Sie sind auf Seite 1von 10

Journal of Food Engineering 109 (2012) 691700

Contents lists available at SciVerse ScienceDirect

Journal of Food Engineering


journal homepage: www.elsevier.com/locate/jfoodeng

An activated-state model for the prediction of solid-phase crystallization growth kinetics in dried lactose particles
Debolina Das , Timothy A.G. Langrish 1
School of Chemical and Biomolecular Engineering (J01), University of Sydney, NSW 2006, Australia

a r t i c l e

i n f o

a b s t r a c t
Molecular-level understanding of solid-phase crystallization growth kinetics has the potential to improve the fundamental basis for understanding this process. This understanding has been developed here as part of an activated-state model, which accounts for the effects of the moisture content and the temperature on the crystallization rate. Water-induced crystallization (WIC) at different temperatures (15 C, 25 C, 40 C) and a constant relative humidity (75%) environment has been used to analyze the changes in enthalpy, entropy and Gibbs free energy of activation for the solid-phase crystallization of lactose. WIC showed that, at higher temperatures, crystallization commences at lower moisture contents. The enthalpy and Gibbs free energy of activation increased during the crystallization process, which suggested that the binding energy needed for the formation of an activated complex increased as the moisture content decreased, making the formation of the activated complex more difcult. The entropy of activation, on the other hand, decreased with the decrease in the moisture content. From the activated rate equation, the energy of activation has been estimated to be 39 2 kJ mol1, which is similar to the literature value (40 kJ mol1) for the solid-phase crystallization of lactose. The reaction rate constant at 25 C is 1.4 104 s1, which is similar to the literature value of 1.3 104 s1. The WLF equation is inconsistent and hence unreliable in predicting the rates of crystallization at the glass-transition temperature from the analysis of experimental data at different temperatures. 2011 Elsevier Ltd. All rights reserved.

Article history: Received 14 July 2011 Received in revised form 9 November 2011 Accepted 11 November 2011 Available online 28 November 2011 Keywords: Solid-phase crystallization Activated-state model Free energy of activation Enthalpy Entropy Kinetics Lactose Reaction engineering Eyring equation WLF and Avrami equation Activated Rate Theory

1. Introduction Amorphous materials have non-aligned molecular structures, are hygroscopic and are in a higher-energy state relative to crystalline ones. Spray-dried materials are mainly amorphous (White and Cakebread, 1966) and potentially sticky (Bhandari et al., 1997). During spray drying, atomized feed in the form of ne droplets comes into contact with hot air in the drying chamber and dries within a fraction of a second, giving insufcient time for the droplets to crystallize. The high hygroscopicity, high solubility, low melting point and low glass-transition temperature of the amorphous spray-dried materials are major factors causing product stickiness, causing them to readily adhere to the walls of spray dryers, hindering product ow and stability. An amorphous material is generally in a metastable state with respect to the crystalline state and hence tends to undergo phase transformation during processing and storage. A system is said to be in a metastable state at a particular temperature if small isothermal changes or uctuations of its thermodynamic variables cause an increase in the free en Corresponding author. Tel.: +61 2 9351 5661; fax: +61 2 9351 2854.
E-mail addresses: debolina.das@sydney.edu.au (D. Das), timothy.langrish@ sydney.edu.au (T.A.G. Langrish). 1 Tel.: +61 2 9351 4568. 0260-8774/$ - see front matter 2011 Elsevier Ltd. All rights reserved. doi:10.1016/j.jfoodeng.2011.11.024

ergy, while larger deviations cause the system to reach a state of minimum thermodynamic potential with low free energy. During crystallization, an amorphous material may be considered to pass through an intermediate state, which acts as a free energy barrier and determines the transformation of the metastable amorphous state into the stable crystalline state (Kauzmann, 1948). Solid-phase crystallization kinetics has been widely studied (Takeuchi et al., 2000; Timmerman et al., 2004; Vautaz, 1988; Ibach and Kind, 2007; Roos and Karel, 1992; Jouppila and Roos, 1994; Buckton and Darcy, 1996, 1997, 1998a,b; Elamin et al., 1995; Vollenbroek et al., 2010; Paterson et al., 1997) and has been modeled using the WilliamsLandelFerry (Williams et al., 1955) and Avrami equations. The WilliamsLandelFerry (WLF) equation suggests that the rate of crystallization is related to the difference between the material temperature (T) and its glass-transition temperature (Tg). The WLF equation is empirical and is a time temperature superposition equation. The equation estimates the temperature shift factor at temperatures other than those for which the material was tested and hence is also called a temperature-shift equation. Although the equation provides a useful correlation between the temperature and rate of crystallization, the function (TTg) is not a physical driving force for the crystallization process. Also, the equation is only valid for temperatures up to 100 C greater than the glass-transition temperature (Williams

692

D. Das, T.A.G. Langrish / Journal of Food Engineering 109 (2012) 691700

et al., 1955). The universal constants used in this equation have been queried (Peleg, 1992) and are most likely to be product specic (Langrish, 2008). Roos and Karel (1992) showed that the WLF equation makes the reaction appear to be a zero-order reaction in terms of the concentration because, in this approach, the crystallization kinetics are expressed in an integrated form rather than as a function of reactant concentration. Therefore, with the above limitations, it is unclear if the WLF equation is able to predict high-temperature crystallization accurately in a quantitative manner. Another approach to modeling solid-phase crystallization is the Avrami equation, which assumes that negligibly small droplets of the stable phase nucleate from the metastable background; the droplets subsequently grow independently without substantial deformation, and the stable phase is pictured as randomly placed, freely overlapping, growing spheres (Roos and Karel, 1990). Although it has been shown, by molecular-dynamics simulations, that there is some theoretical basis for applying the Avrami equation to solid-phase crystallization (Novotny et al., 2000; Ramos et al., 1999), often considerable difculty is faced in applying this theory in practice (Kedward et al., 2000). The applicability of the Avrami equation has been debated, since this equation does not explicitly account for the effect of temperature when used as a correlation for tting data from isothermal crystallization experiments (Langrish, 2008). Roos and Karel (1992) tested the applicability of the Avrami equation by measuring the crystallization rate of amorphous lactose using water-induced crystallization and found that the equation did not t the crystallization process very well. The physical process of crystallization consists of two main steps (Tamman, 1926), nucleation and crystal growth. Both the WLF and Avrami equations do not distinguish between the two processes and instead lump them together. The current level of empiricism means that considerable experimental work is required for new materials due to the lack of physical understanding that restricts the prediction of solid-phase crystallization behavior from rst principles. A new way of quantitatively modeling the kinetics of solid-phase crystallization may be a reaction engineering approach, by viewing it as an activated-rate process, where it might be possible to quantitatively express the kinetics of crystal growth separately from nucleation in a rate form rather than an integrated one (Langrish, 2008). Both the Arrhenius and Eyring equations describe the temperature dependence of reaction rates. The Arrhenius equation is based on the empirical observation that rates of reactions increase with temperature, while the Eyring equation is based on a transitionstate model. According to the transition-state model, the reactants reach an unsteady intermediate state on the reaction pathway between reactants and products. An energy barrier exists between the reactants and the products on the reaction pathway that determines the minimum energy required for the reaction to occur. This minimum energy is known as the activation enthalpy or activation energy. The Eyring equation can be expressed as:

