You are on page 1of 10

Food Hydrocolloids 109 (2020) 106136

Contents lists available at ScienceDirect

Food Hydrocolloids
journal homepage: http://www.elsevier.com/locate/foodhyd

Mixed plant-based emulsifiers inhibit the oxidation of proteins and lipids in


walnut oil-in-water emulsions: Almond protein isolate-camellia saponin
Shulin Zhang a, Li Tian a, Jianhua Yi b, **, Zhenbao Zhu b, Eric Andrew Decker c,
David Julian McClements c, *
a
College of Biology and Food Engineering, Anyang Institute of Technology, Huanghe Road, An yang, Henan, 455000, PR China
b
School of Food and Biological Engineering, Shaanxi University of Science and Technology, Xuefu Road, Xi’an, Shaanxi, 710021, China
c
Department of Food Science, University of Massachusetts, Amherst, MA, 01003, United States

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

Keywords: The oxidation of lipids and proteins often occurs at the oil-water interface in emulsions, and so emulsifiers may
Emulsions influence the oxidative stability of these systems. The objective of this work was to assess the influence of mixed
Protein oxidation plant-based emulsifiers, almond protein isolate (API) and camellia saponin (CS), on the physical and oxidative
Lipid oxidation
stability of walnut oil-in-water emulsions. Initially, a 5 wt% walnut oil emulsion was formulated using 0.5 wt%
Almond protein
API as an emulsifier, then different levels of CS (1.0–2.0 wt%) were added. Experiments were carried out under
Camellia saponins
conditions where the API and CS had different charges: pH 3 (protein positive, saponin negative) and pH 7 (both
negative). Analysis of the surface-potential and interfacial protein concentration indicated that mixed API-CS
interfacial layers were present at the oil droplet surfaces, with the main driving forces for adsorption being
electrostatic and/or hydrophobic interactions. Emulsions prepared with the mixed emulsifier system had greater
resistance to droplet flocculation than those stabilized by API alone when incubated at 45 � C for 0, 3, and 6 days.
Markers of lipid oxidation (hydroperoxides and TBARS) and protein oxidation (carbonyl formation, sulfhydryl
loss, intrinsic fluorescence loss, and electrophoresis) were recorded during storage. These measurements showed
that API-CS-coated droplets were more resistant to oxidation than API-coated ones. Oxidation was faster at pH 3
than pH 7, which was mainly linked to the higher water-solubility and chemical reactivity of the transition
metals under acidic conditions. Our results indicate that a combination of almond protein and camellia saponin is
suitable for forming plant-based emulsions with high physical and chemical stability.

1. Introduction Almond protein isolate (API) has been reported to have good emulsi­
fying properties in previous studies (Sze-Tao et al., 2000). The produc­
Almonds are widely consumed globally because they have good tion of API from whole intact almonds is not economically viable
flavor, nutritional and health attributes (Sze-Tao and Sathe (2000). because they are too expensive. However, damaged almonds or
Indeed, almonds were the most highly produced tree nut in 2016 (1.06 side-streams from almond production can be utilized for the production
million tons), with the United States being responsible for around half of of almond proteins. The API ingredients obtained can then be used as
global production (Zhang, Zhang, Sheng, Wang, & Fu, 2016). They emulsifiers to fabricate food emulsions, such as beverages, dressings,
provide relatively high levels of dietary proteins (16–22%) and lipids creams, sauces, and desserts (Singh & Sarkar, 2011).
(50–55%), with most of these lipids being unsaturated (85%) (Zhu et al., Owing to the perceived benefits of lipids enriched with poly­
2018). In addition, they are also a good source of dietary fibers, min­ unsaturated fatty acids (PUFAs) on human health, they are commonly
erals, and some vitamins (Sathe & Sze, 1997). The kernels from almonds incorporated into emulsion-based foods and beverages. These lipids are,
can be isolated and used as functional ingredients in various processed however, prone to oxidation, thereby having deleterious effects on the
foods (Sathe, 1993). Many desirable functional attributes of almonds are aroma, taste, texture, color, and nutrient value of emulsified foods (Ma
linked to the proteins they contain (Young & Cunningham, 1991). et al., 2012). Both animal and plant proteins have been reported to be

* Corresponding author.
** Corresponding author.
E-mail addresses: yijianhua1971121@126.com (J. Yi), mcclements@foodsci.umass.edu (D.J. McClements).

https://doi.org/10.1016/j.foodhyd.2020.106136
Received 13 May 2020; Received in revised form 23 June 2020; Accepted 26 June 2020
Available online 29 June 2020
0268-005X/© 2020 Elsevier Ltd. All rights reserved.
S. Zhang et al. Food Hydrocolloids 109 (2020) 106136

