Sie sind auf Seite 1von 8

Innovative Food Science and Emerging Technologies 5 (2004) 127134

Molecular mobility around the glass transition temperature: a mini review


G. Roudaut*, D. Simatos, D. Champion, E. Contreras-Lopez, M. Le Meste
Department of molecular and sensory engineering of foods and pharmaceuticals ENSBANA 1, Esplanade Erasme. F 21000 Dijon, France Received 11 January 2003; accepted 8 December 2003

Abstract This paper consists of a non-exhaustive review of molecular mobility in relation to the glass transition, based on both synthetic polymer and food science literature. The glass transition phenomenon is a concept originally developed for synthetic polymers, which has been applied to many food products particularly with a view to predict their stability over time or temperature or water content. The glass transition is considered to be the point at which the material changes from having a solid-like behaviour to a more malleable or liquid-like. This paper proposes an assessment of the different levels of mobility existing below and above the glass transition temperature, with some examples of their consequences on the physical properties of the products. 2004 Elsevier Ltd. All rights reserved.
Keywords: Glass transition; Mobility; Relaxation; Diffusion; Food; Stability Industrial relevance: In addition to extrusion technology the concept of glass transition is one of the most important aspects for food science and technology transfered from the area of synthetic polymers. This mini review summarizes the state of the art and more importantly provides research needs which are clearly in the area of complex products and the interactions of multiple food components.

1. Introduction The knowledge and understanding of dynamical properties of food components are so far quite limited. Indeed although detailed molecular dynamics studies are available for chemically pure or homogeneous materials such as proteins, nucleic acids, or polysaccharides, very few studies have been focused on food systems. Extrapolating the existing data onto the latter should be done cautiously for them and often ignore the heterogeneous nature of food products as well as their interactions with water. The term molecular mobility covers several concepts related to the stability of food products upon storage or processability. It encompasses different types of motions such as: molecular displacement or deformation resulting from a solvent migration or mechanical strain; migration of solvent or solute molecules caused by a chemical potential gradient, or electric field;
*Corresponding author. Tel.: q33-3-80-39-66-58; fax: q33-3-8039-66-47. E-mail address: gaelle.roudaut@u-bourgogne.fr (G. Roudaut). 1466-8564/04/$ - see front matter doi:10.1016/j.ifset.2003.12.003

molecular diffusion reflecting Brownian movements; and rotation of atoms groups or polymeric segments around covalent bonds On the one hand, molecular mobility covers the displacement of reactants towards each other within a product. This displacement might cause crystallisation processes or both chemical and enzymatic reactions, which are of particular interest for the food scientist when the reactions are degradative. However, molecular mobility also relates to the viscosity of the material which in turn controls the flow properties, structure collapse, mechanical properties, and thus, the texture of the product considered. Generally, molecular mobility is mainly driven by hydration and temperature. Indeed water content and interactions between water and other food ingredients control both thermodynamic and dynamic properties of all aqueous phase elements. Moreover, the higher the temperature, the higher is the molecular mobility; such a mobility increase can be enhanced by physical state changes (glass transition, crystal melting).

2004 Elsevier Ltd. All rights reserved.

128

G. Roudaut et al. / Innovative Food Science and Emerging Technologies 5 (2004) 127134

The glass transition concept was introduced to the food science and technology world more than 30 years ago as a controlling factor in the stability of low moisture food products (White & Cakebread, 1966). The application of such a concept to food products required some differences from the synthetic polymer features to be considered. Food products generally exhibit chemical heterogeneity due to their complex nature, moreover, they are particularly sensitive to water. The glass transition was seen relevant to food technology as many dynamic processes are known to result in amorphous materials. As an example, the thermal processing of cereal products causes the loss of the semi-crystalline structure of the biopolymer and results in amorphous matrices. Depending on the cooling or dehydration rate, the products may evolve towards a glassy state, when cooled or dehydrated rapidly, or a crystalline state when cooled slowly. Similarly, this situation is also encountered in the confectionery industry, when melted sugars are rapidly cooled, or in the ice cream making process when the samples are stored at very low temperatures. 1.1. The glass transition: a transition between amorphous solid and rubberyviscous liquid The glass transition relates to the phenomena observed when a supercooled, malleable liquid or rubbery material is changed into a disordered solid glass upon cooling, or conversely when a brittle glass is changed upon heating into a supercooled liquid or a rubbery material. The transition is a kinetic process, the temperature at which the system vitrifies depends on the cooling rate: for a given material, cooling slowly creates a glass at a lower temperature than a fast cooling would do. It is generally associated with the main relaxation (a), which corresponds to a change in dynamic properties. When a material is cooled through its transition, the large amplitude cooperative movements, agitating the material above Tg, stop and the viscosity of the material reaches 1012 Pa s. The glass transition occurs over a temperature range, the width of which is controlled by the heterogeneity of the system: the greater the heterogeneity the wider the transition range (Ferry, 1980; Bosma, Ten, Brinke & Ellis, 1988). In food systems, the glass transition temperature is mainly affected by the water content (Orford, Parker, Ring & Smith, 1989; Blond, 1994; Zeleznak & Hoseney, 1987) and the average molecular weight (Orford, Parker & Ring, 1990). The glass to rubber or liquid transition is accompanied by rather abrupt changes in the physical properties of the material with temperature, e.g. increase in entropy, heat capacity, and volume and a decrease in both rigidity and viscosity. These changes in physical properties can be used to determine the glass transition temperature (Tg ) of the material.

