My watch list
my.chemeurope.com  
Login  

Glass transition temperature



The glass transition temperature, Tg, is the temperature at which a glass becomes brittle on cooling, or soft on heating. More specifically, it defines a pseudo second order phase transition in which a supercooled melt yields, on cooling, a glassy structure and properties similar to those of crystalline materials e.g. of an isotropic solid material.[1] Tg is usually applicable to wholly or partially amorphous solids such as common glasses and plastics (organic polymers).

Below the glass transition temperature, Tg, amorphous solids are in a glassy state[2] and most of their joining bonds are intact. In inorganic glasses, with increased temperature more and more joining bonds are broken by thermal fluctuations so that broken bonds (termed configurons) begin to form clusters. Above Tg these clusters become macroscopic large facilitating the flow of material. In organic polymers, secondary, non-covalent bonds between the polymer chains become weak above Tg. Above Tg glasses and organic polymers become soft and capable of plastic deformation without fracture. This behavior is one of the things which make most plastics useful.[3]

Although being a thermodynamic parameter the glass transition temperature parametrically depends on the melt cooling rate so that the slower melt cooling rate the lower Tg. In addition, Tg depends on the measurement conditions, which are not universally defined.[4]

The bond system of an amorphous material changes its Hausdorff dimension from Euclidian 3 below Tg (where the amorphous material is solid), to fractal 2.55±0.05 above Tg, where the amorphous material is liquid.[5]

Contents

Time dependency

Consider a molecular liquid which is slowly cooling down. At a certain temperature, the average kinetic energy of molecules no longer exceeds the binding energy between neighboring molecules and growth of organized solid crystal begins. Formation of an ordered system takes a certain amount of time since the molecules must move from their current location to energetically preferred points at crystal nodes. As temperature falls, molecular motion slows down further and, if the cooling rate is fast enough, molecules never reach their destination — the substance enters into dynamic arrest and a disordered, glassy solid (or supercooled liquid) forms. In fact, Walter Kauzmann has argued that if such an arrest did not happen, at still lower temperatures a thermodynamically paradoxical situation would arise, where the undercooled liquid would have to be denser and of a lower enthalpy than the crystalline phase. Such arrest apparently takes place at certain temperature, which is called the calorimetric ideal glass transition temperature T0c. This means that 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.[6] The glass transition temperature TgT0c as dTdt → 0.

A full discussion of Tg requires an understanding of mechanical loss mechanisms (vibrational and resonance modes) of specific (usually common in a given material) functional groups and molecular arrangements. Factors such as heat treatment and molecular re-arrangement, vacancies, induced strain and other factors affecting the condition of a material may have an effect on Tg ranging from the subtle to the dramatic. Tg is dependent on the viscoelastic materials properties, and so varies with rate of applied load. The silicone toy 'Silly Putty' is a good example of this: pull slowly and it flows; hit it with a hammer and it shatters.

In contrast to the melting points of crystalline materials the glass transition temperature is therefore somewhat dependent on the time-scale of the imposed change. To some extent time and temperature are interchangeable quantities when dealing with glasses, a fact often expressed in the time-temperature superposition principle. An alternative way to discuss the same issue is to say that a glass transition temperature is only truly a point on the temperature scale if the change is imposed at one particular frequency. This is why the ability to modulate the temperature in a DSC experiment has made determining Tg considerably more precise. Since Tg is cooling-rate (or frequency) dependent as the glass is formed, the glass transition is not considered a true thermodynamic phase transition by many in the field. They reserve this epithet rather for a transition that is sharp and history-independent.

 

The IUPAC Compendium of Chemical Terminology, 1997, 66, 583 defines the glass transition as a second order phase transition in which a supercooled melt yields, on cooling, a glassy structure and properties similar to those of crystalline materials e.g. of an isotropic solid material. Phase transitions are associated with the symmetry breaking[8]. The translation-rotation symmetry in the distribution of atoms and molecules is unchanged at the liquid-glass transition, which retains the topological disorder of fluids. Symmetry changes at glass transition can be viewed when considered not for atoms but for bonds. The disordered material changes its symmetry, namely the Hausdorff dimension of bonds, from Euclidian 3D below to fractal 2.55±0.05- dimensional above the glass transition temperature.[9]

In polymers, 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. With thermoplastics, the stiffness of the material will drop due to this effect. This is shown in the figure below. It can be seen that when the glass temperature has been reached, the stiffness stays the same for a while, until the material melts. This region is called the rubber plateau.

 

Tg can be significantly decreased by addition of plasticisers 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 used to modify interior plastics (e.g., dashboards) to keep them from cracking in the cold, 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 which 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.

In glasses (including amorphous metals and gels), Tg is related to the energy required to break and re-form covalent bonds in a somewhat less than perfect (may be regarded as an understatement) 3D lattice of covalent bonds. The Tg is therefore influenced by the chemistry of the glass. E.g., add B, Na, K or Ca to a silica glass, which have a valency less than 4 and they help break up the 3D lattice and reduce the Tg. Add P which has a valency of 5 and it helps re-establish the 3D lattice, increasing Tg.

The Space Shuttle Challenger disaster was caused by rubber O-rings that were below their glass transition temperature on an unusually cold Florida morning, and thus could not flex adequately to form proper seals between sections of the two solid-fuel rocket boosters.