effective antioxidants in emulsions (Gumus, Decker, & McClements, the desired physicochemical and functional attributes (Dan, Gochev, &
2017; Qiu, Zhao, Decker, & McClements, 2015; Yang & Xiong, 2015). Miller, 2015; Dan et al., 2012). In principle, the composition, thickness,
The antioxidative activity of these proteins has been ascribed to their packing, and charge of the droplet coatings can be modulated by
potential to sequester and inactivate pro-oxidant transition metals (e.g. selecting appropriate protein-surfactant combinations, which would be
iron and copper), to scavenge free radicals, and to generate thick coat­ postulated to influence the physicochemical stability of emulsions. For
ings around the droplets that separate the unsaturated lipids from the instance, it may be possible to form thick, closely packed, charged in­
prooxidants in the aqueous phase (Elias, Kellerby, & Decker, 2008; terfaces that contain antioxidant emulsifiers (e.g., with free radical
Villiere, Viau, Bronnec, Moreau, & Genot, 2005). scavenging or chelating properties) that are effective at inhibiting
Antioxidant proteins slow down the rate of lipid oxidation in foods, oxidation in emulsions. This could be achieved by judicious choice of
but they are often oxidized themselves during this process (Villiere et al., emulsifier types, concentrations, and orders-of-addition.
2005). Moreover, the products of lipid oxidation often have proox­ Here, we examined the possibility of using a natural plant-based
idative effects on proteins (Estevez, Kylli, Puolanne, Kivikari, & Hei­ surfactant, camellia saponin (CS), to enhance the emulsification and
nonen, 2008; Refsgaard, Tsai, & Stadtman, 2000). Oxidation causes antioxidant properties of almond proteins. Camellia saponins are sec­
many physical and chemical alterations in proteins, including confor­ ondary metabolites of Camellia seeds (Cui et al., 2018; Yu & He, 2018).
mational changes, formation of carbonyl derivatives, destruction of Historically, Camellia has been cultivated in East Asia for the extraction
amino acids (especially lysine, cysteine, methionine, histidine, and of commercial tea oil from its seeds. After oil extraction, the seed
tryptophan), crosslinking, polymerization, and β-scission (Mestdagh, pomace was typically discarded without further use. Recently, however,
Kerkaert, Cucu, & Meulenaer, 2011; Schaich & Karel, 1976). These al­ it has been realized that these agricultural waste products contain sub­
terations typically result in changes in the functional attributes of pro­ stantial quantities of triterpenoids (such as saponins) that have potential
teins, such as a reduction in solubility, a loss of essential amino acids, for application as functional ingredients within foods and other prod­
and the generation of possibly toxic reaction products (Chen, Zhao, Sun, ucts. Camellia triterpenoids have already been marketed in some coun­
& Zhao, 2013; Obando, Papastergiadis, Li, & De Meulenaer, 2015; tries as therapeutic agents for the treatment of liver-related diseases,
Schaich, 2014). Protein oxidation also influences the emulsifying and diabetes, and as wound-healing agents (Cui et al., 2018). However, there
interfacial properties of proteins. For instance, highly oxidized whey commercial success in food applications will depend on them being
protein has been reported to form less elastic interfacial layers compared approved for use in each target market. Camellia saponins contain both
to the non-oxidized version (Berton-Carabin, Schroder, hydrophilic groups (e.g., xylose, arabinose, galactose, and glucuronic
Rovalino-Cordova, Schroen, & Sagis, 2016), which may account for the acid) and hydrophobic groups (e.g., triterpene), and therefore have
reported decrease in the resistance of oxidized emulsions to droplet surface activity (Guo, Tong, Ren, Tu, & Li, 2018). Consequently, these
coalescence (Muijlwijk et al., 2017). Consequently, it is desirable to natural plant-based surfactants are receiving more attention as func­
retard both protein and lipid oxidation in food emulsions to improve tional ingredients in foods due to the desire to replace chemically syn­
their stability and extend their shelf lives. thesized or animal-derived ingredients with more “label-friendly”
In multiphase foods, like emulsions, the oil-water interface is a pri­ alternatives (McClements & Gumus, 2016; Ozturk & McClements,
mary location of lipid and protein oxidation reactions (Berton-Carabin, 2016).
Ropers, & Genot, 2014). In this location, amphiphilic lipid hydroper­ In our previous study, we found that CS can form fine oil droplets at
oxides can interact with hydrophilic transition metals to form free rad­ low surfactant-to-oil ratios, which are highly resistant to alterations in
icals, thereby propagating the lipid oxidation reaction (Berton-Carabin, system environmental conditions, such as temperature, pH, or ionic
Genot, Gaillard, Guibert, & Ropers, 2013; Berton-Carabin et al., 2014; strength (Zhu et al., 2019). Moreover, we found that CS was negatively
McClements & Decker, 2000). Moreover, adsorbed proteins are more charged across the pH ranges typically found in foods (pH 2–9). Other
readily oxidized than non-adsorbed ones (Yi et al., 2019). In the recent researchers, have found the isoelectric point (pI) of almond protein to be
decades, numerous scientists have focused on improving the stability of around pH 4.5 (Sathe, 1993). Consequently, API is positive at pH<4.5,
emulsions by using emulsifier combinations rather than single emulsi­ whereas CS is negative. This means that it may be possible to form mixed
fiers. For example, proteins and other oppositely charged biopolymers API-CS interfaces based on electrostatic attraction under sufficiently
have been successfully used to formulate multilayer emulsions by acidic conditions. Conversely, API and CS both have a negative charge at
layer-by-layer electrostatic deposition (Dickinson, 2011; Evans, Rat­ pH > 4.5. Under these conditions, it may still be possible to form mixed
cliffe, & Williams, 2013; Guzey & McClements, 2006). Multilayer interfaces, but the nature of the interactions involved should be
coatings can be designed to prevent lipid oxidation In many respects: (i) different, leading to the creation of interfacial layers with different
including antioxidant components within them; (ii) creating a steric functional attributes. In particular, one would expect hydrophobic in­
barrier; and (iii) making them positively charged so they repel cationic teractions to be much more important under conditions where the two
transition metal ions (Chaprenet, Berton-Carabin, Elias, & Coupland, types of emulsifier electrically repel each other.
2014; Lesmes, Sandra, Decker, & McClements, 2010). The aim of the current work was to establish if a combination of
The functional attributes of protein emulsifiers can often be extended almond protein (API) and tea saponins (CS) could be used to create
by using them in combination with synthetic or natural surfactants. The plant-based emulsions that were resistant to lipid and protein oxidative
nature of the mixed interfacial layers formed at the droplet surfaces reactions. In addition, the influence of pH on the physicochemical sta­
depends on the surface activities, interactions, and concentrations of the bility of the emulsions was evaluated as this would be expected to alter
proteins and surfactants used (Miller, Fainerman, Makievski, Krgel, & the nature of the interfacial complexes formed. The results of this study
Sinyachenko, 2000). For instance, it is possible for two emulsifiers to may aid in the development of plant-based functional ingredients for
form homogeneously mixed interfaces, phase-separated interfacial do­ formulating vegan or vegetarian products enriched with omega-3 fatty
mains, multilayer interfaces, or interfacial aggregates (Mackie & Wilde, acids.
2005; Pugnaloni, Dickinson, Ettelaie, Mackie, & Wilde, 2004). The kind
of interfaces formed mainly depends on the interactions (attractive or 2. Materials & methods
repulsive) between the two emulsifiers, their order of addition, and their
affinity for the oil-water interface. In some cases, one emulsifier may 2.1. Materials
displace another one from the interface due to competitive adsorption,
which depends on the relative concentrations and surface activities of Almond kernels were obtained from a local market. API was pre­
the two emulsifiers (Dickinson, 1999). Consequently, it is important to pared according to the method of Sze-Tao & Sathe (Sze-Tao et al., 2000).
optimize the proteins and surfactants used to form mixed interfaces with The resulting API powder was characterized according to AOAC (1990)

