You are on page 1of 8

Food Hydrocolloids 97 (2019) 105212

Contents lists available at ScienceDirect

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

Effect of amylose content in morphological, functional and emulsification T


properties of OSA modified corn starch
Madai Lopez-Silvaa, Luis A. Bello-Pereza,∗, Edith Agama-Acevedoa, Jose Alvarez-Ramirezb
a
Instituto Politécnico Nacional, CEPROBI, Km. 6.5 Carr. Yautepec-Jojutla Col. San Isidro, Calle CEPROBI No. 8, Yautepec, Morelos, Mexico
b
Departamento de Ingeniería de Procesos e Hidráulica, Universidad Autónoma Metropolitana-Iztapalapa, Apartado Postal 55-534, Ciudad de México, 09340, Mexico

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

Keywords: Octenyl succinic starch esters were prepared from three corn starches with different amylose content; namely,
Amylose content waxy (5.43%), normal (25.16%) and Hylon VII (65.84%). The degree of substitution varies from 0.0103 for
OSA starch waxy starch to 0.0125 for Hylon VII, meaning that the degree of substitution is positively correlated with the
FTIR amylose content. Scanning electron microscopy images showed that the esterification reaction affected only the
Emulsification
granule surface while leaving the integrity of the granules. Solubility and water retention capacity were in-
creased with the OSA modification, with the effect more pronounced for higher amylose contents. X-ray analysis
revealed that the crystallinity type was not affected by esterification. FTIR analysis corroborated the changes in
the hydrated and ordered structures of starch granules. Emulsification tests showed that octenyl succinic starch
ester was an effective emulsifier for stabilization of oil-in-water emulsions. In particular waxy and Hylon VII
starches induced high stabilization degrees as revealed by long-term stability of oil-in-water emulsions.

1. Introduction branching points of amylopectin molecules. Analysis techniques like X-


ray diffraction of wide-angle did not identify the OSA modification
Esterification by octenylsuccinic anhydride (OSA) is an effective since changes in the X-ray diffraction pattern were not observed. In
method for modification of starch. The task is to improve the functional contrast, FTIR spectroscopy showed evidences of the esterification ef-
properties of starch by adding partial hydrophobic character, such that fects (Bai et al., 2009), suggesting that esterification takes place at lo-
the esterified starch presents amphiphilic character. OSA-modified cated functional groups of the starch chains. The distribution of OSA
starch has found sound applications in food products, including im- groups within the starch structure has been explored in the recent
provement of resistant starch fraction, stabilization of emulsions by the years. OSA groups appeared to be homogeneously distributed in waxy
“Pickering” mechanism, and stabilization of thin edible films corn starch where the amorphous regions are located in the branching
(Sweedman, Tizzotti, Schäfer, & Gilbert, 2013). It has been pointed out points of the amylopectin (Shogren, Viswanathan, Felker, & Gross,
that OSA esterification likely impairs the binding of α-amylase and 2000), although further studies suggested that the distribution is un-
decrease the starch digestion (He, Song, Ruan, & Chen, 2006; Zhu, Xie, even (Bai et al., 2014). Potato and normal maize starch with similar
Song, & Ren, 2011). Overall, it has been demonstrated that OSA-mod- amylose content, showed more OSA groups in the inner region of the
ified starch is a good derivative with important applications in the food granule than OSA potato starch. Also, a higher number of OSA groups in
industry (Sweedman et al., 2013). the periphery than in the inner of both starch granules was reported
The preparation method and the physicochemical properties of OSA (Wang, Su, & Wang, 2010, 2013). Overall, these studies indicated that
starch esters have been studied at some detail. However, the mechan- the distribution of OSA groups determine largely the physiochemical
isms underlying the starch modification by OSA have not been com- and functional properties of modified starches. OSA reaction on starch
pletely elucidated (Bai, Shi, & Wetzel, 2009). The issue is important for granules is a complex process affected by amylose content, particle size,
the design of proper procedures aimed to obtain starch esters with and molecular structure. Several reports have focused on the role of
tailored properties. It was postulated that OSA substitution is carried these factors on OSA starch reactions and emulsifying properties. The
out in the amorphous regions of starch (Bai, Kaufman, Wilson, & Shi, report by Sweedman, Hasjim, Schäfer, and Gilbert (2014) showed that
2014), which are located in the periphery of the granules (Jane & Shen, high amylopectin starches with a higher degree of branching (DB) and
1993; Pan & Jane, 2000) and randomly interspersed among the more rigid structure showed the best colloidal stability. Also, Wang,


Corresponding author.
E-mail address: labellop@ipn.mx (L.A. Bello-Perez).