Moisture plays a key role (Ibach and Kind, 2007; Bianchi et al., 1991; Ando et al., 1992) in the kinetics of solid-phase crystallization, along with temperature (Thomsen et al., 2005; Langrish, 2008; Chiou et al., 2008). The activation enthalpy and entropy are directly related to the change in composition of the material, which is moisture in the case of water-induced crystallization. Chen and Xie (1997) created a drying ngerprint, a unique function for each material that expresses the relationship between the changes in activation energy for drying as a function of the moisture content. A similar approach might be taken to create a unique function that relates the activation energy for crystallization to the change in moisture content and temperature, as a type of ngerprint for the crystallization kinetics of a material. Sorption kinetics of various materials, including lactose, have been widely studied (Paterson and Bronlund, 2004; Jouppila and Roos, 1994a) and effectively modeled before (Jouppila and Roos, 1997; Haque and Roos, 2004; Omar and Roos, 2007) using a three-parameter GAB equation (Anderson, 1946; de Boer, 1995; Guggenheim, 1995) based on a multilayer adsorption model. It represents a thermodynamic model based on multi-layer condensation, and it has been successfully applied to a wide range of food and other materials (Van den Berg, 1985). In thermodynamic terms, the enthalpy of activation is related to the moisture content of the material by the energy of interaction between the monolayer and multilayer molecules. The decrease in interaction energy between monolayer and multilayer molecules causes the enthalpy to decrease during the process of crystallization and releases moisture. Crystallization of amorphous sugars releases a quantity of heat, which is dependent on the moisture content (Roos and Karel, 1991a). Since the change in particle moisture content has been successfully modeled before with the GAB equation, the corresponding energy (enthalpy) change, from energy-mass equivalence, can be modeled using the same form of the equation. Also, since the activation enthalpy and entropy follow enthalpyentropy compensation in lactose, the GAB equation can be used to model the activation entropy as well, as a function of the particle moisture content. However, the constants in the equation for modeling the activation enthalpy and entropy for the process are likely to be very different. This paper explores the potential for applying a transition-state reaction engineering approach to analyzing water-induced solid-phase crystallization in comparison with the WLF equation.

2. Materials and methods Samples of dried lactose powders have been produced using a Buchi-290 laboratory-scale spray dryer (Buchi, Flawil, Switzerland). The spray-drying conditions were kept at the following values based on previous work by Chiou et al. (2008) and Chiou and Langrish (2008) to produce highly amorphous powders: Inlet air temperature: 134 C Aspirator rate: 100% (0.0149 kg/s) Atomizer air-ow rate: 52 mm (19.9 L/min) Pump rate: 23% (0.000106 kg/s) Solution concentration: 9.09% w/w The moisture sorption behavior of the freshly spray-dried lactose samples was then determined at different temperatures, ranging from 15 C to 40 C, using water-induced crystallization (WIC). The method for WIC was to place the sample on a Petri dish and take mass measurements every minute with an analytical balance (Mettler-Toledo AB204S, four decimal gure analytical 0.0001 g) to study the change in moisture content as a function of time. A controlled humidity environment with a saturated sodium chloride solution (75% relative humidity, Winston and Bates, 1960)

ln

      k DH 1 kB DS ln T T R h R

where k is the reaction rate constant, kB is Boltzmanns constant, h is Plancks constant, DS and DH are the entropy and enthalpy of activation, respectively, T is the absolute temperature and R is the universal gas constant. A plot of ln (k/T) as a function of 1/T produces a H straight line of the formy =mx+ b, where the slope (m) is DR  S and intercept (b) is ln khB DR . The enthalpy and entropy of activation for the reaction can be calculated from the temperature and the reaction rate constant.

D. Das, T.A.G. Langrish / Journal of Food Engineering 109 (2012) 691700

693

Fig. 1. Water-induced crystallization at a constant temperature showing the sorption peak, the initial moisture content and the equilibrium moisture content.

Fig. 2. Variation of lactose moisture content with time during solid-phase crystallization at three different temperatures in water-induced crystallization with a relative humidity of 75%.

and a constant temperature (15 C, 25 C or 40 C) was used. The moisture contents of the particles were characterized specically in terms of the nal moisture content, which refers to the moisture content of the particles when they are crystalline (the nal equilibrium moisture content), and the moisture sorption percentage, which is the percentage change in weight from the sorption peak to the nal equilibrium moisture level. Fig. 1 shows the sorption peak, the initial and the nal equilibrium moisture content in a general sorption curve from a water-induced crystallization experiment. In order to model the rate process for the solid-phase crystallization of lactose, water-induced crystallization experiments were performed at three different temperatures of 15 C, 25 C and 40 C and a relative humidity of 75% using the same sample of freshly spray-dried lactose.

