Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Food Hydrocolloids 19 (2005) 269–278

www.elsevier.com/locate/foodhyd

Cold gelation of b-lactoglobulin oil-in-water emulsions


Valérie Leung Sok Line, Gabriel E. Remondetto, Muriel Subirade*
Chaire de recherche du Canada sur les protéines, les bio-systèmes et les aliments fonctionnels, Centre de recherche en sciences et technologie du lait (STELA),
Institut des nutraceutiques et des aliments fonctionnels (INAF), Faculté des sciences de l’agriculture et de l’alimentation, Université Laval,
Pavillon Paul-Comtois, QC, Canada G1K 7P4
Received 20 February 2004; revised 7 June 2004; accepted 7 June 2004

Abstract
The cold gelation of b-lactoglobulin emulsion gels was investigated in this work. The effects of oil and calcium concentrations on the
rheological properties and structure of emulsion gels were studied using dynamic small strain rheometry and electron microscopy techniques.
Results show that the storage modulus (G 0 ) is only affected by the oil content, the effect of which is greater than that of calcium. Emulsion
gels with high G 0 and good water-holding capacity (WAC) were obtained by raising the oil concentration. In contrast, an increase in salt
concentration reduced the WHC and changed the structure of the emulsion gel from fine-stranded to random aggregates. Oil and calcium
regulated the process of gel network formation and modulated the functional properties of cold-set emulsion gels.
q 2004 Elsevier Ltd. All rights reserved.

Keywords: b-Lactoglobulin; Emulsion gel; Cold gelation; Rheology; Microstructure

1. Introduction by different rheological and microstructural properties, may


be produced depending on pH conditions (Stading &
Whey proteins (WP) are widely used as food ingredients, Hermansson, 1990, 1991). Opaque particulate gels com-
because of their high nutritional value and their remarkable posed of random aggregates occur at pH values near the
functional properties (de Wit, 1989; Morr & Ha, 1993). isoelectric point of b-lg (pI) (pH 4–6), whereas transparent
Among these is their gel-forming ability, which is fine-stranded gels are obtained by shifting the pH far from
particularly of great importance to the food industry. the pI. Both types of gels have been recently characterized
Gelation of b-lactoglobulin (b-lg), the main WP, is and differentiated at a molecular level (Lefèvre & Subirade,
traditionally achieved through heat treatment. When con- 2000).
centrated b-lg solutions are heated above 65 8C, the globular WP, especially b-lg, also show good emulsifying
protein molecules partially unfold, expose their hydro- properties. Because of their flexibility and amphiphilic
phobic and sulfhydryl groups, and aggregate to create a nature, these proteins rapidly adsorb to the emulsion
three-dimensional network that entraps water by capillary interface where they self-aggregate and form continuous
forces (Twomey, Keogh, Mehra, & O’Kennedy, 1997). and homogeneous membranes around the oil droplets
In salt-free solutions, two types of heat-set b-lg gels (i.e., through intermolecular b-sheets interactions (Lefèvre &
‘fine-stranded’ and ‘particulate’), which are characterized Subirade, 2003). By coating the oil droplets with charged
layers, the protein films provide an electrostatic barrier
Abbreviations: ANOVA, analysis of variance; b-lg, b-lactoglobulin; d, against flocculation and coalescence (Phillips, Whitehead,
phase angle; G 0 , storage modulus; G 00 , loss modulus; pI, isoelectric point; & Kinsella, 1994). Oil-in-water emulsions stabilized by
SEM, scanning electron microscopy; TEM, transmission electron whey protein isolates (WPI) or by b-lg can be converted into
microscopy; WHC, water-holding capacity; WP, whey proteins; WPI, heat-set gels. Moreover, over the past few years, there has
whey protein isolates.
* Corresponding author. Tel.: C1-418-656-2131x4278; fax: C1-418-
been a growing interest in these composite systems,
656-3353. because of their practical application in food formulations
E-mail address: muriel.subirade@aln.ulaval.ca (M. Subirade). (Dickinson & Chen, 1999). In fact, several workers have
0268-005X/$ - see front matter q 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.foodhyd.2004.06.004
270 V.L. Sok Line et al. / Food Hydrocolloids 19 (2005) 269–278

