Sei sulla pagina 1di 10

World Research Journal of Chemistry

Author Galley Proof || Author Galley Proof.


Available online at http://www.bioinfopublication.org/jouarchive.php?opt=&jouid=BPJ0000295

MOLECULAR, KINETIC AND THERMODYNAMIC FEATURES OF SUPERCOOLED LIQUIDS GLASS TRANSITION AND KAUZMANN PARADOX
SIVAPRAKASAM K.*
Department of Chemistry and Physics, St. Cloud State University, 720 4th Ave South, St. Cloud, MN, USA.
*Corresponding Author: Email- ks248siva@gmail.com
Received: January 30, 2014; Accepted: February 12, 2014
Abstract- Supercooled liquid-glass transition has both kinetic and thermodynamic features that make it a truly complex process to understand.
The origin of dynamic heterogeneity as molecules slip into glassy state is intriguing and the molecular mechanism underlying these events is
not clear. This is largely due to lack of reasonable molecular models and experimental techniques to follow the vitrification process at molecular scale. The complex dynamics of glasses involving sluggish, cooperative kinetics leading to structural arrest and freeze in entropy is discussed from various perspectives. Can the slowing down of flow and relaxation behavior at low temperature be addressed by a unified theory?
This review critically analyzes the generalized features of glasses that could pave the way for an elusive universal theory of glass transition.
Keywords- Supercooled liquids, Glass transition, Kauzmann paradox, Viscosity, Ideal glass, Free volume
Introduction
Supercooled liquid-glass transition has been intensively studied due
to its broad significance in physical, material and biological sciences
[1,2]. Despite extensive research spanning several decades [3],
there still remain many unresolved questions: What is the driving
force for glass transition? Is glass transition purely a kinetic phenomenon or is there a thermodynamic factor that underlies vitrification process. Molecular events occurring before, during and after
the phase transformation are not clearly understood. Glass, like
liquid, is both intriguing and frustrating. Both glass and liquid phases do not have a suitable model like crystal or gas. Despite concerted efforts of both experimentalists and theoreticians, molecular
understanding of glass transition and glassy state remains unclear
[4]. Anderson (1977 Nobel laureate in Physics) aptly sums it up,
"the deepest and the most interesting unresolved problem in solid
state theory is probably the theory of nature of glass and glass transition" [5].
Our interest in glass transition originated with an aim to understand
glass formation of long chain plasticizing molecules, specifically
alkanes and carboxylate esters having molecular weight in the
range of 400 - 2000 Daltons [6-11]. Knowledge of glass forming
tendency of these long chain molecules is expected to lead to judicious selection or rational design of plasticizers, besides it would
also help to understand their plasticization mechanism. Furthermore, these molecules can potentially serve as model compounds
for long chain polymers. Despite the technological importance of
these long chain molecules as plasticizers and lubricants it is surprising that adequate attention has not been paid to understand
their glass forming tendency. In this respect, flow, glass transition
and plasticization behavior of long chain carboxylate esters is one
of the few comprehensive studies undertaken [6,9-11]. We had also
examined thermodynamic and kinetic behavior of phosphate esters
which are used as flame-retardant plasticizers [7].
This review examines molecular, kinetic and thermodynamic factors
affecting the formation and stability of glassy state. Paradoxical

situation arising due to intrusion of kinetics into the domain of thermodynamic phase transitions and the significance of Kauzmann
temperature [12] is discussed. Influence of experimental conditions
affecting supercooled liquids and the physical concepts of the theories of glass transition are analyzed. A novel aspect of this review is
the use of the old free volume theory in interpreting recent experimental findings.
Supercooled Liquids and Glass Formation
Liquids at temperatures below their melting point (Tm) are called
supercooled liquids [13,14]. All liquids when cooled fast enough can
escape crystallization: the competition between rates of crystallization and cooling determines the solid state structure at low temperatures [14]. Glass formation is kinetically favored if the rate of nucleus formation and the velocity of crystal growth are low enough to
avoid crystallization [15]. Viscous liquids with high enthalpy of acti#
vation for viscous flow (H ) readily form glass [16]. For many
glass formers, in particular metals, crystallization can be avoided by
using small sample volumes, since this helps to increase cooling
rate and reduce nucleation probability [17]. Kinetic factors not only
influence rate and extent of formation of glassy state, it can also
play a critical role in the stability of the metastable phase.
Glass
Classification of thermodynamic phases into gas, liquid and solid
leaves no room for glassy phase which combines the irregular
structure of a liquid with the rigidity of a solid [18]. Glass can be
defined in various ways. Glassy state is a hybrid form of matter that
maintains the microstructure, energy, and volume of a liquid, but
changes in energy and volume with temperature are similar in magnitude to that of a crystalline solid [19]. Glass can also be considered as an extension of a supercooled liquid state whose viscosity
(h) is above 1013 Poise, since at this viscosity a material would substantially maintain its shape against shearing forces [20]. The local
structure of glass is similar to that of supercooled liquid as is confirmed from X-ray diffraction studies [21]. There is no significant

World Research Journal of Chemistry


|| Author Galley Proof || Author Galley Proof ||
|| Bioinfo Publications ||

Sivaprakasam K. (2014) Molecular, Kinetic and Thermodynamic Features of Supercooled Liquids - Glass Transition and Kauzmann Paradox.
World Research Journal of Chemistry, Author Galley Proof || Author Galley Proof || Author Galley Proof.

change in spatial atomic configuration at glass transition. From this,


it is clear that glass has the external appearance of a solid, but
maintains the molecular arrangement of a liquid, lacking long range
order and the periodic structure of a crystalline solid [22]. Glasses
are amorphous in nature due to the presence of long-range disorder. Absence of the thermal averaging effects present in liquids
ensures freeze of disorder in glasses so that each site contributes
in a unique manner to the ensemble average. Therefore, physical
properties that are single valued in both crystals and liquids, will
have a statistical distribution in glasses [23]. This is evident from the
relaxation studies of chromophores in supercooled liquids which
display large spread of rotational correlation times indicating heterogeneity [24,25]. Irrespective of the chemical nature, most glasses
display unique features that are absent in their crystalline counterparts. Bulk modulus and disorder appears to be the dominant factor
that causes its characteristic properties (packing coefficient, viscosity, wide variation in correlation times, etc.) [13,26]. Recent studies
indicate that glass transition temperature (Tg) depends on thickness
of the material [27,28]. As a material reaches nan dimensions, T g of
the substance is reduced. This is due to the increasingly predominant role played by the mobile surface layer as the size gets reduced.
Ideal Glass
Poor understanding of the nature of glassy state may be partly ascribed to the lack of suitable models. However, some progress has
been made in this aspect with the introduction of ideal glass concept. An ideal glass does not have a unique arrangement like a
crystalline solid [29]; the positions are randomly distributed and
hence do not exhibit translational symmetry [Fig-1]. But both crystal
and ideal glass have fixed geometrical positions about which atoms/
molecules oscillate. For a glassy state, residual entropy is a measure of the deviation from ideal behavior [30]. Gibbs and DiMarzio
[31,32] first suggested ideal glassy state for polymeric systems and
later Cohen and Turnbull [33,34] extended this concept to simpler,
even monatomic glasses. For systems that are reluctant to crystallize, like polymers, ideal glass concept was suggested as a reference state and is used to calculate the Gibbs free energy of formation of glass.

