Download as pdf or txt
Download as pdf or txt
You are on page 1of 69

Glass transition

The glass–liquid transition, or glass


transition, is the gradual and reversible
transition in amorphous materials (or in
amorphous regions within semicrystalline
materials), from a hard and relatively
brittle "glassy" state into a viscous or
rubbery state as the temperature is
increased.[1] An amorphous solid that
exhibits a glass transition is called a glass.
The reverse transition, achieved by
supercooling a viscous liquid into the
glass state, is called vitrification.

The glass-transition temperature Tg of a


material characterizes the range of
temperatures over which this glass
transition occurs. It is always lower than
the melting temperature, Tm, of the
crystalline state of the material, if one
exists.

Hard plastics like polystyrene and


poly(methyl methacrylate) are used well
below their glass transition temperatures,
i.e., when they are in their glassy state.
Their Tg values are well above room
temperature, both at around 100 °C
(212 °F). Rubber elastomers like
polyisoprene and polyisobutylene are used
above their Tg, that is, in the rubbery state,
where they are soft and flexible.[2]

Despite the change in the physical


properties of a material through its glass
transition, the transition is not considered
a phase transition; rather it is a
phenomenon extending over a range of
temperature and defined by one of several
conventions.[3][4] Such conventions include
a constant cooling rate (20 kelvins per
minute (36 °F/min))[1] and a viscosity
threshold of 1012 Pa·s, among others.
Upon cooling or heating through this
glass-transition range, the material also
exhibits a smooth step in the thermal-
expansion coefficient and in the specific
heat, with the location of these effects
again being dependent on the history of
the material.[5] The question of whether
some phase transition underlies the glass
transition is a matter of continuing
research.[3][4][6] In a more recent model of
glass transition, the glass transition
temperature corresponds to the
temperature at which the largest openings
between the vibrating elements in the
liquid matrix become smaller than the
smallest cross-sections of the elements or
parts of them when the temperature is
decreasing. As a result of the fluctuating
input of thermal energy into the liquid
matrix, the harmonics of the oscillations
are constantly disturbed and temporary
cavities ("free volume") are created
between the elements, the number and
size of which depend on the temperature.
The glass transition temperature Tg0
defined in this way is a fixed material
constant of the disordered (non-
crystalline) state that is dependent only on
the pressure. As a result of the increasing
inertia of the molecular matrix when
approaching Tg0, the setting of the
thermal equilibrium is successively
delayed, so that the usual measuring
methods for determining the glass
transition temperature in principle deliver
Tg values that are too high. In principle, the
slower the temperature change rate is set
during the measurement, the closer the
measured Tg value Tg0 approaches.(Karl
Günter Sturm Microscopic-
Phenomenological Model of Glass
Transition I. Foundations of the model
(DOI: 10.13140/RG.2.2.19831.73121)

Introduction IUPAC
The glass transition of a definition
glass
liquid to a solid-like state
transition (in
may occur with either polymer

cooling or compression.[10] science):


Process in
The transition comprises a
which a
smooth increase in the
polymer
viscosity of a material by
melt
as much as 17 orders of changes on
magnitude within a cooling to a
temperature range of 500 polymer
K without any pronounced glass or a
polymer
change in material
glass
structure.[11] The
changes on
consequence of this
heating to a
dramatic increase is a polymer
glass exhibiting solid-like melt.[7]
mechanical properties on Note 1:
the timescale of practical Phenomena
observation. This occurring at

transition is in contrast to the glass


transition of
the freezing or
polymers are
crystallization transition,
still subject
which is a first-order phase
to ongoing
transition in the Ehrenfest scientific
classification and involves investigation
discontinuities in and debate.
thermodynamic and The glass
transition
dynamic properties such
presents
as volume, energy, and
features of a
viscosity. In many
second-
materials that normally order
undergo a freezing transition
transition, rapid cooling since
will avoid this phase thermal
transition and instead studies

result in a glass transition often


indicate that
at some lower
the molar
temperature. Other
Gibbs
materials, such as many
energies,
polymers, lack a well molar
defined crystalline state enthalpies,
and easily form glasses, and the
even upon very slow molar
volumes of
cooling or compression.
the two
The tendency for a
phases, i.e.,
material to form a glass
the melt and
while quenched is called the glass,
glass forming ability. This are equal,
ability depends on the while the
composition of the heat
material and can be capacity and

predicted by the rigidity the


expansivity
theory.[12]
are
Below the transition discontinuou
s. However,
temperature range, the
the glass
glassy structure does not
transition is
relax in accordance with
generally not
the cooling rate used. The regarded as
expansion coefficient for a
the glassy state is roughly thermodyna
equivalent to that of the mic

crystalline solid. If slower transition in


view of the
cooling rates are used, the
inherent
increased time for
difficulty in
structural relaxation (or reaching
intermolecular equilibrium

rearrangement) to occur in a polymer


glass or in a
may result in a higher
polymer
density glass product.
melt at
Similarly, by annealing (and
temperature
thus allowing for slow s close to
structural relaxation) the the glass-
glass structure in time transition
approaches an equilibrium temperature.

