You are on page 1of 12

Food Chemistry 372 (2022) 131354

Contents lists available at ScienceDirect

Food Chemistry
journal homepage: www.elsevier.com/locate/foodchem

Development and characterization of medium and high internal phase


novel multiple Pickering emulsions stabilized by hordein nanoparticles
Sareh Boostani a, b, Masoud Riazi c, Ali Marefati b, Marilyn Rayner b, Seyed Mohammad
Hashem Hosseini a, *
a
Department of Food Science and Technology, School of Agriculture, Shiraz University, Shiraz, Iran
b
Department of Food Technology, Engineering and Nutrition, Faculty of Engineering, Lund University, Lund, Sweden
c
Enhanced Oil Recovery (EOR) Research Center, School of Chemical and Petroleum Engineering, Shiraz University, Shiraz, Iran

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

Keyword: Medium and high internal phase W1/O/W2 multiple Pickering emulsions (MPEs) were fabricated by physically-
Double Pickering emulsion modified hordein nanoparticles. A triphasic system was developed at dispersed phase volume fraction (Φ) of 0.5
Encapsulation with an overrun value of ~40%. No overrun was detected in high internal phase MPEs (Φ 0.8). Optical and
Hordein
confocal laser scanning microscopy confirmed the formation of MPEs. Monomodal droplet size distribution with
MLS
a mean diameter of 32.90 and 21.48 μm was observed for MPEs at Φ 0.5 and Φ 0.8, respectively. Static multiple
Vitamin B12
light scattering confirmed that creaming was the main mechanism behind the instability of MPEs. Both MPEs
revealed pseudo-plastic behavior and predominant storage modulus (G′ ) over the applied frequency range. The
encapsulation efficiency of vitamin B12 in MPEs was 98.3% and remained relatively constant during 28 d. These
results suggested the excellent potential of hordein nanoparticles as appropriate candidate for designing multi-
structural colloidal systems using plant proteins.

1. Introduction desired site of gastrointestinal tract (GIT) are required (Garti, 1997). The
other potential advantages of multiple emulsions in food systems
The innovative design of multi-purpose functional foods is an include the ability to develop taste contrast properties (e.g., decreasing
ongoing research area in the food industry to fulfill consumer demands the total amount of salt while keeping optimum saltiness perception in
and diversify food products. In this regard, colloidal delivery systems the mouth), and masking undesirable flavors (e.g., bitterness and
(CDSs) are an ingenious option for delivery by design purposes. Among metallic aftertaste) (Muschiolik & Dickinson, 2017). Interestingly, W1/
various colloidal entities, emulsion-based delivery systems are the best O/W2 multiple emulsions have also been suggested for formulating low-
candidates for developing new food products. Recent developments in calorie and low-fat products such as reduced-fat white cheese with
the field of complex colloids have led to a renewed interest in multiple similar rheological behavior and appearance to full-fat counterpart
emulsions systems (McClements, 2018; Muschiolik & Dickinson, 2017). through embedding tiny water droplets into the oil droplets (Lobato-
The most common type of multiple emulsion (also known as W1/O/W2 Calleros, Rodriguez, Sandoval-Castilla, Vernon-Carter, & Alvarez-
double emulsion) is the system in which the oil droplets (O) containing Ramirez, 2006).
internal aqueous phase droplets (W1) are dispersed within an external Polyglycerol polyricinoleate (PGPR) and Tweens, as hydrophobic
water phase (W2). A hydrophobic emulsifier is usually applied to sta­ and hydrophilic emulsifiers, respectively, are the most preferred choices
bilize the primary water-in-oil emulsion (W1/O); while, a hydrophilic for long-term stability of double emulsions (McClements & Jafari,
one should be used to stabilize the W1/O emulsion in the W2 2018). There has been a focus on assembling food-grade stabilizers from
(Muschiolik, 2007; Muschiolik & Dickinson, 2017). These systems have proteins, carbohydrates, and lipids, in order to develop functional foods
the potential for simultaneous delivery of both hydrophobic and hy­ containing all natural ingredients (Jafari, Sedaghat Doost, Nikbakht
drophilic bioactive components. W1/O/W2 systems are applied when Nasrabadi, Boostani, & Van der Meeren, 2020). Pickering emulsions,
controlled release and targeted delivery of bioactive compounds at the also known as surfactant-free emulsions, have adsorbed nano/micro-

* Corresponding author.
E-mail address: hhosseini@shirazu.ac.ir (S.M.H. Hosseini).

https://doi.org/10.1016/j.foodchem.2021.131354
Received 13 May 2021; Received in revised form 29 September 2021; Accepted 4 October 2021
Available online 7 October 2021
0308-8146/© 2021 Elsevier Ltd. All rights reserved.
S. Boostani et al. Food Chemistry 372 (2022) 131354

