Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Food Hydrocolloids 96 (2019) 288–299

Contents lists available at ScienceDirect

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

Improvement of the solubility and emulsifying properties of rice bran T


protein by phosphorylation with sodium trimetaphosphate
Zhenying Hua, Liang Qiub, Yong Suna, Hua Xionga,∗, Yasumitsu Ograc
a
State Key Laboratory of Food Science and Technology, NanChang University, Nanchang, 330047, PR China
b
Jiangxi University of Traditional Chinese Medicine, Nanchang, Jiangxi 330004, PR China
c
Laboratory of Toxicology and Environmental Health, Graduate School of Pharmaceutical Sciences, Chiba University, Chuo, Chiba, 260-8675, Japan

A R T I C LE I N FO A B S T R A C T

Keywords: Rice bran protein (RP) could represent an important component of the protein resources in daily life because a
Rice bran protein large amount of it remains after rice processing. At present work, RP underwent phosphorylation using sodium
Phosphorylation trimetaphosphate under hydrothermal treatment at different pH values (3.0, 5.0, 7.0, 9.0 and 11.0), and the
Emulsion utilization of phosphorylated products in emulsions was also assessed. Phosphorylated RP at pH 9.0 showed the
XPS
highest solubility and emulsifying activity. According to the results of the phosphorus content, zeta potential,
Emulsion stability
infrared spectroscopy and X-ray photoelectron spectroscopy analysis, phosphate groups were introduced to RP,
and the phosphate esters (P=O) were detected. With the analysis of surface hydrophobicity, intrinsic fluores-
cence, and circular dichroism spectra, phosphorylation by hydrothermal treatment changed the secondary and/
or tertiary structure of proteins and increased the solubility and emulsifying activity. Emulsions prepared with
phosphorylated RP at pH 9.0 were stable over a range of environmental conditions: pH 7–9, NaCl < 50 mM, and
temperatures < 90 °C were studied. It shows that RP phosphorylated by sodium trimetaphosphate could be
utilized in the formulation and production of natural emulsion-based products as emulsifier.

1. Introduction characteristics of RP are expected to be modified in order to maximize


the utilization in the food industry.
Rice bran is the by-product of rice milling and is considered to be a To the best of our knowledge, the application of phosphorylation in
rich source of protein, fat, carbohydrate and micronutrients, such as protein modification could be an effective and economical way to im-
minerals and phenolic acid (Han, Chee, & Cho, 2015). Almost 90% of prove the functional properties of protein and has already been con-
the rice bran worldwide is utilized as fodder for livestock, and others ducted in soybean, peanut, whey, egg white isolated protein, and others
were used for the extraction of rice bran oil (Zullaikah, Melwita, & Ju, (Cheng et al., 2017; Li, Sun, Ma, Jin, & Sheng, 2018; Sánchez-Reséndiz
2009). Rice bran seemed to be minimally utilized in the food industry et al., 2018; Yu et al., 2015). Based on its chemical versatility, phos-
because of its high fibre content and possible hull contamination phate can form mono-, di- and tri-esters with alkyl and aryl hydroxyl
(Hamada, 1997). Accounting for 5–8% of total grain, rice bran contains groups as well as acid anhydrides, which are known to occur on nine
11–13% crude protein (Gul, Yousuf, Singh, Singh, & Wani, 2015). amino acids in protein including Ser, Thr, Tyr, Arg, Lys, His, Cys, Asp
Meanwhile, rice bran protein (RP) is recognized as a suitable protein and Glu, and are stable in aqueous solution at physiological pH (Hunter,
source for infants, young children and special groups, which is attrib- 2012). According to the previous report, the mechanism of modification
uted to its hypoallergenic nature and reasonable balance of amino acids of food protein by phosphate is that, the protein side chain groups,
(Hou et al., 2017). Thus, RP could represent an important component of including –OH groups of Ser, Thr and Tyr, ε-NH2 of Lys, the 1 and 3
the protein resources in daily life because a large amount of it remains nitrogen atoms of the His imidazole ring and the nitrogen atom of the
after rice processing. Additionally, RP shows good physicochemical Arg guanidine group, were selectively induced by a large number of
characteristic in foaming and emulsifying due to the presence of hy- phosphate groups (Fig. 1) (Sung, Chen, Liu, & Su, 1983). With the in-
drophobic and hydrophilic groups (Fabian & Ju, 2011). Hence, this creasing negative charges introduced by phosphate groups on the pro-
finding suggested that RP could be utilized in drinks and baked pro- tein molecular surface, the protein will be enhanced with respect to
ducts (Sharif, Butt, Anjum, & Khan, 2014). The physicochemical hydration. This step leads to the improvement in water solubility, water


Corresponding author.
E-mail address: huaxiong100@126.com (H. Xiong).

https://doi.org/10.1016/j.foodhyd.2019.05.037
Received 12 January 2019; Received in revised form 25 April 2019; Accepted 21 May 2019
Available online 22 May 2019
0268-005X/ © 2019 Published by Elsevier Ltd.
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

Fig. 1. Phosphoesterification reaction between protein isolates and sodium trimetaphosphate (STMP) as cited in the work of Sung et al. (1983).

