You are on page 1of 8

Food Chemistry 301 (2019) 125268

Contents lists available at ScienceDirect

Food Chemistry
journal homepage: www.elsevier.com/locate/foodchem

Morphological, technological and nutritional properties of flours and T


starches from mashua (Tropaeolum tuberosum) and melloco (Ullucus
tuberosus) cultivated in Ecuador
M. Teresa Pachecoa, F. Javier Morenoa, , Rodrigo Morenob, Mar Villamiela,

Oswaldo Hernandez-Hernandeza
a
Instituto de Investigación en Ciencias de la Alimentación (CIAL) (CSIC-UAM) CEI (CSIC+UAM), Campus de la Universidad Autónoma de Madrid, Nicolás Cabrera 9,
28049 Madrid, Spain
b
Instituto de Cerámica y Vidrio (ICV), Consejo Superior de Investigaciones Científicas (CSIC), Madrid, Spain

ARTICLE INFO ABSTRACT

Keywords: Morphological, technological and nutritional analyses were done in two scarcely studied starches from Andean
Starch tubers (mashua and melloco). The low sedimentation values, and the high zeta potential of mashua and melloco
Morphology starches in cold dispersions, as consequence of their electronegativity, indicated a better behaviour as stabilizer
Viscosity than potato starch. During heating, mashua and melloco starches presented much higher viscosity than potato
Resistant starch
starch, associated with their higher average particle size and greater amylose content. DSC and TGA analyses
Digestibility
indicated that melloco starch had the highest gelatinization enthalpy ΔHgel (12.32 J g−1) and degradation
temperature (270 °C), in comparison with potato starch, which are indicators of a better thermal resistance.
Consequently, extracted mashua and melloco starches could be excellent and cost-effective thickening or gelling
agents in both foods and a wide range of biomaterials. Mashua and melloco starches exhibited a digestion rate
close to 80%, which agreed with the low resistant starch content.

1. Introduction promote social and economic development.


The Andean region is known for its great diversity of roots and tu-
Starch is the most important food polysaccharide, maize, cassava, bers that are consumed by the rural population as staple foods because
potato, wheat and rice being the most important sources for obtaining of the high amount of carbohydrates in them, especially starch. In tu-
the commercially available starch. The global production of starch berous plants, on average, starch accounts for no more than 16–24% of
reached more than 340.1 million Tons in 2017; however, there are 795 their weight, the rest being water and other non-starchy components
million people who still suffer from hunger, with more than two billion (Hammes, Brandt, Rosenheim, Seitter, & Vogelmann, 2005). However,
people undergoing micronutrient deficiencies (FAO, 2017). The hunger some Andean tubers, such as mashua (Tropaeolum tuberosum) and
situation is exacerbated by other global issues, such as uneven demo- melloco (Ullucus tuberosus), have showed to possess more than 35% of
graphic expansion, climate change, intensification of natural disasters, starch, together with a high resistance to pests, low temperatures and
upsurges in pests and diseases, and the need to adjust to major changes drought (Campos, Chirinos, Gálvez Ranilla, & Pedreschi, 2018; Morón,
of the global food systems (Ojuederie & Ogunsola, 2017). In addition, 1999). Mashua and melloco cultivated in Ecuador (at 2800–3600 m
despite the fact that global food production has increased, the Dietary above the sea level, m.a.s.l.) constitute two underutilised sources of
Energy Supply (DES) in middle- and low-income countries (e.g. some of starch whose thermal, technological, functional properties and appli-
the nations of Latin America and the Caribbean) remains below the DES cations are still scarcely described.
observed in high-income countries (FAO, 2017). These trends and Starches can be classified into three main groups according to their
challenges have spurred the interest in seeking new sources of nu- digestibility properties. These categories are: rapidly digestible, slowly
trients, that can contribute to a healthier lifestyle, which, in turn, can digestible and resistant starches. Although the largest portion of starch

Corresponding author at: Instituto de Investigación en Ciencias de la Alimentación (CIAL) (CSIC-UAM), C/ Nicolás Cabrera 9, Campus de la Universidad

Autónoma de Madrid, E-28049 Madrid, Spain.