2
S. Zhang et al. Food Hydrocolloids 109 (2020) 106136

methods (Zhu et al., 2018), and found to contain 89.51% protein, 0.88% respectively) were calculated from the particle size distributions. The
fat, and 0.66% ash. The walnut oil was isolated and purified according to droplet ζ-potential in the emulsions were measured by electrophoresis
our previous method (Yi et al., 2019), where the fatty acid composition (Zetasizer Nano ZS90, Malvern Instruments, UK). Emulsions were
was reported to be: C16:0, 7.1%, C16:1, 0.10%, C18:0, 3.5%, C18:1,17.8%, dispersed into phosphate buffer at a ratio of 1: 200 before measurements
C18:2, 61.9%, C18:3, 9.3%, C20:0, 0.12%, C20:1, 0.22%; tocopherols, to obtain a good signal while avoiding multiple scattering effects.
non-detectable; and hydroperoxides, 1.8 mmol/kg oil. Camellia saponin Microstructural images of the emulsions was acquired by confocal
was donated by Qing Han Bio-technology Co. Ltd. (Hunan, China). It laser scanning microscopy (CLSM, TCS SP2, Leica, Germany) using a
was reported to contain saponins (51.8%), total carbohydrates (36.1%), 65� immersion objective lens with a numerical aperture of 1.0. Emul­
crude proteins (5.6%), ash (2.7%), and water (3.7%). This product was sions were stained with a mixture of 0.02% (w/v) Nile Red (lipid stain)
further dialyzed to removed proteins. Briefly, the Camellia saponin was and 0.1% (w/v) Nile Blue (protein stain). The CLSM device was run in
placed in a dialysis bag (molecular mass cutoff, 8 kDa, Shengya Co., fluorescence mode with excitation being carried out at 633 nm for Nile
Xi’an, China) and dialyzed for 24 h. The external solution was collected Blue and 488 for Nile Red.
and replaced with fresh solution, and then dialysis was repeated two The interfacial protein concentration (mg/m2) in freshly made
more times (3 in total). The dialysates were pooled together for each emulsions was measured using a published method (Yang et al., 2015).
sample and was then concentrated and lyophilized. The dried sample Briefly, emulsions were centrifuged at 4 � C for 1 h at 35,000 g to separate
contained saponins (68.4%), total carbohydrates (25.3%), crude pro­ the serum phase from the cream phase. The non-adsorbed proteins were
teins (0.2%), ash (3.6%), water (2.4%), and total polyphenols 12 mg/g. then recovered from the cream, by washing the cream phase twice with
Isooctane, isopropanol, methanol, 2-propanol, ethyl acetate, buffer solution, and then centrifuging using the same conditions stated
ammonium thiocyanate, 1-butanol, and guanidine hydrochloride 5,5- above. The serum phases were then combined and passed through a
dithio-bis (2-nitrobenzoic acid) were obtained from Aladdin Reagent 0.45 μm Millipore filter to remove any residual oil droplets. The protein
(Shanghai, China). Cumene hydroperoxide, Trichloroacetic acid (TCA), content of the serum phase was detected by the previous method
thiobarbituric acid (TBA), 1,1,3,3-tetraethoxypropane, and 2,4-dinitro­ (Bradford, 1976). The interfacial protein concentration was obtained by
phenylhydrazine (DNPH) were obtained from Adamas Reagent deducting the non-adsorbed serum protein from the proteins initially
(Shanghai, China). All other reagents were purchased from the Yuanye used to produce the emulsions.
Biotechnology Co. Ltd. (Shanghai, China). All chemicals were of
analytical grade unless specified otherwise. Solutions and emulsions 2.4. Determination of lipid and protein oxidation
were prepared using double-distilled and deionized water.
The methods used to monitor lipid and protein oxidative reactions in
2.2. Emulsion preparation emulsions during storage are the same as those described in detail in our
previous study (Zhu et al., 2018). Briefly, lipid oxidation was followed
Preparation of stock emulsions: API and CS were dispersed separately by monitoring the generation of lipid hydroperoxides and 2-thiobarbitu­
into phosphate buffer solutions (10 mM, pH 3 or 7) and stirred slowly ric acid reactive substances (TBARS) during storage. Protein oxidation
overnight at 4 � C to fully hydrate the ingredients. 20 wt % walnut oil was was followed by measuring carbonyl formation, free sulfhydryl groups,
dispersed into 80 wt% emulsifier solution (containing 2.0 wt% API intrinsic tryptophan fluorescence, and electrophoresis (SDS-PAGE)
based on the final stock emulsion weight) with a hand-held homogenizer during storage.
(Model F6/10-10G, FuLuke Fluid Machinery Manufacturing Co. Ltd.
Shanghai, China) for 2 min and coarse emulsions were obtained. Stock 2.5. Statistical analysis
emulsions (pH 3 or 7) were produced by further homogenizing these
coarse emulsions by using a high-pressure homogenization as described All experiments were carried out in triplicate. Statistical analyses
previously (Yi et al., 2019). The stock emulsions were stored at 4 � C were performed utilizing a software package (SPSS 17.0, SPSS Inc.,
prior to use to inhibit lipid oxidation. Chicago, IL, USA). Data are reported as means � standard deviations,
Two sets of primary emulsions (pH 3 and 7) were produced. These and mean values were compared using Duncan’s multiple-range tests to
two sets were produced by diluting the stock emulsions (pH 3 or 7) into identify significant differences (p < 0.05).
phosphate buffer solutions (10 mM, pH 3 or 7) at a mass ratio of 1:4 (w/
w), respectively. The final composition of both primary emulsions was 5 3. Results and discussion
wt % walnut oil and 0.5 wt % API.
Two sets of secondary emulsions (pH 3 and 7) were also produced by 3.1. Physical properties of emulsions
diluting the stock emulsions (pH 3 or 7) into CS solution (containing 1.0
or 2.0 wt % CS, based on the final secondary emulsion weight) at the 3.1.1. Electrical characteristics of emulsion droplets
proportion of 1:4 (w/w) and subsequently adjusting pH to 3 or 7. The Preliminary experiments were carried out by measuring droplet size
final composition of these secondary emulsions was 5 wt % walnut oil, versus API concentration to determine the minimum amount that could
0.5 wt % API, and 0, 1.0, or 2.0 wt % CS. These samples were fully mixed be used to form small droplets under the homogenization conditions
by gently stirring for 4 h and then stored overnight at 4 � C. The oxidation used. These experiments showed that 0.5% API could form stable
stability experiments were carried on by incubating the samples at 45 � C emulsions with little non-absorbed proteins in the aqueous phase. This
in the dark to accelerate lipid and protein oxidation. level of API was therefore used to formulate the emulsions in the
remainder of the study. Preliminary experiments were also carried out to
2.3. Emulsion characterization establish whether stable emulsions could be formed by adding CS to API-
coated lipid droplets. At pH 3, stable emulsions could only be formed
Laser diffraction was used to measure the particle size distributions from 0 to 2 wt% CS, above which appreciable droplet aggregation was
of the emulsions (MasterSizer 2000; Malvern Instruments, Worcester­ observed, presumably attributed to electrostatic charge neutralization
shire, UK) at 25 � C after incubation for 0, 3and 6 days. Small volumes of and bridging flocculation effects, i.e., the sharing of anionic CS mole­
emulsions were added into the test chamber using a pipette. The test cules or micelles between the surfaces of two or more cationic droplets.
chamber was filled with phosphate buffer solution at the appropriate At pH 7, the emulsions were stable even at the highest CS concentration
sample pH. Sufficient emulsion was added to obtain a good light scat­ added (5 wt %), which was probably ascribed to the strong electrostatic
tering signal without introducing multiple light scattering effects. The repulsion between the droplets. For this reason, emulsions were pre­
d4,3 and d3,2 (the volume and surface weighted mean particle diameters, pared containing 1.0 and 2.0 wt % CS so that the effects of pH could be

