Sei sulla pagina 1di 40

4

Fat Crystal Networks


Geoffrey G. Rye, Jerrold W. Litwinenko, and Alejandro G. Marangoni
University of Guelph Guelph, Ontario, Canada

1. INTRODUCTION Fats provide fundamental structural and textural attributes to a wide range of consumer products, including lipstick, chocolate, and everyday products such as butter and margarine (1, 2). Within these fat-based products, certain textural properties are required to meet desirable sensory attributes to gain consumer acceptance (3). This has led to an increase in research efforts on the physical properties of fats, particularly their rheology. The goal of this chapter is to investigate the effects of processing conditions on the physical properties of fats, using anhydrous milkfat (AMF) as an example. The approach proposed may lead to an increased understanding of product quality and characteristics, while offering insight into future methods for the determination and prediction of the rheological and textural attributes of fat-based products.

Baileys Industrial Oil and Fat Products, Sixth Edition, Six Volume Set. Edited by Fereidoon Shahidi. Copyright # 2005 John Wiley & Sons, Inc.

121

122

FAT CRYSTAL NETWORKS

2. MECHANICAL PROPERTIES OF MILKFAT Most mechanical tests developed for fats are empirical in nature and are usually designed for quality control purposes, and they attempt to simulate consumer sensory perception (3, 4). These large-deformation tests measure hardness-related parameters, which are then compared with textural attributes evaluated by a sensory panel (3, 5). These tests include penetrometry using cone, pin, cylinder and several other geometries (3, 612), compression (13), extrusion (13, 14), spreadability (15, 16), texture prole analysis (2), shear tests (13), and sectility measurements (14). These methods are usually simple and rapid, and they require relatively inexpensive equipment (3, 4, 17). The majority of these tests are based on the breakdown of structure and usually yield single-parameter measurements such as hardness, yield stress, and spreadability, among others (4, 1720). The relationship between these mechanical tests and the structure of a fat has, however, not been established. The ultimate aim of any materials science endeavor is to examine the relationship between structure and macroscopic properties. As a starting point, we need a model that can be used to describe the formation and interaction among structural elements that affect the macroscopic properties of a fat crystal network. One such model is shown in Figure 1. This model suggests that lipid composition, directly under the inuence of the processing and storage conditions, will affect the solid fat content, polymorphism, and microstructure of a fat crystal network. The model also suggests that interactions among these

Legend Strong Influence Potential Influence Nucleation & Crystal Growth

Macroscopic Properties

Microscopic Properties

Processing Conditions Solid Fat Content Polymorphism

Storage Time Storage Conditions

Molecular/Lipid Composition
Figure 1. Hierarchical model of the factors affecting the macroscopic properties of a fat crystal network.

LIPID COMPOSITION

123

factors also play an important role, because they all ultimately affect the structure of the fat crystal network, and in turn inuence the physical properties and sensory perception of a fat. Each of the factors within the model will be examined in turn.

3. LIPID COMPOSITION 3.1. Triacylglycerols Fats are composed primarily of triacylglycerols (TAG). A TAG consists of three fatty acid residues esteried to a glycerol (three-carbon sugar alcohol) backbone at specic locations known as sn-1, sn-2, and sn-3. Fatty acid residues exist in a wide variety of forms, including short and long chain, saturated and unsaturated, odd or even carbon number, trans- or cis-, linear or branched, as well as any combination thereof (21). TAG variety in a fat system is very large, as there are potentially hundreds of different fatty acid residues and their isomers available for reaction at any of the three sn-positions.
3.1.1. Anhydrous Milkfat (AMF) Composition and Properties

AMF is heterogeneous in nature and is the most complex naturally derived fat (2224). Previous research has revealed greater than 400 different fatty acids in AMF with carbon chain lengths ranging from 4 to 24 with varying degrees of saturation and molecular arrangement (24, 25). In bovine AMF, fatty acids are primarily in the form of TAGs. TAGs make up 9698% of the total fat and are the primary component in the formation and structure of the fat crystal network (23, 25). The remaining 24% are minor components, which include monoacylglycerols, diacylglycerols, phospholipids, free fatty acids, cholesterol, and some protein (26). Milkfat complexity dramatically increases when one considers the number of combinations for 400 fatty acid species esteried to glycerol at three different bonding sites4003 (64 million) possible TAG structures are theoretically possible (24, 25). In AMF, however, there are typically only 13 fatty acids present that are in concentrations greater than 1% (w/w), and therefore, a theoretical maximum of 133 (2197) TAG isomers typically exist (25). However, if locations sn-1 and sn-3 are considered equivalent, the number of possible TAGs is reduced to 455 (25). For the purpose of investigating the effects of lipid composition on the physical properties of milkfat, we will use the fatty acid concentrations determined using gas-liquid chromatography. Table 1 shows the fatty acid compositions of milkfat and milkfat diluted using canola oil. Dilution with canola oil allows for the controlled variation in solids content. By diluting with canola oil, AMF composition is altered through the addition of the long-chain fatty acids, primarily 18:1, 18:2, and 18:3, with only small modications to the saturated fatty acid concentrations (16:0, 18:0, and 20:0). This allows for the study of the effects of lipid composition on physical properties by altering the TAG makeup using a diluent that neither

124

FAT CRYSTAL NETWORKS

TABLE 1. Fatty Acid Composition (wt%) of Anhydrous Milkfat and Dilutions of Anhydrous Milkfat in Canola Oil. Fatty Acid 4:0 6:0 8:0 10:0 12:0 14:0 14:1 15:0 15:1 16:0 16:1 17:0 17:1 18:0 18:1 18:2 18:3 20:0 22:0 100% AMF 2.76 2.18 1.39 3.16 3.76 12.34 1.92 1.42 0.35 30.37 2.93 0.91 0.48 9.48 21.92 2.79 1.31 0.52 0.00 90% AMF 2.48 1.96 1.25 2.84 3.38 11.15 1.73 1.28 0.32 27.92 2.69 0.82 0.43 8.77 25.69 4.52 2.11 0.61 0.05 80% AMF 2.21 1.74 1.11 3.01 2.53 9.96 1.54 1.14 0.28 25.46 2.44 0.73 0.38 8.05 29.46 6.24 2.92 0.71 0.09 70% AMF 1.93 1.53 0.97 2.63 2.21 8.76 1.34 0.99 0.25 23.01 2.20 0.64 0.34 7.34 33.23 7.97 3.72 0.80 0.14

signicantly solubilizes milkfat TAGs nor crystallizes in the temperature range of interest (usually above 0 C) (27). AMF complexity gives it unique textural, physical, and sensory properties. The crystallization and melting properties of AMF are directly related to its complex composition. The large variety of TAG species that exist in AMF cause it to have a wide melting range, spanning from 40 C to 40 C. Within this melting range, at least three distinct fractions have been identied and designated as lowmelting (LMF), medium-melting (MMF), and high-melting (HMF) fractions (28). These fractions are not well dened; however, the LMF typically contains shortchain (4:0 to 8:0) and long-chain unsaturated (18:1, 18:2, 18:3, 20:1) fatty acids. MMF contains medium-chain (10:0-14:0) fatty acids, and HMF contains longer chain (>16:0) fatty acids. The heterogeneous nature of the material, and its complex phase behavior, causes drastic differences in the solid nature of AMF crystal networks when storage temperature or crystallization conditions are modied. AMF lipid composition is the fundamental property affecting all of the structural characteristics of the crystallized bulk fat (Figure 1), because it will inuence nucleation, crystal growth, and the formation of the nal crystal network. There have been various attempts to modify the lipid composition of AMF as a means of producing butter with improved spreadability. These studies have included the fractionation and removal of portions of the HMF region to provide a more spreadable product at refrigeration temperatures (24). Other attempts

PROCESSING CONDITIONS

125

have included the modication of bovine diets to incorporate particular fatty acids into MF TAGs, and other work has concentrated on the chemical interesterication of different fatty acids to the TAGs of AMF (9, 2932). These methods typically affect the avor of bulk AMF, and labeling the product butter is not permitted because of the chemical adulterations performed. For this reason, processors are continually trying to modify crystal structure, and thereby improve textural characteristics solely through the modication of processing conditions.

4. PROCESSING CONDITIONS The structure of most processed food depends not only on the ingredient formulation, but also on the processing history of the material (33). Processing conditions, such as crystallization, storage temperature, cooling rate, storage time, shear, and tempering, affect the crystal structure and rheological properties of fats. In the model proposed in Figure 1, processing conditions are considered an external factor, and like lipid composition, affect the underlying physical properties of the fat crystal network. Adjusting processing parameters, such as cooling rate and/or crystallization temperature, will cause fats to exhibit differences in physical properties. Altering cooling rate has been shown to induce the formation of different polymorphs in AMF (34). This research concentrated on determining the initial polymorph present in AMF upon cooling at various controlled rates. AMF was cooled from 70 C to 65 C at different cooling rates, and polymorphism was monitored using differential scanning calorimetry (DSC) and powder X-ray diffraction. At temperatures above 0 C, the a-polymorphic form was predominant at all cooling rates above 1 C/min. At 1 C/min, a mixture of both a- and b0 -polymorphs was present, and at 0.5 C/min, only the b0 -polymorph was detectable. This indicates that the processing conditions affect the nucleation and crystal growth of AMF, as indicated by differences in polymorphism. Herrera and Hartel (35, 36) demonstrated differences in microstructural size and structure as a result of cooling rate using milkfat and milkfat blends. Cooling a sample more rapidly resulted in smaller and more numerous crystallites. Additionally, they demonstrated differences in SFC and polymorphism. Differences were induced at the nucleation and crystal growth stages. They further demonstrated the effects of cooling rate on the rheological characteristics of the system using the compression storage modulus (E0 ) as an indicator (37). They found that, under shear, the storage modulus decreased with increases in cooling rate (smaller particle sizes), therefore making a link among processing conditions, solid fat content, and microstructure and macroscopic properties (3537). These various examples depict the effects of processing conditions on the physical properties of fats and demonstrate that external crystallization conditions, such as cooling rate and storage temperature, can have dramatic effects on the nal measurable properties of a fat.

