You are on page 1of 10

Food Hydrocolloids xxx (2018) xxx-xxx

Contents lists available at ScienceDirect

Food Hydrocolloids

F
journal homepage: www.elsevier.com

OO
Pickering emulsions stabilized with native and lauroylated amaranth starch
Everth J. Leal-Castañedaa, Yunia García-Tejedaa, Humberto Hernández-Sáncheza, Liliana Alamilla-Beltrána,
Darío I. Tellez-Medinaa, Georgina Calderón-Domíngueza, Hugo S. García-Galindob, Gustavo F. Gutiérrez-Lópeza, ∗
a
Graduados e Investigación en Alimentos, Escuela Nacional de Ciencias Biológicas, Instituto Politécnico Nacional, Prolongación del Carpio y Plan de Ayala s/n, Col. Santo

PR
Tomás, 11340 Ciudad de México, Mexico
b
Unidad de Investigación y Desarrollo de Alimentos, Instituto Tecnológico de Veracruz, M.A. de Quevedo 2779, Col. Formando Hogar, Veracruz, Ver. 91897, Mexico

ARTICLE INFO ABSTRACT

Article history: Pickering emulsions were prepared by using native and modified (esterification with lauroyl chloride) ama-
Received 29 August 2017 ranth starches as stabilizing particles of an oil phase formed with a mixture of canola oil and α-tocopherol

ED
Received in revised form 16 December 2017 (1% v/v). The effect of different concentrations of native (NS) and modified starch (MS) ranging from 2 to
Accepted 30 January 2018
30 %wt was evaluated in the emulsions regarding average droplet size, emulsion stability, zeta potential (ζ)
Available online xxx
and emulsification index. Optical microscopy showed that both NS and MS were adsorbed at the oil-water
interface forming a barrier that delayed phase separation. The MS gave place to a smaller emulsion droplet
Keywords: size than NS, and no phase separation was detected for the emulsions prepared by using 20 and 30 % wt of
Pickering emulsions NS or MS, producing stable emulsions.
Amaranth starch © 2017.
CT
Lauroylated amaranth starch
Emulsion stability
Lauroylation
Modified starches

1. Introduction bility of the particles which allows their accumulation at the oil-water
interface (Marto, Ascenso, Simoes, Almeida, & Ribeiro, 2016). Edible
RE

Solid-stabilized emulsions, also known as Pickering emulsions particles replacing synthetic surfactants include: colloidal celluloses
(Pickering, 1907), have been recently applied in food related fields and various cellulose derivatives (Duffus, Norton, Smith, Norton, &
given their potential applications (Rayner et al., 2014). Many food Spyropoulos, 2016), proteins (Gao, Wang, Yue, Xiong, Wu, Qiao, &
products are formed of immiscible phases, so that an emulsifier makes Liao, 2017), solid lipid nanoparticles and fat crystals (Weiss et al.,
it possible to form a homogeneous mixture, i.e., an emulsion which 2008), chitin nanoparticles (Tzoumaki, Moschakis, Kiosseoglou, &
may be a suitable option for the delivering of a number of com- Biliaderis, 2011), starch granules from different sources (Li, Sun, &
R

pounds, including vitamins. In particular, α -tocopherol is the most Yang, 2013; Marefati, Bertrand, Sjöö, Dejmek, & Rayner, 2017), and
active form of vitamin E which has a potent antioxidant capacity starch derivatives (Haaj, Thielemans, Magnin, & Boufi, 2014).
(Wysota, Michael, Hiew, Dawson, & Rajabally, 2017) and the prepa- As a potential biopolymer for Pickering emulsions, starch has re-
CO

ration of emulsions containing this vitamin has been reported using ceived increased attention and different starch sources have been eval-
different polymers as wall materials (Granillo et al., 2017). On the uated for their capacity to prepare Pickering emulsions, including:
other hand, a recommended dose of α-tocopherol of 22.4 IU/day has waxy maize, wheat, potato and rice. Li, Li, Sun, and Yang (2013) re-
been reported (Wysota et al., 2017) equivalent to 15 mg/day for adults. ported that starches having small granules (i.e., rice starch) better sta-
This amount may be the basis to prepare Pickering emulsions con- bilize Pickering emulsions. Moreover, native starches enhance their
taining this vitamin, using a suitable vehicle such as vegetable oil in performance in Pickering emulsions by means of esterification such as
which it is soluble. Pickering emulsions are stabilized by edible solid the hydrophobization with octenyl succinic anhydride (n-OSA) which
UN

particles instead of the commonly used surfactants. Stabilization of confers the starch molecule with amphiphilic properties (Bormann,
emulsion droplets by solid particles is attributed to the dual wetta Pierucci, Leite, & Miguez da Rocha, 2013).
The amphiphilic nature of starches obtained by means of hy-
drophobic modifications makes them attractive in a wide and in-
teresting spectrum of applications. For instance, as rheology modi-
Abbreviations: NS, native starch; MS, modified starch; NSE, native starch emul-
fiers, emulsion stabilizers, surface modifiers and drug delivery ve-
sions; MSE, modified starch emulsions; PDI, polydispersity index; ζ, zeta-poten- hicles among others (Namazi & Dadkhah, 2010). Timgren, Rayner,
tial; Global TSI, global Turbiscan stability index; EI, stability index; DS, degree of Dejmek, Marku, and Sjöö (2012) studied n-OSA-starches from dif-
substitution ferent sources as potential stabilizers of Pickering emulsions; they

Corresponding author.
Email address: gusfgl@gmail.com (G.F. Gutiérrez-López)

https://doi.org/10.1016/j.foodhyd.2018.01.043
0268-005/© 2017.
2 Food Hydrocolloids xxx (2018) xxx-xxx

observed that the size of the granule had a big influence on the sta- 2.2. Isolation of starch
bilizing capacity, and in particular, the smallest starch granules from
quinoa had the best emulsifying properties. The chemical modifica- Starch was isolated from amaranth seeds, based on the method
tion of starch by lauroyl chloride produces amphiphilic macromole- reported by Rayner, Timgren, Sjöö, and Dejmek (2012), for quinoa
cules consisting mainly of a hydrophilic backbone and hydrophobic starch isolation. Amaranth seeds were immersed in distilled water dur-
side chains (Namazi, Fathi, & Dadkhah, 2011). These hydrophobic ing 24 h at 4 °C, then the seeds were blended with distilled water (wa-

