You are on page 1of 9

Carbohydrate Polymers 266 (2021) 118148

Contents lists available at ScienceDirect

Carbohydrate Polymers
journal homepage: www.elsevier.com/locate/carbpol

Comparison of binary cress seed mucilage (CSM)/β-lactoglobulin (BLG) and


ternary CSG-BLG-Ca (calcium) complexes as emulsifiers: Interfacial
behavior and freeze-thawing stability
Afsaneh Taheri a, Mahdi Kashaninejad a, *, Ali Mohammad Tamaddon b, Seid Mahdi Jafari a
a
Department of Food Process Engineering, Faculty of Food Science and Technology, Gorgan University of Agricultural Sciences and Natural Resources, Gorgan, Iran
b
Department of Pharmaceutical Nanotechnology and Center for Nanotechnology in Drug Delivery, Shiraz University of Medical Sciences, Shiraz, Iran

A R T I C L E I N F O A B S T R A C T

Keywords: Protein–polysaccharide complexes often exhibit amended techno-functional characteristics when compared to
β-Lactoglobulin their individual participant biomolecules. In this study, a complex coacervation of cress seed mucilage (CSM)/
Complex Coacervation β-lactoglobulin (Blg) was used for stabilizing oil-in-water emulsions; they were characterized in terms of physical
Cress seed
properties, droplet-size distribution and microstructure. Also, a comprehensive study was carried out on inter­
Emulsion
Freeze-thawing
facial rheological responses and on the corresponding emulsion stability of different complexes. Freeze-thaw
stability of the produced emulsions which had from mixtures of CSM-Blg was also evaluated. More than the
size of droplets, interfacial rheological characteristics were associated with the properties of the adsorbed layers
and with the stability of emulsions in storage. Using the CSM-Blg-Ca ultimately resulted in emulsions that proved
stable against creaming, with no sign of phase separation over 3 weeks. These results show pro­
tein–polysaccharide complexes as appropriate emulsifiers that can make emulsion-based products resistant to
unwanted changes caused by freeze-thawing.

1. Introduction (Taheri & Jafari, 2019a). Environmental stresses can include thermal
treatments, freezing, drying and mechanical agitation. The degree of
Interactions between proteins and polysaccharides through covalent stability that is imparted to emulsions, when enhanced with pro­
bonding or electrostatic interactions usually play significant roles in tein–polysaccharide complexes, usually exceeds the stability of emul­
improving the structure and stability of emulsions, thereby highlighting sions in which proteins are used alone (i.e. as a single interfacial layer).
their importance in emulsion-based food industries, personal care For instance, the stability of oil-in-water emulsions can be increased by
products and pharmaceutical applications (McClements, 2012; Mehrnia pectin-whey protein (Esfanjani et al., 2015; Faridi Esfanjani et al., 2017;
et al., 2016). There are often two different approaches in creating Gharehbeglou et al., 2019; Mohammadi et al., 2016), chitin nano­
emulsions with protein-polysaccharide complexes. The first approach crystals-β-lactoglobulin (Gülseren & Corredig, 2013), sodium caseinate-
involves adding an oil phase to an aqueous solution containing pro­ gellan gum mixtures (Sosa-Herrera et al., 2008), maltodextrin-whey
tein–polysaccharide complexes, followed by mixing two phases (i.e. protein (Assadpour et al., 2016a, 2016b), lactoferrin-ι-carrageenan
homogenization or sonication). The second approach involves creating (Shimoni et al., 2013) and β-lactoglobulin-chitosan (Chang et al., 2018).
an initial emulsion first, one that is stabilized by proteins, and then By common knowledge, some products and ingredients can resist
adding polysaccharides which adsorb onto the protein layer and orga­ deterioration after they are repeatedly frozen and thawed (Iwanaga
nize a bilayer/multilayer that can increase the bridging flocculation et al., 2008). Termed freeze-thaw stability, this feature is technically
(Bago Rodriguez et al., 2018; Jafari et al., 2007; Taheri & Jafari, 2019b). important when managing emulsions for food formulation. It can be
Generally, under appropriate conditions of formulation by the two ap­ useful for extending the shelf life of products (e.g. sauces for frozen
proaches, emulsions can become more resistant to environmental meals) and for imparting better quality (e.g. ice cream). Nevertheless,
stresses as a result of adding protein–polysaccharide complexes with carrying out freeze-thaw treatments on emulsion-based products can
multilayered interfacial membranes at the emulsion droplet surface sometimes serve as an environmental stress and, thus, lead to adverse

* Corresponding author.
E-mail address: kashani@gau.ac.ir (M. Kashaninejad).

https://doi.org/10.1016/j.carbpol.2021.118148
Received 2 April 2021; Received in revised form 21 April 2021; Accepted 29 April 2021
Available online 5 May 2021
0144-8617/© 2021 Elsevier Ltd. All rights reserved.
A. Taheri et al. Carbohydrate Polymers 266 (2021) 118148