https://doi.org/10.1016/j.foodhyd.2019.105212
Received 20 February 2019; Received in revised form 3 June 2019; Accepted 3 July 2019
Available online 06 July 2019
0268-005X/ © 2019 Elsevier Ltd. All rights reserved.
M. Lopez-Silva, et al. Food Hydrocolloids 97 (2019) 105212

Tang, Fu, Huang, and Zhang (2016) showed that smaller granules with 0.162 × (A × M )/ W
DS =
larger specific surface area had higher degree of substitution when re-
acted with OSA and showed more uniform distribution of octe-
(
1 − 0.21
A×M
W )
nylsuccinate substituents. where A is the titration volume of NaOH solution (mL), M is the mo-
The effects of amylose content in the properties of OSA-modified larity of NaOH solution, and W is the dry weight (g) of the modified
starch have been scarcely studied. The positive effect of amylose con- starch. The reaction efficiency (RE) was calculated as follows:
tent on the OSA modification was characterized (He, Song, Ruan, &
Chen, 2006), showing that amylose content increased the degree of Measured DS
RE = × 100
substitution and modifies the pasting properties of OSA starches (Song, Theoretical DS
Zhao, Li, Fu, & Dong, 2013). Starches with similar amylose content The theoretical DS was calculated by assuming that all the added
(∼25%) showed different degree of esterification (Whitney, Reuhs, OSA reacted with starch to form the ester derivative.
Martinez, & Simsek, 2016). The results indicated that the OSA groups
were present only in the amylose and perhaps in the amylopectin that
2.5. Polarized light microscopy
was hydrolyzed during the chemical modification.
The results published to date have indicated that amylose is the
A polarized light microscope (Eclipse 80i, Nikon, Japan) equipped
main receptor of OSA esterification. This suggests that the design of
with a 40x objective lens and a digital camera (Digital imaging Head,
starch esters with specific properties should rely on the accurate char-
DC330 camera MTI, Japan) was used. The dried starch was spread on a
acterization of the amylose content effects on esterification. In this re-
slide and a coverslip was added to the slide. A drop of deionized water
gard, the aim of this study was to compare the morphological proper-
was added to the edge of the coverslip and the images were captured
ties, crystallinity, thermal and pasting properties of OSA starch with
under polarized light.
different amylose/amylopectin ratio.

2.6. Scanning electron microscopy (SEM)


2. Materials and methods
Starch samples were fixed to conductive tape mounted on a brass
2.1. Materials
disc. Samples were then coated with gold using Polaron E5100 (Polaron
Equipment Ldt., Watford, UK). Images of starches were captured using a
Waxy corn (Amioca, CODE: 04401106), normal corn (Globe AA
scanning electronic microscope model JSM-5800LV (JEOL, Tokyo,
CODE: 03401037), and high-amylose corn (Hylon VII) starches were
Japan).
obtained from Ingredion®. All other chemicals and solvents were of
analytical grade.
2.7. Particle size distribution

2.2. Preparation of OSA-modified starch The size distribution of the starch granules was determined by laser
diffraction analysis (Mastersizer 2000; Malvern Instruments Ltd.,
The corn starch (30 g, dry weight) was dispersed in 100 mL of dis- Malvern, United Kingdom). The samples were analyzed using the Hydro
tilled water under gently stirring. The pH of the suspension was ad- 2000S accessory. The powders were diluted in water to achieve a sa-
justed to 8.75 using a solution of NaOH (0.1 M). A weighed amount of turation between 14 and 16% (concentration of ∼0.001%). The sam-
OSA (3.0% of the dried starch base) was added slowly over 2 h while ples were sonicated (250 rpm for 2 min) during the analysis to avoid
controlling the pH at 8.75. The reaction was allowed to proceed for a aggregation of the starch granules. The particle size is expressed as the
total of 6 h at 25 °C. After the reaction, the pH was adjusted to 7 with a mean diameter D [v, 0.5], which is defined as the diameter for which
solution of HCl (1 M). The mixture was centrifuged and washed three 50% of the particles by volume are larger and the cumulative volume
times with distilled water and once with acetone. The OSA-modified distribution.
starch was dried in an oven at 35 °C for 24 h, ground and then passed
through a 100 mesh.
2.8. Water retention capacity and solubility