and oil retention capacity, emulsification and foaming properties of the 2. Material and methods
protein (Moure, Sineiro, Domínguez, & Parajó, 2006). To date, the
study of phosphorylation on RP modification is limited, as well as ap- 2.1. Material and chemical reagents
plications in the food industry. Sodium tripolyphosphate (STP), sodium
trimetaphosphate (STMP) and POCl3 are regarded as safe food additives RP was extracted by 0.05 M NaOH solution and isoelectric pre-
and could represent economical reagents for large-scale production of cipitation from defatted rice bran (the basic information of rice bran
modified proteins in the food industry (Kato, Aoki, Kato, Nakamura, & shown in Table S1), which was purchased from local farmer market,
Matsuda, 1995). Meanwhile, STMP can be more efficient in chemical and the protein purity of RP was 88.9% (determined by Kjeldahl Ni-
reactions compared to the straight chain polyphosphates due to its ring- trogen, N × 5.95), of which the contamination was starch and minerals.
shaped molecular structure (Sánchez-Reséndiz et al., 2018). Hence, we 1-anilio-8naphthalenesulphonate (ANS) was of analytical grade. Whey
proposed to improve the functional properties of RP by phosphorylation protein isolate (WPI; Mingrui Chemicals Company, Zhenzhou, Henan
with STMP. province, China) and STMP was purchased from Sigma-Aldrich In-
The proper pH range for the reaction of protein molecules with dustrial Co. Ltd. (Shanghai, China). The Reagent C for the determina-
STMP spans from 7 to 9 regarding thermal treatment in aqueous solu- tion of phosphorylation degree, was prepared freshly each time, by
tion (Cheng et al., 2017). However, RP is composed of 37% albumin mixing 10 mL of 6 M sulfuric acid with 20 mL of distilled water and
(water-soluble), 31% globulin (salt-soluble), 2% prolamin (alcohol-so- 10 mL of 2.5% ammonium molybdate, then adding 10 mL of 10% as-
luble) and 27% glutelin (alkali-soluble protein) (Fabian & Ju, 2011). corbic acid. The soybean oil was bought in the supermarket. All re-
Thus, it seems that the large proportion of protein fractions (primarily agents used here were of analytical grade or higher.
glutelin) could not react sufficiently with STMP under near-neutral
environments due to its low solubility and dispersibility. It has been 2.2. Preparation of phosphorylated RP
reported that the solubility of rice bran glutelin could be improved by
the phosphorylation under hydrothermal treatment at 55 °C, pH 9.7 RP with 2% (w/v) dispersion was added to 0.1 M STMP at different
(Ma, Na, Cheng, & Wang, 2017). Technically, phosphorylation might pH (3.0, 5.0, 7.0, 9.0 and 11.0) with adjustment of 1 M NaOH and/or
rigidify the RP structure by introducing additional favourable contacts HCl. The mixture was continuously stirred at 55 °C for 2 h in water bath.
between RP residues, which would be accompanied by increasing the After the neutralization with NaOH and/or HCl, the samples were
exposure of hydrophobic side chains, resulting in decreasing its solu- dialyzed to remove free STMP for 72 h against deionized water (sam-
bility (Chen, Liu, Wang, Ye, & Shen, 2017). Chen, Liu, Wang, Ye, and ples/deionized water, 1:100), which was changed every 6 h. After
Sheu (2017) conduct phosphorylation to rice protein, of which the lyophilizing, the samples were collected, named as P-RP3, P-RP5, P-
condition was at 35 °C, pH 11.5, and 20.6% of Ser residues were RP7, P-RP9 and P-RP11, respectively, and raw RP as the control. Then
phosphorylated while no enhancement of solubility was observed, or the samples were stored at −20 °C for further studies.
even worse. Hence, the improvement of solubility and emulsifying
properties did not only count on the phosphorylation degree, but also
2.3. Physicochemical properties of phosphorylated RP
depend on the structure of protein after the modification. In the present
work, we aim to study the effect during phosphorylation under hy-
2.3.1. Determination of phosphorylation degree
drothermal treatment on the modification of RP considering the solu-
The phosphorylation degree was represented by the phosphorus
bility and emulsifying activity. The study was also conducted to eluci-
content of sample and determined according to previously report with
date the presence and mechanism of phosphorylation during the
minor modification (Xiong, Zhang, & Ma, 2016). 0.5 g of samples was
chemical modification by the methods of fluorescence, circular di-
suspended in perchloric acid and left overnight at room temperature.
chroism (CD), Fourier transform infrared (FT-IR) spectroscopy, scan-
Thereafter, the suspension was wet-ashed at 170 °C with additional HCl
ning electron microscopy (SEM) and X-ray photoelectron spectroscopy
until it became transparent. The solution was diluted with deionized
(XPS) analysis. Furthermore, we evaluate the potential utilization of the
water to 5 mL and filtered through a 0.45 μm syringe filter. 4 mL of
modified product within emulsions in terms of stability under different
reagent C was added into the digested samples and capped with Par-
environmental condition. It is expected that the work could be helpful
afilm and incubated in a 37 °C water bath for 1.5 h. Then cool to room
to understand the potential utilization of protein modifications by
temperature and read the absorbance in UV/Vis spectrum (TU1810,
phosphorylation.
Beijing Purkinje General Instrument LTD Co., China) at 820 nm against
the blank. The standard curve for the quantitate was prepared by the
phosphorus standard solution. For the determination of inorganic
phosphorus, 5 mL of 12% trichloroacetic acid was added to the same

289
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

volume of 10 g/L sample solution, which was centrifuged at 10,000 g to 400 nm with an excitation wavelength of 280 nm.
for 15 min. The amount of phosphorus bound to proteins was calculated
by the difference between the total phosphorus and the phosphorus 2.4.2. Circular dichroism (CD) measurement
content in supernatant and presented as mg/g in terms of protein. CD profile of samples were obtained from a 0.1 mg/mL of protein
sample dispersed in PBS (50 mM, pH 7.0) by a Bio-Logic MOS-450 CD
2.3.2. Determination of zeta potential spectrometer (French Bio-Logic SAS, Claix, French) with a 1.0 mm path
The zeta potential of samples was measured by a Zetasizer Nano-ZS length quartz cuvette. The samples were scanned from 190 to 250 nm to
instrument (Malvern Instruments, Worcestershire, UK). Protein sample gain the far-UV CD spectra with PBS as the reagent blank. The data was
was added into PBS solution (50 mM, pH 7.4) to a final concentration of analysed by the online SELCON3 method.
0.05% (w/v). After filtration through a 0.45 μm porous membrane, the
diluted samples were induced directly into the chamber of Zetasizer 2.4.3. Fourier transform infrared (FT-IR) analysis
Nano-ZS particle electrophoresis instrument prior to zeta potential The FT-IR information of samples were recorded by an FTIR spec-
analysis at 25 °C. Zeta potentials were calculated from the electro- trometer (Nicolet Nexus 470, Nicolet Co., USA) with scanning the
phoretic mobility data obtained from Smoluchowski’ equation. whole band of (400-4000 cm−1) at room temperature. In brief, a 2 mg
Measurements were carried out in triplicate for the time interval of 30 s of samples was mixed with KBr and pressed into pellet. The transmit-
for 100 zeta runs. tance intensity was obtained in the wavenumber range of
4000–400 cm−1 at 4 cm−1 resolutions. Infrared spectrogram of sample
2.3.3. Protein solubility was a superposition of 32 times and the KBr spectrum as the back-
Lowry method was conduct here to determine protein solubility ground. All samples prior to determination were baseline corrected,
with bovine serum albumin as standard (Lowry, Rosebrough, Farr, & after background correction.
Randall, 1951; (Xu et al., 2016). In brief, protein sample was diluted in
Tris-HCl (50 mM, pH 7.0) to make 1% aqueous solution (w/v) and kept 2.4.4. Scanning electron microscopes (SEM) analysis
stirring magnetically at room temperature for 30 min. After the cen- The microstructure of sample was determined by SEM (Quanta
trifugation at 10,000×g for 15 min, the protein content of supernatant 200F, FEI, Hillsboro, OR, USA) according to our previous work (Peng
was determined by Folin & Ciocalteu's reagent. The result was calcu- et al., 2010). The samples were sprinkled onto a double-sided tape and
lated by the division of protein content in supernatant and total protein sputter-coated with a 5 nm-thick gold layer.
in samples.
2.4.5. X-ray photoelectron spectroscopy (XPS) analysis
2.3.4. Emulsifying activity and emulsion stability of sample after The surface chemical composition was analysed by a VG Multilab
phosphorylation 2000 X-rays at 300 W with the axis of the energy analyzer normal to the
Emulsifying activity and stability were determined by the method plane of the sample surface. Elemental surface compositions were ob-
from Zhao et al. (2012) with some modification. In brief, 16 mL of 0.1% tained at a pass energy of 25 eV with a 0.05 eV resolution. All spectra
protein aqueous solution (50 mM PBS, pH 7.0) and 4 mL of soybean oil were recorded by using double anode Al target and 0.05 eV resolution.
were mixed and homogenized at 12,000 rpm for 1 min with a high-
shear mixer (IKA-T18, IKA-Werke GmbH&Co., Staufen, Germany). A 2.5. Influence of environmental stress on emulsion stability
50 μL volume of pre-emulsion was dispersed into 5 mL of 0.1% SDS (w/
v) and read at 500 nm for the absorbance with 0.1% SDS solution as The preparation of emulsion and the assessment of emulsion stabi-
blank by a UV spectrophotometer (TU-1810, Beijing Puxi General In- lity under environmental stress were performed according to the
strument Co., Beijing, China). The absorbance values measured record method of Xu et al. (2016) with minor modification.
at 0 min (A0) and 10 min (A10) after the formation of pre-emulsion were
used to calculate the emulsifying activity index (EAI) and the emulsion 2.5.1. Preparation of emulsion
stability index (ESI): Protein samples were dispersed into 10 mM PBS (pH 7.0) and stirred
for 3 h to completely hydrate. After centrifugation at 3000×g for
2 × 2.303
EAI (m2 /g) = × A0 × D 15 min, the supernatant was adjusted to the concentration of 5 mg/mL
C × (1 − ∅) × 10 4 (1) determined by Lowry method. Then a 90 mL of solution was mixed with
10 × A0 10 mL of soybean oil and pre-homogenized at 12,000 rpm for 2 min
ESI (%) = with a high-shear mixer (IKA-T18, IKA-Werke GmbH&Co., Staufen,
A0 − A10 (2)
Germany) and emulsified by a Microfluidizer (NCJJ-0.007/200,
Where A0 and A10 represent the absorbance at 500 nm at 0 min and Langfang Tongyong Machinery Manufacturing Co., China) at 80 MPa
10 min, respectively. D was the dilution times and the value was 100 and repeated 3 times. 0.01% (w/v) of sodium azide was added into the
here. C is the protein concentration (g/mL) before homogenizer, and ϕ emulsions to protect from the growth of microbial growth.
is the oil volume fraction (v/v) of the pre-emulsion (ϕ = 0.20).
2.5.2. pH stability
2.4. Structural characteristics of modified RP by phosphorylation After adjusting pH at the range of 3–9 by 1M HCl and NaOH solu-
tions. The emulsions were transferred to 15 mL tube and stored 25 °C
2.4.1. Surface hydrophobicity and intrinsic fluorescence for 24 h prior to particle size and zeta potential analysis.
Surface hydrophobicity (H0) was measured according to the de-
scription of Zhao, Zhao, Chen, and Xiong (2019) with ANS as fluores- 2.5.3. Ionic strength stability
cence probes. The fluorescence intensity was measured at wavelength Freshly prepared emulsions were transferred into 15 mL tubes and
of 390 nm as the excitation and 470 nm as emission by a fluorometer modified the ionic strength by adding 6 M NaCl solution to lead various
(F7000, Hitachi Co., Tokyo, Japan). H0 was the initial slope calculated final NaCl concentrations (0–300 mM). Emulsions were vortexed and
by a linear regression analysis of fluorescence intensity versus protein stored for 24 h at 25 °C prior to particle size.
concentration (%, w/v).
Intrinsic fluorescence was gained from a 0.01 mg/mL of protein 2.5.4. Temperature stability
sample dispersed in PBS (50 mM, pH 7.0) by a fluorometer. The Emulsion were transferred into 15 mL tubes and incubated in a
fluorescent intensity was recorded as the emission wavelength from 300 water bath at different temperatures (30–90 °C) for 30 min. After

