You are on page 1of 7

LWT - Food Science and Technology 117 (2020) 108690

Contents lists available at ScienceDirect

LWT - Food Science and Technology


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

Improvement of calcium sulfate-induced gelation of soy protein via T


incorporation of soy oil before and after thermal denaturation
Haibo Zhaoa,b, Bin Yua, Yacine Hemara, Jie Chenb,∗∗, Bo Cuia,∗
a
State Key Laboratory of Biobased Material and Green Papermaking, College of Food Science and Engineering, Qilu University of Technology, Shandong Academy of
Sciences, Jinan, 250353, China
b
State Key Laboratory of Food Science and Technology, School of Food Science and Technology, Jiangnan University, Wuxi, 214122, China

A R T I C LE I N FO A B S T R A C T

Keywords: Soy protein-based foods have attracted extensive attention for their potential health benefits. In this study, soy
Soy protein isolate protein isolate (SPI) emulsions containing 10–70 mL/L of soy oil were primarily prepared either by adding the
Tofu oil, emulsifying and heat treatment (OEH), or by heat treatment of the soy protein dispersion, addition of oil and
Texture emulsification (HOE). After adding CaSO4, the rheological properties of emulsions during gelation and the
Viscoelasticity
textural properties of the resulting gels were investigated and compared. The results showed that the increase in
Heat denaturation
the volume fraction (Φ) of the soy oil from 10 to 70 mL/L resulted in the increase of the final elastic modulus (G′)
of HOE and OEH emulsion gels from 5216.7 to 5827.3 Pa, and 4184.3–4935.3 Pa, respectively. The gel hardness
and water-holding capacity gradually improved with increasing Φ. HOE gels exhibited greater performance of
gelling properties than OEH gels at the same Φ. Confocal laser scanning microscopy images demonstrated that
HOE gels, with smaller soy oil droplets dispersed within the networks, had more homogeneous and compact
network structures than OEH gels. Reinforced emulsifying capacity of denatured SPI could enhance the protein-
oil interactions and result in improvement of CaSO4-gelation of SPI.

1. Introduction chemical and structural characteristics, which are, in turn, affected by


treatment parameters, processing methods and interactions with other
Soy products have attracted interest worldwide because of their ingredients (Zhao, Wang, Li, Qin, & Chen, 2017). Therefore, under-
potential health benefits, such as reducing the risk of heart and cardi- standing the effects of specific processing factors and other ingredients
ovascular diseases, lowering the cholesterol levels and preventing cer- on the gelation properties of soy protein isolate (SPI) will be very useful
tain cancers (Chang, 2006; Rekha & Vijayalakshmi, 2010). Tofu, also for the development of soy protein-based products, including tofu-type
called soy curd, is a typical gel product normally prepared by coagu- gel.
lating soy milk with acid (glucono-δ-lactone) or salt (e.g., CaSO4) Soy oil, the main constituent of soy milk in addition to soy protein,
(Nishinari, Fang, Guo, & Phillips, 2014; Obatolu, 2008; Ringgenberg, plays an important role in the texture, flavor and nutritional value of
Alexander, & Corredig, 2013). The process is relatively complicated and soy protein foods, including soy protein-based gels, such as tofu
cumbersome, including soaking and grinding of soya beans, heat (Fukushima, 1991). A gel containing soy oil droplets could be con-
treatment, separation of soy milk, coagulation, breaking of the curd and sidered a protein gel matrix in which the soy oil droplets are embedded
pressing to reform the gel (Chang, 2006; Corzo-Martinez, Garcia- (Kim, Renkema, & van Vliet, 2001). The soy oil droplets combine with
Campos, Montilla, & Moreno, 2016). Conversely, soy protein, which is protein particles and participate to form a stable tofu–curd emulsion
free of cholesterol and contains less saturated fat, is a versatile source of network (Guo, Ono, & Mikami, 1999). Murekatete, Zhang,
vegetable protein (Wang et al., 2018). The development of tofu-type gel Hategekimana, Karangwa, and Hua (2016) studied the effects of dif-
using soy protein as the basic material would be an alternative available ferent soy oil volume fractions (50, 150, 250 and 350 mL/L) on the
for the food manufacturers, restaurant, canteen, etc., which are not rheological behavior of acid- or salt-induced SPI gels, and the results
equipped with tofu processing facilities. showed that the storage modulus (G′) of acid-induced SPI gels de-
The gel properties of soy protein are related to their physical, creased with increasing soy oil volume fraction, while the influence was


Corresponding author.
∗∗
Corresponding author.
E-mail addresses: chenjie@jiangnan.edu.cn (J. Chen), cuibopaper@163.com (B. Cui).