particles at the oil/water interface (Melle, Lask, & Fuller, 2005). The encapsulated within the W1 of MIP MPE and the ability of delivery
exclusion volume effect and steric hindrances of particles, accumulated system to control the ejection of Vitamin B12 from W1 was studied over
at the interface, provide stabilization against coalescence. Investigating time.
appropriate food-grade Pickering candidates is a live area of research in
Pickering emulsions (Dickinson, 2017; Xiao, Li, & Huang, 2016). 2. Materials and methods
Plant-based ingredients exhibit some advantages over animal-based
ones due to the low risk of infection and contamination, no re­ 2.1. Materials
strictions on cultural eating habits and vegetarian consumers, as well as
more significant economic benefits and sustainability (Samborska et al., Hordein was isolated from whole barley grain (Boostani et al., 2019).
2021). There have been some reports on the successful application of Purified soybean oil and vitamin B12 were purchased from Sigma-
plant-based ingredients as Pickering stabilizers in double emulsions. Aldrich (St. Louis, MO, USA). PGPR was provided from Danisco Co.
Octenyl succinate derivatives of quinoa starch granules (Lin et al., 2019; (Denmark).
Marefati, Sjöö, Timgren, Dejmek, & Rayner, 2015), gliadin (Chen et al.,
2018), and kafirin (Xiao, Lu, & Huang, 2017) particles were used as 2.2. Fabrication of hordein particles
external interface stabilizer of W1/O/W2 multiple Pickering emulsions
(MPEs), and carnauba wax particles as the stabilizer of both interfaces in Hordein (nano)particles were fabricated according to the procedure
O1/W/O2 MPEs (Szumała & Luty, 2016). A mixture of zein nanoparticles previously explained (Boostani et al., 2019) with slight modifications.
and soybean lecithin has recently been used for the stabilization of W1/ Briefly, hordein was hydrated in ethanol–water mixture (65:35) and
O/W2 high internal phase (HIP) Pickering emulsions (Jiang et al., 2021). thermally treated (75 ◦ C for 5 min). One part (20 mL) of the ethanolic
Various mixtures of colloidal microcrystalline cellulose and surfactant dispersion of hordein was added to three parts of deionized water in 2
were utilized to stabilize MPEs (Spyropoulos, Duffus, Smith, & Norton, min under intense stirring, followed by immediate pH adjustment to pH
2019). The adsorption of soluble and insoluble complexes of whey 3.0. Afterward, the protein aggregates were subjected to evaporation in
protein concentrate/gum Arabic at both interfaces of O1/W/O2 MPEs a rotary evaporator to remove the extra solvents and to reach a final
was reported by Estrada-Fernández et al. (2018). Recently, a concentration of 2% w/v. The dispersion was then sonicated for 2 min
maltodextrin-in-gelatin-in-maltodextrin W/W/W Pickering emulsion by a horn sonicator (model UP200st equipped with sonotrode S26d7,
was fabricated by cooling down of simple gelatin-in-maltodextrin W/W Hielscher Ultrasound Technology, Germany) in the continuous mode.
emulsion with the aid of mucin glycoprotein particles (Beldengrün et al., The frequency, amplitude, and nominal power were 24 kHz, 75%, and
2020). 200 W, respectively. Temperature control was performed using an ice
Among plant-based materials, prolamins (as a group of plant storage tank during sonication.
proteins mainly found in cereals) can be exploited as appropriate can­ After appropriate dilution with deionized water pre-adjusted to pH 3,
didates for the fabrication of micro- and nano-particles. The limited the particle size and zeta-potential of hordein dispersion were deter­
solubility of these proteins in aqueous and oil systems as well as their mined using a Zetasizer Nano (Malvern Instruments Ltd., UK) at 25 ◦ C.
highly hydrophobic nature make prolamins suitable for Pickering sta­ The three-phase contact angle (θwo) of hordein nanoparticles was also
bilization (Samborska et al., 2021). β-Hordein is the main fraction of determined (Boostani et al., 2019). Briefly, lyophilized powder was
barley prolamins. It shows high surface hydrophobicity and thus ag­ pressed into small pellets by a laboratory scale instrument and analyzed
gregation behavior, which result in adsorbing to oil/water and air/water by a drop shape analyzer (DSA 100, KRÜSS GmbH, Hamburg, Germany).
interfaces and forming viscoelastic films around the dispersed phases The surface morphology of protein particles was investigated by a
(Boostani et al., 2020). Recently, we reported that hordein particles are scanning electron microscope (TESCAN-Vega 3, TESCAN Co., Czech
able to develop emulsion-foam (triphasic) Pickering structures with Republic) at an accelerating voltage of 10 kV and magnification of 35.0
higher stability against coalescence and aqueous phase separation kx.
compared to zein- and secalin-based particles (Boostani et al., 2019).
Such a property can be applied for mimicking the desirability and sen­ 2.3. Fabrication of multiple Pickering emulsions
sory properties of whipped cream-like products. In another work, the
ability of hordein particles to stabilize HIP Pickering emulsions was also MPEs were prepared in a total volume of 6 mL in two steps. Firstly,
confirmed (Boostani et al., 2020). Fabricating stable HIP emulsion, the primary emulsions (W1/O) were prepared by mixing the internal
removing water, and then shearing is a common procedure applied for aqueous phase (30% v/v) and the oil phase (70% v/v). The W1 phase
converting low viscosity liquid vegetable oils into solid-like structured consisted of deionized water at pH 3. The oil phase was prepared by
oils without the need to hydrogenation or mixing with saturated fats mixing 5% w/v PGPR in soybean oil under stirring at 65 ◦ C for 20 min.
(Zhou et al., 2018). Structured oils can minimize inappropriate nutri­ Homogenization was performed at 24,000 rpm for 10 min using a high-
tional effects of formulations containing saturated fats, while keeping speed homogenizer equipped with a 6 mm dispersing tool (Ystral Ho­
the original texture and even developing innovative organoleptic mogenizer X-10/25, Germany). The MPEs were prepared by the emul­
perception. Additionally, forming an interfacial layer of biopolymer sification of primary emulsion in the external aqueous phase (W2 pre-
particles around the oil droplets might lead to designing anti-obesity adjusted to pH 3) containing 20 mg of hordein nanoparticle per mL of
formulations via retarding the digestion of triacylglycerols (Xiao et al., the oil phase. For the preparation of medium internal phase (MIP) MPEs
2017). with a dispersed volume fraction (φ) of 0.5, equal volumes (i.e., 3 mL) of
Vitamins are the essential components of daily diet. The majority of primary emulsion (W1/O) and external aqueous phase (W2) were ho­
these microelements cannot be synthesized within the human body. W1/ mogenized at 16000 rpm for 2 min. High internal phase (HIP) MPEs (φ
O/W2 MPEs have the advantages of both multiple and Pickering emul­ 0.8) were prepared by the dropwise addition of W1/O primary emulsion
sions. They can be applied to encapsulate hydrophilic bioactive com­ (4.8 mL) to W2 (1.2 mL) under homogenization at 16000 rpm for 5 min.
pounds in the W1, while providing physicochemical stability during It is known that the shearing conditions in the second step should be
processing and storage, as well as controlling the release in GIT. To the mild enough to avoid destabilization of the internal droplets but at the
best of our knowledge, the design of MPEs stabilized by hordein-based same time being able to create a secondary emulsion (Matos et al.,
particles has not yet been studied. Therefore, the aim of this study was 2013).
the fabrication and characterization of food-grade medium internal
phase (MIP) and high internal phase (HIP) MPEs based on hordein
particles. Vitamin B12, as a model hydrophilic compound, was