290
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

cooling to room temperature, samples were vortexed and stored for which was significantly higher than raw RP (0.049 ± 0.006 mg/g)
24 h at 25 °C prior to particle size. (p < 0.05). According to the amino acid analysis, most of amino acids
were influenced after phosphorylation, especially Ser (Table S2). It
2.5.5. Storage stability seems that amino acids of protein side chain groups reacted with
Emulsion were transferred into 15 mL tubes and incubated at dif- phosphate group by phosphorylation and generated phosphor-amino-
ferent temperatures (4 °C, 25 °C, 60 °C) for 7 days prior to particle size. acid residues, leading to the decrease of amino acid contents, such as
Ser (Fig. 1) (Sung et al., 1983). Hydrothermal phosphorylation has been
2.6. Statistical analysis widely used for protein modification in the food industry and numerous
industrial applications. To the best of our knowledge, the proper pH for
All determination was conduct in triplicate and presented as the reaction of protein molecules with STMP is within a range of 7–9
mean ± SD. Meanwhile, one-way analysis of variance (ANOVA) with under thermal treatment (Cheng et al., 2017). However, the optimized
Duncan test were performed by SPSS statistics 20.0 (SPSS Inc., Chicago, pH of reaction differs among various proteins. Ma et al. (2017) sug-
US). Differences were considered significant at p < 0.05 among the gested that the optimal pH of phosphorylation for rice bran glutelin
samples. treated by STP was 9.7 according to single factor experiments. This
finding was similar to our findings in this study. The phosphorus con-
tents of modified RP reacted above pH 7 were higher than that in the
3. Results and discussion
acidic solution environment (Fig. 2A). The amino or hydroxyl groups in
the side chains of lysine or serine are at least partially deprotonated and
3.1. Effect of phosphorylation on physicochemical properties of modified RP
lead to the enhancement of contact with STMP when above pH 8
(Feeney, 1987). When phosphorylated modification was performed at
3.1.1. Total phosphorus content
pH levels higher than 12, deprotonation of OH groups of serine residues
Hydrothermal phosphorylation was conducted in modifying RP
reached equilibrium, and no further improvement of phosphorylation
under different pH conditions with STMP, and the total phosphorus
was observed (Sánchez-Reséndiz et al., 2018). This finding is poten-
content of modified RP was summarized in Fig. 2A. It has been found
tially the reason why the phosphorus content of P-RP11
that the highest content of phosphorus in modified RP was achieved
(8.16 ± 0.476 mg/g) was significantly lower than P-RP9 here
when the reaction was performed at pH 9.0 (9.53 ± 0.522 mg/g),

Fig. 2. Phosphorus content (A), zeta potential (B), protein solubility (C) and the emulsifying properties (EAI and ESI) (D) of phosphorylated RP under treatments at
pH 3.0, 5.0, 7.0, 9.0 and 11.0. Each of the values was determined in triplicate and presented as the mean ± SD, and different lowercase letters indicate significant
differences between each group (p < 0.05).

291
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

Table 1 Table 2
Secondary structure content of phosphorylated RP compared with native RP by Secondary structure content of phosphorylated RP compared with native RP by
CD analysis. FTIR analysis.
Samples α-helix (%) β-sheet (%) β-turn (%) Random coil (%) Samples α-helix (%) β-sheet (%) β-turn (%) Random coil (%)

RP 20.4 38.6 16.5 24.3 RP 21.2 47.9 3.2 27.8


P-RP3 16.2 28.6 24.0 31.2 P-RP3 21.2 43.5 11.0 24.2
P-RP5 10.7 34.2 20.8 34.2 P-RP5 20.2 43.3 10.8 25.7
P-RP7 23.1 36.8 16.5 23.6 P-RP7 20.9 46.8 6.4 26.0
P-RP9 34.5 11.0 23.1 31.3 P-RP9 22.8 38.4 14.6 24.3
P-RP11 34.3 11.4 22.9 31.3 P-RP11 23.9 36.5 14.7 24.9