https://doi.org/10.1016/j.lwt.2019.108690
Received 11 June 2019; Received in revised form 24 September 2019; Accepted 24 September 2019
Available online 25 September 2019
0023-6438/ © 2019 Elsevier Ltd. All rights reserved.
H. Zhao, et al. LWT - Food Science and Technology 117 (2020) 108690

opposite for salt-induced SPI gels. Kim et al. (2001) indicated that the generated SPI emulsions were denoted as OEH. The second SPI dis-
G′ and the loss modulus (G″) of SPI gels after the temperature cycle persion was initially heated at 90 °C for 15 min and cooled down to
increased as soy oil volume fractions increased from 100 to 300 mL/L. room temperature (25 °C) before incorporating the same volume frac-
(Tang et al., 2011) evaluated the effects of thermal treatment and tions of soy oil, and then SPI emulsions were prepared according to the
emulsification on the properties of CaCl2-or glucono-δ-lactone (GDL)- above procedure. The resultant SPI emulsions were denoted as HOE. A
induced SPI gels and concluded that the physicochemical properties of 70 g/L SPI dispersion free of soy oil under the same conditions was used
the gels have a close relationship with their microstructures. These as a control in this study.
findings strongly suggested that the incorporation of soy oil might be of
vital importance for the gelation behavior and network structures of 2.3. Formation of CaSO4-induced SPI gels
tofu-type gels.
However, it is surprising that in a model system close to traditional A freshly prepared CaSO4 solution (100 g/L) was added to the SPI
soy milk and tofu, information about the effects of different soy oil emulsions to give a final coagulant concentration of 3.5 g/L, and the
volume fractions on the textural properties and microstructures of solutions were stirred gently for 1 min. Next, they were heated at 90 °C
CaSO4-induced SPI gels is scant. Although the effects of soy protein's for 15 min and cooled down to room temperature with iced water to
thermal denaturation on the properties of tofu or tofu-type gels have allow the formation of the gel. The gels were stored at 4 °C for further
been widely reported (Liu, Chang, Li, & Tatsumi, 2004; Shin, analysis.
Yokoyama, Kim, Wicker, & Kim, 2015; Zhao, Li, Qin, & Chen, 2016),
the effects of thermal denaturation of SPI before or after the in- 2.4. Characteristics of SPI emulsion droplets
corporation of different soy oil volume fractions via emulsification on
the gel structure and gelation behavior of CaSO4-induced SPI gels have The droplet size distribution of freshly prepared SPI emulsions was
not been discussed in detail. Therefore, this study aimed to investigate determined using a Microtrac S3500 Particle Size Analyser (Microtrac
the effects of different volume fractions of additional soy oil (within the Inc., Largo, FL, USA) according to the methods of Guo, Hu, Wang, and
scope of soy oil content in traditional tofu) and thermal denaturation of Liu (2018) and Zhuang et al. (2019) with some modifications. The
SPI before and after incorporation of soy oil on the rheological and range of measurement was from 0.01 to 2000 μm. After the preparation
textural properties of CaSO4-induced SPI gels. Based on the earlier of emulsions, each sample was dripped into the sample chamber and
preliminary work (not published), the oil content in the tofu ranges directly diluted with deionized water. Afterwards, the measurement
from 20 g/kg to 50 g/kg. In addition, Wang et al. (2018) and Luo et al. was performed at 25 °C. The droplet size of the SPI emulsions, expressed
(2019), who investigated the effects of pre-aggregation and pre-cross- as the volume-average diameter, was calculated as an average of three
linking of SPI on the CaSO4-induced gelation properties of soy protein freshly prepared SPI emulsion samples. Measurements were taken in
emulsion, chose a 50 mL/L oil content. Therefore, the oil volume fac- duplicate. The refractive indices of soy oil and continuous phase used
tions in this study varied from 10 to 70 mL/L. The dispersion of soy oil were 1.456 and 1.33, respectively.
droplets in the protein gel matrix and microstructures of the gels were
characterized by confocal laser scanning microscopy (CLSM). 2.5. Dynamic rheological measurements

2. Materials and methods An AR-G2 Rheometer (TA Instruments, New Castle, DE, USA) was
used to perform the rheological measurements. The viscosity of the SPI
2.1. Materials emulsions subjected to heat treatment before and after incorporation of
different soy oil volume fractions was measured using a cone plate
Soybeans (Xuelang Seed Station, Wuxi, China) were ground with a geometry (diameter = 40 mm; angle = 2°). Measurements were taken
15B grain mill (Minyi, Yongkang, China). After removal of the hull, the in triplicate at room temperature (25 °C).
flour was defatted using n-hexane/ethanol (10:1, mL:mL) at room G′ and G″ of the SPI emulsions were determined using a parallel
temperature (25 °C). SPI was extracted from the defatted soybean meal plate (diameter = 40 mm). After adding a coagulant, each sample was
following the method of Wang, Liu, Ma, and Zhao (2019) with some immediately loaded into the rheometer, and the excess solution around
modifications. Briefly, SPI was obtained by alkali extraction (pH 8.0) the edge of a plate was removed using tissue paper. The exposed rim of
with 2 mol/L NaOH and acid precipitation (pH 4.5) with 2 mol/L HCl. each sample was covered with silicon oil to prevent evaporation during
The extract was then washed and resuspended with deionized water gelation. The strain applied was 0.01, which was within the linear
and adjusted to pH 7.0 with 2 mol/L NaOH, followed by freeze-drying. viscoelastic region, and the frequency was 1 Hz. The samples were
The protein content was 910 g/kg as measured by Kjeldahl method heated from 25 °C to 90 °C at a heating rate of 5 °C/min, followed by
(N × 6.25). Rhodamine B and Nile Blue were purchased from Sigma- incubation at 90 °C for 10 min and subsequently cooled to 25 °C at a
Aldrich (St. Louis, MO, USA), CaSO4·2H2O from Sinopharm Chemical cooling rate of 5 °C/min. After the temperature cycle, the gels formed
Reagent Co., Ltd. (Shanghai, China), and soy oil from a local super- were subjected to a strain sweep measurement at 25 °C. The strain was
market (Wuxi, China). All other chemicals were of analytical grade. varied from 0.001 to 1, and the frequency was 1 Hz.

