Subscriber access provided by WEBSTER UNIV
Food and Beverage Chemistry/Biochemistry
Construction of Fucoxanthin Vector Based on Binding of Whey Protein Isolate and Its Subsequent Complex Coacervation with Lysozyme Junxiang Zhu, Hao Li, Ying Xu, and Dongfeng Wang J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b06679 • Publication Date (Web): 26 Feb 2019 Downloaded from http://pubs.acs.org on February 27, 2019
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 40
Journal of Agricultural and Food Chemistry
1
Construction of Fucoxanthin Vector Based on Binding of Whey Protein Isolate
2
and Its Subsequent Complex Coacervation with Lysozyme
3
Junxiang Zhu†,‡,§, Hao Li†, Ying Xu†, Dongfeng Wang*, †
4
†
5
Republic of China
6
‡
7
Republic of China
8
§
9
Zhejiang, People's Republic of China
College of Food Science and Engineering, Ocean University of China, Qingdao, 266003, the
Marine Fisheries Research Institute of Zhejiang, Zhoushan, 316021, Zhejiang, People's
Marine and Fisheries Research Institute, Zhejiang Ocean University, Zhoushan, 316021,
10 11
Corresponding author: Dongfeng Wang
12
Email:
[email protected] ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
13
Abstract
14
In this study, a novel vector for fucoxanthin was constructed using the ligand-binding property
15
of whey protein isolate and its subsequent heteroprotein complex coacervation with lysozyme.
16
The results showed that fucoxanthin could quench the intrinsic fluorescence of whey protein
17
isolate by a static mechanism, indicating that they could spontaneously form nanocomplex
18
through non-covalent interactions. Moreover, the structural and electrostatic properties of whey
19
protein were different from before with the binding of fucoxanthin, and this reason could be well
20
explained by molecular dynamic simulation. The size and ζ-potential tests showed that when the
21
whey protein isolate was combined with fucoxanthin and then coacervated with lysozyme, the
22
heteroprotein ratio and pH that affected the coacervation process also changed compared to free
23
whey protein isolate. The FT-IR results showed that fucoxanthin was successfully encapsulated
24
into complex coacervates. In addition, the heteroprotein system exhibited a higher loading
25
efficiency and also provided a better protection for fucoxanthin in heating, storage and simulated
26
gastrointestinal environments.
27
Keywords: fucoxanthin, whey protein isolate, lysozyme, ligand binding, heteroprotein complex
28
coacervation, molecular dynamics
ACS Paragon Plus Environment
Page 2 of 40
Page 3 of 40
Journal of Agricultural and Food Chemistry
29
INTRODUCTION
30
FX is a marine xanthophyll commonly found in macroalgae such as Undaria pinnatifida,
31
Laminaria japonica, and Ecklonia cava.1 It is also the most abundant carotenoid in nature and
32
contributes to more than 10% of the total natural carotenoid products.2 Although FX is not a
33
provitamin A, its special structure endows a variety of biological activities that are beneficial to
34
human health, including hypolipidemic, anti-obesity, antidiabetic, and anticarcinogenic effects.3
35
Among them, the anti-obesity property of FX attracts attention and has received support from
36
many in vitro and in vivo studies.4 A recent study also reported the effectiveness of FX reducing
37
the formation of β-Amyloid fibrils and attenuating β-Amyloid neurotoxicity in the early stage of
38
Alzheimer’s disease.5 These findings indicate that FX can be developed as a nutritional
39
supplement or used to produce functional foods.
40
However, similar to other carotenoids, FX is sensitive to degradation under ambient stresses (light,
41
temperature, oxygen) and gastrointestinal chemistry. To improve its poor stability and low
42
bioavailability, several delivery systems have been designed for FX. For instance, Salvia Trujillo
43
et al. investigated the bioavailability of FX in nanoemulsions and compared three different oils
44
used in carriers, which showed that oil type has a conspicuous impact on the bioavailability of
45
FX.6 Dai et al. constructed FX-loaded O/W microemulsions using medium chain triglycerides as
46
an oil phase, tween 80 as a surfactant, and polyethylene glycol 400 as a co-surfactant.7 The results
47
showed that about 95% of FX in microemulsions was retained for 4 weeks. These studies either
48
use an emulsification process that requires high energy, which reduces FX and is costly for the
49
food industry or uses synthetic surfactants that may be toxic.8,9
50
Whey proteins are GRAS materials according to the Food and Drug Administration of the United
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
51
States.10 In addition to providing nutrition, they also have physicochemical properties suitable for
52
the preparation of delivery systems, including binding ions or small molecules, self-assembly,
53
gelation, emulsification, and complex coacervation.11 Whey proteins have hydrophobic cavities
54
that can bind ligands of different sizes by non-covalent interactions, which are usually located in
55
the surface pocket or internal cavity of protein.12 Recently, some researchers have utilized this
56
binding property to construct the heteroprotein complex coacervates to transport the ligand bound
57
by whey proteins. Chapeau et al. prepared complex coacervates of lactoferrin and β-Lg, which
58
could load hydrophilic vitamin B9 at a content of 10 mg/g protein,13 mainly due to the binding
59
properties of β-Lg.14 This approach was later successfully expanded to the pilot scale.15 However,
60
lactoferrin is relatively expensive and may be replaced with other food protein such as Lyz, which
61
can be easily extracted from egg whites.16 The hydrophobic vitamin D3 was loaded in complex
62
coacervates of β-Lg and Lyz with a high encapsulation efficiency of 90.8 ± 4.8%.17 This study
63
also utilized the fact that β-Lg can bind vitamin D3 to form a complex.18 In their subsequent study,
64
this complex coacervate was shown to be effective in protecting vitamin D3 against UV-light
65
irradiation and increased bioavailability in vivo.19 Our previous study showed that β-Lg, BSA,
66
and α-La can form nanocomplexes with FX,20 suggesting we can try to construct FX vector based
67
on complex coacervates using heteroprotein from whey and other sources. However, FX has a
68
larger molecular weight and is not as heat resistant as vitamin D3.21 Another thing to note is that
69
the presence of ligand can affect the heteroprotein complex coacervation.22 When a protein forms
70
a complex with a ligand, the protonation of ionizable groups and the conformation of the
71
hydrophobic region sometimes change.23 These phenomena can affect the electrostatic and
72
hydrophobic forces between the two proteins, ultimately determining the coacervate size or
ACS Paragon Plus Environment
Page 4 of 40
Page 5 of 40
Journal of Agricultural and Food Chemistry
73
product formulation.
74
Therefore, The objective of this study was to construct a FX delivery system based on the binding
75
property of whey protein and heteroprotein complex coacervation. Food proteins WPI and Lyz
76
were selected in this study owing to their low cost and commercial availability, which might
77
increase the scalability of food applications. WPI was initially used to bind FX and then interact
78
with Lyz. In this process, the effect of FX on complex coacervation of WPI and Lyz was studied
79
by theoretical simulation and experiment. Finally, the protective and release properties of FX in
80
this heteroprotein complex coacervates were evaluated.
81
MATERIALS AND METHODS
82
Materials and Chemicals. WPI (HilmarTM-9410, protein 92.9% dry basis) was obtained from
83
Hilmar Cheese Company, Inc. (Hilmar, CA, USA). Lyz from hen egg white (protein ≥ 95%, > 40000
84
units/mg protein), all-trans FX standard, and ANS-Na (≥ 97.0%, for fluorescence) were purchased
85
from Sigma Chemical Company (St. Louis, MO, USA). FX used in this study was isolated from
86
Undaria pinnatifida and quantified by 1260 HPLC (Agilent Technologies Inc., Santa Clara, CA,
87
USA).24 Solvents for HPLC including acetonitrile and MTBE were procured from Merck & Co.,
88
Inc. (Darmstadt, Germany). Pepsin from porcine gastric mucosa (10000 units/mg protein), trypsin
89
from porcine pancreas (250 units/mg proteins), and porcine bile extracts were purchased from
90
Beijing Solarbio Science & Technology Co., Ltd. (Beijing, China). All other chemicals were of
91
analytical grade and procured from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China).