126

FAT CRYSTAL NETWORKS

5. NUCLEATION AND CRYSTAL GROWTH Crystal shape, size, and density all affect the physical properties of the nal solid fat matrix. Crystal growth, primary nucleation, and secondary nucleation in fat systems are inuenced by many factors, including diffusion, molecular compatibility, TAG structure, nuclei composition and surface properties, number of nuclei, and processing conditions (temperature and/or shear) (38, 39). It is during the crystallization process of fats that the template for the nal physical properties of the material is created. 5.1. Nucleation Mechanisms Crystallization does not take place until the melt becomes supersaturated or undercooled (38). Like all substances, fats cannot crystallize until tiny embryonic crystals, known as nuclei, are formed. Supersaturation and consequent homogeneous nucleation is unlikely in MF because MF is of a heterogeneous nature with virtually no pure TAG species in concentrations greater than 2 mol% (39, 40). In the melt, there are two primary mechanisms by which nucleation takes place. These nucleation processes, illustrated in Figure 2, are referred to as homogeneous and heterogeneous, the latter being predominant in AMF crystallization (39, 40). After nucleation, crystal growth takes place on the surface of existing nuclei (38).

Figure 2. Schematic depicts nucleation and crystal growth processes.

NUCLEATION AND CRYSTAL GROWTH

127

5.1.1. Homogeneous Nucleation Homogeneous nucleation, resulting from

bimolecular reactions between the TAG species, leads to the formation of nuclei in the absence of foreign particles (39, 40). To achieve homogeneous nucleation, no solid substrate or contaminant can exist in the fat. TAGs interact only with one another, usually at higher degrees of supercooling (in most cases are greater than 30 C below the nal melting temperature). AMF typically will not nucleate in a homogeneous fashion because of the following 1. The presence of catalytic impurities such as water, dirt, monoacylglycerols, diacylglycerols, phospholipids, and proteins. 2. The presence of a large variety of different and incompatible TAGs in concentrations of less than 1 mol% of the entire AMF volume (40). 3. The presence of agitation, temperature gradients, inducing inconsistent nucleation, and diffusion-limited gradients (39).
5.1.2. Heterogeneous Nucleation Heterogeneous nucleation is the most com-

mon process in AMF crystallization (39, 40). Nucleation is induced by the presence of foreign particles or catalytic impurities, and therefore, it requires a far lower degree of undercooling (as low as 3 C) than that required for homogeneous nucleation (39, 40). Therefore, because of the large quantity of structurally incompatible TAGs in AMF, nucleation on catalytic impurities is entropically favored over homogeneous nucleation.
5.1.3. Secondary Nucleation and Crystal Growth

Secondary nucleation is the process whereby nuclei are created as a result of inhomogeneous growth on primary crystals, or crystal breakage as a result of processing, which yield new interfaces for the creation of secondary nuclei. Secondary growth is the continued solidication of TAGs onto the surface of already existing nuclei, which allows crystals to form larger structures. During crystal growth, fat crystals take shape by forming complex spherulitic or needle-like crystal structures (41). In AMF, spherulitic crystal structures are predominant. Spherulites tend to have a very dense crystal center with decreasing density as the distance from the crystal center increases (41). Spherulites can continue to grow and aggregate to form a threedimensional network.

5.2. Factors Directly Inuenced by Nucleation and Crystal Growth


5.2.1. Solid Fat Content Many methods for measuring SFC have been devel-

oped. These include dilatometry, calorimetry, and pulsed nuclear magnetic resonance (pNMR). Dilatometry and calorimetry use measurements of volume or heat content ratios between the completely liquid and the completely solid states (42). Dilatometry and calorimetry methods are time consuming and tend to be applicable only when the SFC is less than 50% (42). Therefore, pNMR has become the most commonly used method for SFC determination.

128

FAT CRYSTAL NETWORKS

5.2.1.1. Solid Fat Content by pNMR NMR techniques allow for a rapid determi-

nation of SFC with increased accuracy and precision relative to dilatometry and calorimetry, while offering additional advantages of being noninvasive and rapid (6 seconds per sample) (43). The NMR method for SFC determination is based on the principle that nuclear spins align in a magnetic eld. In a pNMR measurement, the realignment of the nuclear spins of 1H atoms after a strong electromagnetic pulse is measured (43). Application of a short 90 radio frequency electromagnetic pulse provides excitation to the 1H atoms, causing them to leave their equilibrium, aligned state. The pulse is then withdrawn leading to the relaxation of the 1H atoms and a return to the aligned, equilibrium state (43). Differences in the time scale for the relaxation process of the solid and liquid protons are used to determine the SFC. Protons, in general, have less mobility in the solid phase than in the liquid phase, therefore leading to large differences in the relaxation times of the 1H atoms (43). A typical excitation and relaxation cycle measured using NMR is shown in Figure 3. The determination of SFC requires measurements at two time periods, one within the short solid response region, S0 (11 ms), and one within the longer liquid response region, S00 (70 ms). The signal intensity S0 provides an estimate of the amount of both solid and liquid protons. Moreover, as the measurement cannot be taken immediately because of receiver dead time, the true amount of solid and liquid protons cannot be directly measured. For this reason, an extrapolation factor, known as the F-factor (F), is used to approximate the maximum response signal (S) value (43). SFC is calculated using the terms derived from the curve in Figure 3 and the NMR digital offset (D) using the following equation: SFCDirect S00 S0 S00 F : S0 S00 F D 1

Figure 3. Typical signal decay for a partially crystallized fat, following a 90  r.f. electromagnetic pulse. Parameters required for measurements of solid fat content (SFC) are shown.

NUCLEATION AND CRYSTAL GROWTH

129

The value calculated is then compared with the signal obtained from either standard oil or polymer calibration standards to determine the actual SFC value. This method is very reproducible and provides standard deviations of less than 1%, indicating a reliable and rapid method for SFC determination in fats (44).
5.2.1.2. Solid Fat Content and the Fat Crystal Network The solids content of a

fat crystal network is of critical importance to the nal physical properties of the system. Generally, an increase in SFC leads to an increase in fat rmness. The SFC measurement has been widely used as a determinant quantity for the structural properties of fat systems. Estimations for commercial plastic fats, including butter, predict rmness increases of 10% for every percent increase in SFC (45). As a result, models used to describe the rheological properties of fats incorporate references to SFC values. The primary factor affecting the SFC of any fat network is molecular composition. In general, longer chain fatty acids will have higher melting and crystallization temperatures than shorter chain fatty acids. Typically, the larger the amount of saturated, long-chain fatty acids, the greater the SFC. Processing conditions, including cooling rate, agitation, and tempering, have the ability to affect the SFC of fats.
5.2.1.3. Effects of Cooling Rate and Storage Time on SFC in AMF

AMF solid fat content is affected by processing conditions, in particular by cooling rate. The effects of cooling rate and storage time on SFC at 5 C can be appreciated in Figure 4. The effects of cooling rate, storage time, and their interaction have been determined to be statistically signicant (p < 0:05). Slowly cooling AMF at a rate of 0.1 C/min results in a 48% lower SFC than for samples that are cooled more rapidly (1 C/min and 5 C/min). The SFC of slowly cooled samples also

55.0

52.5

SFC (%)

50.0

47.5 0.1C/min 45.0 1C/min 5C/min 42.5 0 2 4 6 8 10 Storage Time (days) 12 14

Figure 4. Solid fat content of anhydrous milkfat cooled at 0.1C/min, 1C/min, and 5 C/min and stored for a period of 14 days at 5 C.

130

FAT CRYSTAL NETWORKS

increased slightly in time. On the other hand, at the higher cooling rates of 1 C/min and 5 C/min, SFC does not statistically change in time. This demonstrates that cooling rate has a signicant effect SFC of AMF, and that small changes in SFC can occur as a result of storage at 5 C for 14 days.
5.2.1.4. Crystallization Kinetics Monitored using SFC by pNMR

Solid fat content measurements can be used to monitor the kinetics of crystallization. SFC was monitored by pNMR as a function of time while samples were being cooled in a water bath. These results are presented in a plot of SFC versus time in Figure 5. The 0.1 C/min crystallization curve indicates that crystal growth begins at approximately 23 C, or 170 min, and continues over time once 5 C is reached. Crystal growth continues at a relatively constant rate until 205 min (19.5 C). At this point, a plateau-like region occurs in the SFC prole, indicating a signicant reduction in crystal growth until a time of 260 min (14 C). After this plateau region, crystal growth begins to occur at a more rapid and constant rate until the nal temperature of 5 C is reached. At this point, the SFC increases very slowly until an equilibrium SFC is achieved. The break in the crystallization curve for cooling at 0.1 C/min possibly indicates that different AMF fractions crystallize at different times, as a result of the slow cooling rate. This fractionation is common in AMF because of the large variety of TAGs and fatty acid species that are present. Within the rst crystal growth region, the longer chain saturated HMF triacylglycerols (16:0 to 20:0) crystallize and contribute to crystal growth and subsequent increase in SFC. Within the plateau-like region, TAGs containing both long- and short-chain fatty acids reach their

60 50 40 SFC (%) 30 20 10
17C 23C 19.5C

reaches 5C

0.1C/min 1C/min 5C/min


14C

0
0 100 200 300 400 1400 1800

Time (min)
Figure 5. Crystallization and crystal growth of anhydrous milkfat at controlled cooling rates of 0.1C/min, 1C/min, and 5 C/min monitored using SFC by pNMR.

