You are on page 1of 10

Food Hydrocolloids 112 (2021) 106467

Contents lists available at ScienceDirect

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

Air-water interfacial and foaming properties of whey protein - sinapic


acid mixtures
Jack Yang a, b, Sarah P. Lamochi Roozalipour b, Claire C. Berton-Carabin c, d,
Constantinos V. Nikiforidis e, Erik van der Linden a, b, Leonard M.C. Sagis b, *
a
TiFN, Nieuwe Kanaal 9A, 6709PA, Wageningen, the Netherlands
b
Laboratory of Physics and Physical Chemistry of Foods, Wageningen University, Bornse Weilanden 9, 6708WG, Wageningen, the Netherlands
c
Laboratory of Food Process Engineering, Wageningen University, Bornse Weilanden 9, 6708WG, Wageningen, the Netherlands
d
INRAE, UR BIA, F-44316, Nantes, France
e
Laboratory of Biobased Chemistry and Technology, Wageningen University, Bornse Weilanden 9, 6708WG, Wageningen, the Netherlands

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

Keywords: Phenols are widely present in plants and often are co-extracted in plant protein extracts, while such components
Whey protein could influence the protein’s interface and foam stabilising properties. In this study, the influence of rapeseed
Sinapic acid phenol sinapic acid (SA) on the interfacial and foaming properties of a well characterised model protein, whey
Protein-phenol
protein isolate (WPI), was investigated. WPI formed strong viscoelastic interfacial layers, and addition of SA
Surface rheology
Atomic force microscopy
reduced the surface dilatational modulus by 25%. Turning SA into its active oxidised form in the WPI-SA mix­
Foam tures led to protein aggregation, resulting in a further decrease of the modulus by 40%. Removal of unbound
phenols induced a slight increase of the dilatational modulus, but the interfacial layer strength did not fully
recover to that made with pure WPI, suggesting binding of phenols to proteins, and thus influencing the protein
interface stabilising properties. Foams stabilised by WPI-SA mixtures had a shorter foam half-life time (130–180
min) than foams stabilised by pure WPI (260 min). Our data are thus in line with the observation that a lower
surface dilatational modulus leads to a lower foam stability. Yet, the foam stability did not recover to the original
values of pure WPI-stabilised foams after removal of unbound phenols. In conclusion, the presence of SA resulted
in a decrease in interfacial layer strength and foam stability. We conclude that in producing protein extracts from
rapeseed or other phenol-rich plant sources, the presence and oxidation of phenols should thus be considered.

1. Introduction are not yet well understood, while it is key to understand such in­
teractions in the utilization of phenol-rich plant sources in food systems.
One of the challenges in utilizing plant protein extracts as functional This work focusses on the main phenol present in rapeseed, which is
food ingredients is the naturally present phenols. For example, rapeseed sinapic acid (SA, Fig. 1), and its influence on interfacial and foaming
proteins have promising emulsifying (Cheung, Wanasundara, & Nick­ properties of proteins. Instead of rapeseed proteins, whey protein isolate
erson, 2014; Ntone, Bitter, & Nikiforidis, 2020; Tan, Mailer, Blanchard, (WPI) was chosen as a model protein system. Whey proteins were
& Agboola, 2011), foaming (Malabat, nchez-Vioque, Rabiller, & Gu already included as model proteins in previous protein-phenol studies,
guen, 2001; J. Yang, Faber, et al., 2020), and gelling properties (Tan, and has well known interfacial properties (Cao & Xiong, 2017; Jia et al.,
Mailer, Blanchard, Agboola, & Day, 2014), but rapeseed may contain up 2016; Jiang, Zhang, Zhao, & Liu, 2018; Jin et al., 2012; Li et al., 2020;
to 3% of phenols (Wanasundara, 2011). Rapeseed phenols are known to Rodríguez, von Staszewski, & Pilosof, 2015; von Staszewski, Pizones
contribute to a bitter taste and astringency in foods, loss of protein di­ Ruiz-Henestrosa, & Pilosof, 2014). Additionally, the removal of phenols
gestibility, and also colour changes of the protein extracts (Cai, Arnt­ from rapeseed proteins requires multiple washing steps with organic
field, & Charlton, 1999; Naczk, Amarowicz, & Shahidi, 1998; Naczk, solvents, resulting in a phenol removal of 75–90% and unknown effects
Oickle, Pink, & Shahidi, 1996). The effects of rapeseed phenols on on the protein structure and properties (Das Purkayastha et al., 2014; Xu
protein functionality (e.g., gelling, emulsifying, and foaming properties) & Diosady, 2002).

* Corresponding author.
E-mail address: leonard.sagis@wur.nl (L.M.C. Sagis).

https://doi.org/10.1016/j.foodhyd.2020.106467
Received 14 September 2020; Received in revised form 3 November 2020; Accepted 5 November 2020
Available online 9 November 2020
0268-005X/© 2020 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
J. Yang et al. Food Hydrocolloids 112 (2021) 106467