2
S. Boostani et al. Food Chemistry 372 (2022) 131354

2.4. Characterization of multiple Pickering emulsions encapsulation properties were assessed similar to the procedure
described by Matos et al. (2015) with some modifications. The encap­
The size of dispersed droplets was measured by laser light scattering sulation efficiency (EE) was defined as the fraction of vitamin B12
technique using a Malvern Mastersizer 2000 (Malvern Instruments Ltd., entrapped in W1 during the emulsification process. In this regard, freshly
UK). Before the measurement, the MPEs were diluted with Milli-Q water fabricated W1/O/W2 MPEs loaded with vitamin B12 were centrifuged at
pre-adjusted to pH 3.0. The emulsion samples were added to dispersion 4000 rpm for 10 min and non-encapsulated vitamins were separated
unit under stirring at 2000 rpm until an obscuration of 10%–20% was using a 0.2 µm PVDF syringe filter. The absorbance of filtrate was
recorded. The refractive index and absorption of soybean oil droplets measured by a UV–visible spectrophotometer (Carry 50, USA) at 361
were 1.46 and 0.001, respectively. Moreover, the refractive index of nm. The vitamin B12 concentration was determined from a standard
1.33 was considered for the dispersant and Mie theory was used to curve (Fig. 6A) prepared by vitamin B12 standard solutions. The EE (%)
compute the results. An optical microscope (Olympus BX50, Japan) was then calculated using equation (3).
equipped with a digital camera (DFK 41AF02, Germany) was used to
Mt − Mp
take optical micrographs. For further analysis of droplet size, the optical EE(%) = × 100 (3)
Mt
micrographs of at least 100 random droplets were analyzed using
ImageJ software (National Institutes of Health, USA). Confocal laser where, Mt is the total amount of vitamin B12 (mg) added to MPE; and Mp
scanning microscopy (CLSM, Nikon A1 + confocal microscope, Nikon is the amount of non-encapsulated (free) vitamin B12 (mg) determined in
Instruments, USA) was used to capture CLSM images of MPEs. Prior to W2. The centrifugal force did not adversely affect the microstructure of
emulsification, the aqueous and oil phases were stained with Nile blue A MPEs.
and Nile red, respectively. The samples (10 μL) were then cast on a single The encapsulation stability (ES%) of vitamin B12 in MPEs was
concave glass slide and monitored with a 60x objective. CLSM was monitored during storage at 4 ◦ C and dark conditions by measuring the
performed using two laser excitation sources (488 and 633 nm) and two EE at given time points (1, 7, 14, 21, and 28 d after preparation).
acquisition channels for Nile red and Nile blue A, respectively, and
combined images under two channels were taken through NIS-elements
software (version 4.6). Owing to the ability of hordein nanoparticles in 2.8. Statistical analysis
the formation of triphasic systems (Boostani et al., 2019), the possible
formation of emulsion-foam structures was assessed by measuring the A completely randomized design was used in this study. Mean values
overrun (i.e., the increase in volume) immediately after the preparation and standard deviations were reported. Tukey’s multiple range post hoc
of emulsions using equation (1): tests were applied by SPSS software (ver. 22, IBM Corp. USA) at a sig­
nificance level of 0.05. All experiments were done in triplicate.
V1 − V0
Overrun(%) = × 100 (1)
V0
3. Results and discussion
where, V1 is the final volume of MPE; and V0 is the total volume of oil
and water phases in non-emulsified form. 3.1. Formation and visual inspection of multiple Pickering emulsions

2.5. Static multiple light scattering (MLS) Spherical-shaped modified hordein nanoparticles with an intensity-
weighted average diameter of approximately 200 nm, zeta potential of
Static MLS was applied to monitor the stability profile of MPEs using about + 30 mV, and three phase contact angle of 64.1◦ were used to
a Turbiscan Lab Expert (Formulation Co., Toulouse, France), similar to stabilize the outer interface of MPEs (Fig. 1A-D). A relatively thick layer
the method described by Matos, Gutiérrez, Iglesias, Coca, and Pazos of hordein nanoparticles around the oil droplets as well as protein ag­
(2015). This instrument works based on the principle of emitting a light gregates in the continuous phase were observed in simple O/W emul­
beam through the whole length of a cylindrical glass vial containing 4.0 sions, confirming the surface activity of hordein nanoparticles and hence
mL of emulsion samples. The transmitted (TS) and backscattered (BS) their ability to adsorb to the O/W interfaces (Fig. 1E-F). This observation
light intensity was monitored as a function of cell height and time (every in simple emulsion led to a conclusion that these nanoparticles are also
30 min for 3 h and after 10 days of storage) at 25 ◦ C. The emulsion index able to form a thick interface around the dispersed multiple W1/O
(EI) was estimated from the BS graphs by dividing the height of the droplets in MPEs.
cream phase by the total height of the sample after 1 h and 10 days. Our previous works have shown that the particle size of hordein
Turbiscan stability index (TSI) was evaluated according to a scani-to- nanoparticles is decreased after thermal and subsequent sonication
scani-1 difference over total sample height (H) using equation (2): treatments (Boostani et al., 2019). The sonication treatment was more
effective than the heat treatment in reducing the hydrophobicity of
∑ ∑ |scani (h) − scani− 1 (h)| particles. This was related to the adaptability and reorganization of
TSI = h
(2)
H sonicated aggregates due to the disruption of noncovalent interactions
and formation of smaller aggregates along with assuming new equilib­
2.6. Rheological properties rium structures via exposing more polar groups buried within the larger
aggregates. In the next study, changes in the aggregation behavior of
A stress-controlled Kinexus rheometer (Malvern Panalytical Ltd, hordein particles under the influence of pH and ionic strength was
UK), equipped with a cup and bob geometry (C14), was used to evaluate investigated (Boostani et al., 2020). The lower intensity-weighted mean
the rheological properties of freshly prepared MPEs at 25 ◦ C. Frequency diameter and narrower distribution of hordein particles at pH 3 and pH 4
sweep test was performed at a range of 0.1–10 Hz within the linear were related to the strong electrostatic repulsion among particles.
viscoelastic region (LVR) at 1 Pa. The storage (G′ ) and loss (G′′ ) moduli Nevertheless, an increase in the particle size was observed at higher pH
were then calculated. The steady shear properties of MPEs were also values close to isoelectric point, and when the ionic strength exceeded a
monitored at a shear rate range of 0.1–100 s− 1. certain level (e.g., 0.05 M at pH 3). Therefore, in this study, physically
modified hordein nanoparticles were fabricated in the optimum condi­
2.7. Encapsulation efficiency and release properties tions and investigated for their ability to develop MPEs.
The visual appearance of MPEs at various time intervals is shown in
In order to evaluate the encapsulation properties of MIP MPEs, Fig. 1G. Interestingly, an overrun value of 41% was observed in MPEs
vitamin B12 (2.5 mg/mL of W1) was used as a hydrophilic model. The prepared at Φ 0.5 (Fig. 1H), indicating the formation of a complex

3
S. Boostani et al. Food Chemistry 372 (2022) 131354

Fig. 1. (A): Intensity-weighted particle size


distribution of physically modified hordein
nanoparticles determined by DLS; (B): Three-
phase contact angle of hordein nanoparticles;
(C): Photographs of isolated hordein and
lyophilized hordein nanoparticles; (D): SEM
micrograph of hordein nanoparticles; (E and
F): CLSM micrographs of hordein nano­
particles adsorption at the interfacial layer of
simple O/W emulsions at two different
excitation wavelength; (G): Visual observa­
tion of hordein nanoparticles-stabilized
MPEs prepared at Φ 0.5 (top) and Φ 0.8
(bottom) over time (from left to right: freshly
prepared, after 1 h, 3 h, 1 d and 10 d,
respectively); (H): Visual observation of W2
and W1/O primary emulsion prior to ho­
mogenization (left) and after homogeniza­
tion and formation of MPE at Φ 0.5 (right)
which indicates obvious overrun; (I): photo­
graphs of MPEs at Φ 0.5 (top) and Φ 0.8
(bottom).