2.2. Emulsion preparation 2.6. Texture analysis of CaSO4-induced SPI gels

SPI was dispersed in deionized water to give a protein concentration The gel hardness was measured using a TA-XT2 texture analyser
of 75 g/L, and stirred using a magnetic stirrer for 2 h at room tem- (Stable Micro Systems, Godalming, Surrey, UK) according to the
perature (25 °C). The pH value of the SPI dispersion was 7.0. methods described by Zhao et al. (2016) and (Wang et al., 2019). The
Subsequently, the SPI dispersion was divided into two parts. One SPI gel size in the containers were 41 mm in diameter and 23 mm in height.
dispersion was directly mixed with soy oil using a T18 basic high-speed Before the measurement, the gels were equilibrated at room tempera-
blender (IKA, Staufen, Germany) for 2 min. The added soy oil volume ture (25 °C) for 30 min. Then, the gels were punctured to 30% of the
fractions were 10, 30, 50 and 70 mL/L. The final protein concentration original height using a cylindrical plunger with a diameter of 10 mm at
in all SPI emulsions was adjusted to 70 g/L. Next, the coarse SPI a pre-test speed of 3 mm/s, test speed of 0.8 mm/s and post-test speed
emulsions were homogenized using a homogenizer (ATS Engineering of 5 mm/s. The trigger force was 9.81 × 10−3 N and the option of re-
Inc., Brampton, Canada) at a pressure of 20 MPa, heated at 90 °C for turn to start was selected. The peak value in the force–distance curve
15 min and subsequently cooled down to room temperature. The was referred to as the gel hardness.

2
H. Zhao, et al. LWT - Food Science and Technology 117 (2020) 108690

2.7. Water-holding capacity of CaSO4-induced SPI gels

The water-holding capacity (WHC) measurement of the gels was


conducted according to the method of Maltais, Remondetto, Gonzalez,
and Subirade (2005) with some modifications. Equal volumes (3 mL) of
emulsions was transformed into centrifuge tubes and the gels are al-
lowed to form following the process mentioned in section 2.3. The gels
were stored at 4 °C overnight, and centrifuged at 10,000 g for 10 min at
4 °C. WHC (%) was calculated using the following formula:

(Weight of water remaining in the gels after centrifugation/Weight


of the total water in the samples) × 100

2.8. Confocal scanning laser microscopy analysis

The microstructure of the SPI gels was evaluated using a Leica TCS
SP8 confocal laser scanning microscope (Leica Microsystems,
Mannheim, Germany). The SPI gels were stained with the fluorescent
dyes Rhodamine B and Nile blue, which stain the protein and oil phases,
respectively, at excitation wavelengths of 488 and 633 nm, respectively.
Both dye solutions (0.1 g/L) were mixed well with the SPI emulsions
prior to gel formation. After adding CaSO4, the mixtures were heated at
90 °C for 15 min to allow gel formation. Subsequently, the gels formed
were placed at the bottom of petri dishes and covered with coverslips.
Images were taken at a certain depth below the coverslip surface. The
protein stained with Rhodamine B appeared red, and the soy oil dro-
plets stained with Nile blue were green.