F
side chains have a greater affinity for the dispersed phase than the ter:seeds ratio 1:1). The blended seeds pulp was sieved through a
native starch, giving place to high-energy interactions at the oil-par- cloth, and the obtained paste was then washed with distilled water for

OO
ticle interface that might be a good alternative for stabilizing emul- at least three times until the washing water was clean. The filtrate,
sions (Rayner et al., 2014). Recently, Marefati et al. (2017) reported which contained the starch suspension, was centrifuged at 3000 × g
the preparation of Pickering emulsions with native and modified ama- during 10 min and the supernatant was removed. Proteins were elim-
ranth starch with different degrees of substitution (DS) of OSA. These inated by washing the starch twice with an alkaline solution (NaOH,
authors demonstrated that DS and the type of granule influenced the 0.3%). Finally, the starch was dried at 40 °C for 24 h in a laboratory
size of the droplet. A number of works suggest that it is possible to oven (SMO1E, Shel-Lab, USA).
create a Pickering emulsion with the n-OSA-starch (Agama & Bello,
2017; Marefati et al., 2017; Saari, Heravifar, Rayner, Wahlgren, &

PR
2.3. Lauroylation reaction
Sjöö, 2016), but the potential use of starches esterified with fatty acid
chlorides has not yet been reported. The use of fatty acid chlorides to
Starch esterification was carried out by following the method of
modify different types of starch has a number of advantages in com-
Namazi et al. (2011). A 50 g sample of starch was dispersed in 500 mL
parison with other classic methods of modification, such as the use of
of NaOH 0.1 M at 25 °C during 10 min; then, lauroyl chloride (1.5 mL/
water instead of organic solvents as the media for the reaction. Re-
g of starch) was added drop by drop to the starch dispersion under stir-
action time is relatively short and the hydrophobic modified products
ring conditions by using a magnetic stirrer at 300 rpm at room tempera-
can precipitate in water and may be separated without the addition of
ture. After the esterification process, the slurry was centrifuged during

ED
solvents (Namazi et al., 2011); besides, the byproducts of the reaction
10 min at 3000 × g and the lauroylated starch was recovered as the sed-
are non-toxic (water and NaCl). Lauroyl chloride, is a medium-size
iment from the tubes. After centrifugation, the lauroylated starch was
chain fatty acid that has proven to be a good candidate for chemical
washed three times with methanol and once with distilled water, then,
modifications of starches, and that allows achieving high degrees of
it was centrifuged and dried in a convection oven (SMO1E, Shel-Lab,
substitution with high yields as compared to the rest of the short and
USA) at 40 °C during 24 h. The dried starch was milled in a Hamil-
large chain fatty acid chlorides (Namazi et al., 2011; Vanmarcke et
ton Beach mill model 80,350 and sieved thorough a mesh having a
al., 2017). Moreover, the Food and Drug Administration (FDA) does
0.149 mm of opening size.
CT
not have published restrictions for lauric acid chloride regarding the
maximum amount that can be used in different food applications. Lau-
royl chloride has been used for the modification of alginate to which 2.4. Qualitative determination of the modification of starch
antimicrobial characteristics have been attributed. The product is cur-
rently patented as Na-lauroyl arginate ethylester, LAE®, which is suit- Qualitative determination of the modification of starch was car-
able for human consumption and was regarded as GRAS by the Food ried out by means of Fourier transformation infra-red (FTIR) spec-
RE

and Drug Administration (FDA) and by the European Food Safety troscopy aiming to demonstrate that the modification of the starch
Authority (EFSA) in 2007 (Coronel-León et al., 2016). In this study, had been carried out. The equipment used was a FTIR module IR2
amaranth starch was selected as stabilizing material of the emulsions equipped with an Indium Gallium Arsenide (InGaAs) detector, cou-
because it forms one of the smallest starch granules found in nature, pled to a Jobin-Yvon LabRam HR800 spectrometer (Horiba, Kioto,
typically 1–3 μm in diameter (Middlewood & Carson, 2012; Xia et al., Japan). The starch samples were placed in a holder and analyzed over
2015), which suggests that native and modified amaranth starch (es- a wave number range between 4000 and 450 m−1 with a spectral res-
olution of 4 cm−1 and performing 36 scans per measurement, using an
R

terified with lauroyl chloride) may be good options for the preparation
of Pickering emulsions. ATR contact objective.
The aim of this work was to prepare and characterize oil in water
(O/W) Pickering emulsions using native and modified (lauroylation)
CO

2.5. Degree of substitution (DS)


amaranth starches to stabilize an oil phase formed by α-tocopherol and
canola oil as the vehicle to supply this vitamin.
The degree of substitution (DS) was determined according to the
method of Jeon, City, Viswanathan, Lowell, and Gross (1999). A 1 g
2. Materials and methods
sample from each starch (NS and MS) was dissolved in 10 mL of di-
methyl sulfoxide during 10 min at 70 °C, left to cool at ambient tem-
2.1. Materials
perature (25 °C) and five drops of phenolphthalein were added. Then,
UN

the obtained solution was titrated with a standard solution of 0.05 M of


Seeds of amaranth (Amaranthus hypochondriacus) were obtained
sodium hydroxide until a pale pink color was obtained. DS was calcu-
in the Xochimilco, Mexico City market, α-tocopherol, lauroyl chlo-
lated with the equation reported by Cao, Song, Deng, and Ragauskas
ride, 1, 2-propanediol (99%), Nile Blue and Nile Red, were acquired
(2011):
from Sigma-Aldrich (Mexico City). Sodium hydroxide 98% and
methanol were acquired from Luzeren Laboratories (Mexico City).
The canola oil was acquired from Mazola (México City). Ultra-pure
water with a resistivity of 18.25 MΩ⋅cm was used throughout the ex-
periment.
(1)
Food Hydrocolloids xxx (2018) xxx-xxx 3