Fig. 1- Schematic representation of crystal and ideal glass


Glass and Crystal
Although both glass and crystal have different structural arrangements of atoms/molecules, there is a correlation among their properties. Tg and Tm of a substance have been found to follow a linear
relationship, though the molecular mechanisms that underlie both
transitions differ; while glass transition has blend of thermodynamic
and kinetic character, fusion is a thermodynamic first order phase
transition. The approximate rule which relates these two tempera-

tures is called Beaman-Kauzmann rule [35] : Tg ~ 2/3 Tm. Packing


of molecules in these two states are also closely related. Packing
and dynamics in glasses is a universal feature [36] and there is a
correlation between the densities of materials in vitreous and crystalline states [37]. This means that the relative change in specific
volume during glass - crystal transition is constant.
Kinetics of Glass Formation
As supercooled liquid is cooled to low temperatures, its viscosity
increases accompanied by an increase in relaxation time for structural rearrangements which results in slow molecular motion [38].
These rearrangements are vital for the formation of a crystalline
lattice and are also responsible for a liquid to attain equilibrium. As
this process is slowed down, rate of crystallization becomes slow
and the system is kinetically entrapped in a thermodynamically
metastable phase. Eventually time scale for molecular rearrangements becomes long compared to time scale of experimental observation. This results in a sudden drop in specific heat (Cp) and thermal expansion coefficient () with freezing of translational degrees
of freedom [39]. The structure of the material is "frozen" for all practical purposes, resulting in formation of glass. A glassy state is
formed when a supercooled liquid "falls out of thermodynamic equilibrium. According to Kauzmann, transition from supercooled liquid
phase to glassy state is really a relaxation phenomenon arising
from slowness with which molecules change their positions at low
temperatures [40]. As a result of this sluggish molecular motion,
supercooled liquid is unable to change its structure appreciably
during the time required to measure the various properties such as
the Cp, density and compressibility (). Hence, the contribution of
changing liquid structure is absent in glasses which results in abrupt
fall of Cp and . The transition temperature from supercooled liquid
to glass varies depending on the experimental conditions.
Cooling a liquid below its Tm is certainly not the only route for glass
formation. Materials can also be obtained in amorphous form by
depositing them onto cold substrate, which is kept at temperatures
lower than their Tg, thereby preventing crystallization. The transition
from a crystalline state to a glassy state is kinetically forbidden, as
the process of rearrangement requires considerable energy to overcome activation barriers caused by molecular structure. But in
methods associated with high energy and temperature such energetic transition do takes place [41]. This has been observed in several metal alloys and minerals by using experimental techniques
such as solidstate reaction [42], mechanical alloying [43], high
pressure [44], ionbeam mixing [41,45], ball milling [46] and hydriding [47]. Among the various ways to glassy state it is the temperature quenching route (rapidly cooling a melt without allowing
enough time for a molecule to orient and form crystal) that has been
the most intensively studied method [48].
As glass formation depends both on experimental time scale and
the time scale for molecular rearrangement, thermodynamic and
dynamic properties of a glass are influenced by the cooling rate and
the time scale of the measurement. Cooling rates for the formation
of glass, however, vary widely depending on the material. For example, it is of the order of 100 degrees/sec for cooling metallic alloys [49] into glass while it is a negligible fraction of a degree/sec for
cooling B2O3 into glass [50]. With an increase in time scale of observation, Tg was shown to decrease steadily [51]. Thus, glass transition is not well defined experimentally. The question of what phenomena would be encountered under longer and longer experimental time scales remains to be properly resolved. Tg represents a

World Research Journal of Chemistry


|| Author Galley Proof || Author Galley Proof ||
|| Bioinfo Publications ||

First Author, Second Author and Third Author (2014) Title of article Title of article Title of article Title of article Title of article Title of article Title of article.
World Research Journal of Chemistry, Author Galley Proof || Author Galley Proof || Author Galley Proof.

practical limit to equilibrium measurements. Theoretically, however,


an ideal glass transition (IGT, To) can be defined to occur at a temperature at which the excess entropy (entropy of supercooled liquid
- entropy of crystal) vanishes. This may occur at the limit of temperature change assuming that the system remains in equilibrium and
crystallization does not intervene.
Glass Forming Temperament
Glass formation is promoted by one or more of the molecular properties such as ionic, hydrogen bonding or dipolar forces [52]. Depending on the type of interaction, glasses can be classified into
different groups [Table-1].
Table 1- Types of glass forming substances
Bond types

Examples

Covalent

Oxides (silicates, borates, phosphates, germanates)


chalcogenides organic polymers

Ionic
Halides, nitrates, carbonates, sulfates
Molecular (van der Waals) Organic liquids
Metallic
Splat-cooled metals and alloys

Glass forming compounds can also be distinguished by their characteristic properties. Departure from Arrhenius law, which governs
many rate processes in nature, is the most important feature of
some of the glass forming liquids. They also exhibit distinct features
like non-Debye relaxation and scaling [53].
Glass forming ability can be qualitatively assessed by the value of
their reduced temperature, Tr (Tg / Tm). Simple nucleation theory
indicates that melts with Tr > 2/3 should readily form glasses [54].
This condition holds good for many simple molecular substances.
Since in metals Tr is low [Table-2] high quench rates of the order of
106 K/sec are necessary to form metallic glasses [55]. In general,
higher the Tr, greater is the glass forming tendency [Table-2].

that can influence glass transition temperature. Most glassy polymers have an inherent structural irregularity that prevents them
from forming crystalline lattices. In the melt, long chain polymers
form statistical coils that strongly interpenetrate each other. Crystallization therefore becomes a very complicated process, strongly
impeded by topological constraints - the coil has to disentangle in
order to form crystalline state [57-59]. In addition, the microstructure
of polymer chains in general introduces further elements of disorder. The structural irregularity can take many forms: (i) in atactic
polymers the side chains are positioned in random directions along
the chain, (ii) double bonds along the main chain backbone can be
linked by substituent at random relative orientations (iii) in copolymers the sequence of the arrangement of the monomers in the
chain affect the solid state structure, for example most of the random copolymers are non-crystalline [58]. Due to these elements of
disorder polymer prefers glassy state upon cooling. A structurally
simple and flexible polymer like polyethylene which possesses a
completely regular microstructure crystallizes partially from the melt.
Even if a polymer possesses the requisite chain structure and is
inherently capable of crystallization, it will fail to crystallize if it is
cooled rapidly.
Glass Transition and Molecular Structure
Molecular weight has a significant effect on the value of Tg [Fig-2].
Tg initially increases with an increase in molecular weight, gradually
becomes less sensitive and finally beyond a critical molecular
weight (Mc) it reaches a constant value. The tapering of the curve
marks the onset of segmental dynamics in long chain molecules
[54]. In polymers, the addition of side chains or pendant groups
alters the Tg. If the pendant group is rigid then Tg increases with the
size of the group [55]. On the other hand, flexible side groups plasticizes the polymer and bring down the Tg. Addition of polar group
brings about greater interaction among the molecules thereby reducing their mobility resulting in higher Tg.

Table 2- Tr of various substances


Substances
3-methylhexane
Ethanol
n-propanol
sec-butyl alcohol
tri-a-napthylbenzene
Sulfur
Glucose
Glycerol
Ni
Au
Al
aAl

Tr

Reference

0.52 - 0.58
0.58 - 0.62
0.59 - 0.68
0.54 - 0.62
0.745
0.622
0.668 - 0.716
0.62 - 0.65
0.23
0.27
0.0a

[103]
[103]
[103]
[103]
[103]
[103]
[103]
[103]
[48]
[48]
[48]

has not been prepared in the glassy state by existing routes [48].