92
Preparation of Protein Samples. WPI powder was dissolved at 0.2% (w/v) in deionized water
93
under a continuous mild stirring for 2 h at 25°C to allow complete rehydration,25 followed by
94
filtration through a 0.22 μm MCE syringe membrane filter to remove undissolved particulates. A
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
95
Lyz solution was prepared at 0.6% (w/v) in deionized water. These stock solutions were kept at 4 °C
96
until diluted to a final concentration and then used within two days.
97
Preparation of WPI–FX Complex. Nanocomplex comprising WPI and FX was prepared based on
98
our previous study.20 FX was pre-dissolved in ethanol at 0.5 mM, and various volumes of the FX
99
solution was respectively dropwise added to 3.0 mL of WPI (0.1%, w/v) followed by the vortex for
100
5 min, corresponding to a final FX concentration of 5.0 μM, 7.5 μM, 10.0 μM, 12.5 μM, and 15.0
101
μM. After standing for different time (0 min, 15 min, 60 min, and 120 min), the samples were
102
measured for fluorescence at three temperatures (300 K, 305 K, 310 K), size and ζ-potential at
103
ambient.
104
Steady-state Fluorescence. Intrinsic fluorescence of sample (with or without FX) was recorded by
105
a F-7000 fluorescence spectrophotometer (Hitachi Ltd., Tokyo, Japan). The excitation wavelength
106
was set at 280 nm, and the emission spectra were recorded in the range of 300–450 nm at room
107
temperature. The excitation and emission slit widths were fixed at 10 nm. To explore the
108
fluorescence quenching mechanism, the Stern−Volmer equation was used: 𝐹0
109
𝐹
= 1 + 𝐾SV[𝑄] = 1 + 𝑘q𝜏0[𝑄]
(1)
110
Where F0 and F represent the steady-state fluorescence intensities in the absence and presence of
111
quencher, respectively. KSV is the Stern–Volmer quenching constant (M−1). [Q] is the concentration
112
of the quencher (mol). kq is the bimolecular quenching constant (M−1 s−1), and τ0 is the average
113
lifetime of the molecule without any quencher (s), and the fluorescence lifetime of the biopolymer
114
is 10–8 s.26
115
When small molecules bind independently to a set of equivalent sites on a macromolecule, the
116
equilibrium between the free and bound molecules is given by the following equation,27
ACS Paragon Plus Environment
Page 6 of 40
Page 7 of 40
117
Journal of Agricultural and Food Chemistry
log
𝐹0 ― 𝐹 𝐹
(2)
= log𝐾𝑎 +𝑛log[𝑄]
118
Where Ka and n are the binding constant (M−1) and the number of binding sites, respectively.
119
For studying thermodynamics, the van’t Hoff equation is used for determining the enthalpy change
120
(ΔH°, kJ mol−1) and entropy change (ΔS°, J mol−1 K−1):
121
ln𝐾a = ―
∆𝐻○ 𝑅𝑇
+
∆𝑆○ 𝑅
(3)
122
where R is the gas constant and T is the temperature (K). Then, the free energy change (ΔG°, kJ
123
mol−1) is calculated according to the following equation:
124
∆𝐺○ = ∆𝐻○ ― 𝑇∆𝑆○
(4)
125
Extrinsic Fluorescence Probe. The PSH was assessed by fluorescence spectroscopy with an ANS
126
probe based on the method.28 The WPI solution (0.1%, w/v) with or without FX was diluted to
127
0.05%, 0.0375%, 0.025%, 0.0125% (w/v). Then, 20 μL of ANS-Na (8 mM) was added to 1.6 mL
128
of each sample, vortexed for 15 s, and kept in the dark for 5 min. The same concentration of protein
129
solution without ANS was determined as the control, and the free ANS was used as the blank. For
130
fluorescence measurement, the excitation wavelength was fixed at 390 nm, and the emission spectra
131
were recorded in the range of 400–600 nm. The excitation and emission slit widths were 5 and 10
132
nm, respectively. The scan speed was 1500 nm/min. The index of PSH was determined from the
133
initial slope of the plot of fluorescence intensity versus protein concentration. Within the protein
134
concentration range used in these experiments, linear relationships were obtained (R2 > 0.99).
135
Molecular Simulation. The molecular simulation was a complementary way to understand the
136
conformation transition and interaction of proteins and its ligand. Three major proteins in WPI,
137
including BSA, β-Lg, and α-La, were selected as model proteins for simulation. Initial
138
conformations of whey proteins with FX were generated by Autodock 4.2 based on our previous
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
139
study,20 and then simulated by MD using the Gromacs 5.1.2 package.29 The topological file of FX
140
with a force field of GROMOS96 54a7 was prepared by the ATB online server.30 Then, a cube box
141
containing each whey protein with its FX ligand was constructed, in which SPC-type water
142
molecules and counterions were filled. After constructing the box, the energy was optimized by the
143
steepest descent method. The maximum number of steps was 50000, and the optimization was to be
144
completed when the maximum force Fmax < 10.0 kJ mol−1. Then, the box was equilibrated under the
145
canonical and isothermal-isobaric ensembles. Upon completion of the two equilibration phases, the
146
system could release the position restraints and run MD with a time length of 20 ns for data
147
collection. Finally, the visualization of three-dimensional structure and electrostatic potential of
148
whey proteins before and after binding FX were carried out by PyMOL. The changes in the
149
secondary structure of the protein are calculated by do_dssp. The titration curve and pI were
150
computed using APBS,31 and H++ web server.32
151
Preparation of WPI–FX–Lyz Coacervates. To study the mass ratio on complex coacervation at
152
25°C, the WPI–FX (0.1%, w/v) solution was mixed with the solution of 0.025%, 0.05%, 0.075%,
153
0.1%, 0.15%, 0.2%, 0.3%, 0.4%, 0.5% (w/v) lysozyme at equal volumes. The solutions were
154
adjusted to pH 6.5 with NaOH or HCl (0.1 M). Then, to study the effect of pH on complex
155
coacervation, the pH 6.5 solutions were also adjusted to pH 5, 5.5, 6, 6.5, 7, 7.5, 8, 8.5, 9, 9.5, 10,
156
10.5, and 11 with NaOH and HCl (0.1 M). After pH adjustment and standing without vortex for 15
157
min, the samples were measured for size and ζ-potential. A mixed solution containing WPI and Lyz,
158
but not including FX, was used as a control.
159
Particles Size and ζ-Potential Measurement. The particle size and ζ-potential of one or two
160
proteins and those with FX were measured with a Zetasizer Nano ZS90 instrument (Malvern
ACS Paragon Plus Environment
Page 8 of 40
Page 9 of 40
Journal of Agricultural and Food Chemistry
161
Instruments Ltd., Malvern, Worcestershire, U.K.) assuming the scattering particles to be spherical,
162
their particles size was calculated from the diffusion parameters using Stokes-Einstein equation.