Where A is the titrated volume of the standard solution of NaOH in used to determine the absorbance at 595 nm in a Thermo Scientific
mL; M is the molarity of the NaOH solution; W is the dry weight of spectrophotometer Nanodrop 2000, (Wilmington, USA).
the esterified starch.
2.9. Preparation of the pickering emulsions
2.6. Particle size measurements of starch granules

F
A high amount of water (70%) was used in respect to the added oil
Particle size distributions (PS) and polydispersity index (PDI) of phase (30 %v) which was prepared as a mixture of canola oil and α-to-
native and modified starch granules were measured by placing 1 mL copherol (99:1). Emulsification of canola oil and α-tocopherol (99:1)

OO
of a water suspension of particles at a concentration of 0.01 %wt, in a in water with the aid of starch granules were formulated as shown in
Malvern Zetasizer Nano ZS (Malvern, USA). Table 1. Starch dispersions were prepared in glass test tubes by using
native or lauroylated amaranth starches at concentrations of 2, 6, 10,
2.7. Scanning electron microscopy (SEM) 20 and 30 % wt, and each one containing 7 mL of ultra-pure water and
3 mL of oil. The mixtures were mixed at 11,000 rpm for 1 min by using
The morphology of starch granules was observed in a Field Emis- a homogenizer IKA T18 Basic Ultra-Turrax at 25 °C.
sion Scanning Electron Microscope JSM-7800F (JEOL, JAPAN), us-
ing an acceleration voltage of 5 kV. The starch samples were fixed in

PR
2.10. Droplet size by optical microscopy
stubs containing a double-faced adhesive metallic tape and they were
cover with gold before fixing them to the microscope stage. Optical microscopy of the Pickering emulsions was performed in
an Eclipse H550S (Nikon, Chiyoda-ku, Japan) optical microscope
2.8. Chemical characterization of starch equipped with a Kodak DC 120 digital camera (Sevier County, Ten-
nessee, USA). The size of the emulsion droplets was obtained by plac-
Starch purity was measured as the total starch content based on the ing a drop of the emulsion between a channeled and a cover slides and
heat-stable α-amylase and amyloglucosidase method, for which the to- capturing images of the emulsions as reported by Li et al. (2013), con-

ED
tal starch measurement kit from Megazyme International Ireland Ltd sidering 200 droplets for each emulsion. Image processing was car-
(Bray Business Park, Ireland) was employed. Procedure A was fol- ried out by using the ImageJ v 1.49 software (Rasband, 1997–2015).
lowed for samples not containing resistant starch. The surface to volume diameter (d32d) and the volume mean diameter
The amylose content (% dry basis) of the starch was determined by (d43d) of the droplets were calculated by applying Eqs. (2) and (3) (Li
the method reported by Williams, Kuzina, and Hlynka (1970), which et al., 2013). Also, the polydispersity index (PDI) was obtained by us-
is based on the reaction of iodine with amylopectin in which a char- ing Eq. (4).
acteristic color is obtained. The concentration of amylose was ob-
CT
tained by interpolating in a standard curve generated by using different
concentrations of amylose standard (Sigma-Aldrich). 5 mL of a 0.5 N
KOH solution were added to 20 mg of starch (20 mg). Subsequently, (2)
this solution was diluted with 100 mL of distilled water and 10 mL of
this solution were transferred to a 50 mL volumetric flask to which
5 mL of 0.5 N HCl and 0.5 mL of iodine solution were added and total
RE

volume completed by adding distilled water. The absorbance was read


at 625 nm after 5 min. In parallel, a blank was run without the starch (3)
and used as a reference. Amylopectin content was calculated by sub-
tracting amylose from the total starch content.
Total protein content of native and modified starch was deter-
mined by Bradford (Thermo Scientific). Starch Solutions were pre-
(4)
R

pared at a concentration of 0.2% (w/v), then a 10 μL aliquot was


added with 10 μL of the Bradford reagent. The mixture was cen-
trifuged at 3000 rpm for 5 min, and after 1 min, a 2 μL sample was In which di is the diameter of a droplet, N is the total number of
droplets.
CO

Table 1
Size parameters of droplets (d32, d43), Polydispersity Index, ζ, Global Turbiscan Stability Index and Emulsification index of emulsions at different concentrations of starch.

Concentration of starch in the emulsion Area (μm2) Perimeter (μm) d32 (μm) d43 (μm) PDI ζ (mV) Global TSI EI
a a a a
NS-2%wt 1041.0 ± 213.4 124.9 ± 13.6 40.0 40.5 1.04 −19.2 ± 0.7 16.79 ± 1.4 0.8 ± 0.2a
NS-6%wt 560.3 ± 105.6b 101.3 ± 11.2b 28.7 28.9 1.02 −20.7 ± 0.5a 11.02 ± 2.8b 0.9 ± 0.2a
UN

NS-10%wt 147.5 ± 29.5c 49.6 ± 5.1b 15.3 15.4 1.02 −21.2 ± 0.7a 3.99 ± 0.64c 1.0 ± 0.1a
NS-20%wt 110.8 ± 22.7c 40.6 ± 9.2b 10.1 10.2 1.05 −22.3 ± 0.9a 1.39 ± 0.13d 1.0 ± 0.1a
NS-30%wt 57.0 ± 14.4d 33.6 ± 5.0b 9.90 10.0 1.04 −24.5 ± 0.1a 2.40 ± 0.35c 1.0 ± 0.0a
MS-2%wt 53.3 ± 13.8d 28.9 ± 4.7b 8.90 9.10 1.05 −12.2 ± 0.5b 11.26 ± 0.8b 0.7±.1a
MS-6%wt 45.4 ± 12.3d 26.2 ± 3.9b 8.30 8.50 1.08 −11.0 ± 0.7b 10.360 ± 1.0b 0.8 ± 0.1a
MS-10%wt 39.3 ± 10.3d 24.0 ± 3.2b 7.70 7.80 1.05 −10.1 ± 0.3b 4.63 ± 0.7c 0.9 ± 0.1a
MS-20%wt 37.8 ± 7.7d 24.1 ± 2.7b 7.50 7.70 1.03 −9.1 ± 0.2b 3.14 ± 0.7c 1.0 ± 0.1a
MS-30%wt 27.5 ± 5.3e 24.4 ± 4.1b 6.80 6.90 1.06 −4.9 ± 0.3c 3.73 ± 0.5c 1.0 ± 0.0a
Each value in the last three columns represents the mean ± standard deviation of three replicates. Different lowercase letters within the same row denote significant differences (≤0.05)
different starch concentration and different type to starch.
4 Food Hydrocolloids xxx (2018) xxx-xxx