Protein-phenol interactions are dictated by the chemical structure of increased, which was probably related to the blocking of lamellae and
the phenols, such as degree of polymerization, conformational flexi­ plateau borders of the foam by aggregates, which reduced the rate of
bility, and hydrophobicity (Seczyk, Swieca, Kapusta, & Gawlik-Dziki, liquid drainage (Cao et al., 2018; Chen et al., 2017; Jia et al., 2016).
2019; Shahidi & Senadheera, 2019). These chemical properties can Another work showed that lower protein surface hydrophobicity after
lead to non-covalent and/or covalent interactions with the proteins. The non-covalent interaction with phenols resulted in higher protein solu­
non-covalent interactions may occur via hydrogen bonding, van der bility, and thus an increase in foam height and stability (Jiang et al.,
Waals forces, electrostatic and hydrophobic interactions (Ozdal, Capa­ 2018). These studies again confirm the complexity of phenol-protein
noglu, & Altay, 2013). Xu et al. showed that 50% of phenols in rapeseed mixtures, as the type of phenol and interaction with the protein deter­
interacted non-covalently with rapeseed proteins at pH 12. Another 40% mine the interface and foam stabilising properties of the mixtures. Also,
exist as unbound phenols in the aqueous solution, and the remaining the influence of SA on proteins with regards to protein functionality (e.g.
10% of the phenols interact covalently with the proteins (Xu & Diosady, stabilisation of interfaces and foams) has not been reported yet. There­
2000). Covalent interactions between phenols and proteins usually fore, the objective of this work was to determine the effect of SA on
occur upon oxidation of the phenols. Hydroxyl groups on phenyl rings interface and foam stabilising properties of WPI. Based on earlier reports
can be converted into quinones by enzymes or auto-oxidation, and the on protein-phenol studies, non-covalent interactions between the WPI
latter is accelerated at alkaline pH or in the presence of oxidizing agents and SA were expected, and covalent interactions were also induced in
(Czubinski & Dwiecki, 2017; Keppler, Schwarz, & van der Goot, 2020). this study by incubation at pH 10. An insight of these interactions and
A similar reaction was also found for SA, as the carboxyl and hydroxyl the link to protein functionality is also useful with regard to the pro­
groups were ionized at a pH above 7. Two oxidised SA could form the cessing of plant proteins, as non-covalent interactions between phenol
molecule thomasidioic acid, and can be further converted into quinones and protein occur directly upon solubilisation of the plant material. A
(Cai et al., 1999) (Fig. 1). Quinones are electrophilic compounds that are conventional protein extraction step is adjusting the pH to alkaline
stabilised by their conjugated cyclic dione structure, and thus very conditions (8− 10) to increase protein solubility, but also induces co­
reactive towards free thiol and amino groups on the proteins. After a valent interactions. This study can give an in-dept understanding of the
Michael addition reaction, a covalent bond between quinone and pro­ influence of such processing steps on the protein functionality, of which
tein is formed (Fig. 1) The phenols on the protein-phenol complex can be air-water interfacial and foaming properties were studied in this work.
oxidised a second time, and thereby crosslink another protein, and could The air-water interfacial properties were studied using large amplitude
ultimately lead to the formation of protein-phenol aggregates (Keppler oscillatory dilatation (LAOD). The protein structures were evaluated
et al., 2020; Kroll, Rawel, & Rohn, 2003). Such complexation into ag­ using differential scanning calorimetry (DSC). The foaming properties of
gregates was demonstrated for several protein sources, and often the WPI/SA mixtures were analysed by monitoring the foam stability
resulted in a decreased protein solubility (Jia et al., 2016; Karefyllakis, and bubble size distribution.
Salakou, Bitter, van der Goot, & Nikiforidis, 2018; Pringent, Voragen,
Visser, van Koningsveld, & Gruppen, 2007; Wei, Yang, Fan, Yuan, & 2. Experimental section
Gao, 2015).
Both non-covalent and covalent phenol-protein interactions can 2.1. Materials
largely influence the protein structure, and thus protein functionality.
First of all, bound phenols might influence the protein surface hydro­ Whey protein isolate (WPI) consisting of 74% β-lactoglobulin, 12.5%
phobicity by several mechanisms, such as the introduction of hydro­ α-lactalbumin, 5.5% bovine serum albumin, and 5.5% IgG1 (Butré,
philic groups from the phenol, or covering of hydrophobic residues on Wierenga, & Gruppen, 2012) (purity 98%) (Davisco Food international,
the protein (Wei et al., 2015; W.; Yang, Liu, Xu, Yuan, & Gao, 2014). France), sinapic acid (SA, 3,5-dimethoxy-4-hydroxy-cinnamic acid)
Additionally, phenol interaction could alter protein structure, and (Sigma-Aldrich, USA), and other chemicals (Sigma-Aldrich, USA) were
thereby exposing hydrophobic residues. A decrease of protein surface all used as received. The samples were prepared in ultrapure water
hydrophobicity was shown for whey proteins upon interaction with (MilliQ Purelab Ultra, Darmstadt, Germany).
phenols (Cao & Xiong, 2017; Cao, Xiong, Cao, & True, 2018; Jiang et al.,
2018; Li et al., 2020). The alteration of protein structure is another
major change in proteins upon phenol binding, which was also 2.2. Sample preparation
confirmed for milk proteins, as non-covalent interactions with phenols
resulted in an increase of unordered secondary protein structures (Cao 2.2.1. Preparation of WPI solution
et al., 2018; Cao & Xiong, 2017; Zhang et al., 2014). WPI was dissolved at 2.5% (w/w) in a sodium phosphate buffer (20
Alteration of a protein’s surface hydrophobicity, structure, and size mM, pH 7.0) for 4 h at room temperature and centrifuged at 16,000×g
may considerably influence its interface and foam stabilising properties. for 30 min. The supernatant was filtered over a 0.45 μm syringe filter,
For instance, phenols that were able to induce aggregation of WPI, and the filtrate was diluted based on dry matter.
resulted in lower surface activity and dilatational moduli compared to
pure WPI (Cao et al., 2018; Rodríguez et al., 2015). Even though the 2.2.2. Preparation of sinapic acid solution
interfacial properties were affected negatively, the foam stability A stock solution of 2% (w/w) sinapic acid (SA) was prepared in
buffer. The pH was corrected with NaOH to pH 7.0 and stirred for 4 h at

Fig. 1. A simplified mechanism of sinapic acid oxidation and covalent binding with proteins.

2
J. Yang et al. Food Hydrocolloids 112 (2021) 106467

room temperature. The pH was adjusted to 7.0 every 15 min, which was was monitored for 10,800 s and fitted with the Young-Laplace equation
necessary in the first hour of dissolving. The SA solution was filtered to obtain the surface tension. After this waiting time, the droplet was
over a 0.45 μm syringe filter after complete dissolution. subjected to dilatational deformations. Amplitude sweeps were per­
formed with amplitudes ranging from 3 to 30% at a fixed frequency of
2.2.3. Preparation of WPI and SA mixtures 0.02 Hz. The droplet was subjected to five cycles at every amplitude.
A WPI solution with a fixed concentration of 0.2% (w/w) was mixed Step-dilatations were performed by a sudden extension or compression
with a SA solution varying from 0.04 to 2% (w/w) in a 1:1 (v/v) ratio, (step time of 2 s) of the area by 10%. All measurements were performed
which resulted in a final concentration of 0.1% (w/w) WPI. This non- at least in triplicate at 20 ◦ C.
oxidised WPI-SA mixture is abbreviated as WS. The WS sample was
oxidised by adjusting the pH to 10.0 with NaOH, followed by incubation
2.5. Rheology data analysis
overnight (16 h) while stirring slowly at room temperature. The pH was
adjusted to 7.0 on the following day, and this solution is the oxidised
The results of the oscillatory amplitude sweeps in dilatation were
sample (WSox). SA was also oxidised separately using the earlier
analysed using Lissajous plots of the surface pressure (Π = γ-γ0) versus
mentioned method (16 h incubation at pH 10), and mixed with WPI at
the deformation ((A-A0)/A0). Here γ and A are the surface tension and
pH 7.0, which yielded the separately oxidised sample (WSsepox). A
area of the deformed interface, γ0 and A0 are the surface tension and area
schematic overview of different samples and their preparation is
of the non-deformed interface. Plots were generated for each deforma­
depicted in Fig. 2.
tion from the middle three oscillation cycles (van Kempen, Schols, van
der Linden, & Sagis, 2013).
2.2.4. Filtration of the mixtures to remove unbound phenols
Selected WS and Wsox mixture solutions were ultra-filtrated with an
Amicon stirred cell (Merck, Germany) equipped with a 10 kDa-pore size 2.6. Preparation of Langmuir-Blodgett films
membrane to remove unbound phenols. An initial volume of 160 mL
WPI-SA mixture was concentrated to 20 mL, which was diluted again Langmuir-Blodgett films of the protein-stabilised interfacial films
with 140 mL of buffer, followed by another filtration step to 20 mL of were prepared using a 243 cm2 Langmuir film balance (Langmuir-
retentate. The retentate was diluted and filtered for a total of four times. Blodgett Trough KN 2002; KSV NIMA/Biolin Scientific Oy, Finland) at
After three dilutions, the conductivity of the filtrate remained constant. room temperature, which was filled with sodium phosphate buffer (20
The filtration step resulted in approximately 5% protein loss due to loss mM, pH 7.0). Protein-phenol mixtures with a 1:10 (w/w) ratio were
on the filtration membrane, and therefore the final retentate was diluted diluted to 0.02% (w/w) protein content and 200 μL of sample was
with buffer to 95% of the initial volume (152 mL). spread on top of the surface using a gas-tight syringe and were equili­
brated for 30 min. The surface layer was compressed at a barrier moving
2.3. Determination of protein thermal stability by DSC speed of 5 mm/min, while the surface pressure was monitored with a
Wilhelmy plate (platinum, perimeter 20 mm, height 10 mm). After
The protein thermal stability was studied by differential scanning reaching a stable surface pressure, the protein layer was transferred on a
calorimetry (DSC) using a Q100 DSC (TA Instruments, USA). About 50 freshly cleaved mica substrate (Highest Grade V1 Mica, Ted Pella, USA),
μL of solution was weighed in a stainless steel high volume pan. All at 1 mm/min withdraw speed. The films were dried in a desiccator, and
mixtures contained a fixed WPI concentration of 10% (w/w) and 1% (w/ produced in duplicate.
w) of SA. An empty stainless-steel volume pan was used as a reference
and nitrogen was used as carrier gas. Samples were equilibrated for 5 2.7. Determination of the interfacial structure by AFM
min at 20 ◦ C, which was followed by heating to 120 ◦ C at a rate of 5 ◦ C/
min. Possible refolding of the proteins was evaluated by a cooling step The interfacial structure of the Langmuir-Blodgett films was studied
from 120 to 20 ◦ C at a rate of 10 ◦ C/min. Samples were prepared in using an atomic force microscope (AFM, MultiMode 8-HR, Bruker, USA).
duplicate and each of these duplicates was measured twice. Images were recorded in tapping mode using the Scanasyst-air model
non-conductive pyramidal silicon nitride probe (Bruker, USA) with a
2.4. Determination of surface tension and surface dilatational properties normal spring constant of 0.40 N/m. A lateral scan frequency of 0.977
Hz was employed for all topographical images. The lateral resolution
The interfacial properties were studied with a drop tensiometer PAT- was set to 512 × 512 pixels2 in a scan area of 2 × 2 μm2. Duplicates of the
1M (Sinterface Technologies, Germany) at 20 ◦ C. The hanging drop Langmuir-Blodgett films were scanned for at least two locations to
method was performed by forming a drop of sample solution with a ensure a good representativeness. The AFM images were analysed using
surface area of 20 mm2 at the tip of a needle. The shape of the droplet Nanoscope Analysis 1.5 software (Bruker, USA).

