You are on page 1of 10

Glass transition

The glass–liquid transitionor glass transition for short is the 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 calledvitrification.

The glass-transition temperature Tg of a material characterizes the range of temperatures over which this glass transition occurs. It
is always lower than themelting 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, that is in their
glassy state. Their Tg values are well above room temperature, both at around 100 °C (212 °F). Rubber elastomers like polyisoprene
[2]
and polyisobutylene are used above their Tg, that is, in the rubbery state, where they are soft and flexible.

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
[3][4][6]
question of whether some phase transition underlies the glass transition is a matter of continuing research.

Contents
Introduction
Transition temperatureTg
Polymers
Silicates and other covalent network glasses
Kauzmann's paradox
In specific materials
Silica, SiO2
Polymers
Mechanics of vitrification
Electronic structure
References
External links

Introduction
The glass transition of a liquid to a solid-like state may occur with either cooling or compression.[7] The transition comprises a
smooth increase in the viscosity of a material by as much as 17 orders of magnitude within a temperature range of 500 K without any
pronounced change in material structure.[8] The consequence of this dramatic increase is a glass exhibiting solid-like mechanical
properties on the timescale of practical observation. This transition is in contrast to the freezing or crystallization transition, which is
a first-order phase transition in the Ehrenfest classification and involves discontinuities in thermodynamic and dynamic properties
such as volume, energy, and viscosity. In many materials that normally undergo a freezing transition, rapid cooling will avoid this
phase transition and instead result in a glass transition at some lower temperature. Other materials, such as many polymers, lack a
well defined crystalline state and easily form glasses, even upon very slow cooling or compression. The tendency for a material to
form a glass while quenched is called glass forming ability. This ability depends on the composition of the material and can be
predicted by the rigidity theory.[9]

Below the transition temperature range, the glassy structure does not relax in accordance with the cooling rate used. The expansion
coefficient for the glassy state is roughly equivalent to that of the crystalline solid. If slower cooling rates are used, the increased time
for structural relaxation (or intermolecular rearrangement) to occur may result in a higher density glass product. Similarly, by
annealing (and thus allowing for slow structural relaxation) the glass structure in time approaches an equilibrium density
corresponding to the supercooled liquid at this same temperature. Tg is located at the intersection between the cooling curve (volume
[10][11][12][13][14]
versus temperature) for the glassy state and the supercooled liquid.

The configuration of the glass in this temperature range changes slowly with time towards the equilibrium structure. The principle of
the minimization of the Gibbs free energy provides the thermodynamic driving force necessary for the eventual change. It should be
noted here that at somewhat higher temperatures than Tg, the structure corresponding to equilibrium at any temperature is achieved
quite rapidly. In contrast, at considerably lower temperatures, the configuration of the glass remains sensibly stable over increasingly
extended periods of time.

Thus, the liquid-glass transition is not a transition between states of thermodynamic equilibrium. It is widely 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][15][16][17]

Transition temperature Tg
Refer to the figure on the 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
[18]
temperature at the intersection of the red regression lines.

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
Measurement of Tg (the temperature
experimentally, this value is close to theannealing point of many glasses.[19]
at the point A) by differential
In contrast to viscosity, the thermal expansion, heat capacity, shear modulus, and scanning calorimetry
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 areTg values characteristic of certain classes of materials.

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

Tire rubber −70 −94[20]

Polyvinylidene fluoride (PVDF) −35 −31[21]

Polypropylene (PP atactic) −20 −4[22]

Polyvinyl fluoride (PVF) −20 −4[21]

Polypropylene (PP isotactic) 0 32[22] Determination of Tg by dilatometry.


Poly-3-hydroxybutyrate(PHB) 15 59[22]

Poly(vinyl acetate) (PVAc) 30 86[22]


Polychlorotrifluoroethylene
45 113[21]
(PCTFE)
47– 117–
Polyamide (PA) Nylon-6,x
60 140
60– 140–
Polylactic acid (PLA)
65 149

Polyethylene terephthalate(PET) 70 158[22]

Poly(vinyl chloride) (PVC) 80 176[22]

Poly(vinyl alcohol) (PVA) 85 185[22]

Polystyrene (PS) 95 203[22]


Poly(methyl methacrylate)(PMMA Plexiglas,
105 221[22]
atactic) Perspex
Acrylonitrile butadiene styrene
105 221[23]
(ABS)

Polytetrafluoroethylene(PTFE) 115 239[24] Teflon

Poly(carbonate) (PC) 145 293[22] Lexan

Polysulfone 185 365

Polynorbornene 215 419[22]