163
The electrophoretic mobility of each sample was determined, then calculating the ζ-potential using
164
the Smoluchowski approximation.33
165
FT-IR. Infrared spectra were obtained from the free FX, WPI and Lyz samples, lyophilized WPI-
166
FX, WPI-Lyz (1:2, w/w), and WPI-Lyz-FX (1:2, w/w). The analyses were performed with an FT-
167
IR spectrometer (Nicolet iS10, Thermo Scientific) using KBr (potassium bromide) and read in the
168
range of 4000–400 cm−1
169
Loading Efficiency of FX in WPI–FX and WPI–FX–Lyz. To determine the loading efficiency of
170
FX in nanocomplex and coacervates, the WPI–FX–Lyz solution was centrifuged at 12000 g and
171
4 °C for 15 min. The precipitate was collected and mixed with an equal volume of acetone/n-hexane
172
(1:1, v/v) solvent. For WPI–FX complex, the solution was centrifuged in Millipore (30 kDa, MWCO)
173
filters at 4000 g for 20 min and washed with deionized water 2–3 times to remove free FX. Then,
174
the concentrate was diluted to 3 mL and also mixed with an equal volume of acetone/n-hexane (1:1,
175
v/v) solvent. After vortex for 20 s, the upper organic phase was collected, concentrated under a
176
stream of nitrogen, and dissolved in acetonitrile for HPLC analysis.24 The molar amounts of FX in
177
the precipitate or concentrate (M, μmol) and that used in sample preparation (M0, μmol) were used
178
to calculate the loading efficiency as follows:
179
Loading efficiency % =
𝑀0 𝑀
× 100%
(5)
180
Stability of FX in WPI–FX and WPI–FX–Lyz. Samples with FX were contained in amber vials
181
and kept at 25 °C for up to 21 d or 75 °C in a water bath for up to 9 h in the dark. Samples were
182
taken periodically and mixed with an equal volume of acetone/n-hexane (1:1, v/v) solvent to extract
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
183
FX for calculation of residual amount. After vortex for 3 min, the upper organic phase was collected,
184
concentrated under a stream of nitrogen, and dissolved in acetonitrile for HPLC analysis. The
185
degradation of FX was evaluated for the zero, first, and second-order kinetic models using Equations
186
(4), (5), and (6), respectively.34
187
𝐹 = 𝐹0 ― 𝑘1𝑡
(6)
188
𝐹 = 𝐹0e ― 𝑘2𝑡
(7)
189
𝐹 = 1 + 𝑘3𝐹0𝑡
𝐹0
(8)
190
where F is the residual amount of FX (μmol) after a treatment time of t (d or h). F0 is the total
191
amount of FX (μmol) before incubation. k1, k2, and k3 are the respective rate constant at the
192
corresponding reaction order, d−1 or h−1.
193
Release Properties of FX after Simulated Digestions. To evaluate in vitro digestion of FX in
194
WPI–FX and WPI–FX–Lyz, the protocol was used to formulate simulated digestive juices and
195
digestion conditions with some modifications.1 A 4 mL of sample solutions containing a certain
196
amount of FX (F0, μmol) was first diluted with 10 mL of an electrolyte solution (pH 5.5) containing
197
120 mM NaCl, 5 mM KCl, and 6 mM CaCl2. The simulated gastric digestion was conducted at
198
37 °C for 2 h in a shaking water bath after mixing the diluted FX sample with 0.5 mL of pepsin
199
solution (0.075 g/mL dissolved in 0.1 M of HCl) and adjusting to pH 2.2 using HCl (10 mM). When
200
the stomach digestion was completed, FX in WPI–FX and WPI–FX–Lyz was extracted by
201
acetone/n-hexane (1:1, v/v). Its amount (F1, μmol) was determined using HPLC. The rate of FX
202
release (Rel) in the stomach was calculated as follows:
203 204
𝐹1
𝑅el(%) = (1 ― 𝐹0) × 100
(9)
The sample after the simulated gastric digestion was treated sequentially to simulate digestions at
ACS Paragon Plus Environment
Page 10 of 40
Page 11 of 40
Journal of Agricultural and Food Chemistry
205
three intestinal phases (duodenum, jejunum, and ileum) in the shaking water bath. To simulate the
206
duodenum phase, 14.5 mL of the sample after the stomach phase treatment was mixed with 250 mg
207
of porcine bile extract, 0.5 mL of pancreatic lipase (0.01 g/mL), and 0.5 mL of trypsin (0.08 g/mL).
208
The mixture was adjusted to pH 6.5 using NaHCO3 (10 mM) and incubated at 37 °C for 30 min. To
209
mimic the jejunum phase, the mixtures after the duodenum phase digestion was changed to pH 6.8
210
using NaHCO3 (10 mM), followed by incubation at 37 °C for 90 min. Subsequently, the mixture
211
was adjusted to pH 7.2 using NaHCO3 (10 mM) and incubated at 37 °C for 5 h to simulate the ileum
212
phase of digestion.
213
After each phase of the simulated intestinal digestion, the mixture was centrifuged at 12000 g for
214
20 min at 4 °C to collect the supernatant to determine the amount of released FX (F2, μmol) using
215
the HPLC. The percentage of FX released (Rel) was then calculated with respect to the total amount
216
used in the simulated digestion, which was consistent with the stomach digestion. 𝐹2
217
𝑅el(%) = 𝐹0 × 100
(10)
218
Statistical Analysis. Data were presented as the mean ± standard deviation of three separate
219
experiments. All statistical analysis was conducted using SPSS software (version 19.0, SPSS Inc.,
220
Chicago, IL, USA). The Shapiro–Wilk test was used to assess the normality of data and the Levene
221
test was used to check the homoscedasticity. The difference between samples was evaluated using
222
the one-way analysis of variance (ANOVA) and Duncan’s multiple comparisons. All statements of
223
significance were based on the 0.05 probability level.
224
RESULTS AND DISCUSSION
225
Fluorescence of WPI quenched by FX. Fluorescence spectroscopy is currently one of the most
226
reliable and convenient methods for studying the interaction between a protein and a ligand, wildly
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
227
applicated in food science, biophysics, and pharmacy. Commonly, the intrinsic fluorescence of a
228
protein is mainly derived from hydrophobic amino acid residues inside the protein molecule. The
229
fluorescence intensity of WPI in Figure S1 exhibited the maximum emission peaks at 335 nm,
230
mainly due to their Tyr and Trp residues.26 As the FX concentration increased, the fluorescence
231
intensity of WPI showed a progressive decrease at three temperatures, indicating that FX quenched
232
the intrinsic fluorescence of WPI mostly owing to molecular interaction resulting in a decrease in
233
the quantum yield of the fluorophore.20 Further, the quenching of WPI fluorescence was studied
234
using the Stern−Volmer model. The plot of F0/F versus FX concentration in Figure 1A showed good
235
linear relationships (R2 > 0.98) and the slope (KSV) decreased with an increasing temperature,
236
indicating that the fluorescence quenching of FX on WPI was static, which also meant that a stable
237
non-covalent complex was formed between WPI and FX because a high temperature usually leading
238
to dissociation of the unstable adduct.26 The bimolecular quenching constant kq calculated in Table
239
S1 also supported static quenching mechanism of WPI by FX, because the kq value far exceeded the
240
diffusion-controlled rate constant of various quenchers with a biopolymer (2.0 × 1010 M−1 s−1).35
241
Binding and thermodynamic behaviors of WPI with FX. After clarifying the static quenching
242
mechanism of WPI fluorescence by FX, the log[(F0–F)/F] versus log[Q] was plotted in Figure 1B
243
to evaluate the binding ability of WPI with FX. Obviously, three curves of binding affinity also
244
showed good linear relationships (R2 > 0.98). The binding constant Ka (300 K) of WPI to FX listed
245
in Table S1 was 3.38 × 104 M−1. This result had the same order (104 M−1) as the norbixin and bixin
246
that were bound to WPI,35,36 and higher than β-carotene (103 M−1)37, showing that WPI has the
247
stronger binding ability to FX than β-carotene. Further, the thermodynamic parameters calculated
248
in Table S1 demonstrated the binding process of FX to WPI was spontaneous owing to the negative
ACS Paragon Plus Environment
Page 12 of 40
Page 13 of 40
Journal of Agricultural and Food Chemistry
249
values of ΔG° (−26.2 kJ mol−1 at 300 K). This result was also higher than β-carotene (−22.2 kJ
250
mol−1 at 298 K), well agreeing with the binding constant result.37 The negative values of ΔH° and
251
ΔS° indicated the major forces of WPI–FX interaction were van der Waals force and hydrogen
252
bond,38 which were different from hydrophobic interactions driving the formation of a complex
253
between WPI and β-carotene because both the ΔH° and ΔS° were positive.38 This result might be
254
due to the fact that the hydroxyl groups of FX had the ability to form hydrogen bonds with the side
255
chain of WPI, but not for β-carotene.