density corresponding to Note 2: In


the supercooled liquid at the case of
this same temperature. Tg polymers,
is located at the conformatio

intersection between the nal changes


of segments,
cooling curve (volume
typically
versus temperature) for
the glassy state and the consisting of

supercooled 10–20 main-


chain atoms,
liquid.[13][14][15][16][17]
become
The configuration of the infinitely
slow below
glass in this temperature
the glass
range changes slowly with
transition
time towards the
temperature.
equilibrium structure. The
principle of the Note 3: In a
partially
minimization of the Gibbs
crystalline
free energy provides the
polymer the
thermodynamic driving glass
force necessary for the transition
eventual change. It should occurs only
be noted here that at in the
somewhat higher amorphous

temperatures than Tg, the parts of the


material.
structure corresponding to
equilibrium at any Note 4: The
temperature is achieved definition is
quite rapidly. In contrast, at different

considerably lower from that in


ref.[8]
temperatures, the
configuration of the glass Note 5: The
remains sensibly stable commonly

over increasingly extended used term


“glass-
periods of time.
rubber
Thus, the liquid-glass transition”
for glass
transition is not a
transition is
transition between states
not
of thermodynamic recommend

equilibrium. It is widely ed.[9]

believed that the true


equilibrium state is always crystalline.
Glass is believed to exist in a kinetically
locked state, and its entropy, density, and
so on, depend on the thermal history.
Therefore, the glass transition is primarily
a dynamic phenomenon. Time and
temperature are interchangeable
quantities (to some extent) when dealing
with glasses, a fact often expressed in the
time–temperature superposition principle.
On cooling a liquid, internal degrees of
freedom successively fall out of
equilibrium. However, there is a
longstanding debate whether there is an
underlying second-order phase transition
in the hypothetical limit of infinitely long
relaxation times.[5][18][19][20]

Transition temperature Tg
This section needs additional citations for
verification. Learn more

Determination of Tg by dilatometry.
Measurement of Tg (the temperature at the point A) by
differential scanning calorimetry

Refer to the figure on the upper right


plotting the heat capacity as a function of
temperature. In this context, Tg is the
temperature corresponding to point A on
the curve. The linear sections below and
above Tg are colored green. Tg is the
temperature at the intersection of the red
regression lines.[21]
Different operational definitions of the
glass transition temperature Tg are in use,
and several of them are endorsed as
accepted scientific standards.
Nevertheless, all definitions are arbitrary,
and all yield different numeric results: at
best, values of Tg for a given substance
agree within a few kelvins. One definition
refers to the viscosity, fixing Tg at a value
of 1013 poise (or 1012 Pa·s). As evidenced
experimentally, this value is close to the
annealing point of many glasses.[22]

In contrast to viscosity, the thermal


expansion, heat capacity, shear modulus,
and many other properties of inorganic
glasses show a relatively sudden change
at the glass transition temperature. Any
such step or kink can be used to define Tg.
To make this definition reproducible, the
cooling or heating rate must be specified.

The most frequently used definition of Tg


uses the energy release on heating in
differential scanning calorimetry (DSC, see
figure). Typically, the sample is first cooled
with 10 K/min and then heated with that
same speed.

Yet another definition of Tg uses the kink


in dilatometry (a.k.a. thermal expansion).
Here, heating rates of 3–5 K/min (5.4–
9.0 °F/min) are common. Summarized
below are Tg values characteristic of
certain classes of materials.