Dry nylon-6 has a glass transition temperature of 47 °C (117 °F).[25] Nylon-6,6 in the dry state has a glass transition temperature of
about 70 °C (158 °F).[26][27] Whereas polyethene has a glass transition range of −130 – −80 °C (−202 – −112 °F)[28] 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[29]


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[30]

Kauzmann's paradox
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 byWalter Kauzmann in 1948.[31]
Entropy difference between crystal
One resolution of the Kauzmann paradox is to say that there must be a phase and undercooled melt
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:

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 anamorphous, 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 theTg.[32]

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.[33] The enthalpy Hd and the entropy Sd of configurons – broken bonds can be found from available experimental data on
viscosity.[34]

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.[35]
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 In ironing, a fabric is heated through
near E2, until the temperature exceeds Tm, and the material melts. This region is the glass-rubber transition.
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 "new-car smell" is due to the initial outgassing of volatile small-molecule plasticizers
(most commonly known as phthalates) used to modify interior plastics (e.g., dashboards) to keep them from cracking in the cold of
winter weather. 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
[22]
automobile engines, and others at low temperatures.

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.
Stiffness versus temperature

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 calleddensity fluctuations. (In polyatomic systems, they may also includecompositional fluctuations.)[36]
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 macroscopicviscous flow in response to the application of an applied shearing force – is accepted by many as
the mechanical distinction between the two.[37][38]

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
[39][40]
described by Bartenev in his mechanical description of the vitrification process.

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.[41]

It has often been suggested thatheat 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).[42] 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).[43][44] 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.[45]

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 increasedscattering 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
[46][47]
mean free path of the electrons is very small (on the order of the interatomic spacing).

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.[48][49][50][51][52]

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 ifchemical bonds are important, the nearly free electron modelsshould 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 tosolids.[53]

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,
[54][55]
suggest that chemical bonding helps to stabilize the amorphous structure.

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.[56]