Fig. 2. Explanation scheme of prepared mixtures, and abbreviations.

3
J. Yang et al. Food Hydrocolloids 112 (2021) 106467

2.8. Determination of foam properties This alteration of secondary protein structure was not reflected in the
DSC results, as WS and pure WPI showed comparable thermal stability.
Foams were produced using a Foamscan foaming device (Teclis IT- However, a comparable thermal stability of a protein before and after
Concept, France) by sparging nitrogen through a metal frit (27 μm pore addition of phenols does not necessarily indicate an unaffected protein
size, 100 μm distance between centres of pores, square lattice). Aliquots structure, which was also shown for BSA and biochanin-A mixtures (Ali,
of 40 mL sample were sparged in a glass cylinder (ø = 60 mm) at a gas flow Alli, Ismail, & Kermasha, 2012). The secondary structure of the BSA was
rate of 400 mL/min to a foam volume of 400 cm3. The decay of foam found to be altered after non-covalent interaction with phenols, as a
volume was monitored by a camera, and from this the foam half-life time decrease of α-helix and β-sheet domains was observed.
was measured, which is the time required for the foam volume to reduce Subsequently, to induce covalent interactions between SA and WPI,
by 50%. The morphology of foam bubbles was captured with a camera, the WS mixtures were oxidised by overnight incubation at pH 10.0, and
and analysed using a custom Matlab script that used DIPlip and DIPimage the pH was adjusted back to 7.0, which yielded WSox. The WSox sample
image analysis software to obtain an average bubble size. All experiments exhibited a slight increase in turbidity (based on visual observation)
were performed at least in triplicate at 20 ◦ C. compared to WPI and WS, and a substantial decline of the denaturation
peak, which was not observed for pure WPI (see Table 1 and Fig. S1 in
3. Results and discussion SI). The peak enthalpy diminished by more than half, and the peak
temperature also shifted slightly to lower values, close to the Tpeak of
3.1. Evaluation of protein denaturation properties by DSC α-lactalbumin. This could suggest the loss of the initial structure for
β-lactoglobulin, while the structure of α-lactalbumin seems to be
The interactions between sinapic acid (SA) and whey protein isolate (partially) present. The substantial decrease in protein denaturation
(WPI) were initially evaluated by probing the protein denaturation enthalpy could be related to the formation of phenol-protein aggregates.
temperatures and enthalpy using differential scanning calorimetry (DSC) As mentioned in earlier sections, SA is converted into quinones at
(Table 1). An overview of the samples and abbreviations is given in Fig. 2. alkaline pH, which can covalently bind to proteins, and thus cross-
The DSC showed a main denaturation peak for WPI (see Fig. S1 in SI) linking multiple proteins into aggregates. The formation of larger ag­
corresponding to a temperature at 74.4 ± 0.1 ◦ C, and to a denaturation gregates was also observed for WPI-epigallocatechin mixtures after
enthalpy of 13.6 ± 0.7 J/g protein. The whey protein isolate used in this inducing covalent interactions by overnight oxidation at alkaline pH (Jia
study is a mixture of proteins with β-lactoglobulin (74%) and α-lactal­ et al., 2016). We also included a negative control, where WPI was given
bumin (12.5%) as the most abundant ones. The denaturation peak is the overnight alkaline pH treatment (to pH 10.0), and we obtained
mainly dominated by the denaturation of β-lactoglobulin. Additionally, a similar results in DSC as non-treated WPI (see Table S1 in SI).
small shoulder peak is present due to α-lactalbumin, with a denaturation To further understand the properties of the WSox mixture, SA was
temperature around 65 ◦ C (Miao and Mao, 2014). first oxidised separately and then added to WPI (WSsepox). The result­
Addition of SA to a WPI solution at pH 7.0 (WS) resulted in a dena­ ing WSsepox mixture had a lower denaturation peak with a roughly 10%
turation peak temperature of 73.9 ± 0.1 ◦ C and a denaturation enthalpy decrease in enthalpy compared to WPI. Oxidizing WPI and SA together
of 13.2 ± 0.9 J/g protein. These values for WS were comparable to those led to a much larger change in protein denaturation temperatures and
of pure WPI (Table 1). Non-covalent interactions between the compounds enthalpy than adding pre-oxidised SA to WPI. One explanation could be
were expected, as it has been reported for BSA (Czubinski & Dwiecki, the polymerization of quinones with each other in the absence of pro­
2017; Jin et al., 2012). SA has a hydroxycinnamic-acid-like structure teins, and this could result in less effective bridging to the proteins
containing a phenyl ring connected to a carboxyl group by a double co­ (Shahidi & Senadheera, 2019). Another explanation could be the dif­
valent bond (Fig. 1), thereby forming a conjugated system (Cai et al., ference in pH of the reaction between SA and WPI, as phenols are more
1999). The interactions between BSA and SA were expected to be mainly reactive at alkaline pH (Cai et al., 1999). In WSox, the WPI and SA
hydrophobic due to the interaction between the hydrophobic phenyl coexist at pH 10.0, where the phenols are potentially more reactive
group and hydrophobic domains on surface of the protein. A similar compared to the pre-oxidised phenols in WSsepox, where WPI and
interaction mechanism was confirmed for hydroxycinnamic acid de­ pre-oxidised phenols are only mixed at pH 7.0. The difference in reac­
rivatives caffeic and ferulic acid and the main whey protein β-lacto­ tivity and possible pre-polymerization probably resulted in less aggre­
globulin (Li et al., 2020). Binding of these molecules on the hydrophobic gation of proteins for WSsepox compared to WSox. Please note that the
domains of the protein surface was confirmed using NMR spectroscopy, ratio of WPI:SA was 10:1 (w/w) for the DSC experiments, which is
and resulted in a decrease of protein surface hydrophobicity up to 35%. different from the experiments in further sections. This ratio was chosen,
Another major alteration of the protein was the secondary protein since a high concentration of WPI (10%) was required for the DSC
structure, as the percentage of β-sheets decreased, and unordered struc­ measurements, and the dissolution of higher concentrations of SA was
tures increased. Coinciding findings on structure and surface hydro­ not feasible. Other experiments required lower WPI concentrations that
phobicity were also observed when epigallocatechin, chlorogenic, allowed us to go to WPI:SA ratios with higher SA contents. In the next
caffeic, ferulic, or coumaric acid were introduced to whey proteins (Cao sections, we focus on the interface stabilising properties of the WPI-SA
et al., 2018; Cao & Xiong, 2017; Jiang et al., 2018). Based on these pre­ mixtures.
vious studies, we expect the SA to bind non-covalently via hydrophobic
interactions of the phenyl-ring with hydrophobic domains on the protein 3.2. Surface oscillatory dilatational rheology
surface. This interaction would probably also lead to a decrease in surface
hydrophobicity, and increase in unordered secondary protein structures. 3.2.1. Non-oxidised mixtures
The surface activity of WPI and WS mixtures with WPI:SA ratios
between 5:1 to 10:1 (w:w) was assessed using a drop tensiometer
Table 1 (Fig. 3A). WPI exhibited an immediate increase in surface pressure
Protein denaturation onset and peak temperatures and enthalpy of WPI and
within a few seconds, and the surface pressure increased continuously to
WPI-SA mixtures dissolved in a 20 mM PO4 buffer, pH 7.0. The averages and
19.0 mN/m after 3 h of adsorption. The addition of SA resulted in a
standard deviations were obtained from four replicates.
surface pressure increase over the whole time range with an additional
WPI WS WSox WSsepox
1.5 mN/m at the highest SA concentrations. The surface-active phenols
Tonset (◦ C) 60.0 ± 0.7 59.9 ± 0.9 54.5 ± 1.3 61.4 ± 0.3 (Fig. 3A) contributed to increased surface activity of the WS mixtures,
Tpeak (◦ C) 74.4 ± 0.1 73.9 ± 0.1 67.5 ± 0.7 74.4 ± 0.2 which could suggests the co-adsorption of SA and WPI on the air-water
Enthalpy (J/g protein) 13.6 ± 0.7 13.2 ± 0.9 5.9 ± 0.4 12.2 ± 0.1
interface. The nature of the stabilised interfacial layer can be further