effects such as oiling off, coalescence or partial coalescence, flocculation (HCl) and sodium hydroxide (NaOH) were purchased from Sigma-
and/or creaming (Degner et al., 2014). Upon freeze-thaw treatments, Aldrich (Germany). All solutions were obtained by using deionized
the physical stability of oil-in-water emulsions (O/W) can be improved water from a Nanopure water system (Nanopure Infinity, USA).
by cryoprotectants by controlling ice crystal growth, or with interactions
between proteins and polysaccharides. This seems to be one of the most
2.2. Cress seed mucilage extraction and purification
effective approaches for obtaining emulsion-based foods that can be
resistant to freeze-thaw treatments (Degner et al., 2014; Ghosh &
Cress seed mucilage was extracted and purified according to methods
Coupland, 2008). Freeze-thaw stability can be improved by protein-
reported by Taheri et al. (2020). Briefly the water seed ratio of 30:1, pH
polysaccharide complexes, and a few examples include relevant en­
10 and soaking period (15 min) were considered as optimized conditions
hancements by steric stabilization in soybean-microcrystalline cellulose
of extraction. The slurry was extracted as mucilage and was purified by
(Xu et al., 2015), interfacial films with high viscoelasticity in whey
ethanol precipitation. Mucilage was stirred for 10 min with ethanol at a
protein isolate-soybean polysaccharides (Cabezas et al., 2019) and
ratio of 1:3 (mucilage volume: ethanol volume). After stirring, the
ovalbumin-sodium alginate (Zhang et al., 2020).
mixture was centrifuged at 15,000 rpm for 3 min at 25 ◦ C. Freeze-drying
In this work, two biopolymers have been selected to stabilize oil-in-
(LD Plus Alpha, Germany) was applied to dehydrate the mucilage.
water emulsions. They are β-lactoglobulin (Blg), a dominant component
of whey protein, and cress seed mucilage (CSM) from garden cress seed.
Blg is a compact, small globular protein, containing 162 amino acids 2.3. Preparation of biopolymer nanocomplexes
with molecular weights of 18.4 kDa (Ali et al., 2016). There is a struc­
tural reassembly of the protein that can partially unfold and expose The stock solutions of CSM and Blg were prepared by dissolving a
hydrophobic residues to the lipophilic phase, thereby making it an proper amount of CSM and Blg in deionized water to reach the final
appropriate ingredient in the stabilization of emulsions (Bos & Van concentration of 30 mg/mL and 20 mg/mL, respectively. CSM-Blg
Vliet, 2001; Damodaran, 2005). CSM is a hybrid polyelectrolyte that complexes were synthesized by the post-blending acidification
contains both protein and carbohydrate polymers, with a molecular method; after dilution, the solution of CSM was added to Blg dropwise.
weight of about 540 kDa subunits which can vary depending on its The final mass ratio (Blg:CSM), total polymer concentration, and pH
origin, age or purification method. In the pH value above 4.0, CSM were adjusted to 1:2, 25 mg/mL, and 3.5, respectively. The selection
carries a net negative charge which can be attributed to its D-galactur­ procedure of the variables was based on our previous study (Taheri
onic acid and D-glucuronic acid residues. Protein composition and et al., 2020). For the preparation of CSM-Blg-Ca complex, CaCl2 was
rheological modification have been considered as reasons for the added to a group of samples after mixing the solutions, and then a final
emulsifying properties of CSM (Taheri & Razavi, 2015a, 2015b). The concentration of 10 mM was reached in each solution. The CSM-Blg and
combination of these two biopolymers can potentially induce a syner­ CSM-Blg-Ca mixtures were then slowly acidified in using by HCl and
gistic effect on emulsion stability which can be examined, although this NaOH so as to reach a final mixture with pH 3.5. The dispersions were
work pursues other aims in addition to the evaluation of emulsion stirred for 4 h and stored for 24 h at ambient temperature prior to freeze-
stability. drying.
In a recent work, researchers optimized the methods of creating
complex coacervation between the cress polyelectrolyte molecule (CSM) 2.4. Preparation of O/W emulsions
and β-lactoglobulin (Blg) in nanoscale ranges. In aqueous solutions, Blg
proteins interacted with CSM to form either soluble or insoluble com­ A two-step homogenization process was applied to produce the
plexes depending on the experimental variables such as pH and con­ emulsions at ambient temperature. Firstly, coarse emulsions were made
centrations of Blg and CSM. The mixture consisted of a dilute solution of by adding corn oil (20% w/w) to aqueous dispersions of CSM-Blg and
Blg (0.1% w/v)-CSM (0.15% w/v) at pH 3.5, where 10 mM calcium ions CSM-Blg-Ca (2.0% w/v). Meanwhile, the solution (20 mL) was stirred at
were either added or not, which eventually led to optimum samples 15,000 rpm for 10 min using a high-speed homogenizer (IKA, T 25,
(Taheri et al., 2020). In this work, we extend the idea that optimized Germany). In the next step, the pre-emulsions were transferred to a
versions of polyelectrolyte nano-complexes, based on CSM and Blg, are beaker and sonicated in an ultrasound homogenizer (Sonoplus HD-4200,
capable of stabilizing corn oil in aqueous emulsions. In this study, Germany) at 50% amplitude, with 1/3 of the standard tip length
emulsions were stabilized by CSM-Blg and CSM-Blg-Ca complexes which immersed in a glass beaker (28 mm diameter) for 3 min. The tempera­
were made of aqueous polymer mixtures (pH = 3.5). The present ture increase during sonication was hindered by putting the beaker in a
research involved measuring several emulsion properties (i.e. size dis­ water-ice bath.
tribution, morphology and emulsion stability) and then the impact of
freeze-thawing was evaluated on stabilized o/w emulsions that con­
tained the nano-complexes. Since the interfacial rheological evolution at 2.5. Freeze–thaw cycling
an oil/complex interface authorizes a better understanding of the
interfacial adsorption and the stabilizing mechanism of these emulsi­ The freeze–thaw treatment was carried out on emulsions. The effects
fiers, diverse interfacial rheological tests were necessary for this study. of the freezing step were evaluated first. A total volume of 5 mL of each
These tests comprising comprised measurements of the dynamic shear at emulsion type was frozen overnight using a freezer (− 18 ◦ C) and 15 h at
various deformation amplitudes and frequencies. The findings can lead (− 30 ◦ C). Thereafter, the emulsions (previously frozen at − 18 ◦ C and
to basic information on the design of efficient nano-delivery systems in − 30 ◦ C) were thawed during 10 h at room temperature.
future applications. New food-grade stabilizers can be introduced based
on polyelectrolytes Blg-CSM mixtures and they can be incorporated into 2.6. Emulsions droplet size and zeta-potential analysis
a wide range of frozen food emulsions.
The droplet size, size distribution and zeta-potential (ζ) of oil drop­
2. Materials and methods lets in each emulsion were measured at 25 ◦ C by a laser-scattering
particle-size analyzer (Horiba SZ-100 analyzer, Kyoto, Japan). All
2.1. Materials emulsions were diluted with deionized water (1:100) before analysis.
The droplet size was as the mean diameter of volume-surface. To specify
Cress seed were purchased from a local medicinal plant market. Blg the distribution width of droplet sizes (based on volume), the ‘span’ had
(purity 93%), CaCl2 (anhydrous), ethanol (99.5%), hydrochloric acid to be computed first from the following formula:

2
A. Taheri et al. Carbohydrate Polymers 266 (2021) 118148

DV,0.9 − DV,0.1 2.11. Statistical analysis


Span = (1)
DV,0.5
All experiments had triplicate samples for analysis. The results were
where Dv,0.5, Dv,0.1 and Dv,0.9 are values of the particle sizes, compared reported as mean values and standard deviation (mean ± SD). Differ­
to which 50%, 10% and 90% of the sample particles are smaller, ences between the mean values were determined using the analysis of
respectively. variance (ANOVA) via SPSS 16.0 (SPSS Inc., Chicago, IL, USA).