Crystallization is assumed to commence after the highest moisture content (peak) in Fig. 2 (Buckton and Darcy, 1995; Burnett et al., 2004). Considering that the degree of crystallinity is zero at the peak of the curve (highest moisture content), corresponding to the start of crystallization, and that the degree of crystallinity is unity at the crystalline state, corresponding to the bottom of the curve, the variation of the degree of crystallinity with time has been shown in Fig. 3. 3.2. Analysis of experimental data In order to analyze the experimental data, it is necessary to consider both the order of the crystallization process and the rate constant, which will be presented and discussed further in the following sections. 3.2.1. Reaction order Experimentally, the order of the reaction for the solid-phase crystallization of lactose can be determined from data on water-induced crystallization at different temperatures. In general, for a reaction in which only the mass of a single component changes, the reaction-rate equation can be written as follows:

3. Results and discussion 3.1. Experimental data The variations in moisture content with time during waterinduced crystallization for the same sample at three different temperatures from 15 C to 40 C have been shown in Fig. 2. It was observed that the sample mass increased due to water vapor sorption and then, after a certain period of time, the sample mass steadily decreased. This characteristic feature of mass loss has been well documented and has been previously assigned to the crystallization of amorphous lactose (Buckton and Darcy, 1995; Burnett et al., 2004; Price and Young, 2004). The lag between the increase in moisture content and the point where moisture loss is observed has been taken as the onset time for crystallization. After the crystallization is complete, the moisture content reaches a constant level. The characteristic moisture sorption peaks were 0.088 g/g, 0.070 g/g and 0.036 g/g dry mass for temperatures of 15 C, 25 C and 40 C, respectively. It was also observed that the time for crystallization decreased with an increase in the temperature. The crystallization times were 1000 min, 600 min and 350 min for temperatures of 15 C, 25 C and 40 C, respectively. Hence, this suggests that less moisture was required to overcome the activation energy barrier for crystallization at higher temperatures. Table 1 shows the comparison between the crystallization times and the moisture sorption peaks at 25 C and different relative humidities for spray-dried lactose obtained from different authors. The differences in the moisture content peaks and the crystallization times here may be mainly due to the differences in the relative humidities.

dX kXn dt

where n is the order of the reaction, [X] is the moisture content at each point in time, t is the time from the start of the reaction (min), and k is the reaction-rate constant. An implicit assumption here is that the extent of water loss is proportional to the extent of the reaction, and the rate of the reaction is the rate of change in the moisture content after the start of the crystallization process (after the peak in the moisture sorption curve). If the reaction rate (d[X]/dt) is plotted against the moisture content [X] at each point in time, then the order of the reaction can be determined from the shape of the curve. From the water-induced crystallization experiments, the solid-phase crystallization results of lactose were analyzed to assess if they followed rst-order reaction kinetics. Figs. 46 shows the results from analyzing the WIC data at 25 C using rst-order, second-order and third-order reaction kinetic models, respectively. The results show that the plot for the reaction rate against the moisture content [X] using a rst-order rate model has given a straight line, suggesting that crystallization follows rst-order kinetics. Plotting the reaction rate with the second and third-order reaction models has given curved lines, indicating that a second or third-order kinetic model is not valid for the solid-phase crystallization of lactose. The results of the analysis were same for the other temperatures (15 C and 40 C) when rst, second and

694

D. Das, T.A.G. Langrish / Journal of Food Engineering 109 (2012) 691700 Table 1 Crystallization time and moisture sorption peaks at 25 C obtained from moisture sorption proles of spray-dried lactose at different relative humidities. Authors Burnett et al. (2004) Price and Young (2004) Das and Langrish (present work) Relative humidity (%) 55 58 75 Moisture sorption peak (%) 12 11.5 7.2 Crystallization time (min) 350 360 560

Fig. 3. Changes in the degree of crystallinity as a function of time during solidphase crystallization for three different temperatures.

Fig. 6. Analysis of WIC data for the reaction rate as a function of the cube of the moisture content [X] using a third-order reaction kinetic model.

3.2.2. Reaction rate constant The statistical thermodynamic version of activated complex theory has motivated an approach in which the activation process is expressed in terms of thermodynamic functions. According to the Eyring equation for activated complex theory (Atkins and de Paula, 2002),

ln

     !   k kB DS DH 1 ln T h R R T

Fig. 4. Analysis of WIC data for the reaction rate as a function of the moisture content [X] using a rst-order reaction kinetic model.

Fig. 5. Analysis of WIC data for the reaction rate as a function of the square of the moisture content [X] using a second-order reaction kinetic model.

third-order reaction models were used. It has been reported by other authors before that solid-phase crystallization of carbohydrates follows a rst-order reaction kinetic model (Roos, 2003). In 1995, Roos found that carbohydrates are generally miscible with water and show rst-order phase transitions, including crystallization. However, Schmitt et al. (1999) used a reaction order of three to estimate the crystallization kinetics of amorphous lactose in the presence and absence of seed crystals. The results here (rst-order kinetics) are in line with the most recent ndings regarding the reaction order.

where kB is Boltzmanns constant (1.381 1023 J/K) and h is Planks constant (6.626 1034 J s), T is the temperature in K, R is the universal gas constant (8.3144 J mol1 K1), DS is the entropy of activation (J mol1 K1) and DH is the enthalpy of activation (J mol1). The reaction rate constants (k) of the WIC experiments at 15 C, 25 C and 40 C have been calculated from the initial concentration using the rst-order reaction kinetic equation (Eq. (2)). The enthalpy and entropy of activation have then been obtained from the slope and intercept of the Eyring equation (Eq. (3)) using leastsquares linear regression of ln (k/T) as a function of 1/T for the three different temperatures of 15 C, 25 C and 40 C. Each of the experiments was repeated three times to measure the uncertainties in the values of the reaction rate constant. The reaction rate constant k for each of the temperatures (15 C, 25 C and 40 C) has been found to be 8 105 3 105 s1, 1.4 104 7 105 s1 and 3 104 1 104 s1, respectively. The predicted k at 60 C has been obtained by extrapolating the line backwards to that temperature and has been found to be 7 104 s1. Table 2 shows the reaction rate constants obtained from the WIC experiments at the different temperatures. Fig. 7 shows the linear regression of Eq. (3) (ln (k/T) as a function of 1/T) with the slope and the intercept. The predicted value of ln (k/T) at 60 C has been found from the Activated Rate Theory in Fig. 7 to be 9 1 by extrapolating the best t line obtained from the experimental data for the three different temperatures. The crystallization rate constant k, has been calculated from the value of the predicted value of ln (k/T) and has been found to be (7 2.5) 104 s1 at a temperature of 60 C and a relative humidity of 75%. From the slope and the intercept of the regression line, the enthalpy of activation for solid-