4
J. Yang et al. Food Hydrocolloids 112 (2021) 106467

Fig. 3. (A) Surface pressure as a function of time at air-water interfacial films stabilised by WPI (black) and WPI-SA mixtures with a WPI:SA 5:1 to 1:10 (w/w) ratio
(blue lines). A darker blue colour indicates a higher concentration of SA compared to WPI. The WPI concentration was 0.1% (w/w) in all samples. An insert of the
same graph for 1% (w/w) sinapic acid (brown) can be found in the bottom-right corner. The surface pressure isotherms represent an average from at least three
replicates. The standard deviation was below 5%. (B) The surface elastic dilatational modulus as a function of deformation amplitude of the same interfacial films as
described in graph A (fixed frequency 0.02 Hz). The average and standard deviations are the result of at least three replicates. (For interpretation of the references to
colour in this figure legend, the reader is referred to the Web version of this article.)

investigated by performing deformations at a range of amplitudes earlier work (J. Yang, Thielen, et al., 2020). The shape of the plots
(3–30%) in so-called amplitude sweeps. became wider and asymmetric at 30% deformation, which is far into the
The Lissajous plots of a WPI-stabilised interface (Fig. 4) at 5% NLVE. At the start of the extension (bottom-left corner), we can observe
deformation showed a narrow and symmetric ellipse. This indicates a steep surface pressure increase that indicates a predominantly elastic
linear viscoelastic behaviour with a dominant elastic component, and response. Afterwards, the elastic component started to diminish, which
implies the formation of a viscoelastic solid-like layer, as shown in was reflected in a gradual decrease in the slope of the surface pressure

Fig. 4. Lissajous plots of surface pressure as a function of the applied deformation, obtained during amplitude sweeps of air-water interfacial films stabilised by WPI
and WPI-SA mixtures at a 1:10 (w/w) ratio. Lissajous plots of a 1% (w/w) SA solution was also included. For clarity, one representative plot is shown for each sample,
but comparable plots were obtained on at least three replicates.

5
J. Yang et al. Food Hydrocolloids 112 (2021) 106467

curve. This is related to a gradual loss of cohesiveness of the interfacial 3.2.2. Oxidised mixtures
microstructure, and an increase in the viscous contribution to the total Oxidised WS (WSox) mixtures were more surface-active compared to
surface pressure response. This phenomenon upon extension of the the non-oxidised WS (Fig. 5A). Additionally, the surface dilatational
interface is called intra-cycle strain softening. The opposite can be modulus (Ed’) decreased by 11–18% over the whole amplitude range
observed upon compression, namely intra-cycle strain hardening, as the after oxidation (Fig. 5B). The Lissajous plots of WSox-stabilised in­
maximum surface pressure reached much higher values compared to terfaces at 30% deformation (Fig. 4, middle) showed even less strain
those reached in extension. In earlier work, we attributed this phe­ hardening in compression compared to WS, implying that phenols were
nomenon to jamming of densely clustered protein regions at the inter­ also adsorbed at the interface, and desorbed into the bulk before the WPI
face (J. Yang, Thielen, et al., 2020). From the Lissajous plots, we can started jamming. In the extension part, we can observe a lower surface
conclude that WPI can form stiff and cohesive interfaces with the pressure increase for WSox than WPI, and the curve for WSox even starts
rheological behaviour of a viscoelastic two-dimensional solid, as flattening, which indicates increased strain-softening of the interfacial
described in several studies (Felix, Yang, Guerrero, & Sagis, 2019; Rühs, layer. This would imply a reduced resistance of the interfacial network
Scheuble, Windhab, & Fischer, 2013; J.; Yang, Thielen, et al., 2020; against extensional deformations, and thus an alteration in the in­
Yang et al., 2020). teractions between proteins at the interface. The proteins in WSox were
On the other hand, the Lissajous plots of SA had a narrow shape, and found to be aggregated, which could explain a weaker interfacial
were more tilted towards the horizontal axis. From this, we can conclude network compared to WPI and WS, as fewer proteins are available to
that SA formed very weak interfaces compared to WPI. The Lissajous interact at the interface. A negative control, where WPI was given the
plots of WS-stabilised (WPI:SA ratio 1:10) interfaces (Fig. 4, left) at 5% same pH-shift treatment, was included and gave identical results as
deformation had a similar shape as WPI-stabilised interfaces, but the untreated WPI in the drop tensiometer (see Fig. S2 in SI). The effect of
plots of WS were more tilted towards the horizontal axis, which in­ phenol oxidation as related to interfacial properties was further studied
dicates a weaker network. A more pronounced difference can be by mixing separately oxidised SA with WPI (WSsepox), which revealed a
observed at 30% deformation, where the strain hardening at compres­ similar surface rheological behaviour as non-oxidised WS (overlapping
sion became less evident compared to WPI, as the maximum surface moduli and Lissajous plots, Figs. 4 & 5B). This proves that the oxidation
pressure value in compression decreased from − 17 mN/m (WPI) to − 13 of WPI and SA together at pH 10.0 had a more significant impact on the
mN/m (WS). This could be due to the presence of SA at the interface, as a interfacial properties than the addition of separately oxidised SA.
reduction of strain hardening would imply that first the SA is expelled
from the interface into the bulk. Afterwards, the WPI starts jamming, 3.2.3. Removal of unbound phenols
and again showed a steep increase in surface pressure at the start of the As mentioned in earlier sections, only a fraction of SA is bound non-
extension. At 30% deformation, both WPI and WS showed a comparable covalently or covalently to the proteins, which implies there is also an
strain-softening behaviour upon extension, as the curves overlap in unbound fraction of phenols (Karefyllakis et al., 2018). To remove SA
extension, which suggests that the WPI in a mixture with SA still in­ that is not interacting with WPI, the WS and WSox mixtures were filtered
teracts to form a cohesive interfacial network similar as the pure WPI to obtain filtered WS (FWS) and filtered WSox (FWSox), respectively.
interface. A comparable protein network upon extension would imply Both filtered and non-filtered samples had a comparable surface activity
that the strain hardening in compression is due to the desorption of (see Fig. S3 in SI). A larger difference was demonstrated in the interfacial
phenols from the interface towards the bulk phase, followed by jamming layer stiffness (Fig. 6). Interfacial films stabilised by FWS revealed a
of the proteins. We should also keep in mind that the SA interacts with slight increase of Ed’ below 10% deformation compared to WS, but both
the WPI by hydrophobic interactions. As a result, hydrophobic regions FWS and WS had overlapping Ed’ at higher deformations. This is also
on the surface of the proteins will be reduced, and less hydrophobic reflected in the Lissajous plots (Fig. 4), as a FWS-stabilised interface
interactions would occur between the proteins on the air-water inter­ showed an increase of maximum surface pressure at 5% deformation,
face. To summarise, the addition of SA to WPI resulted in an increase in and a similar Lissajous plot at 30% deformation compared to WS. The
surface activity and a decrease in interfacial layer strength (Fig. 3A and FWSox-stabilised interface had a comparable rheological behaviour as
B). A WPI:SA ratio of 1:10 gave the most pronounced difference the FWS-stabilised interface, which implied a sizable increase of Ed’ after
compared to WPI alone. Therefore, we chose to continue with this ratio unbound phenol removal, compared to WSox. The Lissajous plot (Fig. 4)
in all further experiments. of the FWSox-stabilised interface also exhibited more strain hardening