NUCLEATION AND CRYSTAL GROWTH

131

supersaturated state and crystallize. Beyond the plateau region, medium-chain saturated MMF triacylglycerols (10:0 to 14:0) become supersaturated and contribute to crystal growth. The implications of this crystal growth pattern will become more apparent in the discussion of the microscopic and macroscopic properties of the crystallized AMF. Beyond the nal temperature of 5 C, any remaining TAGs that can become supersaturated continue to crystallize until an equilibrium state is achieved. The 1 C/min and 5 C/min crystallization proles indicate that crystal mass becomes detectable at approximately 17 C in both cases, which occurs after 23 min and 4 min, respectively (Figure 5). Once this temperature is reached, there is a constant and very rapid rate of crystal growth until the nal temperature of 5 C is reached. At approximately 12 C, there is an observable uctuation in the SFC growth curve that is followed by a continued rapid crystal growth until the equilibrium temperature of 5 C is reached. This minor uctuation in the SFC growth curves may indicate a polymorphic transition or more likely the occurrence of fractionation. The rapid and more spontaneous nucleation and crystal growth that occurs at higher cooling rates leads to less fractionation and the formation of more numerous mixed crystals containing a larger array of TAGs and a nal SFC that is 48% higher than the samples cooled at 0.1 C/min after 24 hours of storage. 5.3. Polymorphism Polymorphism is another characteristic of fat networks that affects their rheological characteristics. Polymorphic forms are crystalline phases with different structural characteristics, but of identical chemical compositions in their liquid state, when melted (45, 46). Polymorphic forms are usually categorized according to characteristic X-ray diffraction patterns, specic volume, and/or melting points (34, 45, 47). The polymorphic form developed during crystallization of bulk fat can be inuenced by several factors, including fat purity, TAG compatibility, temperature, supercooling, cooling rate, catalytic impurities, solvents, and seed crystals (45). Phase transitions from one polymorphic form to the next may also occur during processing and storage depending on many factors (45, 46). The three common polymorphic forms that exist in fat crystal networks are the suba, a, b0 , and b modications. These polymorphic modications and their characteristic crystallization patterns are shown in Figure 6. The suba and a forms are metastable (34). Each polymorphic form yields different crystal structures dependent on the magnitude of the crystallization driving force. The polymorphic modications also have varying thermodynamic stability, which determines their lifetime within a crystal matrix, with the tendency toward greater stability in the order a to b0 to b (24, 34).
5.3.1. Polymorphism and Milkfat High cooling rates (>1 C/min), or high

levels of supercooling (>15 C), lead to the rapid formation of metastable a nuclei (34). The persistence of these unstable nuclei is dependent on thermal treatments that occur after crystallization. These nuclei may remain in the a form or convert

132

FAT CRYSTAL NETWORKS

Figure 6. Polymorphic forms of fat crystals, including the possible polymorphic transitions, subcell packing structures, stability characteristic, and triacylglycerol stacking conformations.

to the more stable b0 structure. At lower degrees of supercooling or low cooling rates (>0 C/min to 1 C/min), b0 crystals predominate with trace amounts of metastable a nuclei detectable (34). Many factors inuence the formation of different polymorphic forms, including catalytic impurities, agitation, viscosity, TAG concentration, among others (48). The a polymorph is most readily formed, but it is very unstable. Under most conditions, it will transform within a short period of time into the b0 polymorphic form (34, 40). The a crystals are composed of TAGs whose long, alkane-like fatty acid chains are packed in a loose hexagonal subcell conformation, as shown in Figure 6 (3, 46). The hexagonal a subcell has a lower density because of the loose packing of the TAG molecules and is characterized by a single X-ray short spacing (wide-angle reection) at 4.15 A (34). This loose packing allows for a large number of reaction sites, the incorporation of a wide variety of TAGs into the solid fat matrix, and ample room for molecular rotation and reorientation. This allows for transitions from the a to b0 polymorphic form to occur readily (3, 24). This open lattice allows mixed crystal formation, incorporating a mixture of all of the supersaturated TAG present in the melt within the crystals (3). In AMF, the formation of the a form takes place during crystallizations at high degrees of supercooling or at high cooling rates, and it may persist as the predominant polymorphic form when the fat is stored at subzero temperatures (34). In most situations, the a form is the rst crystal structure formed during crystallization because of the loose packing and lower supersaturation requirements. However, a
5.3.1.1. The alpha (a) Polymorph

NUCLEATION AND CRYSTAL GROWTH

133

crystals tend to remain small and eventually transform, through the melt or molecular reorientation in the solid state, to the more stable b0 modication (3).
5.3.1.2. The Beta Prime (b0 ) Polymorph The b0 polymorph is the most common

polymorphic form in AMF. It will form directly from the melt or through transformation of the a polymorph (Figure 6) (24). The b0 polymorphic form exists in either a double- or triple-chain length conguration, the latter being more common in AMF, and orients in a more dense orthorhombic subcell structure (Figure 6), char acterized by X-ray short spacings at 3.8 and 4.2 A (45, 46). Processors prefer this form because it is structurally stable and maintains small- to-moderate crystal sizes allowing for soft and smooth products (3). The b0 structure does not allow for the incorporation of a large variety of TAGs within the crystal lattice, compared with the a structure (3). Lower TAG variety is a direct result of the tighter packing required to form the more stable structure, which requires similar length TAGs. During crystallization in the a form, dissimilar TAGs will often cocrystallize temporarily; however, in time, the shorter chain, or more unsaturated, less supersaturated molecules will redissolve. This, in turn, allows the remaining, more similar longer chain supersaturated TAGs to convert to the more stable b0 form (3, 40). TAGs in AMF under moderate undercooling conditions will crystallize in both the a and b0 polymorphic forms.
5.3.1.3. The Beta (b) Polymorph The presence of the b polymorph is uncommon

in AMF, but it has been known to exist in some situations (34). It is the most stable, highest packing density polymorphic form, and it may exist in either double- or triple-chain length conformation. Fatty acid chains within TAGs pack in a hexago nal arrangement (Figure 6), with a characteristic short spacing at 4.6 A (34, 45). The b form is uncommon in AMF because of the dense packing arrangement of similar TAGs. Because of the wide variety of TAG in low individual concentrations (<1 mol%) in AMF, the conditions required to form b structures are only met when the fat is crystallized at very slow rates, or at low degrees of supercooling (34, 46). These crystallization conditions are unusual in commercial practice; thus, the b polymorph is rare. The formation of b solids is also undesirable in AMF because it leads to the development of large crystals and sand-like texture (3, 34).
5.3.1.4. Polymorphism in Milk Fat as a Result of Processing Conditions

Work by ten Groetenhuis et al. (34) has shown that AMF typically crystallizes in two fractions. The highest melting group of TAG species represents one fraction, and the middle and low-molecular-weight TAGs form the second. This group also suggests that the application of different cooling rates lead to variations in the initial polymorphic form during crystallization of MF (34). Two regions of crystallization can be identied from the two major endotherms in the melting proles of samples subjected to different cooling rates, as shown in Figure 7. Immediate melting proles showed differences, suggesting that the initial polymorphic forms differ in AMF crystallized at various rates. Figure 7a depicts the melting prole of samples cooled at various rates after 10 minutes of storage at 5 C.

134

FAT CRYSTAL NETWORKS

Figure 7. Differential scanning calorimetry curves for anhydrous milkfat cooled at rates of 0.1 C/ min, 1 C/min, and 5  C/min to 5  C and stored for time periods (A) 10 minutes, (B) 1 day, (C) 7 days, and (D) 14 days.

The sample cooled at 0.1 C/min was expected to have had sufcient time for rearrangement, resulting in more dense packing of the TAGs, and crystallized in the stable b0 polymorph (34). For this reason, the peak of the endotherm at 17.5 C is assumed to correspond to the melting temperature of the b0 polymorph. The samples cooled at 1 C/min and 5 C/min (Figure 7a) both have minor inections in the curve at this same temperature (17.5 C), but they also have a major peak endotherm at around 15 C. This may correspond to an a-polymorphic melt. It is thus suggested that cooling at 1 C/min results in a mixture of a and b0 polymorphs, and cooling at 5 C/min results predominantly in the a polymorph. These predictions correspond to the ndings of previous studies on AMF employing similar crystallization conditions using DSC and powder X-ray diffraction (34). The samples cooled at the three cooling rates were also stored at 5 C for 1, 7, and 14 days, yielding the melting proles shown in Figures 7b, c, and d, respectively. After storage at 5 C, the melting proles began to look similar irrespective of the cooling rate used during crystallization. As it has been shown that AMF tends to crystallize in the b0 polymorph upon storage (some believe that the a polymorph survives for only minutes), similar shapes and peak melting points would be expected (3, 40).