Classically, the value quoted tends to be the temperature of the onset, midpoint or end of the heat capacity change observed by differential scanning calorimetry (DSC). However, due to the kinetic character of the transition, the Tg value should always be provided with the experimental conditions (heatingycooling rates) (Champion, Le Meste & Simatos, 2000). Such a phenomenon is primarily controlled by a change in the mobility, therefore the text will be further structured around the motions both below and above the glass transition temperature. 1.2. Motions in the glass A glass is made of both dense and less dense regions the latter being called defects (Perez, 1990), heterogeneities (Ediger, Inoue, Cicerone & Blackburn, 1996) or density fluctuations (Stillinger & Hodgdon, 1994), representing mobility islands in which localised motions of the polymer or translational motions of entrapped gas or small molecules remain possible (Chan, Pathmanthan & Johari, 1986). Relaxations studies are classically used to probe molecular motions in the glassy state, dynamical mechanical thermal analysis or dielectric spectroscopy are among the most employed methods. Within the glassy state, however, mobility is mainly local, restricted to vibrations of atoms or bonds, reorientation of small groups of atoms; and thus do not directly involve the surrounding atoms or molecules (Chan et al., 1986). Sub-Tg relaxations are named according to their position relative to the main relaxation a. Even if several secondary relaxations (d, g,) can be observed for biopolymers, only b relaxations have been thoroughly studied, but even if their origin is still a matter of debate. The temperature dependence of dynamic processes in the glassy state have been shown to obey an Arrhenius behaviour, with a characteristic apparent activation energy (Ea) reflecting the degree of cooperativity (or non) of the movement (Starkweather, 1990; Starkweather & Avakian, 1989). Due to the non-cooperative character of the sub-Tg motions, their Ea values are rather low (4050 kJymol) (Perez, 1994; Montes & Cavaille, 1999; Montes, Mazeau & Cavaille, 1998). b relaxations have been reported for starch components (Gidley, Cooke & Ward-Smith, 1993; Roudaut, Maglione & Le Meste, 1999; Borde, Bizot, Vigier & Buleon, 2002) (Fig. 1), and various polysaccharides such as dextran, pullulan etc (Scandola, Ceccorulli & Pizzoli, 1991Montes et al., 1998, 1999). A secondary relaxation (Eas50 kJymol) has been observed between y60 and y40 8C (depending on the frequency) with dielectric spectroscopy in both dry white bread and in gelatinised dry starch, but not in gluten (Roudaut et al., 1999). It was therefore associated with the carbohydrate ingredients (sucrose and starch) of bread. Moreover, in starch

G. Roudaut et al. / Innovative Food Science and Emerging Technologies 5 (2004) 127134

129

Fig. 1. Dielectric spectrum of dry bread measured at (=) 5 kHz, (*) 10 kHz, (q) 20 kHz and (s) 50 kHz.