Fig. 5. (A) Surface pressure as a function of time at air-water interfacial films stabilised by WPI and WPI-SA mixtures. The surface pressure isotherm represents an
average from at least three replicates. The stand deviation was below 5%. (B) The surface elastic dilatational modulus as a function of deformation amplitude of
interfacial films stabilised by WPI and WPI-SA mixtures at a 1:10 ratio (fixed frequency 0.02 Hz). The different samples are given in the legend. The average and
standard deviations are the result of at least three replicates.

6
J. Yang et al. Food Hydrocolloids 112 (2021) 106467

covalently bound phenols could be released from the proteins, when


the phenol concentration in the bulk decreases, and thereby still co-
adsorb at the interface.

3.3. Interfacial microstructure

We also evaluated the interfacial microstructure by transferring the


interfacial film at a surface pressure of 15 or 25 mN/m onto a solid
substrate (Langmuir-Blodgett deposition) and analysing the topography
of the dried film using atomic force microscopy (AFM) (2D images in
Fig. 7 and 3D images in Fig. S6 in SI). A heterogeneous structure was
observed for the WPI-stabilised films at 15 mN/m, where thicker regions
were present. The thicker regions could be densely clustered proteins
that are even more densely compressed at a higher surface pressure of
25 mN/m. At this surface pressure, a heterogeneous structure was still
present, as was also demonstrated by others (Berton-Carabin, Genot,
Gaillard, Guibert, & Ropers, 2013; Rühs et al., 2013), and related to the
rheological nature of the WPI-stabilised interface (J. Yang, Thielen,
et al., 2020), which behaved as a heterogeneous viscoelastic solid. The
Fig. 6. Surface elastic dilatational modulus as a function of deformation WS mixture exhibited a similar heterogeneous structure, and this
amplitude of interfacial films stabilised by WPI, WPI-SA mixtures, and filtered structure remained after filtration (FWS).
WPI-SA mixtures at a 1:10 ratio (fixed frequency 0.02 Hz). The different sam­ Oxidation of SA showed changes in microstructure, as the WSox-
ples are given in the legend. The average and standard deviations are the result stabilised interface at 15 mN/m showed the presence of larger struc­
of at least three replicates. tures. These structures could be the previously mentioned aggregates
that are cross-linked by quinones. The aggregates were still intact and
compared to WSox, suggesting the presence of few phenol molecules at present in the film after filtration. The addition of separately oxidised SA
the interface. Unbound SA seems to play a larger role on interfacial (WSsepox) resulted in more clustered areas compared to pure WPI, but
properties after oxidation together with protein. Oxidised SA was also they were not as large as the aggregates in WSox and FWSox. This
studied separately for the interfacial properties, which were found to suggests that conditions (pH & reaction time) between phenol and WPI
coincide with the non-oxidised version (see Fig. S4 in SI). The are key parameters to influence the interaction between phenols and
strain-softening in extension in the 30% Lissajous plot of FWSox was proteins. The aggregates are not present at a surface pressure of 25 mN/
coinciding with the plot of WSox, and implies that the proteins show less m, which was also observed for interfaces stabilised by heat-denatured
resistance to extension compared to pure WPI, and thus the interactions WPI aggregates (J. Yang, Thielen, et al., 2020). In that study, it was
between the proteins became weaker. shown that the aggregates were pushed out of the interface into the bulk
After the removal of unbound phenols, the samples FWS and FWSox phase upon compression in the Langmuir trough. It should be noted that
still showed a higher surface activity and a decrease of elastic moduli the Langmuir-Blodgett films are produced by spreading the material at
compared to pure WPI. This would imply the non-covalent interaction the interface, and may result in a different microstructure compared to
with phenols affected the proteins by reducing the protein surface hy­ an interfacial layer formed by spontaneous diffusion of surface-active
drophobicity, and induces changes in protein secondary structure. This material, such as those formed in the drop tensiometer. At a surface
was reflected in weaker interfaces for WS compared to WPI. Oxidation of pressure of 25 mN/m, we observed a similar heterogeneous interface for
the phenols resulted in WSox, which showed an even less stiff layer than all samples, which we could relate to the rheological behaviour at higher
WS, and this could be related to the formed protein-phenol aggregates. deformations, as the moduli of all samples started to overlap at 30%
Still, removal of unbound phenols resulted in a slight increase in moduli, deformation in the amplitude sweeps (Figs. 3, 5 and 6). This would
indicating that some unbound phenols could be present at the interfacial suggest that WPI still forms a heterogeneous viscoelastic layer at the
layer. Additionally, we should keep in mind that traces of unbound interface in the presence of SA.
phenols might still be present after extensive filtration. Besides, non-

Fig. 7. AFM images of Langmuir-Blodgett films made from WPI- and WPI-SA mixtures-stabilised interfaces. The surface pressure indicates the conditions during
film sampling.