2.3. Determination of amylose The water retention capacity was determined according to the
Hallgren method (Hallgren, 1985). Briefly, 5 mL of water was added to
The amylose content was determined using an amylose/amylo- 0.25 g of starch samples (native and modified) in pre-weighed cen-
pectin kit from Megazyme (Wicklow, Ireland). trifuge tubes at room temperature and heated at different temperatures
(50–90 °C) for 15 min, with agitation at 5 and 10 min . The tubes were
2.4. Determination of degree of substitution (DS) and reaction efficiency centrifuged for 15 min at 1000×g, 10 min. The supernatant was dec-
(RE) anted and the tubes were allowed to drain for 10 min at a 45Z° angle.
The tubes were weighed and then the weight gain was used to calculate
The DS is the average number of hydroxyl groups substituted per the percentage of gain as the water holding capacity. The experiments
glucose unit. The DS of OSA starch was determined using a titration were performed in triplicate.
method (Timgren, Rayner, Dejmek, Marku, & Sjöö, 2013). The OSA
starch (1.25 g) was dispersed by stirring for 30 min in 12.5 mL of a 2.9. X-ray diffraction analysis
0.1 M HCl solution. Subsequently, the sample was centrifuged at
3000×g for 10 min. The precipitate was washed with ethanol (90%) The X-ray diffraction patterns were obtained using a Rigaku dif-
and with distilled water. The starch was resuspended in 75 mL of dis- fractometer, model MiniFlex600 (Rigaku, Corporation Japan). The
tilled water and subjected to a boiling water bath for 10 min. It was moisture content of the samples was adjusted to 86% by storing the
allowed to cool to room temperature. Then, it was titrated with a samples in a desiccator with a saturated potassium sulfate solution for
standard solution of 0.1 N NaOH until pH 8.3 and as a reference it was 7 day at room temperature (Song and Jane, 2000). The range of the
titled native starch. The DS was calculated using the following equa- scanned diffraction angle (2θ) was 2–60° and the scanning speed was
tion: 2°/min.

2
M. Lopez-Silva, et al. Food Hydrocolloids 97 (2019) 105212

2.10. Thermal properties Table 1


Amylose content (%) of native corn starches and modified with 3% OSA.
The thermal properties were determined using a differential scan- Starch Native (g/100 g) Modified (g/100 g) Change (%)
ning calorimeter (DSC) (Q10, TA Instruments, USA). 2 mg of starch was
placed in an aluminum tray at room temperature (25 °C) and 7 μL Waxy 5.43 ± 0.15 c
4.04 ± 0.64 c
−25.59 ± 1.63b
Normal 25.16 ± 0.51b 14.85 ± 0.45b −59.65 ± 1.87a
(7 mg) of deionized water was added to obtain a moisture content of
Hylon VII 65.84 ± 0.67a 54.12 ± 1.00a −17.82 ± 1.12c
about 70%. The trays were hermetically sealed and allowed to stand for
12 h at room temperature for even distribution of water. The sample The values are the average of six replicates ± standard deviation. Different
trays were heated at a speed of 10 °C/min from 20 to 140 °C. An empty letters in the same column mean statistically significant difference (p < 0.05).
tray was used as a reference. The start temperature, the maximum
temperature, the completion temperature and the enthalpy of gelati- Table 2
nization (ΔH) were recorded. Degrees of substitution and reaction efficiency of corn starch esterification with
3% OSA.
2.11. FTIR spectroscopy
Starch DS × 102 RE (%)

c
FTIR analysis was carried out in a Vertex 70 FT-IR (Bruker Optik Waxy 1.03 ± 0.3 49.43 ± 0.35a
GmbH., Ettlingen, Germany) spectrometer equipped with a mercury Normal 1.12 ± 0.1b 52.58 ± 0.47b
Hylon VII 1.25 ± 0.4a 58.38 ± 0.93c
cadmium telluride detector and KBr beam splitter using a Platinum
Diamond ATR accessory with a diamond crystal at an angle of incidence The values are the average of six replicates ± standard deviation.
of 45°. Spectra were collected using 256 scans at 4 cm−1 resolution over DS = degrees of substitution; RE (%) = Reaction efficiency. Different letters
the entire 4000-400 cm−1 wavenumber region. Two spectra were col- mean statistically significant difference (p < 0.05).
lected for each sample and averaged. The analysis was performed using
OPUS version 7.0 software.

2.12. Preparation of emulsions

Oil-in-water emulsions were prepared using the method of Timgren


et al. (2013) with some modifications. The continuous phase of the
emulsions was made with 5 mmol/L of phosphate butter at pH 7 and
0.2 mol/L of NaCl, and the dispersed phase was liquid paraffin. Four
milliliters of continuous phase, 2 mL of dispersed phase, 1 mg of Solvent
Red 26 oil-soluble dye and starch in variable amounts (0–550 mg of
starch/ml of oil) were emulsified in glass test tubes by means of a
mixer, high shear in a SilentChruser M mixer (Heidolph, Germany) at
25,000 rpm for 120 s. The emulsions were observed for seven days.