E-mail addresses: javier.moreno@csic.es (F.J. Moreno), rmoreno@icv.csic.es (R. Moreno), m.villamiel@csic.es (M. Villamiel),
o.hernandez@csic.es (O. Hernandez-Hernandez).

https://doi.org/10.1016/j.foodchem.2019.125268
Received 26 February 2019; Received in revised form 25 June 2019; Accepted 26 July 2019
Available online 27 July 2019
0308-8146/ © 2019 Elsevier Ltd. All rights reserved.
M.T. Pacheco, et al. Food Chemistry 301 (2019) 125268

is digested in the human small intestine, a substantial part may reach the fact that the diffraction angle is inversely proportional to the par-
the large bowel. Resistant starch (RS) is defined as the portion of starch ticle size. The size distribution was described in terms of d (0.1), d (0.5)
that is not digested in the small intestine but reaches the large intestine and d (0.9) that correspond respectively to 10, 50 and 90% of particles
in its intact form (Lovegrove et al., 2017). Recently, RS has gained more that are smaller than the reported size.
importance because of its positive physiological benefits including the
prebiotic effect, improvement of cholesterol metabolism, and reduction 2.2. Stability of starches in aqueous dispersion
of the risks of ulcerative colitis and colon cancer (Shi & Maningat,
2017). RS can be classified into five types: physically inaccessible starch 2.2.1. Sedimentation
(RS1); ungelatinized starch (RS2); retrograded starch (RS3); chemically Sedimentation of each starch was analysed using a Turbiscan MA
modified starch (RS4); and amylose-lipid complex (RS5). RS3 has been 2000 (AGS Instrument). Glass tube was filled with starch dilution 8%
particularly studied because of its beneficial functions in cereal pro- w/v of deionized water (pH 6.5–6.7) up to a height of 6.5 cm, and it was
ducts, such as thermal stability, high gelatinization temperature, low immediately placed inside the device. The transmittance and back-
water holding properties, improved texture, appearance and organo- scattering were measured every 2 min, during 1 h, at 20 °C.
leptic properties (Shi & Maningat, 2017).
A fundamental characteristic of native starches from different ve- 2.2.2. Zeta potential (ζ)
getal sources is that their granular and molecular structures influence The zeta potential (ξ) was analysed using a Malvern Zeta sizer Nano
their physicochemical and functional properties. The thickening, ZS instrument (Malvern Instruments Ltd., Worcestershire, UK). A sus-
binding and emulsifying properties of starches make them useful in the pension was first prepared by dispersing 250 mg of starch in 250 mL of
food industry, cosmetics manufacturing and in the development of KCl 0.1 M (mixture 1). Mixture 1 was agitated, sonicated during 1 min,
biomaterials (Ogunsona, Ojogbo, & Mekonnen, 2018). Therefore, al- and its pH was measured. Then, 10 mL of mixture 1 and 90 mL of KCl
though several studies have examined the different applications of 0.1 M were agitated, sonicated during 1 min, and its pH was recorded
starch (Vilpoux, Brito, & Cereda, 2019), it deems valuable to evaluate (mixture 2). Then, the cell of the equipment was filled with mixture 2,
new starch sources with specific characteristics to explore possibilities using a syringe, and the zeta potential was measured. The procedure for
for new industrial applications across sectors (Turola Barbi, Lopes the preparation of mixtures was repeated, in order to obtain dilutions of
Teixeira, Silveira Hornung, Ávila, & Hoffmann-Ribani, 2018). In con- mixture 1, at different pH (from 2 to 10), by adding of HCl 0.1 M or
sequence, the aim of this work was to analyze the morphological, KOH 0.1 M, and their respective ζ values were measured.
technological and nutritional properties of flours and starches extracted
from mashua (Tropaeolum tuberosum, var. INIAP-ECU-Izaño) and mel- 2.2.3. Cold flow curve
loco (Ullucus tuberosus, var. INIAP-ECU-amarillo-rosa) cultivated in Flow curves were monitored only for starches, since flours are
Ecuador to determine their possible industrial applications. usually used to prepare food with heat, in order to confer consistency,