investigated the microstructure and rheological properties of determine the water-holding capacity (WHC), permeability,
such emulsion gels (Chen & Dickinson, 1998; Dickinson, texture, and appearance of the gel (Bryant & McClements,
Hong, & Yamamoto, 1996; Jost, Baechler, & Masson, 1986; 2000).
Jost, Dannenberg, & Rosset, 1989; McClements, Monahan, As for all WP gels, modulation of microstructure and
& Kinsella, 1993). They found that droplets covered by WP functional properties of WP emulsion gels is possible by
become an integral part of the network and increase the gel cold gelation. Nevertheless, to our knowledge, no exhaus-
elastic modulus. This behavior is consistent with the tive reports on the cold-induced gelation of WP emulsion
mechanical reinforcement of the gel matrix by strongly gels are available, despite the fact that WP beads have been
interacting filler particles (Van Vliet, 1988). Other studies successfully produced using an emulsification/cold-gelation
have also revealed that the viscoelasticity of these materials process in a previous study (Beaulieu, Savoie, Paquin, &
is mainly controlled by the type of interactions between the Subirade, 2002). Thus, our objective was to develop cold-set
gel matrix and the filler particles, although oil volume b-lg emulsion gels and investigate the effects of oil and
fraction and average droplet size also play a role Ca2C concentrations on their rheological and structural
in determining the texture of emulsion gels (Chen & properties.
Dickinson, 1999a; Chen, Dickinson, Langton, &
Hermansson, 2000; Dickinson & Chen, 1999). The use of
WP emulsion gels extends the possibilities for creating
2. Material and methods
foods with new and improved organoleptic properties.
However, the thermal treatment needed to produce these
emulsion gels limits their application in formulations 2.1. Materials
containing heat-sensitive ingredients. An alternate gelation
method, involving low temperatures (i.e., i cold gelation), b-Lactoglobulin (b-lg) was obtained from Davisco
could be judiciously exploited to overcome this limitation. International, Inc. (Le Sueur, MN). It contained 98.2%
According to Barbut and Foegeding (1993), WPI or b-lg protein determined by semi-micro Kjeldahl method
gel production at 25 8C is achieved through two consecutive (AOAC, 1984) using N-factor 6.38. Sunflower oil was
steps: the preparation of a heat-denatured protein suspen- purchased from a local retailer and calcium chloride (CaCl2)
sion and the subsequent formation of a network through the was provided by Fisher Scientific (Fair Lawn, NJ).
addition of Ca2C to the chilled solution. The preheating
step, during which no gelation occurs, is essential as it 2.2. Preparation of cold-set emulsion gels
causes the protein molecules to open their structure and
interact with each other (Barbut & Foegeding, 1993; Protein suspensions (9.5 wt% protein) were prepared by
Hongsprabhas & Barbut, 1997b). Typically, a low ionic dispersing the powdered b-lg in deionized water and stirring
strength solution of native WP is kept at a pH far from the for 1 h to ensure complete dissolution. After having adjusted
isoelectric point, usually at pH 7, and held for 5–60 min the pH at 7 with 1 N HCl or NaOH, the suspensions
between 70 and 90 8C. Protein concentration is also were degassed for 30 min, heated at 85 8C during 45 min
maintained below that which is required to create a network in a water bath, and cooled to room temperature (w23 8C)
(Bryant & McClements, 1998). After the preheating for 2 h.
treatment, the protein molecules form aggregates without Oil-in-water coarse emulsions (6.65 wt% protein) were
gelation. These aggregates constitute the structural units produced by mixing appropriate amounts of protein solution
responsible for the three-dimensional network of cold-set and sunflower oil with a high speed blender operating at
gels (Remondetto & Subirade, 2003). After cooling, salt is 20,000 rpm for 2 min (Ultra-Turrax, Janke and Kunkel,
added to screen the repulsive forces between the aggregated IKA-Labortechnik, Staufen, Germany). The size of emul-
protein molecules, which can then form a gel. Besides sion droplets was then reduced further using a high-pressure
charge dispersion, a divalent salt, such as CaCl2, induces homogenizer (Emulsiflex-C5, Avestin Inc., Ottawa,
cross-linking of proteins and thus promotes gelation at much Canada). A two-stage homogenization was performed: the
lower concentrations than a monovalent salt (Bryant & first pass was carried out at 100 MPa and the second at
McClements, 2000; Hongsprabhas & Barbut, 1997a). 10 MPa. Emulsions were then degassed for 30 min to
The amount of salt used to form a cold-set gel is likely to remove all the air bubbles entrapped during homogenization
be the major determinant of the structure and spatial as their presence can lead to irreproducibility in subsequent
organization of protein aggregates (Hongsprabhas, Barbut, rheological measurements (Dickinson & Hong, 1995).
& Marangoni, 1999). Many studies confirm that the network To induce cold gelation, emulsions were mixed with
formation processes are governed by salt concentration CaCl2 by vortexing at 23 8C. The levels of oil and CaCl2 to
resulting in different gelation mechanisms and in various be added were determined using a central rotatable
structure types (Hongsprabhas & Barbut, 1997b, 1998; composite experimental design as shown in Table 1. The
Hongsprabhas et al., 1999; Remondetto, Paquin, & final protein concentration in all formulations was adjusted
Subirade, 2002; Remondetto & Subirade, 2003), which to 6.5 wt%.
V.L. Sok Line et al. / Food Hydrocolloids 19 (2005) 269–278 271