structure ((W1/O)/Air-in-W2) after emulsification arising from the with the aid of a thickening agent in the external phase due to the
hordein particles ability to adsorb at both air–water and oil–water in­ emulsion inversion phenomenon (Matos, Guti, Martínez-rey, Iglesias, &
terfaces. This finding was consistent with our previous finding (Boostani Pazos, 2018).
et al., 2019), in which hordein particles resulted in the formation of a In our previous study, the O/W simple Pickering emulsion (at Φ 0.8)
triphasic oil/air-in-water simple Pickering emulsion at Φ 0.5. In the case stabilized by hordein nanoparticles was self-supporting after the vial
of hordein-stabilized MPEs at Φ 0.5, a separated aqueous phase at the inversion and highly stable to coalescence and aqueous phase separa­
bottom was observed after about 1 h and gradually increased over 10 tion. This property was associated with the robust particle-based inter­
days of storage (Fig. 1G and Fig. 3D). The average droplet size of this facial characteristics of hordein around the oil droplets. The three-
sample remained relatively constant during the storage period (Fig. 2A), dimensional solid-like network, resulted from the close proximity and
indicating the relative stability of the concentrated emulsion. MPEs tight packing density of oil droplets, led to the formation of high internal
prepared at Φ 0.8 had no overrun after emulsification and showed phase Pickering emulsions (HIP PEs) stable against creaming and coa­
higher creaming stability than those prepared at Φ 0.5 during storage lescence (Boostani et al., 2020). Moreover, in HIP PE systems the steric
(Fig. 1G, and Fig. 3D), which could be related to the higher viscosity and hindrance and electrostatic repulsion among droplets induce higher
lower droplet size (discussed later). The absence of overrun was also viscosity (Zhu, Gao, Liu, Zou, & McClements, 2019). However, in this
attributed to the increase in the viscosity and elastic behavior of MPEs at study HIP MPEs did not exhibit such exact solid-like character (Fig. 1I),
higher droplet concentrations (section 3.5). Besides, under certain cir­ which might be due to the differences in droplet size distribution (sec­
cumstances, the oil phase can destabilize the foam systems via spreading tion 3.2) and viscosity (section 3.5) of HIP PEs and HIP MPEs. Basically,
at the air/water interface and triggering the foam rupture (Boostani in emulsion systems, the droplets are caged by other ones at Φ 0.58
et al., 2020). The potential of hordein nanoparticles for the stabilization which causes a large increase in viscosity (known as jamming). At higher
of HIP MPEs is notable as other potent Pickering candidates, such as OSA Φ (e.g., 0.64), the droplets may adopt a random close-packed arrange­
quinoa starch and kafirin particles, have shown intense oiling off and ment, and at Φ 0.74, they may adopt a face-centered cubic arrangement,
droplet collapse at (W1/O):W2 ratio of 7:3 (Lin et al., 2019; Xiao et al., where they become entirely close-packed and have a solid-like behavior.
2017). Additionally, it has been reported that the surfactant molecules Additionally, the viscosity of an emulsion may increase by decreasing
are unable to develop concentrated W1/O/W2 emulsion systems even the droplet size, particularly in the case of more concentrated emulsions

4
S. Boostani et al. Food Chemistry 372 (2022) 131354

Fig. 2. Changes in volume-weighted droplet size distribution of MPEs stabilized by hordein nanoparticles at Φ 0.5 (A) and Φ 0.8 (B) during storage; Day 1 indicates
freshly prepared samples. (C): Number-weighted droplet size distribution of freshly prepared MPEs stabilized by hordein nanoparticles at Φ 0.5 and Φ 0.8 determined
by ImageJ software.

(Zhu et al., 2019). of droplets analyzed by ImageJ software was significantly lower than
Free oil formation was not observed in MPEs prepared at both Φ that analyzed by laser diffraction technique and hence less representa­
values during the storage period (Fig. 1G). Chen et al. (2018) and Jiang tive of the system. It should be noted that the scattering signals caused
et al. (2021) similarly reported high stability of HIP MPEs prepared by by the internal aqueous droplets and the external oil droplets might lead
gliadin and zein nanoparticles/lecithin, respectively. Such behavior can to some unavoidable experimental errors. However, taking into account
be related to the formation of prolamin-based rigid interfaces, which the relatively normal distribution of dispersed droplets, appropriate
effectively prevents the dispersed droplets from coalescence and Ost­ correlation was observed between the results of volume- and number-
wald ripening (Chen et al., 2018). Double emulsions are prone to various weighted distributions. The droplet size distribution patterns of both
instability mechanisms including osmotic and chemical-driven coales­ systems were monomodal similar to those previously reported for MPEs
cence, diffusion of internal aqueous droplets into the external water stabilized by plant-based Pickering particles like OSA-modified quinoa
phase, transport of hydrophilic emulsifier into the internal water drop­ starch granules (Lin et al., 2019; Marefati et al., 2015; Matos, Timgren,
lets, flocculation, expulsion, gravitational separation, and Ostwald Sjöö, Dejmek, & Rayner, 2013). Fresh samples prepared at Φ 0.5 had
ripening (Liu et al., 2020; Muschiolik & Dickinson, 2017; Benichou, slightly lower span values than those prepared at Φ 0.8. However, the
Aserin, & Garti, 2007). Compared to small emulsifier molecules, increase in the distribution width of MPEs at Φ 0.5 was higher than that
biopolymer-based particles do not readily pass through the oil droplets of MPEs at Φ 0.8 indicating lower stability. Creaming was the main
and accordingly the coalescence of the internal water droplets and the instability mechanism of such system. These results are in accordance
external water phase is reduced (Schuch, Helfenritter, Funck, & with those reported by Matos et al. (2015), and Fechner, Knoth, Scherze,
Schuchmann, 2015). Marefati et al. (2015) reported that Pickering and Muschiolik (2007). The sample prepared at Φ 0.8 had higher sta­
candidates enhance the stability of W1/O/W2 emulsions against coa­ bility over time.
lescence and Ostwald ripening greater than the non-aggregated indi­ The high resistance of Pickering emulsions against destabilization
vidual macromolecules. phenomena is associated with the larger size of Pickering particles than
the size of surfactant molecules. In these systems, the free energy of
detachment (ΔGD) is much greater than the thermal energy and, thus the
3.2. Size characteristics of multiple Pickering emulsions
particles are irreversibly adsorbed (Dickinson, 2010). In such condi­
tions, the system approaches a steady-state (Estrada-Fernández et al.,
Fig. 2A and 2B show the changes in the volume-weighted droplet size
2018). A thick layer of particles can sterically hinder the close contact of
distribution of MPEs obtained by laser diffraction technique over time.
droplets when the particles have appropriate contact angle (Section 3.1,
The number-weighted droplet size distribution of freshly prepared
Fig. 1A-F). This phenomenon makes the fabrication of highly stable
emulsions determined by ImageJ software is also shown in Fig. 2C. Both
concentrated emulsions possible even in the presence of coarse droplets
the volume mean (D4,3) and number (arithmetic) mean (D1,0) of MPEs
(Marefati & Rayner, 2020; Marefati et al., 2015). The ΔGD (i.e., the
prepared at Φ 0.5 were relatively larger than those of MPEs prepared at
minimum energy needed to detach a spherical particle from the inter­
Φ 0.8. In freshly prepared samples, the volume mean values of MPE
face) is a function of particle size, contact angle, and interfacial tension.
droplets prepared at Φ 0.5 and 0.8 were 32.9 ± 1.8 and 21.5 ± 1.9 μm,
It rapidly increases with increasing the particle size (Dickinson, 2017).
respectively. Considering the equations (or numerator and denomina­
Similar to other prolamin-based Pickering candidates such as gliadin
tor) utilized in calculating D4,3, and D1,0, the variations in the volume
(Hu et al., 2016), kafirin (Xiao et al., 2017) and zein/lecithin (Jiang
and number means might be resulted from the higher contribution of
et al., 2021), hordein nanoparticles were irreversibly anchored onto the
larger droplets than smaller ones to D4,3 values. Moreover, the number

5
S. Boostani et al. Food Chemistry 372 (2022) 131354

Fig. 3. Turbiscan profiles of MPEs prepared at Ф 0.5 (A) and Ф 0.8 (B) over 10 d; (C): Turbiscan profile of PGPR-stabilized W1/O primary emulsion during 30 d; (D):
TSI values of MPEs over time. Different colors indicate different time intervals.

