You are on page 1of 12

Food Chemistry 430 (2024) 137033

Contents lists available at ScienceDirect

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

Effects of glycation method on the emulsifying performance and interfacial


behavior of coconut globulins-fucoidan complexes
Qianqian Zhu a, Haiming Chen a, Weijun Chen a, Ming Zhang a, Qiuping Zhong a, Zixin Chen b,
Jianfei Pei a, *, Wenxue Chen a, *
a
School of Food Science and Engineering, Hainan University, 58 People Road, Haikou 570228, PR China
b
Wenchang Zaineng Industrial Co., Ltd., Dongjiao Town, Wenchang 571300, PR China

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

Keywords: Coconut globulins (CG) possesses potential as an emulsifier but has not been utilized well. In this study, the
Coconut emulsifying performance of glycated CG-fucoidan (CGF) complexes, and the relationship between emulsifying
Globulins stability and interfacial behavior were investigated. The results showed that the grafting of fucoidan increased
Fucoidan
the molecular weight of CG, and decreased the zeta potential and fluorescence intensity. With the higher
Glycosylation
glycosylation degree, the fucoidan modified CG exhibited better emulsifying stability and higher viscosity.
Emulsion
Interfacial adsorption Moreover, the result of adsorption kinetics revealed that elasticity was the main property of the interface layer.
Compared to CG, CGF complexes with high degree of glycosylation had thicker interfacial layer on the oil–water
interface. A thicker elastic interfacial layer may be beneficial to the emulsion stability, owing to the strong
interaction of electrostatic repulsion and steric hindrance between oil droplets. These findings may provide
useful information for glycated CGF complexes as emulsifiers in functional food.

1. Introduction lead to the squandering of resource, but also contaminate the environ­
ment. How to implement the high-value utilization of coconut protein is
Plant-based proteins are of increasing interest to scientists as emul­ a problem that needs to be solved.
sifier of multiphase food system, owing to their advantages of nutritional Currently, it is an attractive research direction to develop coconut
and functional properties (Drusch et al., 2021). Coconut protein can protein as an emulsifier for high-value utilization in the food industry.
quickly adsorb on the oil–water interface to reduce interfacial tension. It Coconut proteins were separated by different solvents, mainly contained
seems to be a promising emulsifier alternative with a well-balanced albumins, globulins, gliadins, glutelins-1 and glutelins-2, of which
amino acids profile, the compositions contain all the nine essential globulins accounted for 60–75% of the total protein isolates (Patil &
amino acids, and non-essential amino acids (Martinez-Padilla et al., Benjakul, 2017). The majority of coconut globulins (CG) is the 11S
2022). The amount of essential amino acid contains 71% to 77%, and it globulins, also called cocosin. It has been reported that CG is a hexamer
can be easily digested and absorbed by the body (Patil & Benjakul, of 55 kDa subunit linked by disulfide bonds, the subunit is consisted of
2018). However, coconut protein is usually discarded in the environ­ an acidic polypeptide and a basic polypeptide (Ma et al., 2022). Previous
ment as a waste by-product after the extraction of virgin coconut oil studies have proved that CG has excellent emulsifying ability and can
(Thaiphanit & Anprung, 2016). Although coconut protein only makes up quickly adsorbed on the oil–water interface to reduce interfacial tension
2.6–4.4% (w/w) of fresh coconut meat, the production is very large (Patil & Benjakul, 2017). However, the solubility of CG is so low that its
(Patil & Benjakul, 2018; Thaiphanit & Anprung, 2016). As reported by emulsifying stability is poor, which limits its applications in food in­
FAO, the production of coconut in 2021 was about 63.68 million tonnes dustry (Raghavendra & Raghavarao, 2012). There are many modifica­
globally (https://www.fao.org). The present treatment may not only tion methods used for promoting properties of plant-based proteins,

Abbreviations: CG, Coconut globulins; CGF, Coconut globulins-fucoidan; CI, Creaming index; DG, The degree of grafting; FTIR, Fourier-transform infrared
spectroscopy; OPA, O-Phthalaldehyde; QCM-D, Dissipative quartz crystal microbalance; SDS-PAGE, Sodium dodecyl sulfate–polyacrylamide gel electrophoresis;
UV–vis spectrophotometer, Ultraviolet–visible spectrophotometer.
* Corresponding authors at: School of Food Science and Engineering, Hainan University, 58 People Road, Haikou 570228, PR China.
E-mail addresses: peijianfei@hainanu.edu.cn (J. Pei), chwx@hainanu.edu.cn (W. Chen).

https://doi.org/10.1016/j.foodchem.2023.137033
Received 4 May 2023; Received in revised form 20 July 2023; Accepted 25 July 2023
Available online 26 July 2023
0308-8146/© 2023 Elsevier Ltd. All rights reserved.
Q. Zhu et al. Food Chemistry 430 (2024) 137033

such as ultrasonic treatment, enzymatic hydrolysis and glycosylation solution (10 mg/mL). Different quantities of fucoidan (0.1, 0.5, 1.0, 2.0,
(Ding et al., 2021; Lad & Murthy, 2012; Li et al., 2022). Glycosylation is 3.0 and 4.0 g) were dissolved in 100 mL deionized water to obtain
a relatively safe, simple method to improve the functional properties of fucoidan stock solution (1, 5, 10, 20, 30, 40 mg/mL).
protein, which can stabilize emulsions by regulating the thickness of (ii) To prepare glycated CGF complexes, 20 mL of CG stock solution
interfacial layer (Ai et al., 2021). The concept has been studied in was mixed with 20 mL fucoidan stock solution. The Millard reaction was
emulsion stability, the encapsulation of bioactive substance and pro­ performed at 60℃ for 2 h, based on the method of Zhang et al. (2022)
tection against oxidation (Zhang et al., 2014; Zhang et al., 2022). At with some modification. CG with or without heat treatment at 60℃ were
present, studies on the emulsifying performance of glycated CG are used as a control group,
poorly reported. Moreover, the mechanism, especially the relationship After the Millard reaction, the samples were immediately transferred
between emulsifying stability and thickness of interfacial layer formed to ice bath to terminate the reaction. One part was used to measure the
by CG-polysaccharide was not studied. degree of grafting, SDS-PAGE, fluorescence spectroscopy, the other part
Fucoidan is a sulfated anionic polysaccharide extracted from brown was freeze-dried at − 50℃ for emulsion preparation and interfacial
algae and echinoderms, mainly composed of repeating fucose unit with behavior study.
sulfated ester groups. Fucoidan has excellent health benefits and phar­ In this study, glycated CGF complexes with different weight ratios of
macological activity, such as anti-inflammatory, antitumor and antiox­ 10:1, 2:1, 1:1, 1:2, 1:3, 1:4 were named CGF10:1, CGF2:1, CGF1:1, CGF1:2,
idant activity (Liu et al., 2020). Many bioactive properties of fucoidan CGF1:3, CGF1:4 respectively.
comes from the presence of sulfated groups (Alghazwi et al., 2019). The
anionic groups of fucoidan has the potential to interact with cationic 2.3. Characterization of the degree of glycosylation and structure of CGF
amino acid residues on CG, thereby improving the physicochemical and complex
emulsifying properties (Liu et al., 2020).
Therefore, the present study focused on the globulins in coconut, the 2.3.1. Degree of grafting
emulsifying stability of glycated coconut globulins by fucoidan, and the The grafting degree (DG) of CGF complexes was measured according
relationship of emulsifying stability and interfacial adsorption behavior to OPA method (Feng et al., 2023). The OPA solution was prepared as
were investigated. First, fucoidan was conjugated on the CG using wet- follows, 40 mg O-Phthalaldehyde (OPA) was dissolved in 1 mL methanol
heating method. The structural characteristics of CGF were analyzed by and then mixed with 2.5 mL SDS (20%, w/v), 25 mL sodium borate
grafting degree, protein pattern, FTIR, fluorescence spectroscopy and solution (0.1 mol/L) and 100 µL β-mercaptoethanol. The final volume
zeta potential. Then, the emulsifying performance of CG and glycated was 50 mL. Next, 200 µL of the CG or CGF complexes sample was added
CGF complexes was studied by particle size, morphology, viscosity and to 4 mL of OPA solution, and incubated at 35℃ for 2 mins. The absor­
centrifugal stability. Finally, the adsorption behavior of glycated CGF bance was determined at 340 nm using UV–vis spectrophotometer
complexes on the oil–water interface was evaluated using interfacial (Mettler Toledo, Switzerland). The DG was calculated by following
dilatational rheology and dissipative quartz crystal microbalance. In equation:
addition, the emulsifying stability-interfacial behavior relationship of
A0 − At
glycated CGF complexes were proposed. DG (%) = × 100% (1)
A0
2. Materials and methods A0 and At were the absorbance before and after the Maillard reaction.