2.7. Storage and creaming stability 3. Results and discussion

The storage stability of each emulsion was reported by testing the The mean droplet size was applied to measure the size of emulsions
droplet size of each emulsion against different time periods (1, 8, 20, 30 quantitatively, considering the fact that samples were nanometer-sized.
days). Four milliliters of emulsion were placed in a glass test tube (with As indicated in Fig. 1, the presence of calcium ion in the oil-water
an internal diameter of 15 mm and a height of 100 mm). These samples interface influenced the droplet size of emulsion. The droplet size of
were stored at 4 ◦ C and 25 ◦ C, and the movement of any creaming emulsion increased by calcium ion (from 164 nm to 326 nm). The
boundary was measured weekly (up to 4 weeks) after the emulsion was emulsion droplets were stabilized by CSM-Blg and they created uniform
produced. The amount of creaming was calculated by using the spheres (Fig. 1a). On the other hand, with a CSM-Blg complex, the
creaming index, which was calculated according to Eq. (2). Where Ht surface of the corn oil droplets enabled the adsorption of more nano­
and Hs are the total height of each emulsion and serum layer height (Hs), particles. This was followed by a gradual dispersion of CSM-Blg within
respectively. the emulsions, along with the production of smaller droplets (Esfanjani
et al., 2015; Linke & Drusch, 2018). The optical images of microscopy
Ht
Cremaimg index = × 100 (2) demonstrated that droplet sizes grow when calcium ions are added to
HS the emulsion. In fact, big drops were observed in the images of emul­
sions where CSM-Blg-Ca was used, as compared to the use of CSM-Blg.
2.8. Optical microscopy measurements On the other hand, using CSM-Blg made the emulsions fluid, with
separate, small and almost monodispersed drops. By CSM-Blg-Ca, the
Optical microscopic observations of the emulsions were performed emulsions harbored big and slightly flocculated drops (Fig. 1b).
using a Raman microscopic microscope (Horiba, Japan) to determine Storage stability is a significant factor of any products because it
the droplet size and the aggregate state of droplets. Samples were placed specifies whether a formulation is appropriate for its planned use. The
on a microscope slide, covered with glass cover slips and observed using creaming, phase separation and alterations in size distribution can be
objective lenses of 5× and 10×. considered as indices of storage stability. Creaming is an initial indicator
of storage stability because, parallel to creaming, coalescence and floc­
2.9. Interfacial rheology culation occur microscopically. The creaming index of CSM-Blg and
CSM-Blg-Ca emulsions were evaluated at 4 ◦ C and 25 ◦ C (Fig. 2). After
Interfacial viscoelasticity measurements were carried out by a monitoring the samples for 14 days, CSM-Blg emulsions exhibited an
rheometer equipped with a Dunoy ring (MCR 302, Anton Paar, Austria). increase in the creaming index at 25 ◦ C (Fig. 2a) which enhanced over
Regarding the liquid/liquid experiments, each nano-complex solution time in storage (from 2.9% to 14.0%). Nonetheless, the creaming index
had a total biopolymer concentration of 0.25% w/v. The solution was increased by a lesser extent at 4 ◦ C, as compared to the increase at 25 ◦ C
poured into a trough, while the ring was carefully immersed in the (from 2.3% to 7.0%) (Fig. 2a). One can conclude that a better stability is
sample to half depth. Corn oil was poured gently into the top of the created by CSM-Blg when emulsions are at 4 ◦ C. Herein, there is a good
solution by a syringe. This was to prevent the mixing of the two phases. agreement between the colloid characteristics of CSM-Blg aqueous
The interfacial rheological results were diverse and were indicated as dispersion. These characteristics include surface charges (Blg = 21 ±
time-dependent evolution of the moduli (elastic (G′ ) and viscous 0.2 mV & CSM = − 68 ± 1.2 mV) and relevant emulsion properties,
modulus (G′′ )), dynamic shear at different deformation amplitudes and while suggesting that the particles adopt various adsorption modes that
frequencies, along with adjusted shear deformation oscillatory runs in rely on their surface charges. In other words, surface charges of CSM-Blg
which the strain amplitude and angular frequency of the oscillations are (− 54 ± 0.4 mV) and CSM-Blg-Ca (− 42 ± 1.1 mV) at pH ~3.0 could be
kept constant at 1% and 1 rad/s, respectively. seen in relation to instability. According to Danov et al. (2006), highly
charged particles are unlikely to be precursors of stability in emulsions
due to a repulsive energy that a charged particle usually acquires in
2.10. Interfacial tension measurements becoming contiguous to the oil-water interface. For highly charged
particles, this force of repulsion could outweigh convective powers
The interfacial tension for each Blg-CSM mixture (0.2% w/v) and the during emulsification. Therefore, charged particles are likely to be
Blg control was measured using a Lauda TD2 Tensiometer (Lauda- adsorbed on the interface of the oil-water blend, thereby making
Königshofen, Germany) equipped with a Du Nüoy ring (20 mm diam­ emulsions unstable. Similar results were reported by De Folter et al.
eter). Within a 40 mm diameter glass sample cup, a 20 mL biopolymer (2012), indicating unstable emulsions that were stabilized by nano­
solution was added, followed by the immersion of the Du Noüy ring and particles with high zeta potentials.
then the addition of an upper corn oil layer (20 mL). The ring was then In contrast, no creaming was observed in emulsions that had been
pulled upward to stretch the interface to measure the maximum force treated with calcium ions at 4 ◦ C for 14 days. After 21 days, a slight
without breaking into the oil phase. Interfacial tension was calculated degree of creaming was observed (5.0%) and, thereafter, the creaming
from the maximum force (Fmax; units: milli-Newton; instrument mea­ index tended to stabilize and no more creaming occurred (Fig. 2b). The
sures force as mg × gravity) using the following equation (Eq. (1)): emulsion stability showed that CSM-Blg-Ca were more capable of hin­
Fmax dering the creaming of emulsions in comparison with the capability of
γ= CSM-Blg at 4 ◦ C and 25 ◦ C. This result could provide a background for
4πRβ
the viscosity enhancements in the continuous phase, thereby making
where, γ is the interfacial tension (mN/m), R is the radius of the ring (20 dispersed droplets move slowly. This result is in agreement with previ­
mm), and β is a correction factor that depends on the dimensions of the ous research by Fan et al. (2017), where adding calcium ions (9 mM) to
ring and the density of the liquid involved (Lam & Nickerson, 2014). low-methoxylated citrus (LMP) solutions ultimately enhanced the

3
A. Taheri et al. Carbohydrate Polymers 266 (2021) 118148

Undersize (%)
Frequency (%)
Diameter (nm)
(a)

Undersize (%)
Frequency (%)

Diameter (nm)

(b)

Fig. 1. Microscopic images of emulsions with added CSM-Blg complexes are taken with optical microscopy; a) CSM-Blg and b) CSM-Blg-Ca. The scale bar is 1 μm.

Fig. 2. Creaming index of emulsion prepared CSM-Blg mixtures a) 25 ◦ C and b) 4 ◦ C.

stability of emulsions modulating the system viscosity. Araiza-Calahorra et al. (2008) where the droplets in nano range enlarged gradually in the
et al. (2018) reported enhancements in binding affinity between absence of phase separation. This led to sedimentation which contrib­
curcumin-whey proteins in emulsions at pH 7.0 and in the presence of uted to Ostwald ripening and mechanisms of coalescence. On the other
10 mM CaCl2. hand, bigger droplets exhibit higher coalescence rates, as compared to
For a better understanding of storage stability, changes in the size of small droplets at the time of storage. Whereas coalescence is less likely
emulsions over time were investigated. Both o/w emulsions had a nar­ to occur in the case of small droplets, medium-sized and big droplets
row distribution (small span) and the distribution moved toward bigger tend to enlarge. Nonetheless, small droplets increase in number. At the
emulsion size through time (Table 1). The appearance suggests that the time of storage, thus, the droplet size grows bigger because of a higher
fresh emulsions remain homogeneous and no phase separation was rate of Ostwald ripening and coalescence (Capek, 2004; Jafari et al.,
observed for both CSM-Blg and CSM-Blg-Ca during storage (Fig. 3). 2008).
These findings are in agreement with previous results reported by Porras The stability behavior of the emulsions, along with size, could be