3
S. Zhang et al. Food Hydrocolloids 109 (2020) 106136

compared under similar surfactant conditions. concentration measurements indicated that the almond proteins were
At pH 3, the ζ-potential of the emulsions stabilized by 1 and 2 wt % not removed from the oil-water interface when CS was added (data not
CS alone were 11.1 and 13.2 mV, respectively (data not shown). shown), suggesting the formation of interfacial API-CS complexes or the
Meanwhile, the ζ-potential of the emulsion stabilized by API alone was co-adsorption of API and CS. One would anticipate an electrostatic
þ16.6 mV (Fig. 1a). The oil droplets were positively charged because the repulsion between the anionic API and anionic CS molecules, indicating
pH was appreciably below the pI of the adsorbed almond proteins. The that other types of attractive interactions are important. CS contains
ζ-potential became less and less positive as more and more CS was both hydrophobic and hydrophilic regions on the molecule, and so it
added. This result is consistent with either the formation of interfacial could interact with API through hydrophobic and/or hydrophilic
CS-API complexes through electrostatic attraction or the partial attraction (McClements & Jafari, 2018). Other researchers have also
displacement of cationic API by anionic CS due to competitive adsorp­ reported that anionic proteins can form complexes with anionic sur­
tion. For this reason, we measured the interfacial protein concentrations factants under neutral conditions, such as pea proteins and quillaja sa­
to provide more insights into the origin of this effect. The addition of 1 or ponins (Reichert, Salminen, Leuenberger, Hinrichs, & Weiss, 2015) or
2 wt % CS did not change the interfacial protein concentration (data not β-lactoglobulin and SDS (Magdassi, Vinetsky, & Relkin, 1996), Alter­
shown), which suggests that the observed changes in ζ-potential were natively, the CS may adsorb directly onto the oil-water interface leading
due to the formation of interfacial CS-API complexes. Other researchers to a mixed interfacial layer with surfactant-rich and protein-rich do­
have also reported that protein-surfactant complexes are formed when mains (Pugnaloni et al., 2004).
ionic surfactants are added to oppositely-charged protein-coated drop­ There were some changes in the droplet potentials in the various
lets (Dan et al., 2015; Dan et al., 2012). emulsions during storage (Fig. 1). In the absence of CS, the magnitude of
At pH 7, the ζ-potential of the emulsions stabilized by 1 and 2 wt % the ζ-potential increased during incubation at pH 3 (becoming more
CS alone were 41.2 and 41.5 mV, respectively (data not shown). positive), but decreased at pH 7 (becoming less negative). These changes
Meanwhile, the ζ-potential of the emulsion coated by API alone was in surface potential are indicative of modifications in the interfacial
24.3 mV (Fig. 1b). The ζ-potential became more negative as the CS composition of the emulsions. Oxidation alters the molecular structure
level was increased, suggesting that some of CS adsorbed to the almond of the adsorbed proteins (discussed later), which may change the elec­
protein-stabilized oil droplets, despite the fact that both the surfactant trical properties of the amino acid residues at the droplet exteriors
and protein were negatively charged. The interfacial protein (Gumus et al., 2017). Alternatively, changes in the droplet aggregation
state during storage (discussed later) might alter the magnitude of the
ζ-potential determined by the electrophoresis instrument (McClements,
2015). Other researchers have also reported that the ζ-potential of oil
droplets coated by plant proteins (lentil, pea, and faba bean) changed
throughout storage under neutral conditions, however, in that case the
ζ-potential became more negative (Gumus et al., 2017). The different
trends observed in various studies may be attributed to the different
protein types used, which may undergo different oxidation reactions. In
the presence of CS, the changes in surface potential during storage were
much smaller (Fig. 1a–b). This suggests that the presence of the saponins
may have inhibited any chemical changes at the droplet surfaces.

3.1.2. Aggregation state of emulsion droplets


The volume-weighted mean particle diameter (d4,3) is highly
dependent on the existence of any large particles in colloidal disper­
sions, and so it is useful for characterizing the aggregation stability of
emulsions (McClements, 2015). As shown in Fig. 2, all of the emulsions
had smaller d4,3 values at pH 7 than pH 3 (p<0.05), suggesting that less
droplet aggregation occurred at the higher pH. This is probably because
the absolute value of the ζ-potential was higher at pH 7, resulting to a
stronger electrostatic repulsion between the droplets. Addition of CS
into the initial emulsions caused a significant reduction in d4,3 at both
pH 3 and 7 (p<0.05), but little change in d3,2 (data not shown). This
suggests that the addition of the saponins promoted the dissociation of
some of the flocs in the emulsions. At pH 7.0, the incorporation of CS
would have increased the droplet electrostatic repulsion by increasing
their negative surface potentials. At pH 3.0, however, the added CS
actually decreased the magnitude of their surface potential, which
means this phenomenon cannot account for the observed results.
Instead, the presence of the saponins may have reduced any hydro­
phobic attraction between the protein-coated droplets by covering any
exposed non-polar regions, or they may have increased the steric
repulsion by forming thicker interfacial layers.
The mean particle size of all emulsions increased during storage,
indicating that some droplet aggregation occurred (Fig. 2a–b). However,
Fig. 1. Changes in the ξ-potential of emulsions during storage. (a) pH 3.0; (b)
the increase was much smaller in the emulsions containing CS than in
pH 7.0. Key: API: emulsions stabilized by API alone; APIþ1%CS or APIþ2%CS:
the ones without CS, and was considerably less at pH 7 than at pH 3. Our
emulsions stabilized by API and either 1% or 2% CS. Samples with different
lowercase superscripts (a–c) differ significantly (p < 0.05) when different results demonstrate that the API-CS coatings enhanced the aggregation
emulsions are compared at the same oxidation time; Samples with different stability of the emulsions, when compared to API coatings alone.
uppercase superscripts (A–C) differ significantly (p < 0.05) when the same
emulsion is compared at different oxidation times.