Apart from the Tr criterion, for metals and metalloids it was suggested [47,48] that there is a correlation between the glass forming
ability and the volume change on melting (Vm = Vl - Vs). It is
shown that metals with the low Vm are easier to quench into a
glassy state [56]. On the other hand, metals with higher Vm resist
transformation to a glassy state, because their viscosity remains
low even at high undercooling. Although a few empirical relations,
TbTm > 2 and Tb/Tm > 2.5 holds good for many glass forming systems [57] it fails in the case of hydrogen bonded substances.
In polymers, although there are no such empirical guidelines to test
the glass forming tendency, there are certain structural parameters

Fig. 2- Schematic representation of dependence of Tg on molecular


weight
Glass Transition of Long Chain Molecules
Except for nalkanes, glass transition of high molecular weight
molecules (400 - 2000 Da) has rarely been studied, despite their
technological importance as plasticizers and lubricants. For the first
time a comprehensive study of the glass transition (both experimentally - using Differential scanning calorimeter- and theoretically, by
Vogel - Tammann - Fulcher (VTF) equation) of long chain carboxylate esters (300 - 1000 Da) were carried out [6, 10, 11]. The influ-

World Research Journal of Chemistry


|| Author Galley Proof || Author Galley Proof ||
|| Bioinfo Publications ||

Sivaprakasam K. (2014) Molecular, Kinetic and Thermodynamic Features of Supercooled Liquids - Glass Transition and Kauzmann Paradox.
World Research Journal of Chemistry, Author Galley Proof || Author Galley Proof || Author Galley Proof.

ence of molecular weight and branching on the glass transition of


these esters was investigated. In the linear carboxylates, Tg decreases continuously with molecular weight [59]. As the reduction of
Tg is a reflection of enhanced flexibility, the molecules are more
flexible with the addition of CH2 groups. On the other hand, in the tri
and tetra esters the value of Tg decreases initially, hits a minimum
and then starts increasing with increase of molecular weight. The
decrease of Tg initially could be attributed to the enhanced flexibility
due to the addition of CH2 groups [59]. But after a critical molecular
weight, as the elongated chains have greater possibility of contacting each other during their motion, thereby limiting the movements
of the molecules. This causes decrease in flexibility of these esters
that is reflected in the increase of their Tg values.

der Waals with non-Arrhenius viscous behavior are a typical example of a fragile liquid [64]. In between strong and fragile extremes
falls a wide variety of intermediate bonding types, all of which contain specific interactions such as hydrogen bonds or weak covalent
interactions.
Apart from the viscosity criteria, the strength of supercooled liquids
and glasses can also be distinguished based on their relaxation
properties. The most commonly used decay function (f (t)) for handling relaxation data of glasses is the Kohlrausch - Williams - Watts
(KWW), also called as stretched exponential function.

Viscosity
Viscosity of a glass is of practical importance during all stages of
the material manufacturing process [60]. Viscosity is a function of
the state of a fluid, like temperature, pressure and volume. It is a
dynamic non-equilibrium property of a fluid and it represents the
macroscopic measure of resistance to flow. It is related to structural
relaxation time () [61]. Viscous flow and relaxation processes are
sensitive kinetic parameters that characterize the dynamic state of
supercooled liquid or glass. Besides, they also function as indicators of glass transition. Viscosity of a liquid increase drastically on
lowering temperature and approaches 1013 Poise as Tg is reached.
It may be recognized that such a high viscosity corresponds to relaxation times of the order of minutes and it exceeds the laboratory
time scales [3,38]. The molecular rotation spreads over more than
14 decades in time for six probes embedded into oterphenyl [62].
This is quite long compared to picosecond and nanosecond rotation
times observed above Tm.
Viscosity is directly related to structure and bonding of the substances. The rate of structural changes in supercooled liquids could
be observed by monitoring their viscosity as a function of temperature and pressure. On the basis of temperature dependence of
viscosity, supercooled liquids and glasses have been classified as
strong and fragile [63] [Fig-3].

where t is the site dependent relaxation time and b is a measure of


the deviation from Arrhenius behavior. When ~ 1 it is strong liquid
whereas < 1 indicates the fragile liquid.

Fig. 3- Schematic representation of strong and fragile liquids


This classification is associated with the degree of departure from
Arrhenius behavior. Essentially, these descriptive terms reflect the
different types of chemical bonding in the glass formers and signify
the sensitivity of the liquid structure to change of temperature. For
instance, SiO2 that possesses strong covalent bond network structure and exhibits Arrhenius viscous behavior is classified as strong
liquid. On the other hand, o-terphenyl molecules linked by weak van

t
(t ) exp

; 0 1

(1)

Strong Liquids
Viscosity (and relaxation) of strong liquids exhibit Arrhenius behavior for a wide range of temperatures. They are stabilized by three
dimensional network structures linked by covalent bonds, like SiO 2,
B2O3, P2O5 and GeO2. The ease with which these network forming
compounds turn to glasses is a result of their extremely high viscosity at Tm. These liquids show an exponential response to perturbations from an equilibrium state, i.e., their relaxation process can be
described by a single relaxation time rather than a distribution of
relaxation time.
[Fig-4] shows typical viscous behavior of strong and fragile liquids.
Strong liquids have a builtin resistance to structural change, and
their vibrational spectra and radial distribution function show little
reorganization despite wide variations of temperature. In strong
liquids, the temperature dependence of viscosity (or relaxation) is
low. Strong liquids can be converted to fragile ones by loosening
their packing [65]. This is brought about by the weakening network
by the incorporation of impurities.
Fragile Liquids
These materials are characterized by non-directional Coulombic
forces or by van der Waals interaction. The viscosity of these liquids
varies in a strongly nonArrhenius fashion over a wide interval of
temperature. The glassy structures of these liquids are very unstable and have high configurational degeneracy. They teeter on the
brink of collapse at Tg and with little provocation from thermal excitation, reorganize to distribution of structures that fluctuate over
different structural arrangements and coordination states. The relaxation processes of these materials follow non-exponential pattern, often according to KWW equation [Eq-1].
In a strong liquid, such as SiO2, the Cp values of the liquid and solid
phases are nearly the same [66]. Hence, the value of Tk is almost
indistinguishable from T = 0 K. On the other hand, in fragile liquids,
the difference between the liquid and crystalline Cp values is relatively large so that Tk often falls not far below Tg. The wide difference in Cp values between liquid and crystalline states of a fragile
liquid reflects significant entropy difference between these states.
Hence, one expects that the entropy catastrophe would be more
likely to occur in fragile liquids than in strong liquids. The fragile
liquids have, therefore, been the system of choice for the experimentalists seeking to investigate the importance of the Kauzmann
argument and the nature of the glass transition in general.

World Research Journal of Chemistry


|| Author Galley Proof || Author Galley Proof ||
|| Bioinfo Publications ||

Sivaprakasam K. (2014) Molecular, Kinetic and Thermodynamic Features of Supercooled Liquids - Glass Transition and Kauzmann Paradox.
World Research Journal of Chemistry, Author Galley Proof || Author Galley Proof || Author Galley Proof.