References
1. ISO 11357-2: Plastics – Differential scanning calorimetry – Part 2: Determination of glass transition temperature
(1999).
2. "The Glass Transition" (http://pslc.ws/macrog/tg.htm). 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 (http://adsabs.harvard.edu/abs/2001Natur .410..259D). doi:10.1038/35065704 (https://
doi.org/10.1038%2F35065704). PMID 11258381 (https://www.ncbi.nlm.nih.gov/pubmed/11258381).
4. Angell, C. A.; Ngai, K. L.; McKenna, G. B.; McMillan, .PF.; Martin, S. W. (2000). "Relaxation in glassforming liquids
and amorphous solids".App. Phys. Rev. 88 (6): 3113–3157. Bibcode:2000JAP....88.3113A (http://adsabs.harvard.ed
u/abs/2000JAP....88.3113A). doi:10.1063/1.1286035 (https://doi.org/10.1063%2F1.1286035).
5. Zarzycki, J. (1991). Glasses and the Vitreous State (https://books.google.com/books?id=D7Z8ywb3QggC&printsec=f
rontcover). Cambridge University Press.ISBN 0521355826.
6. Ojovan, M. I. (2004). "Glass formation in amorphous SiO 2 as a percolation phase transition in a system of network
defects". Journal of Experimental and Theoretical Physics Letters
. 79 (12): 632. Bibcode:2004JETPL..79..632O (htt
p://adsabs.harvard.edu/abs/2004JETPL..79..632O) . doi:10.1134/1.1790021 (https://doi.org/10.1134%2F1.1790021).
7. Hansen, J.-P.; McDonald, I. R. (2007).Theory of Simple Liquids(https://books.google.com/books?id=Uhm87WZBnx
EC&printsec=frontcover). Elsevier. pp. 250–254. ISBN 0123705355.
8. Adam, J-L; Zhang, X. (14 February 2014).Chalcogenide Glasses: Preparation, Properties and Applications(https://b
ooks.google.com/books?id=IuNRAwAAQBAJ&pg=PA94). Elsevier Science. p. 94.ISBN 978-0-85709-356-1.
9. 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 (http://adsabs.harvard.edu/abs/1979JNCS...3
4..153P). doi:10.1016/0022-3093(79)90033-4(https://doi.org/10.1016%2F0022-3093%2879%2990033-4) .
10. Moynihan, C. et al. (1976) inThe Glass Transition and the Nature of the Glassy State, M. Goldstein and R. Simha
(Eds.), Ann. N.Y. Acad. Sci., Vol. 279. ISBN 0890720533.
11. Angell, C. A. (1988). "Perspective on the glass transition".Journal of Physics and Chemistry of Solids. 49 (8): 863.
Bibcode:1988JPCS...49..863A (http://adsabs.harvard.edu/abs/1988JPCS...49..863A) . doi:10.1016/0022-
3697(88)90002-9 (https://doi.org/10.1016%2F0022-3697%2888%2990002-9) .
12. 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 (https://doi.org/10.1021%2Fjp953538d).
13. Angell, C. A. (1995). "Formation of Glasses from Liquids and Biopolymers".
Science. 267 (5206): 1924–35.
Bibcode:1995Sci...267.1924A (http://adsabs.harvard.edu/abs/1995Sci...267.1924A).
doi:10.1126/science.267.5206.1924(https://doi.org/10.1126%2Fscience.267.5206.1924) . PMID 17770101 (https://w
ww.ncbi.nlm.nih.gov/pubmed/17770101).
14. Stillinger, F. H. (1995). "A Topographic View of Supercooled Liquids and Glass Formation". Science. 267 (5206):
1935–9. Bibcode:1995Sci...267.1935S (http://adsabs.harvard.edu/abs/1995Sci...267.1935S) .
doi:10.1126/science.267.5206.1935(https://doi.org/10.1126%2Fscience.267.5206.1935) . PMID 17770102 (https://w
ww.ncbi.nlm.nih.gov/pubmed/17770102).
15. Nemilov SV (1994). Thermodynamic and Kinetic Aspects of the V
itreous State. CRC Press. ISBN 0849337828.
16. Gibbs, J. H. (1960). MacKenzie, J. D., ed.Modern Aspects of the Vitreous State. Butterworth. OCLC 1690554 (http
s://www.worldcat.org/oclc/1690554).
17. 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(http://adsabs.harvard.edu/ab
s/2010JNCS..356.2534O). doi:10.1016/j.jnoncrysol.2010.05.012(https://doi.org/10.1016%2Fj.jnoncrysol.2010.05.01
2).
18. Tg measurement of glasses(http://www.glassproperties.com/tg/). Glassproperties.com. Retrieved on 2012-06-29.
19. glass-transition temperature(http://old.iupac.org/goldbook/G02641.pdf), IUPAC Compendium of Chemical
Terminology, 66, 583 (1984)
20. Galimberti, Maurizio; Caprio, Michela; Fino, Luigi (2001-12-21)."Tyre comprising a cycloolefin polymer, tread band
and elasomeric composition used therein"(http://patentscope.wipo.int/search/en/WO2003053721)(published 2003-
03-07). "country-code =EU, patent-number =WO03053721 "