4
A. Taheri et al. Carbohydrate Polymers 266 (2021) 118148

Table 1 complex had electrostatic repulsion with droplets. Stable emulsions


Changes in droplet size and span number in emulsions stabilized with CSM-Blg- were prepared using particles as stabilizers, demonstrating that the
Ca & CSM-Blg during 30 days at 25 ◦ C. convective power as mechanical energy input to the oil-water interface
Sample Droplet size (nm) Span is larger than the repulsive forces. Some research has reported that
CSM-Blg
particles with the smallest sizes are in a better situation to create sta­
Day 1 178 ± 10d 0.10 ± 0.02c bility (Li et al., 2014). In the current research, however, other parame­
Day 8 225 ± 13c 0.19 ± 0.03c ters such as surface chemistry and interfacial properties affected the
Day 20 363 ± 21b 0.30 ± 0.05b ability of CSM-Blg and CSM-Blg-Ca to stabilize emulsions.
Day 30 443 ± 27a 0.44 ± 0.07a
Many emulsion-based commercial food products such as sauces,
CSM-Blg-Ca mayonnaise, soups and whipped toppings can have a longer shelf life if
Day 1 345 ± 13d 0.09 ± 0.01c they are frozen. Freezing is usually followed by thawing at most stages of
Day 8 485 ± 20c 0.10 ± 0.01bc
processing and also prior to consumption. Nonetheless, poor levels of
Day 20 610 ± 31b 0.14 ± 0.02b
Day 30 824 ± 47a 0.29 ± 0.03a freeze-thaw stability characterize many food emulsions because droplets
usually coalesce as a consequence of crystallization in the oil and/or
a–d: Means followed by the same lower case in the same column are not
water phases (Thanasukarn et al., 2004). In this study, all emulsions
significantly different (P > 0.05).
were initially evaluated in terms of freeze-thaw stability by successive
acts of freezing (− 18 ◦ C, 24 h and − 30 ◦ C, 18 h) and thawing (25 ◦ C, 10
explained by spontaneous desorption. This holds true when radius r is h). The stability was gauged against creaming, coalescence and oiling
considered for spherical particles and r2 corresponds with the free- off. The hypothesis was that the synergistic effect of CSM and Blg in
energy of spontaneous desorption (Dickinson, 2010). This indicates emulsification led to a thick layer around the emulsion droplets, thereby
that the enhanced particle size of the CSM-Blg-Ca also participated in the preventing instability during freezing (Aoki et al., 2005). Regarding
enhanced physical stability of the CSM-Blg-Ca emulsion. In fact, ac­ soluble soybean polysaccharides in mixture with whey protein isolate,
cording to Coehn's rule, in a dispersion system composed of two non- Cabezas et al. (2019) reported that oil-in-water emulsions can become
conductors, a larger amount of dielectric constant in one phase is posi­ more stable against freeze-thaw due to the nature of the said mixture.
tively charged whereas the other is negatively charged (Kotsuki et al., The emulsification ability of the two components in the mixture
2016). The dielectric constant of water (78.4 F/m) is much greater than enhanced the interactions at the oil/water interface in favor of freeze-
that of corn oil (2–3 F/m). When corn oil was added to the water phase thaw stability.
during emulsification, the surfaces of the droplets were negatively All emulsions were hardly stable against oiling off. The freeze-thaw
charged. The values of zeta potential in CSM-Blg and CSM-Blg-Ca par­ treatment caused negligible change in the emulsions resistance to
ticles were − 54 mV and − 42 mV, respectively, and so the particles in a creaming. The observations intelligibly demonstrated that the CSM-Blg-

Day 1 Day 28

CSM-Blg CSM-Blg-Ca CSM-Blg


CSM-Blg-Ca

(a)

Day 1 Day 28

CSM-Blg-Ca CSM-Blg CSM-Blg-Ca CSM-Blg

(b)
Fig. 3. Visual observation of emulsion prepared CSM-Blg mixtures after 28 days storage a) 4 ◦ C and b) 25 ◦ C.