The theories that explain the viscosity - temperature relationship


have relevance in understanding the glass transition too. For example, the concept of free volume is useful in comprehending both the
viscosity and glass transition.
Viscosity - Temperature Relationship
The study of temperature dependence of viscosity of supercooled
liquids and glasses facilitates the understanding of the mechanism
of their flow process. Based on the degree of curvature of ln ()temperature curve, they are classified into two types, (i) Arrhenius
and (ii) non-Arrhenius.
Arrhenius Behavior and Eyrings Theory of Viscosity
Viscous flow is a simple thermally activated process [67]. The event
of flow process can be discretized as follows. A molecule is pictured
as vibrating about an equilibrium position until the following two
events occur.
Creation of an empty site into which the molecule can jump.
The molecule attains sufficient energy to overcome the attractive forces holding it to the neighbors
The probability of a transition or jump from one site to another is
given by Pj = Pe X Pv, where Pe is the probability of attaining sufficient energy to break interaction and Pv is the probability that there
is sufficient local free volume for a jump to occur. Hence, the jump#
ing frequency of a flow unit depends on the H and the size of
molecular/kinetic units that are determined by the specific structure
of the molecule and temperature. If the hole creation necessitates
alteration of equilibrium positions of many molecules flow becomes
a cooperative process.
The temperature dependence of viscosity can be represented by an
exponential function in the first approximation [Eq-2].
H#

A exp
(2)

RT
where A is the pre-exponential factor and H# not only functions as
a quantitative measure of the energy necessary for the flow to take
H #
place, it also reveals the intermolecular forces. For example, 010
reveals the degree of association in aqueous systems [68].
Eyrings analysis of the flow behavior was based on Anrades approach and he consolidated it further with the proposal of a lattice
model for understanding flow in liquid state. According to significant
structure theory of liquid by Eyring, liquid has a semicrystalline lattice in which portion of the molecules have solid-like degrees of
freedom and the rest of them possess gas-like freedom [69]. The
dependence of viscosity of liquids on temperature was treated by
Eyring within the framework of the theory of absolute reaction rate,
the basic propositions of this theory are extended to diffusion and
flow processes. Eyring and Frenkel have justified Eq-3 on the basis
of absolute reaction rate theory [70]. Eyrings absolute reaction rate
theory of viscosity facilitates an understanding of the mechanism of
the flow from physicochemical point of view. Although Eyrings model is built on the basis of the similarity of the liquid to a solid, it uses
the statistical mechanical approach as applied to gas phase to develop the theory. Eyring derived an expression linking viscosity to
free energy of activation for viscous flow (G # ).

G#
Nh

exp

Vm
RT

(3)

where N, h and Vm are Avogadro Number, Plancks constant and

molar volume respectively. Since


G# H# TS#

(4)

where S # denotes entropy of activation for viscous flow. Combining


Eq-3 and Eq-4.

H#
S #
Nh
exp
exp

R
Vm
RT

(5)

It has been widely supposed that most viscosity data can be satisfactorily represented by Eq-5 with single activation energy. This
impression probably has arisen because of the relatively poor accessibility of low temperature viscosity behavior to investigation,
due to crystallization and experimental difficulties. Eyrings theory
seems to be more suited to the high temperature behavior of glasses and liquids [71]. However, spherical molecules (liquefied inert
gases, liquid metals and molten salts which contain monatomic
ions), which can rotate many times about at least two axes during
the time between such translational jump obey Arrhenius equation
over normal liquid range [72]. In the case of other molecules, the H#
is far from constant; it increases continuously and at an accelerating
rate with decreasing temperature. For example, for simple molecular substances H#1shows temperature dependent variation; it increases with decrease in temperature [73]. The Arrhenius equation
accurately describes temperature variation of viscosity of many
liquids of low viscosity. However, it fails at temperatures close to
Tm, and in supercooled and associated liquids [74]. The inadequacy
of the Arrhenius equation has been shown for different systems
over extended temperatures [75,76].
Non-Arrhenius Behavior and Cooperative Motion
At low temperatures, when the supercooled liquid becomes highly
viscous, the dependence of viscosity increases strongly with decreasing temperature than indicated by an exponential function [Eq5] [77]. Complex molecules, such as long chain alkanes and alkyl
benzenes, show a transition from Arrhenius to non-Arrhenius behavior that is caused by restriction of molecular rotation. A molecule
whose rotation is restricted will be unable to align itself into a favorable orientation to squeeze past its neighbors. This effect is prominently seen in the transition zone just above the Tg.
Kinetic phenomena in supercooled liquids are known to be highly
cooperative [78]. A molecule cannot change its position independent of its neighbors implying that motion of a single molecule should
be considered together with the corresponding changes in its sur#
roundings and hence the H
is not representative of the hopping
of a single molecule but that of a cooperative region. These cooperative rearranging regions or structural units gradually expand in
size as temperature is lowered and at a critical temperature, T k,
identified as Kauzmann temperature configuration entropy vanishes
[36]. The entire molecule freezes in a single conformation. The
cooperative phenomenon becomes clear when one considers the
#
#
increase of H
with decrease of temperature, attaining value H
5Hv as temperature approaches Tg [79,80]. The curvilinear behavior of log vs. 1/T is also a reflection of the change in structural
configuration of the flow unit that accompanies lowering of temperature in the supercooled liquid. The continually changing cooperative
rearranging regions causes a significant change in correlation times
with decrease of temperature. At high temperatures, T >> T g, the
correlation times significantly differ according to size of the molecules. But at low temperatures similar correlation times are observed owing to cooperative motion. Hence, supercooling a liquid is

World Research Journal of Chemistry


|| Author Galley Proof || Author Galley Proof ||
|| Bioinfo Publications ||

Sivaprakasam K. (2014) Molecular, Kinetic and Thermodynamic Features of Supercooled Liquids - Glass Transition and Kauzmann Paradox.
World Research Journal of Chemistry, Author Galley Proof || Author Galley Proof || Author Galley Proof.

accompanied by an increase of cooperative motion and the identity


of individual molecules is partially lost. Due to the cooperative dynamics around the Tg complex motional behavior of the molecules
takes place resulting in nonexponential relaxation.
Free Volume Approach
Another approach to the theory of temperature dependence of viscosity is associated with the free volume concept. Batschinski first
advanced an idea that fluidity of liquids is due to the presence of
free volume [81]. In this model, a molecule is assumed to move
freely within the cage formed by its neighbors, unless its kinetic
energy exceeds critical energy when it may push its neighbors
aside and escape. Molecular transport occurs only when free volume redistributes itself locally to reach the volume of a flow unit. It
was observed that for certain liquids the temperature dependence
of viscosity almost disappears when the volume is kept constant.
Based on this fact Batchinski treated viscosity as a function of "free
volume"
Vf = V Vo
(6)
where Vo is the molecular volume at absolute zero. MacLeod [82]
proposed that could be related to Vf by a linear function
= k Vf

(7)

If Vo is independent of pressure then would a function of V alone


= f (V)
(8)
Although the simplicity of this equation is appealing, it at best qualitatively describes the temperature dependence of viscosity. For
many substances, like polymers, this formula is nothing more than a
crude approximation. Careful measurements of viscosity reveals
demonstrated that this relationship is a limiting condition that holds
only for simple molecular substances (it fails in the case of associated and high molecular weight molecules) for a short interval of temperature when their specific volume is not too small [83]. For all the
other substances investigated, the fluidity increases at a substantial
rate with temperature at constant volume [83]. Hence, viscosity can
be related to free volume by:
= f (Vf)
Later, Doolittle [84] proposed a relationship:
bVo

Vf

A exp

(9)

(10)