2.13. Average droplet size

The volume-weighted mean droplet diameter (d4,3) was determined


on fresh emulsions and after 3, 7, 14, 21 and 28 days of storage at room
temperature (∼20 °C) with a laser light diffraction particle size ana-
lyzer (Mastersizer 2000, Malvern Instruments, Ltd.,Malvern,
Worcestershire, UK) using deionized water as dispersant. A relative
refractive index of 1.103 (ratio of the refractive indices of oil and water
phases) was used for all measurements.

2.14. Creaming stability

A dispersion analyzer (LUMiSizer, L.U.M. GmbH, Berlin, Germany) Fig. 1. Micrographs of polarized light of the granules of native and 3% of OSA
was used to determinate the creaming rate of the emulsions. It is a novel modified corn starches. Waxy (A, a), normal (B, b), and Hylon (C,c). Left panel –
instrument employing centrifugal sedimentation to accelerate the oc- native, and right panel - modified with OSA.
currence of instability phenomena such as sedimentation, flocculation
or creaming. Samples (0.4 mL) of emulsions were subjected to cen- 3. Results and discussion
trifugation at 4000 rpm for about 127 min at 25 °C, and 255 measure-
ments were determined at time intervals of 30 s. The creaming rate was 3.1. Amylose content
estimated by the SEPView equipment software, and was correlated to
the emulsion stability; namely, the higher the creaming rate, the lower Table 1 presents the amylose content of the three different un-
the emulsion stability. modified corn starches. As expected, waxy starch exhibited the lowest
amylose content (5.43%), while Hylon VII presented the highest value
2.15. ζ-Potential (65.84%). The OSA esterification decreased the content of detectable
amylose (i.e., apparent amylose), with the highest decrease for the
The ζ-Potential of the fresh emulsions was measured with normal corn starch (59.65%). The results in Table 1 suggest that OSA
ZetasizerNano ZS equipment. Emulsions samples were diluted in deio- esterification took place mainly on amylose chains, such that their
nized water (ratio 1:100 v/v) and then filled into the test cell and presence with standard amylose measurement methods was not de-
placed on the measuring equipment. The ζ-Potential was calculated by tected (Whitney et al., 2016). The issue is of importance given that
instrument using the Smoluchowski model.

3
M. Lopez-Silva, et al. Food Hydrocolloids 97 (2019) 105212

Fig. 2. Scanning electron micrographs (SEM) of native and 3% of OSA-modified


corn starches. The magnification of images was 2500×. Uppercase letters in- Fig. 3. Particle size distribution (PSD) of native (black continuous line)) and
dicate native starches, lowercase letters indicate modified starches. Waxy (A, a) OSA-modified (red dashed line) corn starches: (a) Waxy, (b) normal, and (c)
the arrows signal the pores on the surface and cracks; normal (B, b) the arrows Hylum VII. (For interpretation of the references to colour in this figure legend,
signal the pores on the surface and hylon VII (C, c) the signal arrows agglom- the reader is referred to the Web version of this article.)
erated on the surface after being esterified.

3.4. SEM microscopy


amylose content has large effect in the starch functionality. In fact, even
small differences in amylose content can have large effects in the The left panel of Fig. 2 presents SEM images of the unmodified
functional properties and digestibility features of starch (Jane et al., starch granules. The surface is smooth, without cavities or fractures.
1999). On the other hand, it has been pointed out that changes in ap- However, the OSA modification disrupted slightly the morphology of
parent amylose content could be attributed to OSA esterification or to the starch granule surface. In fact, some granules exhibited holes,
the mild alkaline treatment during the modification procedure (Simsek, cavities and small fractures produced by the starch treatment. However,
Ovando-Martinez, Marefati, & Rayner, 2015). it is not clear at all whether the surface modifications were provoked by
esterification reactions or by mild hydrolysis by the preliminary NaOH
treatment (Yan & Zhengbiao, 2010). It is noted that granules of OSA-
3.2. Degree of substitution and reaction efficiency modified Hylon VII starch (Fig. 3c) presents some protrusions linked
probably to the aggregation of OSA-esterified amylose chains otherwise
The degree of substitution (DS) is presented in the first column of concentrated on the granule surface. The large amylose content of
Table 2. The DS was positively correlated with the amylose content, Hylon VII allowed the extensive action on OSA starch. In contrast, the
suggesting that OSA esterification takes place primordially in the surface of the OSA-modified waxy starch (Fig. 3a) presented large
amorphous regions where amylose chains are concentrated. Similarly, cavities, caused probably by erosion induced by alkaline reactions.
the reaction efficiency (RE) was positively correlated with the amylose
content, as shown by Table 2 where the RE increased with the amylose
content. 3.5. Particle size distribution