(p < 0.05). Nevertheless, some proteins have been reported to contain zeta potential for phosphorylated RP increased from −4.0 mV (RP) to
percentages of phosphorylation that are higher at pH values above 12 −15.12 mV (P-RP3), −16.1 mV (P-RP5), −20.1 mV (P-RP7),
than at pH 11, such as peanut protein (Yu et al., 2015). Peanut protein −23.3 mV (P-RP9) and −19.2 mV (P-RP11). This finding could be at-
was regarded as difficult to modify due to its resistance to denaturation tributed to the different degree of phosphorylation in RPs, as we
(Feeney, 1987). Meanwhile, peanut protein was comparatively more mentioned above (Fig. 2A). Taken together, the greater the abundance
phosphorylated than other proteins because of its higher availability of of phosphate is introduced to RP, the higher the absolute value of zeta
serine residues (Sánchez-Reséndiz et al., 2018). It seems that proteins potential is increased, since the attachment of phosphate increased the
with different structural properties affect the degree of phosphorylation net negative charge. The value of the zeta potential can be significantly
in modifications of proteins under the same conditions. Interestingly, it related to the stability of colloidal dispersions and indicates the degree
has been reported that the solubility of rice bran glutelin could be of repulsion between similarly charged particles (Wu, Zhao, Yang, &
improved by the phosphorylation under hydrothermal treatment at Chen, 2014). In the emulsion system, the intensity of the electric
55 °C, pH 9.7 (Ma et al., 2017) and the phosphorus contents of phos- charges at the droplet surfaces was improved by the introduced nega-
phorylated product were at the range from 7 to 25 mg/g protein, which tively charged phosphate groups with a sufficiently strong electrostatic
is much higher than our results. The different phosphorus contents repulsion force between the droplets, leading to better steric stabiliza-
between Ma et al. (2017) and our results would be attributed to the tion and prevention of the aggregation and coalescence between dro-
objective material, pH of reaction and reagent dose and chemical plets (Khan, Mu, Zhang, & Arogundade, 2014; Xiong et al., 2016).
component for phosphorylation. First of all, the RP was gained from the Hence, the assessment of phosphorylated RP with respect to the utili-
rice bran and comprises albumin, globulin, prolamin and glutelin, zation of the emulsion could be conducted as a next step.
which Ma et al. only focus on the solo glutelin as the objective to study
the phosphorylation reaction. As we mentioned before, both of the 3.1.3. Solubility
composition of amino acid, protein structure and properties would in- The solubility of modified RP products was presented in Fig. 2C. All
fluence the reaction between phosphate and protein molecule. Sec- sample solubilities were improved after phosphorylated treatment,
ondly, the main purpose for Ma et al. is to optimize the condition of especially those of P-RP7 (42.3 ± 2.72%), P-RP9 (58.4 ± 2.61%) and
phosphorylation between rice bran glutelin and sodium tripolypho- P-RP11 (35.3 ± 3.44%), as compared to the control (RP,
sphate to improve the processing characteristics. The response surface 6.67 ± 1.46%). Even though the solubilities of P-RP3 (12.7 ± 1.10%)
method was designed according to the principle of Box-Benhnken and P-RP5 (15.4 ± 0.62%) were significantly higher than RP
central combination design based on the single factor experiments. (p < 0.05), the improvement of this property is insufficient for further
Hence, the results would be higher at optimized condition than ours, of utilization. RP shows good physicochemical characteristics with respect
which the purpose is to evaluate the solubility and emulsifying prop- to foaming and emulsifying due to the presence of hydrophobic and
erties to utilize in the formulation and production of natural emulsion- hydrophilic groups (Fabian & Ju, 2011). The hydrophilicity/hydro-
based products as emulsifier and reveal the possible mechanism of phobicity balance influences the protein solubility, which depends on
phosphorylation in RP. Thirdly, we conduct phosphorylation with 3% the amino acid composition, particularly at the protein surface (Moure
sodium trimetaphosphate, while 9% sodium tripolyphosphate was ap- et al., 2006). Under phosphorylated treatment, protein side chain
plied by Ma et al. The dynamic reaction is different with the dose and groups selectively induced a large number of phosphate groups. With
chemical species. Fourthly, the method of determination of phosphor- the increasing negative charges introduced by phosphate groups along
ylation degree was different. The details of Ma et al.’s method was that, the protein molecular surface, the hydration of protein will be en-
‘molybdenum blue colorimetric method was adopted to determine the hanced and lead to the improvement in water solubility (Moure et al.,
phosphorus content in RBG before and after modification. Using 2006; Sung et al., 1983). RP shows the lowest solubility at pI (iso-
phosphorus standard reserve. Combined phosphorus (mg/g) = (Phos- electric point, approximately 4.5 for RP), leading to the aggregation of
phorus content after modification - Phosphorus content before mod- protein molecules, since hydrophobic interactions between proteins are
ification)/sample mass ×100%’. Obviously, the free sodium tripoly- far greater than the hydrophilic and hydration repulsion forces pro-
phosphate, which could still retain with the protein as contamination, duced by charged residues in this condition. Therefore, the solubility of
was not taken into consideration by authors. This is quite different with RP is poor at acid condition around pI. Fewer amino residues could
ours, resulting the higher values of phosphorus probably. Taken to- react with STMP in acidic conditions near the pI than those in neutral
gether, it is difficult to discuss the differences between the results of Ma and alkaline environments. Phosphorus content and solubility of P-RP3
et al. and our work in terms of phosphorus content. and P-RP5 were lower than those of P-RP7, P-RP9 and P-RP11 (Fig. 2A
and C). Meanwhile, more negative charges were introduced by STMP at
3.1.2. Zeta potential the surface of the protein molecule when the pH value of the reaction is
Fig. 2B summarized the zeta potential values of modified RPs. The far from the pI of RP (Fig. 2B). Hydrophilic and hydration repulsion
zeta potential of RP was electronegative because the isoelectric point of forces can overcome hydrophobic interactions, enabling phosphory-
RP (approximately pH 4.5) was lower than the pH value of the buffer lated RP to maintain high solubility (Yu et al., 2015).
solution (PBS pH 7.0). In addition, the absolute values of the zeta po-
tentials of modified RPs show the same tendency with phosphorus 3.1.4. Emulsifying activity
content of modified RPs. Compared to raw RP, the absolute value of As previously mentioned, RP exhibits good physicochemical

292
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

Fig. 3. Surface hydrophobicity (H0) of phosphorylated RP under pH conditions of 3.0, 5.0, 7.0, 9.0 and 11.0 during treatments (A). Each value was determined in
triplicate and presented as the mean ± SD, and different lowercase letters indicate significant differences between each group (p < 0.05). Also shown are the
intrinsic fluorescence (B) and the FT-IR profile within the wavelength range of 4000–400 cm−1 (C) and the wavelength range of 1700–500 cm−1 (D).

Fig. 4. SEM images for RP (A), P-RP3 (B), P-RP5 (C), P-RP7 (D), P-RP 9 (E) and P-RP 11(F), respectively, with an internal scale of 1 μm.

characteristics with respect to emulsifying due to the presence of hy- higher net negative charge in the surface of RP (Fig. 2B). During the
drophobic and hydrophilic groups (Fabian & Ju, 2011). With the in- preparation of the emulsion, a higher absolute value of the zeta po-
crease of phosphate content introduced to RP molecules, it showed tential of droplets significantly enhances steric stabilization and

293
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

Fig. 5. Surface element content (a) and XPS spectra (536-528 eV) of RP and P-RP (pH 5.0, 7.0, 9.0) and peak fitting spectra of O1s in raw and phosphorylated RP (pH
5.0, 7.0, 9.0) (C–F).

prevents the aggregation and coalescence between droplets (Xiong an emulsion (Benelhadj, Gharsallaoui, Degraeve, Attia, & Ghorbel,
et al., 2016). Therefore, the emulsifying activity of RP should be im- 2016). With the induced negative charges of phosphate groups in
proved after phosphorylation. Fig. 2D shows the emulsifying properties protein chain residues, phosphorylated RP could move rapidly to the
of raw RP and phosphorylated RPs. The EAI of emulsions from raw RP oil/water interfaces and improve the adsorption of the protein onto the
and phosphorylated RP at 0 min were 1.70 m2/g (RP), 3.83 m2/g (P- oil-water interfacial layer, as well as the dispersion of the oil droplets,
RP3), 5.45 m2/g (P-RP5), 6.91 m2/g (P-RP7), 13.72 m2/g (P-RP9) and by increasing the electrostatic repulsion force of droplets (Thaiphanit &
13.01 m2/g (P-RP11), respectively, indicating that the emulsifying ac- Anprung, 2016). However, P-RP3 and P-RP5 showed lower emulsifying
tivity of RP was obviously improved by phosphorylation, as expected, activity when considering EAI values. This may be attributed to the low
and that P-RP9 showed the highest emulsifying activity when compared solubility and minimal negative charge introduced. The protein could
to other treatments. The emulsifying activity reflects the rapid ad- not be sufficiently hydrated to play an important role as emulsifier.
sorption of proteins at the oil/water interfaces during the formation of The ESI of raw and phosphorylated RPs over a time period of 10 min

294
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

Fig. 6. Influence of temperature on the mean particle size of the P-RP9 emul-
sion and WPI emulsion (A), visual graphs of P-RP9 emulsions (B) and WPI (C)
emulsions, respectively.