This exponential relationship is a better representation of the viscosity-temperature behavior than the linear one (eq. 9). Cohen and
Turnbull [85] showed that this exponential relation gives a fairly
good representation of the fluidity behavior of simple molecular
substances over a wide temperature range. Later the formation of
glass transition was associated with different materials converging
with the same fractional free volume [86].
Williams, Landel and Ferry had also shown empirical relation (WLF
equation, Eq-11 and Eq-12) that describes satisfactorily the fluidity
data of many substances in the glass transition region [87]. The
WLF is one of the well-known empirical equations which approximately describe the universal nonArrhenius effect of temperature
on viscosity and relaxation times in polymers.
(T )

log aT log
*
(11)
(T )
log aT

C1 (T T * )
T ( T * C2 )

(12)

aT = shift factor, = relaxation time


T* = reference temperature and C1 & C2 are constants
The ratio aT is a reduced variable making the widely different relaxation times observed in different systems over a relatively wide
range of temperatures conform alike. For organic glasses the constant C1 ~ 17 and C2 ~ 52. Both constants are related to the free
volume of the glass at Tg and above Tg as follows
C1 = (1/2.303) Vf
(13)
C2 = (Vf / )
(14)
An important feature of glass forming systems is that they exhibit
similar transport mechanism in the glass transition range irrespective of the chemical structure. The functional form of the temperature dependence of relaxation times (or the viscosity) for supercooled liquids are often described, at least approximately, by the
VTF equation. The WLF equations [Eq-11 & Eq-12], often used to
describe viscosity or relaxation times in polymers, are mathematically equivalent to the VTF equations [Eq-15 & Eq-16]. The viscosity behavior of glass and supercooled liquid can be more accurately
represented over a wide range of temperature by VTF equation
than by Arrhenius equation. Fragile liquids which tend to exhibit
continually changing activation energy over a wide temperature
range obey VTF equation.
B
(15)

exp
o

T T
o

As t h [Eq-15] can be written as


B
ln A
T To

(16)

where To = VTF temperature. When To is zero, the VTF equation


becomes Arrhenius equation. When To > 0, the temperature dependence is nonArrhenius and the relaxation time is predicted to
become infinite at To. To not only linearize the otherwise nonlinear ln
() vs. (1/T) plots, but has physical significance as well. It functions
as a convergent temperature for the kinetic and thermodynamic
approach to glass transition. The To tends to be equal to ideal glass
transition temperature (IGT), normally falls 50 - 100 K below Tg [88].
Macedo and Litovitz discussed [89] the drawback of free volume
model that concentrates on the importance of holes and neglects
H # They developed a hybrid equation that combines free volthe 111.
ume and Eyring equations. This approach rectified one of the drawbacks of the free volume model. However, this model still predicts
too much curvature for the viscosity behavior at low temperatures.
Thermodynamic Glass Transition
Glass transition, fluidity-superfluidity, conducting-superconducting
transitions in which the second partial derivative of free energy
changes abruptly is second order phase transition.
Glass transition temperature (Tg) can be defined in different ways.
The most unambiguous definition of it is in terms of discontinuity in
a, Cp [Fig-4] [92].

1
V

2G
2G
1 2G

; T
Cp ;

2
2
V PT
P
T

(17)

As seen in Fig-4, transition from supercooled liquid state to glassy


phase is not unique, but occurs rather over a range of temperatures
called the "glass transformation range". Tg varies with cooling rates.
A lower cooling rate allows the sample to stay in equilibrium until
low temperatures. In general, the dependence of Tg on cooling rate

World Research Journal of Chemistry


|| Author Galley Proof || Author Galley Proof ||
|| Bioinfo Publications ||

Sivaprakasam K. (2014) Molecular, Kinetic and Thermodynamic Features of Supercooled Liquids - Glass Transition and Kauzmann Paradox.
World Research Journal of Chemistry, Author Galley Proof || Author Galley Proof || Author Galley Proof.

is relatively low; an order of magnitude change in cooling rate


changes Tg by only 3 5 K [90]. In spite of the dependence of Tg on
experimental conditions, it is still an important material property; It is
most useful in evaluating the mechanical behavior of a polymeric
material.

Fig. 4- Variation in Cp as function of temperature


Thermodynamic Changes at Tg
Thermodynamic properties of a supercooled liquid can be considered as an extrapolation of the properties of a stable equilibrium
liquid phase. Supercooling a glass forming substance preserves its
liquid properties until it is frozen in a specific configuration. Above
Tg, any change in material property of a supercooled liquid has two
contributions: vibrational and configurational changes. Below T g, as
the configuration is largely frozen, only vibrations contribute to the
properties, resulting in either a break in slope or a discontinuity in
thermodynamic and spectroscopic properties. Hence, the temperature dependence of properties of a glass forming substance is low
below Tg. As the configurational contribution of a system cease
below Tg, all properties measured below Tg are called isostructural.
During configurational freezing of a supercooled liquid, Cp usually
changes markedly within a narrow range around Tg [Fig-4]. This
decrease reflects the loss of configurational freedom. More important, the value of Tg and the shape of Cp curve are found to be
very dependent on the kinetics of measurement. If equilibrium with
respect to some degree of freedom is lost during a cooling process,
then the total entropy associated with that degree of freedom cease
to contribute to the Cp according to:
S2 = k Cp
(18)
Thus, glass transition is always accompanied by a decrease of Cp.
This change may, in some cases, be very small, but the inequality
Cp (liquid) Cp (solid) = Cp 0
(19)
cannot be violated. The Cp is a quantitative measure of changes
in molecular ordering which occurs once the glassy structure has
been "unlocked" by sufficient increase of temperature. Supercooled
liquids with degenerative structures will be characterized by large
Cp values, while those with temperature resistant short range
orders will show relatively small Cp values.
Glass transition exhibits unique characteristics of both kinetic and
thermodynamic phenomena. The dependence of glass transition on
the cooling rate shows that the supercooled liquid - glass transition
is a relaxation process. On the other hand, the discontinuity in Cp, a
and b on passing through Tg indicates some analogy between glass

formation process and a second order phase transition. Attempts to


break this impasse and resolve the issue of nature of glass transition resulted in a compromise of kinetics and thermodynamics concepts. Glass transition was accepted as a kinetic phenomenon underlain by thermodynamic phase transition.
Thermodynamic Catastrophe and Kauzmann Paradox
As value of Tg depends on the time scale of observation it is pertinent to raise a question, although hypothetical. Is there an end to
the decrease in Tg as the time scale of an experiment is slowed?
How a supercooled liquid would be expected to behave at very low
temperatures if experimental measurements were made sufficiently
slow enough to permit the liquid structure to be in thermodynamic
equilibrium with the surroundings (but, of course, not so slowly for
the spontaneous crystallization)? It is reasonable to expect a clue to
this behavior by a simple extrapolation to low temperatures of the
known properties of supercooled liquids above their glass transformation temperatures. When such an extrapolation is applied to
entropy vs. temperature curves of several substances (e.g., glucose
and lactic acid), a rather startling result was obtained [12]. Not very
far below Tg, but still far above absolute zero the extrapolated entropy of the supercooled liquid becomes less than that of the crystalline solid. The extrapolated enthalpy vs. temperature and specific
volume vs. temperature curves of supercooled liquid also show
similar tendencies; supercooled liquid appears to strive for a lower
enthalpy and entropy and a smaller volume than crystal at temperatures well above absolute zero. The temperature at which entropy
of supercooled liquid attains the same value as that of the crystal is
termed as Kauzmann temperature. This temperature also marks the
lowest limit for the existence of supercooled liquid. Below this temperature the extrapolated entropy of the supercooled liquid would
have lower value than that of the crystal. This strange situation was
termed as a paradox by Kauzmann as it is a violation of the third
law of thermodynamics, hence known as Kauzmann paradox. The
third law of thermodynamics would be violated if the supercooled
liquid entropy continues to decrease much below Tk without change
in slope, since then it would become negative well above absolute
zero. This line of reasoning by Kauzmann effectively puts a limit on
the value of Tg since glass transition must intervene at or above Tk
in order for the entropy of the glass to remain positive. Therefore,
Kauzmann paradox does not happen in reality due to the occurrence of glass transition at or above Tk.
Theories of Glass Transition
A complete theoretical understanding of the glass transition is not
yet available. However, there are three theories of glass transition:
the freevolume theory, the kinetic theory, and the thermodynamic
theory. These three theories explain certain features of the same
phenomenon.
The FreeVolume Theory
As developed by Eyring and others, molecular motion in the liquid
state depends on the presence of holes. The nature of glass transition can also be understood in terms of free volume. Liquids, whether polymeric or not, possess free volume, which may be present
either as holes of the order of molecular dimensions or smaller
voids due to packing irregularities. For the motion of polymer
chains, more than one "hole may be required in the same locality
as cooperative motions are involved. Thus for a polymer segment to
move from its position to an adjacent site, a critical void volume