4
S. Zhang et al. Food Hydrocolloids 109 (2020) 106136

oxidation products was significantly lower in the emulsions containing


CS (p < 0.05). Our results indicate that the adsorbed API-CS complexes
were better interfacial antioxidants than the API alone, which may have
been due to various reasons. First, the adsorbed CS molecules have some
inherent antioxidant activity, including free radical scavenging and
ferrous ion chelating (Joshi, Sood, Dogra, & Mahendru …, 2013;
Xiao-Ling, Qiu, Sun, & Zhao-Jiang, 2005). Second, the API-CS may form
a physical barrier that sterically inhibits water-soluble pro-oxidants
reaching the emulsified lipids. Interfacial complexes have also been
reported to improve the oxidative stability of other types of emulsions
due to their ability to form thicker coatings around the oil droplets, such
as pectin/fibroin (Chen, Li, Ding, & Rao, 2011) and pec­
tin/β-lactoglobulin (Katsuda, McClements, Miglioranza, & Decker,
2008). An alternative explanation is that the non-adsorbed CS could
have some antioxidant activity. Previous studies with proteins have
shown that non-adsorbed proteins can bind pro-oxidants (such as tran­
sition metal ions) and therefore inhibit their ability to adsorb to the
droplet surfaces and promote oxidation (Gumus et al., 2017).
The droplets in the emulsions were positive at pH 3 but negative at
pH 7 (Fig. 1). Consequently, one might expect slower oxidation at pH 3
than pH 7 due to electrostatic repulsion of the cationic transition metal
ions from the cationic droplets under acidic conditions (McClements
et al., 2000). In practice, however, the emulsions were more resistant to
oxidation under neutral pH conditions. For example, the lipid hydro­
peroxide concentration of the API-emulsions increased around 81%
(6.05–10.95 mmol/L) after 6 days storage at pH 3 (Fig. 4a), but only
61% (5.56–8.93 mmol/L) at pH 7 (Fig. 4b). Similar findings have been
reported for lipid droplets coated by other kinds of proteins, such as
whey proteins and casein (Haahr & Jacobsen, 2008; Yesiltas,
Garcia-Moreno, Sorensen, & Jacobsen, 2011). These authors proposed
that the higher rate of oxidation observed under acidic conditions was
primarily a result of the higher water-solubility and chemical reactivity
of the iron at low pH, as well as pH-dependent changes in interfacial
Fig. 2. Changes in mean particle diameter (d4,3) during storage: (a) pH 3.0; (b) composition. In our study, we also found that the interfacial protein
pH 7.0. See caption to Fig. 1 for description of sample codes and statisti­ composition in the emulsions was markedly different at pH 3 and 7
cal analysis. (discussed below). This phenomenon may therefore have contributed to
the differences in oxidation stability observed at the different pH values.
3.1.3. Microstructure of emulsions In addition, non-adsorbed proteins can play a critical role in lipid
Images of the microstructures of emulsions containing either 0 or 2% oxidation under neutral conditions. Anionic non-adsorbed proteins can
CS were acquired by CLSM (Fig. 3). At pH 3, the microscopy images bind cationic transition metal ions, inhibiting their interactions with the
show that the oil droplets were highly flocculated in the fresh emulsions encapsulated lipids (Gumus et al., 2017).
stabilized by API alone but less so in the ones stabilized by API-CS
(Fig. 3a). At longer storage times, however, extensive aggregation was 3.3. Protein oxidation
observed in all the emulsions under acidic conditions. The instability
observed in these emulsions may be a result of the relatively low surface Proteins adsorbed to oil droplet surfaces may be oxidized more
potential of the droplets or due to the impact of oxidation on their readily than the oil itself, thereby altering their susceptibility to oxida­
physical stability (discussed later). At pH 7, all emulsions were consid­ tion (Berton, Ropers, Guibert, Sol�e, & Genot, 2012; Yang et al., 2015).
erably more stable to droplet aggregation than at pH 3 (Fig. 3b), which For instance, reactive species generated during lipid oxidation, such as
may have been due to the strong electrostatic repulsive forces acting free radicals, can interact with proteins, thereby promoting their
between them under neutral conditions. Interestingly, the microscopy oxidation (Stadtman, 2006, pp. 22–38). For this reason, protein oxida­
images indicated that there was a substantial increase in droplet floc­ tion within the emulsions was also monitored with and without CS.
culation in the emulsions containing CS during storage (Fig. 3), which
was not observed in the light scattering experiments (Fig. 2). The most 3.3.1. Protein-bound carbonyls
likely reason is the different principles of the light scattering and mi­ The formation of carbonyl derivatives is one of the first stages of
croscopy measurements. Light scattering measurements require dilution protein oxidation (Berton et al., 2012). The carbonyl content of the
and stirring of the samples that can disrupt weak flocs, whereas micro­ emulsions was therefore measured to evaluate protein oxidation (Fig. 5).
scopy simply involves placing an emulsion on a slide that might leave The protein-bound carbonyls increased progressively during 6 days of
these flocs intact. storage in all the emulsions, which is indicative of the proteins being
progressively oxidized. No significant differences in carbonyl levels
3.2. Lipid oxidation were observed between the initial emulsions in the absence or presence
of CS at both pH 3 and 7 (p>0.05). During storage, however, the rate of
Lipid oxidation was investigated by monitoring the generation of carbonyl formation was appreciably less in the presence of CS at both pH
primary (lipid hydroperoxides) and secondary (TBARS) reaction prod­ values, suggesting it inhibited protein oxidation. Carbonyls can arise
ucts during storage (Fig. 4). Both the primary and secondary products from the direct attack of arginine, lysine, proline, and threonine, which
increased progressively throughout storage in all emulsions, suggesting is catalyzed by transition metal ions. Carbonyls can also be produced by
that lipid oxidation occurred. At both pH 3 and 7, the formation of lipid reactions between lipid oxidation products and functional groups on the

5
S. Zhang et al. Food Hydrocolloids 109 (2020) 106136

Fig. 3. Confocal micrographs of emulsions with 0 or 2% added CS at different oxidation times. (a) pH 3.0; (b) pH 7.0.

proteins, such as lysyl, cysteyl, and histidyl residues (Chen, Chen, Ren, & oxidation, the rate of protein oxidation was greater at pH 3 than pH 7
Zhao, 2011). The presence of the saponins at the oil droplet surfaces may (Fig. 5a–b), which again may be due to the increased water-solubility
have inhibited protein oxidation due to their antioxidant properties such and chemical reactivity of the transition metals under acidic conditions.
as their capacities for chelating transition metal ions and scavenging free
radicals or the ability of API-CS complexes to form a thick impermeable 3.3.2. Free sulfhydryls
coating that acted as a steric barrier against the contact between pro­ Changes in the number of free sulfhydryl groups within proteins can
teins and the pro-oxidants in the aqueous phase. Moreover, the reduced be used as a measure of their oxidation state: the lower the number, the
generation of lipid oxidation products in the emulsions containing CS greater the extent of oxidation (Zhu et al., 2018). The free sulfhydryl
(Fig. 4) may have reduced the rate of protein oxidation. Same as lipid concentration decreased significantly (p < 0.05) in all the emulsions

6
S. Zhang et al. Food Hydrocolloids 109 (2020) 106136

Fig. 5. Formation of protein carbonyls in different emulsions during storage.


(a) pH 3.0; (b) pH 7.0. See caption to Fig. 1 for description of sample codes and
statistical analysis.

during storage (Fig. 6a–b), indicating that the proteins were undergoing
oxidation. Again protein oxidation was significantly faster at pH 3 than
pH 7, and was slower in the emulsions containing CS (p < 0.05).