2.1. Materials 2.3.2. Sodium dodecyl sulfate–polyacrylamide gel electrophoresis (SDS-


PAGE)
The by-products of coconut skim milk without heat treatment was The SDS-PAGE experiment was performed according to Gerliani’s
kindly provided by Wenchang Zaineng Industrial Co., Ltd (Wenchang, method with some modification (Gerliani et al., 2019). The concentra­
China). Fucoidan (Mw = 190 kDa) was purchased from Macklin tion of separating gel was 12%, and that of stacking gel was 4%. The
Biochemical Co., Ltd. (Shanghai, China). Medium-chain triglyceride samples at 2 mg/mL were added to loading buffer with a ratio of 4:1 and
(MCT) was supplied by the Nisshin Oillio Group (Tokyo, Japan). O- boiled for 10 mins before loading to gel. The voltage of running stacking
Phthalaldehyde, KBr were purchased from Sangon Biotech Co., Ltd. gel was 80 V, and running the separating gel at 120 V. The gel was
(Shanghai, China). Other reagents were analytical grade. stained with Coomassie brilliant blue R250. The molecular weight of CG
and CGF complexes were calculated using a standard protein marker
2.2. Preparation glycated CGF complexes (10–170 kDa).

2.2.1. Purification of coconut globulins 2.3.3. Zeta potential


Purification of coconut globulins was carried out according to the Zeta potential was measured by Zetasizer (Malvern Instrument, UK)
modified method of Ma et al. (2023). The by-product of coconut skim as reported (Zhao et al., 2023). The samples were diluted 10-fold with
milk was dispersed in deionized water with a ratio of 1:20 (w/v), and 35 mM sodium phosphate buffer (pH 7.0) and then added into the
then stirred at 4℃ for 2 h. The mixture was centrifuged at 10000 rpm for sample cell. All measurements were tested in triplicate at 25℃.
15 mins, and there were three phases in the system: cream, aqueous
solution with albumins and precipitation. The precipitation was 2.3.4. Fourier-transform infrared (FTIR) spectroscopy
collected to extract globulins using 0.4 M NaCl in 35 mM sodium The FTIR spectra of CG and glycated CGF complexes were acquired
phosphate buffer (pH 7.0). Extraction procedure with deionized water using FTIR spectroscopy (Bruker Corp., Germany). At first, the sample
and NaCl solution were repeated three times. The supernatant was was grinded with KBr power under the ratio of 1:100 (w/w), and then
collected and dialyzed with deionized water for 48 h (6–8 kDa molecular tableted. The spectra were scanned in the range of 4000–400 cm− 1 with
weight cut off). The dialysate was freeze-dried and stored at − 20℃ for a resolution of 4 cm− 1 (Li et al., 2022).
further analysis.
2.3.5. Fluorescence spectroscopy
2.2.2. Preparation of CGF complexes by Millard reaction The fluorescence spectroscopy of CG and CGF complexes were
Glycated CGF complexes were prepared following the steps below: recorded by F-6000 fluorescence spectrophotometer (Hitachi, Japan).
(i) The CG (2.0 g) was dissolved in 0.4 M NaCl solution in 35 mM The excitation wavelength was set to 280 nm, and emission wavelength
potassium phosphate buffer (200 mL, pH 7.0) to obtain a CG stock was 290–500 nm (Xia et al., 2015). The slit width was 5 nm.