Table 1 2.5. Rheology of cold-set emulsion gels


Experimental range and levels of the independent variables established
according to the Box–Wilson central composite design for cold gelation of
b-lactoglobulin emulsionsa
Small deformation viscoelasticity of emulsion gels was
investigated using dynamic oscillatory measurements per-
Treatment Variables formed by a controlled-stress rheometer equipped with
Oil concentration (%) Calcium concentration (mM) 40 mm diameter parallel plates (SR-5000, Rheometric
Coded X1 Uncoded Coded X2 Uncoded Scientific, Piscataway, NJ). The phase angle d and the
[Oil] [Ca2C] storage and loss moduli, G 0 and G 00 , were monitored as a
1 1 28 1 60 function of time in the linear viscoelastic region at 23 8C and
2 1.414 31.1 0 40 at a frequency of 1 Hz. All measurements were conducted for
3 0 20.5 K1.414 11.7 15 h and a solvent trap was used to prevent sample drying.
4 K1.414 9.9 0 40
5 0 20.5 0 40
6 0 20.5 0 40 2.6. Scanning and transmission electron microscopy
7 0 20.5 0 40 (SEM and TEM) of cold-set emulsion gels
8 1 28 K1 20
9 0 20.5 0 40 The microstructure of cold-set emulsion gels was
10 K1 13 K1 20
11 K1 13 1 60
analyzed by transmission electron microscopy (TEM) and
12 0 20.5 1.414 68.3 scanning electron microscopy (SEM). Samples (1!1!
13 0 20.5 0 40 2 mm) were fixed in 2.5% glutaraldehyde C1.5%
a paraformaldehyde in 0.1 M cacodylate buffer for 24 h,
The experimental runs were performed in random order.
rinsed and post-fixed in 1% OsO4 for 90 min. For a better
2.3. Particle size analysis penetration of the fixative solution in the floating gel
samples, vacuum was applied for 1 h during the fixation
The droplet size distribution of freshly made emulsions step. Dehydration was accomplished in a graded series of
was determined by photon correlation spectroscopy ethanol. For TEM, the samples were immersed in propylene
(Nicomp Submicron Particle Sizer System, Model C370, oxide, propylene oxide/Epon solution (1:1), and finally pure
Pacific Scientific, Manlo Park, CA) with the method of Epon. After infiltration overnight at room temperature, they
Robin and Paquin (1991). Emulsion samples were dispersed were embedded in Epon, with polymerization at 60 8C,
in a dissociating buffer composed of urea (8 M), EDTA thinly sectioned, stained with uranyl and lead acetate, and
(50 mM), and b-mercaptoethanol (10 mM). Dilute emul- viewed at 80 kV (JEOL Model 1200EX, JEOL Ltd.
sions, placed into the measurement cell of the instrument, Akishima, Japan). For SEM, the samples were critical
were illuminated by a focused laser beam and fluctuations in point dried, sputter coated with 30 nm of gold/palladium,
the intensity of the scattered light were measured by a and examined at 15 kV using a JEOL JSM-6360 LV
photomultiplier detector at an angle of 908. Particle instrument.
diffusivity was computed from the autocorrelation function
of the scattered light by the method of cumulants and 2.7. Experimental design
translated into a particle diameter value using the Stokes–
Einstein relation for spheres (Nicoli, McKenzie, & Wu, A response surface methodology study was conducted to
1991). Each determination was performed in triplicate. determine the simultaneous effects of oil and Ca2C
concentrations on the rheological and water-holding proper-
ties of cold-set emulsion gels. A 22 factorial Box-Wilson
2.4. Water-holding capacity (WHC) central composite design with four-star points and five
replicates at the center point, leading to a set of 13
The WHC was evaluated as described by Remondetto experiments performed in random order was used. Five
et al. (2002). Emulsion gels were formed in an ultrafiltration levels of each factor (oil and Ca2C concentrations) were
unit (Ultrafree-CL Filters, Millipore Co., Billerica, MA.) incorporated into the design. Maximum and minimum ranges
consisting of a collection tube (5 ml) and a filter cup (2 ml) of independent variables investigated were chosen by
with an ultrafiltration membrane of 5000 NMWL. After carrying out preliminary tests. The full experimental plan,
24 h, the unit was centrifuged at 4000 g for 20 min. WHC with respect to the variable values in their actual and coded
was calculated as: forms, is listed in Table 1. Assessment of error was derived
WT K WF from replication of the central point treatment conditions.
WHC Z !100%
WT
2.8. Statistical analysis
where WT is the total amount (g) of water in the sample and
WF is the quantity of water (g) released. Each determination Regression analysis was performed on the obtained data.
was carried out in five replicates. The following second-order equation was fitted for each
272 V.L. Sok Line et al. / Food Hydrocolloids 19 (2005) 269–278

factor assessed: Table 2


Results of storage modulus (G 0 ) and WHC for each condition of
Y Z b0 C b1 X1 C b2 X2 C b11 X12 C b22 X22 C b12 X1 X2 experimental design

where Y is the estimated response, b0, b1, b2, b11, b22, b12 Treatment G 0 (Pa) WHC (%)
are constant and regression coefficients of the model, and 1 28,870 90
X1, X2 are the independent variables. Analysis of variance 2 29,680 92
3 10,048 94
(ANOVA) with Fisher’s test for significance was carried out
4 3693 78
to evaluate linear, quadratic, and interactive effects of oil 5 13,884 87
and Ca2C concentrations, and to determine the adequacy of 6 9797 86
the experimental model. P%0.05 was considered to be 7 11,044 85
significant. Proportion of variance explained by the model 8 31,184 94
9 14,839 87
was given by the multiple coefficient of regression R2.
10 8543 85
Statgraphics Plus Professional version 4.1 (Manugistics 11 5096 79
Inc., Rockville, MD) was used to generate and analyze the 12 11,850 87
experimental design. 13 10,972 87