256
Surface hydrophobicity of WPI after binding FX. As discussed above, WPI and FX can form a
257
complex by non-covalent interaction forces. Here, the surface hydrophobicity of WPI with or
258
without FX was further measured by ANS probe. In aqueous solution, ANS fluoresces very weakly,
259
but its quantum yield increases significantly upon binding to some hydrophobic regions of proteins,
260
so it is widely used to characterize surface exposure of hydrophobic patches.39 The results in Figure
261
1C indicate that the PSH of WPI increases with an increasing FX concentration, indicating a tighter
262
binding of ANS to WPI in the presence of FX. One possible reason is the binding of FX exposed
263
some of the previously buried hydrophilic regions of WPI, generating new high-affinity ANS
264
binding sites and increasing the number of bound ANS per each protein. This was similar to Shahlaei
265
et al. and Jia et al. who studied the binding of sertraline with human serum albumin and ferulic acid
266
with β-Lg, respectively.39,40 It was worth noting that if the surface hydrophobicity of protein was
267
increased, the internal hydrophobic pocket would be exposed, enhancing the hydrophobic
268
interaction between the protein molecules.20 This result might promote the self-aggregation of WPI,
269
and affect the complex coacervation with Lyz. This inference was further studied in the following
270
experiments.
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
271
Size and ζ-Potential Change Upon Forming WPI–FX Complex. As described above, the PSH of
272
WPI increased after binding FX, meaning some hydrophobic areas of WPI had been exposed.
273
Besides, the previous research reported some ionizable residues of protein may change their charge
274
state during the binding process.23 For complex coacervation of two oppositely charged polymers,
275
this phenomenon can affect the electrostatic interaction. So, this section investigated the effects of
276
FX concentration and standing time on the size and ζ-potential of WPI. The particle size in Figure
277
2A indicated that the native WPI (183 ± 3 nm) became larger (> 200 nm) as FX concentration
278
enhanced. In Figure 2B, as the preparation time prolonged, the size of WPI was further increased
279
and eventually stabilize at about 220 nm. The above results found WPI aggregated after binding
280
FX, supporting the inference from surface hydrophobicity change. On the other hand, the data in
281
Figure 2C showed the ζ-potential of natural WPI was −18.83 ± 0.39 mV, which did not change
282
significantly after adding ethanol. This was because the ethanol volume was controlled within 3%
283
of the total volume of the WPI solution, which substantially preserved the original conformation of
284
the protein.41 After adding 5 μM, 10 μM, and 15 μM of FX, the ζ-potential significantly decreased
285
to −23.90 ± 0.57, −25.87 ± 0.33 and −25.40 ± 0.94 mV, respectively, suggesting the change in
286
surface potential of WPI after binding FX. In Figure 2D, the ζ-potential of the WPI–FX was further
287
reduced after sample solution was allowed to stand for different time, and then became gentler
288
similar to the size change, indicating that the binding process was gradually stabilized.
289
MD simulation. As previously mentioned, the PSH, size, and ζ-potential of WPI changed when
290
interacting with FX. To further explore its mechanism in detail, the complex conformation obtained
291
by docking three main proteins (β-Lg, α-La, and BSA) of WPI with FX was used as the initial
292
conformation of the MD simulation, and the simulation time was 20 ns. Both pure whey proteins
ACS Paragon Plus Environment
Page 14 of 40
Page 15 of 40
Journal of Agricultural and Food Chemistry
293
and the protein–FX complexes were simulated in the meantime, which could display the dynamic
294
alterations of three proteins after the molecular recognition of FX. The variations of the RMSD
295
regarding β-Lg, α-La and BSA with or without FX were shown in Figure 3. Commonly, if the
296
fluctuations of RMSD value for a typical dynamic system was kept within 0.1 nm, the system could
297
be considered to reach a stable state of dynamic equilibrium.42 Obviously, the three non-covalent
298
whey protein–FX complexes became stable from 12 ns. Comparatively, the pure protein systems
299
were equilibrated with some mild fluctuations over a time period of 20 ns. When the MD simulations
300
were completed, equilibrium conformation of three complexes was also obtained in Figure 3. It was
301
noteworthy that hydrogen bonds between Leu87 in β-Lg, Ser286 in BSA and FX were formed, well
302
agreeing with the thermodynamic measurement as discussed earlier. Other amino acid residues with
303
hydrophobic side chain, e.g., Leu39, Val41, Ile71, Ala86, Met107 in β-Lg, Leu259, Ala290, Val432
304
in BSA, Phe31, Ala40, Ile41, Val42 in α-La, constituted the hydrophobic pockets to envelop the FX
305
molecule.
306
Then, in order to verify the hydrophobic region exposure of the WPI after binding FX, the changes
307
in the radius of gyration (Rg) and SASA during the MD simulation were calculated and shown in
308
Figure 4. The Rg is a physical quantity that describes the tightness of the protein structure. The larger
309
its value, the lower the protein density, also indicating that the protein has expanded.43 The data in
310
Figure 4A-4C showed the average Rg of complex systems was higher than the free whey proteins,
311
revealing the protein tightness was reduced and the structure transformed into a loose state caused
312
by FX binding. To further realize the variation of the radius of gyration, the SASA values of free
313
and FX-binding whey proteins were determined. Normally, during protein folding, the hydrophobic
314
amino acid residues tend to be buried inside the protein molecule to minimize the accessible surface
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
315
of their non-polar groups. Therefore, the surface hydrophobicity of proteins can be studied by
316
calculating their SASA, and there is a certain correlation between them.44 The data in Figure 4D-4F
317
showed the average SASA of systems containing FX was higher than that of the pure whey proteins,
318
indicating that the structure became loose after the protein was combined with FX, causing an
319
increase in SASA. Both SASA and Rg indicated that the structures of whey proteins were folded in
320
the presence of FX, exposing more hydrophobic areas.
321
Furthermore, time-dependent secondary structure fluctuation analysis provided some additional
322
information about protein structure upon binding a ligand. The secondary structures of β-Lg, BSA,
323
and α-La before and after binding FX were calculated and shown in Figure 4G-4I, respectively. In
324
the initial conformation of BSA, the main secondary structure was α-helix, accounting for 72.6%,
325
decreased to 69.5% after simulation, while the contents of random coil and β-turn increased from
326
20.9% to 23.2%. For β-Lg, the content of main β-sheet structure decreased from 40.7% to 34.0%.
327
Other structures, such as α-helix and random coil, accounted for 6.8% and 19.1%, respectively.
328
After MD, they were increased to 13.6% and 25.9%, respectively. Similar to BSA, the main
329
secondary structure of α-helix was also α-helix, whose content was reduced from 33.3% to 26.0%,
330
while the random coil increased from 18.7% to 25.2%. The above results showed that the major
331
secondary structure components of the whey proteins after binding FX were reduced, indicating the
332
presence of FX caused the conspicuous loosening and unfolding of polypeptide chain backbone
333
structure in whey proteins, resulting in exposure of the hydrophobic region. This also confirmed the
334
enhancement of PSH and particle size after spontaneous assembly of WPI and FX.