2
Q. Zhu et al. Food Chemistry 430 (2024) 137033

2.4. Emulsifying properties of CGF complexes where C0 is the concentration of the continuous phase, K is the
Boltzman constant, T is the absolute temperature, t is the adsorption
2.4.1. Emulsions preparation time, and D is the diffusion coefficient.
Different glycated CGF samples (CGF10:1, CGF2:1, CGF1:1, CGF1:2, Furthermore, the penetration and rearrangement rate of CG and CGF
CGF1:3, CGF1:4) and CG with or without heat treatment (1%, w/v) were at O/W interface were analyzed by the first-order equation:
hydrated overnight, then 20 wt% MCT oil was added to the solution. The [( )/( )]
ln πf − π t π f − π 0 = − ki t (4)
mixture was sheared at 12000 rpm for 2 mins by a high speed disperser
(IKA T18 digital ULTRA-TURRAX, Germany) to prepare emulsions (Ma where πf, π0, πt represent the interfacial pressure at the final
et al., 2023). adsorption time (10800 s), at the initial time and t time respectively.
The interfacial viscoelastic modulus of CG and CGF solutions (15 µL)
2.4.2. Droplet size and CI index of emulsions at the O/W interface was measured, the deformation amplitude (ΔA/A)
The droplet size of emulsions was measured by a static laser light was set to 10% at 0.1 Hz frequency for 3 h. All measurements were
scattering instrument (Malvern Instruments, UK). Fresh emulsion was carried out at 25℃. tan φ was calculated according to following Eqs:
added to the sink and dispersed automatically until the appropriate laser
obscuration. The refractive index of MCT oil and water were set to 1.45 tanφ = Ev /Ed (5)
and 1.33, respectively. The surface-weighted mean diameter (d3,2) and where φ represents phase angle between stress and strain. Ed is the
volume-weighted mean diameter (d4,3) were recorded (Niu et al., interfacial elastic modulus, and Ev is the interfacial viscous modulus.
2022b).
Creaming index (CI) index was calculated as follows: CI% = (height 2.5.2. Amplitude sweep
of the cream layer / total height of the emulsion) × 100%. The deformation amplitude (ΔA/A) from 2.5% to 30% was set to
study the perturbation of amplitude sweep on the structure of interfacial
2.4.3. Fluorescence micrograph of emulsions layer, the frequency was set to 0.1 Hz, and the droplet volume was 15 µL.
The microstructure of emulsion was observed by inverted fluores­ The interface was balanced for 3 h before measurement. To intuitively
cence system (MODEL ECLIPSE Cl-S, OPTIKA, Italy). Adding 40 µL Nile show the structure change of interface caused by drop deformation, the
red solution (0.2 wt%, dissolved in isopropanol) into 1 mL fresh emul­ Lissajous plots were created (Ma et al., 2023).
sion, mixed and stained at dark for 2 h. The sample was distributed
across a microslide, and immediately covered with a cover slip, imaged 2.5.3. Frequency sweep
at 10 × magnification (Ma et al., 2023). The structure change of interface layer under frequency perturbation
was measured from 0.005 to 1 Hz. The droplet volume was 15 µL, and
2.4.4. Viscosity measurement deformation amplitude (ΔA/A) was 10%. The interface on oil–water was
Apparent viscosity of emulsions was determined by HAAKE MARS III balanced for 10800 s before measurement (Niu et al., 2022b).
rotary rheometer (ThermoFisher, USA). The emulsion was added to the
parallel plate, and scraped off the excess sample outside the plate. The
gap of measurement was set to 1.000 mm. The apparent viscosity of 2.6. QCM-D analysis
emulsions was measured at the shear rate in the range of 1–100 s− 1. All
measurements were performed at 25℃ (Chen et al., 2023). The adsorption behavior of CG and CGF on the oil–water interface
were studied by dissipative quartz crystal microbalance (QCM-D) (Biolin
2.4.5. Long-term stability measurement Scientific AB, Sweden). Frequency and dissipation signal were recorded
The long-term stability of emulsions was evaluated using LUMiSizer simultaneously. According to previously established procedure (Niu
stability analyzer (Berlin, Germany). About 400 µL fresh emulsion was et al., 2022a), 300 µL MCT oil (0.5%, dissolved in chloroform) was
added to the bottom of the sample cell, and then placed horizontally into coated on the gold quartz sensor (QSX 301, 5 MHz), then the sensor was
the instrument. According to the standard procedure, the wavelength dried overnight in a vaccum drying oven at 40℃. At experiment, the
was 870 nm, and emulsions were centrifuged at 4000 rpm for 50 mins. sensors were cleaned and calibrated the baseline using buffer solution
All measurements were taken at 25℃. The transmission profile was firstly, and then the sample solutions were pumped into the measure­
recorded with time evolution. Integral transmission and instability index ment cell by a peristaltic pump, the flow rate was set to 50 µL/min. After
were calculated by SEPView 6.0 software after measurement (Niu et al., adsorption equilibrium, the sensors were rinsed by buffer solution to
2022b). remove the unbound samples. The data was calculated by the smartfit
model of Qsense Dfind. The frequency and dissipation shifts were based
2.5. Dilatational rheology measurement on the 3rd and 3th overtones.

2.5.1. Dynamic interfacial tension and interfacial dilatational rheology


Dilatational rheology measurement was performed at an optical 2.7. Molecular docking
contact anglemeter (Dataphysics Instruments GmbH, Germany) ac­
cording to a previously established procedure with some modification The interaction between CG and fucoidan was simulated by molec­
(Niu et al., 2022a). The droplet volume on the test of dynamic interfacial ular docking method (Feng et al., 2023). The main globular protein
tension was 25 µL. The interfacial pressure (π, mN/m) was calculated by (cocosin) in CG and the oligosaccharide structure simplified from
the equation as follows (Yang et al., 2023): fucoidan were chosen to build the CGF conjugated structure. Crystal
structure of cocosin (5WPW) was downloaded from the Protein Data
π = γ0 − γ (2)
Bank (https://www.rcsb.org/). The oligosaccharide structure of fucoi­
where γ0 is the interfacial tension of buffer solution and MCT oil dan was mapped by the software of ChemOffice. First, we used PyMOL
(20.35 mN/m), and γ represents the interfacial tension of CG and CGF to remove water molecules and add hydrogen atoms on cocosin struc­
complexes with MCT oil. ture, and then the protein and oligosaccharide unit structure were
The diffusion rate of CG and CGF at oil–water interface was analyzed further modified by AutoDock Tools 1.5.6 software. The molecular
by Ward and Tordai equation: docking was launched in AutoGrid and AutoDock. After 3D molecular
docking procedure was finished, the pdbqt format file was imported in
π = 2C0 KT(Dt/3.14)1/2 (3) LigPlot software to generate 2D graph of CGF molecular docking map.

3
Q. Zhu et al. Food Chemistry 430 (2024) 137033

Fig. 1. The grafting degree and SDS-PAGE pattern (A), Zeta potential (B), FTIR spectroscopy (C) and fluorescence emission spectra (D) of CG and glycated CGF
complexes. Different letters in each column indicate significant differences (P < 0.05).

2.8. Statistical analysis fucoidan. The DG of glycated CGF complexes increased gradually as the
ratio of fucoidan increased (P < 0.05), and reached the maximum at a
All the experiments were conducted in triplicate, and the data was ratio of 1:4, which was 55.70%. DG could be affected by the properties
expressed as average values ± standard deviations. Statistical differ­ of proteins and polysaccharides (Zhao et al., 2023), it reported that the
ences between the experimental groups were calculated using the least DG of apple pectin, citrus pectin, mango pectin, sugar beet pectin and
significant difference (LSD) test (P < 0.05) by Origin 2023 software (EI- coconut protein isolate were 59.11%, 52.80%, 41.39%, and 39.2%
Hadary et al., 2023). respectively (Li et al., 2022). The differences of DG maybe due to
different molecular weight of polysaccharide. Generally, the lower
3. Results and discussion molecular weight will cause the greater degree of glycosylation at the
same reaction condition.
3.1. Characterize glycated CGF complexes SDS-PAGE was used to evaluated the glycosylation degree of CG and
fucoidan. As reviewed previous study, CG are composed of 11S globulins
3.1.1. Glycation degree of CG with fucoidan or cocosin (86%), and 7S globulins (14%). Cocosin is a hexamer of 55
A train of glycated CGF samples with different mass ratios (10:1, 2:1, kDa subunit linked by disulfide bonds. The 7S globulins is a trimeric
1:1, 1:2, 1:3, 1:4) was prepared. The characterization of glycated CGF protein comprised by 16, 22 and 24 kDa subunits (Ma et al., 2022). As
complexes was shown in Fig. 1, including the degree of grafting (DG), shown in Fig. 1A, the main bands of CG were mainly distributed in
SDS-PAGE pattern, Zeta potential, FTIR and fluorescence emission 15–55 kDa, a clear 11S globulins subunit was found in the lane of un­
spectra. The glycosylation of CG and fucoidan depends on the covalent treated CG and CG-heated, which was consistence with the results of Ma
reaction of the amino groups of protein with the carbonyl groups of the et al. (2022). For the glycated CGF complexes, the intensity of 55 kDa
reducing sugar (Ai et al., 2021). Therefore, DG can be evaluated by band decreased gradually with increasing the DG. Compared with CG or
measuring the amount of free amino groups in the system. Fig. 1A de­ CG-heated, there were obviously darker bands at the top of the stocking
picts the DG values of glycated samples with different mass ratio of gel and separation gel, which may be attributed to the conjugation of CG