Polymers
Material Tg (°C) Tg (°F) Commercial name

Tire rubber −70 −94[23]

Polyvinylidene fluoride (PVDF) −35 −31[24]

Polypropylene (PP atactic) −20 −4[25]

Polyvinyl fluoride (PVF) −20 −4[24]

Polypropylene (PP isotactic) 0 32[25]

Poly-3-hydroxybutyrate (PHB) 15 59[25]

Poly(vinyl acetate) (PVAc) 30 86[25]

Polychlorotrifluoroethylene (PCTFE) 45 113[24]

Polyamide (PA) 47–60 117–140 Nylon-6,x

Polylactic acid (PLA) 60–65 140–149

Polyethylene terephthalate (PET) 70 158[25]

Poly(vinyl chloride) (PVC) 80 176[25]

Poly(vinyl alcohol) (PVA) 85 185[25]

Polystyrene (PS) 95 203[25]

Poly(methyl methacrylate) (PMMA atactic) 105 221[25] Plexiglas, Perspex

Acrylonitrile butadiene styrene (ABS) 105 221[26]

Polytetrafluoroethylene (PTFE) 115 239[27] Teflon

Poly(carbonate) (PC) 145 293[25] Lexan

Polysulfone 185 365

Polynorbornene 215 419[25]

Dry nylon-6 has a glass transition


temperature of 47 °C (117 °F).[28] Nylon-6,6
in the dry state has a glass transition
temperature of about 70 °C (158 °F).[29][30]
Whereas polyethene has a glass transition
range of −130 – −80 °C (−202 –
−112 °F)[31] The above are only mean
values, as the glass transition temperature
depends on the cooling rate and molecular
weight distribution and could be
influenced by additives. For a semi-
crystalline material, such as polyethene
that is 60–80% crystalline at room
temperature, the quoted glass transition
refers to what happens to the amorphous
part of the material upon cooling.

Silicates and other covalent


network glasses
Material Tg (°C) Tg (°F)

Chalcogenide GeSbTe 150 302[32]

Chalcogenide AsGeSeTe 245 473

ZBLAN fluoride glass 235 455

Tellurium dioxide 280 536

Fluoroaluminate 400 752

Soda-lime glass 520–600 968–1,112

Fused quartz (approximate) 1,200 2,200[33]

Kauzmann's paradox

Entropy difference between crystal and undercooled


melt

As a liquid is supercooled, the difference in


entropy between the liquid and solid phase
decreases. By extrapolating the heat
capacity of the supercooled liquid below
its glass transition temperature, it is
possible to calculate the temperature at
which the difference in entropies becomes
zero. This temperature has been named
the Kauzmann temperature.

If a liquid could be supercooled below its


Kauzmann temperature, and it did indeed
display a lower entropy than the crystal
phase, the consequences would be
paradoxical. This Kauzmann paradox has
been the subject of much debate and
many publications since it was first put
forward by Walter Kauzmann in 1948.[34]
One resolution of the Kauzmann paradox
is to say that there must be a phase
transition before the entropy of the liquid
decreases. In this scenario, the transition
temperature is known as the calorimetric
ideal glass transition temperature T0c. In
this view, the glass transition is not merely
a kinetic effect, i.e. merely the result of
fast cooling of a melt, but there is an
underlying thermodynamic basis for glass
formation. The glass transition
temperature:
The Gibbs-DiMarzio model specifically
predicts that a supercooled liquid's
configurational entropy disappears in the
limit , where the liquid's
existence regime ends, its microstructure
becomes identical to the crystal's, and
their property curves intersect in a true
second-order phase transition. This has
never been experimentally verified due to
the difficulty of realizing a slow enough
cooling rate while avoiding accidental
crystallization.

Alternative resolutions
There are at least three other possible
resolutions to the Kauzmann paradox. It
could be that the heat capacity of the
supercooled liquid near the Kauzmann
temperature smoothly decreases to a
smaller value. It could also be that a first
order phase transition to another liquid
state occurs before the Kauzmann
temperature with the heat capacity of this
new state being less than that obtained by
extrapolation from higher temperature.
Finally, Kauzmann himself resolved the
entropy paradox by postulating that all
supercooled liquids must crystallize
before the Kauzmann temperature is
reached.
In specific materials
Silica, SiO2

Silica (the chemical compound SiO2) has a


number of distinct crystalline forms in
addition to the quartz structure. Nearly all
of the crystalline forms involve tetrahedral
SiO4 units linked together by shared
vertices in different arrangements. Si-O
bond lengths vary between the different
crystal forms. For example, in α-quartz the
bond length is 161 picometres
(6.3 × 10−9 in), whereas in α-tridymite it
ranges from 154–171 pm (6.1 × 10−9–
6.7 × 10−9 in). The Si-O-Si bond angle also
varies from 140° in α-tridymite to 144° in α-
quartz to 180° in β-tridymite. Any
deviations from these standard
parameters constitute microstructural
differences or variations that represent an
approach to an amorphous, vitreous or
glassy solid. The transition temperature Tg
in silicates is related to the energy required
to break and re-form covalent bonds in an
amorphous (or random network) lattice of
covalent bonds. The Tg is clearly
influenced by the chemistry of the glass.
For example, addition of elements such as
B, Na, K or Ca to a silica glass, which have
a valency less than 4, helps in breaking up
the network structure, thus reducing the
Tg. Alternatively, P, which has a valency of
5, helps to reinforce an ordered lattice, and
thus increases the Tg.[35]