NUCLEATION AND CRYSTAL GROWTH

135

5.4. Microstructure The textural properties of a fat are inuenced by all levels of structure, particularly microstructure. The microstructure includes the spatial distribution of mass, particle size, interparticle separation distance, particle shape, and interparticle interaction forces (4951). Methods that can be used for the characterization of microstructure in fat systems include, among others, small deformation rheology and polarized light microscopy, employing a fractal approach (4951).
5.4.1. Fat Crystal Network Theory Fat crystal growth is dictated by external

processing conditions and TAG composition. TAGs nucleate and grow from the melt into certain polymorphic and polytypic states. These primary crystals then aggregate into larger polycrystalline particles, also known as microstructural elements. Aggregation continues, leading to the formation of larger clusters, or microstructures, until a space-lling three-dimensional network is formed through interactions among microstructures. Crystalline mass within microstructures is distributed in a heterogeneous, disordered fashion, which can be characterized using fractal scaling principles (4951). Perfect fractal objects display exactly self-similar at all levels of magnication. Natural systems do not exhibit exact self-similarity, but statistical self-similarity; i.e., the microstructure, on average, is similar in appearance, distribution, and structure within a limited range of magnications. In particles networks, fractal scaling is usually encountered within the range of magnications corresponding to the size of microstructural elements to the size of microstructures (51). An example of statistical self-similarity uses the analogy of a tree branch structure. The tree begins as a trunk, the tree trunk has branches, these branches have branches, and so on. When the scale of observation is changed, a statistically self-similar pattern is observed. This theory was rst developed for colloidal aggregate networks and was later adapted to fat crystal networks (5254). In colloidal systems (with a disordered distribution of mass and statistical self-similar patterns), the mass of a fractal aggregate (or the distribution of mass within a network), M, is related to the size of the object or region of interest (R) in a power-law fashion: M $ RD ; 2

where D is the mass fractal dimension of the object, or the distribution of mass within a region of the network. The elasticity of a fat crystal network is dependent on the microstructure of the fat crystal network, particularly the spatial distribution of mass. The shear modulus (G) scales with the volume fraction of solids in a power-law fashion (4954): G $ m ; 3

where is the volume fraction of solid fat, or SFC/100, and m is dependent on D (see below).

136

FAT CRYSTAL NETWORKS

Figure 8. Idealized fat crystal network under extension. Particles (a) are packed in a fractal fashion within ocs (x). A force (F) acting on the network causes the links between ocs to yield, and the original length of the system in the direction of the applied force (L) to increase L. Thus, the inter-oc separation distance (l), also increases.

The scaling behavior of the elastic modulus was rst characterized for colloidal gels within two specic rheological regimes (55) and was later adapted to fat crystal networks (5254). These two regimes depend on the strength of the links that exist between individual clusters, relative to inner cluster strength. These regimes are referred to as the strong-link regime and weak-link regime (55). Figure 8 illustrates the behavior of a fat crystal network under extension in relation to these theoretical rheological regimes. The strong-link regime is only applicable at very low solid fat contents (usually below 10%). In the strong-link regime, crystal clusters grow large, and the links between these ocs are stronger than the ocs themselves. The elastic response of the material in this regime is a function of the elastic response of the ocs, and it is thus dependent on internal aggregate structure. The contrary is true for the weak-link regime, which is observed at high solid fat contents, where ocs are small and stronger than the links between them. In this regime, the elastic response of the material is a function of the elastic response of the links between ocs, and it is not dependent on the structure within the ocs. In both regimes, the macroscopic elastic modulus (K) of a system of size L is the sum of the elastic constants of the ocs (kx ) or the links between ocs (kL) in the direction of the applied stress (one-dimensional treatment). In the weak-link regime,   L K$ 4 kL $ x1 ; x where x is the size of a oc. Floc size (x) scales with the volume fraction of solids () in a power-law fashion: x $ aDd ;
1

NUCLEATION AND CRYSTAL GROWTH

137

where d is the Euclidean dimension (usually d 3), and a is the size of a primary particle. By inserting Equation 5 into Equation 4, we obtain the following relationship: K $ 3D $ G: This equation can also be expressed as G l3D ;
1 1

where l is a constant independent of the volume fraction, but dependent on several primary particle structural parameters as well as intermolecular forces. Our group has derived expressions for the Youngs modulus (E) of a fat as it relates to the structure of the material (5658): E$ 6d 1 A 1 3D 3D ; 2 ae 2pae d0 8

where d is the crystal-melt interfacial tension (about 0.01 J/m2 for TAGs), a is the primary particle size, is the volume fraction of solids (SFC/100), D is the fractal dimension, A is Hamackers constant (about 5 1020 J for alkanes), e is the extensional/compressional strain at the limit of linearity, and do is the equilibrium inter-microstructural (ocs or clusters) separation distance. Values for the shear modulus (G) could be obtained from knowledge of the Poisson ratio of the material. For a material where no volume change takes place when it is stretched or compressed, the Poissons ratio is 0.5 and E 3 G. Careful scrutiny of Equation 8 would suggest that the stress at the limit of linearity (s ) can be determined from the product of the Youngs modulus (E) and the strain at the limit of linearity (e ), s E e , yielding the expression: s 6d 1 3D : a 9

This expression for the stress at the limit of linearity, proposed in Marangoni and Rogers (58), provides an approximation to the yield stress (and thus hardness) of a fat. This model would allow for the prediction of the yield stress of a fat based on easily determined structural characteristics.
5.4.2. Fractal Dimension Evaluation using Small Deformation Rheology (Dr) To determine Dr for a particular system by rheological methods, the shear

storage modulus G0 must be measured at different solids volume fractions. The solids volume fraction, which is equal to the SFC/100, is varied by diluting the fat with an inert solvent that will not cocrystallize, or otherwise alter the behavior of the system. Canola oil has been used as a suitable AMF diluent for these purposes. A loglog plot of G0 versus yields a slope m from which Dr can be

138

FAT CRYSTAL NETWORKS

TABLE 2. Rheologically Determined Fractal Dimensions Dr and PreExponential Terms (l) for Anhydrous Milkfat Crystallized at Various Rates of Cooling and Storage Times at 5 C. Processing Conditions Cooling Rate  C=min 0.1 Storage Time (days) 1 7 14 1 7 14 1 7 14 Microstructural Parameters Dr 2.82 2.78 2.79 2.67 2.59 2.60 2.57 2.50 2.47 lMPa 1150 376 400 91 65 64 65 52 49

1.0

5.0

determined. The pre-exponential term l, which is inuenced primarily by particle properties, can be determined from the value of G0 , where 1:0 (SFC 100%). The effects of cooling rate on Dr are shown in Table 2. In general, an increase in cooling rate results in a decrease in Dr. Previous work in our laboratory has also shown that higher fractal dimensions occur in networks that are more ordered; thus, it can be expected that lower cooling rates would display this trend. The higher the Dr, the more ordered the crystal packing. For this reason, the value for Dr is expected to vary in the order of Dr0:1 C=min > Dr1 C=min > Dr5 C=min . When observed by time-lapsed microscopy, the samples cooled slowly (0.1 C/min) and yielded less numerous (see Figure 12 below), but more sporadically formed nuclei relative to the more rapidly cooled samples. This allows for a more ordered crystal growth. The 1 C/min sample initially demonstrated sporadic nucleation and controlled crystal growth of more numerous and smaller crystals than the slower cooled sample until reaching 13 C. Beyond 13 C, the AMF began to nucleate spontaneously and ll space very rapidly, leading to a more disordered nal system. Finally, samples cooled at 5 C/min demonstrated spontaneous nucleation and rapid crystal growth at 16 C. This resulted in a more disordered system with a fractal dimension lower than for slower cooling rates. The values for l, shown in Table 2, are also affected by cooling rate and storage time at 5 C. As with the Dr, the values for l decrease with increases in cooling rate. During storage at 5 C, the l values decrease moderately at each cooling rate and then appear to equilibrate at 7 days, with only minor differences detectable between the 7-day and 14-day values. This decrease may be attributable to continued crystal growth (demonstrated by small increases in SFC) and by the restructuring that occurs during polymorphic transitions and molecular rearrangement.
5.4.2.1. Fractal Dimension by Microscopy The use of fractal geometry as a

means of characterizing microstructure has been extensive in recent years. Various

NUCLEATION AND CRYSTAL GROWTH

139

applications have included investigating the structure of protein gels (59), depicting the irregular nature of solid adsorbents such as zeolites (60), and predicting optical responses of colloid-adsorbate lms (61). Recent work by Kim and Berg (62) has focused on the use of fractal scaling as a link between aggregation kinetics and structure that result from both diffusion and reaction-limited cluster aggregation processes of colloidal materials. To obtain the fractal dimension of a network of particles, acquiring images of the microstructure is necessary. Many forms of microscopy can be used, including brighteld microscopy, confocal laser scanning microscopy, scanning electron microscopy, and in the case of fat crystal networks, polarized light microscopy. Grayscale images of the network (usually consisting of pixel intensities ranging from pure white to pure black) must be converted to a binary format prior to image analysis. The discriminatory conversion of grayscale images to binary, requires a process termed thresholding. Images of fat crystal networks are typically obtained as 8-bit (256 color) images and consist of a histogram of pixel values that span the range of 0 (pure white) and 255 (pure black). Thresholding involves the selection of a particular pixel value within the range of 0 and 255 that will result in a binary image that best isolates the features of the network. This process is subject to much controversy because the selection of the cutoff pixel value is subjective. To minimize subjectivity, an autothresholding algorithm was used that provides accurate and reproducible results.1 This algorithm works by scanning the histogram to nd an intensity value where the average moments of the histogram counts about an intensity value are balanced. This means that a threshold value is chosen where the average pixel intensities are equal above and below the threshold (63). The general process of autothresholding can be seen in Figure 9. Once the image is converted to binary and the features isolated, it can be subjected to image analysis. Features are measured, and result in data, which can then be interpreted and related to structure. A ow chart outlining the general process from start to nish is depicted in Figure 10 (64). Emphasis is put on acquiring initial images that are of high quality (high distribution of pixel values, no saturation of white features, and evenly distributed lighting) to avoid unnecessary image processing to enhance features, thereby altering pixel values. Assuming a statistically constant microstructural element (or particle) size, the relationship between radius and mass (Equation 2) can be used to determine the fractality of crystal networks from two-dimensional PLM images (54). In this case, the scaling relationship between the number of discrete particles (N) and the length of the region of interest containing the particles (L) can be expressed as: N cLD : 10

1 Performed on a Macintosh computer using the public domain NIH Image program, developed at the U.S. National Institutes of Health and available on the Internet at http://rsb.info.nih.gov/nih-image/

140

FAT CRYSTAL NETWORKS

Figure 9. The conversion of 256-color grayscale micrograph to a binary image suitable for image analysis. A threshold value is chosen to isolate the featues that will be measured (AdobePhotoshop1 6.0).