World Research Journal of Chemistry


|| Author Galley Proof || Author Galley Proof ||
|| Bioinfo Publications ||

Sivaprakasam K. (2014) Molecular, Kinetic and Thermodynamic Features of Supercooled Liquids - Glass Transition and Kauzmann Paradox.
World Research Journal of Chemistry, Author Galley Proof || Author Galley Proof || Author Galley Proof.

must exist before a segment can jump. The variation of free volume
of a liquid as a function of temperature could be determined by
monitoring their a value that represents primarily the creation of
additional free volume with rising temperature. This theory provides
a relationship between coefficients of thermal expansion below and
above Tg with free volume. According to this theory glass transition
results from the reduction of molecular mobility as the temperature
falls, slowing the collapse of free volume. In other words, mobility at
any temperature depends on the free volume remaining.
The free volume of a liquid cannot be directly measured, but must
be deduced from the measurements of specific volume. The occupied volume has been obtained by many ways. It has been calculated from the van der Waals radii of the atoms. Others have used the
crystalline volume at absolute zero (eq. 6). A third approach is to
take the difference between the total volume and the fluctuation
volume. Doolittle used the extrapolated specific volume of liquids at
absolute zero as the occupied volume (Vo).
Tg as an isoFree Volume State
The free volume theory states that glass transition is characterized
by an iso-free volume state. Fox and Flory examined the relation
between glass transition and free volume of polystyrene as a function of molecular weight and relaxation time [91]. For limiting molecular weight, it was found that the free volume could be expressed
above Tg as:
Vf = K + (aR aG ) R
(20)
Where K is related to free volume at absolute zero, aR and aG represent the volume expansion coefficient in the rubbery and glassy
states respectively and R is the universal gas constant. It was also
established that below Tg the same specific volume temperature
relationship holds for all polystyrenes, independent of molecular
weight. From this study it was established that (i) below Tg the local
conformational arrangement of the polymer segments is independent of both molecular weight and temperature and (ii) the glass
transition temperature is an isofree volume state.
To form a glass, the free volume must be reduced to the value

V f (Tg ) 0.025Vl 0 V0

(21)

where is the volume of the compound and Vo is the volume at


absolute zero.
Vl0

The Kinetic Theory


The kinetic theory of the glass transition considers the molecular
and macroscopic response within a varying time frame. According
to this theory, there is no thermodynamic glass transition, the phenomenon is purely kinetic because changing the time scale of the
experiment can alter Tg. The glass transition of supercooled liquids
occurs when the rate of molecular (structural, configurational) rearrangements becomes the same order of magnitude as the rate of
cooling. This theory treats the transition of a kinetic unit from one
state to another as an individual, independent process.
The effect of rate of change of temperature (dq) on Tg is related to
the activation energy for the relaxation (h).
d ln | q |
h

(22)
d Tg
R Tr2
where Tr represents the temperature in the middle of the transition
region.
The major drawback of this theory is that it does not take into ac-

count the cooperative phenomenon of segmental motion when


glass transition takes place.
Thermodynamic Theory
The Gibbs-DiMarzio theory (GD) is the first physical theory of the
glass transition based on the use of the statistical mechanics of
polymers. This theory rests on the application of a lattice model,
similar to the well-known Flory-Huggins lattice model. The thermodynamic properties of a polymer system are determined by considering a canonical ensemble of conformations. It states that the molecular mechanism of glass transition is governed by two factors (i)
the energy difference between the gauche- and trans-isomers (E
= E2 - E1) and (ii) energy of hole formation (i.e., the energy of intermolecular interaction). At high temperatures a molecule would have
numerous modes of packing in the amorphous phase. When a polymer is cooled, the macromolecular conformation energy falls and
transforms into glass. Two processes accompany this transformation: (i) low energy molecular conformations start to predominate
which makes the chain rigid and (ii) the volume decreases (the free
volume too decreases). E characterizes the stiffness of the polymer chain. The value of the E for stiff chains is greater than that
for flexible chains. This theory states that despite the observed
glass transition being a kinetic phenomenon, the underlying true
phase transition undergoes discontinuity in equilibrium properties
according to Ehrenfest classifications. As per this theory, in infinitely
slow experiments a glassy phase will eventually emerge whose
entropy is negligibly higher than that of the crystal. The configurational entropy is related to the viscosity and it is postulated by Gibbs
and DiMarzio that the configurational entropy is zero at the thermodynamic glass transition (T2). The GD theory adequately explains
the experimental observation of Tg variation as a function of molecular weight, cross-linking, tacticity, deformation, and compositional
dependence (in the case of blends or copolymers).
Unified Treatment
The Gibbs-DiMarzio theory has exerted a strong impact on investigations on the nature and molecular mechanism of glass transition.
Adam and Gibbs have utilized the ideas of Gibbs-DiMarzio theory
for describing the temperature dependence of the relaxation time
upon glass transition [92]. This theory introduced the concept of a
"cooperatively rearranging region", the smallest region capable of
conformational changes independently. Adam and Gibbs attempted
to unify the kinetic and the thermodynamic theories by linking viscosity with the cooperatively rearranging regions. This region is a
subsystem of the sample in which enthalpy fluctuations are likely to
occur. The size of this region expands gradually with the lowering of
temperature and finally at T2, this region becomes equal to the size
of the sample; the entire molecule would acquire a single conformation. The variation of t as a function of temperature is determined
by the change in the dimensions of the cooperatively rearranging
regions. It has been shown that the size of these regions may be
described in terms of the Sc.
An expression used to describe the non-Arrhenius temperature
dependence of the relaxation data by Adam-Gibbs equation (AG)
[Eq-23 and Eq-24] reflects the entropy catastrophe as pointed by
Kauzmann [12]. The Sc decreases with temperature and at T = Tk,
Sc becomes unity and hence the viscosity attains singularity. Conceptually, this model resolves the Kauzmann paradox by showing
that the structural relaxation process diverges at Tk, kinetically arresting a phase transition so that Sc equals to unity point is never