Tg is directly proportional to bond strength,


e.g. it depends on quasi-equilibrium
thermodynamic parameters of the bonds
e.g. on the enthalpy Hd and entropy Sd of
configurons – broken bonds: Tg = Hd / [Sd
+ Rln[(1-fc)/ fc] where R is the gas constant
and fc is the percolation threshold. For
strong melts such as SiO2 the percolation
threshold in the above equation is the
universal Scher-Zallen critical density in
the 3-D space e.g. fc = 0.15, however for
fragile materials the percolation
thresholds are material-dependent and fc
<< 1.[36] The enthalpy Hd and the entropy
Sd of configurons – broken bonds can be
found from available experimental data on
viscosity.[37]

In ironing, a fabric is heated through the glass-rubber


transition.

Polymers
In polymers the glass transition
temperature, Tg, is often expressed as the
temperature at which the Gibbs free
energy is such that the activation energy
for the cooperative movement of 50 or so
elements of the polymer is exceeded. This
allows molecular chains to slide past each
other when a force is applied. From this
definition, we can see that the introduction
of relatively stiff chemical groups (such as
benzene rings) will interfere with the
flowing process and hence increase Tg.[38]
The stiffness of thermoplastics decreases
due to this effect (see figure.) When the
glass temperature has been reached, the
stiffness stays the same for a while, i.e., at
or near E2, until the temperature exceeds
Tm, and the material melts. This region is
called the rubber plateau.

In ironing, a fabric is heated through this


transition so that the polymer chains
become mobile. The weight of the iron
then imposes a preferred orientation. Tg
can be significantly decreased by addition
of plasticizers into the polymer matrix.
Smaller molecules of plasticizer embed
themselves between the polymer chains,
increasing the spacing and free volume,
and allowing them to move past one
another even at lower temperatures. The
addition of nonreactive side groups to a
polymer can also make the chains stand
off from one another, reducing Tg. If a
plastic with some desirable properties has
a Tg that is too high, it can sometimes be
combined with another in a copolymer or
composite material with a Tg below the
temperature of intended use. Note that
some plastics are used at high
temperatures, e.g., in automobile engines,
and others at low temperatures.[25]

Stiffness versus temperature


In viscoelastic materials, the presence of
liquid-like behavior depends on the
properties of and so varies with rate of
applied load, i.e., how quickly a force is
applied. The silicone toy Silly Putty
behaves quite differently depending on the
time rate of applying a force: pull slowly
and it flows, acting as a heavily viscous
liquid; hit it with a hammer and it shatters,
acting as a glass.

On cooling, rubber undergoes a liquid-glass


transition, which has also been called a
rubber-glass transition.

Mechanics of vitrification
Molecular motion in condensed matter
can be represented by a Fourier series
whose physical interpretation consists of
a superposition of longitudinal and
transverse waves of atomic displacement
with varying directions and wavelengths.
In monatomic systems, these waves are
called density fluctuations. (In polyatomic
systems, they may also include
compositional fluctuations.)[39]

Thus, thermal motion in liquids can be


decomposed into elementary longitudinal
vibrations (or acoustic phonons) while
transverse vibrations (or shear waves)
were originally described only in elastic
solids exhibiting the highly ordered
crystalline state of matter. In other words,
simple liquids cannot support an applied
force in the form of a shearing stress, and
will yield mechanically via macroscopic
plastic deformation (or viscous flow).
Furthermore, the fact that a solid deforms
locally while retaining its rigidity – while a
liquid yields to macroscopic viscous flow
in response to the application of an
applied shearing force – is accepted by
many as the mechanical distinction
between the two.[40][41]