2.11. Confocal laser scanning microscopy (CLSM)


(6)
A Confocal Laser Scanning Microscope (CLSM) (LSM 710 NL0
Carl Zeiss, Germany) was used to visualize the structure of the Picker-
Where Vemuls is the volume occupied by the emulsion, and Vtotal is the
ing emulsion droplets. Nile Blue dye was used to stain the native and
total volume of all phases (oil phase, added starch, and water).
lauroylated amaranth starches (Yusoff & Murray, 2011) in an aque-

F
ous solution and Nile Red dye was used to stain the oil phase (Yusoff
2.14. Statistical analysis
& Murray, 2011). Pickering emulsions were then prepared by mixing

OO
stained Nile Blue and Nile Red amaranth and oil samples respectively,
Data were reported as mean values of three replicates for each ana-
as described in Section 2.9. Then, the emulsions were immediately
lytical determination. One-way ANOVA and Tukey's significant mul-
placed between two cover glasses for observation in the CLSM.
tiple comparison test were used to determine statistical differences be-
The CLSM was operated in fluorescence mode with an Apochro-
tween treatments (p ≤ 0.05).
mat- Plan objective (63×/1.4 Oil DIC M27) at 60×/1.4. The fluores-
cent dyes were excited by using an argon laser at 488 nm for Nile
3. Results and discussion
Red or a helium neon (He-Ne) laser at 633 nm for Nile Blue. The pro-
jected area of the droplets was evaluated using the software ImageJ

PR
3.1. Characterization of starches
v 1.49 software (Rasband, 1997–2015), applying the method reported
by Garcia-Armenta et al. (2016).
3.1.1. Chemical characterization of starch
The obtained amylose contents of NS and MS were 22.3 ± 0.3 and
2.12. ζ of starch particles 21.2 ± 0.1 as %dry basis, respectively. Recently, Marefati et al. (2017)
reported 20.95% of amylose content for amaranth hypocondriacus
To determine ζ to the NS and MS starch particles, they were dis- starch, native from Mexico. Esterification did not generate significant
persed (0.01 wt%) in water and measured by using a Zetasizer ZS differences in the amylose content. Similar results were reported by

ED
(Malvern, USA) at 25 °C. He, Song, Ruan, and Chen (2006) for different varieties of rice starch
modified with n-OSA. The authors obtained protein contents of 4.9
2.13. Stability kinetics of the emulsion and emulsifying capacities of and 3.2 as % dry basis for NS and MS, respectively. Similar results
the starches were also reported by Marefati et al. (2017) for amaranth starch mod-
ified with n-OSA at different DS for which the protein content con-
Stability of the emulsions with time was analyzed by light scat- comitantly decreased with esterification: 0.112 to 0.032 for native and
CT
tering in a Turbiscan Lab Expert (Formulation, Toulouse, France). modified starch, respectively. Reported esterification with fatty acid
The detection head is composed of a pulsed near-infrared light source chlorides and n-OSA were carried out in alkaline media which gener-
(λ = 880 nm) and two synchronous detectors. Stability of the emulsions ated a release of the protein during chemical modification (Rayner et
with time was analyzed by light scattering in a Turbiscan Lab Expert al., 2012). On the other hand, total starch contents were 86% ± 0.9 and
(Formulation, Toulouse, France) by evaluating the Turbiscan Stabil- 90% ± 1.1 for NS and MS, respectively.
ity Index (TSI). The detection head of the equipment is composed of
3.1.2. Analysis of lauroylated starch by FTIR
RE

a pulsed near-infrared light source (λ = 880 nm) and two synchronous


detectors. The calculation of this index was carried out with equation The FTIR spectra of amaranth starch and lauroylated starch are
(5): shown in Fig. 1. The native amaranth starch presented its characteris
R

(5)
CO

Where xi is the average backscatter for each minute of the measure-


ment, xbs is the mean value of xi and n is the number of scans. TSI
considers all the variations detected in the samples in terms of size
and/or concentration with time so that it allows to infer on the stability
of the sample and compare among different systems. The higher the
TSI value, the less stable the emulsion is (Kang et al., 2011). A 25 mL
sample of the emulsion was poured into a glass cell and placed in the
UN

equipment to obtain the stability kinetics during 6 h and the Global


Turbiscan Stability Index (Global TSI) of the emulsion at the end of
the experiment is then evaluated as the average value of the individu-
als TSI: at the bottom, in the middle and in the top sections of the cell.
The emulsifying capacities of the starches were determined one
day after their preparation as the emulsion index (EI) (Saari et al.,
2016) as follows:
Fig. 1. The FT-IR spectra of native (a) and lauroylated (b) amaranth starches. Numbers
in the spectra are the wavenumbers in the corresponding axis.
Food Hydrocolloids xxx (2018) xxx-xxx 5

tic fingerprint region (Fang, Fowler, Sayers, & Williams, 2004), ad- face of the granule which could reduce its size (as reported by us).
ditionally, the absorption band for NS at 1544 cm−1 was associated In our work, protein content decreased by 35% (4.9 and 3.2 for NS
to a residual protein amide II. These absorption band has been re- and MS, respectively). The damage generated by the chemical modi-
ported for amaranth protein (Aceituno-Medina, Mendoza, Lagaron, & fication depends on many factors, such as the source of the starch, the
López-Rubio, 2015). After esterification, lauroylated amaranth starch type of modification made, the media of reaction, the extent of agita-
showed a characteristic band of the carbonyl group at 1624 cm−1 of tion, among others (Namazi et al., 2010, 2011). Chemical conditions

F
the fatty ester. In addition, the C-H stretching vibrations of the alkyl at modification may affect the amorphous region of starch granules by
groups of the fatty ester chain were clearly present at 2920 and exposing the -OH groups of the amylose which in turn as are exposed

