relaxation, dcconductivity, and viscosity
Lunkenheimer, A. Loidl, Pressure induced change in the relaxation dynamics data, J. Chem. Phys. 104 (1996) 2043–2055.
of glycerol, JETP Lett. 92 (2010) 479–483. [80] C. Hansen, F. Stickel, P. Berger, R. Richert, E.W. Fischer, Dynamics of glass-
[46] K. Sou, K. Nishikawab, Y. Kogac, K. Tozakid, High-resolution calorimetry on forming liquids. III. Comparing the dielectric a- and b-relaxation of 1-propanol
thermal behaviour of glycerol (I): glass transition, crystallization and melting, and o-terphenyl, J. Chem. Phys. 107 (1997) 1086–1093.
and discovery of a solid–solid transition, Chem. Phys. Lett. 506 (2011) 217– [81] J.C. Martinez-Garcia, J. Martinez-Garcia, S.J. Rzoska, J. Hulliger, The new insight
220. into dynamic crossover in glass forming liquids from the apparent enthalpy
[47] J. Swenson, K. Elamin, H. Jansson, S. Kittaka, Why is there no clear glass analysis, J. Chem. Phys. 137 (2012) 064501–64508.
transition of confined water?, Chem Phys. 424 (2013) 20–25. [82] R. Richert, C. Angell, Dynamics of glass-forming liquids. V. On the link between
[48] M.V. Roussenova, M.A. Alam, A.S. Townrow, J.D. Kilburn, P.E. Sokol, R. Guillet- molecular dynamics and configurational entropy, J. Chem. Phys. 108 (1998)
Nicolas, F. Kleitz, A nano-scale free volume perspective on the glass transition 9016–9026.
of supercooled water in confinement, New J. Phys. 16 (2014) 1–14. [83] Nieves M.C.T. Prieto, Thiago A. Souza, Ana P. Egas, Abel G.M. Ferreira, Lélio Q.
[49] G.E. Gibson, W.F. Giauque, The third law of thermodynamics. Evidence from Lobo, António T.A. Portugal, Liquid glycerol: experimental densities at
the specific heats of glycerol that the entropy of a glass exceeds that of a pressures up to 25 MPa, and some derived thermodynamic properties, J.
crystal at the absolute zero, J. Am. Chem. Soc. 45 (1922) 93–104. Chem. Thermodyn. 101 (2016) 64–77.
[50] J.V. Blazhnov, N.P. Malomuzh, S.V. Lishchuk, Temperature dependence of [84] Y.-M. Chen, A.J. Pearlstein, Viscosity-temperature correlation for glycerol-
density, thermal expansion coefficient and shear viscosity of supercooled water solutions, Chem. Res. 26 (1987) 1670–1672.
glycerol as a reflection of its structure, J. Chem. Phys. 121 (2004) 6435–6441. [85] E. Donth, Characteristic lengh of the glass transition, J. Polym. Sci., Part B:
[51] W. Kauzmann, The nature of the glassy state and the behavior of liquids at low Polym. Phys. 34 (1996) 2881–2892.
temperatures, Chem. Rev. 43 (1948) 219–256. [86] M. Aniya, M. Ikeda, A study on the correlation between fragility and
[52] G.E. McDuffie, M.V. Kelly, Effect of pressure on the viscosity and dielectric cooperativity in wide class of glass-forming substances, Physics Procedia 48
relaxation time in glycerol, J. Chem. Phys. 41 (1964) 2666–2670. (2013) 113–119.
[53] J. Huck, A. Bondeau, G. Noyel, L. Jorat, Transport and relaxation phenomena in [87] D. Fragiadakis, R. Casalini, C.M. Roland, Comparing dynamic correlation
highly viscous liquids, IEEE Trans. Elect. Ins. 23 (1988) 615–625. lengths from an approximation to the four-point dynamic susceptibility and
[54] C. Klieber, T. Hecksher, T. Pezeril, D.H. Torchinsky, J.C. Dyre, K.A. Nelson, from the picosecond vibrational dynamics, Phys. Rev. E 84 (2011) 042501–
Mechanical spectra of glass-forming liquids. II. Gigahertz-frequency 42504.
182 A.G.M. Ferreira et al. / J. Chem. Thermodynamics 113 (2017) 162–182

[88] J. Jonas, T. DeFries, D.J. Wilbur, Molecular motions in compressed liquid water, [91] R. Casalini, C.M. Roland, Thermodynamical scaling of the glass transition
J. Chem. Phys. 65 (1976) 582–588. dynamics, Phys. Rev. E 69 (2004) 62501–62503.
[89] R.F. Marzke, D.P. Raffaelle, K.E. Halvorson, G.H. Wolf, Second International
Discussion Meeting on Relaxations in Complex Systems, Edt, K. L. Ngai,
Alicante, Spain, 1993.
[90] L.J. Root, B.J. Berne, Effect of pressure on hydrogen bonding in glycerol: a
molecular dynamics investigation, J. Chem. Phys. 107 (1997) 4350–4357. JCT 17-125

Anda mungkin juga menyukai