4
Q. Zhu et al. Food Chemistry 430 (2024) 137033

and fucoidan. With the increase of fucoidan ratio, the band of 55 kDa Table 1
gradually diminished, indicating the increase of glycation degree. The Droplet size and CI index of emulsion stabilized by CG and glycated CGF
SDS-PAGE pattern was consistent with the result of DG. Protein pattern complexes.
of glycated coconut protein isolates with pectin also exhibited a similar Sample Stored for 24 h
phenomenon (Li et al., 2022). d3,2 (µm) d4,3 (µm) CI (%)

CG 2.88 ± 0.07a 13.43 ± 0.73a 19.27 ± 0.17a


3.1.2. Zeta potential CG-heated 2.95 ± 0.04a 13.48 ± 0.41a 18.35 ± 0.10b
The zeta potential was used to measure the surface charge density of CGF10:1 2.49 ± 0.33b 10.09 ± 2.89bcd 17.50 ± 0.13c
CGF2:1 2.34 ± 0.01b 8.73 ± 0.06d 15.34 ± 0.11e
particles, which could offer an indication of electrostatic repulsion
CGF1:1 2.37 ± 0.01b 8.90 ± 0.04 cd 16.35 ± 0.13d
interaction. Fig. 1B shows the zeta potential of CG and CGF complexes CGF1:2 2.77 ± 0.03a 10.60 ± 0.17bc ND
after glycosylation. The higher absolute value of zeta potential can in­ CGF1:3 2.82 ± 0.02a 10.35 ± 0.13bcd ND
crease the electrostatic repulsion interaction, which lead to the greater CGF1:4 2.88 ± 0.04a 11.29 ± 0.17b ND
distance between particles. On the contrary, particles will tend to All data are expressed as mean ± standard deviation, and different letters
coagulation or flocculation. The zeta potential of CG without treatment indicate significant difference (P < 0.05). ND: not detected.
was − 14.4 mV, and the CG-heated was − 11.0 mV, which reflected the
solution aggregation after heat treatment to some extent. The absolute emulsions. As shown in Table 1, the CI index of emulsions prepared by
value of zeta potential of CG heated at 80℃ has also been reported to be CGF1:2, CGF1:3, CGF1:4 could not be detected after storage for 24 h,
smaller than the control, the electrostatic repulsion interactions weak­ which were relatively stable. The mean size of emulsion stabilized by CG
ened (Ma et al., 2022). When the value of DG and fucoidan ratio was 2.88 ± 0.07 µm. As the ratio of fucoidan gradually increased (from
increased, the absolute value of zeta potential increased significantly, 10:1 to 1:1), the droplet size of oil was significantly (P < 0.05) decreased
which was attributed to the increase of fucoidan grafting on the CG to 2.49 ± 0.33 µm (CGF10:1), 2.34 ± 0.01 µm (CGF2:1) and 2.37 ± 0.01
surface. It is generally considered that when zeta potential value is µm (CGF1:1). This was due to the glycated CGF complexes could quickly
greater than ± 30 mV, the system is more stable (Lu et al., 2023). adsorb on the surface of oil droplets during high speed dispersion, and
Therefore, glycated CGF complexes with ratios of 1:1, 1:2, 1:3 and 1:4 formed a viscoelastic film to inhibit the small oil droplets from merging
were more stable with zeta potentials of − 35.53 ± 1.79 mV, − 46.03 ± into larger droplets (Niu et al., 2023). Interestingly, the size of oil
2.51 mV, − 49.93 ± 4.31 mV and − 46.63 ± 5.20 mV, respectively. droplets increased to 2.77 ± 0.03 µm (CGF1:2), 2.82 ± 0.02 µm (CGF1:3),
2.88 ± 0.04 µm (CGF1:4) as the ratio of fucoidan increase continuously.
3.1.3. Structural characteristics of CGF The trend was consistence with the value of d4,3. It is also believed that a
FTIR spectra reflect the secondary structure and hydrogen bonding good emulsifier tended to provide a small droplet size. However, a
forces of protein, which is widely used to analyze the glycosylation similar phenomenon has been found in the high internal phase emulsion
modification of protein. To verify the conjugate formation of CG and gels stabilized by insoluble soy polysaccharide nanoparticles (c = 0.25
fucoidan, FTIR analysis was used to reveal the characteristics of the wt%) (Yang et al., 2020b), which may be explained by the phenomenon
modified proteins. As shown in Fig. 1C, untreated CG showed typical of bridging flocculation. Flocculation refers to the combination of two or
diffraction peaks at wavenumber of 3437.9, 1648.3, 1157.5, 863.8 more oil droplets, but maintain their integrity. In general, flocculation is
cm− 1, that represented the O–H stretching vibrational, amide I bands detrimental to the stability of emulsion. However, moderate bridging
(C–– O stretching vibrational), NH2 inward osway and bending vibration
flocculation will cause an increase of emulsions viscosity, which is
of NH groups, respectively. These bands are consistent with the reported beneficial to the stability of emulsions.
spectrum of coconut protein isolate (Li et al., 2022). There was not To confirm our hypothesis, the microstructure of oil droplets in
obvious change in FTIR spectrum between CG-heated and natural CG. emulsions was observed by inverted fluorescence microscopy. The oil
After glycosylation modification, a new diffraction peak appeared at the phase was stained by Nile red. As depicted in Fig. 2A, O/W emulsions
wavenumber of 1422.8 cm− 1, which represented the formation of C–N were successfully prepared. Compared to the emulsion stabilized by CG
covalent bond at the CG to fucoidan mass ratios of 10:1, 2:1, 1:1, 1:2, or CG-heated, the oil droplet size stabilized by glycated CGF complexes
1:3, 1:4 (Su et al., 2010). This was a typical characteristic of carbonyl­ decreased as the ratio of CG and fucoidan from 10:1 to 1:1. However, the
amine binding between protein and carbohydrate by Millard reaction, oil droplets occurred flocculation phenomenon when the CG and
which was consistent with the phenomenon of canola protein isolate- fucoidan ratio was 1:2, 1:3, 1:4. This could explain why the droplet size
gum Arabic Maillard conjugate (Pirestani et al., 2018). of oil phase increased. In addition, the binding activity between oil
The fluorescence of aromatic amino acid residues is very sensitive to droplets was strong to resist the stirring and shaking during the light
surrounding environment (Li et al., 2019). Conformational changes of scattering measurements. The strong interaction between two or more
CG modified by grafting fucoidan can be revealed by fluorescence neighboring droplets may be attributed to the viscoelastic interfacial
quenching experiment. The emission spectra of CG and CGF are shown layer on the oil–water interface, which can hinder the movement of oil
in Fig. 1D, CG exhibited the maximum emission peak at 339.6 nm when droplets and prevent the breakage of oil droplets (Niu et al., 2023).
stimulated by 280 nm excitation light. After glycosylation, the intensity
of fluorescence decreased with the increase in fucoidan ratio, which was 3.2.2. Rheology properties of emulsions
attributed to the shielding effect of polysaccharide chain (Fan et al., The rheological properties of emulsions play an important role in
2017). For the covalent grafting between CG and fucoidan, carbonyl their stability and functionality for the application. The emulsions with
groups could act as quenching agents causing the fluorescence 20% MCT oil were prepared to study the rheological properties of CG
quenching. Zhang et al. (2022) also showed that glycated egg protein- and glycated CGF complexes. As can be seen in Fig. S1, the apparent
based complexes caused the fluorescence quenching phenomenon, viscosity of emulsions with shear rate was measured. The viscosity
when conjugated with lactose, fructose, xylose and glucose. gradually decreased with the increasing shear rate in the range of 1–100
s− 1, the difference of viscosity between CGF and CG were significant in
3.2. Emulsifying properties of CGF complexes high speed rate (10–100 s− 1) than that at low shear rate (1–10 s− 1). The
phenomenon of shear thinning demonstrated that all samples were non-
3.2.1. Droplet size and microstructure of emulsions Newtonian fluid (Liu et al., 2023). Moreover, the viscosity of emulsions
Droplet size is one of important indicators to evaluate the stability of stabilized by glycated CGF complexes was higher than CG and CG-
emulsions. The effect of different concentration of fucoidan glycated CG heated, which may be beneficial to the greater stability. With CGF1:4
on the droplet size (d3,2 and d4,3) and CI index were studied in the O/W