OO
2833 cm−1. The decrease of hydroxyl vibrations in the range of to the alkaline media that damage the granule surface (Wang et al.,
3000–3600 cm−1 and at 1624 cm−1 suggested the esterification of hy- 2014). Extreme erosion of the amorphous region can practically dis-
droxyl groups of starch (Namazi et al., 2011). In this work, FTIR was appear such region so giving place to crystals as reported by Li et al.
used as a first step for determining if chemical modification of starch (2013) who reported that emulsions can be stabilized by starch crys-
had occurred. The analysis of FTIR has not been regarded as a reli- tals. The surface of the modified granule, unlike the native, showed a
able technique to quantitatively evaluate DS (Freire, Silvestre, Neto, smoother and damaged surface, while the native starch had a polygo-
& Rocha, 2005). nal shape and a rougher surface. These morphometric changes might
be attributed to the chemical modification although, to the best of our

PR
3.1.3. Determination of the degree of substitution of lauroylated knowledge, there is no evidence of the morphometry of starch gran-
starch ules modified with fatty acid chlorides. Song, He, Ruan, and Chen
Lauroylated amaranth starches have not been employed as Pick- (2006) reported surface damage of rice starch granules after treatment
ering emulsion stabilizers before. It has been reported that the incor- with OSA. In this case, native rice starch granules were polygonal
poration of hydrophobic groups to the amylose, results in surface ac- with well-defined edges, much like the lauroylated starch granules
tive properties, it is believed that functional groups are located mainly of amaranth. Starch OSA granules exhibited rough surfaces and their
at the surface of the granules (Whitney, Reuhs, Ovando, & Simsek, edges lost definition. Pores and cavities were generated on the granule

ED
2016.), which is useful for stabilizing emulsions. The properties of es- surface. Gao et al. (2014) reported no obvious differences in the shape
terified amaranth starch depend on the degree of substitution (DS). In of waxy corn starch granules after lipase esterification and only slight
this work, the DS obtained for lauroylated amaranth starch was 0.01. morphometric changes could be appreciated on the surface of esteri-
Variables such as the source of the starch, length of the fatty acid chain fied starch granules as compared to that of the native starch. Esterified
and type of chemical modification carried out have an influence on starch granules exhibited slightly rough surfaces with appearance of
DS. Fang, Fowler, Tomkinson, and Hill (2002) chemically modified apophysis and pores. Marefati et al. (2017), on the other hand, did not
four starches (corn, Hylon VII, Hylon, Amioca) and reported that the report significant changes in the diameter of amaranth starch granules
CT
esterification reaction was successful only with acid chlorides contain- modified with different OSA concentrations.
ing 6–10 carbon chains. Longer and shorter fatty acids chains were hy-
drolyzed due to the conditions of reaction. Namazi et al. (2013) carried 3.2. Characterization of emulsions
out the esterification reaction of potato and corn starch with longer
carbon chain fatty acid chlorides with some modifications with respect 3.2.1. Droplet size of pickering emulsions
to the work by Fang et al. (2002) such as a lower NaOH concentration, Two key factors influencing the emulsion droplet size during Pick-
RE

temperature and reaction time. These authors reported a lower DS for ering emulsion preparation process are: (i) the concentration of starch
the modification with lauryl chloride than for octenyl chloride (0.35 particles initially dispersed in the aqueous phase and (ii) the amount
and 0.48, respectively) and even lower for palmitoyl chloride (0.11) of oil phase (Frelichowska, Bolzinger, & Chevalier, 2010). The influ-
and corn starch. Gao et al. (2014) esterified corn starch with lauric acid ence of starch concentration in the water phase was assessed by mea-
using lipase as a catalyst and reported similar DS (0.0109 and 0.0151) surements of droplet size distributions over a wide range of starch con-
than those obtained in our work for two different concentrations of tent. Pickering stabilization does not require the complete coverage
R

lauric acid with an amylose content of 47.31 and 48.69%, respectively of closely packed particles in the interface (Dickinson, 2013). Emul-
and using the same technique reported in our work. Morphology and sions prepared at starch concentrations lower than 10% wt (Table 1)
size of native and modified starches. were unsuccessful. In this case, the surface coating by solid particles
was not large enough to form a compact layer in the oil-water inter-
CO

3.1.4. Morphology and size of native and modified starches face and coalescence occurred, which could be observed by optical
Granule size plays an important role in the formation of an emul- microscopy. On the other hand, efficient stabilization against coales-
sion, it has been reported that small granules with smooth surfaces, cence was achieved with 20 and 30% wt of starch for both NS and
have a better contact in the oil-water interface (Saari et al., 2016). MS. Pickering emulsion droplets prepared at the following starch con-
The shapes and size of starch granules are dependent on their source. centrations: NS-2%wt, NS-2% wt, NS-30% wt and WS-30% wt are
Scanning electron micrographs of native starch are shown in Fig. 2, shown in Fig. 3. The droplets are conformed by an oil core coated
in which it is possible to observe granules with a polyhedral morphol- by granules of starch. In the micrographs, it is difficult to observe the
UN

ogy; the size of the granules ranged from 0.6 to 1.7 μm and the av- granules of starch because of their small size. This observation has
erage particle size was 1.13 μm. After modification, the size of ob- also been reported by Rayner et al. (2012). At NS-2% wt, only a few
tained granules was 0.6–1.2 μm with an average size of 1.09 μm and starch granules on the surface of droplets can be seen (Fig. 3a). As
PDI found were larger for NS than MS (0.87 and 0.49, respectively. the solids concentration increased from 2% wt to 30 %wt, larger parti-
Jackson, Choto-Owen, Waniska, and Rooney (1988) reported that cles were observed in Fig. 3c, the droplet interface was more densely
sodium hydroxide could depolymerize starch, generating a smaller di- coated with particles at 30% wt than at 2%wt for both NSE and MSE.
ameter of the starch granule. Rayner et al. (2012) and Wang et al. Average size of individual droplets was related to the concentra-
(2014) found that NaOH gave place to proteins removal from the sur tion of the solid particles forming the emulsions shown in Fig. 3. As
6 Food Hydrocolloids xxx (2018) xxx-xxx

F
OO
PR
ED
CT
Fig. 2. Particle size distribution of native (NS) and lauroylated (MS) starches granules (top) and scanning electron microscopy images of a) native and b) modified amaranth starch
RE

granules (bottom). Scale bar = 1 μm.