2.9. Statistical analysis

All results were expressed as the mean ± standard deviation of


three independent tests, each with at least triplicate analyses. Data were
subjected to nonparametric test using IBM SPSS Statistical 22.0 soft-
ware (Armonk, New York, USA). Significant differences (P < 0.05)
between individual means within OEH or HOE group were identified by
the Kruskal-Wallis test followed by the Dunn-Bonferroni post hoc test.
The comparisons between OEH and HOE groups at given volume
fractions were performed using Wilcoxon signed-rank test.
Fig. 1. Oil droplet size distribution of OEH emulsions (A) and HOE emulsions
(B) containing different soy oil volume fractions: 0 mL/L (●); 10 mL/L (◯);
3. Results and discussion
30 mL/L (▼); 50 mL/L (△); 70 mL/L (■). OEH emulsions: soy protein emul-
sions prepared following the process of addition of oil, emulsification and heat
3.1. Soy oil droplet size distribution of SPI emulsions treatment; HOE emulsions: soy protein emulsions prepared following the pro-
cess of heat treatment, addition of oil and emulsification.
The droplet size distribution of the freshly formed SPI emulsions
was measured using laser diffraction and the results were shown in
emulsions is larger compared to HOE emulsions.
Fig. 1. Generally, the volume mean diameter (VMD) of all the SPI
emulsions was in the range of 1.60–3.09 μm. As the soy oil volume
fraction gradually increases from 10 to 70 mL/L, the mean droplet
3.2. Viscosity of SPI emulsions
diameter of OEH and HEO emulsions increases correspondingly. This
might be because of the increased hydrophobic interactions between
Fig. 2 illustrates the viscosities of OEH and HOE emulsions at soy oil
individual soy oil droplets (Wong & Kitts, 2003). In addition, at a given
volume fractions of 10, 30, 50 and 70 mL/L. The results showed that the
protein concentration in the continuous phase, a possible coalescence of
viscosity decreased with increasing shear rate for all emulsions, in-
soy oil droplets in the SPI emulsion system would occur more easily at a
dicating a typical shear-thinning (pseudoplastic) behavior (Li et al.,
higher soy oil volume fraction (Yang, Liu, & Tang, 2013).
2011). The apparent increase in viscosity could also be observed at a
The mean droplet diameter of the HOE emulsions ranged from 1.60
specific shear rate with increasing soy oil volume fractions, exhibiting a
to 1.92 μm. In contrast, the mean droplet size of the OEH emulsions was
parallel trend like the droplet size. This phenomenon could be attrib-
slightly higher (1.96–3.09 μm). Thermal denaturation of SPI before the
uted to the interactions between proteins and soy oil droplets promoted
addition of soy oil can cause unfolding of protein molecules and ex-
by the thermo-denatured protein molecules (Taha & Mohamed, 2004).
posure of hydrophobic groups initially located in the interior of SPI
Moreover, the viscosity of HOE emulsions is higher than that of OEH
molecules. The emulsifying properties of denatured proteins would
emulsions at the same soy oil volume fraction and shear rate. Based on
improve accordingly (Dissanayake & Vasiljevic, 2009; Keerati-u-rai &
the results shown in Fig. 1, HOE emulsions with a smaller droplet size at
Corredig, 2009). In contrast, heat treatment of SPI dispersions after
a given soy oil volume fraction (compared to OEH emulsions) comprise
incorporation of soy oil and emulsification usually leads to flocculation
a larger number of droplets. Therefore, the probability that any two soy
of soy oil droplets (Iordache & Jelen, 2003), possibly from interactions
oil droplets enter the attraction region increases, leading to an increase
between adsorbed proteins at a droplet interface or between droplets in
in viscosity of HOE emulsions.
emulsions (Li, Kong, Zhang, & Hua, 2011). Therefore, the VMD of OEH

3
H. Zhao, et al. LWT - Food Science and Technology 117 (2020) 108690

Fig. 3. Storage modulus (G′) and temperature (T) as a function of time for OEH
emulsions (A) and HOE emulsions (B) containing different soy oil volume
fractions (0 mL/L: solid line; 10 mL/L: dotted; 30 mL/L: dash-dot; 50 mL/L:
short dash; 70 mL/L: dash-dot-dot) during CaSO4-induced gelation. OEH
emulsions: soy protein emulsions prepared following the process of addition of
oil, emulsification and heat treatment; HOE emulsions: soy protein emulsions
Fig. 2. Viscosity of OEH emulsions (A) and HOE emulsions (B) containing prepared following the process of heat treatment, addition of oil and emulsi-
different soy oil volume fractions (0 mL/L: solid line; 10 mL/L: dotted; 30 mL/L: fication.
dash-dot; 50 mL/L: short dash; 70 mL/L: dash-dot-dot) as a function of shear
rate. OEH emulsions: soy protein emulsions prepared following the process of viscoelastic properties of CaSO4-induced SPI gels, G′ and G″ were ob-
addition of oil, emulsification and heat treatment; HOE emulsions: soy protein
tained by applying a temperature-time profile during the rheological
emulsions prepared following the process of heat treatment, addition of oil and
measurements. The typical profiles of G′ after the addition of CaSO4 are
emulsification.
presented in Fig. 3. The curves of G″ which are similar, to those of G′ are
not shown. The final G′ and G″ of the SPI gels after the temperature
3.3. Viscoelastic measurement cycles are summarized in Fig. 4. A CaSO4-induced SPI gel free of soy oil
was used as the control.
According to a previous report (Zhao et al., 2016), CaSO4-induced After the addition of a coagulant, all the SPI emulsions underwent
gelation of SPI emulsions is considered a two-step process. Step 1 is the sol to gel transition as the temperature increased. During the heating
thermal denaturation of soy proteins, during which hydrophobic re- period and holding time at 90 °C, the rigidity (elastic modulus) gradu-
gions and –SH groups initially located inside the protein molecules are ally increased for all the SPI gels among which the differences were not
exposed to the outside (Guo et al., 2015). Meanwhile, protein ag- obvious along with the gels made from SPI emulsions with a 70 mL/L
gregates are formed via hydrophobic interactions and disulfide bonds soy oil level. The increase in G′ during the heating period could be
among the SPI molecules. After the addition of a coagulant, negative attributed to the protein–oil and protein–protein physical interactions
charges on the surface of protein molecules are neutralized and elec- combined with the structural rearrangements of the SPI gel network
trostatic repulsion among them is reduced, leading to further ag- (Chen & Dickinson, 2000). During the cooling period, a sharp increase
gregation with Ca2+ acting as a bridge between the charged carboxylic in the elasticity of all the SPI gels is observed, and the differences in G′
groups on adjacent SPI molecules (Kao, Su, & Lee, 2003). In step 2, the resulting from the incorporation of different soy oil volume fractions
SPI aggregates interact more strongly with one another through pro- were distinguishable on the curves. The dramatic escalation of this ri-
tein–protein interactions, eventually promoting the formation of a gidity during the cooling stage indicated that the gel structures could be
three-dimensional gel network. Because of the addition of soy oil, the further enhanced by the formation of stronger bonds. Typically, hy-
potential interactions could be more complicated in the SPI emulsions. drogen bond formation is generally considered more likely at lower
To investigate the effects of thermal denaturation of SPI before and/ temperatures. Fig. 4 indicates that the viscoelastic properties of the SPI
or after incorporation of soy oil with different Φ values on the gels improved after the addition of soy oil. For CaSO4-induced OEH