3. Results that the significant terms are both the linear and quadratic
components of oil and Ca2C concentrations. Fig. 2 displays
3.1. Emulsion properties the response surface plot. It can be seen that oil and Ca2C
had antagonistic effects on the ability of emulsion gels to
The particle size distribution of freshly prepared emul- physically hold water. Indeed, salt addition decreased the
sions containing various concentrations of oil was WHC values by promoting the expulsion of water, while the
measured. The volume frequency of oil droplets as a presence of oil improved the water retention of the emulsion
function of size is presented in Fig. 1. All emulsions had a gels. In this case, the regression model correctly described
mean particle diameter ranging from 0.68 to 0.80 mm and the collected data (lack of fit, pO0.05). Its coefficients are
a monomodal particle size distribution with an almost listed in Table 4. According to the R2 value, 98% of the
symmetrical peak. Fig. 1 reveals that the droplet size variability in the response can be attributed to the model. It
distribution was not significantly different between the is then possible to use this model to estimate the cold-set
emulsions, suggesting here a negligible effect of the amount b-lg emulsion gel WHC values when experimenting with
of oil added. These preliminary results ensured that different oil and calcium concentrations.
subsequent changes observed in emulsion gel oil droplet
size were attributed only to the cold gelation process. 3.3. Rheological measurements of emulsion gels

3.2. Water-holding capacity The small deformation rheological properties of emul-


sion gels can be characterized by the storage modulus (G 0 ),
Results of WHC for all the experimental conditions are the loss modulus (G 00 ), and the phase angle (d). G 0 is a
summarized in Table 2. Variance analysis (Table 3) shows measure of the energy stored per cycle of deformation and
its value reflects the solid-like or elastic behavior of the
tested material, while G 00 refers to the energy dissipated by
the sample and is associated with the fluid-like or viscous

Table 3
Analysis of variance of storage modulus (G 0 ) and WHC

Source of Degrees of Sum of squares


variation freedom G0 WHC
8*
X1:[Oil] 1 8.65!10 1.98!102*
X2:[Ca2C] 1 1.29!106 4.95!101*
X1$X1 1 8.24!107** 5.48**
X1$X2 1 3.21!105 1.00
X2$X2 1 2.28!106 2.41!101*
Lack of fit 3 6.27!107 1.13
Error 4 1.84!107 3.20
Total (corr.) 12 1.03!109 2.86!102
R2 92.12 98.49
Fig. 1. Effect of oil on particle size distribution of 6.65 wt% b-lactoglobulin
oil-in-water emulsions (pHZ7). *p!0.05; **p!0.01.
V.L. Sok Line et al. / Food Hydrocolloids 19 (2005) 269–278 273

Fig. 2. Response surface graph of the effect of oil and Ca2C concentrations
on the water-holding capacity of cold-set b-lg emulsion gels containing
6.5 wt% protein (pHZ7).

characteristics of the material (Bell, 1989; Tadros, 1996).


The phase angle, which is the arc tangent of the ratio G 00 /G,
corresponds to the response delay of strain to the applied Fig. 3. Evolution of G 0 , G 00 , and tan d during cold-set b-lg emulsion gel
stress and indicates the relative importance of viscous and formation (6.5 wt% protein, 28 wt% oil, and 20 mM Ca2C) at 23 8C
elastic elements in the emulsion gel. For example, d values and 1 Hz.
of 0 and 908 are typical for perfect elastic solids and viscous
liquids, respectively. When the value of G 0 is far larger than the incorporation of oil droplets significantly improved
the value of G 00 , d is very small and the material is said to be the elasticity of cold-set b-lg gels at any given Ca2C
predominantly elastic (Chen & Dickinson, 1999a). concentration. For example, adding up to 30 wt% oil in the
These three parameters were recorded during the protein network enhanced about 10-fold the gel storage
emulsion gel formation. All samples showed the same modulus. Although Ca2C was necessary to build up the
type of gelation curves: both G 0 and G 00 increased similarly three-dimensional network, it had a negligible effect on the
to reach a maximum value, while d decreased to a plateau elasticity of the emulsion gels. The regression coefficients
(Fig. 3). Because G 0 values were considerably greater in of the model used to describe the evolution of the storage
magnitude than G 00 values and because they suggested the modulus as a function of oil and Ca2C concentrations are
formation of elastic emulsion gels, only the results of G 0 at displayed in Table 4. Since the lack of fit test was not
the end of the gelation process are presented and discussed significant (pO0.05), the model appears to be adequate for
hereafter (Table 2). the observed data at the 95% confidence level and can
The analysis of variance, presented in Table 3, indicates explain 92% of the variability in the response, as indicated
that, among the variables studied, only the oil concen- by the R2 value. Thus, this model can be used to predict the
tration significantly affected the rheological properties of rheological behavior of cold-set b-lg emulsion gels at
the cold-set emulsion gels. Varying oil levels had a strong various oil concentrations.
positive effect on the storage modulus, the linear and
quadratic components both being significant. As shown in
3.4. Microstructure analysis
Fig. 4, the greater was the oil content, the more important
was the value of G 0 . Such increase in the modulus of the
Electron micrographs were taken using TEM and SEM in
emulsion gels suggests that the dispersed oil droplets act as
order to get insight into the structure of cold-set emulsion
active fillers that interact with the gel matrix. This is
gels obtained with different levels of added oil and Ca2C.
confirmed by Fig. 5 where the storage modulus of cold-set
b-lg emulsions gels is compared with that of an equivalent
gel without oil droplets. As it can be observed,