5
Q. Zhu et al. Food Chemistry 430 (2024) 137033

Fig. 2. Microstructure images (A), Evolution of transmission profiles (B), Evolution of integral transmission (C), the change rates of the integral transmission (D),
evolution of instability index (E) and the instability index (F) of emulsions stabilized by 1% CG and glycated CGF complexes.

6
Q. Zhu et al. Food Chemistry 430 (2024) 137033

Fig. 3. Time dependence of interfacial tension (A), the square root of time dependence of interfacial pressure (π, mN/m) (B), time dependence of ln[(π10800-πt) /
(π10800-π0)] (C), interfacial viscoelastic modulus as a function of interfacial pressure (D), time evolution of interfacial elastic modulus (E) and time evolution of tanφ
(F), interfacial viscoelastic modulus as a function of amplitude (G) and frequency (H) of 1% CG and glycated CG complexes.

7
Q. Zhu et al. Food Chemistry 430 (2024) 137033

having the highest viscosity, which was consistence with the macro constant (Kdiff). The rate of adsorption slowed down after the short
observation (data not shown). Chen et al. (2023) also verified that the disffusion-controlled adsorption process, which was attributed to the
droplets movement was more easily limited by higher concentration of presence of energy barriers (Niu et al., 2022a). Fig. 3C showed the
Mesona chinensis Benth polysaccharide, and improved the stability of curves of ln[(π10800-πt)/(π10800-π0)] over time for CG and glycated CGF
emulsions. complexes. The curve exhibited two linear regions generally, the initial
and second slope represented the penetration rate (Kp) and rearrange­
3.2.3. Long-term stability analysis of emulsions ment rate (Kr), respectively. The addition of fucoidan reduced the value
The emulsion was initially an opaque and uniform system. According of Kdiff in solutions, Kp and Kr also exhibited the same tendency
to the standard procedure, the emulsions were gradually centrifuged (Table S1). It might be attributed to the molecular entanglement and
using LUMiSizer to accelerate instability. The typical transmission pro­ high viscosity in the system prolonged the diffusion, penetration and
file change of CG and glycated CGF complex with time is presented in rearrangement rate of CGF molecules.
Fig. 2B. During the process of rapid centrifugation, the oil phase As shown in Fig. 3D, the E-π curves were used to present the visco­
migrated upward owing to low density, while the aqueous phase with elasticity of interfacial layers on oil–water interface. E increased with
higher transparency remained at the bottom. As the area of transmission the increase in interfacial pressure, indicating the continuous adsorption
profile decreased, the emulsion tended to be more stable. The first of CG and glycated CGF complexes at interface, forming viscoelastic
transmission profile of emulsions prepared by natural CG and CG-heated films. When the ratio of CG to fucoidan was 1:3 and 1:4, the E-π plots
were far away from zero, which reflected that emulsions were extremely shifted towards lower π values, which indicated an increased accumu­
unstable. Emulsions stabilized by glycated CGF complexes became sta­ lation of the CGF complex on the interface. In addition, the slopes of E-π
ble with an increase in the ratio of fucoidan. In other words, the physical curves were greater than 1, which indicated strong molecular interac­
stability of the emulsions was enhanced (Ma et al., 2023). tion to form viscoelastic film (data not shown). Fig. 3E showed the
To facilitate the comparison of transmission changes, the integral curves of elastic modulus over time. Ed gradually increased with
transmission profiles with time were plotted in Fig. 2C. The slopes of continuous adsorption of CG and glycated CGF complexes on the
transmission profiles were shown in Fig. 2D, which reflected the rates of interface, and reached equilibrium finally. Ed showed a slight growth in
transmission. The value of slope is inversely proportional to the stability later stages, indicating the molecules reorganization on the interface.
of emulsion. The rate of integral transmission decreased with the in­ These results were also found in interfacial behavior of Mesona chinensis
crease ratio of fucoidan to CG in glycated system, and emulsions became Benth polysaccharide (Chen et al., 2023). Fig. 3F showed the change of
more stable than the control. The change in the instability index over tanφ over time for CG and glycated CGF complexes. The tanφ value of all
time and instability index at the ending of separation were shown in samples is<1 (data not shown), indicating elasticity is the main property
Fig. 2E and F. The range of the instability index is from 0 to 1, 0 means of the interface (Bouyer et al., 2012). The smaller value of tanφ means
that the sample is very opaque and stable, while 1 is extremely unstable the better elastic properties.
(Fernandes et al., 2017). Initially, the emulsion was relatively uniform,
which separated with time, so the clarification index increased. The 3.3.2. Amplitude and frequency sweeps
instability index of emulsion prepared by glycated CGF complexes Amplitude and frequency sweeps were adopted to evaluate me­
decreased compared to the control. It reflects that emulsion stabilized by chanical properties of the interface layer. The change of viscoelastic
glycated CGF complexes was more stable. The instability index of modulus with amplitude sweeps was shown in Fig. 3G. Viscoelastic
emulsion prepared by 1% CGF1:4 was 0.439 ± 0.029, which may be modulus of all samples gradually decreased with the amplitude defor­
owing to the higher viscosity and the thicker film on the interface. Ac­ mation from 5% to 30%, which indicated the interfacial layer was
cording to previous study, moderate bridging flocculation is beneficial damaged (Niu et al., 2022b). According to the principle of Fourier
to the stability of emulsions owing to the higher viscosity (Li et al., transform that was used to calculate viscoelastic modulus, the nonline­
2021). arity existed in original signal were ignored. Thus, the Lissajous plots
retain original data were used to illustrate the nonlinear dilatational
3.3. Interfacial properties of CGF complexes on the oil–water interface behavior (Fig. S2). As the deformation of ΔA/A increased, the curve of
Lissajous plots became wider, which was attributed to the response of
3.3.1. Dilatational interfacial tension and interfacial dilatational rheology viscosity (Yang et al., 2020a). When the amplitude sweep (ΔA/A =
analysis 30%), the slope of all sample curves decreased during extension,
To better understand the behavior of CG and glycated CGF com­ implying that the interfacial layers underwent strain softening. In
plexes on oil–water interface, the interfacial tension was measured for 3 contrast, the slope of curves increased during compression deformation,
h. As shown in Fig. 3A, the interfacial tension of all samples decreased which suggested that the interfacial layer experienced strain stiffening.
gradually with the time due to the continuous and spontaneous Tang & Shen (2015) also showed BSA molecules exhibited the similar
adsorption of CG and glycated CGF complexes on the surface of oil phenomenon. Fig. 3H displayed the change of viscoelastic modulus with
droplets. After adsorption equilibrium for 3 h, the interfacial tension of frequency, E increased with frequency. This phenomenon can be
CG and CGF1:4 were 6.68 mN/m and 7.91 mN/m respectively. Glycated explained by diffusion relaxation. At low frequency of perturbation,
CGF complexes showed higher interfacial tension than CG, indicating there were enough time for the CG and CGF complexes to the interface,
the decrease in affinity on interface. A similar phenomenon was resulting a small viscoelastic modulus exhibited by the interfacial layer.
demonstrated in soy protein isolate-gum Arabic composites stabilized However, when subjected to high frequency perturbation, the short
emulsion (Feng et al., 2023). This result suggested that reducing inter­ detection time limited the diffusion relaxation of the interfacial layer,
facial tension may not be the main reason for better emulsifying per­ showed a large viscoelastic modulus (Ma et al., 2023).
formance of CGF complexes.
Fig. 3B displayed the curves of interfacial pressure (π) verus the 3.3.3. QCM-D analysis
square root of time (t1/2) of CG and glycated CGF complexes. Compared QCM-D is an effective tool to study the adsorption processing of
to CG, glycated CGF complexes had lower interfacial pressure, which molecules on the interface. According to Sauerbrey equation, there is a
may be attributed to the larger hydrodynamic radii of glycated CGF linear relationship between the change of frequency and mass, so the
complexes. The characteristic limited the adsorption of protein on the thickness of adsorption layer can be calculated. Moreover, dissipation
surface of oil droplets. Another study demonstrated that glycated whey and energy loss are directly proportional, and can reflect the rheological
protein isolate and chitooligosaccharide exhibited lower initial π values information about the adsorption layer. Based on macroscopic obser­
(Qin et al., 2022). In addition, slope of the curves is called diffusion rate vation, we selected CG and four glycated CGF complexes that were