sucrose mixtures, the amplitude of the secondary relaxation exhibited sensitivity to sucrose concentration (Roudaut et al., 1999). In polysaccharides, the b process has been linked with the rotation of lateral groups (b relaxation at low temperature) or with local conformation changes of the main chain (b relaxation at Tb)Tg) (Scandola et al., 1991). The Ea value range for a b relaxation in sugars (maltose or glucose) is between 40 and 70 kJymol (Noel, Parker & Ring, 1996). Together with the b relaxation, secondary relaxations can be represented on mobility maps which report log (relaxation frequency) vs. reciprocal values of absolute temperature (Roudaut et al., 1999). This can be used to

characterise the different motions taking place in a material in a given temperature and frequency range. Thus, in the case of dry crispy bread (Fig. 2), several types of motions can be pointed out: a first one occurring at approximately y60 8C and then others at higher temperature, which may be associated to the glass transition. Several cereal-based products have exhibited a drastic change in both mechanical properties and texture associated with a loss of crispness, with increasing hydration, even when the samples were still in the glassy state (Hutchinson, Mantle & Smith, 1989; Attenburrow, Davies, Goodband & Ingman, 1992; Kaletung & Breslauer, 1993; Le Meste, Roudaut & Davidou, 1996;

Fig. 2. Relaxation map for dry bread.

130

G. Roudaut et al. / Innovative Food Science and Emerging Technologies 5 (2004) 127134

Fig. 3. Evolution of mechanical properties of extruded starch at 11% water stored at 22 8C (after Shogren (1992)).

Nicholls, Appelqvist, Davies, Ingman & Lillford, 1995; Li, Kloeppel & Hsieh, 1998). In synthetic polymers, such a brittle-ductile transition preceding the glass transition has been attributed to secondary relaxations (Wu, 1992), therefore the previously described sub-Tg motions were suggested as responsible for such texture changes. Published works on the influence of hydration on these secondary relaxations have not been yet reached a clear agreement. Depending on the cooling conditions, a glassy material may not be formed into its most stable state, the values of enthalpy, entropy and volume being above those corresponding to the most stable glass. It is therefore described as out of equilibrium. Hence, if there is enough molecular mobility, these characteristics will tend to evolve towards their equilibrium values, a process which is called physical ageing. This ability to evolve towards a lower energy level is facilitated by a storage at temperature close to Tg. It is generally assumed that below the temperature of secondary b relaxations, the mobility is so limited that such a phenomenon cannot take place on an experimental or practical timescale (Struik, 1987). As a result of this microstructural change, the more compact molecular organisation and the strengthened interactions induce changes in both mechanical through an increased rigidity (Struik, 1987; Lourdin, Colonna, Brownsey, Noel & Ring, 2002) and in transport properties through a decreased diffusivity (Tiemblo, Guzman, Riande, Mijangos & Reinecke, 2001). Most of the studies investigating physical ageing report calorimetric or mechanical data obtained with synthetic polymers. However, there are some interesting data obtained on model foods. Fig. 3 shows the effect of ageing on the Youngs modulus of glassy extruded starch, equilibrated at 11% water and stored at 22 8C (Shogren, 1992). The Young modulus, associated with the rigidity of the system, increases with storage time. Similarly, for the same products, the elongation exhibits an important decrease from 12 to 3% after 7 days of storage. The effect of water observed on the extent of physical ageing is related to the plasticizing effect of water on Tg. Increasing the mois-

ture content shifts Tg to a lower value, thus for samples stored at the same temperature and at a smaller wTyTgx, the structural relaxation is facilitated in glasses with increasing moisture content (Perera, 2002). In addition to these motions, if the glass consists of a mixture of both polymer and smaller solutes or gas, the latter have been observed to exhibit translational motion within the glassy material. Translational diffusion of gas (e.g. nitrogen, oxygen, carbon dioxide) in a glassy matrix remains significant: between 10y10 and 10y16 m2 sy1) (Tiemblo et al., 2001, Schoonman, Ubbink, Bisperink, Le Meste & Karel, 2002), but due to both the important influence of the sample structure on the diffusion coefficient and the difficulty of measuring at low water content, the literature data on gas diffusion show large variations. Similarly, water remains mobile even within glassy systems (Ablett, Darke, Izzard & Lillford, 1993; Roudaut, Maglione, van Duschotten & Le Meste, 1999; Hemminga and van der Berg, 1999). Akele, Thominette, Paris, Pays and Verdu (1996) describe the sorption of water in glassy polycarbonate as controlled by secondary relaxations. 1.3. Molecular mobility above Tg As far as macromolecules are concerned, the glass transition, or the associated a relaxation, is accompanied by segmental motions involving several tens of carbon atoms. These motions are cooperative (with an apparent activation energy approx. 300500 kJymol). Their temperature dependence can no longer be accurately described by an Arrhenius law; indeed in the temperature region between Tg and Tgq100, the Williams Landel Ferry (WLF) expression better reflects the evolution of the physical properties with temperature. Applied to viscosity, the expression reads as follow: yC1TyTg. logh s log hg C2qTyTg (1)