Figure 10. Flowchart depicts the steps involved in obtaining and evaluating the microstructure of materials from digitally acquired images.

NUCLEATION AND CRYSTAL GROWTH

141

The rst method used was a particle counting algorithm performed using NIHImage2 as follows: c is a constant, and the number of distinct particles (N) are rst counted within the entire image of known length (L). At 5% increments, the image is cropped, thereby making a new, smaller length L, and the number of particles in each new image are counted. This process is repeated until the length of the image is 35% of the original size. A diagram depicting the overlaying of the boxes of decreasing size can be appreciated in Figure 11. The number of particles counted within each box size of various lengths (L) is plotted on a loglog scale, and the slope of the line is determined by linear regression. This slope corresponds to Df. This method for calculating Df was modied from previous procedures in our laboratory to eliminate artifacts and improve accuracy. This was accomplished by performing these iterative counts twice on each imageonce including all particles touching the edge of each region of interest, and once excluding those that touch the edges. By taking the average of these two counts, a Df that best represents the spatial distribution of mass is obtained. This improved method is equivalent to well-established methods that involve counting features that touch two sides of the region of interest (64). The fractal dimension arrived at by this method reveals information on the degree of order in the packing of the microstructural elements within the microstructures. Systems that display a high degree of order have characteristically lower

Figure 11. Schematic diagram shows the incremental decreases in box size used for the particle counting method for the determination of the microscopic fractal dimension D f .

2 Analysis performed on a Macintosh computer using the public domain NIH Image program, developed at the U.S. National Institutes of Health and available on the Internet at http://rsb.info.nih.gov/nih-image/

142

FAT CRYSTAL NETWORKS

fractal dimensions than those of a more disordered network (54). Previous simulations and studies employing mass-radius techniques for the determination of D have found that fast, diffusion-limited cluster aggregation (DLCA) typically results in D values of 1.751.8, whereas slow, reaction-limited cluster aggregation (RLCA) processes result in D values of 2.02.1 (60). Another method used to examine microstructure is the box counting or grid dimension analysis. This dimension is referred to as a grid dimension because for mathematical convenience, the boxes are usually part of a grid that is laid over the image (65). The box dimension is dened as the exponent D in the relationship: NL $ 1 ; LD 11

where N(L) is the number of boxes of linear size L necessary to cover a data set of points distributed in a two-dimensional plane. If the network is indeed fractal, plotting the logarithm of N(L) versus the logarithm of L results in a linear plot with a negative slope equal to D (65). This analysis appears to be sensitive to the degree of ll of the solids in a crystal network. It can be expected that a network that is more empty (many large void spaces) will result in a lower proportion of full boxes counted, and thus, it will result in a lower fractal dimension and vice versa. It is important to note that in some cases, the fractal dimensions estimated by different methods are not the same. It is therefore necessary that one be aware of the particular method used and its interpretation; this is mandatory for comparisons of dimensions of different data sets (65). One of the future goals within our laboratory is to continue investigating and quantifying the link between the fractal dimensions determined by rheology and microscopy.
5.4.3. Polarized Light Microscopy Microscopy allows for the visual observa-

tion of crystallization, and crystal growth of fat networks in realtime and subsequent image analysis can be used to quantify particle size, degree of order, and space-lling mass. Polarized light microscopy was used to examine crystal network properties of AMF during nucleation and crystal growth, and over time as a result of the processing conditions and storage time. Cooling at 0.1 C/min, 1 C/min, and 5 C/min resulted in detectable birefringent crystal mass at onset temperatures of 26.8 C, 20 C, and 16 C, respectively. The onset temperatures of the 0.1 C/min and 1 C/min samples do not correspond exactly to those determined by DSC or NMR. This is because of the small quantity of matter crystallizing at the onset of nucleation, which although visible by PLM, does not release enough heat to be resolved by the DSC or enough solid mass to be detected using pNMR. The sample cooled at 5 C/min, on the other hand, shows visual signs of crystallization at 16 C, which corresponds closely to the values obtained in the DSC and pNMR experiments. This similarity in measurement

NUCLEATION AND CRYSTAL GROWTH

143

and resolution by each of the methods is attributable to the large change in state that occurs at higher cooling rates. Microscopy also allows for the observation of dynamic changes that occur during nucleation and crystal growth. Figure 12 depicts still images of crystallizations at cooling rates of 0.1 C/min, 1 C/min, and 5 C/min. The images shown are at 5 C intervals in the range of 30 C to 5 C and portray dramatic differences in kinetics, and crystal properties, as a result of altering the cooling rates. The crystallization prole for AMF cooled at 0.1 C/min shown in Figure 12 spans a total time period of 250 minutes. At 26.8 C, sporadic nucleation begins, and as time progresses, these early nuclei intermittently grow outward in a radial fashion as the temperature decreases. Until late in the crystallization period (at 5 C), there is no visible evidence of secondary nucleation. Between the temperatures of 26.8 C and 23 C, there is a discrete and continuous growth region where the microstructures increase dramatically in size as a function of time. This is followed by a period of very little growth (little evidence of changing microstructure sizes) until 14.5 C, when signicant growth occurs once again up until a temperature of 12 C is reached. From approximately 8 C until 5 C, minor space-lling crystal growth occurs. These intermittent, temperature-dependentrelated growth regions are likely attributable to the crystallization of the AMF fractions of HMF (26.8C24 C), followed by MMF (14.5C12 C), and then the space-lling crystallization of some of the LMF. This is made possible by the very slow change in supercooling that occurs over time when cooling AMF at 0.1 C/min, which permits adequate time for diffusion-related interaction to occur, leading to the fractionationmediated growth of the network. The resulting fat crystal network is made up of a small number of very large crystal structures and relatively large regions of void space as seen in Figure 12. The crystallization prole for AMF cooled at 1 C/min shown in Figure 12 spans a real-time period of 25 minutes. The rst sign of crystal structure occurs at 20 C and is followed by a large amount of spontaneous nucleation and radial crystal growth until a temperature of 16 C is reached. After the rapid growth period, there is minimal visible growth until approximately 12 C is reached and a very rapid, spontaneous, space-lling secondary nucleation and growth occurs and continues until 5 C is attained. This rapid period of space lling crystallization leads to the masking of the larger microstructures formed during the earlier stages of crystallization. Like the system cooled at 0.1 C/min, AMF cooled at 1 C/min demonstrates fractionation-mediated growth where most of the HMF and some MMF likely crystallizes during the early stages, followed by crystallization of the MMF and LMF during the later stages. The resulting microstructure consists of a large number of small crystal structures; however, previous micrographs collected at the higher temperatures indicate the presence of somewhat larger structures (Figure 12). The crystallization prole of AMF cooled at 5 C/min spans a real-time period of just 5 minutes (Figure 12). The rst visible sign of a crystal structure occurs at a temperature of 16 C, after which very spontaneous, and rapidly space-lling, nucleation and crystal growth occurs. The progression of growth until 5 C shows

144

FAT CRYSTAL NETWORKS

Figure 12. Polarized light micrographs of anhydrous milkfat cooled at 0.1C/min, 1C/min, and 5 C/min. Images were acquired during crystallization in the range of 30 C to 5 C at intervals of 5 C.

NUCLEATION AND CRYSTAL GROWTH

145

minimal differences over time with the exception of the occurrence of signicantly increasing birefringence with decreases in temperature. This is an indicator of the densication and continued crystal growth of the microstructural elements within the crystal network. The resulting network is a densely packed, space-lling matrix of very small crystal structures. In general, as cooling rate increases, the number of crystal aggregates also increases, and the size of the constituent particles decreases. The effects of storage time on microstructure at 5 C are represented in Figure 13. For each cooling rate, the network structure and particle size established during crystallization remain relatively unchanged during storage. Minor changes in the appearance of the 0.1 C/min sample occur as a result of the lling of void space, during storage time, by small crystal structures similar to those present in the originally formed crystal matrix. Rapidly cooled samples demonstrate no signicant change in network or particle appearance as a function of storage time at 5 C. The effects of cooling rate and storage time on mean particle size, as determined by image analysis, are shown in Table 3. Cooling at 0.1 C/min resulted in average particle sizes that were approximately two times larger than those resulting from cooling at 1 C/min, and 5 C/min. Also, the minor changes in the appearance of samples cooled at 0.1 C/min can be further characterized by a slight decrease

Figure 13. Polarized light micrographs of anhydrous milkfat cooled at 0.1C/min, 1C/min, and 5 C/min followed by storage for 1 day, 7 days, and 14 days at 5 C.