The inadequacies of this conclusion,


however, were pointed out by Frenkel in his
revision of the kinetic theory of solids and
the theory of elasticity in liquids. This
revision follows directly from the
continuous characteristic of the structural
transition from the liquid state into the
solid one when this transition is not
accompanied by crystallization—ergo the
supercooled viscous liquid. Thus we see
the intimate correlation between
transverse acoustic phonons (or shear
waves) and the onset of rigidity upon
vitrification, as described by Bartenev in
his mechanical description of the
vitrification process.[42][43]
The velocities of longitudinal acoustic
phonons in condensed matter are directly
responsible for the thermal conductivity
that levels out temperature differentials
between compressed and expanded
volume elements. Kittel proposed that the
behavior of glasses is interpreted in terms
of an approximately constant "mean free
path" for lattice phonons, and that the
value of the mean free path is of the order
of magnitude of the scale of disorder in
the molecular structure of a liquid or solid.
The thermal phonon mean free paths or
relaxation lengths of a number of glass
formers have been plotted versus the
glass transition temperature, indicating a
linear relationship between the two. This
has suggested a new criterion for glass
formation based on the value of the
phonon mean free path.[44]

It has often been suggested that heat


transport in dielectric solids occurs
through elastic vibrations of the lattice,
and that this transport is limited by elastic
scattering of acoustic phonons by lattice
defects (e.g. randomly spaced
vacancies).[45] These predictions were
confirmed by experiments on commercial
glasses and glass ceramics, where mean
free paths were apparently limited by
"internal boundary scattering" to length
scales of 10–100 micrometres (0.00039–
0.00394 in).[46][47] The relationship
between these transverse waves and the
mechanism of vitrification has been
described by several authors who
proposed that the onset of correlations
between such phonons results in an
orientational ordering or "freezing" of local
shear stresses in glass-forming liquids,
thus yielding the glass transition.[48]

Electronic structure

The influence of thermal phonons and


their interaction with electronic structure is
a topic that was appropriately introduced
in a discussion of the resistance of liquid
metals. Lindemann's theory of melting is
referenced, and it is suggested that the
drop in conductivity in going from the
crystalline to the liquid state is due to the
increased scattering of conduction
electrons as a result of the increased
amplitude of atomic vibration. Such
theories of localization have been applied
to transport in metallic glasses, where the
mean free path of the electrons is very
small (on the order of the interatomic
spacing).[49][50]

The formation of a non-crystalline form of


a gold-silicon alloy by the method of splat
quenching from the melt led to further
considerations of the influence of
electronic structure on glass forming
ability, based on the properties of the
metallic bond.[51][52][53][54][55]

Other work indicates that the mobility of


localized electrons is enhanced by the
presence of dynamic phonon modes. One
claim against such a model is that if
chemical bonds are important, the nearly
free electron models should not be
applicable. However, if the model includes
the buildup of a charge distribution
between all pairs of atoms just like a
chemical bond (e.g., silicon, when a band
is just filled with electrons) then it should
apply to solids.[56]

Thus, if the electrical conductivity is low,


the mean free path of the electrons is very
short. The electrons will only be sensitive
to the short-range order in the glass since
they do not get a chance to scatter from
atoms spaced at large distances. Since
the short-range order is similar in glasses
and crystals, the electronic energies
should be similar in these two states. For
alloys with lower resistivity and longer
electronic mean free paths, the electrons
could begin to sense that there is disorder
in the glass, and this would raise their
energies and destabilize the glass with
respect to crystallization. Thus, the glass
formation tendencies of certain alloys may
therefore be due in part to the fact that the
electron mean free paths are very short, so
that only the short-range order is ever
important for the energy of the electrons.

It has also been argued that glass


formation in metallic systems is related to
the "softness" of the interaction potential
between unlike atoms. Some authors,
emphasizing the strong similarities
between the local structure of the glass
and the corresponding crystal, suggest
that chemical bonding helps to stabilize
the amorphous structure.[57][58]

Other authors have suggested that the


electronic structure yields its influence on
glass formation through the directional
properties of bonds. Non-crystallinity is
thus favored in elements with a large
number of polymorphic forms and a high
degree of bonding anisotropy.
Crystallization becomes more unlikely as
bonding anisotropy is increased from
isotropic metallic to anisotropic metallic to
covalent bonding, thus suggesting a
relationship between the group number in
the periodic table and the glass forming
ability in elemental solids.[59]