8
Q. Zhu et al. Food Chemistry 430 (2024) 137033

Fig. 4. Frequency change Δf (A) and dissipation change ΔD (B), ΔD-Δf curves of adsorption (C) and desorption (D) at n = 3 overtones, mass (E) and thickness (F) of
interfacial layer of glycated CGF complexes (1%).

unstable (CGF2:1) and relatively stable (CGF1:2, CGF1:3, CGF1:4) on with MCT oil were balanced by buffer solution. With the evolution of
emulsions for QCM-D analysis. Fig. 4 shows the curve of Δf (A), ΔD (B), time, there was a continuous absorption and reorganization of CG and
ΔD-Δf of adsorption (C) and desorption (D), the change of mass (E) and CGF covalent complexes at the interface between oil and water, which
thickness (F) about 1% CG and glycated CGF complexes on the led to a gradual decrease in frequency and an increase in dissipation
adsorption layer (n = 3 overtones). Initially, the gold sensors coated (Fig. 4A, 4B). The dissipation increased rapidly with the molecules

9
Q. Zhu et al. Food Chemistry 430 (2024) 137033

Fig. 5. Schematic mechanism of emulsions stabilized by CG and glycated CGF complexes.

adsorption on the interface continuously (Chen et al., 2023). The ΔD-Δf emulsifying performance of glycated CGF complexes.
curves of adsorption and desorption were shown in Fig. 4C and 4D, the
slope of K value indicates the energy loss with the change of per fre­ 4. Conclusion
quency. In our results, the curves were continuous, which indicated that
glycated CGF complexes were a slow adsorption and desorption process In conclusion, glycated coconut globulins-fucoidan (CGF) complexes
on the oil–water interface (Wei et al., 2020). In addition, the slopes of was fabricated to evaluate their emulsifying performance and interfacial
ΔD-Δf curves were discontinuous, K2<K1 (data not shown), indicated behavior. The results showed that glycated CGF complexes with high
that the conformation of glycated CGF complexes has been changed, grafting degree exhibited better emulsifying stability due to the higher
maybe owing to the multilayer adsorption of CGF covalent complexes on viscosity and a thicker elastic interfacial layer formed on the interface,
the interface. And K3>K4 at the desorption curve, the interfacial layer simultaneously provided higher electronegativity in emulsions. It was
became dense from loose. As reported before, whey protein also un­ suggested that appropriate glycosylation of CG with fucoidan could
dergone a conformational transition on the interface (Li et al., 2021). improve the physicochemical characterize and long-term stability of
The change of interfacial layer mass and thickness was consistent with emulsions. These findings are useful for the application of glycated
the Δf (Fig. 4E, F). The covalent complexes with high-ratio fucoidan had protein-polysaccharide in food emulsions, and can serve as reference for
obvious advantages in interfacial layer thickness, the layer thickness high-value utilization of coconut protein.
was 32.05 ± 0.03 nm stabilized by 1% CGF1:4. The layer thickness of CG,
CGF2:1, CGF1:2, CGF1:3 were 15.87 ± 0.01 nm, 15.16 ± 0.03 nm, 20.23 CRediT authorship contribution statement
± 0.02 nm, 29.33 ± 0.04 nm, respectively. This could explain why
glycated CGF complexes with the better emulsifying stability. Qianqian Zhu: Methodology, Investigation, Data curation, Writing –
original draft. Haiming Chen: Writing – review & editing, Funding
3.4. Schematic mechanism of glycated CGF complexes formation and acquisition. Weijun Chen: Data curation. Ming Zhang: Methodology,
their behavior at o/w interface Data curation. Qiuping Zhong: Validation. Zixin Chen: Investigation.
Jianfei Pei: Software. Wenxue Chen: Conceptualization, Supervision.
The interaction of fucoidan with cocosin was simulated by molecular
docking method. As shown in Fig. 5, the aldehyde group of fucoidan Declaration of Competing Interest
might be attached to the amino acid residue (Arg 196) on the surface of
cocosin, with a distance of 2.94 Å. The binding energy was − 5.42 kcal/ The authors declare that they have no known competing financial
mol, indicating that it was easy to form glycated conjugates between interests or personal relationships that could have appeared to influence
cocosin and fucoidan at these locations. In addition, as confirmed by the the work reported in this paper.
interfacial dilatational rheology and QCM-D analysis, glycated CGF
complexes with a high ratio of fucoidan could form a thicker elastic film Data availability
at the O/W interface. In this regard, stronger electrostatic repulsions and
steric hindrance between droplets were beneficial to the better No data was used for the research described in the article.