Fig. 3 presents the particle size distribution of unmodified and OSA-


3.3. Polarized light microscopy modified starches. The mean diameter of waxy starch exhibited a slight
increase, from 18.23 μm to 19.35 μm (Fig. 3a). In contrast, the Hylon
Fig. 1 presents images of polarized light microscopy for unmodified VII mean diameter was unchanged. However, the mean diameter of the
and OSA-modified starches. Visible changes in the morphology of the normal corn starch exhibited a large increase, from 17.54 μm to
starch granules were not observed, indicating that OSA modification 25.42 μm. The increase of the mean diameter could be due to ag-
did not affect the integrity of the granules. It is apparent that OSA es- gregation effects induced by changes in the amphiphilic nature of OSA-
terification has the major effects in the fine structure of starch chains, modified starches.
while leaving untouched the solid structure of starch granules.

4
M. Lopez-Silva, et al. Food Hydrocolloids 97 (2019) 105212

Fig. 4. Solubility of native (■) and OSA-modified (□) corn starches: (a) Waxy,
(b) normal, and (c) Hylum.
Fig. 6. X-ray diffraction pattern of native (black line) and OSA-modified (red
line) of corn starches: (a) Waxy, (b) normal, and (c) Hylum. (For interpretation
of the references to colour in this figure legend, the reader is referred to the
Web version of this article.)

Table 3
Gelatinization characteristics of native and 3% of OSA modified corn starches.
Starch T0 (°C) Tp (°C) Tc (°C) ΔH(J/g sample)

a a a
NW 63.20 ± 0.64 71.98 ± 0.86 79.76 ± 0.26 9.61 ± 0.88a
MW 63.18 ± 0.62a 71.67 ± 0.28a 77.88 ± 0.35b 6.55 ± 0.63b
NN 68.40 ± 0.42a 73.81 ± 0.37a 81.50 ± 0.29a 10.44 ± 0.37a
MN 67.72 ± 0.08a 72.97 ± 0.09a 79.39 ± 0.17a 6.94 ± 0.10b
NH 71.52 ± 0.42a 89.72 ± 0.20a 107.80 ± 0.15a 10.45 ± 0.14a
MH 72.59 ± 0.08a 83.16 ± 0.46b 100.48 ± 0.05b 5.48 ± 0.20b

NW: native waxy, MW: modified waxy, NN: native normal, MN: modified
normal, NH: native Hylon VII, MH: modified Hylon VII. T0, Tp and Tc indicate
the temperatures of the onset, peak and conclusion of gelatinization, respec-
tively. ΔH indicates the enthalpy change of gelatinization. All data reported on
dry basis and represent the mean of triplicates. Means inside the columns of NW
and MW; NN and MN; NH and MH with different letters are significantly sta-
tistically different (p < 0.05).

3.6. Solubility and water retention capacity

The solubility of the unmodified and OSA-modified starches is dis-


played in Fig. 4. The solubility as function of temperature increased
with the OSA treatment, although the increase was quite low (less than
15%) for waxy and normal corn starches. In contrast, the increase was
more important for Hylon VII, with increases of the order of 120% for
90 °C. The high content of amylose and leaching of amylose chains
could be behind the relatively high solubility of Hylon VII. The water
Fig. 5. Water retention capacity (WRC) of native (■) and OSA-modified (□) retention capacity (WRC) is presented in Fig. 5. The WRC of the waxy
corn starches: (a) Waxy, (b) normal, and (c) Hylum VII.
starch was greatly increased after the OSA treatment. In contrast, Hylon
VII presented the smaller increase of the WRC. Temperature in WRC