D. Das, T.A.G. Langrish / Journal of Food Engineering 109 (2012) 691700 Table 2 Crystallization rate constants (k) from the WIC experiments using the Activated Rate Theory. 15 C Reaction rate constant k (s
1

695

25 C
5

40 C
4

60 C
4

) from WIC experiments using Activated Rate Theory

8 10

1.4 10

3 10

7 104 s1 (predicted)

Fig. 7. Linear regression of the Eyring equation showing the slope and the intercept for the WIC data at three different temperatures.

phase crystallization has been found to be 39 2 kJ mol1. The entropy of activation is 156 6 Jmol1 K1. Table 3 shows the standard uncertainties in the slope and intercept of the regression line and the effect of these uncertainties on the calculated enthalpy and entropy of activation. The standard uncertainty u(y) of a measurement result y is the estimated standard deviation of y. In 2007, Ibach and Kind studied the crystallization kinetics of amorphous lactose, whey-permeate and whey powders and found that the crystallization rate constant k increased with increasing temperature. They found that the activation energy for the solidphase crystallization of pure lactose at relative humidities between 70% and 80% was 40 kJ mol1, which is similar to the experimental value obtained from the WIC experiments here using the activated rate theory (39 2 kJ mol1). 3.3. Modeling the activation enthalpy and activation entropy of crystallization The GAB equation has been used to model the thermodynamic variables, including the enthalpy and entropy of activation, as a function of the moisture content during the crystallization process. The equation for the enthalpy of activation is shown in Eq. (4).

values at the peak for the WIC experiments at the three temperatures with the activation enthalpy value found experimentally (39 2 kJ mol1) discussed in the previous section. The moisture contents (X) when crystallization began in the WIC experiment (at the peak) were 0.036, 0.070 and 0.088 (g/g dry weight) for the temperatures of 40 C, 25 C and 15 C, respectively. The activation enthalpy should t all the three X (moisture content) values in Eq. (4) by using the three parameters m1, c1 and k1. Using the three watersorption records for the three temperatures, three equations were obtained as follows:

39000 2000

m1 c1 k1 0:088 1 0:088k1 1 0:088k1 0:088c1 k1 m1 c1 k1 0:07 1 0:07k1 1 0:07k1 0:07c1 k1 m1 c1 k1 0:036 1 0:046k1 1 0:046k1 0:046c1 k1

39000 2000

39000 2000

DH

m1 c1 k1 X 1 k1 X 1 k1 X c1 k1 X

Solving Eqs. (5)(7), the values of the constants m1, c1 and k1 of the enthalpy equation are 21,000 2000, 16.7 2 and 4.7 0.6, respectively. Similarly, the equation for the activation entropy, as modeled with the GAB equation, has been shown in Eq. (8), where m2, c2 and k2 are the three constants:

Here DH is the activation enthalpy; X is the moisture content of the particle; m1, c1 and k1 are the three parameters of the GAB equation. In the equation, m1 is the monolayer activation enthalpy, while c1 and k1 are parameters that depend on the temperature (Van den Berg, 1985). The values of the constants m1, c1 and k1 in the enthalpy equation can be obtained by correlating the moisture content

DS

m2 c2 k2 X 1 k2 X 1 k2 X c2 k2 X

Table 3 Standard uncertainties in enthalpy and entropy of activation. Coefcients Slope, (m) y-intercept, (c) Correlation, (r) DH (J mol1) DS (J mol1 K1) 4600 4.99 0.8 39,000 156 Standard uncertainties 230 0.77 2000 6

Here X is the moisture content of the material (g/g dry weight). The parameter m2 refers to the activation entropy of the monolayer, while c2 and k2 are parameters related to monolayer and multilayer properties (Kaymak-Ertekin and Gedik, 2004). The values of the constants m2, c2 and k2 for the entropy equation can be obtained by comparing the moisture content values at the peaks for the WIC experiments at the three temperatures with the activation entropy value found experimentally (156 6 J mol1K1). Equating the experimental entropy similarly, in Eq. (8), with the moisture content peaks for the three temperatures, three equations have been obtained:

156 6

m2 c2 k2 0:088 1 0:088k2 1 0:088k2 0:088c2 k2

696

D. Das, T.A.G. Langrish / Journal of Food Engineering 109 (2012) 691700

156 6

m2 c2 k2 0:070 1 0:072k2 1 0:072k2 0:072c2 k2 m2 c2 k2 0:036 1 0:036k2 1 0:036k2 0:036c2 k2

10

156 6

11

From the least-squares analysis of the experimental moisture contents and activation entropy for three different temperatures, the constants m2, c2 and k2 of the activation entropy equation are 95 10, 16.7 2 and 4.7 0.6, respectively. Fig. 8 shows the variation in the enthalpy of activation with the moisture content during crystallization using Eq. (4). All the results for different temperatures overlap on the same curve, suggesting that this theory accounts for these variations in enthalpy with changing temperature and moisture content. Fitting a function to these data and to this curve, the relationship between the enthalpy of activation and the moisture content can be written as,

Fig. 9. The variation in the entropy of activation for lactose crystallization as a function of moisture content.

DH

2100016:74:7X 1 4:7X1 4:7X 16:74:7X

12

The increase in the activation enthalpy (activation energy) with the decrease in the moisture content during crystallization suggested that the binding energy needed for the formation of an activated complex decreased as the moisture content decreased during the crystallization process. Fig. 9 shows the variation in the entropy of activation with the moisture content during crystallization using Eq. (8). As with Fig. 8, the consistency of the theory is suggested by all the data lying on a single smooth curve for different temperatures and moisture contents. Fitting a function to these data and to this curve, the relationship between the entropy of activation and the moisture content can be written as,

Fig. 10. The variation in the Gibbs free energy of activation for lactose crystallization as a function of moisture content for three different temperatures of 15 C, 25 C and 40 C.