Table 4
Coefficients of the regression models for storage modulus (G 0 ) and WHC

Treatment G0 WHC
4
b0 1.40!10 8.13!101
b1 –1.20!103* 1.18*
b2 K1.73!102 K5.65!10K1*
b1$b1 6.12!101** K1.58!10K2**
b1$b2 1.89 3.33!10K3
b2$b2 1.43 4.66!10K3* Fig. 4. Response surface graph of the effect of oil and Ca2C concentrations
on the storage modulus of cold-set b-lg emulsion gels containing 6.5 wt%
*
p!0.05; **p!0.01. protein (pHZ7).
274 V.L. Sok Line et al. / Food Hydrocolloids 19 (2005) 269–278

that increasing the oil volume fraction, while keeping Ca2C


concentration constant, led to emulsion gels with higher
density and smaller pores (Fig. 6a and b), mainly due to
space filling by the oil droplets and their interaction with the
protein matrix. At 9.9% oil and 40 mM Ca2C (Fig. 6a),
the gelled emulsion has an inhomogeneous and porous
structure. However, when the oil content was elevated to
31.1%, the matrix, in which the oil droplets were evenly
distributed, was denser (Fig. 6b). Micrographs from TEM
agreed with those from SEM: when the oil concentration
was more important, the network structure became more
compact (Fig. 6c and d). Fig. 6c and d also show that
increasing the oil volume fraction leads to changes in the
structure of emulsion gels containing the same Ca2C
concentration from a particulate to a mixed-type gel,
where both fine-stranded and random aggregates are found.
Fig. 5. Storage modulus of cold-set b-lg gels containing 6.5 wt% protein, On the other hand, raising Ca2C concentration and
with or without oil, and at different calcium concentrations after 15 h. maintaining the oil volume fraction constant resulted in
larger pores and larger protein aggregates. For example, at
A selection of conditions was tested rather than the full 11.7 mM Ca2C, fine protein aggregates were uniformly
experimental design. Four samples were then produced by dispersed in the matrix and linked together by thin strands
combining the central point of a given factor with the (Fig. 7a). At 68 mM Ca2C, the structural difference is
extreme levels of the other factor. The microstructure of the obvious as the aggregates clumped together to form larger
analyzed samples is presented in Figs. 6 and 7. TEM shows particles, which were well separated from each other by

Fig. 6. Transmission (TEM) and scanning (SEM) electron micrographs of cold-set b-lg emulsion gels containing 6.5 wt% protein, prepared at pH 7 and
constant Ca2C concentration of 40 mM in the presence of: 9.9 wt% oil, TEM (a); 31.1 wt% oil, TEM (b); 9.9 wt% oil, SEM (c); and 31.1 wt% oil, SEM (d).
V.L. Sok Line et al. / Food Hydrocolloids 19 (2005) 269–278 275

Fig. 7. Transmission (TEM) and scanning electron micrographs (SEM) of cold-set b-lg emulsion gels containing 6.5 wt% protein, prepared at pH 7 and
constant oil concentration of 20.5 wt% in the presence of: 11.7 mM Ca2C, TEM (a); 68 mM Ca2C, TEM (b); 11.7 mM Ca2C, SEM (c); 68 mM Ca2C, SEM (d).

the aqueous phase (Fig. 7b). As shown by TEM, exclusion or non-interactive systems (Mor, Shoemaker, & Rosenberg,
of some oil droplets from the protein matrix and their 1999). It has been suggested that, in non-interactive
tendency to flocculate together were also observed at higher systems, lipid globules only fill pores in the gel matrix
Ca2C concentration (Fig. 7b). Such a phenomenon is likely without interacting with the protein-based network. Such
related to the excessive calcium bridging between the particles are termed ‘inactive fillers’ and are known to act as
protein molecules and to the subsequent development of structure breakers because they have little or no affinity for
tightly aggregated structures. Varying Ca2C concentrations the gel matrix and cause its disruption and weakening.
had a notable effect on the aspect of the emulsion gels. This Increasing the volume fraction of these inactive fillers
effect can be appreciated in SEM micrographs, where a generally results in lower gel strengths. By contrast,
filamentous network, with small pore size, is evident at low interactive systems are characterized by strong interactions
calcium concentrations (Fig. 7c) and where the formation of between the lipid globules and the gel matrix. These
particulate structure composed of random aggregates is so-called ‘active fillers’ greatly enhance gel strength as they
observed upon further salt addition (Fig. 7d). are fully incorporated as a part of the network. This
The microstructure study indicates that the network reinforcement effect becomes more pronounced with greater
formation processes are governed by Ca2C concentration amounts of active fillers (Brownsey, Ellis, Ridout, & Ring,
and by oil volume fraction during the cold gelation of b-lg 1987; Jost et al., 1989; McClements et al., 1993; Ring &
emulsions.
Stainsby, 1982).
Cold-set b-lg emulsion gels produced in this work belong
to the category of interactive systems. In fact, the difference
4. Discussion observed between the elasticity of protein-only gel and that
of emulsion gels, and the steep increase in emulsion gel
Composite gels consisting of a protein matrix and a storage modulus, along with higher oil concentration, mean
dispersed lipid phase can be classified as either interactive that the dispersed oil droplets not only function as space
276 V.L. Sok Line et al. / Food Hydrocolloids 19 (2005) 269–278