rather than a gelling ingredient. Flow curves of 8% w/v of starch dis-
2. Materials and methods persions in water (23 °C) were recorded by performing hysteresis loop
tests using a Rheometer (Modular Advanced Rheometer system MARS,
Mashua (Tropaeolum tuberosum, var. INIAP-ECU-Izaño) and melloco Thermo Haake, Germany). Shear rate was continuosly increased from 0
(Ullucus tuberosus, var. INIAP-ECU-amarillo-rosa) were cultivated in to 1000 s−1, maintained constant at 1000 s−1 and decreased to 0 s−1;
Cotopaxi-Ecuador (2800–3600 m.a.s.l.) and harvested in mid-April. The the duration of each step was 300, 60 and 300 s, respectively. The
tubers were cleaned, lyophilized, ground to obtain the flours, and measuring geometry used was a double gap cylinder system DG41.
maintained at −20 °C. Starch was obtained by rinsing the flour with
deionized water (1:5 w/v) through a 250 µm-mesh sieve. After de- 2.3. Behaviour during heating
canting the filtered liquid, the precipitate was resuspended in water and
washed again several times until the supernatant was clear, by cen- 2.3.1. Viscosity during heating
trifugation at 10,000 × g for 15 min at 20 °C, separating the brownish The applied methodology has been previously described by de
layer from the upper part. Flours and starches were preserved at −20 °C Souza Gomes et al. (2018). Briefly, the gelation temperature during
until the analysis. Potato starch and corn flakes, used as standard of heating was determined using the rheometer described above. Starch
resistant starch, were purchased from Sigma-Aldrich Ltd (Switzerland) dispersions were prepared in water to a concentration of 8% w/v and
and Kellogg Company (Svendborg, Denmark), respectively. Chemicals stabilized at 50 °C for 1 min and then heated from 50 °C to 90 °C, at a
were reagent grade from J.T. Baker (Phillipsburg, NJ) and enzymes heating rate of 6 °C min−1, maintaining a shear rate of 50 s−1.
were from Sigma (Sigma Co., St. Louis, MO, USA).
2.3.1.1. Differential scanning calorimetry (DSC). The thermal properties
2.1. Morphological analysis and particle properties were measured using a differential scanning calorimeter (Discovery
DSC T. A. Instruments) previously calibrated with 99.99% purity
2.1.1. Scanning electron microscopy (SEM) indium, melting point of Tp = 156.6 °C, ΔH = 28.56 J/g. The samples
Morphological observations of flours and starches were performed (3 mg) and 7 mg of Milli-Q water were placed in a DSC hermetic
using scanning electron microscopy (SEM). Samples were mounted on aluminium pan. The sample was sealed, and maintained at room
aluminium stubs with sticky double-sided conductive metal tape and temperature during 4 h, then heated from 20 °C to 95 °C (10 °C/min).
vacuum-metalized with gold-palladium. Microscopic images were elu- An empty pan was used as a reference. From the DSC curve, the onset
cidated using a DSM 950 scanning electron microscope (Zeiss Iberia, (To), peak (Tp), conclusion (Tc) and enthalpy (ΔH) were calculated.
Madrid, Spain) at 7 kV accelerating voltage, a magnification of 1000×
and a distance of 11 mm. Carbon adhesive tabs and aluminium mount 2.3.1.2. Thermogravimetric analysis (TGA). Thermogravimetric
were acquired from Aname (Madrid, Spain). analyses of flours and starches were carried out with a
Thermobalance TGA Q 500 TA Instrument (USA). Analyses were
2.1.2. Particle size distribution conducted initially under a nitrogen flow of 10 and 90 mL/min for
Particle size distribution of starch samples was determined at room the purge and dehydration, respectively; and then, with oxygen for
temperature in water dispersion using a laser scattering analyser carbon combustion. The assays were carried out in platinum pans
(Malvern Instruments, Ltd, UK, Model Master Sizer S), and a refractive without lids. Thermograms were obtained over a temperature range
index (nD) of 1.53 (20 °C). The method used in this instrument relies on comprised between 20 and 900 °C with a heating rate of 10 °C/min. The