4
H. Zhao, et al. LWT - Food Science and Technology 117 (2020) 108690

Fig. 5. Strain sweep of storage modulus (G′) of CaSO4-induced gels made from
Fig. 4. Final storage modulus (G′) (A) and loss modulus (G″) (B) of CaSO4- OEH emulsions (A) and HOE emulsions (B) with different soy oil volume
induced soy protein gels made from OEH emulsions (□) and HOE emulsions fractions: 0 mL/L (●); 10 mL/L (◯); 30 mL/L (▼); 50 mL/L (△); 70 mL/L (■).
(■) as a function of oil volume fraction. Means with different letters sig- The measurement was performed at room temperature (25 °C). OEH emulsions:
nificantly differ (P < 0.05). The means between OEH and HOE groups (not soy protein emulsions prepared following the process of addition of oil, emul-
shown) differ significantly (P < 0.05). OEH emulsions: soy protein emulsions sification and heat treatment; HOE emulsions: soy protein emulsions prepared
prepared following the process of addition of oil, emulsification and heat following the process of heat treatment, addition of oil and emulsification.
treatment; HOE emulsions: soy protein emulsions prepared following the pro-
cess of heat treatment, addition of oil and emulsification.
& Subirade, 2008). The strain value at which G′ decreased to 95% of its
maximum value was considered the critical strain of the gel (Bi, Li,
emulsion gels containing 10–70 mL/L of soy oil, the increase in the final Wang, & Adhikari, 2013). The results demonstrated all SPI gels exhibit
G′ ranged from 20.4% to 42.0% compared to the control. In contrast, similar behavior (Fig. 5). Furthermore, at the critical strain, G′ showed
the final G′ of HOE emulsion gels containing soy oil in the same range of an upward trend with increasing soy oil volume fraction, and the ri-
volume fraction increased by 50.1%–67.7% compared to the control. gidity of the HOE emulsion gels was stiffer compared to OEH emulsion
The reinforced protein–oil interactions with increasing soy oil volume gels, in good agreement with the results shown in Figs. 3 and 4.
fractions led to a higher degree of cross-linking in the gel matrix. As
depicted in Fig. 1, the droplet size of the HOE emulsions was less than
that of the OEH emulsions at the same volume fraction of soy oil. 3.4. Hardness and WHC of CaSO4-induced SPI gels
Therefore, the contact area between protein molecules and soy oil
droplets in HOE emulsions tended to be more predominant, and the Fig. 6 show the hardness and WHC of CaSO4-induced SPI gels made
resultant interactions corresponding to the gel network were stronger. from OEH and HOE emulsions containing different soy oil volume
In this study, a strain sweep test was performed after the formation fractions. The results indicated that incorporation of soy oil at a volume
of the SPI gels, and the results are shown in Fig. 5. For all SPI gels, a fraction of 10–70 mL/L increases both the hardness and WHC of the SPI
linear response at strain less than 10% was observed wherein G′ was gels to different degrees. In addition, the SPI gels made from HOE
independent of strain. This implies that gelation measurements con- emulsions have significantly greater (P < 0.05) hardness and WHC
ducted at 1% strain fall within the linear viscoelastic region. When the compared to OEH emulsion gels at the same soy oil volume fraction.
strain was increased further, G′ decreased sharply, indicating a The results of hardness and WHC of CaSO4-induced SPI gels are
breakage of bonds in the gel network and transition from linear to non- consistent with the results of G′ discussed earlier. The increase in
linear behavior (Eleya, Ko, & Gunasekaran, 2004; Maltais, Remondetto, hardness may arise from the formation of a greater number of cross-
links in the system of emulsion gel at higher oil concentrations.