the amaranth starch concentration increased, the droplet size of Pick- 3.2.2. Laser scanning confocal microscopy
ering emulsions decreased from 40.0 to 9.8 μm and from 8.9 to 6.8 μm In Fig. 4a and d, the CLSM images of NS and MS stabilized emul-
for NS and MS respectively. Authors have shown that the average size sions are shown; blue regions correspond to amaranth starch and red
of the droplet increases at short times after preparation of the emulsion areas are those of the oil phase. Upon separation of the blue and red
and quickly reaches a limit value that depends on the amount of solid channels, it was possible to observe that droplets of emulsion prepared
R

particles. It is likely that the studied emulsions showed limited coa- with NS (Fig. 4c) had a lower coverage of the oil phase in respect to
lescence (Arditty, Whitby, Binks, & Schmitt, 2003). When concentra- those prepared with MS (Fig. 4f). Oil phases in both cases are shown
tion of particles was low they did not adsorb on the oil surface in suf- in Fig. 4b and e for NS and MS emulsions respectively. All droplets
CO

ficient amount to stabilize the emulsion, larger droplets were formed in Fig. 4 have an average thickness which is dependent on the diam-
which had sufficient amount of particles to stabilize them as compared eter of the adsorbed particles. Additionally, according to Dickinson
with droplets stabilized with larger amount of particles in which lim- (2013) due to the increased hydrophobicity of the modified particles,
ited coalescence gave place to smaller droplets (see Table 1) (Binks they might tend to aggregate with each other in the aqueous phase so
& Olusanya, 2017; Leal-Calderon & Schmitt, 2008; Tzoumaki et al., that sometimes, aggregates adsorb on the oil surface rather than indi-
2011). When comparing the droplet size of the NS and MS, it was ob- vidual particles giving place to local larger thicknesses in the sites of
served that MS emulsions had a smaller droplet size than NS. Marefati adsorption.
UN

et al. (2017) reported an increment in the emulsifying capacity of mod-


ified amaranth starch with OSA, which generated an increase in the 3.2.3. Stability of emulsions determined by backscattering
thickness of the emulsion starch layer as the modification level in- Pickering emulsions usually present phase and reverse sedimen-
creased. Also, Saari et al. (2016) reported droplet diameters of 13.4 μm tation (Rayner et al., 2014). For a conventional emulsion, this type
and 4.7 μm, when using 200 and 1200 mg of n-OSA quinoa starch/mL of phenomena is considered as destabilizing, which significantly af-
of oil, respectively. fects the physical properties of the emulsion. However, for Pickering
emulsions, sedimentation, flocculation and Ostwald maturation are
not considered important factors that modify physical properties since
they are commonly found in these types of emulsions and in which
Food Hydrocolloids xxx (2018) xxx-xxx 7

F
OO
PR
ED
Fig. 3. Amaranth starch granule stabilized Pickering emulsions (optic microscopy image 40× magnification) at different concentrations a) 2 % wt of NS, b) 2 % wt of MS, c) 30 %
wt of NS and d) 30 % wt of MS. Scale bar = 10 μm.
CT
R RE
CO
UN

Fig. 4. CLSM micrographs of NSE prepared with 20%wt of starch (a, b and c) and of MSE prepared with 20 % wt of starch (d, e and f).
8 Food Hydrocolloids xxx (2018) xxx-xxx

coalescence has to be avoided (Rayner et al., 2014). The backscatter- (EI6h = 1.0) upwards in comparison with MSE at the same concentra-
ing profile of emulsions stabilized by NS and MS are shown in Fig. 5. tion (EI6h = 0.9). Particle size and its density influenced the EI (Rayner
The X-Axis represents the height of the tube, while Y-axis represents et al., 2012). Successful Pickering emulsions with different sizes and
the percent of change of BS relative to the initial stage. As shown in thickness of their starch cover were found. The smaller the size of the
Fig. 5, there are few changes in NSE-30% wt and NSE-10% wt, indi- particles, the larger the thickness of their starch covers which caused
cating that there is no particle migration in the emulsion. The MS-2% an eventual droplets flocculation. In this case, high concentrations of

F
wt and NS-2% wt emulsion samples, which contain a low concentra- starch generated droplets with different densities (Rayner et al., 2012),
tion of starch, showed the decrease of the size of the emulsified region thus generating a stable emulsion without separation of visible phases

OO
with time. Phase separation was larger for MSE and NSE at 2% wt and with an IE6h = 1 for the concentrations of 20 and 30%wt of native
starch concentration. and modified starches.
The Global TSI after 6 h of storage under different starch concen-
tration is shown in Table 1. The Global TSI of NSE decreased from
16.8 to 2.4 as starch concentration increased from 2 to 30 % wt indi- 4. Conclusions
cating increased stability and the Global TSI of MSE decreased from
11.26 to 3.73, as starch concentration increased from 2 to 30 % wt; Food grade Pickering emulsions stabilized by native and lauroy-
which indicated improved stability. The increase of solids content pro- lated amaranth starches were prepared. It was found that both, na-

PR
duced a lower Global TSI after storage, which was attributed to small tive and modified amaranth starches acted as good stabilizers of Pick-
sizes of particles that were associated to more stable emulsions. Global ering emulsions. The starch concentration had an influence on the
TSI values of 1.39 and 2.4, obtained for NS-20% wt and NS-30% wt droplet size of the emulsion, modified starch generated smaller droplet
emulsions respectively, indicated the emulsions with the highest sta- size and a greater emulsion stability attributed to the affinity of the
bility with time. particle with the oil phase. NSE-10% wt, NSE-20% wt, NSE-30%
Finally, to compare the effect of the lauroylation of amaranth wt, MSE-20% wt, MSE-30% wt were stable during 6 h. The stabil-
and starch concentration in the emulsions, the emulsification index ity studies presented in this work showed that the lauroylated ama-

ED
(EI) evaluated at 6 h after preparation is depicted in Table 1. Ele- ranth starch improved emulsifying capacity as compared with the na-
vated starch concentration favored the formation of a higher EI6h in tive amaranth starch because it produced smaller droplet sizes of the
both types of starch. However, NS produced a greater stability for emulsion, slower creaming and a better stabilization of the Pickering
the emulsified region (Fig. 5) as from 10% of starch concentration emulsion as compared with emulsions prepared with native starch.
CT
R RE
CO
UN