6
S. Boostani et al. Food Chemistry 372 (2022) 131354

Fig. 3. (continued).

interface and provided stability against the coalescence of oil droplets. systems, the BS profile is more representative than the TS profile for the
The differences in such emulsion systems, reported in literature, is stability analysis owing to the ability of oil droplets to scatter 880-nm
rooted in the experimental conditions, as well as the characteristics of near-infrared light beam. Since the BS intensity of W1/O primary
the protein aggregates. Hoffmann, and Reger (2014) stated that the emulsion remained constant over the whole monitoring time (Fig. 3C), it
unique properties of the emulsions are a reflection of the properties of can be concluded that PGPR was able to stabilize water droplets within
the proteins in the bulk solution. the continuous oil phase. Based on the BS profile, Matos et al. (2015)
The emulsifying ability of particles depends not only on the particle similarly reported high stability of W1/O emulsion stabilized by PGPR as
size but also on the contact angle and interfacial tension (Dickinson, well as no significant changes in the water droplet size.
2017). The three-phase contact angle is an important parameter which The BS intensity of both MPEs at time zero exhibited a flat line along
indicates the wettability of particles by continuous phase and/or oil the measuring cell (Fig. 3A and B). A decrease in the BS intensity of
phase (Jiang et al., 2021). Pickering particles remain in the continuous MPEs at Φ 0.5 was observed at the height range of 10–19 mm during the
phase when it contributes more than the oil phase to the wetting of measuring time (Fig. 3A), which implied creaming and clarification
particles, and thus better steric stabilization is achieved (Dickinson, instability mechanisms. The TS intensity was kept constant along the
2017). The contact angle of hordein nanoparticle was about 64.1◦ measuring cell during the monitoring period. The movement of W1/O
indicating the preferential wetting by water. Chen et al. (2018) reported droplets toward the top of the measuring cell results in the clarification
that the water–air-gliadin contact angle was about 60◦ and gliadin process. This causes an increase in the BS intensity (also known as
nanoparticles were able to stabilize W1/O/W2 emulsion gels against creaming phenomena) (Matos et al., 2018). Zero BS signal was not
phase separation. Practically, conventional double emulsions stabilized detected at the top of the BS graph, which was interpreted as the absence
by surfactants have large and polydisperse droplets that are unstable of free oil in the creamed layer (Fig. 3A). As mentioned previously, the
against coalescence, flocculation, and creaming phenomena. Therefore, size of dispersed droplets was larger than the wavelength of the light
the increase in the droplet size is common during the storage. Addi­ source; nevertheless, the BS variation in the middle part of the
tionally, the span values are increased with time (Giroux et al., 2013; measuring cell probably indicated the growth in the emulsion droplet
Benichou et al., 2007). The osmotic pressure difference between two size. The existence of air bubbles could also have some influences on the
aqueous phases is also another mechanism behind the instability of results. A slight reduction in the BS intensity accompanied by the fluc­
double emulsions and is responsible for the changes in the size of tuations in the BS line was possibly due to the foam destruction in the
dispersed droplets owing to the diffusion of water molecules from the case of MPEs at Φ 0.5 during the test period (Fig. 3A), even though the
external phase to the internal phase and vice versa. Therefore, control­ collapse of bubbles was not significant. A decrease in the BS intensity
ling the osmotic balance by appropriate salt solutions can maintain the along the whole measuring cell was attributed to the increase in the
stability of W1/O/W2 emulsions over time (Matos et al., 2018). In the droplet size due to coalescence (Matos et al., 2015).
current study, the ionic strength (I) of both aqueous phases were not MPEs at Φ 0.8 revealed remarkable stability since the BS signal was
adjusted as the values above 0.05 M caused large aggregation of hordein almost steady over time (Fig. 3B). After 1 h, the emulsion index (EI) of
particles mainly due to the charge screening effects (Boostani et al., MPEs prepared at Φ 0.8, and 0.5 was 100%, and 83.9%, and after 10
2020). The consequence of small protein particles’ affinity to either d reached to 96.1%, and 76.7%, respectively. TSI is inversely propor­
adsorb at the interface or form large aggregates with other particles tional to the stability of emulsions. After 1 h, this value was 0.90 and
determines the efficacy of particulate entities in the Pickering stabili­ 11.86 in the emulsions prepared at Φ 0.8 and 0.5 and reached to 6.56
zation (Fig. 1E-F). and 28.13 after 10 d, respectively. The TSI values further confirmed the
higher stability of MPEs prepared at Φ 0.8 (Fig. 3D). These results are in
3.3. Turbiscan stability profile accordance with the results of droplet size (section 3.2) and rheological
properties (section 3.5). Similar to our findings, Matos et al. (2018) re­
Monitoring the profiles of backscattering (BS) and transmission (TS) ported that by increasing the ratio of W1/O to W2, the creaming rate was
was performed for further studying the stability of MPEs (Fig. 3). Since decreased. This observation can be explained by the larger volume
the average size of colloidal particles was larger than the wavelength of occupied by the W1/O droplets and hence having less space for free
the incident light, a turbid appearance was observed both in the movement within the network (i.e., higher stability). Matos et al. (2018)
creamed layer and at the bottom of the measuring cell (Fig. 1G). In such also reported that the clarification occurred in a higher rate in double

7
S. Boostani et al. Food Chemistry 372 (2022) 131354

emulsions with larger droplet sizes and higher BS variations (Matos micrographs also confirmed that the water droplets were dispersed
et al., 2018). The BS intensity of MPEs prepared at Φ 0.5 was lower than within the oil (red) droplets. In MPE systems stabilized by zein nano­
that of emulsions prepared at Φ 0.8 due to the less compact structure particles and lecithin, the W1 droplets tend to be located near the O/W
resulting from the presence of air bubbles (Fig. 1H). MPEs at Φ 0.8 interface, which was attributed to the density difference and hydro­
showed a slight decrease of BS at the top of the measuring cell after 10 phobic interactions (Jiang et al., 2021). Chen et al. (2018) reported a
d of storage, which indicated minor creaming. Instability mechanisms high degree of oil droplets flocculation which was related to a decrease
derived from bridging or depletion flocculation are not typically a in the electrical charge on gliadin nanoparticles near their isoelectric
matter of emulsions with high Φ value, since the droplets are packed point.
tightly together. A highly viscous or gelled continuous phase also pre­
vents the flocculated droplets from moving (Zhou et al., 2018). 3.5. Rheological properties