where h and hg are the viscosities at T and Tg, respectively; hg, C1 and C2 are fitting parameters. C1 and C2 values fluctuate slightly around their universal values (C1s17.4 and C2s51.6) given by Williams, Landel and Ferry (1955) depending on the material considered (Ferry, 1980). Above Tg, the amplitude, frequency, or rate of the movements, as well as their sensitivity to temperature increases (Fig. 2). This marked change in the molecular chain flexibility is accompanied by important changes in the rheological behaviour of the material (such as deformability, fracture mechanism or flow behaviour). Processes such as collapse of porous matrices or stickiness are consequences of the glass transition (Levi and Karel, 1995).

G. Roudaut et al. / Innovative Food Science and Emerging Technologies 5 (2004) 127134

131

Fig. 4. Translational diffusion coefficients of fluorescein in sucrose water mixtures as function of (TyTg ), Tg being calculated from equation1 for different concentrations: (e) for 30% sucrose, (n) 43.5%, (=) 57.5%, (s) 65.3%, (h) 84% (*) 87%, (y) 90%.The line represents the WLF equation fitted to the experimental data for the concentration below 85% (C1 gs17.4, C2 gs51.6 and DTgs8=10y24 m2 sy1).

1.4. Translational diffusion around Tg In homogeneous liquid media exhibiting Brownian motions, the translational diffusion coefficient is related to the viscosity of the diffusion media through the Stokes Einstein expression: Dtranss kT 6phr (2)

Dtrans: Translational diffusion coefficient, ksBoltzmann constant, Tsabsolute temperature, hsviscosity of the diffusion medium, rshydrodynamic radius of the diffusing molecule. The equation would predict more than 106 years for a water molecule to diffuse at room temperature over a 1 mm distance in a glassy material. This might suggest that glassy food products should be completely stable over their normal shelf life, but this is known not to be the case. Faced with a much shorter stability of glassy food material, several studies have been investigated thoroughly the diffusion properties of small molecules around Tg (Champion, Hervet, Blond & Simatos, 1995; Parker & Ring, 1995a; Champion, Hervet, Blond, Le Meste & Simatos, 1997; Gunning, Parker & Ring, 2000). Depending on the water content range, the translational diffusion coefficient of fluorescein (size similar to the sucrose molecule) was collected with the concentration profile method in media containing ice or with the fluorescence recovery after photobleaching method (FRAP) for light transmitting media (Fig. 4). The experimental data (measured Dtrans) were compared to

calculated data combining both WLF and Stokes Einstein equations to predict the viscosity and the diffusion coefficient, respectively. The fit describes satisfactorily the translational diffusion of fluorescein in sucrose solutions at T)1.2 Tg. Indeed, above wTgq20x, there is a good agreement between the experimental diffusion coefficients and the values predicted from viscosity. By extrapolation of the fit, D value at Tg is 8=10y24 m2 sy1, but experimentally below wTgq20x, there is a marked decoupling between the model extrapolated from viscosity and the experimental translational diffusion. The diffusion values measured are much higher than those predicted by the model. This decoupling between diffusion and viscosity has been observed by several authors (Fujara, Geil, Sillescu & Fleischer, 1992; Gunning et al., 2000, Le Meste, Champion, Roudaut, Contreras-Lopez, Blond & Simatos, 1999; Blackburn, Wang & Ediger, 1996) and probably reflects a change in the diffusion mechanism approximately 1.2 Tg. Around Tg, diffusion would no longer be controlled by viscosity, but possibly by local motions rather than collective motions (Le Meste et al., 1999; Contreras-Lopez, Champion, Hervet, Blond & Le Meste, 2000). One has to question the validity of the relation Eq. (2), especially since the macroscopic viscosity of the material may not represent the microviscosity of the diffusing medium, and the coupling existing between the diffusing molecule and the medium depends on the composition of the latter (i.e. water content). The relation between viscosity and translational diffusion has been investigated through the effect of polymer (dextran, pullulan, and gum arabic) addition to a 57.5% (wyw) sucrose solution (Contreras-Lopez, Champion, Hervet, Blond & Le Meste, 2000). At 1% wyw, the diffusion coefficient of fluorescein was not significantly affected by the addition of polymer. With 10% polymer, the effect was variable depending on the polymer used: the Dtrans values determined for dextran, pullulan and gum arabic were, respectively, equal, lower and greater than the D values of a sucrose solution at the same dry matter content (Fig. 5). Within the concentrations studied, the biopolymers do not appear to hinder the diffusion of the small probe such as fluorescein; suggesting that the main factors controlling the solute diffusion are the dry matter content and the temperature. Moreover, it is likely that the viscosity of the polymer affects the viscosity of the solution at a macroscopic level, i.e. creating a network-like structure within which fluorescein would diffuse. 1.5. Enzymatic reactions in frozen systems Enzymatic reactions are known to be strongly dependent on temperature, but if reactions can be slowed down by low temperatures, they are generally not completely inhibited. The reaction rate has been considered to be