Fig. 5. Turbiscan Lab Expert backscattering light graphs at different concentrations of starch: 2, 10 and 30 %wt of NS; and 2, 10 and 30 %wt of MS; and images of the corresponding
emulsions.
Food Hydrocolloids xxx (2018) xxx-xxx 9

Acknowledgments of a colocalization of its components. Revista Mexicana de Ingeniería Química


1–7.
Haaj, S.B., Thielemans, W., Magnin, A., Boufi, S., 2014. Starch nanocrystal stabilized
Author Leal-Castañeda acknowledges the support from CONA- pickering emulsion polymerization for nanocomposites with improved perfor-
CYT-México. Thanks go to Center of Nanosciences and micro and mance. ACS Applied Materials & Interfaces 6 (11), 8263–8273.
nanotechnologies from IPN and to Ms. Ilse Monroy for help with He, G.Q., Song, X.Y., Ruan, H., Chen, F., 2006. Octenyl succinic anhydride modified
formatting references. Authors acknowledge the financial support of early Indica rice starches differing in amylose content. Journal of Agricultural and

F
Food Chemistry 54 (7), 2775–2779.
CONACYT-Mexico through the grants (CB-2014-01-242371) and Jackson, D.S., Choto-Owen, C., Waniska, R.D., Rooney, L.W., 1988. Characterization
SIP-20171502. of starch cooked in alkali by aqueous high-performance size-exclusion chromatog-

OO
raphy. Cereal Chemistry 65 (6), 493–496.
Appendix A. Supplementary data Jeon, Y., City, I., Viswanathan, A., Lowell, M., Gross, R., 1999. Studies of starch es-
terification: Reactions with alkenyl-succinates in aqueous slurry systems. Starch
Staerke 3, 90–93.
Supplementary data related to this article can be found at https:// Kang, W., Xu, B., Wang, Y., Li, Y., Shan, X., An, F., et al., 2011. Stability mechanism
doi.org/10.1016/j.foodhyd.2018.01.043. of W/O crude oil emulsion stabilized by polymer and surfactant. Colloids and Sur-
faces A: Physicochemical and Engineering Aspects 384 (1–3), 555–560.
Leal-Calderon, F., Schmitt, V., 2008. Solid-stabilized emulsions. Current Opinion in
Uncited references Colloid & Interface Science 13 (4), 217–227.

PR
Li, C., Li, Y., Sun, P., Yang, C., 2013. Pickering emulsions stabilized by native starch
French et al., 2016, Quintanilla et al., 2011, U.S. Food and Drug granules. Colloids and Surfaces A: Physicochemical and Engineering Aspects 431,
Administration, 2012. 142–149.
Marefati, A., Bertrand, M., Sjöö, M., Dejmek, P., Rayner, M., 2017. Storage and diges-
tion stability of encapsulated curcumin in emulsions based on starch granule Pick-
References ering stabilization. Food Hydrocolloids 63, 309–320.
Marto, J., Ascenso, A., Simoes, S., Almeida, A.J., Ribeiro, H.M., 2016. Pickering
Aceituno-Medina, M., Mendoza, S., Lagaron, J.M., López-Rubio, A., 2015. Photopro- emulsions: Challenges and opportunities in topical delivery. Expert Opinion on
tection of folic acid upon encapsulation in food-grade amaranth (Amaranthus Drug Delivery 13 (8), 1093–1107.
hypochondriacus L.) protein isolate – pullulan electrospun fibers. Lebensmit- Middlewood, P.G., Carson, J.K., 2012. Extraction of amaranth starch from an aqueous

ED
tel-Wissenschaft und -Technologie- Food Science and Technology 62 (2), medium using microfiltration: Membrane characterization. Journal of Membrane
970–975. Science 405–406, 284–290.
Agama, E., Bello, L., 2017. Starch an emulsions stability: The case of octenyl succinic Namazi, H., Dadkhah, A., 2010. Convenient method for preparation of hydrophobi-
anhydride (OSA) starch. Current Opinion in Food Science 13, 78–83. cally modified starch nanocrystals with using fatty acids. Carbohydrate Polymers
Arditty, S., Whitby, C.P., Binks, B.P., Schmitt, V., 2003. Some general features of lim- 79, 731–737.
ited coalescence in solid-stabilized. Physical Journal 281, 273–281. Namazi, H., Fathi, F., Dadkhah, A., 2011. Hydrophobically modified starch using
Binks, B.P., Olusanya, S.O., 2017. Pickering emulsions stabilized by coloured organic long-chain fatty acids for preparation of nanosized starch particles. Scientia Iranica
pigment particles. Chemical Science 8 (1), 708–723. 18 (3), 439–445.
CT
Bormann, D., Pierucci, A.P.T.R., Leite, S.G.F., Miguez da Rocha, M.E., 2013. Mi- Pickering, S.U., 1907. CXCVI.-Emulsions. Journal of the Chemical Society Transac-
croencapsulation of passion fruit (Passiflora) juice with n-octenylsuccinate-deriva- tions 91 (0), 2001–2021.
tised starch using spray-drying. Food and Bioproducts Processing 91 (1), 23–27. Quintanilla, M., Meraz, L., Alamilla, L., Chanona, J., Terres, E., Hernández, H., et al.,
Cao, S., Song, D., Deng, Y., Ragauskas, A., 2011. Preparation of starch fatty acid 2011. Morphometric characterization of spray-dried microcapsules before and af-
modified clay and its application in packaging papers. Industrial & Engineering ter a-tocopherol extraction. Revista Mexicana de Ingeniería Química 10 (2),
Chemistry Research 50 (9), 5628–5633. 301–312.
Coronel-León, J., López, A., Espuny, M.J., Beltran, M.T., Molinos-Gómez, A., Ro- Rasband, W.S., 1997. ImageJ. US National Institutes of Health, Bethesda, MD.
cabayera, X., et al., 2016. Assessment of antimicrobial activity of Nα -lauroyl Rayner, M., Marku, D., Eriksson, M., Sjöö, M., Dejmek, P., Wahlgren, M., 2014. Bio-
RE