3.4. Microstructure of multiple Pickering emulsions 3.5.1. Apparent viscosity


Both samples exhibited shear thinning behavior (Fig. 5A), which
The formation of MPEs was also confirmed by microscopic evalua­ indicates the disruption of network structure or deflocculation of
tion. In both samples, relatively large oil droplets containing numerous droplets by increasing the shear rate (Boostani et al., 2020; Xi, Liu,
tiny water droplets were present (Fig. 4A). The optical micrographs also McClements, & Zou, 2019). Higher droplet concentration increases the
revealed that the adjacent droplets were packed with each other to form friction forces within the network. Therefore, the sample prepared at Φ
a network. The higher packing density of W1/O droplets in MPEs pre­ 0.8 had higher apparent viscosity (Fig. 5A). In the emulsions with high
pared at Φ 0.8 led to the formation of irregular polyhedral morphology oil content, there is an increase of droplet flocculation owing to the
(Fig. 4A) which increased the viscoelastic properties and developed higher volume fraction occupied by dispersed droplets and possibly a
structured emulsion system (Fig. 5B). The visual observation of these higher sensitivity to increasing shear rate (Xi et al., 2019).
emulsions are in good agreement with the experimental data reported in
Fig. 2 and Fig. 3. Confocal laser scanning microscopy (CLSM) was also 3.5.2. Frequency sweep test
performed (Fig. 4B). Due to the complexity of the system, the samples The viscoelastic properties of MPEs prepared at Φ 0.5 and 0.8 are
were highly diluted with W2 to capture separate droplets and for better illustrated in Fig. 5B. In both samples, the elastic character was domi­
illustration, CLSM 3-D images were taken at different depths. The CLSM nant (G′ > G′′ ) over the studied frequency range. A minor dependency of

Fig. 4. (A): Optical micrographs (magnification 100 x; scale bar = 10 μm) of MPEs at different Ф values; (B): 3D CLSM images of scanned droplets (magnification 60
x) at three layers from bottom to top at different Ф values.

8
S. Boostani et al. Food Chemistry 372 (2022) 131354

Fig. 5. (A): Apparent viscosity vs. shear rate plots of MPEs at Ф 0.5 and Ф 0.8; (B): Frequency sweep test of MPEs at a frequency range of 0.1–10 Hz.

both moduli on the applied frequency (0.1–10 Hz) was also noted. In significant (p > 0.05), which demonstrated appropriate encapsulation
concentrated emulsions, the droplets are tightly packed together and stability (ES) or very low ejection of vitamin B12 from W1 to W2 during
become trapped in “cages” by their neighbors and thus elastic-like fea­ the time period. The decrease in the EE vs. time was well fitted with a
tures are observed (Zhu et al., 2019). As shown in Fig. 5B, the storage logarithmic regression (Fig. 6C). The measurement of EE% over a
modulus of MPEs at Φ 0.8 was significantly higher than that of emul­ defined time is an indicator of losing internal water droplets by coa­
sions prepared at Φ 0.5 which can be related to the higher degree of lescence (Schuch et al., 2015). From ES point of view, a double emulsion
inter-droplet interactions of hordein-stabilized droplets. Additionally, system can be considered stable if the initial EE is around 95% and
sharing hordein nanoparticles at the interface might result in the floc­ reaches to 70–80% after some weeks of storage (Xiao et al., 2017). Type
culation of tightly packed oil droplets and increase the viscoelastic of bioactive component, encapsulant matrix, and surrounding medium
properties (Boostani et al., 2020). are the key parameters that affect the release of bioactive compounds
from the encapsulating materials. Owing to the molecular dimension as
well as the concentration gradient between W1 and W2, vitamin B12 can
3.6. Encapsulation properties rapidly diffuse from W1, pass through the oil medium and ultimately
reach to W2 during homogenization or storage. The main challenges
The standard curve of vitamin B12 is shown in Fig. 6A and the associated with common W1/O/W2 double emulsions are controlling the
encapsulation efficiency (EE) of vitamin B12 in W1 during 28 d is shown diffusion of water and emulsifier molecules across different phases
in Fig. 6C. In freshly prepared samples, the EE was 98.3% and reached to during preparation and storage. Hydrophilic surfactants tend to migrate
97.1% at the end of storage. Changes in the EE over time were not

9
S. Boostani et al. Food Chemistry 372 (2022) 131354

Fig. 6. (A): Standard curve of vitamin B12; (B): Photographs of multiple Pickering emulsion loaded with vitamin B12 (from left to right: W1/O primary emulsion (top)
and W2 phase (bottom) prior to the second emulsification step; Vitamin B12-loaded MPE; and centrifuged MPE for calculating encapsulation efficiency (EE)); (C):
Logarithmic regression of EE over time.

from the external water phase to the oil medium and aggregate as hordein nanoparticles revealed appropriate encapsulation properties in
reverse micelles (Bou et al., 2014). Such challenges may be overcome terms of both EE and ES (Fig. 6). The interfacial film, created by the
using macromolecules instead of surfactants; although, biopolymer adsorbed particles, developed a physical barrier. This mechanism is not
stabilized double emulsions may themselves undergo some instability principally different from that provided by the individual protein and
mechanisms. As reported by Fechner et al. (2007), the initial release of polysaccharide molecules (Estrada-Fernández et al., 2018). However,
vitamin B12 from W1 to W2 in double emulsions was remarkably considering the theory of Pickering stabilization, the free energy of
decreased by changing the outer interface stabilizer from sodium spontaneous desorption of a solid particle is several orders larger than
caseinate to sodium caseinate-dextran conjugate (Fechner et al., 2007). that of small surfactant molecules; therefore, the particles are assumed
The EE% of vitamin B12 in freshly prepared W1/O/W2 emulsion stabi­ to be irreversibly anchored at the interface and stable against coales­
lized by sodium caseinate was about 16.6% and increased to about cence (Dickinson, 2017; Dickinson, 2010). Additionally, it seems that
27.4% when sodium caseinate-maltodextrin conjugates were used as an the stability of W1 can be directly associated with the EE (Dias et al.,
emulsifier (O’Regan & Mulvihill, 2010). A total reduction in the EE over 2018). Similar results regarding the high EE and ES of MPEs stabilized
7 d (ranging from 39.2% to 27.7%) was observed in all types of emul­ by plant-based particles were reported by other researchers for various
sions. The ES over 7 d was better in the presence of conjugates (O’Regan cargoes. The EE of carmine was 98.5% and 90% in freshly prepared and
& Mulvihill, 2010). Giroux et al. (2013) reported high EE of vitamin B12 21-d old MPEs stabilized by OSA-modified starch granules, respectively
(98.4%) in W1/O/W2 emulsions prepared in skim milk or sodium (Matos et al., 2013). The EE values of 96.6% and 98.5% for liquid and
caseinate dispersion (as W2). In conclusion, the individual biopolymer solid shea nut oil, respectively, and the ES value of 97% were reported
molecules have advantages over small surfactant molecules; however, by Marefati et al. (2015) in MPEs stabilized by OSA-modified quinoa
some instability phenomena such as the leakage of W1 into W2 may still starch particles. The EE of anthocyanin in freshly prepared MPEs sta­
occur. Pickering particles-stabilized multiple emulsions containing bilized by kafirin particles was 85.1% and reached to 75.9% after 2 w
water-soluble markers generally show better EE and ES compared to (Xiao et al., 2017). HIP Pickering double emulsion/gels stabilized by
macromolecules- or surfactant-stabilized systems. MPEs stabilized by gliadin particles showed an EE of 65.5% for epigallocatechin gallate and