DS

9516:74:7X 1 4:7X1 4:7X 16:74:7X

13

The entropy of activation decreased with the decrease in the moisture content during crystallization. This behavior is expected because entropy measures the degree of randomness or disorder in a system, and, as crystallization proceeds, the particles become drier, with more restricted molecular movement. The Gibbs free energy of activation has been calculated for each point of moisture content during crystallization using the GibbsHelmholtz equation Eq. (14).

DG DH T DS

14
Fig. 11. Enthalpyentropy compensation effect during solid-phase crystallization of dried lactose particles.

The variation in the Gibbs free energy of activation with moisture content during crystallization for different temperatures of 15 C, 25 C and 40 C has been shown in Fig. 10. From the tted function in Eqs. (12) and (13), the relationship between the Gibbs free energy of activation and the moisture content can be written as,

DG

2100016:74:7X 1 4:7X1 4:7X 16:74:7X 9516:74:7X T: 1 4:7X1 4:7X 16:74:7X

15

Given that m1/m2 = 221 K, from Eqs. (4) and (8), Eq. (15) may be simplied to give,

  T DG DH 1 221

16

Fig. 8. The variation in the enthalpy of activation for lactose crystallization as a function of moisture content.

It is observed that the free energy increases as the moisture content decreases, which suggests that the formation of the activated complex for solid-phase crystallization becomes more difcult as the reaction proceeds. Temperature does not have signicant effect on the Gibbs free energy of activation, since the

D. Das, T.A.G. Langrish / Journal of Food Engineering 109 (2012) 691700

697

three temperature curves completely overlap in Fig. 10, as would be expected from previous experience (Crofts and Wang, 1989; Sharma and First, 2009) for activated reactions. Given that temperature does not have signicant effect on the Gibbs free energy an isokinetic temperature (compensation effect) was evaluated. In Fig. 8 from Eq. (4), the enthalpy of activation increases with a decrease in the moisture content during crystallization, whereas in Fig. 9 from Eq. (8), the entropy of activation decreases with a decrease in the moisture content, showing an enthalpyentropy compensation effect during the process of lactose solid-phase crystallization. Hence, solid-phase crystallization of lactose follows enthalpyentropy compensation and the Gibbs free energy, being equal to the difference between the enthalpy of activation and the product of the temperature with the entropy of activation, is not affected by the temperature. Fig. 11 shows the enthalpyentropy compensation effect for lactose during solidphase crystallization. 3.4. Inconsistencies in the WLF equation and its outcome The WLF equation (Eq. (16)) suggests that the ratio of the time for crystallization at a particular point (hcr) to the time for crystallization at the glass-transition temperature (hg) decreases when the difference between the material temperature and its glasstransition temperature becomes sufciently high. The inverse of this ratio, r, is a measure of the impact of the particle temperature (T) and its glass-transition temperature (Tg) (Chiou et al., 2008):

Fig. 12. Inconsistency in the estimated logarithmic values of the rates of crystallization at the glass-transition temperature as a function of moisture content from the analysis of experimental data at 15 C, 25 C and 40 C.

log r log

    hcr kg 17:44T T g log 51:6 T T g hg kcr

glass-transition temperature (Tg) and the material temperature (T). The kg values for all the crystallization temperatures were compared, and it was observed that the values of kg varied widely for 15 C, 25 C and 40 C. Fig. 12 shows the inconsistent values for the rate of crystallization at the glass-transition temperature from the analysis of experimental data at the three different temperatures using the WLF equation. The rate of crystallization at the glass-transition temperature would be expected to be constant regardless of how it has been obtained, so the inconsistent values of kg obtained experimentally here illustrate the unreliability of the WLF equation. 3.5. Comparing the WLF equation and Activated Rate Theory The equation(s) for the Activated Rate Theory relates the rate of solid-phase crystallization to the moisture content and the temperature of the material and shows how the moisture content in turn changes the enthalpy and entropy of activation, thus affecting the Gibbs free energy of activation for the crystallization process. The Activated Rate Equation(s) has been summarized below:

17

The effect of the moisture content on this glass-transition temperature can be estimated using the Gordon Taylor equation Eq. (17):

Tg

w1 T g1 kw2 T g2 w1 kw2

18

Here, w1 ; T g1 and w2 ; T g2 are the weight fractions and glass-transition temperatures of the individual components, water and lactose, respectively. The glass-transition temperature of pure lactose is 101 C (Roos and Karel, 1991b) and that for pure water is 137 C (Johari et al., 1987). The curvature constant k is equal to 7.42 (Roos, 1993), which is determined empirically. The weight fraction of water (w1) is related to the moisture content, X, expressed on a dry basis, through the equation w1 = X/(1 + X). Since the moisture content varies throughout adsorption and desorption (a typical sorption curve is given in Fig. 1), the corresponding glass-transition temperature varies accordingly. To check for consistency in the WLF equation, the rate of crystallization at the glass-transition temperature (kg) estimated from the experimental data should be same for all the temperatures. The rates of crystallization (k) for three different temperatures (15 C, 25 C and 40 C) were assessed using the relation, k dh=dt where h is the degree of crystallinity at time t, assuming that the material was completely amorphous (degree of crystallinity, h = 0) at the peak of the sorption curve (Fig. 1) and one hundred percent crystalline (degree of crystallinity, h = 1) at the bottom of the sorption curve when the moisture content becomes constant with respect to time. The WLF equation can be expressed as follows:

ln

     !   k kB DS DH 1 ln T h R R T 2100016:74:7X 1 4:7X1 4:7X 16:74:7X 9516:74:7X 1 4:7X1 4:7X 16:74:7X

20

DH

21

DS

22

17:44T T g log kg logk 51:6 T T g

19

The logarithmic value of the rate of crystallization at the glasstransition temperature (log kg) was calculated from the rate of crystallization at the each local point (k) of the sorption curve, the