arginate ethylester (LAE®) against Yersinia enterocolitica and Lactobacillus plan- mass-based particles for the formulation of Pickering type emulsions in food and
tarum by flow cytometry and transmission electron microscopy. Food Control 63, topical applications. Colloids and Surfaces A: Physicochemical and Engineering
1–10. Aspects 458, 48–62.
Dickinson, E., 2013. Stabilising emulsion-based colloidal structures with mixed food Rayner, M., Timgren, A., Sjöö, M., Dejmek, P., 2012. Quinoa starch granules: A can-
ingredients. Journal of the Science of Food and Agriculture 93 (4), 710–721. didate for stabilising food-grade pickering emulsions. Journal of the Science of
Duffus, L.J., Norton, J.E., Smith, P., Norton, I.T., Spyropoulos, F., 2016. A compara- Food and Agriculture 92 (9), 1841–1847.
tive study on the capacity of a range of food-grade particles to form stable O/W Saari, H., Heravifar, K., Rayner, M., Wahlgren, M., Sjöö, M., 2016. Preparation and
R

and W/O Pickering emulsions. Journal of Colloid and Interface Science 473, 9–21. characterization of starch particles for use in pickering emulsions. Cereal Chem-
Fang, J.M., Fowler, P.A., Sayers, C., Williams, P.A., 2004. The chemical modification istry Journal 93 (2), 116–124.
of a range of starches under aqueous reaction conditions. Carbohydrate Polymers Song, X., He, G., Ruan, H., Chen, Q., 2006. Preparation and properties of octenyl suc-
55 (3), 283–289. cinic anhydride modified early Indica rice starch. Starch Staerke 58 (2), 109–117.
CO

Fang, J.M., Fowler, P.A., Tomkinson, J., Hill, C.A.S., 2002. The preparation and char- Timgren, A., Rayner, M., Dejmek, P., Marku, D., Sjöö, M., 2012. Emulsion stabilizing
acterisation of a series of chemically modified potato starches. Carbohydrate Poly- capacity of intact starch granules modified by heat treatment or octenyl succinic
mers 47, 245–252. anhydride. Food Sciences and Nutrition 1 (2), 157–171.
Freire, C.S.R., Silvestre, A.J.D., Neto, C.P., Rocha, R.M.A., 2005. An efficient method Tzoumaki, M.V., Moschakis, T., Kiosseoglou, V., Biliaderis, C.G., 2011. Oil-in-water
for determination of the degree of substitution of cellulose esters of long chain emulsions stabilized by chitin nanocrystal particles. Food Hydrocolloids 25 (6),
aliphatic acids. Cellulose 12 (5), 449–458. 1521–1529.
Frelichowska, J., Bolzinger, M., Chevalier, Y., 2010. Effects of solid particle content U.S. Food and Drug Administration, 2012. Title 21 Code of federal regulations Vol. 2,
on properties of o/w Pickering emulsions. Journal of Colloid and Interface Science Office of the Federal Register. U.S. Government Printing Office, Washington, DC.
351 (2), 348–356. Vanmarcke, A., Leroy, L., Stoclet, G., Duchatel-Crépy, L., Lefebvre, J., Joly, N., et al.,
UN

French, D.J., Brown, A.T., Schofield, A.B., Fowler, J., Taylor, P., Clegg, P.S., 2016. 2017. Influence of fatty chain length and starch composition on structure and prop-
The secret life of pickering emulsions: Particle exchange revealed using two colors erties of fully substituted fatty acid starch esters. Carbohydrate Polymers 164 (1),
of particle. Scientific Reports 6, 1–9. 249–257.
Gao, Y., Wang, L., Yue, X., Xiong, G., Wu, W., Qiao, Y., et al., 2014. Physicochemi- Wang, S., Luo, H., Zhang, J., Zhang, Y., He, Z., Wang, S., 2014. Alkali-induced
cal properties of lipase-catalyzed lauroylation of corn starch. Starch - Stärke 66 changes in functional properties and in vitro digestibility of wheat starch: The role
(5–6), 450–456. of surface proteins and lipids. Journal of Agricultural and Food Chemistry 62 (16),
Garcia-Armenta, E., Téllez, D., Sánchez, L., Alamilla, L., Hernández, H., Gutiérrez, 3636–3643.
L., 2016. Multifractal breakage pattern of tortilla chips as related to moisture con- Weiss, J., Decker, E.A., McClements, D.J., Kristbergsson, K., Helgason, T., Awad, T.,
tent. Journal of Food Engineering 168, 96–104. 2008. Solid lipid nanoparticles as delivery systems for bioactive food components.
Granillo, V., Villalobos, J., Alamilla, L., Téllez, D., Hernández, H., Dorantes, L., et al., Food Biophysics 3 (2), 146–154.
2017. Optimization of the formulation of emulsions prepared with a mixture of vit- Whitney, K., Reuhs, B., Ovando, M., Simsek, S., 2016. Analysis of octenylsuccinate
amins D and E by means of an experimental design simplex centroid and analysis rice and tapioca starches: Distribution of octenylsuccinic anhydride groups in
starch granules. Food Chemistry 2011, 608–615.
10 Food Hydrocolloids xxx (2018) xxx-xxx

Williams, P., Kuzina, F., Hlynka, L., 1970. A rapid colorimetric procedure for estimat- Xia, X., Li, G., Liao, F., Zhang, F., Zheng, J., Kan, J., 2015. Granular structure and
ing the amylose content of starches and flours. Cereal Chemistry 47, 411–420. physicochemical properties of starches from amaranth grain. International Journal
Wysota, B., Michael, S., Hiew, F.L., Dawson, C., Rajabally, Y.A., 2017. Severe but re- of Food Properties 18 (5), 1029–1037.
versible neuropathy and encephalopathy due to vitamin E deficiency. Clinical Yusoff, A., Murray, B.S., 2011. Modified starch granules as particle-stabilizers of
Neurology and Neurosurgery 160, 19–20. oil-in-water emulsions. Food Hydrocolloids 25 (1), 42–55.

F
OO
PR
ED
CT
R RE
CO
UN

You might also like