10
S. Boostani et al. Food Chemistry 372 (2022) 131354

97.2% for quercetin (Chen et al., 2018). The lower stability of some Chen, X., McClements, D. J., Wang, J., Zou, L., Deng, S., Liu, W., … Liu, C. (2018).
Coencapsulation of (-)-Epigallocatechin-3-gallate and Quercetin in particle-
systems might be associated with the mass transport, driven by the
stabilized W/O/W emulsion gels: Controlled release and bioaccessibility. Journal of
gradient of osmotic pressure and chemical potential difference between Agricultural and Food Chemistry, 66(14), 3691–3699.
W1 and W2 (Muschiolik & Dickinson, 2017; Dickinson, 2017), which Dias, S. V. E., Züge, L. C. B., Santos, A. F., & Scheer, A.de P. (2018). Effect of surfactants
increases the oil droplet size owing to the diffusion of W1 to W2. As and gelatin on the stability, rheology, and encapsulation efficiency of W1/O/W2
multiple emulsions containing avocado oil. Journal of Food Process Engineering, 41,
reported by Xiao et al. (2017) in MPEs stabilized by kafirin particles, the e12684.
growth in oil droplets size was concomitant with a decrease in the Dickinson, Eric (2010). Food emulsions and foams: Stabilization by particles. Current
packing density of nanoparticles at the interface which led to the Opinion in Colloid and Interface Science, 15(1-2), 40–49.
Dickinson, Eric (2017). Biopolymer-based particles as stabilizing agents for emulsions
collapse of emulsion droplets. In some systems, it is not crucial to pre­ and foams. Food Hydrocolloids, 68, 219–231.
cisely balance the osmotic pressure difference to obtain high stability Estrada-Fernández, A. G., Román-Guerrero, A., Jiménez-Alvarado, R., Lobato-
(Matos et al., 2013). Calleros, C., Alvarez-Ramirez, J., & Vernon-Carter, E. J. (2018). Stabilization of oil-
in-water-in-oil (O1/W/O2) Pickering double emulsions by soluble and insoluble
whey protein concentrate-gum Arabic complexes used as inner and outer interfaces.
4. Conclusion Journal of Food Engineering, 221, 35–44.
Fechner, A., Knoth, A., Scherze, I., & Muschiolik, G. (2007). Stability and release
properties of double-emulsions stabilised by caseinate–dextran conjugates. Food
In this study, the fabrication of concentrated (Φ 0.5) and high in­ Hydrocolloids, 21(5-6), 943–952.
ternal phase (Φ 0.8) MPEs stabilized by physically-modified hordein Garti, Nissim (1997). Double emulsions—scope, limitations and new achievements.
nanoparticles was investigated. The microstructure, and physical sta­ Colloids and Surfaces A: Physicochemical and Engineering Aspects, 123-124, 233–246.
Giroux, Hélène J., Constantineau, Stéphane, Fustier, Patrick, Champagne, Claude P., St-
bility of emulsions were affected by the droplet size distribution, dis­ Gelais, Daniel, Lacroix, Monique, & Britten, Michel (2013). Cheese fortification using
tribution width, and rheology of the systems. Increasing the amount of water-in-oil-in-water double emulsions as carrier for water soluble nutrients.
dispersed phase (i.e., W1/O primary emulsion) led to an increase in the International Dairy Journal, 29(2), 107–114.
Hoffmann, H., & Reger, M. (2014). Emulsions with unique properties from proteins as
apparent viscosity and the elastic character. Creaming and clarification emulsifiers. Advances in Colloid and Interface Science, 205, 94–104.
phenomena occurred during storage but not oiling off. Excellent Hu, Ya-Qiong, Yin, Shou-Wei, Zhu, Jian-Hua, Qi, Jun-Ru, Guo, Jian, Wu, Lei-Yan, …
encapsulation efficiency and stability were observed after the entrap­ Yang, Xiao-Quan (2016). Fabrication and characterization of novel Pickering
emulsions and Pickering high internal emulsions stabilized by gliadin colloidal
ment of vitamin B12 (as a model hydrophilic component) in W1. These particles. Food Hydrocolloids, 61, 300–310.
observations indicated the great potential of hordein nanoparticles in Jafari, Seid Mahdi, Sedaghat Doost, Ali, Nikbakht Nasrabadi, Maryam, Boostani, Sareh, &
stabilizing MPEs. Hordein nanoparticles can therefore be utilized as Van der Meeren, Paul (2020). Phytoparticles for the stabilization of Pickering
emulsions in the formulation of novel food colloidal dispersions. Trends in Food
green Pickering candidates for developing complex multi-purpose
Science & Technology., 98, 117–128.
colloidal systems with innovative techno-functional applications. Jiang, Hang, Zhang, Tong, Smits, Joeri, Huang, Xiaonan, Maas, Michael, Yin, Shouwei, &
Ngai, To (2021). Edible high internal phase pickering emulsion with double-
emulsion morphology. Food Hydrocolloids, 111, 106405. https://doi.org/10.1016/j.
CRediT authorship contribution statement foodhyd.2020.106405
Lin, Xiaoying, Li, Songnan, Yin, Juhua, Chang, Fengdan, Wang, Chan, He, Xiaowei, …
Sareh Boostani: Formal analysis, Investigation, Writing - original Zhang, Bin (2020). Anthocyanin-loaded double Pickering emulsion stabilized by
octenylsuccinate quinoa starch: Preparation, stability and in vitro gastrointestinal
draft. Masoud Riazi: Validation, Writing - review & editing. Ali Mar­ digestion. International Journal of Biological Macromolecules, 152, 1233–1241.
efati: Validation, Writing - original draft. Marilyn Rayner: Validation, Liu, Jinning, Zhou, Hualu, Muriel Mundo, Jorge L., Tan, Yunbing, Pham, Hung, &
Resources, Supervision. Seyed Mohammad Hashem Hosseini: McClements, David Julian (2020). Fabrication and characterization of W/O/W
emulsions with crystalline lipid phase. Journal of Food Engineering, 273, 109826.
Conceptualization, Resources, Supervision, Writing - review & editing. https://doi.org/10.1016/j.jfoodeng.2019.109826
Lobato-Calleros, C., Rodriguez, E., Sandoval-Castilla, O., Vernon-Carter, E. J., & Alvarez-
Ramirez, J. (2006). Reduced-fat white fresh cheese-like products obtained from W1/
Declaration of Competing Interest O/W2 multiple emulsions: Viscoelastic and high-resolution image analyses. Food
Research International, 39(6), 678–685.
Marefati, A., & Rayner, M. (2020). Starch granule stabilized Pickering emulsions: An
The authors declare that they have no known competing financial eight-year stability study. Journal of the Science of Food and Agriculture., 100,
interests or personal relationships that could have appeared to influence 2807–2811.
the work reported in this paper. Marefati, Ali, Sjöö, Malin, Timgren, Anna, Dejmek, Petr, & Rayner, Marilyn (2015).
Fabrication of encapsulated oil powders from starch granule stabilized W/O/W
Pickering emulsions by freeze-drying. Food Hydrocolloids, 51, 261–271.
Acknowledgement Matos, María, Timgren, Anna, Sjöö, Malin, Dejmek, Petr, & Rayner, Marilyn (2013).
Preparation and encapsulation properties of double Pickering emulsions stabilized
by quinoa starch granules. Colloids and Surfaces A: Physicochemical and Engineering
This work was financially supported by Shiraz University (Grant Aspects, 423, 147–153.
number: 95GCU2M194065). Authors are very thankful to Lund Uni­ Matos, M., Gutiérrez, G., Iglesias, O., Coca, J., & Pazos, C. (2015). Enhancing
versity for technical support. encapsulation efficiency of food-grade double emulsions containing resveratrol or
vitamin B12 by membrane emulsification. Journal of Food Engineering, 166, 212–220.
Matos, M., Guti, G., Martínez-rey, L., Iglesias, O., & Pazos, C. (2018). Encapsulation of
References resveratrol using food-grade concentrated double emulsions : Emulsion
characterization and rheological behaviour. Journal of Food Engineering, 226, 73–81.
McClements, David Julian (2018). Recent developments in encapsulation and release of
Beldengrün, Y., Dallaris, V., Jaén, C., Protat, R., Miras, J., Calvo, M., … Esquena, J.
functional food ingredients: Delivery by design. Current Opinion in Food Science, 23,
(2020). Formation and stabilization of multiple water-in-water-in-water (W/W/W)
80–84.
emulsions. Food Hydrocolloids, 102, 105588. https://doi.org/10.1016/j.
McClements, David Julian, & Jafari, Seid Mahdi (2018). Improving emulsion formation,
foodhyd.2019.105588
stability and performance using mixed emulsifiers: A review. Advances in Colloid and
Benichou, A., Aserin, A., & Garti, N. (2007). W/O/W double emulsions stabilized with
Interface Science, 251, 55–79.
WPI-polysaccharide complexes. Colloids and Surfaces A: Physicochemical and
Melle, Sonia, Lask, Mauricio, & Fuller, Gerald G. (2005). Pickering emulsions with
Engineering Aspects, 294(1-3), 20–32.
controllable stability. Langmuir, 21(6), 2158–2162.
Boostani, S., Hosseini, S. M. H., Golmakani, M.-T., Marefati, A., Abdul Hadi, N. B., &
Muschiolik, Gerald (2007). Multiple emulsions for food use. Current Opinion in Colloid &
Rayner, M. (2020). The influence of emulsion parameters on physical stability and
Interface Science, 12(4-5), 213–220.
rheological properties of Pickering emulsions stabilized by hordein nanoparticles.
Muschiolik, Gerald, & Dickinson, Eric (2017). Double emulsions relevant to food systems:
Food Hydrocolloids, 101, 105520. https://doi.org/10.1016/j.foodhyd.2019.105520
Preparation, stability, and applications. Comprehensive Reviews in Food Science and
Boostani, S., Hosseini, S. M. H., Yousefi, G., Riazi, M., Tamaddon, A.-M., & Van der
Food Safety, 16(3), 532–555.
Meeren, P. (2019). The stability of triphasic oil-in-water Pickering emulsions can be
O’Regan, Jonathan, & Mulvihill, Daniel M. (2010). Sodium caseinate–maltodextrin
improved by physical modification of hordein- and secalin-based submicron
conjugate stabilized double emulsions: Encapsulation and stability. Food Research
particles. Food Hydrocolloids, 89, 649–660.
International, 43(1), 224–231.
Bou, R., Cofrades, S., & Jiménez-Colmenero, F. (2014). Physicochemical properties and
Samborska, Katarzyna, Boostani, Sareh, Geranpour, Mansoureh, Hosseini, Hamed,
riboflavin encapsulation in double emulsions with different lipid sources. LWT –
Dima, Cristian, Khoshnoudi-Nia, Sara, … Jafari, Seid Mahdi (2021). Green
Food Science and Technology, 59(2), 621–628.

