Вы находитесь на странице: 1из 8

Glass transition or vitrification refers to the transformation of a glass-forming liquid into a glass, which usually occurs upon rapid

cooling. It is a dynamic phenomenon occurring between two distinct states of matter (liquid and glass), each with different physical properties. Upon cooling through the temperature range of glass transition (a "glass transformation range"), without forming any long-range order or significant symmetry of atomic arrangement, the liquid contracts more continuously at about the same rate as above the melting point until there is a decrease in the thermal expansion coefficient (TEC).[1] The glass transition temperature, Tg, is lower than melting temperature, Tm, due to supercooling. Tg depends on the time scale of observation which must be defined by convention. One approach is to agree on a standard cooling rate of 10 K/min. Another approach is by requiring a viscosity of 1012 Pas. Otherwise, one can only talk about a glass transformation range.[2] Contents [hide] 1 Introduction 2 Transition temperature Tg 3 Kauzmann's paradox 4 The glass transition in specific materials 4.1 Silica, SiO2 4.2 Polymers 4.3 Biomaterials 5 See also 6 References 7 Further reading 8 External links

Introduction
The glassy or vitreous state of matter is typically formed by rapid cooling and solidification from the molten (or liquid) state. If the liquid were allowed to crystallize on cooling, then according to the Ehrenfest classification of first-order phase transitions, there would be a discontinuous change in volume (and thus a discontinuity in the slope or first derivative with respect to temperature, dV/dT) at the melting point. In this context, glass and melt are distinct phases with an interfacial discontinuity having a surface of tension with a positive surface energy. Thus, a metastable parent phase is always stable with respect to the nucleation of small embryos or droplets from a daughter phase, provided it has a positive surface of tension. Such first-order transitions must proceed by the advancement of an interfacial region whose structure and properties vary discontinuously from the parent phase.[3][4][5][6] 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 versus temperature) for the glassy state and the supercooled liquid.[7][8][9][10][11]

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 (see Physics of glass). 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.[12][13][14]

Transition temperature Tg

Measurement of Tg by differential scanning calorimetry

etermination of Tg by dilatometry. 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 temperature at the intersection of the red regression lines.[15] 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 Pas). As evidenced experimentally, this value is close to the annealing point of many glasses.[16] In contrast to viscosity, the thermal expansion, heat capacity, 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. Here, heating rates of 3-5 K/min are common. Summarized below are Tg values characteristic of certain classes of materials. Material Tyre rubber Polypropylene (atactic) Poly(vinyl acetate) (PVAc) Polyethylene terephthalate (PET) Poly(vinyl alcohol) (PVA) Poly(vinyl chloride) (PVC) Polystyrene Polypropylene (isotactic) Poly-3-hydroxybutyrate (PHB) Poly(methylmethacrylate) (atactic) Poly(carbonate) Chalcogenide GeSbTe Chalcogenide AsGeSeTe ZBLAN glass Tellurium dioxide Polynorbornene Fluoroaluminate Soda-lime glass Fused quartz Tg (C) 70[17] 20[18] 30[18] 70[18] 85[18] 80[18] 95[18] 0[18] 15[18] 105[18] 145[18] 150[19] 245 235 280 215[18] 400 520-600 ~1200[20]

These are only mean values, as the glass transition temperature depends on the cooling rate, molecular weight distribution and could be influenced by additives. Note also that for a semicrystalline 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 upon cooling.

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.[21] One resolution of the Kauzmann paradox is to say that there must be a phase change 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: Tg T0c as
dT

dt 0.

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.

The glass transition 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 -quartz the bond length is 161 pm, whereas in tridymite it ranges from 154-171 pm. The Si-O-Si bond angle also varies from 140 in tridymite to 144 in -quartz to 180 in -tridymite. Any deviations from these standard parameters constitute microstructural differences or variations which 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.[22] 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[23]. The enthalpy Hd and the entropy Sd of configurons broken bonds can be found from available experimental data on viscosity [24].

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[citation needed]. 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. [25] 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 "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.[18] 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.

Stiffness versus temperature On cooling, rubber undergoes a liquid-glass transition, which has also been called a rubberglass transition. For example, the Space Shuttle Challenger disaster was caused by rubber O-

rings that were being used well 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.

Biomaterials
When sucrose is cooled slowly, the result is crystal sugar (or rock candy), but, when cooled rapidly, the result can be in the form of syrupy cotton candy (candyfloss). Vitrification can also 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. Additives used in cryobiology or produced naturally by organisms living in polar regions are called cryoprotectants. Arctic frogs and some other ectotherms naturally produce glycerol or glucose in their livers to reduce ice formation. When glucose is used as a cryoprotectant by Arctic frogs, massive amounts of glucose are released at low temperature,[26] and a special form of insulin allows for this extra glucose to enter the cells. When the frog rewarms during spring, the extra glucose must be rapidly removed from the cells and recycled via renal excretion and storage in the bladder. Arctic insects also use sugars as cryoprotectants. Arctic fish use antifreeze proteins, sometimes appended with sugars, as cryoprotectant. Vitrification technology is being used to cryopreserve cells, tissues and organs for transplantation. For years, glycerol has been used in cryobiology as a cryoprotectant for blood cells and bull sperm, allowing storage at liquid nitrogen temperatures. However, glycerol cannot be used to protect whole organs from damage. Instead, many biotechnology companies are currently[update] researching the development of other cryoprotectants more suitable for such uses. A successful discovery may eventually make possible the bulk cryogenic storage (or "banking") of transplantable human and xenobiotic organs. A substantial step in that direction has already occurred. At the July 2005 annual conference of the Society for Cryobiology,[27] Twenty-First Century Medicine announced the vitrification of a rabbit kidney to -135C with their proprietary vitrification cocktail. Upon rewarming, the kidney was successfully transplanted into a rabbit, with complete functionality and viability. In the context of cryonics, especially in preservation of the human brain, vitrification of tissue is thought to be necessary to prevent destruction of the tissue or information encoded in the brain. At present, vitrification techniques have only been applied to brains (neurovitrification) by Alcor and to the upper body by the Cryonics Institute, but research is in progress by both organizations to apply vitrification to the whole body. The scientific basis behind the cryonics is that proteins 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,[28] but can also be modeled in the absence of explicit water molecules, suggesting that part of the transition is due to internal protein dynamics.[29]