References
1. ISO 11357-2: Plastics – Differential
scanning calorimetry – Part 2:
Determination of glass transition
temperature (1999).
2. "The Glass Transition" . Polymer
Science Learning Center.
3. Debenedetti, P. G.; Stillinger (2001).
"Supercooled liquids and the glass
transition". Nature. 410 (6825): 259–
267. Bibcode:2001Natur.410..259D .
doi:10.1038/35065704 .
PMID 11258381 .
4. Angell, C. A.; Ngai, K. L.; McKenna, G.
B.; McMillan, P. F.; Martin, S. W. (2000).
"Relaxation in glassforming liquids and
amorphous solids" . Appl. Phys. Rev.
88 (6): 3113–3157.
Bibcode:2000JAP....88.3113A .
doi:10.1063/1.1286035 .
5. Zarzycki, J. (1991). Glasses and the
Vitreous State . Cambridge University
Press. ISBN 978-0521355827.
6. Ojovan, M. I. (2004). "Glass formation
in amorphous SiO2 as a percolation
phase transition in a system of
network defects". Journal of
Experimental and Theoretical Physics
Letters. 79 (12): 632–634.
Bibcode:2004JETPL..79..632O .
doi:10.1134/1.1790021 .
7. Meille Stefano, V.; Allegra, G.; Geil
Phillip, H.; He, J.; Hess, M.; Jin, J.-I.;
Kratochvíl, P.; Mormann, W.; Stepto, R.
(2011). "Definitions of terms relating to
crystalline polymers (IUPAC
Recommendations 2011)" (PDF). Pure
Appl Chem. 83 (10): 1831.
doi:10.1351/PAC-REC-10-11-13 .
8. IUPAC, Compendium of Chemical
Terminology, 2nd ed. (the "Gold Book")
(1997). Online corrected version:
 (2006–) "glass transition ".
doi:10.1351/goldbook.G02640
9. Hess, M.; Allegra, G.; He, J.; Horie, K.;
Kim, J.-S.; Meille Stefano, V.;
Metanomski, V.; Moad, G.; Stepto
Robert, F. T.; Vert, M.; Vohlídal, J.
(2013). "Glossary of terms relating to
thermal and thermomechanical
properties of polymers (IUPAC
Recommendations 2013)" (PDF). Pure
Appl Chem. 85 (5): 1017.
doi:10.1351/PAC-REC-12-03-02 .
10. Hansen, J.-P.; McDonald, I. R. (2007).
Theory of Simple Liquids . Elsevier.
pp. 250–254. ISBN 978-0123705358.
11. Adam, J-L; Zhang, X. (14 February
2014). Chalcogenide Glasses:
Preparation, Properties and
Applications . Elsevier Science. p. 94.
ISBN 978-0-85709-356-1.
12. Phillips, J.C. (1979). "Topology of
covalent non-crystalline solids I: Short-
range order in chalcogenide alloys".
Journal of Non-Crystalline Solids. 34
(2): 153.
Bibcode:1979JNCS...34..153P .
doi:10.1016/0022-3093(79)90033-4 .
13. Moynihan, C. et al. (1976) in The Glass
Transition and the Nature of the
Glassy State, M. Goldstein and R.
Simha (Eds.), Ann. N.Y. Acad. Sci., Vol.
279. ISBN 0890720533.
14. Angell, C. A. (1988). "Perspective on
the glass transition". Journal of
Physics and Chemistry of Solids. 49
(8): 863–871.
Bibcode:1988JPCS...49..863A .
doi:10.1016/0022-3697(88)90002-9 .
15. Ediger, M. D.; Angell, C. A.; Nagel,
Sidney R. (1996). "Supercooled Liquids
and Glasses". The Journal of Physical
Chemistry. 100 (31): 13200.
doi:10.1021/jp953538d .
16. Angell, C. A. (1995). "Formation of
Glasses from Liquids and
Biopolymers". Science. 267 (5206):
1924–35.
Bibcode:1995Sci...267.1924A .
doi:10.1126/science.267.5206.1924 .
PMID 17770101 .
17. Stillinger, F. H. (1995). "A Topographic
View of Supercooled Liquids and
Glass Formation". Science. 267
(5206): 1935–9.
Bibcode:1995Sci...267.1935S .
doi:10.1126/science.267.5206.1935 .
PMID 17770102 .
18. Nemilov SV (1994). Thermodynamic
and Kinetic Aspects of the Vitreous
State. CRC Press. ISBN 978-
0849337826.
19. Gibbs, J. H. (1960). MacKenzie, J. D.
(ed.). Modern Aspects of the Vitreous
State. Butterworth. OCLC 1690554 .
20. Ojovan, Michael I; Lee, William (Bill) E