3.3.3. Intrinsic tryptophan fluorescence


A decrease in the intrinsic tryptophan fluorescence of proteins is also
indicative of their oxidation (Berton et al., 2012), which is a result of a
modification of the tryptophan indole ring through free radical re­
actions, as well as changes in the environment of the tryptophan groups
caused by conformational changes (Rampon et al., 2003). The loss of
intrinsic tryptophan fluorescence of the emulsions was therefore moni­
tored during oxidation (Fig. 7a–b). The fluorescence intensity in the
primary emulsions decayed faster than in the secondary emulsions at
both pH 3 and 7. In contrast, the reduction of intrinsic tryptophan
fluorescence was less marked in the secondary emulsion stabilized by
the API-CS complexes. For instance, at pH 3, the intrinsic tryptophan
fluorescence intensity at the peak of emission decreased 18% after 6
days of incubation, while it only reduced 11% in the secondary emulsion
stabilized by the API-CS complexes (Fig. 7a). These observations
therefore further supported the above results that the proteins in the
secondary emulsions were more oxidatively stable than their counter­
parts in the primary emulsions. In addition, a greater decrease in tryp­
tophan fluorescence intensity was observed in the emulsions at pH 7
than at pH 3 (Fig. 7a and b), which is consistent with the results dis­
cussed earlier.

Fig. 4. Formation of lipid hydroperoxides (a, b) and TBARS (c, d) in different 3.3.4. SDS-PAGE analysis
emulsions during storage. (a, c) pH 3.0; (b, d) pH 7.0. See caption to Fig. 1 for Information about the impact of CS addition on the characteristics of
description of sample codes and statistical analysis. the almond proteins in the emulsions was obtained using gel

7
S. Zhang et al. Food Hydrocolloids 109 (2020) 106136

Fig. 6. Changes in free sulfhydryls in different emulsions during storage. (a) pH


3.0; (b) pH 7.0. See caption to Fig. 1 for description of sample codes and sta­
tistical analysis.

electrophoresis. Proteins isolated from emulsions stabilized by API or


API-CS were characterized by non-reducing SDS-PAGE after 0, 3 and 6
days storage in the dark at 37 � C (Fig. 8a and b). Interestingly, the nature
of the proteins in the emulsions was strongly pH-dependent. At pH 3, the
proteins formed a number of bands with molecular weights below 40
kDa. Conversely, at pH 7, the proteins formed five major bands with
molecular weights of 62, 40, 29 27, and 15 kDa. The reason for this big Fig. 7. Changes in intrinsic tryptophan fluorescence in different emulsions
difference in protein composition is unknown. It is possible that an during storage. (a) pH 3.0; (b) pH 7.0.
acidic environment promoted the disruption of some chemical bonds
between or within the protein molecules, thereby generating a series of 4. Conclusion
smaller polypeptides. This difference in protein composition may at least
partially account for the difference in physical and chemical stability of To summarize, we have shown that adsorbing tea saponins onto
the emulsions observed at pH 3 and 7. almond protein-coated lipid droplets improves their physical and
The band intensities for all the emulsions decreased during incuba­ chemical stability. In particular, the mixed protein/saponin-coatings
tion, suggesting that the proteins had become oxidized and degraded. inhibit lipid and protein oxidation more effectively than protein-
Before incubation, the emulsions had similar SDS-PAGE profiles under coatings alone. The driving force for saponin adsorption depends on
the same pH conditions in the absence or presence of CS, indicating that pH. Both electrostatic and hydrophobic attraction are likely to be
the surfactant itself did not directly alter protein composition. However, important when the proteins and saponins have opposite charges (pH <
the reduction in band intensities during storage was slower in the pI), but hydrophobic attraction is likely to be the most important when
presence of CS, again demonstrating its ability to inhibit protein they have similar charges (pH > pI). There are at least two possible
oxidation. For example, at pH 3, there were only two obscure protein mechanisms of action for the antioxidant properties of the saponins: (i)
bands remaining in the absence of CS after 6 days storage, while there the saponins have free radical scavenging and metal chelation proper­
were four relatively clear bands in the presence of CS (see the green ties; (ii) the protein-saponin coatings create a steric barrier that prevents
square in Fig. 8a). The reduction in band intensities was also slower at water-soluble pro-oxidants reaching the lipids. Oxidation was faster at
pH 7 than at pH 3, again indicating that protein degradation was faster pH 3 than at pH 7, which was mainly attributed to the relatively high
under acidic conditions. water-solubility and chemical reactivity of the metal ions under acidic
conditions. In addition, there were differences in the protein composi­
tion at the different pH values, which may also have contributed to
differences in oxidative stability. Our results may lead to the formulation

8
S. Zhang et al. Food Hydrocolloids 109 (2020) 106136

References

AOAC, Association of Official Analytical Chemists. (1990). Official methods of analysis