were 50.0% (RP), 49.4% (P-RP3), 53.9% (P-RP5), 85.2% (P-RP7),


83.6% (P-RP9) and 89.5% (P-RP11), respectively. The highest ESI was
observed at pH 7.0, 9.0 and 11.0, and no significant difference was Fig. 7. Influence of pH on the mean particle sizes of the P-RP9 emulsion and
found among them (p > 0.05). This finding indicated that ESI im- WPI emulsion (A), visual graphs of P-RP9 emulsions (B) and WPI (C) emulsions,
proved with increasing phosphorylation pH and reached a platform respectively.
after pH 7.0. ESI referred to the ability of emulsions to remain dispersed
without coalescing, flocculating and creaming, and ESI values were
3.2. Mechanism of phosphorylation by STMP in modified RP
found to be positively correlated with protein surface charge and so-
lubility (Karaca, Low, & Nickerson, 2011; Patel & Kilara, 1990).
3.2.1. Surface hydrophobicity and intrinsic fluorescence
Meanwhile, the hydrophilicity/hydrophobicity balance influences the
To confirm whether the hydrothermal treatment with STMP influ-
protein emulsifying activity, as well. After the proteins are absorbed
enced the secondary and tertiary structures of protein, surface hydro-
and move rapidly to oil/water interfaces, the surface hydrophobicity of
phobicity (H0) and intrinsic fluorescence were taken into consideration.
the protein molecule is important to stabilize the oil droplet through
Surface hydrophobicity is a measure of the exposure of hydrophobic
hydrophobic interaction. Even though P-RP11 shows lower solubility
amino acids in the protein chain, which were otherwise buried within
and phosphorus content than that of P-RP9 (Fig. 2A and C), both EAI
the interior (Yerramilli, Longmore, & Ghosh, 2017). ANS was utilized
and ESI of P-RP11 exhibit similarity to P-RP9 (p > 0.05). This result
here as a fluorescent probe, which non-covalently binds to the surface
might be attributed to protein unfolding and increasing exposure of
hydrophobic regions of protein. The H0 values of raw RP and phos-
hydrophobic moieties, leading to the improvement of the emulsifying
phorylated RP under pH treatment were shown in Fig. 2A. Compared to
stability at pH 11.0. With better hydrophilic-lipophilic balance, the oil
raw RP and P-RP5, phosphorylated RP at pH 3.0, i.e., P-RP3, was ex-
droplet surface is easier to anchor by protein (Wan, Wang, Wang, Yuan,
posed to more hydrophobic moieties, an effect which was primarily due
& Yang, 2014). If the hypothesis is correct, the hydrothermal treatment
to aggregation of protein at acidic condition (p < 0.05). Compared to
with STMP should affect the secondary and/or tertiary structure as
raw RP (213), it was found that H0 values of phosphorylated RP im-
well, which could be proven by further analysis.
proved with increasing phosphorylation pH within the range of
5.0–11.0, indicating that protein unfolding leads to the change of
protein tertiary structure upon hydrothermal treatment with STMP.
Fluorescence intensity could play an important role as index of the

295
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

extent of exposure of aromatic amino acids to water, which is related to


changes of protein tertiary conformation (Cui, Zhao, Yuan, Zhang, &
Ren, 2013). The emission fluorescence spectroscopy profile of raw and
phosphorylated RPs were shown in Fig. 2B. The fluorescence intensities
of phosphorylated RPs were higher than that of raw RP, except P-RP5,
because of more aromatic groups being exposed at the surface of pro-
tein molecules and increasing emission fluorescence (Xu et al., 2016).
The fluorescence intensity of phosphorylated RP improved with in-
creasing phosphorylation pH within the range of 5.0–11.0. This result
was coincident with previous surface hydrophobicity data (Fig. 2A).
Through hydrothermal treatment with STMP, aromatic groups of RP
were exposed to water, indicating the changes of tertiary conformation
and partial unfolding of protein. However, the maximum emission
wavelength (λmax) of phosphorylated RP shifted from 350 nm to
343 nm. This indicates that the environment of aromatic amino acids of
protein chain residues was changed, and suggest covalent binding to
other groups (Iosin, Toderas, Baldeck, & Astilean, 2009). It has been
reported that λmax was blueshifted when the aromatic amino acid be-
came buried within the interior of the protein, as shown by the decrease
of fluorescence intensity (Chen, Zhao, Sun, Ren, & Cui, 2013). In the
present work, we supposed that aromatic amino acids would be exposed
to water after the hydrothermal treatment with STMP and reacted with
phosphorylated groups at the surface of protein molecules, as we
mentioned before, considering the high fluorescence intensity and
blueshifted λmax in phosphorylated protein here.

3.2.2. Secondary structure content


A deeper understanding of raw and phosphorylated RP secondary
structure was analysed using the far-UV CD spectrum, which is a gen-
eral technique to predict protein secondary structure which is particu-
larly useful for soluble proteins (Wang, Zhang, Wang, Wang, & Chen,
2015; Xu et al., 2016). Only slight differences were observed in phos-
phorylated RP (P-RP3, P-RP5 and P-PR7) comparing to raw RP (Fig.
S1). It seems that the secondary structure of protein was changed a little
under hydrothermal treatment by phosphorylation at acidic or around
neutral condition. However, the CD profiles of P-RP9 and P-RP11
showed quit different than others with strong negative bands from 200
to 240 nm observed. Generally, α-helical proteins have negative bands
Fig. 8. Influence of NaCl on the mean particle sizes of the P-RP9 emulsion and at 222 nm and 208 nm and a positive band at 193 nm (Greenfield,
WPI emulsion (A), visual graphs of P-RP9 emulsions (B) and WPI (C) emulsions, 2006). It indicated that the α-helix content increased after hydro-
respectively. thermal treatment by phosphorylation. It has been reported that the β-
sheet structures are relatively stable, whereas the α-helix, β-turn, and
random coil structures are relatively flexible and open, as reported
previously (Yong, Yamaguchi, & Matsumura, 2006). To evaluate the

Fig. 9. Influence of storage time on the mean particle sizes of P-RP9 emulsion and WPI emulsion (A), visual graphs of P-RP9 emulsions (B) and WPI (C) emulsions
with different temperatures after 7 days of storage.