5
A. Taheri et al. Carbohydrate Polymers 266 (2021) 118148

Ca complex can better assist highly-stable emulsions to take form at both At the CSM-Blg, the interfacial tension reduction seems somewhat less
− 18 and − 30 ◦ C. After thawing at room temperature, the samples extended (from 47.87 to 30.12 mN/m) at CSM-Blg compared to the
showed that both emulsions with CSM-Blg and CSM-Blg-Ca acquire CSM-Blg-Ca (from 47.87 to 28.01 mN/m at the same concentration),
some amount of susceptibility to the aggregation of droplets after probably due to the relatively lower surface charge of the complex in the
freezing and thawing at − 30 ◦ C. There was an increase in the average aqueous phase which affected the surface pressure. A considerable
diameter of particles from 164 nm to 669 nm by CSM-Blg and from 326 reduction was observed in the interfacial tension (from 30.12 to 5.22
nm to 1150 nm by CSM-Blg-Ca after the freeze-thaw treatment. Since mN/m at CSM-Blg) compared to the CSM-Blg-Ca (from 28.01 to 5.18
CSM-Blg-Ca (larger particles) can better prevent re-coalescence through mN/m at the same concentration). In other words, the Blg-CSM mixtures
the emulsifying process, resistance to creaming is improved in associa­ resulted in slightly lower interfacial tension values, accounting for a
tion with coalescence stability. Therefore, it is reasonable to suggest that synergy between the two emulsifiers. Dickinson (2003) also presented
the emulsion does not expand significantly by thawing. This is especially that protein-polysaccharide electrostatic complexes have improved
true in the case of emulsions treated with CSM-Blg-Ca, whereby the interfacial properties compared to the single ingredients (Dickinson,
emulsion is likewise more stable (Chen et al., 2019; Jafari et al., 2008). 2003).
In this study, slow cooling (− 18 ◦ C freezer), as compared to fast cooling One mechanical feature of emulsions is the viscoelasticity of fluid-
(− 30 ◦ C), led to a greater stability of samples, especially because of the fluid interfaces. It is an indicative kinetic of polymer adsorption by an
relatively large CSM-Blg-Ca which increased the viscosity of the unfro­ interface of oil/water and determines the strength of oil-water films
zen aqueous phase. It reduced crystallization and maintained the quality when molecules arrange and stabilize the films. According to Benjamins
of samples through freeze–thaw. Tippetts and Martini (2009) reported and Lucassen-Reynders (1998), the interfacial viscoelastic behavior can
that large and stable crystals are more likely to form when the cooling be considered as important as the interfacial tension value when
rate is slow, thereby resulting in higher stability of the emulsion, assessing interface characteristics, exclusively with the aim of gauging
particularly when samples contain smaller amounts of oil (20%, v/v). how oil droplets are stabilized. On the other hand, the viscoelastic na­
The nature of the interfacial layer between the oil droplets and the ture of an interface can improve how stable emulsions are, while
complexes is a significant parameter that affects the stability of emul­ droplets in the emulsion coalesce and the rate of film thinning is reduced
sions in the freeze-thaw process and also influences indices that define (Tambe & Sharma, 1994). In the case of CSM-Blg-Ca and CSM-Blg
stability, e.g. creaming index (Mun et al., 2008). Thus, the characteris­ (Fig. 5a), the kinetics of surface viscoelasticity and the expansion of
tics of interfacial oil/complexes were evaluated in this study. CSM-Blg-Ca make progress through two steps. It involves a primary and
Fig. 4 demonstrates the alterations of interfacial tension as a function sharp increase in G′ and G′′ , while a slower increase occurs in the second
of adsorption time. After approximately 2 h, the graph showed an im­ step. This highlights the importance of processes in molecular adsorp­
mediate and significant reduction in the oil-water interfacial tension tion and by an assembly between molecules at the nanocomplex/oil
from 23.0 to 17.6 mN/m, upon adding 0.2 (w/w)% Blg to water. This interface. A similar phenomenon exists (i.e. two-step interfacial ageing)
was in agreement with the expected surface characteristics of Blg in protein-adsorbed films and lipophilic surfactants (Opawale &
(Bouyer et al., 2011). Whereas CSM did not indicate any immediate Burgess, 1998).
considerable oil-water interfacial tension decrease upon CSM addition. In contrast, CSM-Blg films can stage rapid kinetics of formation
It is also worth noting that the lowest interfacial tension values recorded (Fig. 5a). Obviously, a predominant elasticity exists in the case of
for CSM alone were similar to those of Blg alone. It means that CSM adsorbed layers in CSM-Blg-Ca and CSM-Blg (G′ > G′′ ). Compared with
displayed good surface characteristics, but it required longer times than CSM-Blg, the network structure of CSM-Blg-Ca was denser, and so the
Blg to reach equilibrium. Comparing the time evolution of the interfacial interface storage modulus (G′ ) and interface loss modulus (G′′ ) of CSM-
tension for oil-Blg, oil-CSM, and oil-Blg-CSM dispersions at 0.2 (w/w)%, Blg-Ca differed from that of CSM-Blg. As a result, its G′ indicated that a
all adsorption isotherms demonstrated similar profiles. For CSM-Blg and densely packed interfacial layer is expanding. The oil-water interfacial
CSM-Blg-Ca mixtures, interfacial behavior indicated an immediate viscoelasticity is a determinant factor by which the stabilization per­
reduction which is similar to that calculated with Blg alone. These formance can be measured. For instance, a higher elastic interface can
findings can be related to the act that CSM is a polysaccharide and re­ reduce the level of breaking off in the oil phase (Jin et al., 2017). The
quires time to be adsorbed on the interface and thus to decrease the strong elastic feature of the interfacial films with CSM-Blg-Ca usually
interfacial tension. A variation could be found when adding the CaCl2. results in a better resistance against coalescence (Romero et al., 2011)
and promotes the stability of emulsions during storage. This result is in
good agreement with previous results about the higher stability of CSM-
Blg-Ca, compared to the CSM-Blg emulsion. At the L/L interface, the
time evolution of G* can be compared with the complexes in the
adsorption films (Fig. 5b). The results suggest that complexes have film-
creating ability at the L/L interface, while G* has a greater value than
the corn/pure water interface in the absence of additives (Fig. 5b).
Layers of adsorption have interfacial features at the liquid/liquid
interfaces. To evaluate their features, amplitude sweeps had to be car­
ried out at a constant frequency of 1 rad/s (Fig. 6a) Even though
emulsifiers stabilized the structure of interfacial layers, this structure is
commonly susceptible to distortion. Thus, the amplitude sweep in CSM-
Blg-Ca and CSM-Blg films are used for evaluating how the interface can
tolerate maximum permitted shear stress, without having to annihilate
the structure. Selected frequencies determine how the amplitude sweep
can function. G′′ is smaller than G′ as an order of magnitude, and this
indicates elasticity and stability for the interfaces. CSM-Blg-Ca and CSM-
Blg film layers demonstrate linear viscoelastic patterns at deformations
of γ < 20% and γ < 5%. In fact, CSM-Blg films create interfacial networks
that are disrupted remarkably at high degrees of deformation (Rühs
Fig. 4. Time evolution of the interfacial tension for oil-0.2 (w/w)% biopolymer et al., 2012). Angular frequencies (ω) were used (5 to 100 rad/s) for
aqueous solutions. measuring frequency-dependent-dynamics at a small strain amplitude of

6
A. Taheri et al. Carbohydrate Polymers 266 (2021) 118148

Fig. 5. Time dependency of the interfacial shear moduli of G′ , G′′ and G* of the complexes (CSM-Blg-Ca & CSM-Blg) at same concentration (0.5% w/w) at liquid/
liquid interfaces (T = 20 ◦ C, deformation amplitude γ 0 = 1%, frequency ω = 1 rad/s).

Fig. 6. Deformation dependence of interfacial shear moduli at a frequency of 1 rad/s (a) and frequency sweep measured at a strain amplitude γ0 = 1% (b) for
complexes (CSM-Blg-Ca & CSM-Blg).

1% (Fig. 6b). CSM-Blg and CSM-Blg-Ca make the elastic behavior


dominant by G′ > G′′ and by being strongly dependent on frequency at
the tested range. No crossover frequency was demonstrated in the case
of G′ and G′′ by complexes adsorption films anywhere in the frequencies
being tested. This suggests that the created layers have viscoelastic
properties which cannot change easily.
The CSM-Blg-Ca and CSM-Blg can stabilize emulsions which then
η (Pa.s)

exhibit interfacial shear viscosities (0.5% w/w) over an extensive


domain of shear values (10− 2 ≤ γ_ ≤ 102) (Fig. 7). They showed different
levels of viscosity and considerable shear-thinning behavior. Further­
more, the evolution of η revealed hysteresis; the amounts that were
calculated in an increasing shear-rate sweep appear to be larger in
comparison with those in a descending shear-rate sweep. This contrib­
uted to a self-disassociation or association at the interface, accompanied
by a breakdown or a reconstruction at the microstructural level (Fig. 7).
On the other hand, the films developed renewed microstructures which
are unlikely to rebound thoroughly after the ruin of interfacial films.
This phenomenon (shear-thinning along with hysteresis) is usually a
general description in some small molecular emulsifiers and in protein- Fig. 7. Steady interfacial shear viscosity of oil/CSM-Blg-Ca and oil/CSM-Blg
based films (Erni et al., 2003; Sharma et al., 2011). during ascending and descending shear rate at the liquid/liquid interfaces.
There is a general agreement about the mechanism which allows
proteins in combination with polysaccharides to develop efficiently into
emulsifiers. The oil surface is where hydrophobic protein-rich parts are