2
M.T. Pacheco, et al. Food Chemistry 301 (2019) 125268

non-oxidative decomposition (heating in nitrogen atmosphere) of


starch granules was evaluated directly by the mass loss between % Resistant starch =
glucose ( )
ug
mL
volume 100
ug
0.9( mL )
1000 dry sample weight
235 °C and 500 °C. To characterize the temperature range and the
peak of combustion, the first derivative of mass loss according to the
combustion temperature was used.
2.4.4. Retrograded resistant starch (RRS)
RRS was determined in gelatinized and cooled samples, according to
2.4. Nutritional properties
the method purposed by Saura-Calixto, Goñi, Bravo, and Mañas (1993)
with small modifications. After the gelatinization of 50 mg of sample,
2.4.1. Starch content and available starch (AS)
and successive incubations with 5 µL of α-amylase thermostable
The starch content of flours was expressed as the extraction yield,
from Bacillus licheniformis (≥500 units/mg protein; 4 µL/mL of po-
determined by the ratio between the final weight of the extracted
tassium phosphate buffer pH 6.9), 50 µL of pepsin (pig gastric mucosa –
starch, over the initial weight of the powder tuber sample. The avail-
∼2500 units/mg protein; 4 µL/mL of potassium phosphate buffer pH
able starch (AS) content of flours and starches was determined by the
6.9) and 30 µL of amyloglucosidase (Aspergillus niger – ∼14 units/mg
method described by Holm, Björck, Asp, Sjöberg, and Lundquist (1985),
protein), the supernatant was disregarded by centrifugation (3000g/
heating the diluted sample with α-amylase from Bacillus licheniformis
15 min) and the residue was washed with Milli-Q water, ethanol and
(Termamyl® Novozymes) at 90 °C for 20 min and then the resulting
acetone, through a centrifugation. A volume of 1.5 mL of water was
digesta were subjected to a second digestion with 25 µL of amyloglu-
added to each residue, followed by the addition of 3 mL of 4 M KOH.
cosidase from Aspergillus niger (∼14 units/mg protein) diluted in so-
The mixtures were stirred vigorously during 30 min at room tempera-
dium acetate buffer 0.1 M pH 4.75, (1:2.5), at 60 °C for 30 min. Re-
ture and then the pH was adjusted to 4.75 and 30 µL of amyloglucosi-
leased glucose was quantified with the reaction kit of glucose oxidase/
dase (Aspergillus niger – ∼14 units/mg protein) were added. Mixtures
peroxidase (Nzytech genes & enzymes, Portugal) at 510 nm. Available
were incubated at 60 °C for 30 min, and centrifuged (3000 g during
starch was expressed as g of available starch/100 g of pure starch.
15 min), collecting the supernatants in volumetric flasks and rinsing
with water at least once. Flasks were brought to volume and resistant
2.4.2. In vitro starch digestion rate
starch values were estimated by the equation displayed above.
In vitro starch digestion rate was analysed following the method
proposed by Holm, Björck, Drews, and Asp (1986) with slight mod-
ifications. Initially the equivalent of 100 mg of available starch was 2.5. Statistical analysis
diluted with 15 mL of 0.1 M sodium/potassium phosphate buffer pH
6.9. The samples were gelatinized at 95 °C for 20 min. Then, the sam- All the analyses were made in triplicate and the data were expressed
ples were placed in a bath at 37 °C and once the temperature was stable, as mean ± standard deviation. The Tukey’s test was used to compare
one aliquot of 20 µL was taken (time 0). Then 250 µL of α-amylase from the sample means at the 95% confidence level (p < 0.05) using the
hog pancreas (∼50 U/mg) prepared in sodium/potassium phosphate SPSS 22.0 software.
buffer pH 6.9 (2.3 mg/mL) was added and aliquots of 20 µL were then
collected at 5, 15, 30 and 60 min. All the aliquots were placed in tubes 3. Results and discussion
containing 80 µL of Milli-Q water and 100 µL of 3,5-dinitrosalicylic acid