132

G. Roudaut et al. / Innovative Food Science and Emerging Technologies 5 (2004) 127134

Fig. 5. Translational diffusion of fluorescein in solutions with 57.5% sucrose and 1 or 10% polysaccharide: dextran (h) 104, (s) 4=104,(y) 5=105, (n) 2=106, (=) gum arabic and (q) pullulan) and in reference solutions: (e) Sucrose 58.5% (or 67.5%). Lines are guide for samples, which are not significantly different (as0.01).

diffusion-controlled (Kerr, Lim, Reid & Chen, 1993) particularly around Tg9 or Tg, where viscosity reaches a very high value, resulting in a limitation of the reaction rate by the mobility of either substrate, enzyme, or enzyme segments (Karel, Buera & Roos, 1993; Champion, Blond & Simatos, 1997; Champion et al., 2000). We have studied this, investigating the hydrolysis of di-Sodium-p-nitrophenyl phosphate catalysed by the alkaline phosphatase (Champion, Blond, Le Meste & Simatos, 2000). Reaction rates were measured in sucrose solutions with different initial concentrations: from 30 to 57.5%

between y24 and 20 8C. Even when the substrate, sucrose and enzyme concentrations are modified by cryoconcentration, the ratio enzyme to substrate remains constant. Once frozen and at a given temperature, the concentrations of the different components will be identical in the non-frozen liquid phase whatever the initial concentration, only the quantity of ice is different. The evolution of the initial diffusion-controlled reaction rate kd as a function of temperature is presented Fig. 6. In the frozen state, kd is identical for all initial sucrose concentrations (Fig. 6), whereas in the absence of ice, and at a given temperature, the reaction rate decreased with increasing initial concentration. It was verified that there is no effect of temperature on the affinity of the enzyme for the substrate. The temperature effect on the reaction rate appears to be an increase in viscosity as the temperature decreases, induced from both the WLF effect described by Eq. (1) and cryoconcentration. The viscosity increase results in a reduction of the mobility of the reacting species (enzyme flexibility and solute diffusivity) (Le Meste et al., 1999). The measured reaction rates were compared to those predicted assuming that the reaction rate followed the Atkinsmodel (Atkins, 1998). Such a model was developed for bimolecular diffusion-controlled reactions based on the StokesEinstein relation. In Atkinss theory, the reaction rate kd can be determined for equimolar reaction with same size reactants, using the viscosity of the medium as a variable: 8RT 3h (3)

kds

where h is the viscosity of the medium, T the absolute temperature.

Fig. 6. Initial reaction rate constant kd of alkaline phosphatase as a function of temperature. Symbols: experimental values obtained in different sucrose solutions: (e) 30% sucrose (wyw), (h) 43.5% and (n) 57.5%. Dashed lines: kd predicted from the measured viscosity (samples without ice). Plain lines: kd calculated from the viscosity predicted for the freeze-concentrated phase.