146

FAT CRYSTAL NETWORKS

TABLE 3. Particle-Counting Fractal Dimension Df , Box-Counting Fractal Dimension Db , and Mean Microstructural Element Area MEA for Anhydrous Milkfat Crystallized at Various Rates of Cooling and Storage Times at 5 C. Processing Conditions Cooling Rate Storage Time  C=min (days) 0.1 1 7 14 1 7 14 1 7 14 Df 1.98 0.11 1.98 0.09 1.98 0.08 1.92 0.08 1.88 0.04 1.86 0.07 1.91 0.03 1.91 0.04 1.89 0.04 Microstructural Parameters (n 1012) Db 1.72 0.01 1.73 0.01 1.73 0.01 1.85 0.01 1.85 0.01 1.85 0.01 1.87 0.01 1.88 0.00 1.88 0.00 MEAmm2 24.76 7.83 15.71 1.43 15.72 1.42 4.12 0.47 3.91 0.22 3.44 0.17 3.88 0.40 3.50 0.19 3.26 0.07

1.0

5.0

in mean particle area (MEA total crystal area/no. of particles) as a function of time. Samples cooled at higher rates of 1 C/min and 5 C/min remain relatively constant in time. Fractal dimensions of micrographs obtained at different cooling rates, determined using the particle counting method, reveal information about the degree of order in the resulting networks (Table 3). The Df values decreased in the order of Df0:1 C=min > Df5 C=min ! Df1 C=min , with no signicant changes detectable over time. Fractal dimensions determined via the box-counting or grid dimension method are also shown in Table 3. Db values decrease in the order of Db1 C=min > Db5 C=min > Db0:1 C=min , and they do not signicantly change in time. Db is sensitive to the degree of ll within the network; therefore, higher values indicate an increase in space-lling mass (i.e., the smaller, more-numerous particles of the 1 C/min sample ll more space relative to the larger, less-numerous 0.1 C/min particles).

5.4.4. Importance of Nucleation and Crystal Growth Kinetics on the Final Crystal Properties and Network As stated in the hierarchical model (Figure 1),

the processing and storage conditions affect the physical properties, including the SFC, polymorphism, and microstructure during nucleation and crystal growth. Therefore, processing conditions exert their most dramatic effect during crystallization, during which the nal crystal network properties are developed. Scrutiny of the micrographs in Figure 12 demonstrates that the crystal structures formed during the early stages of nucleation and crystal growth dictate the nal fat crystal network structure. The samples cooled at 0.1 C/min show fewer, more sporadic nuclei, which appear at higher temperatures. These nuclei then grow slowly to form

NUCLEATION AND CRYSTAL GROWTH

147

large crystal structures that continue to grow until the nal temperature of 5 C is reached. Similar trends are seen in the faster cooled samples, where spontaneous nucleation occurs, leading to a large number of small crystals. This distribution remains at the nal temperature of 5 C/min. Analysis of these micrographs using the fractal dimensions (Df and Db) also indicates that the crystal network structure is determined before the nal temperature is reached. These results can be seen in Figure 14. These results demonstrate that at 10 C15 C prior to the nal temperature of 5 C, the nal equilibrium value (or close to the nal value) for

2.5 2.0 Fractal Dimension (Df)

1.5

1.0 0.1C/min 0.5 1C/min 5C/min 35 25 15 5

0.0

Temperature (C) (a) 2.0

Fractal Dimension (Db)

1.5

1.0 0.1C/min 0.5 1C/min 0.0 35 25 15 5C/min 5

Temperature (C) (b)


Figure 14. Cooling prole analysis of polarized micrographs of 100% AMF collected during the static crystallization process at the cooling rates of 0.1C/min, 1C/min, and 5 C/min depecting the microscopic properties (A) D f and (B) D b monitored against changing temperature.

148

FAT CRYSTAL NETWORKS

microscopically determined fractal dimensions are reached. This indicates that the nal structure is predetermined far before the nal equilibrium state is reached.

6. MECHANICAL PROPERTIES 6.1. Small Deformation Small deformation rheometry refers to testing procedures that do not cause structural damage to the sample. Constant stress rheometers, such as dynamic mechanical analyzers or oscillatory constant stress rheometers, are often used. There have been a number of studies that demonstrate that crystallized AMF and butter exhibit linear (ideal) viscoelastic behavior at low levels of stress or strain (4), where the strain is directly proportional to the applied stress. For most materials, this region occurs when the critical strain (strain where structure breaks down) is less than 1.0%, but for fat networks, the strains typically exceed 0.1% (4, 66). Ideally, within the LVR, milkfat crystal networks will behave like a Hookean solid where the stress is directly proportional to the strain (i.e., s / g), as shown in Figure 15 (66, 68). Within the elastic region, stress will increase linearly with strain up to a critical strain. Beyond that critical strain (strain at the limit of linearity), deformation of the network will occur at a point known as the yield point. The elastic limit quickly follows, beyond which permanent deformation and sample fracture occurs. Beyond these points, the structural integrity of the network is compromised and the sample breaks down. From small-deformation oscillatory methods, several useful parameters can be obtained to describe the mechanical properties of a material. These measurements

Figure 15. Stress-strain behavior of a typical elastic system, including (A) yield point, (B) elastic limit, (C) irreversible deformation, and (D) fracture.

MECHANICAL PROPERTIES

149

Figure 16. Time prole of an applied sinusoidal stress wave and the corresponding resulting sinusoidal strain wave as they apply to small deformation rheological testing.

include the complex modulus (G*), the shear storage modulus G0 , the shear loss modulus G00 , and the tangent of the phase shift or phase angle tan d. These rheological parameters can be determined using controlled stress rheometers using dynamic oscillatory testing within the LVR region (68). The oscillatory method collects strain information by applying a controlled stress via the application of a sinusoidal stress wave. The rheometer measures the variation in strain as a function of the applied stress, in terms of the magnitude of the strain and the phase angle (d) between the applied stress wave and the resulting strain wave. A typical stress-strain sinusoidal relationship is shown in Figure 16. The vectorial resolution, shown in Figure 17, of the stress-strain ratio is used to calculate the complex modulus G*, which is derived from the following equation: jG j p G02 G002 :

12

The strain response can be broken down into its elemental components of stress, which are in phase or out of phase, to derive the values for G0 and G00 . The storage modulus G0 is the ratio of the applied stress that is in phase with the strain (d 0 ). This means that G0 is an expression of the magnitude of the energy stored in the material, recoverable per deformation cycle (68). The loss modulus G00 is the ratio of the applied stress that is out of phase with the strain (d 90 ), meaning that it is a measurement of the energy lost as viscous dissipation per deformation cycle (66 68). These two moduli are dependent on the phase angle of the system and are

150

FAT CRYSTAL NETWORKS

Figure 17. Vectorial resolution of the complex modulus components.

derived from the vectorial components of G* and are calculated using the following relationships:  s0 G cos d; g  0 s0 sin d: G00 g0
0

13 14

The tangent of the phase angle tan d can be expressed as the ratio of the loss to the storage moduli, and it represents the relative balance of elastic to viscous components in a material: tan d G00 : G0 15

These rheological parameters have been successfully correlated to textural attributes of hardness and spreadability and provide information pertaining to the fat crystal network (69). The value of G0 is useful in assessing the solid-like structure of the fat crystal network. Increases in the value of G0 typically correspond to a stronger network and a harder fat (66). Alternatively, G00 represents the uid-like behavior of the fat system. This value can be related to the spreadability of a fat system, because increases in G00 indicate more uid-like behavior under an applied shear stress. The tan d is the ratio of these two values. As the value of d approaches 0 (stress wave in phase with stress wave), the G00 value approaches zero, and therefore, the sample behaves like an ideal solid and is referred to as perfectly elastic (68). As d approaches 90 (stress is completely out of phase relative to the strain),

MECHANICAL PROPERTIES

151

the G0 value approaches zero; all of the energy will be dissipated as heat, and the sample will behave predominantly as a uid (68). At intermittent values, the samples are considered to be viscoelastic in nature. Therefore, tan d is an excellent indicator of the structural integrity of the sample, indicating the proportion of the material structure attributable to the crystal network and to the liquid phase.
6.1.1. Effects of Processing ConditionsSmall Deformation Rheology The

effects of cooling rate and time on small deformation rheological measurements of G0 , G00 , and tand are shown in Figures 18, 19, and 20, respectively. Figure 18 illustrates the effects of cooling rate and storage time on the storage modulus G0 measured using small deformation rheology at 5 C. There are no signicant effects of cooling rate or storage time on the G0 of AMF samples, indicating the absence of differences in solid-like properties and potentially hardness. Figure 19 illustrates the effects of cooling rate and storage time on the loss modulus G00 , measured using small deformation rheology at 5 C. Results show lower G00 for samples cooled at 5 C/min and 1 C/min than at 0.1 C/min. Additionally, the G00 values for the 5 C/min and 0.1 C/min cooling rates do not change signicantly with time, and the values for the 1 C/min samples decrease slightly with time, which may be indicative of time-dependent hardening. The tangent of the phase angle tan d offers a better indicator of structural integrity than either the G0 or G00 measurements do individually. The phase angle (d) or its tangent tan d are an indicator of the systems structural behavior or integrity because it indicates whether the system behaves predominantly as a solid, liquid, or viscoelastic material. The greater the tan d, the more liquid-like the samples will behave, and conversely, low values indicate more solid-like properties. A tan d value of 1 is indicative of a viscoelastic material.

20

15 G' (MPa)

10

0.1C/min 1C/min 5C/min

0 0 2 4 6 8 10 Storage Time (days) 12 14

Figure 18. Storage moduli G 0 of anhydrous milkfat cooled at rates of 0.1C/min, 1C/min, and 5 C/min followed by storage for 14 days at 5 C.

152

FAT CRYSTAL NETWORKS

2.0

1.5 G(MPa)

1.0

0.5

0.1C/min 1C/min 5C/min

0.0 0 2 4 6 8 10 Storage Time (days) 12 14

Figure 19. Loss moduli G 00 of anhydrous milkfat cooled at rates of 0.1C/min, 1C/min, and 5 C/min followed by storage for 14 days at 5 C.