10
Q. Zhu et al. Food Chemistry 430 (2024) 137033

Acknowledgements Liu, X., Zhang, X., Ding, L., Jin, H., Chen, N., Huang, X., Jin, Y., & Cai, Z. (2023). Natural
egg yolk emulsion as wall material to encapsulate DHA by two-stage
homogenization: Emulsion stability, rheology analysis and powder properties. Food
This work was supported by Hainan Province Science and Technol­ Research International, 167, Article 112658. https://doi.org/10.1016/j.
ogy Special Fund (ZDYF2022XDNY166) and Hainan Provincial Natural foodres.2023.112658
Science Foundation of China (223RC402). Lu, F., Ma, Y., Zang, J., Qing, M., Ma, Z., Chi, Y., & Chi, Y. (2023). High-temperature
glycosylation modifies the molecular structure of ovalbumin to improve the freeze-
thaw stability of its high internal phase emulsion. International Journal of Biological
Appendix A. Supplementary data Macromolecules, 233, Article 123560. https://doi.org/10.1016/j.
ijbiomac.2023.123560
Ma, J., Chen, H., Chen, W., Wu, J., Li, Z., Zhang, M., Zhong, Q., & Chen, W. (2022).
Supplementary data to this article can be found online at https://doi. Effects of heat treatment and pH on the physicochemical and emulsifying properties
org/10.1016/j.foodchem.2023.137033. of coconut (Cocos nucifera L.) globulins. Food Chemistry, 388, Article 133031.
https://doi.org/10.1016/j.foodchem.2022.133031
Ma, J., Pan, C., Chen, H., Chen, Y., Chen, W., Pei, J., … Chen, W. (2023). Interfacial
References: behavior of coconut (Cocos nucifera L.) globulins at different pH: Relation to
emulsion stability. Food Hydrocolloids, 144, Article 108958. https://doi.org/
10.1016/j.foodhyd.2023.108958
Ai, M., Xiao, N., & Jiang, A. (2021). Molecular structural modification of duck egg white
Martinez-Padilla, L. P., Hernandez-Rojas, F. S., Sosa-Herrera, M. G., & Juliano, P. (2022).
protein conjugates with monosaccharides for improving emulsifying capacity. Food
Novel application of ultrasound and microwave-assisted methods for aqueous
Hydrocolloids, 111, Article 106271. https://doi.org/10.1016/j.
extraction of coconut oil and proteins. Journal of Food Science and Technology, 59
foodhyd.2020.106271
(10), 3857–3866. https://doi.org/10.1007/s13197-022-05409-0
Alghazwi, M., Smid, S., Karpiniec, S., & Zhang, W. (2019). Comparative study on
Niu, H., Chen, X., Luo, T., Chen, H., & Fu, X. (2022a). Relationships between the
neuroprotective activities of fucoidans from Fucus vesiculosus and Undaria pinnatifida.
behavior of three different sources of pectin at the oil-water interface and the
International Journal of Biological Macromolecules, 122, 255–264. https://doi.org/
stability of the emulsion. Food Hydrocolloids, 128, Article 107566. https://doi.org/
10.1016/j.ijbiomac.2018.10.168
10.1016/j.foodhyd.2022.107566
Bouyer, E., Mekhloufi, G., Rosilio, V., Grossiord, J. L., & Agnely, F. (2012). Proteins,
Niu, H., Chen, X., Luo, T., Chen, H., & Fu, X. (2022b). The interfacial behavior and long-
polysaccharides, and their complexes used as stabilizers for emulsions: Alternatives
term stability of emulsions stabilized by gum arabic and sugar beet pectin.
to synthetic surfactants in the pharmaceutical field? International Journal of
Carbohydrate Polymers, 291, Article 119623. https://doi.org/10.1016/j.
Pharmaceutics, 436(1–2), 359–378. https://doi.org/10.1016/j.ijpharm.2012.06.052
carbpol.2022.119623
Chen, X., Niu, H., Chen, H., Zhu, X., Shen, M., & Xie, J. (2023). Interfacial behavior of
Niu, H., Wang, W., Dou, Z., Chen, X., Chen, X., Chen, H., & Fu, X. (2023). Multiscale
Mesona chinensis Benth polysaccharide at the oil-water interface and the stability of
combined techniques for evaluating emulsion stability: A critical review. Advances in
the emulsion. Food Hydrocolloids, 143, Article 108847. https://doi.org/10.1016/j.
Colloid and Interface Science, 311, Article 102813. https://doi.org/10.1016/j.
foodhyd.2023.108847
cis.2022.102813
Ding, Y., Chen, L., Shi, Y., Akhtar, M., Chen, J., & Ettelaie, R. (2021). Emulsifying and
Patil, U., & Benjakul, S. (2017). Characteristics of albumin and globulin from coconut
emulsion stabilizing properties of soy protein hydrolysates, covalently bonded to
meat and their role in emulsion stability without and with proteolysis. Food
polysaccharides: The impact of enzyme choice and the degree of hydrolysis. Food
Hydrocolloids, 69, 220–228. https://doi.org/10.1016/j.foodhyd.2017.02.006
Hydrocolloids, 113, Article 106519. https://doi.org/10.1016/j.
Patil, U., & Benjakul, S. (2018). Coconut milk and coconut oil: Their manufacture
foodhyd.2020.106519
associated with protein functionality. Journal of Food Science, 83(8), 2019–2027.
Drusch, S., Klost, M., & Kieserling, H. (2021). Current knowledge on the interfacial
https://doi.org/10.1111/1750-3841.14223
behaviour limits our understanding of plant protein functionality in emulsions.
Pirestani, S., Nasirpour, A., Keramat, J., Desobry, S., & Jasniewski, J. (2018). Structural
Current Opinion in Colloid & Interface Science, 56, Article 101503. https://doi.org/
properties of canola protein isolate-gum Arabic Maillard conjugate in an aqueous
10.1016/j.cocis.2021.101503
model system. Food Hydrocolloids, 79, 228–234. https://doi.org/10.1016/j.
EI-Hadary, A. R. E., Sulieman, A. M., & EI-Shorbagy, G. A. (2023). Comparative effects of
foodhyd.2018.01.001
hibiscus leaves and potato peel extracts on characteristics of fermented orange juice.
Tang, C.-H., & Shen, L. (2015). Dynamic adsorption and dilatational properties of BSA at
Journal of Food Quality and Hazards Control. 10, 39-50. https://doi.org/10.18502/
oil/water interface: Role of conformational flexibility. Food Hydrocolloids, 43,
jfqhc.10.1.11988.
388–399. https://doi.org/10.1016/j.foodhyd.2014.06.014
Fan, Y., Yi, J., Zhang, Y., Wen, Z., & Zhao, L. (2017). Physicochemical stability and in
Qin, X., Yu, J., Wang, Q., Zhang, H., Chen, H., Hu, Z., Lv, Q., & Liu, G. (2022).
vitro bioaccessibility of β-carotene nanoemulsions stabilized with whey protein-
Preparation of camellia oil Pickering emulsion stabilized by glycated whey protein