11
S. Boostani et al. Food Chemistry 372 (2022) 131354

biopolymers from by-products as wall materials for spray drying microencapsulation Xiao, Jie, Li, Yunqi, & Huang, Qingrong (2016). Recent advances on food-grade particles
of phytochemicals. Trends in Food Science & Technology, 108, 297–325. stabilized Pickering emulsions: Fabrication, characterization and research trends.
Schuch, A., Helfenritter, C., Funck, M., & Schuchmann, H. P. (2015). Observations on the Trends in Food Science & Technology, 55, 48–60.
influence of different biopolymers on coalescence of inner water droplets in W/O/W Xiao, Jie, Lu, Xuanxuan, & Huang, Qingrong (2017). Double emulsion derived from
(water-in-oil-in-water) double emulsions. Colloids and Surfaces A: Physicochemical kafirin nanoparticles stabilized Pickering emulsion: Fabrication, microstructure,
and Engineering Aspects, 475, 2–8. stability and in vitro digestion profile. Food Hydrocolloids, 62, 230–238.
Spyropoulos, Fotis, Duffus, Laudina J., Smith, Paul, & Norton, Ian T. (2019). Impact of Zhou, F. Z., Zeng, T., Yin, S. W., Tang, C. H., Yuan, D. B., & Yang, X. Q. (2018).
Pickering intervention on the stability of W1/O/W2 double emulsions of relevance to Development of antioxidant gliadin particle stabilized Pickering high internal phase
foods. Langmuir, 35(47), 15137–15150. emulsions (HIPEs) as oral delivery systems and the: In vitro digestion fate. Food and
Szumała, Patrycja, & Luty, Natalia (2016). Effect of different crystalline structures on W/ Function, 9, 959–970.
O and O/W/O wax emulsion stability. Colloids and Surfaces A: Physicochemical and Zhu, Y., Gao, H., Liu, W., Zou, L., & McClements, D. J. (2019). A review of the rheological
Engineering Aspects, 499, 131–140. properties of dilute and concentrated food emulsions. Journal of Texture Studies, 51,
Xi, Zewen, Liu, Wei, McClements, David Julian, & Zou, Liqiang (2019). Rheological, 45-44.
structural, and microstructural properties of ethanol induced cold-set whey protein
emulsion gels: Effect of oil content. Food Chemistry, 291, 22–29.

12

You might also like