The graphs in Figure 20 demonstrate the effects of cooling rate and storage time on tan d at 5 C. Values of tan d increase in the order of tan d5 C/min < tan d1 C/min ( tan d0.1 C/min. The tan d value for AMF cooled at 0.1 C/min is signicantly higher than the values at 1 C/min and 5 C/min. The values do not show large increases with time; therefore, the changes during storage at 5 C are minimal. The data from small deformation rheology indicate that at higher cooling rates, the fat samples are more solid-like in nature. Conversely, the low cooling rate of 0.1 C/min results in a softer, less elastic system.
0.15

0.10 tan() 0.05

0.1C/min 1C/min 5C/min

0.00 0 2 4 6 8 10 Storage Time (days) 12 14

Figure 20. Values of tan d of anhydrous milkfat cooled at rates at 0.1C/min, 1C/min, and 5 C/min followed by storage for 14 days at 5 C.

MECHANICAL PROPERTIES

153

6.2. Large Deformation Large deformation methods usually are based on the determination of the amount of applied force required to induce a change in a sample. Parameters determined include hardness, spreadability, cutting force, or yield force (4). One such method involves the compression of a sample between two parallel plates to determine relative hardness values via the measurement of yield force. The compression of uniform samples to the point where the force exceeds the structural capacity causes it to permanently deform and essentially break (4). A typical load-deformation curve can be used to derive values for yield stress, yield strain, and compressive yield work, and depending on the linearity of the onset of compression, a compressive modulus may be obtained (4). These measurements can be used to provide an index of hardness for fats, which have been successfully correlated to the textural attributes of hardness and spreadability obtained through sensory evaluation (4). Unfortunately, these tests are destructive in nature and yield minimal information about the native microstructure of the system. The effects of cooling rate and time on the yield force of AMF and its dilutions are shown in Figure 21. Yield force measurements (Fy) demonstrate that cooling rate and, to a lesser extent, storage time affect the hardness of AMF. In general, the Fy increases with increasing cooling rate; however, the difference between 1 C/min and 5 C/min is relatively small compared with the difference observed between 0.1 C/min and 1 C/min. AMF samples cooled at 0.1 C/min have yield force values (3035 N) that are approximately half that of the higher cooling rates of 1 C/min (5863 N) and 5 C/min (6472 N). Over time, the yield force increases slightly, by
6.2.1. Effect of Processing ConditionsLarge Deformation Rheology

75

Yield Force (N)

50

25

0.1C/min 1C/min 5C/min

0 0 2 4 6 8 10 12 14 Storage Time (days)


Figure 21. Yield force F y of anhydrous milkfat cooled at rates of 0.1C/min, 1C/min, and 5 C/min followed by storage for 14 days at 5 C.

154

FAT CRYSTAL NETWORKS

a few Newtons, at each of the cooling rates. This indicates that the samples may be hardening slightly; however, the yield force does not signicantly increase p > 0:05 during the 14-day storage period.

7. ASSESSING THE VALIDITY OF THE MODEL: CORRELATING EXPERIMENTALLY DETERMINED PARAMETERS In order to further demonstrate the existence of inter-relationships between parameters that inuence macroscopic properties, linear correlations were performed on data obtained from the study of cooling rate and storage time (Table 4). As the physical properties of a fat system are dictated primarily by lipid composition, it is interesting to observe the effects of composition of other measured parameters. For example, a signicant correlation exists between the saturated fatty acid content (SFAC) and the nal SFC attained (Figure 22a). This agrees with expectations, because the saturated FAs generally have higher melting points, and thus, they are likely to crystallize and contribute to the measured SFC at any given temperature. Signicant positive correlations also exist for relationships between SFAC and

55 50 SFC (%) 45 40 35

A
G' (MPa)

20 15 10 5 r =0.905 55 60 65 70
2

30 50

0 50

r2=0.785 55 60 65 70

Saturated Fatty Acid Concentration (%) 75 Yield Force (N)

Saturated Fatty Acid Concentration (%)

50

25 r2=0.392 60 65 70

0 55

Saturated Fatty Acid Concentration (%)


Figure 22. Linear correlations between saturated fatty acid content (%) and (A) solid fat content, (B) storage modulus, and (C) yield force. Data shown represent all points collected at all cooling rates and storage times.

TABLE 4. First Order Linear Regression Coefcients (2) for Various Physical Properties of Anhydrous Milkfat. Correlations were Performed After Pooling all Results from Experiments Involving Various Cooling Rates and Storage Times at 5 C. SFAC SFC Df Db Dr MEA l G0 G00 tan (d) Yield Force Work Strain
Legend na bold SFAC () or ()

SFC 0.80 () 0.89 () 0.78 () 0.98 () 0.86 () 0.87 () 0.57 () 0.23 () 0.77 () 0.81 () 0.02 ()

Df

Db

Dr

MEA

G0

G00

tan (d)

Yield Force

0.91 () 0.37 () 0.62 () na 0.31 () na 0.78 () 0.71 () 0.08 () 0.39 () 0.58 () 0.04 ()

0.78 () 0.70 () 0.77 () 0.54 () 0.19 () 0.80 () 0.81 () 0.87 () 0.66 () 0.58 ()

0.90 () 0.90 () 0.60 () 0.11 () 0.92 () 0.95 () 0.97 () 0.48 () 0.72 ()

0.77 () 0.54 () 0.04 () 0.87 () 0.87 () 0.90 () 0.28 () 0.73 ()

0.90 () 0.15 () 0.76 () 0.79 () 0.86 () 0.53 () 0.61 ()

0.11 () 0.50 () 0.52 () 0.61 () 0.36 () 0.44 ()

0.90 () 0.34 () 0.61 () 0.63 () 0.01 ()

0.10 () 0.02 () 0.14 () 0.29 ()

0.63 () 0.31 () 0.52 ()

0.76 () 0.23 ()

comparisons unachievable due to method of calculation p<0.05 saturated fatty acid concentration (%) indicates positive or negative correlation

156
Microstructural Element Area (m2)

FAT CRYSTAL NETWORKS

30 25 20 15 10 5

r2=0.983 G' (MPa)

20 15 10 5

0 45.0

47.5

50.0 SFC (%) 75 Yield Force (N)

52.5

55.0

0 25

r2=0.874 30 35 40 SFC (%) 45 50 55

50

25 r2=0.773 35 40 45 50 55 SFC (%)

0 30

Figure 23. Linear correlations between solid fat content and (A) microstructural element size, (B) storage modulus, and (C) yield force. Data shown represent all points collected at all cooling rates and storage times.

measures of apparent hardness. SFAC is linearly correlated with both the storage modulus G0 and yield force (Fy) (Figure 22b,c). Higher saturate levels, their subsequent crystallization, and contribution to the solids content of the network translate into increases in the materials hardness. SFC, which is commonly used as a primary indicator of hardness, shows a high positive correlation to both the storage modulus G0 and yield force (Fy) in milkfat (Figure 23a,b). This relationship forms the basis of methods used to determine the fractal dimension of fat crystal networks. Evidence of a relationship between SFC and microstructural element size was also established (Figure 23c). Smaller crystal aggregates are a result of higher cooling rates, which also dictate the nal SFC. This relationship between factors further exemplies the interdependence of processing conditions, microstructure, and SFC. Correlations shown in Figure 24 highlight the relationships between various microstructural parameters and yield force. As stated previously, microstructure strongly inuences macroscopic hardness. Networks consisting of smaller crystal aggregates are generally harder than those made up of larger microstructures (Figure 24a). Correspondingly, networks consisting of particle distributions that are more disorderly (lower Df and Dr) and more space lling (higher Db) are also harder, as indicated by higher yield force values (Figure 24.b, c, and d).

ASSESSING THE VALIDITY OF THE MODEL: CORRELATING

157

80 Yield Force (N)

60 40 20 0 0 5 10 15 20 25 30 Microstructural Element Area (m2) 80

Yield Force (N)

r2=0.865

80 60 40 20 0

r2=0.865

1.85

1.90

1.95

2.00

Fractal Dimension (Df) 80 Yield Force (N)

Yield Force (N)

D
60 40 20 0 2.4 r2=0.897 2.5 2.6 2.7 2.8 2.9 Fractal Dimension (Dr)

60 40 20 0 1.70 r =0.969 1.75 1.80 1.85 1.90 Fractal Dimension (Db)


2

Figure 24. Linear correlations between yield force and (A) microstructural element area, (B) fractal dimension by particle-counting. (C) fractal dimension by box-counting, and (D) fractal dimension by rheology. Data shown represent all points collected at all cooling rates and storage times.

7.1. Summary of the Effect of Processing Conditions on the Physical Properties of AMF The schematic illustration of the combined effects of the majority of the physical properties examined can be appreciated in Figure 25. Figure 25a illustrates the effect of rapid cooling (5 C/min) on the physical properties of AMF. The diagram depicts a large number of small, randomly packed crystal structures. These structures are highly space lling and therefore have a larger number of particleparticle interactions. The result is a crystal network with an even higher SFC (5153%), high Db (highly space lling), and low Dr and Df values, resulting in an even harder fat crystal network. Figure 25b illustrates the effects of the midrange cooling rate (1 C/min) on the physical properties of AMF. This diagram shows a combination of both small- and medium-sized crystal structures that are more randomly distributed in space. The increased number of smaller crystal structures lead to increases in the number of particleparticle interactions and a decrease in void space. This resulted in a system with a higher SFC (5153%), higher Db (more space lling), and decreased Dr and Df (less packing order), which all translated into a harder fat.

158

FAT CRYSTAL NETWORKS

Figure 25.