World Research Journal of Chemistry


|| Author Galley Proof || Author Galley Proof ||
|| Bioinfo Publications ||

Sivaprakasam K. (2014) Molecular, Kinetic and Thermodynamic Features of Supercooled Liquids - Glass Transition and Kauzmann Paradox.
World Research Journal of Chemistry, Author Galley Proof || Author Galley Proof || Author Galley Proof.

reached. AG equation predicts an ideal glass transition at T k and


argues that the smallness of Sc at metastable equilibrium for a supercooled liquid reflects its sluggish behavior. The AG expression
linking relaxation times with configurational entropy gives an excellent account of the non-linearity observed in enthalpy relaxation of
amorphous polymeric, inorganic and simple molecular materials
near and below Tg
E
A exp 1
(23)
TS c
Since , Eq-23 can be written in terms of viscosity as

E
(24)
h A exp 2
TS c
where A is a temperature independent constant, E1 and E2 are the
energy required for the reorientation and viscous flow of a cooperative region respectively. AG equation enables one to estimate quantitatively the variation of the Tg of the amorphous polymers with a
change of crystallinity. Increasing the dsgree of crystallinity leads to
decrease in the value of Sc. From Eq-23 it is clear that decrease of
Sc will lead to exponential increase in t value resulting in an increase in Tg value. Thus the AG theory provides a satisfactory explanation of the change of relaxation time of segmental motion and
the Tg variation with the change in crystallinity in polymers.
Conclusions
The study of glass transition has received wide attention and many
theories were developed to explain this phenomenon. In spite of the
extensive investigations on the glass transition, the nature of this
transition remains contentious; there is no indication for the emergence of a concurrent view. Nevertheless, these theories explain
certain features of glass transition. Hence, an insight into glass
transition necessitates integration of kinetic, thermodynamic and
free volume approaches.
One of the fundamental questions concerning glass formation is to
know the factors that cause a material to crystallize or lead to the
appearance of the amorphous state. The ability to form glass is not
solely a material property. Kinetics can always control the outcome
of cooling of liquid state. Complexity as a result of entry of kinetics,
with an array of possibilities, is a serious deterrent to the understanding of the glass forming tendency of molecules.
Tk, though not an alternative to Tg, serves as an ideal glass transition temperature. Unlike Tg, Tk is not dependent on kinetic factors.
References
[1] Debenedetti P.G. (1996) Metastable Liquids: Concepts and
Principles, Princeton N.J., Princeton University Press.
[2] Angell C.A. (1995) Proc. Natl. Acad. Sci. USA, 92(15), 66756682.
[3] Karmakar S., Dasgupta C., Sastry S. (2009) Proc. Nat. Acad.
Sci. USA, 106(10), 3675-3679.
[4] Kawasaki T., & Tanaka H. (2010) J. Phys. Condens. Matter, 22
(23), 232102.
[5] Anderson P.W. (1995) Science, 267(5204), 1615-1616.
[6] Kishore K. & Shobha H.K. (1992) J. Phys. Chem., 96(20), 81618168.
[7] Kannan S. & Kishore K. (1999) J. Chem. Eng. Data, 44(4), 649655.
[8] Kishore K., Bharat S., & Kannan S. (1996) J. Chem. Phys., 105,

11364.
[9] Kishore K. & Mattamal G.J. (1986) J. Polym. Sci. Polym. Lett.,
24, 53-55.
[10] Kishore K., Shobha H.K. & Mattamal G.J. (1990) J. Phys.
Chem., 94(4), 1642-1648.
[11] Shobha H.K., Kishore K. (1992) Macromolecules, 25(25), 67656769.
[12] Kauzmann W. (1948) Chem. Rev., 43(2), 219-256.
[13] Stevenson J.D., Wolynes P.G. (2009) Nat. Phys., 6(1), 62-68.
[14] Zondervan R., Xia T., van der Meer H., Storm C., Kulzer F., van
Saarloos W. & Orrit M. (2008) Proc. Natl. Acad. Sci. USA, 105
(13), 4993-4998.
[15] Gutzow I.S. & Schmelzer J. (2013) The vitreous state: thermodynamics, structure, rheology, and crystallization, Springer,
Berlin.
[16] Yang Z., Fujii Y., Lee F.K., Lam C.H. & Tsui O.K.C. (2010) Science, 328(5986), 1676-1679.
[17] Suryanarayana C. & Inoue A. (2011) Bulk Metallic Glasses,
CRC Press, Boca Raton, FL.
[18] Leuzzi L. & Nieuwenhuizen T.M. (2008) Thermodynamics of the
glassy state, Taylor & Francis, New York.
[19] Ojovan M.I. (2013) J. Non-Cryst. Solids, 382, 79-86.
[20] Ge S., Pu Y., Zhang, W., Rafailovich, M., Sokolov, J., Buenviaje, C., Buckmaster R., Overney R.M. (2000) Phys. Rev. Lett.,
85(11), 2340-2343.
[21] Hsiung L.M., Cai W. & Wadley H.N.G. (1992) Acta Metall. Mater., 40(11), 3035-3049.
[22] Yavari A.R. (2006) Nature, 439(7075), 405-406.
[23] Thorpe M.F. & Tich L. (2001) Properties and Applications of
Amorphous Materials, Springer.
[24] Paeng K., Powell C.T., Yu L. & Ediger M.D. (2012) J. Phys.
Chem. Lett., 3(18), 2562-2567.
[25] Richert R. (2011) Annu. Rev. Phys. Chem., 62(1), 65-84.
[26] Jiang M.Q., Wilde G., Gao J.B., Dai L.H. (2013) Scr. Mater., 69
(10), 760-763.
[27] Yin H., Napolitano S. & Schnhals A. (2012) Macromolecules,
45(3), 1652-1662.
[28] Tress M., Erber M., Mapesa E.U., Huth H., Muller J., Serghei
A., Schick C., Eichhorn K.J., Voit B., Kremer F. (2010) Macromolecules, 43(23), 9937-9944.
[29] Stachurski Z.H. (2011) Materials, 4(9), 1564-1598.
[30] Stillinger F.H. & Debenedetti P.G. (2013) Annu. Rev. Condens.
Matter Phys., 4(1), 263-285.
[31] Gibbs J.H., DiMarzio E.A. (2004) The Journal of Chemical
Physics, 28(3), 373-383.
[32] Dimarzio E.A., & Gibbs, J. H. (1963) J. Polym. Sci. Part A, 1(4),
1417.
[33] Cohen M.H. & Turnbull D. (2004) The Journal of Chemical
Physics, 31(5), 1164-1169.
[34] Cohen M.H., Turnbull D. (1961) Nature, 189(4759), 131-132.
[35] Thorpe M.F., Mitkova M.I. (1997) Amorphous Insulators and
Semiconductors, Springer.
[36] Berthier L., Biroli G., Bouchaud J.P., Cipelletti L. & Saarloos W.
van. (2011) Dynamical Heterogeneities in Glasses, Colloids,

World Research Journal of Chemistry


|| Author Galley Proof || Author Galley Proof ||
|| Bioinfo Publications ||

Sivaprakasam K. (2014) Molecular, Kinetic and Thermodynamic Features of Supercooled Liquids - Glass Transition and Kauzmann Paradox.
World Research Journal of Chemistry, Author Galley Proof || Author Galley Proof || Author Galley Proof.

and Granular Media, Oxford University Press.