7
A. Taheri et al. Carbohydrate Polymers 266 (2021) 118148

adsorbed and attached, while the aqueous solution is where hydrophilic- Aoki, T., Decker, E. A., & McClements, D. J. (2005). Influence of environmental stresses
on stability of O/W emulsions containing droplets stabilized by multilayered
carbohydrate moieties leak. Bigger sizes of particles usually cause a slow
membranes produced by a layer-by-layer electrostatic deposition technique. Food
absorption of the polysaccharide chains into the interface. After being Hydrocolloids, 19(2), 209–220. https://doi.org/10.1016/j.foodhyd.2004.05.006
located at the interface, the hydrophobic amino acids are exposed when Araiza-Calahorra, A., Akhtar, M., & Sarkar, A. (2018). Recent advances in emulsion-
protein parts are unfolded. In realigning themselves with the oil-phase, based delivery approaches for curcumin: From encapsulation to bioaccessibility. In ,
vol. 71. Trends in food science and technology (pp. 155–169). Elsevier Ltd. https://doi.
these proteins extend along a viscoelastic film where the aqueous phase org/10.1016/j.tifs.2017.11.009.
has carbohydrate moieties. Consequently, a clear increase in G′ occurs in Assadpour, E., Maghsoudlou, Y., Jafari, S. M., Ghorbani, M., & Aalami, M. (2016a).
testing for interfacial shear-rheology. Once a complex is adsorbed on the Optimization of folic acid nano-emulsification and encapsulation by maltodextrin-
whey protein double emulsions. International Journal of Biological Macromolecules,
oil surface, adsorbed molecules can improve in terms of repulsive in­ 86, 197–207. https://doi.org/10.1016/j.ijbiomac.2016.01.064
teractions because polysaccharide macromolecules exhibit electrostatic Assadpour, E., Maghsoudlou, Y., Jafari, S. M., Ghorbani, M., & Aalami, M. (2016b).
and steric hindrance which can contribute to electrostatic stability Evaluation of folic acid nano-encapsulation by double emulsions. Food and Bioprocess
Technology, 9(12), 2024–2032. https://doi.org/10.1007/s11947-016-1786-y
(Dickinson, 2003). By evaluating the interface viscoelasticity or by Bago Rodriguez, A. M., Binks, B. P., & Sekine, T. (2018). Emulsion stabilisation by
measuring how much an adsorbed layer is characterized by mechanical complexes of oppositely charged synthetic polyelectrolytes. Soft Matter, 14(2),
strength, significant information can be made regarding how stable 239–254. https://doi.org/10.1039/c7sm01845b
Benjamins, J., & Lucassen-Reynders, E. H. (1998). Surface dilational rheology of proteins
emulsions can become (Humblet-Hua et al., 2013). However, creating adsorbed at air/water and oil/water interfaces. Studies in Interface Science, 7(C),
simple relations among emulsion stability and interfacial flow behavior 341–384. https://doi.org/10.1016/S1383-7303(98)80056-2
does not seem to be a simplified process (Dickinson et al., 1988; Wil­ Bos, M. A., & Van Vliet, T. (2001). Interfacial rheological properties of adsorbed protein
layers and surfactants: A review. In , vol. 91(3). Advances in colloid and interface
liams et al., 1997). Each adsorbed film has a quantity of interfacial-
science (pp. 437–471). Elsevier. https://doi.org/10.1016/S0001-8686(00)00077-4.
viscoelastic reaction, in addition to rheological characteristics (i.e. Bouyer, E., Mekhloufi, G., Le Potier, I., Kerdaniel, T.d. F.d., Grossiord, J. L., Rosilio, V., &
predominantly elastic or viscos modulus). These values correspond Agnely, F. (2011). Stabilization mechanism of oil-in-water emulsions by
efficiently with the correlation-emulsion stability (i.e., CSM-Blg-Ca > β-lactoglobulin and gum arabic. Journal of Colloid and Interface Science, 354(2),
467–477. https://doi.org/10.1016/j.jcis.2010.11.019
CSM-Blg). The ultimate purpose of the interfacial rheological charac­ Cabezas, D. M., Pascual, G. N., Wagner, J. R., & Palazolo, G. G. (2019). Nanoparticles
teristics is to associate the properties with adsorbed layers to the pro­ assembled from mixtures of whey protein isolate and soluble soybean
cessing and storage-behavior of actual emulsions. Arriving at this aim polysaccharides. Structure, interfacial behavior and application on emulsions
subjected to freeze-thawing. Food Hydrocolloids, 95, 445–453. https://doi.org/
needs more fundamental research where model-systems can be applied, 10.1016/j.foodhyd.2019.04.040
as in this study, and thus valued correlations can arise between bulk Capek, I. (2004). Degradation of kinetically-stable o/w emulsions. Advances in Colloid
stability and interfacial features. and Interface Science, 107(2–3), 125–155. https://doi.org/10.1016/S0001-8686(03)
00115-5
Chang, H. W., Tan, T. B., Tan, P. Y., Abas, F., Lai, O. M., Wang, Y., … Tan, C. P. (2018).
4. Conclusions Physical properties and stability evaluation of fish oil-in-water emulsions stabilized
using thiol-modified β-lactoglobulin fibrils-chitosan complex. Food Research
International, 105, 482–491. https://doi.org/10.1016/j.foodres.2017.11.034