7
J. Yang et al. Food Hydrocolloids 112 (2021) 106467

3.4. Step-dilatation experiments β-lactoglobulin aggregates (Chen et al., 2017; Rullier, Novales, & Axe­
los, 2008). These studies suggested that these colloids had little contri­
The structural heterogeneity of the WPI- and WPI-SA mixture-sta­ bution to the air bubble surface, but in fact the micelles/aggregates were
bilised interfaces were further investigated by performing step- trapped in the lamellae and plateau borders of the foam, and slowed
dilatations in the drop tensiometer (example in Fig. S5 in SI). The down the drainage of liquid from the foam. Nevertheless, the less stiff
relaxation response was characterised with a combination of a Kohlraus- interfacial layer of WSox compared to WPI seems to be more detrimental
William-Watts (KWW) stretch exponential and a regular exponential for foam stability.
term (Equation (1)) (Watts & Davies, 1969). Removal of unbound phenols (FWS and FWSox) resulted in a larger
bubble size, comparable to that obtained with pure WPI, and in a slight
γ(t) = ae− (t/τ1 )β
+ be− t/τ2
+c (1) increase of the foam half-life time to 181–188 min. FWS and FWSox still
The stretch exponent β and relaxation time τ1 in the equation are formed less stable foams than WPI (258 min), which aligned with the
presented in Table 2. The characteristic time of the second term τ2 and lower surface elastic modulus of films formed by FWS and FWSox
fitting parameters a, b, and c can be found in Table S2 in the SI. compared to pure WPI (Fig. 6). The presence of unbound phenols and
The KWW model was applied previously to characterise the relaxa­ the bound phenols on proteins led to the formation of weaker interfacial
tion behaviour of WPI-stabilised interfaces (Felix et al., 2019; Sagis layers, and thus a less stable foam compared to pure WPI. Generally,
et al., 2019; Sagis et al., 2014; J.Yang, Thielen, et al., 2020), which was previous studies also showed an increase in foam ability after addition of
described by the first term of the equation. The regular exponential term phenols to protein, which was related to the co-adsorption of phenols
was included to separate the aging of the interface from the relaxation and proteins, and an alteration of the protein structure and surface hy­
process. All interfaces showed β′ s between 0.50 and 0.56 for extension drophobicity (Cao et al., 2018; Jia et al., 2016; Jiang et al., 2018). The
and 0.55–0.62 for compression, and relaxation times (τ1) between 19.7 type of interaction also influenced the foam stability, as the formation of
and 27.4 s. A value of β < 1 is an indication of dynamic heterogeneity, protein aggregates seemed to increase foam stability due to blockage of
and reveals the presence of local variations in the relaxation kinetics, the lamellae and plateau borders in the foams (Jia et al., 2016; Rodrí­
which is also common in disordered solids (Klafter & Shlesinger, 1986; guez et al., 2015). This phenomena was slightly present in WSox foams,
Phillips, 1996). This dynamic heterogeneity could be the result of the as the stability was higher compared to WS foams, but still lower than
structural heterogeneity that was revealed in the interfacial micro­ pure WPI. The size of the aggregates are largely determining the effec­
structure (Fig. 7). The β- and τ1-values of all interfaces were in a com­ tiveness in lamellae and plateau border blockage (Rullier et al., 2008),
parable range, which was also expected, as all AFM images at a surface which suggests that the aggregate size in WSox was not optimal
pressure of 25 mN/m gave comparable heterogeneous structures. The compared to those in the earlier reported works (Jia et al., 2016;
addition of SA to WPI leads to a similar relaxation behaviour of the in­ Rodríguez et al., 2015). Other works also demonstrated an increase in
terfaces, which shows that the WPI in the presence of SA still forms foam stability, as the protein solubility increased due to a decrease in
solid-like structures at the air-water interface, as also demonstrated in surface hydrophobicity upon addition of phenols (Cao et al., 2018; Jiang
the amplitude sweeps, and AFM images. et al., 2018). This cannot occur in our work, as we prepared the WPI
solutions by centrifugation and filtration to obtain a soluble fraction.
Therefore, we can largely attribute the foaming properties to the studied
3.5. Foams interfacial properties.

Foams were produced by N2-gas injection and were monitored for 4. Conclusions
the foam-ability and stability, which were expressed by the average
bubble size, and foam half-life time, where half of the foam volume Interfacial and foaming properties of whey protein isolate were
collapsed (Fig. 8). WPI is often described as an excellent foaming agent studied in the presence of sinapic acid. A schematic overview of the
with high foam stability (Cao et al., 2018; Lech, Delahaije, Meinders, proposed interfacial microstructure stabilised by WPI-SA mixtures can
Gruppen, & Wierenga, 2016). In our work, WPI exhibited a high foam be found in Fig. 9. Whey protein isolate forms a strong viscoelastic solid-
half-life time of 258 ± 24 min. The presence of SA (WS, WSox, WSsepox) like layer at the air-water interface (Fig. 9A), and thereby exhibits high
diminished the half-life time to 138–172 min. This lower foam stability foam stability. The addition of sinapic acid results in weaker interfacial
could be related to the surface dilatational elastic modulus decrease layers (Fig. 9B), and oxidation of the phenols further enhances this ef­
(Figs. 3, 5 and 6), as the strength of the interfacial layer around the air fect. Oxidation of sinapic acid increases the reactivity of sinapic acid
bubble is often related to foam stability (Bos & van Vliet, 2001). The towards proteins and cross-links WPI into aggregates (Fig. 9C), as shown
increased surface activity after the addition of SA resulted in a smaller by AFM. Removal of unbound phenols caused a slight recovery in
air bubble size (Fig. 8B). WSox-stabilised foam showed a slightly longer interfacial layer strength and foam stability, but it was still less
half-life time compared to WS and WSsepox, which is probably due to compared to whey protein. From this, we suggest that the unbound
the smaller bubble size of WSox, and can be linked to its higher surface phenols and proteins are both present at the air-water interface, and
activity. Additionally, the formation of aggregates/particles could in­ form a weaker interfacial layer than pure protein (Fig. 9B). Additionally,
crease foam stability, as was demonstrated for casein micelles and phenols are bound to the proteins, and could presumably induce changes
to surface and structural properties of the proteins, which should be
Table 2 addressed in further research. The phenol sinapic acid can negatively
The stretch exponent β and relaxation time τ1 for WPI and WPI-SA mixtures influence the functionality of (whey) proteins, and this negative effect is
obtained from the step-dilatation experiments. The averages and standard de­ further enhanced upon oxidation of the phenols. The presence of phe­
viations are the result of at least three replicates. nols, and the possible oxidation upon protein extraction (at alkaline
Extension Compression conditions) should be considered in the production of plant protein
β τ1 β τ1 extracts from phenol-rich plant sources, and also in the utilization of
these sources in our food systems.
WPI 0.51 ± 0.05 25.9 ± 2.3 0.56 ± 0.01 27.4 ± 6.7
WS 0.54 ± 0.01 23.0 ± 3.9 0.59 ± 0.03 26.5 ± 1.5
WSox 0.50 ± 0.04 23.7 ± 3.6 0.59 ± 0.04 19.7 ± 2.3 CRediT authorship contribution statement
WSsepox 0.51 ± 0.02 23.1 ± 2.4 0.62 ± 0.03 24.2 ± 7.0
FWS 0.56 ± 0.01 25.9 ± 2.1 0.60 ± 0.02 26.6 ± 2.1 Jack Yang: Conceptualisation, Methodology, Investigation, Valida­
FWSox 0.53 ± 0.03 22.2 ± 3.4 0.55 ± 0.02 21.6 ± 3.6
tion, Visualisation, Writing - original draft. Sarah P. Lamochi

8
J. Yang et al. Food Hydrocolloids 112 (2021) 106467

Fig. 8. The half-life time (A), and bubble diameter (B) of foams stabilised by WPI and WPI-SA mixtures. Averages and standard deviations are the result of at least
three replicates.

Fig. 9. Schematic overview of the proposed interfacial composition of WPI-SA mixtures. The illustrated molecules are not on scale.