G. Roudaut et al. / Innovative Food Science and Emerging Technologies 5 (2004) 127134

133

The viscosity at T was predicted from the WLF equation, with C1 and C2 determined from viscosity data for sucrose solutions. The concentration of the freezeconcentrated phase and the corresponding Tg at the storage temperature T were deduced from the state diagram (Blond, Simatos, Catte, Dussap & Gros, 1997). As shown in Fig. 6, the predicted temperature dependence of the reaction rate constant was very similar to that of the experimental data. The chosen system probably represented a favourable situation for modelling according to WLF kinetics probably because the enzyme exhibited a relatively high catalytic activity even in concentrated sucrose solutions and at low temperature (Champion et al., 2000). Other enzymatic reactions have been investigated in similar conditions (Schebor, Buera & Chirife, 1996; Cardona, Schebor, Buera, Karel & Chirife, 1997; Mazzobre, Buera & Chirife, 1997; Manzocco, Nicoli, Anese, Pitotti & Maltini, 1999; Taragano & Pilosof, 2001). No absolute trend pointing towards or against a prediction of the reaction rate by the glass transition can be concluded from these studies. 2. Conclusion Although the glass transition is known to mark a change in the dynamic properties of amorphous materials, this short review showed that some molecular mobility still remains in glassy materials. Depending on the processes considered, the mobility changes related to the glass transition can be regarded as either suitable or inadequate predictors to understand the dynamics of the observed phenomena. Research on complex products is needed, for the coexistence of multiple phases, and thus of multiple dynamics, in food products requires a thorough understanding of the components behaviour, considered not only separately, but also through their different interactions. Acknowledgments This work has been supported the European Commission Contract ERBFAIRCT 1085 and was partly presented at the IFT National meeting, New Orleans in June 2001. References
Ablett, S., Darke, A. H., Izzard, M. J., & Lillford, P. J. (1993). Studies of the glass transition on malto-oligomers. In J. M. V. Blanshard, P. J. Lillford, The glassy state in foods (pp. 189 206). Nottingham: Nottingham Press. Akele, N., Thominette, F., Paris, D., Pays, M. F., & Verdu, J. (1996). Physical ageing and water sorption in polycarbonate. Journal of Materials Science Letters, 15, 1001 1002. Atkins, P. W. (1998). Molecular reaction dynamics. Physical chemistry (6th ed) (pp. 819 848). UK: Oxford University Press.

Attenburrow, G. E., Davies, A. P., Goodband, R. M., & Ingman, S. J. (1992). The fracture behaviour of starch and gluten in the glassy state. Journal of Cereal Science, 16, 1 12. Blackburn, F. R., Wang, C. Y., & Ediger, M. D. (1996). Translational and rotational probes in supercooled 1,3,5-tris(naphtyl)benzene. Journal of Physical Chemistry, 100, 18249 18257. Blond, G. (1994). Mechanical properties of frozen model solutions. Journal of Food Engineering, 22, 253 269. Blond, G., Simatos, D., Catte, M., Dussap, C. G., & Gros, J. B. (1997). Modeling of water-sucrose state diagram below 0 8C. Carbohydrate Research, 298, 139 145. Borde, B., Bizot, H., Vigier, G., & Buleon, A. (2002). Calorimetric analysis of the structural relaxation in partially hydrated amorphous polysaccharides. II. Phenomenological study of physical ageing. Carbohydrate Polymers, 48, 111 123. Bosma, M., Ten Brinke, G., & Ellis, T. S. (1988). Polymer-polymer miscibility and enthalpy relaxations. Macromolecules, 21, 1465 1470. Cardona, S., Schebor, C., Buera, M. P., Karel, M., & Chirife, J. (1997). Thermal stability of invertase in reduced-moisture amorphous matrice in relation to glassy state and trehalose crystallization. Journal of Food Science, 62, 105 112. Champion, D., Blond, G., & Simatos, D. (1997). Reaction rates at subzero-temperatures in frozen sucrose solutions: a diffusioncontrolled reaction. Cryo-Letters, 18, 251 260. Champion, D., Hervet, H., Blond, G., Le Meste, M., & Simatos, D. (1997). Translational diffusion in sucrose solutions in the vicinity of their glass transition. Journal of Physical and Chemistry B, 10, 10674 10679. Champion, D., Hervet, H., Blond, G., & Simatos, D. (1995). Comparison between two methods to measure translational diffusion of a small molecule at subzero temperature. Journal of Agricultural and Food Chemistry, 43, 2887 2891. Champion, D., Blond, G., Le Meste, M., & Simatos, D. (2000). Reaction rate modelling in cryoconcentrated solutions: alkaline phosphatase catalyzed DNPP hydrolysis. Journal of Agricultural and Food Chemistry, 48, 4942 4947. Champion, D., Le Meste, M., & Simatos, D. (2000). Towards an improved understanding of glass transition and relaxations in foods: molecular mobility in the glass transition range. Trends in Food Science and Technology, 11, 41 55. Chan, R. K., Pathmanthan, K., & Johari, G. P. (1986). Dielectric relaxations in the liquid and glassy states of glucose and its water mixtures. Journal of Physical Chemistry, 90(63), 6358 6362. Contreras-Lopez, E., Champion, D., Hervet, H., Blond, G., & LeMeste, M. (2000). Rotational and translational mobility of small molecules in sucrose plus polysaccharide solutions. Journal of Agricultural and Food Chemistry, 48, 1009 1015. Ediger, M. D., Inoue, T., Cicerone, M. T., & Blackburn, F. R. (1996). Probe rotation near and below Tg: relationship to viscoelasticity and physical aging. Macromolecules Symposium, 101, 139 146. Ferry, J. D. (1980). Viscoelastic properties of polymers. New York: John Wiley and Sons. Fujara, F., Geil, B., Sillescu, H., & Fleischer, G. (1992). Translational and rotational diffusion in supercooled orthoterphenyl close to the glass transition. Zeitschrift fur Physik B Condensed Matter, 88, 195 204. Gidley, M. J., Cooke, D., & Ward-Smith, S. (1993). Low moisture polysaccharide systems: thermal and spectroscopic aspects. In J. M. V. Blanshard, P. J. Lillford, The glassy state of foods (pp. 303 316). Nottingham, UK: Nottingham University Press. Gunning, Y. M., Parker, R., & Ring, S. G. (2000). Diffusion of short chain alcohols from amorphous maltose-water mixtures above and below their glass transition temperature. Carbohydrate Research, 329, 377 385.