Eq. (3) shows the correlation between the rate of crystallization (k) with the enthalpy (DH) and entropy of activation (DS) and the temperature (T) based on the transition state model of Eyring equation. Eqs. (12) and (13) correlate the enthalpy and entropy of activation, respectively, as a function of the moisture content for solid-phase crystallization, and the values for the enthalpy and entropy of activation are specic to the solid-phase crystallization of lactose. Table 4 compares the crystallization rate constant (k) obtained from WIC experiments using the Activated Rate Theory, and the k value from the WLF equation, with that (k) of the experimental data from Haque and Roos (2005). The reaction rate constant (k) values were calculated experimentally from the WIC experiments at 15 C, 25 C and 40 C using the Activated Rate Theory, as shown in Table 2. At 60 C, k has been predicted by extrapolating the line in Fig. 7 backwards to that temperature, giving a value for k of 7 104 s1. The rates of crystallization (k) for three different temperatures (15 C, 25 C and 40 C) were assessed using the WLF equation

698

D. Das, T.A.G. Langrish / Journal of Food Engineering 109 (2012) 691700

Table 4 Comparison of crystallization rate constants from WIC experiments, using the Activated Rate Theory, with the k value from the WLF equation. 15 C Reaction rate constant k (s ) from WIC experiments using Activated Rate Theory Reaction rate constant k (s1) using WLF equation Experimental reaction rate constant k (s1) from Haque and Roos (2005)
1

25 C
5 1

40 C
4 1

60 C
4 1

8 10 s 2.5 105 s1

1.4 10 s 6 105 s1 1.3 104 s1

3.2 10 s 1.1 104 s1

7 104 s1 (predicted) 1.7 104 s1 (predicted)

Eq. (14) described earlier in Section 3.4. Using the WLF equation, the k values were found to be 2.5 105 s1, 6 105 s1 and 1.1 104 s1 at temperatures of 15 C, 25 C and 40 C, respectively. The reaction rate constant at 60 C was found to be 1:7 104 s1 by extrapolating the line to that temperature. Comparing the reaction rate constants in Table 2 for lactose solid-phase crystallization from the experimental data of Haque and Roos (2005) shows that the value of k (0.48 h1 equivalent to 1.3 104 s1) at a relative humidity of 76.1% at room temperature is close to the k value (1.4 104 s1) at 25 C and 75% RH from the Activated Rate Theory, but very different from the k value obtained using the WLF equation. The WLF equation is an empirical equation that takes into account the effects of time and temperature on the rate of crystallization, but the function (TTg) does not represent a physical driving force for the crystallization process. The WLF equation is based on the principle that the rate of crystallization varies with this function (TTg), which itself varies as crystallization proceeds, because the moisture content varies and changes the glass-transition temperature simultaneously. The equation gives inconsistent values of kg experimentally, while it is expected that the crystallization rate at the glass-transition temperature (kg) would be constant regardless of the method by which it is obtained. Unlike the WLF equation, the Activated Rate equation (based on a transition-state model) overcomes these limitations and accounts for the effects of both the temperature and moisture content (composition). Moisture content acts as a key factor in the physical process of crystallization and affects the enthalpy and entropy of activation for the process. The experimental crystallization rate constant k (1.3 104 s1) in Table 4 obtained by Haque and Roos is closely similar to the value of k obtained from the Activated Rate Theory (1.4 104 s1) at the same temperature and relative humidity but signicantly different from the k value obtained from the WLF equation 6 105 s1 , which supports the better predictive performance of the new equation. The Activated Rate equation overcomes the limitations of the WLF equation and provides a unied solution that takes into account the fundamental effects of both temperature and moisture content on the kinetics of solidphase crystallization. The Activated Rate Theory was not developed to t data to the Eyring equation since the Eyring equation itself is a universally-accepted equation that describes the temperature and thermodynamic dependence on reaction rates. We have found that experiments conducted on crystallization appear to follow the pattern of the Eyring equation. Moreover, the use of lactose for the study enabled the contribution of moisture to crystal formation to be understood, which physically corresponds to the activated state of the particles from the Eyring equation. In Fig. 2, the solid-phase crystallization behavior of lactose was monitored and plotted using Water Induced Crystallization (WIC) at three different temperatures of 15 C, 25 C and 40 C with a constant relative humidity of 75%. In all the three cases, it was observed that the plotted curves followed a similar pattern, which is a constant rise in moisture content to a peak followed by a steady decline in the moisture content until a constant value is reached, representing the formation of crystals. The experiments suggested the behavior of reaching an activated state before subsequently forming crystals. Considering these factors, it was deduced that the particles

at the peak moisture content have sufcient activation energy to initiate the process of crystallization, which starts by expelling the moisture stored within the high energy state particles as depicted in Fig. 2 of the paper. This phenomenon enables the molecules of the particles to arrange themselves from a high energy amorphous state to a low energy crystalline state by expelling moisture. So, the mechanism for the process of solid-phase crystallization in dried materials can be viewed as an Activated Rate process wherein the amorphous particles adsorbs moisture until it reaches a peak moisture content, marked by a high energy activated state, followed by the desorption of water and a steady decrease in the energy stored within the particles, which corresponds to them being arranged into a low-energy structured crystalline form until a constant moisture content is reached, that marks the completion of crystallization. Fig. 13 shows the schematic energy levels within the particles during crystallization, corresponding to the adsorption of moisture, to reach a high energy activated state at the highest moisture content, and subsequent expulsion of moisture and a decrease in the internal energy of the particles causing them to form lower energy structured crystals. The enthalpy and entropy are related to the activity of thermal energy lost and gained by the particles in a high energy activated state. The enthalpy being the total heat content, and the entropy being the measure of the randomness or disorder within the particles, both the components are Activated Rate equations directly related to the thermal energy. The moisture content, among other components, is a contributing factor to the phenomenon of thermal energy lost and gained during the process of crystallization. The process of solid-phase crystallization depends on both the temperature and moisture content, and the moisture in itself cannot dene the physical process of crystallization. The GAB equation has been used to relate the energy level of the particle to the moisture content. Thus the Activated Rate Equations dene the phenomenon of solid-phase crystallization, considering the energy level contained within the particles irrespective of their chemical composition. Hence, we may also say that the chemical composition relates the thermodynamic factors (enthalpy and entropy) to the rate of crystallization.

Fig. 13. Schematic diagram showing the energy level within the particles during crystallization.