21. Ibeh, Christopher C. (2011).THERMOPLASTIC MATERIALS Properties, Manufacturing Methods, and Applications.
CRC Press. pp. 491–497.ISBN 978-1-4200-9383-4.
22. Wilkes, C. E. (2005). PVC Handbook (https://books.google.com/books?id=YUkJNI9QYsUC&pg=PT1&lpg=PT1)
.
Hanser Verlag. ISBN 1-56990-379-4.
23. ABS (https://web.archive.org/web/20120227150812/https://www
.nrri.umn.edu/NLTC/ABS07.pdf). nrri.umn.edu
24. Nicholson, John W. (2011). The Chemistry of Polymers(https://books.google.com/books?id=5XFsT69cX_YC&prints
ec=frontcover&dq=polymer+chemistry#v=onepage&q=polymer%20chemistry&f=false) (4, Revised ed.). Royal
Society of Chemistry. p. 50. ISBN 9781849733915. Retrieved 10 September 2013.
25. nylon-6 information and properties(http://www.polymerprocessing.com/polymers/PA6.html). Polymerprocessing.com
(2001-04-15). Retrieved on 2012-06-29.
26. Jones, A (2014). "Supplementary Materials for Artificial Muscles from Fishing Line and Sewing Thread".
Science.
343 (6173): 868 Fig. S12. Bibcode:2014Sci...343..868H (http://adsabs.harvard.edu/abs/2014Sci...343..868H)
.
doi:10.1126/science.1246906(https://doi.org/10.1126%2Fscience.1246906) .
27. Measurement of Moisture Effects on the Mechanical Properties of 66 Nylon(https://web.archive.org/web/201505201
10712/http://www.tainstruments.co.jp/application/pdf/Thermal_Library/Applications_Briefs/T
A133.PDF). TA
Instruments Thermal Analysis Application Brief TA-133
28. PCL | Applications and End Uses | Polythene(http://www.polyesterconverters.com/pcl_apps/stage1/stage2/applicati
ons_and_enduses/polyethene.htm). Polyesterconverters.com. Retrieved on 2012-06-29.
29. EPCOS 2007: Glass Transition and Crystallization in Phase Change Materials(http://www.epcos.org/library/papers/p
df_2007/paper13_Salinga.pdf). Retrieved on 2012-06-29.
30. Bucaro, J. A. (1974). "High-temperature Brillouin scattering in fused quartz".
Journal of Applied Physics. 45 (12):
5324–5329. Bibcode:1974JAP....45.5324B (http://adsabs.harvard.edu/abs/1974JAP ....45.5324B).
doi:10.1063/1.1663238 (https://doi.org/10.1063%2F1.1663238).
31. Kauzmann, Walter (1948). "The Nature of theGlassy State and the Behavior of Liquids at Low e
Tmperatures".
Chemical Reviews. 43 (2): 219. doi:10.1021/cr60135a002 (https://doi.org/10.1021%2Fcr60135a002).
32. Ojovan M.I. (2008). "Configurons: thermodynamic parameters and symmetry changes at glass transition"
(http://ww
w.mdpi.org/entropy/papers/e10030334.pdf)(PDF). Entropy. 10 (3): 334–364. Bibcode:2008Entrp..10..334O (http://ad
sabs.harvard.edu/abs/2008Entrp..10..334O). doi:10.3390/e10030334 (https://doi.org/10.3390%2Fe10030334).
33. Ojovan, M.I. (2008). "Configurons: thermodynamic parameters and symmetry changes at glass transition"
(http://ww
w.mdpi.org/entropy/papers/e10030334.pdf)(PDF). Entropy. 10 (3): 334–364. Bibcode:2008Entrp..10..334O (http://ad
sabs.harvard.edu/abs/2008Entrp..10..334O). doi:10.3390/e10030334 (https://doi.org/10.3390%2Fe10030334).
34. 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(http://adsabs.harvard.edu/abs/2007JPCM...19O5107O) . doi:10.1088/0953-
8984/19/41/415107 (https://doi.org/10.1088%2F0953-8984%2F19%2F41%2F415107) .
35. Cowie, J. M. G. and Arrighi, V., Polymers: Chemistry and Physics of Modern Materials, 3rd Edn. (CRC Press, 2007)
ISBN 0748740732
36. Slater, J.C., Introduction to Chemical Physics(3rd Ed., Martindell Press, 2007)ISBN 1178626598
37. Born, Max (2008). "On the stability of crystal lattices. I".Mathematical Proceedings of the Cambridge Philosophical
Society. 36 (2): 160. Bibcode:1940PCPS...36..160B (http://adsabs.harvard.edu/abs/1940PCPS...36..160B) .
doi:10.1017/S0305004100017138(https://doi.org/10.1017%2FS0305004100017138) .
38. Born, Max (1939). "Thermodynamics of Crystals and Melting".The Journal of Chemical Physics. 7 (8): 591.
Bibcode:1939JChPh...7..591B (http://adsabs.harvard.edu/abs/1939JChPh...7..591B). doi:10.1063/1.1750497 (http
s://doi.org/10.1063%2F1.1750497).
39. Frenkel, J. (1946). Kinetic Theory of Liquids. Clarendon Press, Oxford.
40. Bartenev, G. M., Structure and Mechanical Properties of Inorganic Glasses (Wolters – Noordhoof, 1970)
ISBN 9001054501