A nanoscale (CSM-Blg and CSM-Blg-Ca) stabilized emulsion was Chen, Y. B., Zhu, X. F., Liu, T. X., Lin, W. F., Tang, C. H., & Liu, R. (2019). Improving
formulated with a narrow droplet size distribution. It had no additive freeze-thaw stability of soy nanoparticle-stabilized emulsions through increasing
particle size and surface hydrophobicity. Food Hydrocolloids, 87, 404–412. https://
based on the combined stabilization effects of the mixed proteins and
doi.org/10.1016/j.foodhyd.2018.08.020
polysaccharides. The dense interfacial layers that packaged the droplets Damodaran, S. (2005). Protein stabilization of emulsions and foams. In , vol. 70(3).
played an important role in stabilizing the emulsions. The results Journal of Food Science (pp. R54–R66). Institute of Food Technologists. https://doi.
org/10.1111/j.1365-2621.2005.tb07150.x.
revealed that emulsions having the CSM-Blg-Ca complex had better
Danov, K. D., Kralchevsky, P. A., Ananthapadmanabhan, K. P., & Lips, A. (2006).
physicochemical characteristics, good storage stability and creaming Particle-interface interaction across a nonpolar medium in relation to the production
than emulsions containing CSM-Blg which contributed to strong elastic of particle-stabilized emulsions. Langmuir, 22(1), 106–115. https://doi.org/
features of the interfacial films. Also, the viscosity of the continuous 10.1021/la052273j
De Folter, J. W. J., Van Ruijven, M. W. M., & Velikov, K. P. (2012). Oil-in-water Pickering
phase increased parallel to the increase in CSM-Blg-Ca content. The ef­ emulsions stabilized by colloidal particles from the water-insoluble protein zein. Soft
fect of the freeze-thaw on the stability of the emulsions demonstrated Matter, 8(25), 6807–6815. https://doi.org/10.1039/c2sm07417f
that CSM-Blg and CSM-Blg-Ca caused high levels of freeze-thaw stability Degner, B. M., Chung, C., Schlegel, V., Hutkins, R., & McClements, D. J. (2014). Factors
influencing the freeze-thaw stability of emulsion-based foods. Comprehensive Reviews
in emulsions. This mixed stabilized emulsion system can have good in Food Science and Food Safety, 13(2), 98–113. https://doi.org/10.1111/1541-
potential for achieving desirable properties in functional foods. 4337.12050
Dickinson, E. (2003). Hydrocolloids at interfaces and the influence on the properties of
dispersed systems. In , vol. 17(1). Food hydrocolloids (pp. 25–39). Elsevier. https://
CRediT authorship contribution statement doi.org/10.1016/S0268-005X(01)00120-5.
Dickinson, E. (2010). Food emulsions and foams: Stabilization by particles. In , vol. 15
(1–2). Current opinion in colloid and interface science (pp. 40–49). Elsevier. https://
Afsaneh Taheri: Conceptualization, Formal analysis, Investigation,
doi.org/10.1016/j.cocis.2009.11.001.
Methodology, Writing – original draft. Mahdi Kashaninejad: Funding Dickinson, E., Murray, B. S., & Stainsby, G. (1988). Coalescence stability of emulsion-
acquisition, Formal analysis, Project administration, Resources, Super­ sized droplets at a planar oil-water interface and the relationship to protein film
surface rheology. Journal of the Chemical Society, Faraday Transactions 1: Physical
vision, Validation, Writing – review & editing. Ali Mohammad Tam­
Chemistry in Condensed Phases, 84(3), 871–883. https://doi.org/10.1039/
addon: Project administration, Resources, Supervision, Validation, F19888400871
Writing – review & editing. Seid Mahdi Jafari: Project administration, Erni, P., Fischer, P., Windhab, E. J., Kusnezov, V., Stettin, H., & Läuger, J. (2003). Stress-
Supervision, Visualization, Writing – review & editing. and strain-controlled measurements of interfacial shear viscosity and viscoelasticity
at liquid/liquid and gas/liquid interfaces. Review of Scientific Instruments, 74(11),
4916–4924. https://doi.org/10.1063/1.1614433
Acknowledgments Esfanjani, A. F., Jafari, S. M., Assadpoor, E., & Mohammadi, A. (2015). Nano-
encapsulation of saffron extract through double-layered multiple emulsions of pectin
and whey protein concentrate. Journal of Food Engineering, 165, 149–155. https://
This research did not receive any specific grant from funding doi.org/10.1016/j.jfoodeng.2015.06.022
agencies in the public, commercial, or not-for-profit sectors. Fan, C., Han, S., Liu, F., Liu, Y., Wang, L., & Pan, S. (2017). Influence of calcium lactate
and pH on emulsification of low-methoxylated citrus pectin in a Pickering emulsion.
Journal of Dispersion Science and Technology, 38(8), 1175–1180. https://doi.org/
References 10.1080/01932691.2016.1230065
Faridi Esfanjani, A., Jafari, S. M., & Assadpour, E. (2017). Preparation of a multiple
emulsion based on pectin-whey protein complex for encapsulation of saffron extract
Ali, A., Mekhloufi, G., Huang, N., & Agnely, F. (2016). β-Lactoglobulin stabilized
nanodroplets. Food Chemistry, 221, 1962–1969. https://doi.org/10.1016/j.
nanemulsions - formulation and process factors affecting droplet size and
foodchem.2016.11.149
nanoemulsion stability. International Journal of Pharmaceutics, 500(1–2), 291–304.
https://doi.org/10.1016/j.ijpharm.2016.01.035