(2010). "Connectivity and glass
transition in disordered oxide
systems". Journal of Non-Crystalline
Solids. 356 (44–49): 2534.
Bibcode:2010JNCS..356.2534O .
doi:10.1016/j.jnoncrysol.2010.05.012 .
21. Tg measurement of glasses .
Glassproperties.com. Retrieved on
2012-06-29.
22. IUPAC, Compendium of Chemical
Terminology, 2nd ed. (the "Gold Book")
(1997). Online corrected version:
 (2006–) "glass-transition
temperature ".
doi:10.1351/goldbook.G02641
23. Galimberti, Maurizio; Caprio, Michela;
Fino, Luigi (2001-12-21). "Tyre
comprising a cycloolefin polymer,
tread band and elasomeric
composition used therein" (published
2003-03-07). "country-code =EU,
patent-number =WO03053721"
24. Ibeh, Christopher C. (2011).
THERMOPLASTIC MATERIALS
Properties, Manufacturing Methods,
and Applications. CRC Press. pp. 491–
497. ISBN 978-1-4200-9383-4.
25. Wilkes, C. E. (2005). PVC Handbook .
Hanser Verlag. ISBN 978-1-56990-379-
7.
26. ABS . nrri.umn.edu
27. Nicholson, John W. (2011). The
Chemistry of Polymers (4, Revised
ed.). Royal Society of Chemistry. p. 50.
ISBN 9781849733915. Retrieved
10 September 2013.
28. nylon-6 information and properties .
Polymerprocessing.com (2001-04-15).
Retrieved on 2012-06-29.
29. Jones, A (2014). "Supplementary
Materials for Artificial Muscles from
Fishing Line and Sewing Thread".
Science. 343 (6173): 868–72.
Bibcode:2014Sci...343..868H .
doi:10.1126/science.1246906 .
PMID 24558156 .
30. Measurement of Moisture Effects on
the Mechanical Properties of 66
Nylon . TA Instruments Thermal
Analysis Application Brief TA-133
31. PCL | Applications and End Uses |
Polythene . Polyesterconverters.com.
Retrieved on 2012-06-29.
32. EPCOS 2007: Glass Transition and
Crystallization in Phase Change
Materials . Retrieved on 2012-06-29.
33. Bucaro, J. A. (1974). "High-
temperature Brillouin scattering in
fused quartz". Journal of Applied
Physics. 45 (12): 5324–5329.
Bibcode:1974JAP....45.5324B .
doi:10.1063/1.1663238 .
34. Kauzmann, Walter (1948). "The Nature
of the Glassy State and the Behavior of
Liquids at Low Temperatures".
Chemical Reviews. 43 (2): 219–256.
doi:10.1021/cr60135a002 .
35. Ojovan M.I. (2008). "Configurons:
thermodynamic parameters and
symmetry changes at glass
transition" (PDF). Entropy. 10 (3):
334–364.
Bibcode:2008Entrp..10..334O .
doi:10.3390/e10030334 .
36. Ojovan, M.I. (2008). "Configurons:
thermodynamic parameters and
symmetry changes at glass
transition" (PDF). Entropy. 10 (3):
334–364.
Bibcode:2008Entrp..10..334O .
doi:10.3390/e10030334 .
37. Ojovan, Michael I; Travis, Karl P; Hand,
Russell J (2007). "Thermodynamic
parameters of bonds in glassy
materials from viscosity–temperature
relationships". Journal of Physics:
Condensed Matter. 19 (41): 415107.
Bibcode:2007JPCM...19O5107O .
doi:10.1088/0953-
8984/19/41/415107 .
PMID 28192319 .
38. Cowie, J. M. G. and Arrighi, V.,
Polymers: Chemistry and Physics of
Modern Materials, 3rd Edn. (CRC
Press, 2007) ISBN 0748740732
39. Slater, J.C., Introduction to Chemical
Physics (3rd Ed., Martindell Press,
2007) ISBN 1178626598
40. Born, Max (2008). "On the stability of
crystal lattices. I". Mathematical
Proceedings of the Cambridge
Philosophical Society. 36 (2): 160.
Bibcode:1940PCPS...36..160B .
doi:10.1017/S0305004100017138 .
41. Born, Max (1939). "Thermodynamics
of Crystals and Melting". The Journal
of Chemical Physics. 7 (8): 591–603.
Bibcode:1939JChPh...7..591B .
doi:10.1063/1.1750497 .
42. Frenkel, J. (1946). Kinetic Theory of