335
Further, to investigate the ζ-potential alteration of WPI, the electrostatic potential of whey proteins
336
with and without FX in Figure 5A, 5B, 5D, 5E, 5G, and 5H was calculated by APBS. The results
ACS Paragon Plus Environment
Page 16 of 40
Page 17 of 40
Journal of Agricultural and Food Chemistry
337
showed that the surface potential of the three whey proteins had undergone some changes when
338
interacting with the FX molecule, among which β-Lg and α-La were more obvious than BSA. In
339
addition, calculations of titration curves by H++ server in Figure 5C, 5F, and 5I showed the pI of β-
340
Lg and α-La reduced obviously after binding FX, while the change of BSA was minimal, indicating
341
the negative charges of β-Lg and α-La with FX were more than the parent proteins above a certain
342
pH of the isoelectric point. This result supported the ζ-potential change after adding FX to WPI and
343
revealed this change was mainly derived from β-Lg and α-La. Aguilar et al. also reported that
344
proteins often adjusted their conformation when they bound ligand, modifying the
345
microenvironment of the amino acids and affecting their pK values.45 In more detail, the electrically
346
charged amino acid residues of three whey proteins involved in interaction with FX molecule
347
(Figure 3) were selected to estimate their pKa change by pK_(1/2) calculation. In Figure S2, the
348
protonation probability of these residues against pH was plotted. The pK_(1/2) was the pH at which
349
the protonation probability was 0.5. In most cases, the protonation probability versus pH was
350
algebraically equivalent to the Henderson-Hasselbalch equation, in which case pK_(1/2) = pKa.46
351
The results showed that the pKa of the five charged residues from β-Lg (Lys69, Lys70, Asp85) and
352
α-La (His32, Asp37) participating in stabilizing FX were reduced, however, there was no change
353
for BSA (data now shown). This was also in line with changes in the isoelectric point of each whey
354
protein (Figure 5), indicating that the introduction of FX decreased the pKa of some amino acid
355
residues that interacted with it, leading to the shift of pI of the entire proteins.
356
Formation of complex coacervates between WPI–FX and Lyz. Based on the successful
357
formation of nanocomplex between WPI and FX, this study then constructed a FX-loaded vector
358
formed by WPI–FX and Lyz. These two proteins are commonly used proteins for complex
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
359
coacervation, one of which had an isoelectric point of 4.6 and the other had an isoelectric point
360
of about 11.0.46,47 As discussed above, WPI underwent a series of changes in its properties after
361
binding FX, such as increased particle size and reduced ζ-potential, which would inevitably affect
362
its complex coacervation with other proteins. Therefore, this experiment studied the effects of
363
WPI/Lyz mass ratio and pH on the heteroprotein complex coacervation process in the presence
364
and absence of FX. As shown in Figure 6A, at a fixed pH 6.5, the particle size of WPI–Lyz
365
continued to increase at WPI/Lyz ratios of 4:1–2:3 (w/w), then reached a plateau at ratios of 2:3–
366
1:5 (w/w) where the complex solutions have the highest size. This result indicated that complex
367
coacervation could occur between WPI and Lyz drove by electrostatic interaction at a suitable
368
mass ratio because pH 6.5 was between the pI of whey protein and lysozyme. After loading FX,
369
the size of WPI–FX–Lyz did not change significantly and also had the largest size at WPI/Lyz
370
ratios of 2:3–1:5 (w/w). The data of ζ-potential in Figure 6A showed something different. As the
371
WPI/Lyz ratio increased, the ζ-potentials of WPI–Lyz and WPI–FX–Lyz exhibited a progressive
372
rise and were close to 0 mV at the ratio from 2:3 (w/w) to 1:2 (w/w). However, after the addition
373
of FX, the negative charge carried by WPI increased, resulting in a lower potential of WPI–FX–
374
Lyz than WPI–Lyz at the same WPI/Lyz ratio. For example, under the WPI/Lyz condition of 1:2
375
(w/w), the ζ-potentials of WPI–Lyz and WPI–FX–Lyz were 1.24 ± 0.94 mV and 0.69 ± 0.05 mV,
376
respectively.
377
As previously mentioned, the pI of the WPI sample was 4.6, while the pI of Lyz sample was 11.0.
378
Thus, to study the effect of pH on the formation of the WPI–Lyz co-precipitates, all the pH values
379
were selected within the range of 5.0–11 to avoid the self-aggregation of a single protein and
380
ensure the opposite charge state of WPI and Lyz. The dynamic light scattering measurement in
ACS Paragon Plus Environment
Page 18 of 40
Page 19 of 40
Journal of Agricultural and Food Chemistry
381
Figure 6B showed the solutions of WPI–Lyz and WPI–FX–Lyz between pH 6.5 to 10.0 were
382
multiple dispersion systems and large aggregates existed. The particles of both systems exhibited
383
the micron levels, and there was no significant difference between them. Besides, the ζ-potential
384
of WPI–Lyz in Figure 6B showed close to zero at pH 8.0, which confirmed the larger size
385
particles produced by complex coacervation. However, after adding FX, there was a significant
386
change in the ζ-potential of complex coacervates formed by WPI–FX and Lyz. The pH of ζ-
387
potential closest to 0 mV had shifted to 7.5, which was due to the increased negative charge of
388
WPI. This was somewhat similar to the case where the Lyz concentration was fixed and the WPI
389
concentration was increased, resulting in a decrease in the pI of the complex coacervates.49
390
Combined with the results of the previous sections, this study indicated that the non-covalent
391
binding of FX to WPI altered the size and charge of whey proteins, which had no obvious effect
392
on the particle size of the complex coacervates subsequently formed with Lyz, but altered the
393
optimal pH of complex coacervation process.
394
FT-IR analysis was performed to evaluate the interaction between proteins and the encapsulation of
395
FX in complex coacervates. Infrared spectra of the single protein and the complex at the ratio of 1:2
396
were shown in Figure 7. The major FTIR spectra of WPI and Lyz were between the bands 1300
397
cm−1 and 1700 cm−1, corresponding to amides I, II, and III. In the FTIR spectrum of WPI–Lyz, the
398
similar structures were observed but the intensities were low. The reduction in the band (1300 cm−1
399
to 1700 cm−1) corresponding to amides was caused by the electrostatic interaction between the –
400
COO− (C=O) cluster of one protein and the –NH3+ (NH) cluster of the other protein.49 In addition,
401
the stretching of the N–H and O–H groups could be identified by the band near the 3300 cm−1. A
402
decrease in the strength of this band was still observed in the WPI–Lyz composite, indicating that
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
403
the formation of the complex coacervates not only involved electrostatic interaction but also
404
hydrogen bonding. In the spectrum of FX, the band at 1923 cm−1 was assigned as an allenic bond
405
(C=C=C), which was considered to be a representative group of FX. The bands at 3373 cm−1 and
406
1737 cm−1 were respectively identified as hydrogen bonded O−H stretching vibrations and the
407
ketones with −C=O bonds. Bands between 2800 and 3100 cm−1 showed the presence of alkanes
408
with C−H bonds. In addition, the bands at 970 cm−1 was the external torsional pendulum vibration
409
of the C−H bond in the double bond, which was the characteristic absorption band of the trans
410
substituted ethylene. In the spectrum of WPI–FX, the bands at 1637 and 1539 cm−1 was
411
corresponding to the carbonyl groups (C=O). However, the major bands in FX were not found in
412
this spectrum, indicating that the FX was well bound and encapsulated within WPI. Similarly, in
413
the spectrum of WPI–FX–Lyz, and the absorption peaks of FX could not be found. Taking all of
414
the results obtained by FT-IR together, it was concluded that FX had been successfully encapsulated
415
in heteroprotein complex coacervates.
416
Ability to deliver FX by WPI–FX–Lyz, and WPI–FX. Based on the above results, the
417
heteroprotein complex coacervates formed by WPI–FX and Lyz as a carrier for FX were constructed.
418
Then, the delivering ability of WPI–FX–Lyz (WPI/Lyz = 1:2, pH = 7.5), including loading
419
efficiency, chemical stability, and digestion in vitro, were investigated in the following experiments.
420
For comparison, WPI–FX was also used for the study. As shown in Figure 8A, when the amount of
421
added FX was 0.045 μmol, the loading efficiency of WPI–FX–Lyz was 82.36 ± 3.85%, which was
422
significantly higher than that of WPI–FX, corresponding to 55.60 ± 2.25%. This result indicated
423
that the spontaneous co-assembly of FX, WPI, and Lyz into coacervates could provide a higher FX-
424
loading capacity than WPI itself.