D. Das, T.A.G. Langrish / Journal of Food Engineering 109 (2012) 691700

699

The paper deals with several industrial engineering aspects within the area of spray-drying, crystallization and food engineering. The implication of the ndings of this paper is found in various sectors of food engineering, including lactose crystallization, storage of dairy powders, production of non-sticky dairy powders, and operation of spray-dryers in the dairy industry. The basic purpose of this paper is to understand the contributing factors of the crystallization process in spray-drying, which is frequently used in the food industry. Understanding the various design components for crystallization of spray-dried food materials has been a key motivation. Previous studies done on crystallization during spray-drying (Chiou et al., 2008) including simulations and experiments has given useful insight of the process of solid-phase crystallization and understanding the phenomena more fundamentally from both thermodynamic and kinetic point of view can be useful for improving spray-drying processes for lactose and better design the processes of crystallization and drying by including the theory in the simulations described in Chiou et al. (2008) paper. The results are thoroughly technical and consist of numerical data and theory sufcient to back the ndings of this paper which the industry may use to sufciently broaden the designing of dried food materials.

4. Conclusions This Activated Rate approach, based on a transition-state model, explains the kinetics of crystallization and relates the rate of solidphase crystallization to both the temperature and moisture content (composition). Water induced crystallization (WIC) experiments have shown that, at higher temperatures, crystallization starts at lower moisture contents. The enthalpy and the Gibbs free energy of activation increased as the moisture content decreased during lactose crystallization. On the other hand, the entropy of activation decreased with decreasing moisture content. The enthalpy, entropy and the Gibbs free energy of activation overlap on the same curve (as a function of moisture content) for all the temperatures, showing that the theory accounts for the variations in enthalpy and entropy at different temperatures and moisture contents. The energy of activation for lactose crystallization has been found by the Activated Rate equation to be 39 2 kJ mol1, which is similar to the activation energy value (40 kJ mol1) found by Ibach and Kind in 2007. The entropy of activation was 156 6 J mol1 K1. The reaction rate constant (k) from the activated rate theory at 25 C is 1.4 104 s1, which is similar to that found by Haque and Roos (2005) of 1.3 104 s1, but the value from the WLF equation has been found to be very different. Since the WLF equation is a temperature-shift equation, it does not take into account the fundamental effect of moisture content on the rate of crystallization. The WLF equation is inconsistent in predicting the rates of crystallization at the glass-transition temperature from the analysis of experimental data at different temperatures, showing the unreliability of the equation.

References
Anderson, R., 1946. Modications of the BET equation. Journal of the American Chemical Society 68 (4), 686691. Ando, H., Ishii, M., Kayano, M., Ozawa, H., 1992. Effect of moisture on crystallization of theophylline in tablets. Drug Development and Industrial Pharmacy 18 (4), 453467. Atkins, P.W., de Paula, J., 2002. Molecular reaction dynamics. In: Atkins, (Ed.), Physical Chemistry, seventh ed. New York, Oxford University Press, 956961. Bhandari, B.R., Datta, N., Howes, T., 1997. Problems associated with spray drying of sugar-rich foods. Drying Technology 15 (2), 671684. Bianchi, R., Chiavacci, P., Vosa, R., Guerra, G., 1991. Effect of moisture on the crystallization behavior of PET from the quenched amorphous phase. Journal of Applied Polymer Science 43 (6), 10871089.

Buckton, G., Darcy, P., 1995. The use of gravimetric studies to assess the degree of crystallinity of predominantly crystalline powders. International Journal of Pharmaceutics 123 (2), 265271. Buckton, G., Darcy, P., 1996. Water mobility in amorphous lactose below and close to the glass transition temperature. International Journal of Pharmaceutics 136 (12), 141146. Buckton, G., Darcy, P., 1997. The inuence of heating/drying on the crystallisation of amorphous lactose after structural collapse. International Journal of Pharmaceutics 158 (2), 157164. Buckton, G., Darcy, P., 1998a. Crystallization of bulk samples of partially amorphous spray-dried lactose. Pharmaceutical development and Technology 3 (4), 503 507. Buckton, G., Darcy, P., 1998b. Quantitative assessments of powder crystallinity: estimates of heat and mass transfer to interpret isothermal microcalorimetry data. Thermochimica Acta 316, 2936. Burnett, D.J., Thielmann, F., Booth, J., 2004. Determining the critical relative humidity for moisture-induced phase transitions. International Journal of Pharmaceutics 287 (12), 123133. Chen, X.D., Xie, G.Z., 1997. Fingerprints of the drying behavior of particulate or thin layer food materials established using a reaction engineering model. Transactions of the Institution of Chemical Engineers and the Chemical Engineer Part C 75 (4), 213222. Chiou, D., Langrish, T.A.G., 2008. Stabilization of moisture sorption in spray-dried bioactive compounds by using novel bre carriers to crystallize the powders. International Journal of Food Engineering 4 (1), article 12. Chiou, D., Langrish, T.A.G., Braham, R., 2008. Partial crystallization behavior during spray drying: simulations and experiments. Drying Technology 26 (1), 2738. Crofts, A.R., Wang, Z., 1989. How rapid are the internal reactions of the ubiquinol:cytochrome c2 oxidoreductase? Photosynthesis Research 22 (1), 6987. de Boer, J., 1995. The dynamic character of adsorption. In: Rao, M., Rizvi, S. (Eds.), Engineering Properties of Food, second ed. New York, Marcel Dekker Inc., pp. 115251. Elamin, A., Sebhatu, T., Ahlneck, C., 1995. The use of amorphous model substances to study mechanically activated materials in the solid state. International Journal of Pharmaceutics 119 (1), 2536. Guggenheim, E., 1995. Applications of statistical mechanics. In: Rao, M., Rizvi, S. (Eds.), Engineering Properties of Food, second ed. Marcel Dekker Inc., New York, pp. 115251. Haque, M.K., Roos, Y.H., 2004. Water sorption and plasticization behavior of spraydried lactose/protein mixtures. Journal of Food Science 69 (8), 385391. Haque, M.K., Roos, Y.H., 2005. Crystallization and X-ray diffraction of spray-dried and freeze-dried amorphous lactose. Carbohydrate Research 340 (2), 293301. Ibach, A., Kind, M., 2007. Crystallization kinetics of amorphous lactose, whey permeate and whey powders. Carbohydrate Research 342 (10), 13571365. Johari, G.P., Hallbrucker, A., Mayer, E., 1987. The glass-liquid transition of hyperquenched water. Nature (London) 330, 552. Jouppila, K., Roos, Y.H., 1994. Water sorption properties and time-dependent phenomena of milk powders. Journal of Dairy Science 77 (7), 17981808. Jouppila, K., Roos, Y., 1997. Water sorption isotherms of freeze-dried milk products: applicability of linear and non-linear regression analysis in modelling. International Journal of Food Science and Technology 77 (6), 29072915. Kaymak-Ertekin, F., Gedik, A., 2004. Sorption isotherms and isosteric heat of sorption for grapes, apricots, apples and potatoes. Food Science and Technology 37 (4), 429438. Kauzmann, W., 1948. The nature of the glassy state and the behavior of liquids at low temperatures. Chemical Reviews 43 (2), 219256. Kedward, C.J., MacNaughtan, W., Mitchell, J.R., 2000. Crystallization kinetics of amorphous lactose as a function of moisture content using isothermal differential scanning calorimetry. Journal of Food Science 65 (2), 324328. Langrish, T.A.G., 2008. Assessing the rate of solid-phase crystallization for lactose: the effect of the difference between material and glass-transition temperatures. Food Research International 41 (6), 630636. Novotny, M.A., Rikvold, P.A., Kolesik, M., Townsley, D.M., Ramos, R.A., 2000. Simulations of metastable decay in two- and three-dimensional models with microscopic physics. Journal of Non-Crystalline Solids 274 (13), 356363. Omar, E.A.M., Roos, Y.H., 2007. Water sorption and time-dependent crystallization behavior of freeze-dried lactosesalt mixtures. International Journal of Food Science and Technology 40 (3), 520528. Paterson, A.H.J., Bronlund, J.E., ODonnell, A.M., 1997. Amorphous Lactose Determination in Lactose Powders by Adsorption. Paper presented at the CHEMECA Conference, Rotorua, New Zealand. Paterson, T., Bronlund, J., 2004. Moisture sorption isotherms for crystalline, amorphous and predominantly crystalline lactose powders. International Dairy Journal 14 (3), 247254. Peleg, M., 1992. On the use of the WLF model in polymers and foods. Critical Reviews in Food Science and Nutrition 32 (1), 5966. Price, R., Young, P.M., 2004. Visualization of the crystallization of lactose from the amorphous state. Journal of Pharmaceutical Sciences 93 (1), 155164. Ramos, R.A., Rikvold, P.A., Novotny, M.A., 1999. Test of the KolmogorovJohnson MehlAvrami picture of metastable decay in a model with microscopic dynamics. Physical Review B 59 (14), 90539069. Roos, Y.H., Karel, M., 1990. Differential scanning calorimetry study of phase transitions affecting the quality of dehydrated materials. Biotechnology Progress 6 (2), 159163.