(15th ed.) Arlington, Vol. A.
Berton-Carabin, C., Genot, C., Gaillard, C., Guibert, D., & Ropers, M. H. (2013). Design of
interfacial films to control lipid oxidation in oil-in-water emulsions. Food
Hydrocolloids, 33(1), 99–105.
Berton-Carabin, C. C., Ropers, M. H., & Genot, C. (2014). Lipid oxidation in oil-in-water
emulsions: Involvement of the interfacial layer. Comprehensive Reviews in Food
Science and Food Safety, 13(5), 945–977.
Berton-Carabin, C. C., Schroder, A., Rovalino-Cordova, A., Schroen, K., & Sagis, L.
(2016). Protein and lipid oxidation affect the viscoelasticity of whey protein layers at
the oil-water interface. European Journal of Lipid Science and Technology, 118(11),
1630–1643.
Berton, C., Ropers, M. H., Guibert, D., Sol� e, V., & Genot, C. (2012). Modifications of
interfacial proteins in oil-in-water emulsions prior to and during lipid oxidation.
Journal of Agricultural and Food Chemistry, 60(35), 8659–8671.
Bradford, M. M. (1976). A rapid and sensitive method for the quantitation of microgram
quantities of protein utilizing the principle of protein-dye binding. Analytical
Biochemistry, 72(1-2), 248–254.
Chaprenet, J., Berton-Carabin, C. C., Elias, R. J., & Coupland, J. N. (2014). Effect of
interfacial properties on the reactivity of a lipophilic ingredient in multilayered
emulsions. Food Hydrocolloids, 42, 56–65. pt.1.
Chen, L., Chen, J., Ren, J., & Zhao, M. (2011). Effects of ultrasound pretreatment on the
enzymatic hydrolysis of soy protein isolates and on the emulsifying properties of
hydrolysates. Journal of Agricultural and Food Chemistry, 59(6), 2600–2609.
Chen, B., Li, H., Ding, Y., & Rao, J. (2011). Improvement of physicochemical stabilities of
emulsions containing oil droplets coated by non-globular protein–beet pectin
complex membranes. Food Research International, 44(5), 0-1475.
Chen, N., Zhao, Q., Sun, W., & Zhao, M. (2013). Effects of malondialdehyde modification
on the in vitro digestibility of soy protein isolate. Journal of Agricultural and Food
Chemistry, 61(49), 12139–12145.
Cui, C. J., Zong, J. F., Sun, Y., Zhang, L., Ho, C. T., Wan, X. C., et al. (2018). Triterpenoid
saponins from the genus camellia: Structures, biological activities, and molecular
simulation for structure-activity relationship. Food & Function, 9(6), 3069–3091.
Dan, A., Gochev, G., & Miller, R. (2015). Tensiometry and dilational rheology of mixed
β-lactoglobulin/ionic surfactant adsorption layers at water/air and water/hexane
Fig. 8. SDS-PAGE electrophoretic profiles of proteins from different emulsions interfaces. Journal of Colloid and Interface Science, 449, 383–391.
Dan, A., Kotsmar, C., Ferri, J. K., Javadi, A., Karbaschi, M., Kr€ agel, J., Wüstneck, R., &
under nonreducing conditions. M: marker; 0%CS: emulsions stabilized by API
Miller, R. (2012). Mixed protein–surfactant adsorption layers formed in a sequential
alone; 1%CS: emulsions stabilized by API and 1% CS; and, 2%CS: emulsions and simultaneous way at water–air and water–oil interfaces. Soft Matter, 8(22),
stabilized by API and 2% CS. 6057.
Dickinson, E. (1999). Adsorbed protein layers at fluid interfaces: Interactions, structure
and surface rheology. Colloids and Surfaces B: Biointerfaces, 15(2), 161–176.
of plant-based foods fortified with polyunsaturated lipids that are suit­ Dickinson, E. (2011). Mixed biopolymers at interfaces: Competitive adsorption and
able for vegans and vegetarians. multilayer structures. Food Hydrocolloids, 25(8), 1966–1983.
Elias, R. J., Kellerby, S. S., & Decker, E. A. (2008). Antioxidant activity of proteins and
peptides. Critical Reviews in Food Science and Nutrition, 48, 430–441.
Declaration of competing interest Estevez, M., Kylli, P., Puolanne, E., Kivikari, R., & Heinonen, M. (2008). Fluorescence
spectroscopy as A novel approach for the assessment of myofibrillar protein
The authors state that they have no conflicts of interest related to this oxidation in oil-in-water emulsions. Meat Science, 80(4), 1290–1296.
Evans, M., Ratcliffe, I., & Williams, P. A. (2013). Emulsion stabilisation using
study. polysaccharide-protein complexes. Current Opinion in Colloid & Interface Science, 18
(4), 272–282.
Gumus, C. E., Decker, E. A., & McClements, D. J. (2017). Impact of legume protein type
CRediT authorship contribution statement
and location on lipid oxidation in fish oil-in-water emulsions: Lentil, pea, and faba
bean proteins. Food Research International, 100, 175–185. S0963996917304684.
Shulin Zhang: Conceptualization, Data curation, Formal analysis, Guo, N., Tong, T., Ren, N., Tu, Y., & Li, B. (2018). Saponins from seeds of genus camellia:
Phytochemistry and bioactivity. Phytochemistry, 149, 42–55.
Investigation, Methodology. Li Tian: Conceptualization, Data curation,
Guzey, D., & McClements, D. J. (2006). Formation, stability and properties of multilayer
Formal analysis, Investigation, Methodology. Jianhua Yi: Conceptual­ emulsions for application in the food industry. Advances in Colloid and Interface
ization, Data curation, Formal analysis, Investigation, Methodology, Science, 128, 227–248.
Writing - original draft. Zhenbao Zhu: Conceptualization, Funding Haahr, A. M., & Jacobsen, C. (2008). Emulsifier type, metal chelation and pH affect
oxidative stability of n-3-enriched emulsions&nbsp. European Journal of Lipid Science
acquisition, Writing - original draft. Eric Andrew Decker: Conceptu­ and Technology, 110(10), 949–961.
alization, Writing - review & editing. David Julian McClements: Joshi, R., Sood, S., Dogra, P.Mahendru, M., … (2013). In vitro cytotoxicity,
Conceptualization, Writing - review & editing. antimicrobial, and metal-chelating activity of triterpene saponins from tea seed
grown in Kangra valley, India. Medicinal Chemistry Research, 22(8), 4030–4038.
Katsuda, M. S., McClements, D. J., Miglioranza, L. H. S., & Decker, E. A. (2008). Physical
Acknowledgements and oxidative stability of fish oil-in-water emulsions stabilized with β-lactoglobulin
and pectin. Journal of Agricultural and Food Chemistry, 56(14), 5926–5931.
Lesmes, U., Sandra, S., Decker, E. A., & McClements, D. J. (2010). Impact of surface
Financial support was received from the National Natural Science deposition of lactoferrin on physical and chemical stability of omega-3 rich lipid
Foundation of China under Grants 31671888 and science and technol­ droplets stabilised by caseinate. Food Chemistry, 123(1), 99–106.
ogy plan project of Shaanxi under Grants 2019NY-122. This material Mackie, A., & Wilde, P. (2005). The role of interactions in defining the structure of mixed
protein-surfactant interfaces. Advances in Colloid and Interface Science, 117(1-3),
was also partly based upon research supported by the National Institute 3–13.
of Food and Agriculture, USDA, Massachusetts Agricultural Experiment Ma, H., Forssell, P., Kylli, P., Lampi, A. M., Buchert, J., Boer, H., et al. (2012).
Station (MAS00491). Transglutaminase catalyzed cross-linking of sodium caseinate improves oxidative
stability of flaxseed oil emulsion. Journal of Agricultural and Food Chemistry, 60(24),
6223–6229.
Appendix A. Supplementary data Magdassi, S., Vinetsky, Y., & Relkin, P. (1996). Formation and structural heat-stability of
β-lactoglobulin/surfactant complexes. Colloids and Surfaces B: Biointerfaces, 6(6),
353–362.
Supplementary data to this article can be found online at https://doi.
org/10.1016/j.foodhyd.2020.106136.