(DNS). Mixtures were boiled for 10 min and cooled in an ice-water bath. 3.1. Morphological analysis and particles properties
After adding 1.5 mL of Milli-Q water, the absorbances were read at
530 nm. The hydrolysis rate was expressed as mg of maltose/100 mg of 3.1.1. Particle size and SEM morphology
pure starch. Fig. 1 presents the scanning electron microscope (SEM) images of
starches and flours from mashua and melloco. The more abundant
2.4.3. Resistant starch (RS) particles corresponded to the starch granules. The non-starchy com-
The RS content was measured in raw and gelatinized samples using ponents, present in both flours could be mainly fiber and/or protein,
the methodology proposed by Goñi, García-Diz, Mañas, and Saura- generally present in these types of Andean tubers (Campos et al., 2018;
Calixto (1996) with slight modifications. Samples (50 mg) were diluted Pacheco, Hernández-Hernández, Moreno, & Villamiel, in press). Fig. 1B
in 5 mL of 0.2 M KCl-HCl buffer pH 1.5 and incubated with 100 µL of points out that mashua starch presented three shapes of particles:
pepsin from the pig gastric mucosa (∼2500 units/mg protein) (0.1 g/ spherical small, spherical-truncated medium and ellipsoidal-truncated
mL buffer KCl-HCl) at 40 °C for 60 min. Then, 4.5 mL of 0.1 M Tris- large, while melloco starch (Fig. 1D) presented spherical small and oval
maleate buffer pH 6.9 was added and the pH was adjusted to 6.9, fol- medium and large particles. The electron micrographs in Fig. 1B and
lowed by the addition of 500 µL of α-amylase from hog pancreas Fig. 1D of starch preparations showed individual granules in the range
(∼50 U/mg) (40 mg/mL Tris buffer). The samples were incubated at of 5–30 µm, which were not as large as those estimated by light scat-
37 °C during 16 h in a thermostatic bath with constant agitation. Later, tering (Table S1, supplementary material). Nevertheless, the lower left
the samples were centrifuged (3000g/15 min), the supernatants were side of Fig. 1D provided an explanation by showing a large number of
removed, and the precipitate was washed with 5 mL of water and starch granules agglomerated together with thin sheet-like material
centrifuged again. A volume of 1.5 mL of water was added and stirred, which is likely derived from tuber cell walls. This finding is supported
followed by 1.5 mL of 4 M KOH; and the mixture was vigorously stirred by the fact that the total starch contents determined enzymatically in
at room temperature for 30 min. Then, 2.75 mL of 2 M HCl and 1.5 mL the samples of extracted starch were 92.42 g/100 g DM of mashua
of 0.4 M sodium acetate buffer were added, the pH was adjusted to 4.75 starch and 89.91 g/100 g DM of melloco starch (Table 1) (Pacheco
and 40 µL of amyloglucosidase from Aspergillus niger (∼14 units/mg et al., in press). Therefore, around a 10% of the extracted product could
protein) (5 mg/mL of sodium acetate buffer) was added and incubated correspond, as above indicated, to other non-starch compounds that
at 60 °C for 45 min with continuous stirring. The mixtures were cen- could be responsible for fusing/agglomerating the individual starch
trifuged (3000 × g/15 min), 5 mL of Milli-Q water were added on the granules into larger particles.
residues, and the centrifugation was repeated collecting the super- The determination of particle size performed by laser scattering
natants in 25 mL volumetric flasks. Finally, the glucose content was (Table S1, supplementary material) allowed the definition of three
measured, with the reaction kit of glucose oxidase/peroxidase (Nzytech different size distributions of particles. The average size distribution
genes & enzymes) at 510 nm. Resistant starch values were calculated (< 50%) observed for potato starch (control standard) was 46.9 μm,
using the following formula: similar value to that reported by Choi, Baik, and Kim (2017). In