fillers, but also help build up the gel structure. This is Foegeding, 1993; Hongsprabhas & Barbut, 1997b, 1998;
consistent with reported effects of WP-covered oil droplets Remondetto et al., 2002), where raising the mineral
on the rheological properties of heat-set emulsion gels content resulted in increasing protein aggregate size,
(Chen & Dickinson, 1998, 1999b; Dickinson & Chen, in forming a looser and more porous structure, and in
1999). Incorporation of oil globules into the matrix of these diminishing the WHC. Since the ability of gels to physically
heat-set gels is achieved through interactions between the hold water varies, among other, with the size of inner
protein molecules adsorbed at a droplet surface as well as structure spaces, gels with large and heterogeneous pores
those of the gel network during thermal treatment, resulting tended to bind water to a lesser extent, because of low
from a combination of disulfide, hydrophobic, and hydrogen capillary forces (Hermansson, 1986). High protein–protein
bonds (McClements et al., 1993). However, for the interactions caused by an increase in Ca2C level also
formation of cold gels, it has been shown that physical explain why water was more easily extracted from
interactions play a major role in the first stage of the process particulate gels.
while disulfide bonds are mainly involved in the second In this work, excessive Ca2C bridging between the
stage of cold-set gel formation either for acid-induced cold protein molecules led not only to the expulsion of water but
gelation (Alting, Hamer, de Kruif, Paques, & Visschers, also to the exclusion of some oil droplets that flocculated
2003, Alting, Hamer, de Kruif, & Visschers, 2003) or for afterwards. Studies on heat-set emulsion gels have revealed
cation-induced cold gelation (Remondetto & Subirade, that the degree of incorporation of oil droplets into the
2003). Since the gelation of our emulsions was conducted protein network affect the gel rheological properties
at room temperature (23 8C) upon calcium salt addition, (McClements et al., 1993; Chen & Dickinson, 1999b). It
inclusion of dispersed oil droplets in the protein network is seems therefore reasonable to infer that aggregated droplets,
thought to be realized via these interactions. obtained at high Ca2Cconcentrations, have disruptive
Increasing oil content in the emulsion led to a greater effects on the protein matrix and hence weaken cold-set
number of oil globules fitted into the gel matrix, which emulsion gels. One would then expect a decrease in G 0
served as many anchor points that strengthened the three- values by adding more Ca2C, especially as particulate gels
dimensional network. This explains, in part, the high are also known to be less elastic than fine-stranded ones
values of G 0 obtained on elevating the quantity of oil (Stading & Hermansson, 1991). However, as seen earlier,
added in emulsion gels as the mechanical properties of the effect of calcium on the emulsion gel storage modulus
this type of composite material may also originate from was not significant. A possible explanation for the
the packing of oil droplets in the network. Indeed, when rheological behavior at high Ca2C concentrations is the
the oil globules are concentrated above the random close greater effective volume fraction for a system of aggregated
packing limit, they behave as perfectly elastic solids droplets when compared to the sum of the volumes of
(Bressy, Hébraud, Schmitt, & Bibette, 2003; Hemar & individual oil droplets (Van Vliet, 1988). A consequence of
Horne, 2000). Although the critical oil volume fraction this is a higher emulsion gel storage modulus that
was not reached in this work, shifting from the lowest to overcomes the negative effect of aggregated structures
the highest oil concentration was sufficient for the droplets created at high Ca2C levels.
to sustain more stress and possess more elastic Oil and calcium are both clearly involved in the cold-set
characteristics. b-lg emulsion gel network building process. Thus, it is
The elasticity provided by the oil droplets to the possible to manipulate the final gel characteristics by
emulsion gels appears precisely to be the main reason for adjusting the proportion of oil included in the matrix and
the greater WHC observed at higher oil content as it enabled the amount of added calcium. In fact, for given calcium
the samples to withstand the mechanical forces during concentration, emulsion gel elasticity and water retention
centrifugation. As suggested by other studies (Alzagtat & can be enhanced by increasing the oil volume fraction. On
Alli, 2002), the improvement of WHC could also be the other hand, for given oil concentration, emulsion gel
attributed, to a certain degree, to an enhancement of the gel structure type and water-holding abilities can be modified
structure density by the presence of oil droplets, which by varying the calcium level. If a filamentous emulsion gel
causes the water to be trapped inside the pores by capillary with superior water retention capabilities is required, a low
forces. calcium concentration should be used whereas for a
Changes in the emulsion gel microstructures and in the syneresing particulate emulsion gel, the use of high calcium
WHC with the addition of calcium salt confirmed that Ca2C concentration should be considered.
was also an integral part of the protein network building
process. Emulsion gels obtained at low salt concentrations
were fine-stranded and retained more water than those 5. Conclusion
containing particulate type structures and formed at high
Ca2C concentrations. The pattern observed for the cold-set This study demonstrates that gelation of b-lactoglobu-
emulsion gels agrees with previous studies on lin emulsion gels can be achieved at room temperature
divalent cation-induced WP gels (Barbut, 1995; Barbut & and that modulation of their microstructure and functional
V.L. Sok Line et al. / Food Hydrocolloids 19 (2005) 269–278 277