Finally, Figure 25c depicts an illustrative model for the effects of the slow (0.1 C/min) cooling rate on the physical properties of AMF. As shown, the crystal network is composed of a small number of large highly ordered crystal structures that are homogeneously distributed in space. The decrease in crystal structure numbers and increased size lends results in fewer particleparticle interactions and a larger amount of void space. The nal result was a lower SFC (46 48%), lower values for Db (less space lling crystal mass), and increases in Dr and Df (increased structural packing order), which all resulted in a softer fat (50% lower yield force than samples cooled at 1 C/min and 5 C/min). Therefore, through experimentation using AMF as a model fat, it has been shown that processing conditions affect the underlying physical properties of a fat, and it is these physical properties and their interrelationships that ultimately effect the nal macroscopic properties of a fat crystal network. ACKNOWLEDGMENTS The authors acknowledge the nancial support of the Natural Sciences and Engineering Research Council of Canada, the Ontario Ministry of Agriculture and Food, the Canadian Foundation for Innovation, and the Ontrario Innovation Trust. REFERENCES
1. M. Kalab, J. Dairy Sci., 68, 32343248. 2. A. L. Halmos, Food Australia, 49(4), 169173 (1997). 3. W. D. de Bruikne and A. Bot, in A. J. Rosenthal, ed., Food Texture Measurement and Perception, Aspen, Gaithersburg, Maryland, 1999 pp. 185227. 4. A. J. Wright, M. G. Scanlon, R. W. Hartel, and A. G. Marangoni, J. Food Sci., 66(8), 1056 1071 (2001). 5. D. Rousseau and A. G. Marangoni, Food Res. Int., 31, 381388 (1998). 6. J. H. Prentice, J. Texture Studies, 3, 415458 (1972).

REFERENCES

159

7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22.

L. de Man, J. M. deMan, and B. Blackman, J. Amer. Oil Chem. Soc., 66, 128132 (1987). J. M. deMan and A. M. Beers, J. Texture Studies, 18, 303318 (1987). D. Rousseau, A. R. Hill, and A. G. Marangoni, J. Amer. Oil Chem. Soc., 73, 983989 (1996). A. M. Fearon and D. E. Johnston, J. Food Quality, 12, 2338 (1989). A. J. Haighton, J. Amer. Oil Chem. Soc., 36, 345348 (1959). K. R. Davey and P. N. Jones, J. Texture Studies, 16, 7584 (1985). M. Kawanari, D. D. Hamann, K. R. Swartzel, and A. P. Hansen, J. Texture Studies, 12, 483505 (1981). H. Rohm and F. Ulberth, J. Texture Studies, 20, 409418. V. R. Huebner and L. C. Thomsen, Amer. Dairy Sci., 40, 834838 (1957). C. Pompei, E. Casiraghi, M. Lucisano, and B. Zanoni, J. Food Sci., 53, 597602 (1988). H. T. Lawless and H. Heymann, in Sensory Evaluation of Food, Aspen, Gaithersburg, Maryland, 1999, pp. 379405. B. K. Mortensen and H. Danmark, Milchwissenschaft 36, 393395 (1982). H. Corey, CRC Crit. Rev. Food Technol., May, 161198 (1970). A. S. Szczesniak, Food Technol., October, 5258 (1966). D. M. Small, in D. J. Hanahan, ed., The Physical Chemistry of Lipids, from Alkanes to Phospholipids, Vol. 4, Plenum Press, New York, 1966, pp. 475522. A. Kuksis, L. Marai, and J. J. Myher, J. Amer. Oil Chem. Soc., 50, 193201 (1973).

23. J. Gresti, M Bugaut, G. Maniongui, and J. Bezard, J. Dairy Sci., 76, 18501869 (1993). 24. B. Breitschuh and E. J. Windhab, J. Amer. Oil Chem. Soc., 75, 897904 (1998). 25. H. E. Swaisgood, in O. R. Fennema, Food Chemistry, 3rd ed., 1996, pp. 841878. 26. W. W. Christie, in P. F. Fox, ed., Advanced Dairy Chemistry, Vol. 2, 2nd ed., Chapman & Hall, London, 1995, pp. 136. 27. A. J. Wright, S. E. McGauley, S. S. Narine, R. W. Lencki, and A. G. Marangoni, J. Agric. Food Chem., 48, 10331040 (2000). 28. A. G. Marangoni and R. W. Lencki, J. Agric. Food Chem., 46, 38793884 (1998). 29. D. Rousseau and A. G. Marangoni, J. Agric. Food Chem., 46, 23682374 (1998). 30. D. Rousseau and A. G. Marangoni, J. Agric. Food Chem., 46, 23752381 (1998). 31. A. G. Marangoni and D. Rousseau, J. Amer. Oil Chem. Soc. 75, 16331636 (1998). 32. A. G. Marangoni and D. Rousseau, Food Res. Int., 31, 595599 (1998). 33. A. G. F. Stapley, H. Tewkesbury, and P. J. Fryer, J. Amer. Oil Chem. Soc., 6, 677685 (1999). 34. E. ten Groetenhuis, G. A. van Aken, K. F. van Malssen, and H. Schenk, J. Amer. Oil Chem. Soc., 76, 10311039 (1999). 35. M. L. Herrera and R. W. Hartel, J. Amer. Oil Chem. Soc., 77, 11771187 (2000). 36. M. L. Herrera and R. W. Hartel, J. Amer. Oil Chem. Soc. 77, 11971204 (2000). 37. M. L. Herrera and R. W. Hartel, J. Amer. Oil Chem. Soc. 77, 11891195 (2000). 38. J. Garside, in J. M. V. Blanshard and P. Lillford, eds., Food Structure and Behaviour, Academic Press, London, 1987, pp. 3549. 39. D. S. Grall and R. W. Hartel, J. Amer. Oil Chem. Soc., 69, 741747 (1992). 40. P. Walstra, T. van Vliet, and W. Kloek, P. F. Fox, eds., Advanced Dairy Chemistry, Vol. 2, 2nd ed., Chapman & Hall, London, 1995, pp. 179211.

160

FAT CRYSTAL NETWORKS

41. A. G. Marangoni and D. Rousseau, N. Widlak, ed., Physical Properties of Fats, Oils and Emulsiers, AOCS Press, Champaign, Illinois, 1999, pp. 96111. 42. B. L. Madison and R. C. Hill, J. Amer. Oil Chem. Soc., 55, 328331 (1978). 43. J. van Duynhoven, G. J. Goudappel, M. C. M Gribnau, and V. K. S. Shukla, Inform, 105, 479484 (1999). 44. L. van Putte, I. Bruker minispec documentation. 45. W. W. Nawar, in O. R. Fennema, ed., Food Chemistry, 3rd ed., 1996, pp. 226319. 46. J. M. deMan and L. deMan, in N. Widlak, R. Hartel, S. Narine, eds., Crystallization and Solidication Properties of Lipids, AOCS Press, Champaign, Illinois, 2001, 225235. 47. C. W. Hoerr, J. Amer. Oil Chem. Soc., 37, 539546 (1960). 48. P. Walstra, T. J. Geurts, A. Noomen, A. Jellema, and M. A. J. A. van Boekel, in P. Walstra, ed., Dairy Technology: Principles of Milk Properties and Processes, Marcel Dekker, New York, 1999, pp. 5070. 49. S. S. Narine and A. G. Marangoni, Food Res. Int., 32, 227248 (1999). 50. S. S. Narine and A. G. Marangoni, Adv. Food Nutr. Res., 44, 33145 (2002). 51. A. G. Marangoni, Trends Food Sci. Technol., 13, 3747 (2002). 52. R. Vreeker, L. L. Hoekstra, D. C. den Boer, and W. G. M. Agterof, Food Hydrocolloids, 6, 423435 (1992). 53. A. G. Marangoni and D. Rousseau, J. Amer. Oil Chem. Soc., 73, 991994 (1996). 54. S. S. Narine and A. G. Marangoni, Phys. Rev. E, 59, 19081920 (1999). 55. W. H. Shih, W. Y. Shih, S. I. Kim, J. Liu, and I. A. Aksay, Phys. Rev. A, 42, 47724779 (1990). 56. S. S. Narine and A. G. Marangoni, Phys. Rev. E, 60, 69917000 (1999). 57. A. G. Marangoni, Phys. Rev. B, 62, 1395113955 (2000). 58. A. G. Marangoni and M. A. Rogers, Appl. Phys. Lett., 82, 13 (2003). 59. T. Hagiwara, H. Kumagi, and K. Nakamura, Food Hydrocolloids, 12, 2936 (1998). 60. M. Tather and A. Erdem-Senatalar, J. Phys. Chem. B, 103, 43604365. 61. K. Solecka-Cermakova, B. Vlckova, and F. Lednicky, J. Phys. Chem., 100, 49544960 (1996). 62. A. Y. Kim and J. C. Berg, Langmuir, 16, 21012104 (2000). 63. G. Joss. 2002. Available: http://list.nih.gov/cgi-bin/wa?A2ind0101&Limagej&D 0&P5222. 64. J. C. Russ, in D. B. Williams, A. R. Pelton, and R. Gronsky, eds., Images of Materials, Oxford University Press, New York, 1991, 338373. 65. Trusoft International Incorporated. (2000). Available: http://www.trusoft.com. 66. H. Rohm and K. H. Weidinger, J. Texture Studies, 24, 157172 (1993). 67. J. F. Steffe, in Rheological Behaviour of Butter at Small Deformations, 2nd ed., Freeman Press, East Lansing, Michigan, 1996, pp. 294349. 68. M. A. Rao, in M. A. Rao, ed., Rheology of Fluid and Semisolid Foods: Principles and Applications, Chapman & Hall, Gaithersburg, Maryland, 1999, pp. 59152. 69. S. S. Narine and A. G. Marangoni, Lebensmittel Wissenschaft und Technologie, 34, 3340 (2001).

Potrebbero piacerti anche