5
M. Lopez-Silva, et al. Food Hydrocolloids 97 (2019) 105212

Fig. 8. (a) Illustration of the deconvolution of the FTIR signal in the starch
fingerprint region. (b) Intensity ratio R1047/1022 reflecting the ratio between
ordered and amorphous structures. (c) Intensity ratio R995/1022 reflecting the
ratio between hydrated and amorphous structures.
Fig. 7. FTIR diffraction pattern of native (black line) and OSA-modified (red
line) of corn starches: (a) Waxy, (b) normal, and (c) Hylon VII. (For inter-
pretation of the references to colour in this figure legend, the reader is referred The changes of the thermal properties in Table 2 suggest, as has been
to the Web version of this article.) already postulated by other researchers (Bai et al., 2014; Shogren et al.,
2000) that OSA treatment affects primordially the amylose chains.
exhibited a mixed effect in the increase of the WRC. Waxy corn starch However, the OSA treatment also affected the internal ordering (double
presented high WRC increase for low temperatures, while the opposite helices of amylopectin chains) of the starch granules.
behavior was exhibited by Hylon VII.
3.9. FTIR analysis
3.7. X-ray diffraction
Fig. 7 presents the FTIR spectra of the unmodified and OSA-mod-
Fig. 6 presents the XRD pattern of unmodified and OSA-modified ified starches. The interpretation of the FTIR spectrum follows the de-
starches. Waxy corn starch (Fig. 6a) shows a typical A-type starch scription by Simsek et al. (2015). The broad peak at 3500-3000 cm−1 is
crystallinity, with d-spacings at 5.8, 5.2 and 3.8 A (Vasanthan & Bhatty, linked to the presence of hydroxyl groups, reflecting the presence of
1996). Similar pattern was exhibited by the normal corn starch hydrating structures. The intensity of this peak increased after the OSA
(Fig. 6b). Hylon VII exhibited a different pattern (Fig. 5c), with a B-type treatment, and the increase was positively linked to the amylose con-
crystallinity pattern. As already reported (Bai et al., 2014), the OSA tent. In turn, the change in the intensity of the O–H stretching peak
modification did not change the crystallinity of the starch granules. This reflects the increase in the water retention capacity exhibited in Fig. 5.
is in line with the SEM image observations (Fig. 2), showing that the Peaks located at 2930 and 1650 cm−1 are attributed to the C–H
OSA modification only affected the surface of the starch granules. stretching vibration and bound water present in the starch, respec-
tively. The fingerprint region of starch is the large peak at 1100-
3.8. Thermal properties 900 cm−1 (C–O bond stretching). Here, the fingerprint peak is com-
posed by individual overlapping peaks that can be obtained by de-
The thermal properties, obtained via DSC analysis, are presented in convolution methods. Fig. 8.a illustrates the deconvolution results for
Table 3. The onset, peak and conclusion temperatures increased with the starch fingerprint region. The peak at about 1047 cm−1 has been
the amylose content, indicating that amylose is closely linked to the linked to ordered structures, and the peak at about 1022 cm−1 to
internal ordering of starch chains within the granules. These tempera- amorphous structures. On the other hand, the peak at 995 cm−1 re-
tures suffered a slight decrease, of about 2 °C, after the OSA treatment. flected hydrated structures. The ratios R1047/1022 and R995/1022 have
This suggests that the internal ordering of the starch granules was been proposed as indicators of ordered-to-amorphous and hydrated-to-
slightly disrupted by the OSA treatment. However, the effect could be amorphous starch structures (Van Soest, Tournois, de Wit, &
linked more to the NaOH conditioning of starch granules, rather than to Vliegenthart, 1995). Fig. 8b and 8.c exhibits the estimated ratios R1047/
the OSA reactions per se. The effect can be better assessed by the ge- 1022 and R995/1022 for the different starch samples. The waxy starch
latinization enthalpy, as this parameter decreased about 30–40% after exhibited the largest ratio R1047/1022, suggesting that the internal or-
treatment. The greater enthalpy decrease was exhibited by Hylon VII. dering is linked to the amylopectin structure. The value of this ratio

6
M. Lopez-Silva, et al. Food Hydrocolloids 97 (2019) 105212

Fig. 11. (a) Graphs of integrated transmission (%) versus time (s), and (b)
creaming velocity (μm/s) of emulsions.

Table 4
ζ-Potential of fresh emulsions.
Fig. 9. Photomicrographs of transmitted light (A) and polarized light (B) mi-
croscopy of the emulsions stabilized with native corn starches and OSA (3%) Emulsion ζ-Potential (mV)
modified starch at seven days. NW: native waxy, MW: modified waxy, NN:
MW −10.1 ± 0.2c
native normal, MN: modified normal, NH: native Hylon VII, MH: modified
MN −12.9 ± 0.5c
Hylon VII. Conditions: 550 mg/mL oil, 0.2 mol/L NaCl. Scale bar 10 μm. MH −93.7 ± 0.7a

Different alphabets indicate significant difference


(p < 0.05) between column values.

increased slightly after the OSA treatment, supporting the proposal that
OSA affects principally the amorphous regions of the starch granule.

3.10. Stability of emulsions

The emulsification capacity of OSA-modified starches is an im-


portant property for application in the food industry. Fig. 9 presents
images of the oil-in-water emulsion stabilized with unmodified and
OSA-modified starches. The stability of the emulsions is due to the
hydrophobic nature of starch particles, which migrate to the oil/water
interface to form a barrier against droplet coalescence and coagulation.
Images in the left panel of Fig. 9 illustrate the stability of the emulsions,
while the right panel uses light microscopy images for assessing the
morphology of droplets. OSA-modified starches led to emulsions with
improved stability, as showed by the absence of coalescence, creaming
and coagulation. In contrast, unmodified starches were unable to sta-
Fig. 10. Variation of mean droplet diameter (d4,3) of the emulsions monitored
bilize the emulsions. For instance, the normal unmodified starch gave
for 10 days of storage at room temperature.
the worst emulsifying properties as illustrates by the sharp separation of
phases. Emulsion made with unmodified starch exhibited coagulation
and sedimentation, while unmodified Hylon VII led to partial separa-
tion of phases and coagulation. It is apparent that the best emulsifying
properties were obtained for OSA-modified waxy starch since no phase
separation was observed.