Roozalipour: Methodology, Investigation. Claire C. Berton-Carabin: Appendix A. Supplementary data


Methodology, Writing - review & editing. Constantinos V. Nikiforidis:
Conceptualisation, Methodology, Writing - review & editing. Erik van Supplementary data to this article can be found online at https://doi.
der Linden: Writing – Reviewing & Editing, Funding acquisition. Leo­ org/10.1016/j.foodhyd.2020.106467.
nard M.C. Sagis: Conceptualisation, Methodology, Supervision, Writing
- review & editing. References

Ali, H., Alli, I., Ismail, A., & Kermasha, S. (2012). Protein-phenolic interactions in food.
Eurasian Journal of Analytical Chemistry, 7(3), 123–133.
Declaration of competing interest Berton-Carabin, C., Genot, C., Gaillard, C., Guibert, D., & Ropers, M. H. (2013). Design of
interfacial films to control lipid oxidation in oil-in-water emulsions. Food
The authors have declared that no competing interest exist. This Hydrocolloids, 33(1), 99–105. https://doi.org/10.1016/j.foodhyd.2013.02.021
Bos, M. A., & van Vliet, T. (2001). Interfacial rheological properties of adsorbed protein
manuscript has not been published and is not under consideration for layers and surfactants: A review. Advances in Colloid and Interface Science, 91(3),
publication in any other journal. All authors approve this manuscript 437–471. https://doi.org/10.1016/S0001-8686(00)00077-4
and its submission to Food Hydrocolloids. Butré, C. I., Wierenga, P.a., & Gruppen, H. (2012). Effects of ionic strength on the
enzymatic hydrolysis of diluted and concentrated whey protein isolate. Journal of
Agricultural and Food Chemistry, 60(22), 5644–5651. https://doi.org/10.1021/
Acknowledgements jf301409n
Cai, R., Arntfield, S. D., & Charlton, J. L. (1999). Structural changes of sinapic acid during
alkali-induced air oxidation and the development of colored substances. JAOCS,
The authors have declared that no competing interest exists. The Journal of the American Oil Chemists’ Society, 76(6), 757–764. https://doi.org/
project is funded by TiFN, a public-private partnership on pre­ 10.1007/s11746-999-0172-6
Cao, Y., & Xiong, Y. L. (2017). Interaction of whey proteins with phenolic derivatives
competitive research in food and nutrition. The public partners are
under neutral and acidic pH conditions. Journal of Food Science, 82(2), 409–419.
responsible for the study design, data collection and analysis, decision to https://doi.org/10.1111/1750-3841.13607
publish, and preparation of the manuscript. The private partners have Cao, Y., Xiong, Y. L., Cao, Y., & True, A. D. (2018). Interfacial properties of whey protein
contributed to the project through regular discussion. This research was foams as influenced by preheating and phenolic binding at neutral pH. Food
Hydrocolloids, 82, 379–387. https://doi.org/10.1016/j.foodhyd.2018.04.020
performed with additional funding from the Netherlands Organisation Chen, M., Sala, G., Meinders, M. B. J., van Valenberg, H. J. F., van der Linden, E., &
for Scientific Research (NWO), and the Top Consortia for Knowledge and Sagis, L. M. C. (2017). Interfacial properties, thin film stability and foam stability of
Innovation of the Dutch Ministry of Economic Affairs (TKI). NWO casein micelle dispersions. Colloids and Surfaces B: Biointerfaces, 149, 56–63. https://
doi.org/10.1016/j.colsurfb.2016.10.010
project number: ALWTF. 2016.001. The authors would like to thank
Carol Shek for her contribution in proofreading.