8
A. Taheri et al. Carbohydrate Polymers 266 (2021) 118148

Gharehbeglou, P., Jafari, S. M., Hamishekar, H., Homayouni, A., & Mirzaei, H. (2019). Porras, M., Solans, C., González, C., & Gutiérrez, J. M. (2008). Properties of water-in-oil
Pectin-whey protein complexes vs. small molecule surfactants for stabilization of (W/O) nano-emulsions prepared by a low-energy emulsification method. Colloids
double nano-emulsions as novel bioactive delivery systems. Journal of Food and Surfaces A: Physicochemical and Engineering Aspects, 324(1–3), 181–188. https://
Engineering, 245, 139–148. https://doi.org/10.1016/j.jfoodeng.2018.10.016 doi.org/10.1016/j.colsurfa.2008.04.012
Ghosh, S., & Coupland, J. N. (2008). Factors affecting the freeze-thaw stability of Romero, A., Beaumal, V., David-Briand, E., Cordobés, F., Guerrero, A., & Anton, M.
emulsions. Food Hydrocolloids, 22(1), 105–111. https://doi.org/10.1016/j. (2011). Interfacial and oil/water emulsions characterization of potato protein
foodhyd.2007.04.013 isolates. Journal of Agricultural and Food Chemistry, 59(17), 9466–9474. https://doi.
Gülseren, I., & Corredig, M. (2013). Interactions of chitin nanocrystals with org/10.1021/jf2019853
β-lactoglobulin at the oil-water interface, studied by drop shape tensiometry. Colloids Rühs, P. A., Scheuble, N., Windhab, E. J., Mezzenga, R., & Fischer, P. (2012).
and Surfaces B: Biointerfaces, 111, 672–679. https://doi.org/10.1016/j. Simultaneous control of pH and ionic strength during interfacial rheology of
colsurfb.2013.06.058 β-lactoglobulin fibrils adsorbed at liquid/liquid interfaces. Langmuir, 28(34),
Humblet-Hua, N. P. K., Van Der Linden, E., & Sagis, L. M. C. (2013). Surface rheological 12536–12543. https://doi.org/10.1021/la3026705
properties of liquid-liquid interfaces stabilized by protein fibrillar aggregates and Sharma, V., Jaishankar, A., Wang, Y. C., & McKinley, G. H. (2011). Rheology of globular
protein-polysaccharide complexes. Soft Matter, 9(7), 2154–2165. https://doi.org/ proteins: Apparent yield stress, high shear rate viscosity and interfacial
10.1039/c2sm26627j viscoelasticity of bovine serum albumin solutions. Soft Matter, 7(11), 5150–5160.
Iwanaga, D., Gray, D., Decker, E. A., Weiss, J., & McClements, D. J. (2008). Stabilization https://doi.org/10.1039/c0sm01312a
of soybean oil bodies using protective pectin coatings formed by electrostatic Shimoni, G., Shani Levi, C., Levi Tal, S., & Lesmes, U. (2013). Emulsions stabilization by
deposition. Journal of Agricultural and Food Chemistry, 56(6), 2240–2245. https:// lactoferrin nano-particles under invitro digestion conditions. Food Hydrocolloids, 33
doi.org/10.1021/jf073060y (2), 264–272. https://doi.org/10.1016/j.foodhyd.2013.03.017
Jafari, S. M., Assadpoor, E., He, Y., & Bhandari, B. (2008). Re-coalescence of emulsion Sosa-Herrera, M. G., Berli, C. L. A., & Martínez-Padilla, L. P. (2008). Physicochemical and
droplets during high-energy emulsification. In , vol. 22(7). Food hydrocolloids (pp. rheological properties of oil-in-water emulsions prepared with sodium caseinate/
1191–1202). Elsevier. https://doi.org/10.1016/j.foodhyd.2007.09.006. gellan gum mixtures. Food Hydrocolloids, 22(5), 934–942. https://doi.org/10.1016/j.
Jafari, S. M., He, Y., & Bhandari, B. (2007). Production of sub-micron emulsions by foodhyd.2007.05.003
ultrasound and microfluidization techniques. Journal of Food Engineering, 82(4), Taheri, A., & Jafari, S. M. (2019a). Nanostructures of gums for encapsulation of food
478–488. https://doi.org/10.1016/j.jfoodeng.2007.03.007 ingredients. In Biopolymer nanostructures for food encapsulation purposes (pp.
Jin, Q., Li, X., Cai, Z., Zhang, F., Yadav, M. P., & Zhang, H. (2017). A comparison of corn 521–578). Elsevier. https://doi.org/10.1016/B978-0-12-815663-6.00018-5.
fiber gum, hydrophobically modified starch, gum arabic and soybean soluble Taheri, A., & Jafari, S. M. (2019b). Gum-based nanocarriers for the protection and
polysaccharide: Interfacial dynamics, viscoelastic response at oil/water interfaces delivery of food bioactive compounds. In , vol. 269. Advances in colloid and interface
and emulsion stabilization mechanisms. Food Hydrocolloids, 70, 329–344. https:// science (pp. 277–295). Elsevier B.V.. https://doi.org/10.1016/j.cis.2019.04.009
doi.org/10.1016/j.foodhyd.2017.03.005 Taheri, A., Kashaninejad, M., Tamaddon, A. M., & Jafari, S. M. (2020). Vitamin D3 cress
Kotsuki, K., Obata, S., & Saiki, K. (2016). Self-aligned growth of organic semiconductor seed mucilage-β-lactoglobulin nanocomplexes: Synthesis, characterization,
single crystals by electric field. Langmuir, 32(2), 644–649. https://doi.org/10.1021/ encapsulation and simulated intestinal fluid in vitro release. Carbohydrate Polymers. ,
acs.langmuir.5b03975 Article 117420. https://doi.org/10.1016/j.carbpol.2020.117420
Lam, R. S. H., & Nickerson, M. T. (2014). The properties of whey protein-carrageenan Taheri, A., & Razavi, S. M. A. (2015a). Fabrication of cress seed gum nanoparticles, an
mixtures during the formation of electrostatic coupled biopolymer and emulsion anionic polysaccharide, using desolvation technique: An optimization study.
gels. Food Research International, 66, 140–149. https://doi.org/10.1016/j. BioNanoScience, 5(2), 104–116. https://doi.org/10.1007/s12668-015-0169-6
foodres.2014.08.006 Taheri, A., & Razavi, S. M. A. (2015b). The conformational transitions in organic solution
Li, M., Ma, Y., & Cui, J. (2014). Whey-protein-stabilized nanoemulsions as a potential on the cress seed gum nanoparticles production. International Journal of Biological
delivery system for water-insoluble curcumin. LWT - Food Science and Technology, 59 Macromolecules, 80, 424–430. https://doi.org/10.1016/j.ijbiomac.2015.06.056
(1), 49–58. https://doi.org/10.1016/j.lwt.2014.04.054 Tambe, D. E., & Sharma, M. M. (1994). The effect of colloidal particles on fluid-fluid
Linke, C., & Drusch, S. (2018). Pickering emulsions in foods - opportunities and interfacial properties and emulsion stability. Advances in Colloid and Interface Science,
limitations. Critical Reviews in Food Science and Nutrition, 58(12), 1971–1985. 52(C), 1–63. https://doi.org/10.1016/0001-8686(94)80039-1
https://doi.org/10.1080/10408398.2017.1290578 Thanasukarn, P., Pongsawatmanit, R., & McClements, D. J. (2004). Influence of
McClements, D. J. (2012). Nanoemulsions versus microemulsions: Terminology, emulsifier type on freeze-thaw stability of hydrogenated palm oil-in-water
differences, and similarities. In , vol. 8(6). Soft matter (pp. 1719–1729). The Royal emulsions. Food Hydrocolloids, 18(6), 1033–1043. https://doi.org/10.1016/j.
Society of Chemistry. https://doi.org/10.1039/c2sm06903b. foodhyd.2004.04.010
Mehrnia, M. A., Jafari, S. M., Makhmal-Zadeh, B. S., & Maghsoudlou, Y. (2016). Crocin Tippetts, M., & Martini, S. (2009). Effect of cooling rate on lipid crystallization in oil-in-
loaded nano-emulsions: Factors affecting emulsion properties in spontaneous water emulsions. Food Research International, 42(7), 847–855. https://doi.org/
emulsification. International Journal of Biological Macromolecules, 84, 261–267. 10.1016/j.foodres.2009.03.009
https://doi.org/10.1016/j.ijbiomac.2015.12.029 Williams, A., Janssen, J. J. M., & Prins, A. (1997). Behaviour of droplets in simple shear
Mohammadi, A., Jafari, S. M., Assadpour, E., & Faridi Esfanjani, A. (2016). Nano- flow in the presence of a protein emulsifier. Colloids and Surfaces A: Physicochemical
encapsulation of olive leaf phenolic compounds through WPC-pectin complexes and and Engineering Aspects, 125(2–3), 189–200. https://doi.org/10.1016/S0927-7757
evaluating their release rate. International Journal of Biological Macromolecules, 82, (96)03972-6
816–822. https://doi.org/10.1016/j.ijbiomac.2015.10.025 Xu, A. Y., Melton, L. D., Jameson, G. B., Williams, M. A. K., & McGillivray, D. J. (2015).
Mun, S., Cho, Y., Decker, E. A., & McClements, D. J. (2008). Utilization of polysaccharide Structural mechanism of complex assemblies: Characterisation of beta-lactoglobulin
coatings to improve freeze-thaw and freeze-dry stability of protein-coated lipid and pectin interactions. Soft Matter, 11(34), 6790–6799. https://doi.org/10.1039/
droplets. Journal of Food Engineering, 86(4), 508–518. https://doi.org/10.1016/j. c5sm01378j
jfoodeng.2007.11.002 Zhang, Z. K., Xiao, J. X., & Huang, G. Q. (2020). Pickering emulsions stabilized by
Opawale, F. O., & Burgess, D. J. (1998). Influence of interfacial properties of lipophilic ovalbumin-sodium alginate coacervates. Colloids and Surfaces A: Physicochemical and
surfactants on water- in-oil emulsion stability. Journal of Colloid and Interface Science, Engineering Aspects, 595, Article 124712. https://doi.org/10.1016/j.
197(1), 142–150. https://doi.org/10.1006/jcis.1997.5222 colsurfa.2020.124712

You might also like