dextran conjugates. Food Hydrocolloids, 63, 256–264. https://doi.org/10.1016/j.
isolate and chitooligosaccharide: Effect on interfacial behavior and emulsion
foodhyd.2016.09.008
stability. LWT, 153, Article 112515. https://doi.org/10.1016/j.lwt.2021.112515
Feng, S., Guo, Y., Liu, F., Li, Z., Chen, K., Handa, A., & Zhang, Y. (2023). The impacts of
Raghavendra, A. N. S. N., & Raghavarao, K. S. M. S. (2012). Production of coconut
complexation and glycated conjugation on the performance of soy protein isolate-
protein powder from coconut wet processing waste and its characterization. Applied
gum Arabic composites at the o/w interface for emulsion-based delivery systems.
Biochemistry and Biotechnology, 167(5), 1290–1302. https://doi.org/10.1007/
Food Hydrocolloids, 135, Article 108168. https://doi.org/10.1016/j.
s12010-012-9632-9
ijbiomac.2023.123554
Su, J., Huang, Z., Yuan, X., Wang, X., & Li, M. (2010). Structure and properties of
Fernandes, A. R., Ferreira, N. R., Fangueiro, J. F., Santos, A. C., Veiga, F. J., Cabral, C.,
carboxymethyl cellulose/soy protein isolate blend edible films crosslinked by
Silva, A. M., & Souto, E. B. (2017). Ibuprofen nanocrystals developed by 22 factorial
Maillard reactions. Carbohydrate Polymers, 79(1), 145–153. https://doi.org/
design experiment: A new approach for poorly water-soluble drugs. Saudi
10.1016/j.carbpol.2009.07.035
Pharmaceutical Journal, 25(8), 1117–1124. https://doi.org/10.1016/j.
Thaiphanit, S., & Anprung, P. (2016). Physicochemical and emulsion properties of edible
jsps.2017.07.004
protein concentrate from coconut (Cocos nucifera L.) processing by-products and the
Gerliani, N., Hammami, R., & Aïder, M. (2019). Assessment of the extractability of
influence of heat treatment. Food Hydrocolloids, 52, 756–765. https://doi.org/
protein-carbohydrate concentrate from soybean meal under acidic and alkaline
10.1016/j.foodhyd.2015.08.017
conditions. Food Bioscience, 28, 116–124. https://doi.org/10.1016/j.
Wei, Y., Xie, Y., Cai, Z., Guo, Y., Wu, M., Wang, P., Li, R., & Zhang, H. (2020). Interfacial
fbio.2019.01.004
and emulsion characterisation of chemically modified polysaccharides through a
Lad, V. N., & Murthy, Z. V. P. (2012). Enhancing the stability of oil-in-water emulsions
multiscale approach. Journal of Colloid and Interface Science, 580, 480–492. https://
emulsified by coconut milk protein with the application of acoustic cavitation.
doi.org/10.1016/j.jcis.2020.07.048
Industrial & Engineering Chemistry Research, 51(11), 4222–4229. https://doi.org/
Xia, S., Li, Y., Xia, Q., Zhang, X., & Huang, Q. (2015). Glycosylation of bovine serum
10.1021/ie202764f
albumin via Maillard reaction prevents epigallocatechin-3-gallate-induced protein
Li, S., Sun, J., Yan, J., Zhang, S., Shi, C., McClements, D. J., Liu, X., & Liu, F. (2021).
aggregation. Food Hydrocolloids, 43, 228–235. https://doi.org/10.1016/j.
Development of antibacterial nanoemulsions incorporating thyme oil: Layer-by-layer
foodhyd.2014.05.022
self-assembly of whey protein isolate and chitosan hydrochloride. Food Chemistry,
Yang, H., Wang, S., Xu, Y., Wang, S., Yang, L., Song, H., He, Y., & Liu, H. (2023). Storage
339, Article 128016. https://doi.org/10.1016/j.foodchem.2020.128016
stability and interfacial rheology analysis of high-internal-phase emulsions stabilized
Li, Y., Liu, B., Jiang, L., Regenstein, J. M., Jiang, N., Poias, V., … Wang, Z. (2019).
by soy hull polysaccharide. Food Chemistry, 418, Article 135956. https://doi.org/
Interaction of soybean protein isolate and phosphatidylcholine in nanoemulsions: A
10.1016/j.foodchem.2023.135956
fluorescence analysis. Food Hydrocolloids, 87, 814–829. https://doi.org/10.1016/j.
Yang, J., Thielen, I., Berton-Carabin, C. C., van der Linden, E., & Sagis, L. M. C. (2020a).
foodhyd.2018.09.006
Nonlinear interfacial rheology and atomic force microscopy of air-water interfaces
Li, Z., Xi, J., Chen, H., Chen, W., Chen, W., Zhong, Q., & Zhang, M. (2022). Effect of
stabilized by whey protein beads and their constituents. Food Hydrocolloids, 101,
glycosylation with apple pectin, citrus pectin, mango pectin and sugar beet pectin on
Article 105466. https://doi.org/10.1016/j.foodhyd.2019.105466
the physicochemical, interfacial and emulsifying properties of coconut protein
Yang, T., Li, X.-T., & Tang, C.-H. (2020b). Novel edible pickering high-internal-phase-
isolate. Food Research International, 156, Article 111363. https://doi.org/10.1016/j.
emulsion gels efficiently stabilized by unique polysaccharide-protein hybrid
foodres.2022.111363
nanoparticles from Okara. Food Hydrocolloids, 98, Article 105285. https://doi.org/
Liu, Q., Chen, J., Qin, Y., Jiang, B., & Zhang, T. (2020). Zein/fucoidan-based composite
10.1016/j.foodhyd.2019.105285
nanoparticles for the encapsulation of pterostilbene: Preparation, characterization,
Zhang, Y., Tan, C., Abbas, S., Eric, K., Zhang, X., Xia, S., & Jia, C. (2014). The effect of soy
physicochemical stability, and formation mechanism. International Journal of
protein structural modification on emulsion properties and oxidative stability of fish
Biological Macromolecules, 158, 461–470. https://doi.org/10.1016/j.
ijbiomac.2020.04.128

11
Q. Zhu et al. Food Chemistry 430 (2024) 137033

oil microcapsules. Colloids and surfaces. B, Biointerfaces, 120, 63–70. https://doi.org/ Zhao, C., Chu, Z., Mao, Y., Xu, Y., Fei, P., Zhang, H., … Liu, J. (2023). Structural
10.1016/j.colsurfb.2014.05.006 characteristics and acid-induced emulsion gel properties of heated soy protein
Zhang, Y., Zhang, Y., Chen, N., Xin, N., Li, Q., Ye, H., … Zhang, T. (2022). Glycated isolate–soy oligosaccharide glycation conjugates. Food Hydrocolloids, 137, Article
modification of the protein from Rana chensinensis eggs by Millard reaction and its 108408. https://doi.org/10.1016/j.foodhyd.2022.108408
stability analysis in curcumin encapsulated emulsion system. Food Chemistry, 382,
Article 132299. https://doi.org/10.1016/j.foodchem.2022.132299

12

You might also like