5
H. Zhao, et al. LWT - Food Science and Technology 117 (2020) 108690

homogeneous microstructures with smaller pores than OEH gels, as will


be discussed later, causing an increase of the capillary forces and hence
resulting in gel matrix retained more water (Xi et al., 2019).

3.5. CLSM analysis of CaSO4-induced SPI gels

Microstructures of CaSO4-induced SPI gels made from OEH and/or


HEO emulsions containing different soy oil volume fractions were
characterized by CLSM. Fig. 7 reveals that the network structure is
somewhat loose and heterogeneous for the gel prepared by SPI dis-
persion without soy oil (control). The addition of soy oil before thermal
denaturation of soy protein improves the structure of CaSO4-induced
SPI gels to some extent after gelation. As shown in Fig. 7B–E, the SPI
gels formed exhibit a slightly dense structure because of the enrichment
of soy oil droplets, which could act as active fillers (Xi et al., 2019).
Moreover, larger soy oil droplets coated with protein molecules on the
surface are more likely to be observed on the images of former gels
(Fig. 7B–E) with increasing soy oil volume fractions, further demon-
strating the occurrence of flocculation within the gel matrix
(Murekatete et al., 2016; Xi et al., 2019). Fig. 7F–I presents the network
structure of SPI gels made from HOE emulsions, featuring a more
homogeneous and compact structure than that of SPI gels made from
OEH emulsions. In addition, the distribution of soy oil droplets within
the SPI gel structure resulting from HOE emulsions appears more uni-
form than that of the soy oil droplets within the SPI gels structure re-
sulting from OEH emulsions, and the visible droplet sizes are consistent
with the results of the VMD described in Fig. 1. The system could be
regarded as a network of particles formed by chemical and physical
bonds. The oil droplets will not only act as filler particles but will be
also component of the gel network (Xi et al., 2019). The enhanced
emulsifying properties resulting from the thermal denaturation of SPI
before incorporation of oil could strengthen the interaction force be-
tween emulsion droplets and between the emulsion droplets and the
protein matrix and explain the improvement of the gel microstructures.
In brief, the microstructure results further explain the differences of the
WHC and the viscoelastic properties among various gels.

Fig. 6. Hardness (A) and water holding capacity (B) of CaSO4-induced soy 4. Conclusions
protein gels made from OEH emulsions (□) and HOE emulsions (■) as a
function of oil volume fraction. Means with different letters significantly differ The rigidity (final G′), hardness and WHC of the gels were all gra-
(P < 0.05). The means between OEH and HOE groups (not shown) differ sig- dually enhanced with the increase of soy oil from 10 mL/L to 70 mL/L.
nificantly (P < 0.05). OEH emulsions: soy protein emulsions prepared fol- The HOE emulsion gel displayed greater performance of gelation be-
lowing the process of addition of oil, emulsification and heat treatment; HOE havior and textural properties and exhibited more compact and
emulsions: soy protein emulsions prepared following the process of heat homogeneous structures compared to OEH emulsion gels at the same
treatment, addition of oil and emulsification. volume fraction of soy oil, possibly resulting from the reinforced pro-
tein-oil interactions and active filler effect of smaller oil droplets within
Furthermore, the viscous properties of the interfacial layers in the the network structures. The results of this study could potentially fa-
system could dissipate energy. Hence, the emulsion gels require more cilitate the development of soy protein-based tofu-type gels with a
force to break at the higher oil volume fractions (Xi, Liu, McClements, & firmer texture and finer microstructure, thereby providing new alter-
Zou, 2019). In addition, because of the role that soy oil plays in en- natives for the making of tofu. However, detailed study is still required
hancing SPI elasticity with increasing volume fraction, soy oil droplets to further understand the structural alteration of SPI aggregate under
are normally recognized as active fillers (Dickinson, 2001; Xi et al., various thermal treatment conditions on the characteristics of SPI
2019). Soy oil interacts with the gel matrix, with stronger interactions emulsion and CaSO4-induced SPI gels.
at higher volume fractions, resulting in reinforcement of the capacity of
the gel to resist compression and puncture. The relatively lower mean Declaration of competing interest
droplet size of HOE emulsions (compared to OEH emulsions at the same
soy oil volume fractions) could improve the filling effect of the active None.
particles and reduce the interstitial spaces of particles within the gel
matrix. Considering an earlier work (Zhao et al., 2016), the denatura-
Acknowledgement
tion of soy protein could improve the compactness of gel network due
to the enhanced protein-protein interactions. In the present study, the
The authors would like to thank the Special Funds for Taishan
larger value of WHC corresponding to higher oil volume fractions (more
Scholars Project, the Innovation Team of Jinan City (No.
oil droplets) could be caused by the enhanced density of gel structure,
2018GXRC004), the Nature Science Foundation of China (No.
leading to the water to be trapped within the interior of the pores by
31901645) and the Natural Science Foundation of Shandong Province
capillary forces. Moreover, the HOE emulsion gels exhibit more
(ZR2019BCE088) for their financial support.