7
M. Lopez-Silva, et al. Food Hydrocolloids 97 (2019) 105212

The results in Fig. 9 showed that OSA-modified starch has the ability Acknowledgments
of forming stable emulsions. Fig. 10 presents the variation of the mean
droplet size for emulsions from 0 to 10 days. Emulsion made with MW We appreciate the economic support from SIP, IPN, EDI-IPN
exhibited the smallest drop size (∼1.4 μm). In contrast, emulsion from CONACYT-México and COFAA-IPN. MLS also acknowledges the scho-
MH (ENWS) presented the highest drop size (∼3.0–5.0 μm). All emul- larship from CONACYT-Mexico.
sions were stable in terms of the droplet size, showing only slight in-
crease of about 10% in the first 10 days. The mean drop size of the Appendix A. Supplementary data
emulsion made with MH increased about 15%, to ∼4.8 μm, suggesting
that the emulsion was affected to some extent by drop coalescence. It is Supplementary data to this article can be found online at https://
apparent that drops tended to be aggregated for this emulsion, sug- doi.org/10.1016/j.foodhyd.2019.105212.
gesting that stability is not sufficient in terms of electrostatic repulsion
to maintain droplets distant from each other. References
Creaming can be also present in the evolution of emulsions. The
tendency to creaming was assessed by accelerating conditions, and the Bai, Y., Kaufman, R. C., Wilson, J. D., & Shi, Y. C. (2014). Position of modifying groups on
results are exhibited in Fig. 11.a. The integral transmission quantifies starch chains of octenylsuccinic anhydride-modified waxy maize starch. Food
Chemistry, 153, 193–199.
the amount of transmitted light through the emulsion bulk, and is Bai, Y., Shi, Y. C., & Wetzel, D. L. (2009). Fourier transform infrared (FT-IR) micro-
considered as an index of creaming. In this way, the emulsion made spectroscopic census of single starch granules for octenyl succinate ester modifica-
with MW presented the slowest creaming rate (Fig. 11b), with the other tion. Journal of Agricultural and Food Chemistry, 57(14), 6443–6448.
Hallgren, L. (1985). Physical and structural properties of cereals, sorghum in particular in
two emulsions exhibiting similar values of the creaming rate relation to milling methods and product use. Copenhagen-I: Carlsberg Research
(∼90–110 μm/s). Some insights on the origin of creaming can be ob- Laboratory, Technical University of Denmark.
tained by analyzing the ζ-potential of the emulsions. The measurement He, G. Q., Song, X. Y., Ruan, H., & Chen, F. (2006). Octenyl succinic anhydride modified
early indica rice starches differing in amylose content. Journal of Agricultural and Food
of the ζ-potential is based on the hydraulic mobility of particles in a
Chemistry, 54(7), 2775–2779.
fluid. Since oil drops dispersed in the aqueous phase are relatively large Jane, J. L., Chen, Y. Y., Lee, L. F., McPherson, A. E., Wong, K. S., Radosavljevic, M., et al.
(> 1.0 μm), the charges carried out by the oil-water interface hardly (1999). Effects of amylopectin branch chain length and amylose content on the ge-
latinization and pasting properties of starch. Cereal Chemistry, 76(5), 629–637.
contributed to the measured electrostatic charge. Instead, the larger
Jane, J., & Shen, J. (1993). Internal structure of the potato starch granule revealed by
contribution is made by the particles dispersed in the continuous phase. chemical gelatinization. Carbohydrate chemistry, 247, 279–290.
Table 4 shows that the emulsion made with MW exhibited the smallest Pan, D. D., & Jane, J. (2000). Internal structure of normal maize starch granules revealed
ζ-potential value (−10.1 mV), while the emulsion made with MH ex- by chemical surface gelatinization. Biomacromolecules, 1, 126–132.
Shogren, R. L., Viswanathan, A., Felker, F., & Gross, R. A. (2000). Distribution of octenyl
hibited the highest value (−23.7 mV). In turn, the emulsion made with succinate groups in octenyl succinic anhydride modified waxy maize starch. Starch
MH presented high value of the creaming rate. High ζ-potential values Staerke, 52(6‐7), 196–204.
suggest that the continuous phase is a repulsive medium that probably Simsek, S., Ovando-Martinez, M., Marefati, A., & Rayner, M. (2015). Chemical compo-
sition, digestibility and emulsification properties of octenyl succinic esters of various