properties is possible by varying the oil and calcium Chen, J., & Dickinson, E. (1998). Viscoelastic properties of heat-set whey
concentrations. Emulsion gels with a high storage protein emulsion gels. Journal of Texture Studies, 29, 285–304.
Chen, J., & Dickinson, E. (1999a). Effect of monoglycerides and
modulus and a good WHC are produced by increasing diglycerol-esters on viscoelasticity of heat-set whey protein
the oil content from 9.9 to 31.1 wt%. However, raising emulsion gels. International Journal of Food Science and Technology,
Ca2C levels from 11.7 to 68.3 mM causes the structure of 34, 493–501.
emulsion gels to shift from filamentous to particulate type Chen, J., & Dickinson, E. (1999b). Interfacial aging effect on the rheology
and leads to a decrease in the water retention capacity of heat-set protein emulsion gel. Food Hydrocolloids, 13, 363–369.
Chen, J., Dickinson, E., Langton, M., & Hermansson, A.-M. (2000).
without affecting the gel rheological properties. The Mechanical properties and microstructure of heat-set whey protein
balance between these two antagonistic effects results in emulsion gels: effect of emulsifiers. Lebensmittel-Wissenschaft und-
emulsion gels of unique properties that can be tailored to Technologie, 33, 299–307.
specific food applications. de Wit, J. N. (1989). Functional properties of whey proteins. In P. N. Fox
(Ed.), Functional properties of whey proteins (pp. 285–321). London,
NY: Applied Science Publishers.
Dickinson, E., & Chen, J. (1999). Heat-set whey protein emulsion gels: role
Acknowledgements of active and inactive filler particles. Journal of Dispersion Science and
Technology, 20, 197–213.
The authors acknowledge Anne-Françoise Allain for her Dickinson, E., & Hong, S.-T. (1995). Influence of water-soluble nonionic
technical assistance. This work was supported by a Natural emulsifier on the rheology of heat-set protein-stabilized emulsion gels.
Journal of Agricultural and Food Chemistry, 43, 2560–2566.
Sciences and Engineering Research Council of Canada Dickinson, E., Hong, S.-T., & Yamamoto, Y. (1996). Rheology of heat-set
(NSERC) fellowship (Valérie Leung Sok Line) and by the emulsion gels containing (b-lactoglobulin and small-molecule surfac-
NSERC Canada Research Chairs Program (Muriel tants. Netherlands Milk and Dairy Journal, 50, 199–207.
Subirade). Hemar, Y., & Horne, D. S. (2000). Dynamic rheological properties of
highly concentrated protein-stabilized emulsions. Langmuir, 16, 3050–
3057.
Hermansson, A.-M. (1986). Water and fat holding. In J. R. Mitchell, &
References D. A. Ledward (Eds.), Water and fat holding (pp. 273–314). London,
NY: Elsevier Applied Science.
Alting, A. C., Hamer, R. J., de Kruif, C. G., & Visschers, R. W. (2003). Hongsprabhas, P., & Barbut, S. (1997a). Protein and salt effects on Ca2C-
Cold-set globular protein gels: interactions, structure and rheology as a induced cold gelation of whey protein isolate. Journal of Food Science,
function of protein concentration. Journal of Agricultural and Food 62, 382–385.
Chemistry, 51, 3150–3156. Hongsprabhas, P., & Barbut, S. (1997b). Structure-forming processes in
Alting, A. C., Hamer, R. J., de Kruif, C. G., Paques, M., & Visschers, R. W. Ca2C-induced whey protein isolate cold gelation. International Dairy
(2003). Number of thiol groups rather than the size of the aggregates Journal, 7, 827–834.
determines the hardness of cold set whey protein gels. Food Hongsprabhas, P., & Barbut, S. (1998). Ca2C-Induced cold gelation of
Hydrocolloids, 17, 469–479. whey protein isolate: effect of two-stage gelation. Food Research
Alzagtat, A. A., & Alli, I. (2002). Protein–lipid interactions in food International, 30, 523–527.
systems: a review. International Journal of Food Sciences and Hongsprabhas, P., Barbut, S., & Marangoni, A. G. (1999). The structure of
Nutrition, 53, 249–260. cold-set whey protein isolate gels prepared with CaCC. Lebensmittel-
AOAC (1984). Official methods of analysis. Arlington: Association of Wissenschaft und-Technologie, 32, 196–202.
Official Analytical Chemists. Jost, R., Baechler, R., & Masson, G. (1986). Heat gelation of oil-in-water
Barbut, S. (1995). Effects of calcium level on the structure of pre-heated emulsions stabilized by whey protein. Journal of Food Science, 51,
whey protein isolate gels. Lebensmittel-Wissenschaft und-Technologie, 440–444.
28, 598–603. Jost, R., Dannenberg, F., & Rosset, J. (1989). Heat-set gels based on
Barbut, S., & Foegeding, E. A. (1993). Ca2C-induced gelation oil/water emulsions: an application of whey protein functionality. Food
of pre-heated whey protein isolate. Journal of Food Science, 58, Microstructure, 8, 23–28.
867–871. Lefèvre, T., & Subirade, M. (2000). Molecular differences in the formation
Beaulieu, L., Savoie, L., Paquin, P., & Subirade, M. (2002). Elaboration and structure of fine-stranded and particulate b-lactoglobulin gels.
and characterization of whey protein beads by an emulsification/cold Biopolymers, 54, 578–586.
gelation process: application for the protection of retinol. Biomacro- Lefèvre, T., & Subirade, M. (2003). Formation of intermolecular b-sheet
molecules, 3, 239–248. structures: a phenomenon relevant to protein film structure at oil–water
Bell, A. E. (1989). Gel structure and food biopolymers. In T. Hardman interfaces of emulsions. Journal of Colloid and Interface Science, 263,
(Ed.), Gel structure and food biopolymers (pp. 251–274). New York: 59–67.
Elsevier Applied Science. McClements, D. J., Monahan, F. J., & Kinsella, J. E. (1993). Effect of
Bressy, L., Hébraud, P., Schmitt, V., & Bibette, J. (2003). Rheology of emulsion droplets on the rheology of whey protein isolate gels. Journal
emulsions stabilized by solid interfaces. Langmuir, 19, 598–604. of Texture Studies, 24, 411–422.
Brownsey, G. J., Ellis, H. S., Ridout, M. J., & Ring, S. G. (1987). Elasticity Mor, Y., Shoemaker, C. F., & Rosenberg, M. (1999). Compressive
and failure in composite gels. Journal of Rheology, 31, 635–649. properties of whey protein composite gels containing fractioned
Bryant, C. M., & McClements, D. J. (1998). Molecular basis of protein milkfat. Journal of Food Science, 64, 1078–1083.
functionality with special consideration of cold-set gels derived Morr, C. V., & Ha, E. Y. W. (1993). Whey protein concentrates and
from heat-denatured whey. Trends in Food Science and Technology, isolates: processing and functional properties. Critical Reviews in Food
9, 143–151. Science and Nutrition, 33, 431–476.
Bryant, C. M., & McClements, D. J. (2000). Influence of NaCl and CaCl2 Nicoli, D., McKenzie, D., & Wu, J.-S. (1991). Application of dynamic light
on cold-set gelation of heat-denatured whey proteins. Journal of Food scattering to particle size analysis of macromolecules. American
Science, 65, 801–804. Laboratory , 32–40.
278 V.L. Sok Line et al. / Food Hydrocolloids 19 (2005) 269–278