ACS Paragon Plus Environment
Page 20 of 40
Page 21 of 40
Journal of Agricultural and Food Chemistry
425
Subsequently, the chemical stabilities of WPI–FX–Lyz and WPI–FX were evaluated by the
426
degradation of FX under the heating and storage conditions. As shown in Figure 8B, for the thermal
427
stability test at 75 ℃, FX in the control group degraded 91.2% after 2 h of heating, while the WPI–
428
FX–Lyz and WPI–FX groups retained 74.4% and 20.9%, respectively. For storage stability test at
429
25 ℃ in Figure 8C, WPI–FX–Lyz group could still retain 63.4% of the total FX after storage for 2
430
d, while the WPI–FX group had degraded the vast majority, only 27.8% of the initial content was
431
left. These results indicated that WPI had a better protective effect on FX after self-assembled with
432
Lyz. Furthermore, the degradation dynamics illustrated that the degradation of FX in all samples
433
followed the second-order kinetic model (n = 2, R2> 0.96). The degradation rates k2 of FX in WPI–
434
FX–Lyz and WPI–FX were less than that of FX with no encapsulation, also indicating the excellent
435
protection property of WPI or WPI–Lyz carriers in the high-temperature or ambient environment.
436
Additionally, it could be found that the delivery system prepared by heteroprotein complex
437
coacervation had a better ability to prevent the degradation of FX than the carrier prepared by a
438
simple protein binding.
439
Finally, the release properties of WPI–FX–Lyz and WPI–FX in the stomach and various intestinal
440
segments were evaluated using the in vitro simulated digestion method. The digestion of FX-loaded
441
vector in the gastrointestinal tract was a complex process, which played a major role in the uptake,
442
distribution as well as metabolism of FX. The results in Figure 8D indicated that the encapsulated
443
FX was very stable and rarely released from the WPI–FX–Lyz and WPI–FX by the action of pepsin.
444
Nearly 90% of the total FX was still retained in the heteroprotein vehicle. The resistance of this
445
carrier toward pepsin digestion might be attributed to the presence of β-Lg above 50% of WPI. It
446
was well documented that the native β-Lg was resistant to hydrolysis in the gastric compartment
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
447
following simulated digestion owing to its compact globular structure.50 In addition, Lyz was
448
particularly stable to changes in pH, and there was no significant conformational transition in a
449
moderately dilute solution within the pH range of 1.2–11.3.51 In the digestion of small intestinal
450
phase, the duodenal incubation resulted in destabilization of the WPI–FX–Lyz and 62.56 ± 2.50%
451
of the encapsulated FX was released. However, after jejunal and ileal incubation, the release rate
452
increased to 66.76 ± 4.00% and 70.82 ± 2.83%, respectively. Yonekura and Nagao revealed that the
453
carotenoids in the food were mainly combined with proteins, which were released from the protein
454
complexes through the catalysis of the digestive enzyme in an animal, transformed into the
455
chylomicrons with the other lipids in the duodenum, and then absorbed by the cholesterol transfer
456
carrier or passive diffusion pathway.52 As an oxygenated carotenoid, the main release of FX
457
occurred in the small intestinal phase, which might be beneficial to the absorption process. Besides,
458
Fu et al. reported that β-Lg could be almost completely digested by pancreatic enzymes,53 which
459
might indicate that the heteroprotein complex coacervates formed by WPI and Lyz was unstable to
460
trypsin. All of these the clues helped explain the effectively cumulative release of FX in the small
461
intestine. It should be noted that the release rate of FX from WPI–FX in the intestine was
462
significantly lower than that from WPI–FX–Lyz, reconfirming the advantages and potential of
463
heteroprotein complex coacervation as the vector to deliver FX.
464 465
ABBREVIATIONS USED
466
α-La, α-lactalbumin; β-Lg, β-lactoglobulin; ANS-Na, 8-anilino-1-naphthalenesulfonic acid
467
ammonium salt; BSA, bovine serum albumin; FT-IR, Fourier transform infrared spectrometry; FX,
468
fucoxanthin; GRAS, generally recognized as safe; Lyz, lysozyme; MD, molecular dynamics;
ACS Paragon Plus Environment
Page 22 of 40
Page 23 of 40
Journal of Agricultural and Food Chemistry
469
MTBE, methyl tert-butyl ether; PSH, protein surface hydrophobicity; RMSD, Root-mean-square
470
deviation; SASA, solvent accessible surface area; WPI, whey protein isolate; WPI–FX, FX-
471
binding WPI; WPI–FX–Lyz, complex coacervates of FX-binding WPI and Lyz; WPI–Lyz,
472
complex coacervates of WPI and Lyz
473 474
ACKNOWLEDGMENTS
475
This work was funded by National Natural Science Foundation of China (31871786), Zhejiang
476
Public Welfare Technology Application Research Project (2018C37023), and Open Project of
477
Key Laboratory of Sustainable Utilization of Technology Research for Fishery Resource of
478
Zhejiang Province.
479 480
SUPPORTING INFORMATION DESCRIPTION
481
Stern–Volmer and binding parameters for the interaction of FX with WPI at three temperatures
482
(Table S1). Fluorescence emission spectra of whey protein isolate with different concentrations of
483
fucoxanthin at 300 K, 305 K, and 310 K (Figure S1). Amino acid residues with electrically charged
484
side chains involved in the interaction of whey proteins and fucoxanthin (Figure S2).
485 486
REFERENCES
487
(1) Koo, S. Y.; Mok, I. K.; Pan, C. H.; Kim, S. M. Preparation of fucoxanthin-loaded nanoparticles
488
composed of casein and chitosan with improved fucoxanthin bioavailability. J. Agric. Food
489
Chem. 2016, 64, 9428−9435.
490
(2) Miyashita, K.; Nishikawa, S.; Beppu, F.; Tsukui, T.; Abe, M.; Hosokawa, M. The allenic
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
491
carotenoid fucoxanthin, a novel marine nutraceutical from brown seaweeds. J. Sci. Food Agr.
492
2011, 91, 1166−1174.
493
(3) Ravi, H.; Baskaran, V. Biodegradable chitosan-glycolipid hybrid nanogels: A novel approach
494
to encapsulate fucoxanthin for improved stability and bioavailability. Food Hydrocolloid.
495
2015, 43, 717−725.
496 497
(4) D’Orazio, N.; Gemello, E.; Gammone, M.; de Girolamo, M.; Ficoneri, C.; Riccioni, G. Fucoxantin: A treasure from the sea. Mar. Drugs 2012, 10, 604−616.
498
(5) Xiang, S.; Liu, F.; Lin, J.; Chen, H.; Huang, C.; Chen, L.; Zhou, Y.; Ye, L.; Zhang, K.; Jin, J.;
499
Zhen, J.; Wang, C.; He, S.; Wang, Q.; Cui, W.; Zhang, J. Fucoxanthin inhibits β-amyloid
500
assembly and attenuates β-amyloid oligomer-induced cognitive impairments. J. Agric. Food
501
Chem. 2017, 65, 4092−4102.
502
(6) Salvia-Trujillo, L.; Sun, Q.; Um, B. H.; Park, Y.; McClements, D. J. In vitro and in vivo study
503
of fucoxanthin bioavailability from nanoemulsion-based delivery systems: Impact of lipid
504
carrier type. J. Funct. Foods 2015, 17, 293−304.
505
(7) Dai, J.; Kim, S. M.; Shin, I. S.; Dai Kim, J.; Lee, H. Y.; Shin, W. C.; Kim, J. C. Preparation
506
and stability of fucoxanthin-loaded microemulsions. J. Ind. Eng. Chem. 2014, 20, 2103−2110.
507
(8) Bhatti, H. S.; Khalid, N.; Uemura, K.; Nakajima, M.; Kobayashi, I. Formulation and
508
characterization of food grade water‐in‐oil emulsions encapsulating mixture of essential amino
509
acids. Eur. J. Lipid Sci. Tech. 2017, 119, 1600202.