41. Reynolds, C. L. Jr. (1979). "Correlation between the low temperature phonon mean free path and glass transition
temperature in amorphous solids".J. Non-Cryst. Sol. 30 (3): 371. Bibcode:1979JNCS...30..371R (http://adsabs.harv
ard.edu/abs/1979JNCS...30..371R). doi:10.1016/0022-3093(79)90174-1(https://doi.org/10.1016%2F0022-3093%28
79%2990174-1).
42. Rosenburg, H. M. (1963)Low Temperature Solid State Physics. Clarendon Press, Oxford.
43. Kittel, C. (1946). "Ultrasonic Propagation in Liquids".J. Chem. Phys. 14 (10): 614. Bibcode:1946JChPh..14..614K (ht
tp://adsabs.harvard.edu/abs/1946JChPh..14..614K) . doi:10.1063/1.1724073 (https://doi.org/10.1063%2F1.1724073).
44. Kittel, C. (1949). "Interpretation of the Thermal Conductivity of Glasses".
Phys. Rev. 75 (6): 972.
Bibcode:1949PhRv...75..972K (http://adsabs.harvard.edu/abs/1949PhRv ...75..972K). doi:10.1103/PhysRev.75.972
(https://doi.org/10.1103%2FPhysRev.75.972).
45. 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: 449. Bibcode:1985JNCS...75..449C (http://adsabs.harvard.edu/ab
s/1985JNCS...75..449C). doi:10.1016/0022-3093(85)90256-X(https://doi.org/10.1016%2F0022-3093%2885%29902
56-X).
46. Mott, N. F. (1934). "The Resistance of LiquidMetals". Proceedings of the Royal Society A. 146 (857): 465.
Bibcode:1934RSPSA.146..465M(http://adsabs.harvard.edu/abs/1934RSPSA.146..465M) .
doi:10.1098/rspa.1934.0166(https://doi.org/10.1098%2Frspa.1934.0166).
47. Lindemann, C. (1911). "On the calculation of molecular natural frequencies".
Phys. Z. 11: 609.
48. Klement, W.; Willens, R. H.; Duwez, POL (1960). "Non-crystalline Structure in Solidified Gold–Silicon Alloys".Nature.
187 (4740): 869. Bibcode:1960Natur.187..869K (http://adsabs.harvard.edu/abs/1960Natur .187..869K).
doi:10.1038/187869b0 (https://doi.org/10.1038%2F187869b0).
49. Duwez, Pol; Willens, R. H.; Klement, W. (1960). "Continuous Series of Metastable Solid Solutions in Silver-Copper
Alloys". Journal of Applied Physics. 31 (6): 1136. Bibcode:1960JAP....31.1136D (http://adsabs.harvard.edu/abs/1960
JAP....31.1136D). doi:10.1063/1.1735777 (https://doi.org/10.1063%2F1.1735777).
50. 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 (http://adsabs.harvard.edu/abs/1960JAP....31.1137D).
doi:10.1063/1.1735778 (https://doi.org/10.1063%2F1.1735778).
51. Chaudhari, P; Turnbull, D (1978). "Structure and properties of metallic glasses".Science. 199 (4324): 11–21.
Bibcode:1978Sci...199...11C (http://adsabs.harvard.edu/abs/1978Sci...199...11C) . doi:10.1126/science.199.4324.11
(https://doi.org/10.1126%2Fscience.199.4324.11) . PMID 17841932 (https://www.ncbi.nlm.nih.gov/pubmed/1784193
2).
52. Chen, J. S. (1980). "Glassy metals".Reports on Progress in Physics. 43 (4): 353. Bibcode:1980RPPh...43..353C (htt
p://adsabs.harvard.edu/abs/1980RPPh...43..353C) . doi:10.1088/0034-4885/43/4/001(https://doi.org/10.1088%2F00
34-4885%2F43%2F4%2F001).
53. Jonson, M.; Girvin, S. M. (1979). "Electron-Phonon Dynamics and ransport
T Anomalies in Random Metal Alloys".
Phys. Rev. Lett. 43 (19): 1447. Bibcode:1979PhRvL..43.1447J (http://adsabs.harvard.edu/abs/1979PhRvL..43.1447
J). doi:10.1103/PhysRevLett.43.1447(https://doi.org/10.1103%2FPhysRevLett.43.1447) .
54. Turnbull, D. (1974). "Amorphous Solid Formation and Interstitial Solution Behavior in Metallic Alloy System".J. Phys.
C. 35 (C4): C4–1. doi:10.1051/jphyscol:1974401(https://doi.org/10.1051%2Fjphyscol%3A1974401) .
55. Chen, H. S.; Park, B. K. (1973). "Role of chemical bonding in metallic glasses".
Acta Met. 21 (4): 395.
doi:10.1016/0001-6160(73)90196-X(https://doi.org/10.1016%2F0001-6160%2873%2990196-X) .
56. 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(http://adsabs.harvard.edu/abs/1977PSSAR..39..697
W). doi:10.1002/pssa.2210390240(https://doi.org/10.1002%2Fpssa.2210390240) .

External links
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=830818042


"

This page was last edited on 17 March 2018, at 03:29.

Text is available under theCreative Commons Attribution-ShareAlike License ; additional terms may apply. By using this
site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of theWikimedia
Foundation, Inc., a non-profit organization.

You might also like