700

D. Das, T.A.G. Langrish / Journal of Food Engineering 109 (2012) 691700 Thomsen, M.K., Lauridsen, L., Skibsted, L.H., Risbo, J., 2005. Temperature effect on lactose crystallization, maillard reactions, and lipid oxidation in whole milk powder. Journal of Agricultural and Food Chemistry 53 (18), 70827090. Timmerman, I.L., Bolzen, N., Trunk, M., Steckel, H., 2004. A calorimetric study on amorphous lactose during and after re-crystallisation at different relative humidity. Paper presented at the Respiratory Drug Delivery IX. Van den Berg, C., 1985. Development of BET-like models for sorption of water on foods: theory and relevance. In: Simatos, D., Multon, J.L. (Eds.), Properties of Water in Foods in Relation to Quality and Stability. Martinus Nijhoff Publishers, Dordrecht, pp. 119131. Vautaz, G., 1988. Preservation of skim-milk powders: role of water activity and temperature in lactose crystallisation and lysine loss. In: Seow, C.C. (Ed.), Food Preservation by Moisture Content. Elsevier Applied Science, New York. Vollenbroek, J., Hebbink, G.A., Ziffels, S., Steckel, H., 2010. Determination of low levels of amorphous content in inhalation grade lactose by moisture sorption isotherms. International Journal of Pharmaceutics 395 (12), 6270. White, G.W., Cakebread, S.H., 1966. The glassy state in certain sugarcontaining food products. Journal of Food Technology 1 (1), 7383. 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 (14), 37013707. Winston, P.W., Bates, D.H., 1960. Saturated solutions for the control of humidity in biological research. Ecology 41 (1), 232237.

Roos, Y.H., Karel, M., 1991a. Water and molecular weight effects on glass transition in amorphous carbohydrates and carbohydrate solutions. Journal of Food Science 56 (6), 16761681. Roos, Y.H., Karel, M., 1991b. Plasticizing effect of water on thermal behavior and crystallization of amorphous food models. Journal of Food Science 56 (1), 38 43. Roos, Y.H., Karel, M., 1992. Crystallization of lactose. Journal of Food Science 57 (3), 715777. Roos, Y.H., 1993. Water activity and physical state effects on amorphous food stability. Journal of Food Processing and Preservation 16 (6), 433447. Roos, Y.H., 2003. Thermal analysis, state transitions and food quality. Journal of Thermal Analysis and Calorimetry 71 (1), 197203. Schmitt, E.A., Law, D., Zhang, G.G.Z., 1999. Nucleation and crystallization kinetics of hydrated amorphous lactose above the glass transition temperature. Journal of Pharmaceutical Sciences 88 (3), 291296. Sharma, G., First, E.A., 2009. Thermodynamic analysis reveals a temperaturedependent change in the catalytic mechanism of bacillus stearothermophilus tyrosyl-trna synthetase. The Journal of Biological Chemistry 284 (7), 4179 4190. Takeuchi, H., Yasuji, T., Yamamoto, H., Kawashima, Y., 2000. Temperature and moisture-induced crystallization of amorphous lactose in composite particles with sodium alginate prepared by spray-drying. Pharmaceutical Development and Technology 5 (3), 355363. Tamman, G., 1926. The States of Aggregation. Transactions. In: Mehl, R.F. (Ed.), Van Nostrand, New York, pp. 220.

Das könnte Ihnen auch gefallen