9
J. Yang et al. Food Hydrocolloids 112 (2021) 106467

Cheung, L., Wanasundara, J., & Nickerson, M. T. (2014). The effect of pH and NaCl levels Pringent, S. V. E., Voragen, A. G. J., Visser, A. J. W. G., van Koningsveld, G. A., &
on the physicochemical and emulsifying properties of a cruciferin protein isolate. Gruppen, H. (2007). Covalent interactions between proteins and oxidation products
Food Biophysics, 9(2), 105–113. https://doi.org/10.1007/s11483-013-9323-2 of caffeoylquinic acid (chlorogenic acid). Journal of the Science of Food and
Czubinski, J., & Dwiecki, K. (2017). A review of methods used for investigation of Agriculture, 87, 2502–2510. https://doi.org/10.1002/jsfa
protein–phenolic compound interactions. International Journal of Food Science and Rodríguez, S. D., von Staszewski, M., & Pilosof, A. M. R. (2015). Green tea polyphenols-
Technology, 52(3), 573–585. https://doi.org/10.1111/ijfs.13339 whey proteins nanoparticles: Bulk, interfacial and foaming behavior. Food
Das Purkayastha, M., Gogoi, J., Kalita, D., Chattopadhyay, P., Nakhuru, K. S., Goyary, D., Hydrocolloids, 50, 108–115. https://doi.org/10.1016/j.foodhyd.2015.04.015
et al. (2014). Physicochemical and functional properties of rapeseed protein isolate: Rühs, P. A., Scheuble, N., Windhab, E. J., & Fischer, P. (2013). Protein adsorption and
Influence of antinutrient removal with acidified organic solvents from rapeseed interfacial rheology interfering in dilatational experiment. The European Physical
meal. Journal of Agricultural and Food Chemistry, 62(31), 7903–7914. https://doi. Journal - Special Topics, 222(1), 47–60. https://doi.org/10.1140/epjst/e2013-01825-
org/10.1021/jf5023803 0
Felix, M., Yang, J., Guerrero, A., & Sagis, L. M. C. (2019). Effect of cinnamaldehyde on Rullier, B., Novales, B., & Axelos, M. A. V. (2008). Effect of protein aggregates on
interfacial rheological properties of proteins adsorbed at O/W interfaces. Food foaming properties of β-lactoglobulin. Colloids and Surfaces A: Physicochemical and
Hydrocolloids, 97, 105235. https://doi.org/10.1016/j.foodhyd.2019.105235 Engineering Aspects, 330(2–3), 96–102. https://doi.org/10.1016/j.
Jiang, J., Zhang, Z., Zhao, J., & Liu, Y. (2018). The effect of non-covalent interaction of colsurfa.2008.07.040
chlorogenic acid with whey protein and casein on physicochemical and radical- Sagis, L. M. C., Humblet-Hua, K. N. P., & van Kempen, S. E. H. J. (2014). Nonlinear stress
scavenging activity of in vitro protein digests. Food Chemistry, 268(May), 334–341. deformation behavior of interfaces stabilized by food-based ingredients. Journal of
https://doi.org/10.1016/j.foodchem.2018.06.015 Physics: Condensed Matter, 26(46), 9. https://doi.org/10.1088/0953-8984/26/46/
Jia, Z., Zheng, M., Tao, F., Chen, W., Huang, G., & Jiang, J. (2016). Effect of covalent 464105
modification by (-)-epigallocatechin-3-gallate on physicochemical and functional Sagis, L. M. C., Liu, B., Li, Y., Essers, J., Yang, J., Moghimikheirabadi, A., et al. (2019).
properties of whey protein isolate. Lebensmittel-Wissenschaft und -Technologie- Food Dynamic heterogeneity in complex interfaces of soft interface-dominated materials.
Science and Technology, 66, 305–310. https://doi.org/10.1016/j.lwt.2015.10.054 Scientific Reports, 9(1), 1–12. https://doi.org/10.1038/s41598-019-39761-7
Jin, X. L., Wei, X., Qi, F. M., Yu, S. S., Zhou, B., & Bai, S. (2012). Characterization of Seczyk, L., Swieca, M., Kapusta, I., & Gawlik-Dziki, U. (2019). Protein–phenolic
hydroxycinnamic acid derivatives binding to bovine serum albumin. Organic and interactions as a factor affecting the physicochemical properties of white bean
Biomolecular Chemistry, 10(17), 3424–3431. https://doi.org/10.1039/c2ob25237f proteins. Molecules, 24(3), 1–24. https://doi.org/10.3390/molecules24030408
Karefyllakis, D., Salakou, S., Bitter, J. H., van der Goot, A. J., & Nikiforidis, C. V. (2018). Shahidi, F., & Senadheera, R. (2019). Encyclopedia of food chemistry: Protein–phenol
Covalent bonding of chlorogenic acid induces structural modifications on sunflower interactions. Encyclopedia of Food Chemistry, 2. https://doi.org/10.1016/b978-0-08-
proteins. ChemPhysChem, 19(4), 459–468. https://doi.org/10.1002/ 100596-5.21485-6. Elsevier.
cphc.201701054 von Staszewski, M., Pizones Ruiz-Henestrosa, V. M., & Pilosof, A. M. R. (2014). Green tea
van Kempen, S. E. H. J., Schols, H. A., van der Linden, E., & Sagis, L. M. C. (2013). Non- polyphenols-β-lactoglobulin nanocomplexes: Interfacial behavior, emulsification and
linear surface dilatational rheology as a tool for understanding microstructures of oxidation stability of fish oil. Food Hydrocolloids, 35, 505–511. https://doi.org/
air/water interfaces stabilized by oligofructose fatty acid esters. Soft Matter, 9(40), 10.1016/j.foodhyd.2013.07.008
9579. https://doi.org/10.1039/c3sm51770e Tan, S. H., Mailer, R. J., Blanchard, C. L., & Agboola, S. O. (2011). Extraction and
Keppler, J. K., Schwarz, K., & van der Goot, A. J. (2020). Covalent modification of food characterization of protein fractions from Australian canola meals. Food Research
proteins by plant-based ingredients (polyphenols and organosulphur compounds): A International, 44(4), 1075–1082. https://doi.org/10.1016/j.foodres.2011.03.023
commonplace reaction with novel utilization potential. Trends in Food Science & Tan, S. H., Mailer, R. J., Blanchard, C. L., Agboola, S. O., & Day, L. (2014). Gelling
Technology, 101(March), 38–49. https://doi.org/10.1016/j.tifs.2020.04.023 properties of protein fractions and protein isolate extracted from Australian canola
Klafter, J., & Shlesinger, M. F. (1986). On the relationship among three theories of meal. Food Research International, 62, 819–828. https://doi.org/10.1016/j.
relaxation in disordered. Proceedings of the National Academy of Sciences of the United foodres.2014.04.055
States of America, 83(4), 848–851. https://doi.org/10.1073/pnas.83.4.848 Wanasundara, J. P. D. (2011). Proteins of brassicaceae oilseeds and their potential as a
Kroll, J., Rawel, H. M., & Rohn, S. (2003). Reactions of plant phenolics with food proteins plant protein source. Critical Reviews in Food Science and Nutrition, 51(7), 635–677.
and enzymes under special consideration of covalent bonds. Food Science and https://doi.org/10.1080/10408391003749942
Technology Research, 9(3), 205–218. https://doi.org/10.3136/fstr.9.205 Watts, C., & Davies, E. (1969). Non-symmetrical dielectric relaxation behaviour arising
Lech, F. J., Delahaije, R. J. B. M., Meinders, M. B. J., Gruppen, H., & Wierenga, P. A. from a simple empirical decay function. Transactions of the Faraday Society, 66(1),
(2016). Identification of critical concentrations determining foam ability and 80–85. https://doi.org/10.1039/TF9706600080
stability of β-lactoglobulin. Food Hydrocolloids, 57, 46–54. https://doi.org/10.1016/ Wei, Z., Yang, W., Fan, R., Yuan, F., & Gao, Y. (2015). Evaluation of structural and
j.foodhyd.2016.01.005 functional properties of protein-EGCG complexes and their ability of stabilizing a
Li, X., Dai, T., Hu, P., Zhang, C., Chen, J., Liu, C., et al. (2020). Characterization the non- model β-carotene emulsion. Food Hydrocolloids, 45, 337–350. https://doi.org/
covalent interactions between beta lactoglobulin and selected phenolic acids. Food 10.1016/j.foodhyd.2014.12.008
Hydrocolloids, 105, 105761. https://doi.org/10.1016/j.foodhyd.2020.105761 Xu, L., & Diosady, L. L. (2000). Interactions between canola proteins and phenolic
Malabat, C., nchez-Vioque, R. I. S., Rabiller, C., & Gu guen, J. (2001). Emulsifying and compounds in aqueous media. Food Research International, 33(9), 725–731. https://
foaming properties of native and chemically modified peptides from the 2S and 12S doi.org/10.1016/S0963-9969(00)00062-4
proteins of rapeseed (Brassica napus L.). Journal of the American Oil Chemists’ Society, Xu, L., & Diosady, L. L. (2002). Removal of phenolic compounds in the production of
78(3), 235–242. https://doi.org/10.1007/s11746-001-0251-x high-quality canola protein isolates. Food Research International, 35(1), 23–30.
Miao, S., & Mao, L. (2014). DSC application to characterizing food emulsions. In https://doi.org/10.1016/S0963-9969(00)00159-9
Differential scanning calorimetry applications in fat and oil technology (1st ed., pp. Yang, J., Faber, I., Berton-Carabin, C. C., Nikiforidis, C. V., van der Linden, E., &
243–272). CRC. https://doi.org/10.1201/b17739-14. Sagis, L. M. C. (2020). Foams and air-water interfaces stabilised by mildly purified
Naczk, M., Amarowicz, R., & Shahidi, F. (1998). Role of phenolics in flavor of rapeseed rapeseed proteins after defatting. Food Hydrocolloids, 112(August 2020), 106270.
protein products. Developments in Food Science, 40, 597–613. https://doi.org/ https://doi.org/10.1016/j.foodhyd.2020.106270
10.1016/S0167-4501(98)80080-0 Yang, W., Liu, F., Xu, C., Yuan, F., & Gao, Y. (2014). Molecular interaction between
Naczk, M., Oickle, D., Pink, D., & Shahidi, F. (1996). Protein precipitating capacity of (-)-epigallocatechin-3-gallate and bovine lactoferrin using multi-spectroscopic
crude canola tannins: Effect of pH, tannin, and protein concentrations. Journal of method and isothermal titration calorimetry. Food Research International, 64,
Agricultural and Food Chemistry, 44(8), 2144–2148. https://doi.org/10.1021/ 141–149. https://doi.org/10.1016/j.foodres.2014.06.001
jf960165k Yang, J., Thielen, I., Berton-Carabin, C. C., van der Linden, E., & Sagis, L. M. C. (2020).
Ntone, E., Bitter, J. H., & Nikiforidis, C. V. (2020). Not sequentially but simultaneously: Nonlinear interfacial rheology and atomic force microscopy of air-water interfaces
Facile extraction of proteins and oleosomes from oilseeds. Food Hydrocolloids, 102, stabilized by whey protein beads and their constituents. Food Hydrocolloids, 101,
105598. https://doi.org/10.1016/j.foodhyd.2019.105598 105466. https://doi.org/10.1016/j.foodhyd.2019.105466
Ozdal, T., Capanoglu, E., & Altay, F. (2013). A review on protein-phenolic interactions Zhang, H., Yu, D., Sun, J., Guo, H., Ding, Q., Liu, R., et al. (2014). Interaction of milk
and associated changes. Food Research International, 51(2), 954–970. https://doi. whey protein with common phenolic acids. Journal of Molecular Structure, 1058(1),
org/10.1016/j.foodres.2013.02.009 228–233. https://doi.org/10.1016/j.molstruc.2013.11.009
Phillips, J. C. (1996). Stretched exponential relaxation in molecular and electronic
glasses. Reports on Progress in Physics, 59(9), 1133–1207. https://doi.org/10.1088/
0034-4885/59/9/003

10

You might also like