296
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

changes, the contents of secondary structure were calculated by the caused the changes in secondary structure (Tables 1 and 2). To verify
online SELCON3 method with CD data. Compared with raw RP, the β- changes in protein structure due to the phosphorylation in RP, the
sheet content in phosphorylated RP decreased from 38.6% to 11.4% details in the major IR-region (1400-800 cm−1) were traced in Fig. 3D.
upon hydrothermal treatment with STMP, whereas the contents of α- The bands at ∼1395 cm−1, which were assigned to the stretching vi-
helix, β-turn and random coil increased from 20.4% to 34.3%, bration of C–N, showed differential absorption, indicating that the
16.5%–22.9% and 24.3%–31.3%, respectively, indicating that hydro- atoms or groups that were bound to C–N of RP changed after phos-
thermal treatment with STMP affected the secondary structure and led phorylation. Meanwhile, the peaks at approximately 1238 cm−1 were
to a more extended form (Table 1). However, the precision of analysis assigned to the stretching vibration of P=O, while the peaks at ap-
should be based on the purity of samples (⩾ 95%) and the accurate proximately 923 cm−1 were attributed to the stretching vibration of
molecular weight of protein (Greenfield, 2006). Only imprecise data P–O in phosphorylated RP (Xiong et al., 2016). Taken together, it was
about secondary structure contents could be obtained here in terms of shown that STMP changed the FT-IR profiles of RP due to modifications
the mixture samples. within their chemical structures.
Considering the low solubility of raw RP and some phosphorylated
forms of RP, far-UV CD spectra cannot provide sufficient information 3.2.5. XPS analysis of modified RP product
regarding the overall changes in proteins, since far-UV CD spectra The surface element content (A) and the XPS spectra (536-524 eV)
merely reflect water-soluble subunits and only the rough data for (B) of raw and phosphorylated RP (pH 5.0, 7.0 and 9.0) and peak fitting
mixture protein showed here. FTIR is a powerful technique that pro- spectra of O1s in raw and phosphorylated RP (pH 5.0, 7.0 and 9.0) (C–F)
vides comprehensive insights into integrated structures, especially for were presented in Fig. 5. The surface phosphorus content in RP in-
solid samples (Wang et al., 2015). Amide Ⅰ bands (1700-1600 cm−1), creased after phosphorylation, and the surface phosphorus atomic
according to FTIR analysis, are highly related to the secondary structure percentages were 0.46% (P-RP5), 0.65% (P-RP7) and 0.6% (P-RP9),
of protein (Hu et al., 2018). Accordingly, the deconvolved spectrum of respectively, as compared to the control (0.14% RP). This suggested
the amide Ⅰ band was iteratively fitted with Gaussian band shapes that the phosphate group was successfully introduced to RP molecule
(Byler & Susi, 1986), and the corresponding secondary structures were surfaces. Meanwhile, it can be observed that the binding energy of O1s
summarized in Table 2. Compared with raw RP, the β-sheet content in in phosphorylated RP (pH 5.0 and pH 9.0) shifted to ∼532.2 and
phosphorylated RP decreased from 47.9% to 36.5 by hydrothermal ∼532.4 eV, respectively, compared with raw RP (∼532.6 eV) ac-
treatment with STMP, whereas the contents of α-helix and β-turn in- cording to Fig. 4B, suggesting that the chemical environment of oxygen
creased from 21.2% to 23.9% and 3.2%–14.7%, respectively. When atoms changed after phosphorylation through a distortion of the elec-
exposed to extreme pH (⩾11) conditions, many globular proteins un- tron clouds of oxygen atoms (Xiong & Ma, 2017). The binding energy
dergo significant conformational changes, known as the molten globule and chemical state of O1s in raw and phosphorylated RP (∼536 eV-
(MG) state. The MG state maintains most secondary structures but tends ∼524 eV) were summarized in Fig. 5 (C–F) by peak fitting. Two peaks
to unfold and lose some of the tertiary structures (Goto, Calciano, & at ∼532.9 and ∼531.7 eV were assigned to the chemical states of
Fink, 1990). Generally, β-sheets could be formed in aggregated protein C–OH and C=O in raw RP (Fig. 5C), respectively. With respect to the
molecules, whereas turn structures were considered to be a product of effect of phosphorylation on RP, three peaks were observed in P-RP7
the protein unfolding of any higher ordered structures (Ellepola, Choi, and P-RP9 at ∼539, ∼532.8 and ∼531.7 eV. The peaks around
& Ma, 2005). Even though the data regarding secondary structure ∼532.8 and ∼531.7 eV were assigned to the chemical states of C–OH
contents showed slightly different results between FTIR and CD ana- and C=O by comparison to that in raw RP. However, Xiong and Ma
lysis, both of the results tended towards the decrease of relatively stable (2017) conjectured that the peak at ∼528.8 eV might be attributed to
structure (β-sheet) and increase of flexible and open structure (such as the O–P bond which may be formed between ovalbumin and STP. In
α-helix and β-turn) within protein secondary structure. This finding addition, it has been reported that the peaks of bridging (P–O–P) and
suggested that phosphorylated RP forms were unfolded and that RP was non-bridging (PO-) oxygen were ∼534 and ∼532 eV, respectively
disaggregated by hydrothermal treatment with STMP. (Brow, Kirkpatrick, & Turner, 1990). Hence, the third peak at ∼539 eV
might be attributed to the O=P bond, considering the analytical results
3.2.3. Microstructure of FT-IR and XPS.
Fig. 4 shows scanning electron micrographs (SEM) of raw and
phosphorylated RP: raw RP (A), phosphorylated RP at pH 3.0 (B), 5.0 3.3. Assessment of emulsifying stability of RP by phosphorylation under
(C), 7.0 (D), 9.0 (E) and 11.0 (F), respectively. Compared to raw RP, different environmental stress
aggregated protein surfaces were observed in P-RP3, P-RP5 and P-RP7.
This was coincident with our findings with respect to secondary The physical state of oil-in-water emulsions under various en-
structure change, since β-sheet content in phosphorylated RP below vironmental stresses was studied in this section. We assumed a com-
neutral pH was shown to be higher than that in the alkaline environ- parison between phosphorylated RP (P-RP5, P-RP7 and P-RP9) with
ment (Tables 1 and 2). Interestingly, the surface profiles of P-RP9 and raw RP in terms of emulsifier. Unfortunately, preliminary experiments
P-RP11 were different from the aggregation pieces formed. Ma et al. showed that RP could not form emulsions, and that the emulsion sys-
(Ma et al., 2017) showed the same results when phosphorylation was tems separated rapidly into oil and aqueous phases. Furthermore, the
performed in rice bran glutelin at pH 9.7. This indicated that RP were emulsion consisting of P-RP5 and P-RP7 is not stable during storage
unfolded during phosphorylation an d that the changes in protein over 7 days (data not shown). This finding suggested that the raw RP, P-
conformation differed with various pH conditions. RP5 and P-RP7 were unsuitable as emulsifiers with the notable excep-
tion of P-RP9. To assess the emulsion regularly, whey protein isolate
3.2.4. FT-IR analysis (WPI), which is widely utilized in the food industry as a commercial
FT-IR analysis has been widely used to study the protein chains emulsifier, was used for the emulsion stability studies as contrast.
when the presence of phosphate esters is known or presumed (Sánchez-
Reséndiz et al., 2018). The FT-IR profiles of raw and phosphorylated RP 3.3.1. Temperature
at the wavelength range of 4000–400 cm−1 were presented in Fig. 3C. The effect of temperature on the mean particle diameter and
Clearly, the band in the range of 1700–1600 cm−1 was assigned to the creaming stability of the emulsions was analysed and shown in Fig. 6.
Amide Ⅰ bands, which were quite different between raw and phos- The initial mean diameters of the droplets coated by P-RP9 and WPI
phorylated RPs. As we discussed with respect to the secondary structure after homogenization were 981 nm and 731 nm at room temperature
of phosphorylated RP, the conformation of RP was influenced and (25 °C). This suggested that phosphorylated RP could be absorbed to the