9
S. Zhang et al. Food Hydrocolloids 109 (2020) 106136

McClements, D., & Jafari, S. (2018). Improving emulsion formation, stability and Sathe, S. K., & Sze, K. W. C. (1997). Thermal aggregation of almond protein isolate. Food
performance using mixed emulsifiers: A review. Advances in Colloid and Interface Chemistry, 59(1), 95–99.
Science, 251, 55–79. Schaich, K. M., & Karel, M. (1976). Free radical reactions of peroxidizing lipids with
McClements, D. J. (2015). Food emulsions: Principles, practice, and techniques. Boca Raton, amino acids and proteins: An ESR study. Lipids, 11(5).
FL: CRC Press. Schaich, K., & Schaich, K. M. (2014). Lipid co-oxidation of proteins: One size does not fit
McClements, D. J., & Decker, E. A. (2000). Lipid oxidation in oil-in-water emulsions: all. Inform, 25(3): 134-139. Inform, 25(3), 134–139, 2014.
Impact of molecular environment on chemical reactions in heterogeneous food Singh, H., & Sarkar, A. (2011). Behaviour of protein-stabilised emulsions under various
systems. Journal of Food Science, 65(8), 1270–1282. physiological conditions. Advances in Colloid and Interface Science, 165(1), 47–57.
McClements, D. J., & Gumus, C. E. (2016). Natural emulsifiers - biosurfactants, Stadtman, E. R. (2006). Protein oxidation in aging and age-related Diseases (Vol. 928).
phospholipids, biopolymers, and colloidal particles: Molecular and physicochemical Annals of the New York Academy of Sciences. HEALTHY AGING FOR FUNCTIONAL
basis of functional performance. Advances in Colloid and Interface Science, 234, 3–26. LONGEVITY: MOLECULAR AND CELLULAR INTERACTIONS IN SENESCENCE.
Mestdagh, F., Kerkaert, B., Cucu, T., & Meulenaer, B. D. (2011). Interaction between Sze-Tao, K. W. C., & Sathe, S. K. (2000). Functional properties and in vitro digestibility of
whey proteins and lipids during light-induced oxidation. Food Chemistry, 126(3), almond (Prunus dulcis L.) protein isolate. Food Chemistry, 69(2), 153–160.
1190–1197. Villiere, A., Viau, M., Bronnec, I., Moreau, N., & Genot, C. (2005). Oxidative stability of
Miller, R., Fainerman, V. B., Makievski, A. V., Krgel, J., & Sinyachenko, O. V. (2000). bovine serum albumin- and sodium caseinate-stabilized emulsions depends on metal
Dynamics of protein and mixed protein/surfactant adsorption layers at the water/ availability. Journal of Agricultural and Food Chemistry, 53(5), 1514–1520.
fluid interface. Advances in Colloid and Interface Science, 86(1-2), 39–82. Xiao-Ling, L. U., Qiu, S. S., Sun, X. X., & Zhao-Jiang, L. I. (2005). Preliminary study on
Muijlwijk, K., Colijn, I., Harsono, H., Krebs, T., Berton-Carabin, C., & Schroen, K. (2017). the capability of antioxidation and scavenging free radicals of sasanquasaponins.
Coalescence of protein-stabilised emulsions studied with microfluidics. Food Food Science, 26(11), 86–90.
Hydrocolloids, 70, 96–104. Yang, J., & Xiong, Y. L. (2015). Inhibition of lipid oxidation in oil-in-water emulsions by
Obando, M., Papastergiadis, A., Li, S., & De Meulenaer, B. (2015). Impact of lipid and interface-adsorbed myofibrillar protein. Journal of Agricultural and Food Chemistry,
protein Co-oxidation on digestibility of dairy proteins in oil-in-water (O/W) 151005142720003.
emulsions. Journal of Agricultural & Food Chemistry. acs.jafc.5b03563. Yesiltas, B., Garcia-Moreno, P., Sorensen, A., & Jacobsen, C. (2011). Oxidative stability of
Ozturk, B., & McClements, D. J. (2016). Progress in natural emulsifiers for utilization in 70% fish oil-in-water emulsions: Impact of emulsifiers and pH. European Journal of
food emulsions. Current Opinion in Food Science, 7, 1–6. Lipid Science and Technology, 113(10), 1243–1257.
Pugnaloni, L. A., Dickinson, E., Ettelaie, R., Mackie, A. R., & Wilde, P. J. (2004). Yi, J., Ning, J., Zhu, Z., Cui, L., Decker, E. A., & McClements, D. J. (2019). Impact of
Competitive adsorption of proteins and low-molecular-weight surfactants: Computer interfacial composition on co-oxidation of lipids and proteins in oil-in-water
simulation and microscopic imaging. Advances in Colloid and Interface Science, 107 emulsions: Competitive displacement of casein by surfactants. Food Hydrocolloids, 87
(1), 27–49. (FEB), 20–28.
Qiu, C., Zhao, M., Decker, E. A., & McClements, D. J. (2015). Influence of protein type on Young, C. K., & Cunningham, S. (1991). Exploring the partnership of almonds with cereal
oxidation and digestibility of fish oil-in-water emulsions: Gliadin, caseinate, and foods. Cereal Foods World, 36(5), 412–&.
whey protein. Food Chemistry, 175, 249–257. Yu, X. L., & He, Y. (2018). Tea saponins: Effective natural surfactants beneficial for soil
Rampon, V., Genot, C., Riaublanc, A., Anton, M., Axelos, M. A. V., & McClements, D. J. remediation, from preparation to application. RSC Advances, 8(43), 24312–24321.
(2003). Front-face fluorescence spectroscopy study of globular proteins in emulsions: Zhang, Y., Zhang, J., Sheng, W., Wang, S., & Fu, T.-J. (2016). Effects of heat and high-
? Influence of droplet flocculation. Journal of Agricultural and Food Chemistry, 51(9), pressure treatments on the solubility and immunoreactivity of almond proteins. Food
2490–2495. Chemistry, 199, 856–861.
Refsgaard, H. H. F., Tsai, L., & Stadtman, E. (2000). Modifications of proteins by Zhu, Z., Wen, Y., Yi, J., Cao, Y., Liu, F. G., & Mcclements, D. J. (2019). Comparison of
polyunsaturated fatty acid peroxidation products. Proceedings of the National natural and synthetic surfactants at forming and stabilizing nanoemulsions: Tea
Academy of Sciences, 97(2), 611–616. saponin, Quillaja saponin, and Tween 80. Journal of Colloid and Interface Science,
Reichert, C. L., Salminen, H., Leuenberger, B. H., Hinrichs, J. R., & Weiss, J. (2015). 536, 80–87.
Miscibility of quillaja saponins with other Co-surfactants under different pH values. Zhu, Z., Zhao, C., Yi, J., Liu, N., Cao, Y., Decker, E. A., et al. (2018). Impact of interfacial
Journal of Food Science, 80(10-12), 2495–2503. composition on lipid and protein co-oxidation in oil-in-water emulsions containing
Sathe, S. K. (1993). Solubilization, electrophoretic characterization and in vitro mixed emulisifers. Journal of Agricultural & Food Chemistry, 66(17), 4458–4468. acs.
digestibility of almond (prunus-amygdalus) proteins. Journal of Food Biochemistry, 16 jafc.8b00590.
(4), 249–264.

10

You might also like