6
H. Zhao, et al. LWT - Food Science and Technology 117 (2020) 108690

Fig. 7. CLSM images of CaSO4-induced gels made from emulsions with different soy oil volume fractions. (A) Gel made without soy oil. (B–E) SPI gels made from
OEH emulsions containing 10 mL/L, 30 mL/L, 50 mL/L and 70 mL/L soy oil, respectively. (F–I) SPI gels made from HOE emulsions containing 10 mL/L, 30 mL/L,
50 mL/L and 70 mL/L soy oil, respectively. OEH emulsions: soy protein emulsions prepared following the process of addition of oil, emulsification and heat
treatment; HOE emulsions: soy protein emulsions prepared following the process of heat treatment, addition of oil and emulsification.

References Maltais, A., Remondetto, G. E., & Subirade, M. (2008). Mechanisms involved in the for-
mation and structure of soya protein cold-set gels: A molecular and supramolecular
investigation. Food Hydrocolloids, 22(4), 550–559.
Bi, C. H., Li, D., Wang, L. J., & Adhikari, B. (2013). Viscoelastic properties and fractal Murekatete, N., Zhang, C., Hategekimana, J., Karangwa, E., & Hua, Y. (2016). Soybean oil
analysis of acid-induced SPI gels at different ionic strength. Carbohydrate Polymers, volume fraction effects on the rheology characteristics and gelation behavior of
92(1), 98–105. glucono‐δ‐lactone and calcium sulfate‐induced tofu gels. Journal of Texture Studies,
Chang, K. C. (2006). Chemistry and technology of tofu making. In Y. H. Hui (Ed.). 47(2), 112–130.
Handbook of food science, technology, and engineering (pp. 1–24). Boca Raton, FL: CRC Nishinari, K., Fang, Y., Guo, S., & Phillips, G. O. (2014). Soy proteins: A review on
Press. composition, aggregation and emulsification. Food Hydrocolloids, 39, 301–318.
Chen, J. S., & Dickinson, E. (2000). On the temperature reversibility of the viscoelasticity Obatolu, V. A. (2008). Effect of different coagulants on yield and quality of tofu from
of acid-induced sodium caseinate emulsion gels. International Dairy Journal, 10(8), soymilk. European Food Research and Technology, 226(3), 467–472.
541–549. Rekha, C. R., & Vijayalakshmi, G. (2010). Influence of natural coagulants on isoflavones
Corzo-Martinez, M., Garcia-Campos, G., Montilla, A., & Moreno, F. J. (2016). Tofu whey and antioxidant activity of tofu. Journal of Food Science and Technology-Mysore, 47(4),
permeate is an efficient source to enzymatically produce prebiotic fructooligo- 387–393.
saccharides and novel fructosylated alpha-galactosides. Journal of Agricultural and Ringgenberg, E., Alexander, M., & Corredig, M. (2013). Effect of concentration and in-
Food Chemistry, 64(21), 4346–4352. cubation temperature on the acid induced aggregation of soymilk. Food Hydrocolloids,
Dickinson, E. (2001). Milk protein interfacial layers and the relationship to emulsion 30(1), 463–469.
stability and rheology. Colloids and Surfaces B: Biointerfaces, 20(3), 197–210. Shin, W. K., Yokoyama, W. H., Kim, W., Wicker, L., & Kim, Y. (2015). Change in texture
Dissanayake, M., & Vasiljevic, T. (2009). Functional properties of whey proteins affected improvement of low-fat tofu by means of low-fat soymilk protein denaturation.
by heat treatment and hydrodynamic high-pressure shearing. Journal of Dairy Science, Journal of the Science of Food and Agriculture, 95(5), 1000–1007.
92(4), 1387–1397. Taha, F. S., & Mohamed, S. S. (2004). Effect of different denaturating methods on lipid-
Eleya, M. M. O., Ko, S., & Gunasekaran, S. (2004). Scaling and fractal analysis of vis- protein complex formation. Lebensmittel-Wissenschaft Und-Technologie-Food Science
coelastic properties of heat-induced protein gels. Food Hydrocolloids, 18(2), 315–323. and Technology, 37(1), 99–104.
Fukushima, D. (1991). Recent progress of soybean protein foods: Chemistry, technology, Wang, Y., Liu, C., Ma, T., & Zhao, J. (2019). Physicochemical and functional properties of
and nutrition. Food Research International, 7(3), 323–351. γ-aminobutyric acid-treated soy proteins. Food Chemistry, 295, 267–273.
Guo, Y. L., Hu, H., Wang, Q., & Liu, H. Z. (2018). A novel process for peanut tofu gel: Its Tang, C. H., Chen, L., & Foegeding, E. A. (2011). Mechanical and water-holding properties