References
1. ^ Kingery, W,D., Bowen, H.K., and Uhlmann, D.R., Introduction to Ceramics, 2nd Edn. (John Wiley & Sons, New York, 2006) 2. ^ Richerson, D.W., Modern Ceramic Engineering,2nd Ed., (Marcel Dekker Inc., 1992) ISBN 0-8247-8634-3. 3. ^ Atkins, P.W., Physical Chemistry (W.H. Freeman & Co., New York, 1994) 4. ^ Hilliard, J.E. and Cahn, J.W., On the Nature of the Interface Between a Solid Metal and Its Melt, Acta Met., Vol. 6, p. 772 (1958) 5. ^ Cahn, J.W., Theory of crystal growth and interface motion in crystalline materials, Acta Met, Vol. 8, p. 554 (1960)

6. ^ Cahn, J.W., Hillig, W.B. and Sears, G.W., The molecular mechanism of solidification, Acta Met., Vol. 12, p. 1421 (1964) 7. ^ Moynihan, C. et al. in The Glass Transition and the Nature of the Glassy State, Eds. M. Goldstein and R. Simha, Ann. N.Y. Acad. Sci., Vol. 279 (1976) 8. ^ Angell, C.A., J. Phys. Chem. Solids, Vol. 49, p. 863 (1988) 9. ^ Angell, C.A. and Nagel, S.R., J. Phys. Chem., Vol. 100, p. 13200 (1996) 10.^ Angell, C.A., Science, Vol. 267, p. 1924 (1995) 11.^ Stillinger, F., Science, Vol. 267, p. 1935 (1995) 12.^ Nemilov, S.V., (1994). Thermodynamic and Kinetic Aspects of the Vitreous State. CRC Press. 13.^ Zarzycki, J. (1991). Glasses and the Vitreous State. Cambridge University Press. 14.^ J.H. Gibbs (1960). J.D. MacKenzie. ed. Modern Aspects of the Vitreous State. Butterworth. OCLC 1690554. 15.^ Tg measurement of glasses 16.^ IUPAC Compendium of Chemical http://old.iupac.org/goldbook/G02641.pdf Terminology, 66, 583 (1984),

17.^ Tyre comprising a cycloolefin polymer, tread band and elasomeric composition used therein. 03.07.2003. 18.^ a b c d e f g h i j k l Charles E. Wilkes, James W. Summers, Charles Anthony Daniels, Mark T. Berard (2005). PVC Handbook. Hanser Verlag. ISBN 1569903794. http://books.google.com/?id=YUkJNI9QYsUC&pg=PT1&lpg=PT1. 19.^ EPCOS 2007: Glass Transition and Crystallization in Phase Change Materials 20.^ Bucaro, J. A. (1974). "High-temperature Brillouin scattering in fused quartz". Journal of Applied Physics 45: 53241974. doi:10.1063/1.1663238. edit 21.^ Walter Kauzmann, The Nature of the Glassy State and the Behavior of Liquids at Low Temperatures; Chemical Reviews 43 (2), 1948.[1] 22.^ Ojovan M.I. (2008). "Configurons: thermodynamic parameters and symmetry changes at glass transition". Entropy 10: 334364. doi:10.3390/e10030334. http://www.mdpi.org/entropy/papers/e10030334.pdf. 23.^ M.I. Ojovan. Configurons: thermodynamic parameters and symmetry changes at glass transition . Entropy, 10, 334-364 (2008). http://www.mdpi.org/entropy/papers/e10030334.pdf 24.^ M.I. Ojovan, K.P. Travis, and R.J. Hand. Thermodynamic parameters of bonds in glassy materials from viscosity-temperature relationships. J. Phys.: Condensed Matter, 19, 415107, 12p (2007). 25.^ Cowie, J.M.G. and Arrighi, V., Polymers: Chemistry and Physics of Modern Materials, 3rd Edn. (CRC Press, 2007) 26.^ J.R. Layne, Jr., R.E. Lee, Jr. (1995). "Adaptations of frogs to survive freezing" (PDF). Climate Research 5: 5359. doi:10.3354/cr005053. http://www.intres.com/articles/cr/5/c005p053.pdf. 27.^ "Plenary Session: Fundamentals of Biopreservation". CRYO 2005 Scientific Program. Society for Cryobiology. July 24, 2005. Archived from the original on 2006-08-30.

http://web.archive.org/web/20060830143259/http://www.me.umn.edu/events/cryo200 5/program.html. Retrieved 2006-11-08. 28.^ D. Vitkup, D. Ringe, G.A. Petsko, M. Karplus (2001). "Solvent mobility and the protein 'glass' transition". Nature Structural Biology 7 (1): 3438. doi:10.1038/71231. PMID 10625424. Entrez Pubmed 10625424 29.^ F.R. Salsbury, W.G. Han, L. Noodleman, C.L. Brooks CL (2003). "Temperaturedependent behavior of protein-chromophore interactions: A theoretical study of a blue fluorescent antibody". Chemphyschem 4 (8): 848855. doi:10.1002/cphc.200300694. PMID 12961983. Entrez Pubmed 12961983

Вам также может понравиться