297
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

droplet surfaces more rapidly during homogenization, or that they were the exploration of phosphorylated proteins in food emulsions, further
more effective at preventing re-coalescence inside the homogenizer (Xu research should be conducted.
et al., 2016). Both of the emulsions coated by P-RP9 and WPI showed
relative stability to thermal processing within the range of 25–60 °C, 4. Conclusions
with a slight change in mean particle diameter. However, a sharp in-
crease in the particle size of the emulsions prepared with P-RP9 was This work demonstrated that the solubility and emulsifying activity
observed from 985 nm to 1400 nm, in contrast to the mild change in of RP were improved by hydrothermal phosphorylation with STMP,
emulsions coated by WPI from 731 nm to 914 nm. The increase in almost 8.7 times and 8.1 times, respectively, which could represent an
droplet aggregation in samples may be attributed to an increase in effective emulsifier candidate utilized in the food industry. Through the
exposure of hydrophobic groups during heating, leading to an increase phosphorylation, large amounts of phosphate accumulated and were
in electrostatic attraction and/or formation of disulfide bonds and at- introduced into the surface of protein molecules, increasing the abso-
tenuation of the electrostatic repulsion between the droplets (Charoen lute zeta potential value. Meanwhile, the protein structure unfolded,
et al., 2011; Zhang, Wu, & Yang et al., 2012). Fortunately, visual ob- showed more flexibility, and was transformed from the insoluble ag-
servation of the emulsions showed that they were stable with respect to gregates present in the raw material to soluble components in the
creaming and that no distinct phase separation occurred. phosphorylated protein under alkaline condition with thermal treat-
ment. In addition, buried hydrophobic groups were exposed to the
3.3.2. pH outer surface of the molecule and/or reacted with phosphate groups,
The influence of pH on the mean particle diameter and visual graphs leading to the increase of surface hydrophobicity and intrinsic fluor-
of the emulsions were presented in Fig. 7. Both of the emulsions pre- escence with blue shift. FT-IR spectra and XPS analysis indicated the
pared by P-RP9 and WPI showed similar trends in the particle size presence of phosphate esters (P=O), confirming the modification of
versus pH profiles. The mean particle diameter of the emulsion pre- protein structure. Phosphorylated RP prepared at pH 9.0 showed good
pared by P-RP9 decreased from 2285 nm to 981 nm, with an increase in emulsifying activity and stability. However, oil droplets stabilized by P-
the pH value, and the emulsion prepared by WPI decreased from RP9 were unstable with respect to aggregation at pH values in the vi-
1849 nm to 750 nm. Extensive droplet aggregation occurred from pH 3 cinity of the isoelectric point of the modified product and with the
to 5, and the emulsions were relatively stable from pH 7 to 9. It was also addition of salt and high temperature. Further research is required to
observed that the emulsions separated into a bottom transparent serum optimize the preparation of phosphorylated RP and improve the sta-
layer and a top white cream layer from 3 to 5 but were stable in re- bility of the emulsions.
sponse to creaming at pH 7 and 9. The protein emulsifier showed low
emulsifying activity, and relevant emulsions are highly unstable to Conflicts of interest
aggregation and creaming near the protein isoelectric points, which
occurs at pH values of approximately 3–5 for RP and WPI. Furthermore, The authors declare no conflicts of interest.
increasing creaming velocity with particle size leads to the creaming
instability at lower pH (McClements, 2015). Acknowledgement

3.3.3. Ionic strength This study was supported by grants from the Research Program of
The influence of ionic strength on the particle sizes of the emulsions State Key Laboratory of Food Science and Technology, Nanchang
was measured (Fig. 8). The WPI emulsion is stabilized through the University (Project No. SKLF-ZZA-201609), the Natural Science
addition of NaCl. The emulsion prepared by P-RP9 showed little change Foundation of Jiangxi Province of China (20171BBF60047), Key project
in particle size with increasing salt content in the range from 0 to ‘5511’ for Science and Technology Research of Jiangxi Province
50 mM NaCl. In contrast, a sharp increase of droplet sizes occurred in (S2018ZDYFE0040) and Major science and technology project of
the range from 100 to 300 mM NaCl with phase separation. Droplet academy of sciences in Jiangxi (2018-YZD1-05).
aggregation can be attributed to the ability of the counter ions in the
salt to screen the electrostatic repulsion between the negatively charged Appendix A. Supplementary data
droplets (Xu, Luo, Liu, & McClements, 2017). When the NaCl con-
centration was above a critical level, the electrostatic repulsion could Supplementary data to this article can be found online at https://
no longer overcome the attractive interactions (van der Waals and doi.org/10.1016/j.foodhyd.2019.05.037.
hydrophobic) between the droplets, thereby leading to aggregation.
References
3.3.4. Storage stability
Through a 7 days of storage test period at temperatures of 4 °C, 25 °C Benelhadj, S., Gharsallaoui, A., Degraeve, P., Attia, H., & Ghorbel, D. (2016). Effect of pH
and 60 °C, changes in the sizes of the droplets in the emulsions were on the functional properties of Arthrospira (Spirulina) platensis protein isolate. Food
Chemistry, 194(1), 1056–1063.
measured and presented in Fig. 9. Compared to the WPI emulsions, Brow, R. K., Kirkpatrick, R. J., & Turner, G. L. (1990). Local-structure of xal2o3.(1-x)
which were stable throughout 7 days of storage at 4 °C and 25 °C, P-RP9 napo3 glasses - an nmr and xps study. Journal of the American Ceramic Society, 73(8),
emulsions were shown to be stable systems, with aggregation of dro- 2293–2300.
Byler, D. M., & Susi, H. (1986). Examination of the secondary structure of proteins by
plets occurring from the 3rd day. Even though both of the emulsions deconvolved FTIR spectra. Biopolymers, 25(3), 469–487.
coated by P-RP9 and WPI showed relative stability to thermal proces- Charoen, R., Jangchud, A., Jangchud, K., Harnsilawat, T., Naivikul, O., & McClements, D.
sing within the range of 25–60 °C, with a slight change in mean particle J. (2011). Influence of biopolymer emulsifier type on formation and stability of rice
bran oil-in-water emulsions: Whey protein, gum Arabic, and modified starch. Journal
diameter (Fig. 6), the particle sizes of P-RP9 and WPI emulsions in- of Food Science, 76(1), E165–E172.
creased to approximately 2000 nm according to the extent of storage. Cheng, J., Xie, S., Yin, Y., Feng, X., Wang, S., Guo, M., et al. (2017). Physiochemical,
Slight phase separation was observed in the P-RP9 emulsion on the 7th texture properties, and the microstructure of set yogurt using whey protein–sodium
tripolyphosphate aggregates as thickening agents. Journal of the Science of Food and
day (Fig. 9B). Storage time may alter interfacial properties of emulsions
Agriculture, 97(9), 2819–2825.
due to the rearrangement or desorption of the protein molecules at the Chen, L. Y., Liu, S. C., Wang, Y. Z., Ye, C. Q., & Sheu, F. (2017). Phosphorylation does not
droplet surfaces (Huang et al., 2018). It has been reported that the improve the solubility of rice protein. Cereal Chemistry Journal, 94(4), 733–739.
addition of anionic polysaccharides, such as xanthan gum, could im- Chen, N., Zhao, M., Sun, W., Ren, J., & Cui, C. (2013). Effect of oxidation on the emul-
sifying properties of soy protein isolate. Food Research International, 52(1), 26–32.
prove the thermal stability of emulsion due to a stronger electrostatic Cui, C., Zhao, M., Yuan, B., Zhang, Y., & Ren, J. (2013). Effect of pH and pepsin limited
and steric repulsion between the droplets (Xu et al., 2017). To complete hydrolysis on the structure and functional properties of soybean protein hydrolysates.