texture, microstructure and protein behavioral changes affected by processing con- and microstructures of soy protein isolate emulsion gels induced by CaCl2, glucono-
ditions. Lwt-Food Science and Technology, 96, 140–146. delta-lactone (GDL), and transglutaminase: Influence of thermal treatments before
Guo, S. T., Ono, T., & Mikami, M. (1999). Incorporation of soy milk lipid into protein and/or after emulsification. Journal of Agricultural and Food Chemistry, 59(8),
coagulum by addition of calcium chloride. Journal of Agricultural and Food Chemistry, 4071–4077.
47(3), 901–905. Wang, X. F., Zeng, M. M., Qin, F., Adhikari, B., He, Z. Y., & Chen, J. (2018). Enhanced
Guo, F. X., Xiong, Y. L. L., Qin, F., Jian, H. J., Huang, X. L., & Chen, J. (2015). Surface CaSO4-induced gelation properties of soy protein isolate emulsion by pre-aggrega-
properties of heat-induced soluble soy protein aggregates of different molecular tion. Food Chemistry, 242, 459–465.
masses. Journal of Food Science, 80(2), C279–C287. Wang, Y. S., Zhao, J., Liu, C. Q., & Li, W. W. (2019). Influence of gamma-aminobutyric
Iordache, M., & Jelen, P. (2003). High pressure microfluidization treatment of heat de- acid on gelling properties of heat-induced whey protein gels. Food Hydrocolloids, 94,
natured whey proteins for improved functionality. Innovative Food Science & Emerging 287–293.
Technologies, 4(4), 367–376. Wong, P. Y. Y., & Kitts, D. D. (2003). A comparison of the butter milk solids functional
Kao, F. J., Su, N. W., & Lee, M. H. (2003). Effect of calcium sulfate concentration in properties to nonfat dried milk, soy protein isolate, dried egg white, and egg yolk
soymilk on the microstructure of firm tofu and the protein constitutions in tofu whey. powders. Journal of Dairy Science, 86(3), 746–754.
Journal of Agricultural and Food Chemistry, 51(21), 6211–6216. Xi, Z., Liu, W., McClements, D. J., & Zou, L. (2019). Rheological, structural, and micro-
Keerati-u-rai, M., & Corredig, M. (2009). Heat-induced changes in oil-in-water emulsions structural properties of ethanol induced cold-set whey protein emulsion gels: Effect of
stabilized with soy protein isolate. Food Hydrocolloids, 23(8), 2141–2148. oil content. Food Chemistry, 291, 22–29.
Kim, K. H., Renkema, J. M. S., & van Vliet, T. (2001). Rheological properties of soybean Yang, M., Liu, F., & Tang, C. H. (2013). Properties and microstructure of transglutami-
protein isolate gels containing emulsion droplets. Food Hydrocolloids, 15(3), 295–302. nase-set soy protein-stabilized emulsion gels. Food Research International, 52(1),
Li, F., Kong, X. Z., Zhang, C. M., & Hua, Y. F. (2011). Effect of heat treatment on the 409–418.
properties of soy protein-stabilised emulsions. International Journal of Food Science Zhao, H. B., Li, W. W., Qin, F., & Chen, J. (2016). Calcium sulphate-induced soya bean
and Technology, 46(8), 1554–1560. protein tofu-type gels: Influence of denaturation and particle size. International
Liu, Z. S., Chang, S. K. C., Li, L. T., & Tatsumi, E. (2004). Effect of selective thermal Journal of Food Science and Technology, 51(3), 731–741.
denaturation of soybean proteins on soymilk viscosity and tofu's physical properties. Zhao, H. B., Wang, Y. S., Li, W. W., Qin, F., & Chen, J. (2017). Effects of oligosaccharides
Food Research International, 37(8), 815–822. and soy soluble polysaccharide on the rheological and textural properties of calcium
Luo, K., Liu, S., Miao, S., Adhikari, B., Wang, X., & Chen, J. (2019). Effects of transglu- sulfate-induced soy protein gels. Food and Bioprocess Technology, 10(3), 556–567.
taminase pre-crosslinking on salt-induced gelation of soy protein isolate emulsion. Zhuang, X. B., Jiang, X. P., Zhou, H. Y., Han, M. Y., Liu, Y. F., Bai, Y., ... Zhou, G. H.
Journal of Food Engineering, 263, 280–287. (2019). The effect of insoluble dietary fiber on myofibrillar protein emulsion gels: Oil
Maltais, A., Remondetto, G. E., Gonzalez, R., & Subirade, M. (2005). Formation of soy particle size and protein network microstructure. Lwt-Food Science and Technology,
protein isolate cold-set gels: Protein and salt effects. Journal of Food Science, 70(1), 101, 534–542.
C67–C73.

You might also like