Measurement of Tg for glasses

 

 

In contrast to the viscosity the thermal expansion, heat capacity, and many other properties of inorganic glasses show a relatively sudden change at the glass transition temperature. This effect is used for measurement by Differential scanning calorimetry (DSC) and dilatometry.

The viscosity at the glass transition temperature depends on the sample preparation (especially the cooling curve), the heating or cooling curve during measurement and the chemical composition.[4] In general, the glass transition temperature is close to the annealing point of glasses at 1013 poise = 1012 Pa·s. For dilatometric measurements heating rates of 3-5 K/min are common, for DSC measurements 10 K/min, considering that the heating rate during measurement should equal the cooling rate during sample preparation.[4]

Biophysics

Proteins also possess a glass transition temperature below which both anharmonic motions and long-range correlated motion within a single molecule are quenched. The origin of this transition is primarily due to "caging" by glassy water[11], but can also be modeled in the absence of explicit water molecules, suggesting that part of the transition is due to internal protein dynamics.[12]

Vitrification (glass formation below the melting point) can occur when starting with a liquid such as water, usually through very rapid cooling or the introduction of agents that suppress the formation of ice crystals. This is in contrast to ordinary freezing which results in ice crystal formation. Additives used in cryobiology or produced naturally by organisms living in polar regions are called cryoprotectants. Vitrification technology is being used to cryopreserve cells, tissues and organs for transplantation.

Glass transition temperature of some materials

PolymerTg (°C)
Polyethylene (LDPE)−105 or −30 also cited
Polypropylene (atactic)−20
Poly(vinyl acetate) (PVAc)28
Polyethylene terephthalate (PET)69
Poly(vinyl alcohol) (PVA)85
Poly(vinyl chloride) (PVC)81
Polystyrene95
Polypropylene (isotactic)0
Poly-3-hydroxybutyrate (PHB)0
Poly(methylmethacrylate) (atactic)105
Chalcogenide AsGeSeTe245
ZBLAN235
Tellurite279
Avatrel; Polynorbornene215
Fluoroaluminate400
Tyre Rubber160
Soda-lime glass520-600
Quartz glass1175

These are only mean values, as the glass transition temperature depends on the cooling-ratio, molecular weight distribution and could be influenced by additives.

Note also that for a semi-crystalline material such as Polyethylene that is 60-80% crystalline at room temperature the quoted glass transition refers to what happens to the amorphous part of the material as the temperature is dropped

References and footnotes

  1. ^ The IUPAC Compendium of Chemical Terminology, 66, 583 (1997).
  2. ^ A. Varshneya. Fundamentals of inorganic glasses. Boston, Academic Press (1994).
  3. ^ But such behavior is not exhibited by crosslinked thermosetting plastics which, once cured, are set for life and will shatter rather than deform, never becoming plastic again when heated, nor melting.
  4. ^ a b c O. V. Mazurin, Yu. V. Gankin: "Glass transition temperature: problems of measurements and analysis of the existing data"; Proceedings, International Congress on Glass, July 1-6, 2007, Strasbourg, France.
  5. ^ M.I. Ojovan, W.E. Lee. J. Phys.: Condensed Matter, 18, 11507-11520 (2006). http://eprints.whiterose.ac.uk/1958/
  6. ^ Baeurle SA, Hotta A, Gusev AA (2006). "On the glassy state of multiphase and pure polymer materials". Polymer 47: 6243-6253. doi:10.1016/j.polymer.2006.05.076.
  7. ^ M. I. Ozhovan (November 2006). "Topological characteristics of bonds in SiO2 and GeO2 oxide systems upon a glass-liquid transition". Journal of Experimental and Theoretical Physics 103 (5): 819-829. doi:10.1134/S1063776106110197.
  8. ^ A.Z. Patashinskioe and V.L. Pokrovskioe, Fluctuation Theory of Phase Transitions (Pergamon, Oxford, 1979).
  9. ^ M. Ojovan and W. Lee (2006). "Topologically disordered systems at the glass transition". Journal of Physics: Condensed Matter 18 (50): 11507. doi:10.1088/0953-8984/18/50/007.
  10. ^ a b Tg measurement of glasses
  11. ^ Vitkup D, Ringe D, Petsko GA, Karplus M (2001). "Solvent mobility and the protein 'glass' transition". Nature Structural Biology 7: 34–38. Entrez PubMed 10625424
  12. ^ Salsbury FR, Han WG, Noodleman L, Brooks CL (2003). "Temperature-dependent behavior of protein-chromophore interactions: A theoretical study of a blue fluorescent antibody". CHEMPHYSCHEM 4: 848–855. Entrez PubMed 12961983
  • For glass transition temperatures of various resins, see Engineered Materials Handbook—Desk edition. (1995). ASM International. ISBN 0871702835. p. 369.
  • For glass transition temperatures of various glasses, see Mazurin, O.V. Handbook of Glass Data. (1993). Elsevier. ISBN 0444816356.
  • Prediction of high weight polymers glass transition temperature using RBF neural networks Journal of Molecular Structure: THEOCHEM, Volume 716, Issues 1-3, 7 March 2005, Pages 193-198 Antreas Afantitis, Georgia Melagraki, Kalliopi Makridima, Alex Alexandridis, Haralambos Sarimveis and Olga Iglessi-Markopoulou
 
This article is licensed under the GNU Free Documentation License. It uses material from the Wikipedia article "Glass_transition_temperature". A list of authors is available in Wikipedia.
Your browser is not current. Microsoft Internet Explorer 6.0 does not support some functions on Chemie.DE