134

G. Roudaut et al. / Innovative Food Science and Emerging Technologies 5 (2004) 127134 Orford, P. D., Parker, R., & Ring, S. G. (1990). Aspects of the glass transition behaviour of mixtures of carbohydrates of low molecular weight. Carbohydrate Research, 196, 11 18. Parker, R., & Ring, S. G. (1995a). Diffusion in maltose-water mixtures at temperatures close to the glass transition. Carbohydrate Research, 273, 147 155. Perera, D. Y. (2002). Effect of thermal and hygroscopic history on physical ageing of organic coatings. Progress in Organic Coatings, 44(1), 55 62. Perez, J. (1990). Quasi-punctual defects in vitreous solids and liquidglass transition. Solid State Ionics, 39, 69 79. Perez, J. (1994). Theories of liquid-glass transition. In P. Fito, A. Mulet, B. Mc Kenna, Water in foods (pp. 89 114). New York: Elsevier Applied Science. Roudaut, G., Maglione, M., & Le Meste, M. (1999). Sub-Tg relaxations in cereal-based systems. Cereal Chemistry, 76(1), 78 81. Roudaut, G., Maglione, M., van Duschotten, D., & Le Meste, M. (1999). Molecular mobility in glassy bread: a multi spectroscopic approach. Cereal Chemistry, 76(1), 70 77. Scandola, M., Ceccorulli, G., & Pizzoli, M. (1991). Molecular motions of polysaccharides in the solid state: dextran, pullulan and amylose. International Journal of Biological Macromolecules, 13, 254 260. Schebor, C., Buera, M. P., & Chirife, J. (1996). Glassy state in relation to the thermal inactivation of the enzyme invertase in amorphous dried matrices of trehalose, maltodextrin and PVP. Journal of Food Engineering, 30, 269 282. Schoonman, A., Ubbink, J., Bisperink, C, Le Meste, M., & Karel, M. (2002). Solubility and diffusion of nitrogen in maltodextriny protein tablets. Biotechnology Progress, 18, 139 154. Shogren, R. L. (1992). Effect of moisture content on the melting and subsequent physical aging of corn starch. Carbohydrate Polymers, (19), 83 90. Starkweather, H. W. (1990). Distribution of activation enthalpies in viscoelastic relations. Macromolecules, 23, 328 332. Starkweather, H. W., & Avakian, P. (1989). relaxations in phenylene polymers. Macromolecules, 22, 4060 4062. Stillinger, F. H., & Hodgdon, J. A. (1994). Translation-rotation paradox for diffusion in fragile glass-forming liquids. Physical Review: E, 50(3), 2064 2068. Struik, L. C. E. (1987). Effect of thermal history on secondary relaxation processes in amorphous polymers. Polymer, 28, 57 68. Taragano, V. M., & Pilosof, A. M. R. (2001). Calorimetric studies on dry pectinlyase preparations: impact of glass transition on inactivation kinetics. Biotechnology Progress, 17, 775 777. Tiemblo, P., Guzman, J., Riande, E., Mijangos, C., & Reinecke, H. (2001). Effect of physical aging on the gas transport properties of PVC and PVC modified with pyridine groups. Polymer, 42(11), 4817 4824. White, G. W., & Cakebread, S. H. (1966). The glassy state in certain sugar containing food products. Journal of Food Technology, 1, 73 82. Williams, M. L., Landel, R. F., & Ferry, J. D. (1955). The temperature dependence of relaxation mechanisms in amorphous polymers and other glass forming liquids. Journal of the American Chemical Society, 77, 3700 3706. Wu, S. (1992). Control of intrinsic brittleness and toughness of polymers and blends by chemical structure: A review. Polymer International, 29, 229 247. Zeleznak, K. J., & Hoseney, R. C. (1987). The glass transition in starch. Cereal Chemistry, 64, 121 124.