led aggregation of the dispersed phase. In contrast, a continuous starches. Food Research International, 75, 41–49.
medium with relatively low ζ-potential allows the free (Brownian) Song, Y., & Jane, J. (2000). Characterization of barley starches of waxy, normal and high
mobility of the dispersed phase. amylose varieties. Carbohydrate Polymers, 41, 365–377.
Song, X., Zhao, Q., Li, Z., Fu, D., & Dong, Z. (2013). Effects of amylose content on the
paste properties and emulsification of octenyl succinic starch esters. Starch Staerke,
4. Conclusions 65(1‐2), 112–122.
Sweedman, M. C., Hasjim, J., Schäfer, C., & Gilbert, R. G. (2014). Structures of octe-
nylsuccinylated starches: Effects on emulsions containing β-carotene. Carbohydrate
The effects of amylose content (waxy: 5.43%, normal: 25.16%, and Polymers, 112, 85–93.
Hylon VII: 65.84%) on the physicochemical, thermal and emulsification Sweedman, M. C., Tizzotti, M. J., Schäfer, C., & Gilbert, R. G. (2013). Structure and
properties of OSA corn starch esters have been studied. The amylose physicochemical properties of octenyl succinic anhydride modified starches: A re-
view. Carbohydrate Polymers, 92(1), 905–920.
content is positively correlated with the effect of the OSA treatment. In
Timgren, A., Rayner, M., Dejmek, P., Marku, D., & Sjöö, M. (2013). Emulsion stabilizing
fact, solubility, water retention capacity and emulsification ability were capacity of intact starch granules modified by heat treatment or octenyl succinic
improved by OSA treatment, and the increase was positively linked to anhydride. Food Sciences and Nutrition, 1(2), 157–171.
amylose content. In turn, this indicates that OSA reactions affected Van Soest, J. J., Tournois, H., de Wit, D., & Vliegenthart, J. F. (1995). Short-range
structure in (partially) crystalline potato starch determined with attenuated total
extensively the amylose chains, and to a minor extent to amylopectin reflectance Fourier-transform IR spectroscopy. Carbohydrate Research, 279, 201–214.
chains. Besides, FTIR analysis showed that OSA modification increased Vasanthan, T., & Bhatty, R. S. (1996). Physicochemical properties of small-and large-granule
the short-range ordering, which is probably linked to the idea that es- starches of waxy, regular, and high-amylose barleys. USA: Cereal Chemistry.
Wang, C., He, X. W., Huang, Q., Luo, F. X., & Fu, X. (2013). The mechanism of starch
terification focuses on the amorphous regions of the starch granules. granule reacted with OSA by phase transition catalyst in aqueous medium. Food
On the other hand, stable emulsions were obtained for waxy starch, Chemistry, 141(4), 3381–3385.
while emulsions with limited stability were obtained for starches con- Wang, J., Su, L., & Wang, S. (2010). Physicochemical properties of octenyl succinic an-
hydride‐modified potato starch with different degrees of substitution. Journal of the
taining a large fraction of amylose. It was suggested that the stabili- Science of Food and Agriculture, 90(3), 424–429.
zation of the oil-water interface was achieved by the formation of Wang, C., Tang, C. H., Fu, X., Huang, Q., & Zhang, B. (2016). Granular size of potato
complex layers attached onto the interface. ζ-potential measurements starch affects structural properties, octenylsuccinic anhydride modification and
flowability. Food Chemistry, 212, 453–459.
indicated that OSA-modified starch formed repulsive medium that Whitney, K., Reuhs, B. L., Martinez, M. O., & Simsek, S. (2016). Analysis of octe-
likely led to creaming for long storage times. Overall, the results nylsuccinate rice and tapioca starches: Distribution of octenylsuccinic anhydride
showed that OSA-modified starch chains have the ability of stabilizing groups in starch granules. Food Chemistry, 211, 608–615.
Yan, H., & Zhengbiao, G. U. (2010). Morphology of modified starches prepared by dif-
oil-water interfaces, with amylopectin leading to more stable conditions
ferent methods. Food Research International, 43(3), 767–772.
in the long term. Zhu, W., Xie, H. L., Song, X. Y., & Ren, H. T. (2011). Production and physicochemical
properties of 2‐octen‐1‐ylsuccinic derivatives from waxy corn starch. Journal of Food
Conflicts of interest Science, 76(3), C362–C367.

The authors declare no conflict of interest.

You might also like