[37] Robertson R.E. (1975) Annu. Rev. Mater. Sci., 5(1), 73-97.
[38] Rault J. (2012) Eur. Phys. J., E, 35(4), 1-28.
[39] Fegley B. (2013) Practical Chemical Thermodynamics for Geoscientists, Academic Press.
[40] Kitamura T. (2012) Liquid Glass Transition: A Unified Theory
From the Two Band Model, Newnes.
[41] Jain R., Bhandari D., Saxena N.S., Sharma S.K. & Tripathi A.
(2001) Bull. Mater. Sci., 24(1), 27-33.
[42] Somiya S. (2013) Handbook of Advanced Ceramics: Materials,
Applications, Processing, and Properties, Academic Press.
[43] Movahedi B., Enayati M.H., & Wong C.C. (2010) Mater. Lett.,
64(9), 1055-1058.
[44] Kurakevych O.O., Le Godec Y., Hammouda T., Goujon C.
(2012) High Press. Res., 32(1), 30-38.
[45] England J., Phaneuf M.W., Laquerre A., Smith A. & Gwilliam R.
(2012) Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At., 272, 409-413.
[46] Paes S.S., Sun S., MacNaughtan W., Ibbett R., Ganster J.,
Foster T.J. & Mitchell J.R. (2010) Cellulose, 17(4), 693-709.
[47] Sharma K., Ponomarev M.V., Verheijen M.A., Kunz O.,
Tichelaar F.D., Van de Sanden M.C.M. & Creatore M. (2012) J.
Appl. Phys., 111(10), 103510-103510-5.
[48] Chryssikos G.D., Patsis A.P., Kamitsos E.I., Kapoutsis J.A.,
Karakassides M.A., Trapalis C., Mylonas E. & Kordas G. (1995)
J. Mater. Sci. Lett., 14(4), 268-270.
[49] Telford M. (2004) Mater. Today, 7(3), 36-43.
[50] Carter C.B. & Norton M.G. (2013) Ceram. Mater., Springer New
York, 389-409.
[51] Patel A.T., Pratap A. (2012) J. Therm. Anal. Calorim., 110(2),
567-571.
[52] Freed K.F. (2011) Acc. Chem. Res., 44(3), 194-203.
[53] Wang L.M., Tian Y., Liu R., Richert R. (2008) J. Chem. Phys.,
128(8), 084503.
[54] Ping W., Paraska D., Baker R., Harrowell P., Angell C.A. (2011)
J. Phys. Chem., B, 115(16), 4696-4702.
[55] Bruna P., Baldi G., Pineda E., Serrano J., Duarte M.J., Crespo
D., & Monaco G. (2011) J. Alloys Compd., 509(1), S95-S98.
[56] Mukherjee S., Schroers J., Zhou Z., Johnson W.L. & Rhim W.K.
(2004) Acta Mater., 52(12), 3689-3695.
[57] Wales D. (2003) Energy Landscapes: Applications to Clusters,
Biomolecules and Glasses, Cambridge University Press.
[58] Leone G., Giovanella U., Bertini F., Hoseinkhani S., Porzio W.,
Ricci G., Bottaa C., Galeotti F. (2013) J. Mater. Chem., 100(1),
6585-6593.
[59] Shobha H.K. & Kishore K. (1992) J. Chem. Eng. Data, 37(4),
371-376.
[60] Tucker N. & Lindsey K. (2002) An Introduction to Automotive
Composites, iSmithers Rapra Publishing.
[61] Weeks E.R., Crocker J.C., Levitt A.C., Schofield A. & Weitz D.
A. (2000) Science, 287(5453), 627-631.
[62] Cicerone M., Blackburn F.R. & Ediger M.D. (1995) J. Chem.
Phys., 102(1), 471-479.
[63] Martinez L.M., Angell C.A. (2001) Nature, 410(6829), 663-667.

[64] Idzikowski B., Miglierini M. & vec P. (2005) Properties and


applications of nanocrystalline alloys from amorphous precursors, Springer.
[65] Wolynes P.G., Lubchenko V. (2012) Structural Glasses and
Supercooled Liquids: Theory, Experiment, and Applications.
John Wiley & Sons.
[66] Angell C.A. (2008) Science, 319(5863), 582-587.
[67] Avramov I. (2007) Phys. Chem. Glas.-Eur. J. Glass Sci. Technol. Part B, 48(1), 61-63.
[68] Marczak W., Adamczyk N. & niak M. (2012) Int. J. Thermophys., 33(4), 680-691.
[69] Eyring H. & Ree T. (1961) Proc. Natl. Acad. Sci. USA., 47(4),
526.
[70] Ferrer M.L., Lawrence C., Demirjian B.G., Kivelson D., AlbaSimionesco C. & Tarjus G. (1998) J. Chem. Phys., 109(18),
8010-8015.
[71] Sanditov D.S., Munkueva S.B., Mashanov A.A. & Sanditov B.
D. (2012) Glass Phys. Chem., 38(4), 379-385.
[72] Davies D.B. & Matheson A.J. (1966) J. Chem. Phys., 45, 1030.
[73] Sorgenfrei S., Chiu C.Y., Gonzalez Jr.R.L., Yu Y.J., Kim P.,
Nuckolls C. & Shepard K.L. (2011) Nature Nanotechnology, 6
(2), 126-132.
[74] Doster W. (2008) Eur. Biophys. J. EBJ, 37(5), 591-602.
[75] Isayev A.I. (1987) Injection and Compression Molding Fundamentals, CRC Press.
[76] Smith I.W.M. (2008) Chem. Soc. Rev., 37(4), 812-826.
[77] Kincs J., Martin S.W. (1996) Phys. Rev. Lett., 76(1), 70-73.
[78] Forrest J.A. & Mattsson J. (2000) Phys. Rev. E, 61(1), R53R56.
[79] Wei J., Denn M.M., Seinfeld J.H., Chakraborty A., Ying J., Peppas N. & Stephanopoulos G. (2001) Molecular Modeling and
Theory in Chemical Engineering, Academic Press.
[80] Totten G.E. (1999) Handbook of Hydraulic Fluid Technology,
CRC Press.
[81] Latini G., Grifoni R.C., Passerini G. (2006) Transport Properties
of Organic Liquids, WIT Press.
[82] Briant J., Denis J., Parc G. (1989) Rheological Properties of
Lubricants, TECHNIP ed.
[83] Barlow A.J., Lamb J. & Matheson A.J. (1966) Proc. R Soc.
Lond. Math. Phys. Sci., 292(1430), 322-342.
[84] Gedde U.W. (1995) Polymer Physics, Chapman & Hall, London.
[85] Rao K.J. (2002) Structural Chemistry of Glasses, Elsevier.
[86] Wypych G. (2013) Handbook of Plasticizers, William Andrew.
[87] Utracki L.A. & Jamieson A.M. (2011) Polymer Physics: From
Suspensions to Nanocomposites and Beyond, John Wiley &
Sons.
[88] Harris K.R., Kanakubo M. & Woolf L.A. (2007) J. Chem. Eng.
Data, 52(3), 1080-1085.
[89] Bourhis E.L. (2008) Glass: Mechanics and Technology, John
Wiley & Sons.
[90] Thomas L.C. (2001) Am. Lab., 33(1), 26-31.
[91] Fox T.G. & Flory P.J. (1954) J. Polym. Sci., 14(75), 315-319.
[92] Hale A., Macosko C.W. & Bair H.E. (1991) Macromolecules, 24
(9), 2610-2621.

World Research Journal of Chemistry


|| Author Galley Proof || Author Galley Proof ||
|| Bioinfo Publications ||

10

Potrebbero piacerti anche