Hemminga, M. A., van der Berg, C. (1999). Molecular mobility in food components as studied by magnetic resonance spectroscopy. In Y. H. Roos, R. B. Leslie, P. J. Lillford, Water management in the design and distribution of quality foods, ISPOW 7 (pp. 225 265). Lancaster: Technomic. Hutchinson, R. J., Mantle, S. A., & Smith, A. C. (1989). The effect of moisture content in the mechanical properties of extruded food foams. Journal of Materials Science, 24, 3249 3253. Kaletung, G., & Breslauer, K. J. (1993). Glass transition of extrudates: relationship with processing-induced fragmentation and end-product attributes. Cereal Chemistry, 70(5), 548 552. Karel, M., Buera, M. P., & Roos, Y. H. (1993). Effects of the glass transitions on processing and storage. In J. M. V. Blanshard, P. J. Lillford, The glassy state in foods (pp. 13 34). Nottingham: Nottingham University Press. Kerr, W. L., Lim, M. H., Reid, D. S., & Chen, H. (1993). Chemical reaction kinetics in relation to glass transition temperatures in frozen food polymers solutions. Journal of Science of Food and Agriculture, 61, 51 56. Le Meste, M., Champion, D., Roudaut, G., Contreras-Lopez, E., Blond, G., & Simatos, D. (1999). Mobility and reactivity in low moisture and frozen foods. In Y. H. Roos, R. B. Leslie, P. J. Lillford, Water management in the design and distribution of quality foods. Isopow 7 (pp. 267 284). Lancaster: Technomic Publishing Company, Inc. Le Meste, M., Roudaut, G., & Davidou, S. (1996). Thermomechanical properties of glassy cereal foods. Journal of Thermal Analysis, 47, 1361 1376. Levi, G., & Karel, M. (1995). Volumetric shrinkage (collapse) in freeze-dried carbohydrates above their glass transition temperature. Food Research International, 28(2), 145 151. Li, Y., Kloeppel, K. M., & Hsieh, F. (1998). Texture of glassy corn cakes as funtion of moisture content. Journal of Food Science, 63(5), 869 872. Lourdin, D., Colonna, P., Brownsey, G. J., Noel, T. R., & Ring, S. G. (2002). Structural relaxation and physical ageing of starchy materials. Carbohydrate Research, 337(9), 827 833. Manzocco, L., Nicoli, M. C., Anese, M., Pitotti, A., & Maltini, E. (1999). Polyphenoloxidase and peroxidase activity in partially frozen systems with different physical properties. Food Research International, 31(5), 363 370. Mazzobre, M. F., Buera, M. P., & Chirife, J. (1997). Protective role of trehalose on thermal stability of lactase in relation to its glass and crystal forming properties and effect of delaying crystallization. Lebensmittel-Wissenschaft und-Technologie, 30, 324 329. Montes, H., Mazeau, K., & Cavaille, J. Y. (1998). The mechanical b relaxation in amorphous cellulose. Journal of Non-Crystalline Solids, 235-237, 416 421. Montes, H., & Cavaille, J. Y. (1999). Secondary dielectric relaxations in dried amorphous cellulose and dextran. Polymer, 40(10), 2649 2657. Nicholls, R. J., Appelqvist, I. A. M., Davies, A. P., Ingman, S. J., & Lillford, P. J. (1995). Glass transitions and fracture behaviour of gluten and starches within the glassy state. Journal of Cereal Science, (21), 25 36. Noel, T. R., Parker, R., & Ring, S. G. (1996). A comparative study of dielectric relaxation behaviour of glucose, maltose, and their mixtures with water in the liquid and glassy states. Carbohydrate Research, 282, 193 206. Orford, P. D., Parker, R., Ring, S. G., & Smith, A.C (1989). Effect of water as a diluent on the glass transition behaviour of maltooligosaccharides, amylose and amylopectine. International Journal of Biological Macromolecules, 11, 91 96.

Das könnte Ihnen auch gefallen