Liquids. Clarendon Press, Oxford.
43. Bartenev, G. M., Structure and
Mechanical Properties of Inorganic
Glasses (Wolters – Noordhoof, 1970)
ISBN 9001054501
44. Reynolds, C. L. Jr. (1979). "Correlation
between the low temperature phonon
mean free path and glass transition
temperature in amorphous solids". J.
Non-Cryst. Solids. 30 (3): 371.
Bibcode:1979JNCS...30..371R .
doi:10.1016/0022-3093(79)90174-1 .
45. Rosenburg, H. M. (1963) Low
Temperature Solid State Physics.
Clarendon Press, Oxford.
46. Kittel, C. (1946). "Ultrasonic
Propagation in Liquids". J. Chem.
Phys. 14 (10): 614.
Bibcode:1946JChPh..14..614K .
doi:10.1063/1.1724073 .
hdl:1721.1/5041 .
47. Kittel, C. (1949). "Interpretation of the
Thermal Conductivity of Glasses".
Phys. Rev. 75 (6): 972.
Bibcode:1949PhRv...75..972K .
doi:10.1103/PhysRev.75.972 .
48. Chen, Shao-Ping; Egami, T.; Vitek, V.
(1985). "Orientational ordering of local
shear stresses in liquids: A phase
transition?". Journal of Non-Crystalline
Solids. 75 (1–3): 449.
Bibcode:1985JNCS...75..449C .
doi:10.1016/0022-3093(85)90256-X .
49. Mott, N. F. (1934). "The Resistance of
Liquid Metals". Proceedings of the
Royal Society A. 146 (857): 465.
Bibcode:1934RSPSA.146..465M .
doi:10.1098/rspa.1934.0166 .
50. Lindemann, C. (1911). "On the
calculation of molecular natural
frequencies". Phys. Z. 11: 609.
51. Klement, W.; Willens, R. H.; Duwez, POL
(1960). "Non-crystalline Structure in
Solidified Gold–Silicon Alloys". Nature.
187 (4740): 869.
Bibcode:1960Natur.187..869K .
doi:10.1038/187869b0 .
52. Duwez, Pol; Willens, R. H.; Klement, W.
(1960). "Continuous Series of
Metastable Solid Solutions in Silver-
Copper Alloys" (PDF). Journal of
Applied Physics. 31 (6): 1136.
Bibcode:1960JAP....31.1136D .
doi:10.1063/1.1735777 .
53. Duwez, Pol; Willens, R. H.; Klement, W.
(1960). "Metastable Electron
Compound in Ag-Ge Alloys". Journal of
Applied Physics. 31 (6): 1137.
Bibcode:1960JAP....31.1137D .
doi:10.1063/1.1735778 .
54. Chaudhari, P; Turnbull, D (1978).
"Structure and properties of metallic
glasses". Science. 199 (4324): 11–21.
Bibcode:1978Sci...199...11C .
doi:10.1126/science.199.4324.11 .
PMID 17841932 .
55. Chen, J. S. (1980). "Glassy metals".
Reports on Progress in Physics. 43 (4):
353. Bibcode:1980RPPh...43..353C .
doi:10.1088/0034-4885/43/4/001 .
56. Jonson, M.; Girvin, S. M. (1979).
"Electron-Phonon Dynamics and
Transport Anomalies in Random Metal
Alloys". Phys. Rev. Lett. 43 (19): 1447.
Bibcode:1979PhRvL..43.1447J .
doi:10.1103/PhysRevLett.43.1447 .
57. Turnbull, D. (1974). "Amorphous Solid
Formation and Interstitial Solution
Behavior in Metallic Alloy System". J.
Phys. C. 35 (C4): C4–1.
CiteSeerX 10.1.1.596.7462 .
doi:10.1051/jphyscol:1974401 .
58. Chen, H. S.; Park, B. K. (1973). "Role of
chemical bonding in metallic glasses".
Acta Metall. 21 (4): 395.
doi:10.1016/0001-6160(73)90196-X .
59. Wang, R.; Merz, D. (1977).
"Polymorphic bonding and thermal
stability of elemental noncrystalline
solids". Physica Status Solidi A. 39 (2):
697. Bibcode:1977PSSAR..39..697W .
doi:10.1002/pssa.2210390240 .

External links
Wikimedia Commons has media
related to Glass-liquid transitions.

Fragility
VFT Eqn.
Polymers I
Polymers II
Angell: Aqueous media
DoITPoMS Teaching and Learning
Package- "The Glass Transition in
Polymers"

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Glass_transition&oldid=896469286"
Last edited 19 days ago by JCW-Cle…

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like