298
Z. Hu, et al. Food Hydrocolloids 96 (2019) 288–299

Journal of Food Science, 78(12), C1871–C1877. chitosan microspheres for controlled release of resveratrol. Food Chemistry, 121(1),
Ellepola, S. W., Choi, S. M., & Ma, C. Y. (2005). Conformational study of globulin from 23–28.
rice (Oryza sativa) seeds by Fourier-transform infrared spectroscopy. International Sánchez-Reséndiz, A., Rodríguez-Barrientos, S., Rodríguez-Rodríguez, J., Barba-Dávila,
Journal of Biological Macromolecules, 37(1–2), 12–20. B., Serna-Saldívar, S. O., & Chuck-Hernández, C. (2018). Phosphoesterification of
Fabian, C., & Ju, Y.-H. (2011). A review on rice bran protein: Its properties and extraction soybean and peanut proteins with sodium trimetaphosphate (STMP): Changes in
methods. Critical Reviews in Food Science and Nutrition, 51(9), 816–827. structure to improve functionality for food applications. Food Chemistry, 260,
Feeney, R. E. (1987). Chemical modification of proteins: comments and perspectives. 299–305.
International Journal of Peptide and Protein Research, 29(2), 145–161. Sharif, M. K., Butt, M. S., Anjum, F. M., & Khan, S. H. (2014). Rice bran: A novel func-
Goto, Y., Calciano, L. J., & Fink, A. L. (1990). Acid-induced folding of proteins. tional ingredient. Critical Reviews in Food Science and Nutrition, 54(6), 807–816.
Proceedings of the National Academy of Sciences, 87(2), 573. Sung, H. Y., Chen, H. J., Liu, T. Y., & Su, J. C. (1983). Improvement of the functionalities
Greenfield, N. J. (2006). Using circular dichroism spectra to estimate protein secondary of soy protein isolate through chemical phosphorylation. Journal of Food Science,
structure. Nature Protocols, 1(6), 2876–2890. 48(3), 716–721.
Gul, K., Yousuf, B., Singh, A. K., Singh, P., & Wani, A. A. (2015). Rice bran: Nutritional Thaiphanit, S., & Anprung, P. (2016). Physicochemical and emulsion properties of edible
values and its emerging potential for development of functional food—a review. protein concentrate from coconut ( Cocos nucifera L.) processing by-products and the
Bioactive Carbohydrates and Dietary Fibre, 6(1), 24–30. influence of heat treatment. Food Hydrocolloids, 52(4), 756–765.
Hamada, J. S. (1997). Characterization of protein fractions of rice bran to devise effective Wang, T., Zhang, H., Wang, L., Wang, R., & Chen, Z. (2015). Mechanistic insights into
methods of protein solubilization. Cereal Chemistry, 74(5), 662–668. solubilization of rice protein isolates by freeze–milling combined with alkali pre-
Han, S.-W., Chee, K.-M., & Cho, S.-J. (2015). Nutritional quality of rice bran protein in treatment. Food Chemistry, 178, 82–88.
comparison to animal and vegetable protein. Food Chemistry, 172, 766–769. Wan, Z.-L., Wang, J.-M., Wang, L.-Y., Yuan, Y., & Yang, X.-Q. (2014). Complexation of
Hou, F., Ding, W., Qu, W., Oladejo, A. O., Xiong, F., Zhang, W., et al. (2017). Alkali resveratrol with soy protein and its improvement on oxidative stability of corn oil/
solution extraction of rice residue protein isolates: Influence of alkali concentration water emulsions. Food Chemistry, 161, 324–331.
on protein functional, structural properties and lysinoalanine formation. Food Wu, L., Zhao, W., Yang, R. J., & Chen, X. C. (2014). Effects of pulsed electric fields
Chemistry, 218, 207–215. processing on stability of egg white proteins. Journal of Food Engineering, 139, 13–18.
Hu, Z., Cheng, Y., Suzuki, N., Guo, X., Xiong, H., & Ogra, Y. (2018). Speciation of sele- Xiong, Z., & Ma, M. (2017). Enhanced ovalbumin stability at oil-water interface by
nium in Brown rice fertilized with selenite and effects of selenium fertilization on rice phosphorylation and identification of phosphorylation site using MALDI-TOF mass
proteins. International Journal of Molecular Sciences, 19(11). spectrometry. Colloids and Surfaces B: Biointerfaces, 153, 253–262.
Hunter, T. (2012). Why nature chose phosphate to modify proteins. Philosophical Xiong, Z. Y., Zhang, M. J., & Ma, M. H. (2016). Emulsifying properties of ovalbumin:
Transactions of the Royal Society B: Biological Sciences, 367(1602), 2513–2516. Improvement and mechanism by phosphorylation in the presence of sodium tripo-
Iosin, M., Toderas, F., Baldeck, P. L., & Astilean, S. (2009). Study of protein–gold nano- lyphosphate. Food Hydrocolloids, 60, 29–37.
particle conjugates by fluorescence and surface-enhanced Raman scattering. Journal Xu, X., Liu, W., Liu, C., Luo, L., Chen, J., Luo, S., et al. (2016). Effect of limited enzymatic
of Molecular Structure, 924, 196–200. hydrolysis on structure and emulsifying properties of rice glutelin. Food Hydrocolloids,
Karaca, A. C., Low, N., & Nickerson, M. (2011). Emulsifying properties of chickpea, faba 61, 251–260.
bean, lentil and pea proteins produced by isoelectric precipitation and salt extraction. Xu, X., Luo, L., Liu, C., & McClements, D. J. (2017). Utilization of anionic polysaccharides
Food Research International, 44(9), 2742–2750. to improve the stability of rice glutelin emulsions: Impact of polysaccharide type, pH,
Kato, Y., Aoki, T., Kato, N., Nakamura, R., & Matsuda, T. (1995). Modification of oval- salt, and temperature. Food Hydrocolloids, 64, 112–122.
bumin with glucose 6-phosphate by amino-carbonyl reaction. Improvement of pro- Yerramilli, M., Longmore, N., & Ghosh, S. (2017). Improved stabilization of nanoemul-
tein heat stability and emulsifying activity. Journal of Agricultural and Food Chemistry, sions by partial replacement of sodium caseinate with pea protein isolate. Food
43(2), 301–305. Hydrocolloids, 64, 99–111.
Khan, N. M., Mu, T. H., Zhang, M., & Arogundade, L. A. (2014). The effects of pH and high Yong, Y. H., Yamaguchi, S., & Matsumura, Y. (2006). Effects of enzymatic deamidation by
hydrostatic pressure on the physicochemical properties of a sweet potato protein protein-glutaminase on structure and functional properties of wheat gluten. Journal
emulsion. Food Hydrocolloids, 35, 209–216. of Agricultural and Food Chemistry, 54(16), 6034–6040.
Li, P., Sun, Z., Ma, M., Jin, Y., & Sheng, L. (2018). Effect of microwave-assisted phos- Yu, L., Yang, W. Q., Sun, J., Zhang, C. S., Bi, J., & Yang, Q. L. (2015). Preparation,
phorylation modification on the structural and foaming properties of egg white characterisation and physicochemical properties of the phosphate modified peanut
powder. Lebensmittel-Wissenschaft & Technologie, 97, 151–156. protein obtained from Arachin Conarachin L. Food Chemistry, 170, 169–179.
Lowry, O. H., Rosebrough, N. J., Farr, A. L., & Randall, R. J. (1951). Protein measurement Zhang, J.-B., Wu, N.-N., Yang, X.-Q., He, X.-T., & Wang, L.-J. (2012). Improvement of
with the Folin phenol reagent. Journal of Biological Chemistry, 193(1), 265–275. emulsifying properties of Maillard reaction products from beta-conglycinin and
Ma, Y. Q., Na, Z. G., Cheng, W. H., & Wang, X. (2017). Study on phosphorylation of rice dextran using controlled enzymatic hydrolysis. Food Hydrocolloids, 28(2), 301–312.
bran glutelin. Journal of Biobased Materials and Bioenergy, 11(4), 313–320. Zhao, Q., Xiong, H., Selomulya, C., Chen, X. D., Zhong, H., Wang, S., et al. (2012).
McClements, D. J. (2015). Food emulsions: Principles, practice, and techniques (3rd ed.). Enzymatic hydrolysis of rice dreg protein: Effects of enzyme type on the functional
Bocan Raton, FL: CRC Press. properties and antioxidant activities of recovered proteins. Food Chemistry, 134(3),
Moure, A., Sineiro, J., Domínguez, H., & Parajó, J. C. (2006). Functionality of oilseed 1360–1367.
protein products: A review. Food Research International, 39(9), 945–963. Zhao, X., Zhao, Q., Chen, H., & Xiong, H. (2019). Distribution and effects of natural
Patel, M. T., & Kilara, A. (1990). Studies on whey protein concentrates. 2. Foaming and selenium in soybean proteins and its protective role in soybean β-conglycinin (7S
emulsifying properties and their relationships with physicochemical Properties1. globulins) under AAPH-induced oxidative stress. Food Chemistry, 272, 201–209.
Journal of Dairy Science, 73(10), 2731–2740. Zullaikah, S., Melwita, E., & Ju, Y.-H. (2009). Isolation of oryzanol from crude rice bran
Peng, H., Xiong, H., Li, J., Xie, M., Liu, Y., Bai, C., et al. (2010). Vanillin cross-linked oil. Bioresource Technology, 100(1), 299–302.

299

You might also like