Phillips, L. G., Whitehead, D. M., & Kinsella, J. E. (1994). Structure– Stading, M., & Hermansson, A.-M. (1990). Viscoelastic behaviour of b-
function properties of food proteins. San Diego: Academic Press. lactoglobulin gel structures. Food Hydrocolloids, 4, 121–135.
Remondetto, G., Paquin, P., & Subirade, M. (2002). Cold gelation of Stading, M., & Hermansson, A.-M. (1991). Large deformation properties of
b-lactoglobulin in the presence of iron. Journal of Food Science, 67, b-lactoglobulin gel structure. Food Hydrocolloids, 5, 339–352.
586–595. Tadros, T. F. (1996). Correlation of viscoelastic properties of stable and
Remondetto, G., & Subirade, M. (2003). Molecular mechanisms of Fe2C- flocculated suspensions with their interparticle interactions. Advances
induced b-lactoglobulin cold gelation. Biopolymers, 69, 461–469. in Colloid and Interface Science, 68, 97–200.
Ring, S., & Stainsby, G. (1982). Filler reinforcement of gels. Progress in Twomey, M., Keogh, M. K., Mehra, R., & O’Kennedy, B. T. (1997). Gel
Food and Nutrition Science, 6, 323–329. characteristics of b-lactoglobulin, whey protein concentrate and whey
Robin, O., & Paquin, P. (1991). Evaluation of the particle size of fat protein isolate. Journal of Texture Studies, 28, 387–403.
globules in a milk model emulsion by photon correlation spectroscopy. Van Vliet, T. (1988). Rheological properties of filled gels. Influence of filler
Journal of Dairy Science, 74, 2440–2447. matrix interaction. Colloid and Polymer Science, 266, 518–524.

You might also like