510
(9) Bouyer, E.; Mekhloufi, G.; Rosilio, V.; Grossiord, J. L.; Agnely, F. Proteins, polysaccharides,
511
and their complexes used as stabilizers for emulsions: alternatives to synthetic surfactants in
512
the pharmaceutical field?. Int. J. Pharmaceut. 2012, 436, 359−378.
ACS Paragon Plus Environment
Page 24 of 40
Page 25 of 40
513 514 515 516 517 518
Journal of Agricultural and Food Chemistry
(10) Chen, L.; Remondetto, G. E.; Subirade, M. Food protein-based materials as nutraceutical delivery systems. Trends Food Sci. Tech. 2006, 17, 272−283. (11) Elzoghby, A. O.; Samy, W. M.; Elgindy, N. A. Protein-based nanocarriers as promising drug and gene delivery systems. J. Control. Release 2012, 161, 38−49. (12) Kimpel, F.; Schmitt, J. J. Milk proteins as nanocarrier systems for hydrophobic nutraceuticals. J. Food Sci. 2015, 80, R2361−R2366.
519
(13) Chapeau, A. L.; Tavares, G. M.; Hamon, P.; Croguennec, T.; Poncelet, D.; Bouhallab, S.
520
Spontaneous co-assembly of lactoferrin and β-lactoglobulin as a promising biocarrier for
521
vitamin B9. Food Hydrocolloid. 2016, 57, 280−290.
522
(14) Tavares, G. M.; Croguennec, T.; Lê, S.; Lerideau, O.; Hamon, P.; Carvalho, A. F.; Bouhallab,
523
S. Binding of folic acid induces specific self-aggregation of lactoferrin: thermodynamic
524
characterization. Langmuir 2015, 31, 12481−12488.
525
(15) Chapeau, A. L.; Hamon, P.; Rousseau, F.; Croguennec, T.; Poncelet, D.; Bouhallab, S. Scale-
526
up production of vitamin loaded heteroprotein coacervates and their protective property. J.
527
Food Eng. 2017, 206, 67−76.
528
(16) Jiang, C. M.; Wang, M. C.; Chang, W. H.; Chang, H. M. Isolation of lysozyme from hen egg
529
albumen by alcohol-insoluble cross-linked pea pod solid ion-exchange chromatography. J.
530
Food Sci. 2001, 66, 1089−1093.
531
(17) Diarrassouba, F.; Remondetto, G.; Garrait, G.; Alvarez, P.; Beyssac, E.; Subirade, M. Self-
532
assembly of β-lactoglobulin and egg white lysozyme as a potential carrier for nutraceuticals.
533
Food Chem. 2015, 173, 203−209.
534
(18) Yang, M. C.; Guan, H. H.; Liu, M. Y.; Lin, Y. H.; Yang, J. M.; Chen, W. L.; Chen, C. J.; Mao,
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
535
S. J. Crystal structure of a secondary vitamin D3 binding site of milk β‐lactoglobulin. Proteins.
536
2008, 71, 1197−1210.
537
(19) Diarrassouba, F.; Garrait, G.; Remondetto, G.; Alvarez, P.; Beyssac, E.; Subirade, M. Food
538
protein-based microspheres for increased uptake of vitamin D3. Food Chem. 2015, 173,
539
1066−1072.
540
(20) Zhu, J.; Sun, X.; Wang, S.; Xu, Y.; Wang, D. Formation of nanocomplexes comprising whey
541
proteins and fucoxanthin: Characterization, spectroscopic analysis, and molecular docking.
542
Food Hydrocolloid. 2017, 63, 391−403.
543
(21) Wagner, D.; Rousseau, D.; Sidhom, G.; Pouliot, M.; Audet, P.; Vieth, R. Vitamin D3
544
fortification, quantification, and long-term stability in Cheddar and low-fat cheeses. J. Agric.
545
Food Chem. 2008, 56, 7964−7969.
546
(22) Tavares, G. M.; Croguennec, T.; Hamon, P.; Carvalho, A. F.; Bouhallab, S. How the presence
547
of a small molecule affects the complex coacervation between lactoferrin and β-lactoglobulin.
548
Int. J. Biol. Macromol. 2017, 102, 192−199.
549 550
(23) Onufriev, A. V.; Alexov, E. Protonation and pK changes in protein–ligand binding. Q. Rev. Biophys. 2013, 46, 181−209.
551
(24) Zhu, J.; Sun, X.; Chen, X.; Wang, S.; Wang, D. Chemical cleavage of fucoxanthin from
552
Undaria pinnatifida and formation of apo-fucoxanthinones and apo-fucoxanthinals identified
553
using LC-DAD-APCI-MS/MS. Food Chem. 2016, 211, 365−373.
554
(25) Hébrard, G.; Hoffart, V.; Cardot, J. M.; Subirade, M.; Beyssac, E. Development and
555
characterization of coated-microparticles based on whey protein/alginate using the
556
Encapsulator device. Drug Dev. Ind. Pharm. 2013, 39, 128−137.
ACS Paragon Plus Environment
Page 26 of 40
Page 27 of 40
557 558
Journal of Agricultural and Food Chemistry
(26) Quenching of Fluorescence. In: Principles of fluorescence spectroscopy; Lakowicz, J. R., Eds.; Springer: Boston, MA, 2006, pp. 277−330.
559
(27) Delavari, B.; Saboury, A. A.; Atri, M. S.; Ghasemi, A.; Bigdeli, B.; Khammari, A.; Maghami,
560
P.; Moosavi-Movahedi, A. A.; Haertlé, T.; Goliaei, B. Alpha-lactalbumin: A new carrier for
561
vitamin D3 food enrichment. Food Hydrocolloid. 2015, 45, 124−131.
562
(28) Sun, C.; Wu, T.; Liu, R.; Liang, B.; Tian, Z.; Zhang, E.; Zhang, M. Effects of superfine grinding
563
and microparticulation on the surface hydrophobicity of whey protein concentrate and its
564
relation to emulsions stability. Food Hydrocolloid. 2015, 51, 512−518.
565
(29) Pronk, S.; Páll, S.; Schulz, R.; Larsson, P.; Bjelkmar, P.; Apostolov, R.; Shirts, M. R.; Smith,
566
J. C.; Kasson, P. M.; van der Spoel, D.; Hess, B.; Lindahl, E. GROMACS 4.5: a high-
567
throughput and highly parallel open source molecular simulation toolkit. Bioinformatics 2013,
568
29, 845−854.
569
(30) Malde, A. K.; Zuo, L.; Breeze, M.; Stroet, M.; Poger, D.; Nair, P. C.; Oostenbrink, C.; Mark,
570
A. E. An automated force field topology builder (ATB) and repository: version 1.0. J. Chem.
571
Theory Comput. 2011, 7, 4026−4037.
572
(31) Baker, N. A.; Sept, D.; Joseph, S.; Holst, M. J.; McCammon, J. A. Electrostatics of
573
nanosystems: application to microtubules and the ribosome. P. Natl. Acad. Sci. USA 2001, 98,
574
10037−10041.
575
(32) Anandakrishnan, R.; Aguilar, B.; Onufriev, A. V. H++ 3.0: automating pK prediction and the
576
preparation of biomolecular structures for atomistic molecular modeling and simulations.
577
Nucleic Acids Res. 2012, 40, W537−W541.
578
(33) Alexander, M.; Dalgleish, D. G. Dynamic light scattering techniques and their applications in
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
579
food science. Food Biophys. 2006, 1, 2−13.
580
(34) Colle, I. J.; Lemmens, L.; Knockaert, G.; Van Loey, A.; Hendrickx, M. Carotene degradation
581
and isomerization during thermal processing: a review on the kinetic aspects. Crit. Rev. Food
582
Sci. 2016, 56, 1844−1855.
583 584
(35) Zhang, Y.; Zhong, Q. Probing the binding between norbixin and dairy proteins by spectroscopy methods. Food Chem. 2013, 139, 611−616.
585
(36) Zhang, Y.; Zhong, Q. Binding between bixin and whey protein at pH 7.4 studied by
586
spectroscopy and isothermal titration calorimetry. J. Agric. Food Chem. 2012, 60, 1880−1886.
587
(37) Allahdad, Z.; Varidi, M.; Zadmard, R.; Saboury, A. A.; Haertlé, T. Binding of β-carotene to
588
whey proteins: Multi-spectroscopic techniques and docking studies. Food Chem. 2019, 277,
589
96−106.
590 591
(38) Ross, P. D.; Subramanian, S. Thermodynamics of protein association reactions: forces contributing to stability. Biochemistry-US, 1981, 20, 3096−3102.
592
(39) Shahlaei, M.; Rahimi, B.; Nowroozi, A.; Ashrafi-Kooshk, M. R.; Sadrjavadi, K.; Khodarahmi,
593
R. Exploring binding properties of sertraline with human serum albumin: Combination of
594
spectroscopic and molecular modeling studies. Chem-biol. Interact. 2015, 242, 235−246.
595
(40) Jia, J.; Gao, X.; Hao, M.; Tang, L. Comparison of binding interaction between β-lactoglobulin
596
and three common polyphenols using multi-spectroscopy and modeling methods. Food Chem.
597
2017, 228, 143−151.
598
(41) Chattoraj, S.; Mandal, A. K.; Bhattacharyya, K. Effect of ethanol-water mixture on the
599
structure and dynamics of lysozyme: a fluorescence correlation spectroscopy study. J. Chem.
600
Phys. 2014, 140, 03B619_1.
ACS Paragon Plus Environment
Page 28 of 40
Page 29 of 40
Journal of Agricultural and Food Chemistry
601
(42) Peng, W.; Ding, F.; & Peng, Y. K. In vitro evaluation of the conjugations of neonicotinoids
602
with transport protein: photochemistry, ligand docking and molecular dynamics studies. RSC
603
Adv. 2016, 6, 1826−1843.
604
(43) Gholami, S.; Bordbar, A. K.; Akvan, N.; Parastar, H.; Fani, N.; Gretskaya, N. M.; Bezuglovc,
605
V. V.; Haertlé, T. Binding assessment of two arachidonic-based synthetic derivatives of
606
adrenalin with β-lactoglobulin: Molecular modeling and chemometrics approach. Biophys.
607
Chem. 2015, 207, 97−106.
608
(44) Heldt, C. L.; Zahid, A.; Vijayaragavan, K. S.; Mi, X. Experimental and computational surface
609
hydrophobicity analysis of a non-enveloped virus and proteins. Colloid. Surface. B 2017, 153,
610
77−84.
611
(45) Aguilar, B.; Anandakrishnan, R.; Ruscio, J. Z.; Onufriev, A. V. Statistics and physical origins
612
of pK and ionization state changes upon protein-ligand binding. Biophys. J. 2010, 98, 872−880.
613
(46) Onufriev, A.; Case, D. A.; Ullmann, G. M. A novel view of pH titration in biomolecules.
614 615 616 617 618
Biochemistry-US 2001, 40, 3413−3419. (47) Zeeb, B.; Mi-Yeon, L.; Gibis, M.; Weiss, J. Growth phenomena in biopolymer complexes composed of heated WPI and pectin. Food Hydrocolloid. 2018, 74, 53−61. (48) Wang, J.; Sun, H.; Li, J.; Dong, D.; Zhang, Y.; Yao, F. Ionic starch-based hydrogels for the prevention of nonspecific protein adsorption. Carbohyd. Polym. 2015, 117, 384−391.
619
(49) Santos, M. B.; de Carvalho, C. W. P.; Garcia-Rojas, E. E. Heteroprotein complex formation of
620
bovine serum albumin and lysozyme: Structure and thermal stability. Food Hydrocolloid. 2018,
621
74, 267−274.
622
(50) Sari, T. P.; Mann, B.; Kumar, R.; Singh, R. R. B.; Sharma, R.; Bhardwaj, M.; Athira, S.
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
623
Preparation and characterization of nanoemulsion encapsulating curcumin. Food Hydrocolloid.
624
2015, 43, 540−546.
625
(51) Radford, S. E.; Woolfson, D. N.; Martin, S. R.; Lowe, G.; Dobson, C. M. A three-disulphide
626
derivative of hen lysozyme. Structure, dynamics and stability. Biochem. J. 1991, 273, 211−217.
627
(52) Yonekura, L.; Nagao, A. Intestinal absorption of dietary carotenoids. Mol. Nutr. Food Res.
628
2007, 51, 107−115.
629
(53) Fu, T. J.; Abbott, U. R.; Hatzos, C. Digestibility of food allergens and nonallergenic proteins
630
in simulated gastric fluid and simulated intestinal fluid a comparative study. J. Agric. Food
631
Chem. 2002, 50, 7154−7160.
632 633
Figure captions
634
Figure 1. Stern–Volmer (A) and binding model (B) calculated by intrinsic fluorescence of whey
635
protein isolate (WPI) quenched with different concentrations of fucoxanthin (FX, 5–15 μM) at three
636
temperatures (300, 305, 310 K). Plots of ANS fluorescence intensity versus concentration of WPI
637
with or without FX (C). Protein surface hydrophobicity (PSH) is determined the slope of the linear
638
regression model.
639 640
Figure 2. Particle size (A) and ζ-potential (C) measured immediately after vortex mixing different
641
concentrations of fucoxanthin (FX) in whey protein isolate (WPI) solution (0.1%, w/v). Effects of
642
standing time on particle size (B) and ζ-potential (D) of WPI solution (0.1%, w/v) upon adding FX.
643
The same letters in all figures represent no significant difference (p > 0.05).
644
ACS Paragon Plus Environment
Page 30 of 40
Page 31 of 40
Journal of Agricultural and Food Chemistry
645
Figure 3. RMSD for the backbone of β-lactoglobulin (A), bovine serum albumin (B), and α-
646
lactalbumin (C) with fucoxanthin (FX) and their detailed interaction after molecular dynamic
647
simulation. The green dash lines represent hydrogen bonds, and the red eyelash models represent
648
hydrophobic interactions between FX and whey proteins.
649 650
Figure 4. Changes in radius of gyration (Rg), solvent accessible surface area (SASA), and secondary
651
structures of β-lactoglobulin (A, D, G), bovine serum albumin (B, E, H), and α-lactalbumin (C, F,
652
I) with fucoxanthin (FX) during process of molecular dynamics.
653 654
Figure 5. Positive (blue) and negative electrostatic potential (red) for whey proteins before and after
655
binding fucoxanthin (FX) calculated by APBS: (A) free β-lactoglobulin, (B) FX-binding β-
656
lactoglobulin, (D) free α-lactalbumin, (E) FX-binding α-lactalbumin, (G) free bovine serum albumin,
657
(H) FX-binding bovine serum albumin. Titration curves for β-lactoglobulin (C), α-lactalbumin (F),
658
and bovine serum albumin (I) with or without FX were also calculated using the H++ web server.
659 660
Figure 6. Size and ζ-potential of heteroprotein complex coacervates formed by whey protein isolate
661
(WPI) and lysozyme (Lyz) in the presence and absence of fucoxanthin (FX): effect of WPI/Lyz
662
ratio at a fixed pH of 6.5 (A), and effect of pH at a fixed WPI/Lyz ratio of 1:2 (B).
663 664
Figure 7. FT-IR of free fucoxanthin (FX), whey protein isolate (WPI), lysozyme (Lyz), FX-binding
665
WPI (WPI–FX), complex coacervates of WPI and Lyz (WPI–Lyz), and complex coacervates of
666
WPI–FX and Lyz (WPI–FX–Lyz).
ACS Paragon Plus Environment
Journal of Agricultural and Food Chemistry
667 668
Figure 8. Loading efficiency (A), thermal stability (B), storage stability (C), and digestion property
669
(D) of fucoxanthin-binding whey protein isolate (WPI–FX) and complex coacervates formed by
670
WPI–FX and lysozyme (Lyz). ** represents the significant at p