3
M.T. Pacheco, et al. Food Chemistry 301 (2019) 125268

Fig. 1. Scanning electron micrographs of mashua flour (1A), mashua starch (1B), melloco flour (1C), melloco starch (1D).

contrast, the average particle size distribution (< 50%) observed in the starches in aqueous dispersion as compared to potato starch (p < 0.05)
mashua starch sample was higher to than observed in the melloco (Table S1, supplementary material). The lowest sedimentation values of
starch sample (190.2 vs. 146.0 μm); however, in the fraction < 90% the mashua and melloco starches could be due to a high repulsion force
sample of starch extracted from melloco presented the greater mean between the particles (Mandala & Bayas, 2004).
particle size (576.6 μm) followed closely by the mean particle size of the The electrostatic repulsive force (Zeta potential) can also predict the
mashua starch sample (560.2 μm). potential stability of the hydrocolloids system and also help to provide
The particle size and nature of the native starch granules derived stable formulations. If the particles have large negative or positive zeta
from mashua and melloco can dramatically influence the behaviour of potential values (that is, from −30 to −60 mV or from 30 to 60 mV),
the extracted starch during processing, as thermoplastic, stabilizing the system is considered stable and anti-cohesive (Genovese & Lozano,
and/or gelling agent. In general, the larger the particle size the better 2001). The zeta potential values obtained in this study indicated that
technological and/or functional properties can be found (Alcázar-Alay potato starch in aqueous dispersion is more stable at pH values above
& Meireles, 2015). 7.0 (Fig. 2). Nevertheless, the majority of foods have a pH lower than
7.0 (Badui, 2006), and at pH values from 6.5 to 3.0, mashua and
3.2. Stability of starches in aqueous dispersion melloco starches presented the highest absolute values of zeta potential,
which are indicative of aqueous dispersions more stable than those
3.2.1. Sedimentation and zeta potential obtained with potato starch (p < 0.5) (Fig. 2). These data are in good
The variation of light transmittance in the dispersions (8% w/v) agreement with the sedimentation behaviour, which was determined at
showed the sedimentation values for potato, mashua and melloco a pH range of 6.5–6.7 (Table S1, supplementary material).
starches, which indicated a higher stability of mashua and melloco On the other hand, when starch particles present greater

Table 1
Starch content (g/100 g DM), available starch (AS) and resistant starch (RS) (%) of mashua and melloco flours and starches.
Tuber Sample Starch AS RS RS RRS
content Raw Cooked Cooked and cooled

Mashua Flour 56.22 ± 1.05b 20.91 ± 0.45a 10.0 ± 0.02d 4.70 ± 0.01d 5.23 ± 0.01d
Starch *92.42 ± 1.87 cd 78.57 ± 2.59c 0.62 ± 0.03a 0.27 ± 0.02a 0.51 ± 0.02a

Melloco Flour 41.07 ± 0.90 a 60.95 ± 0.87b 7.73 ± 0.06c 0.63 ± 0.02bc 1.83 ± 0.03bbc
Starch *89.91 ± 0.53c 85.83 ± 1.64d 5.45 ± 0.18b 0.52 ± 0.03b 1.72 ± 0.02b

Values expressed as means ± SD (n = 3). RS Raw: Resistant starch in raw samples, RS Cooked: Resistant starch in cooked samples, RRS: Retrograded resistant
starch: Resistant starch in cooked and cooled samples. Different letters superscripts denote significant differences between values of the same column (p < 0.05).
* Data reported in Pacheco et al. (in press).

4
M.T. Pacheco, et al. Food Chemistry 301 (2019) 125268

Fig. 2. Zeta potential (ζ) of mashua and melloco starch dispersions at different pH.

electronegativity, they also can exhibit greater facility to retrogradation melloco flour presented a better behaviour during heating than mashua
(Genovese & Lozano, 2001). Therefore, mashua and melloco starches flour, showing a viscosity value of 3678 mPa.s vs.247 mPa.s
could have high potential to retrograde at the mentioned pH acquiring (p < 0.05). Recently, Pacheco et al. (in press) found a higher amylose
higher resistance to enzymatic digestion. content in melloco starch (51.70%) in comparison with potato starch
In addition, greater stability of cold aqueous dispersions of starch (18–29%) (Vamadevan & Bertoft, 2015). This high amylose content and
could be achieved by combination with stabilizers that inhibit the de- the high average particle size of melloco starch (Table S1, supplemen-
struction of starch granule and the leaching of amylose, and that in- tary material) can be related to its high viscosity at low gelatinization
crease the electrostatic repulsion between particles (Cai, Hong, Gu, & temperature. In the case of mashua starch, these properties seem to be
Zhang, 2011). related mainly to the larger average particle size in comparison with
According to the obtained results, mashua and melloco starches as potato starch, since both tubers have shown similar amounts of amylose
produced in the present study could be better food stabilizers of cold (29.55% vs. 18.00–29.00% of amylose) (Pacheco et al., in press).
mixtures than potato starch, even at low pH (e.g., some sausages, jams Based on these results, the starches extracted in the present study,
and fruit drinks), with the possibility to retrograde after the pasteur- that is, mashua starch, and especially, melloco starch, could confer
ization process and cold storage. higher viscosity than potato starch in foods prepared by heating, which
could allow the use of less amount of starch, to achieve the same
3.2.2. Cold flow curve thickener effect in foods, and consequently, having lower caloric value.
Shear stress and viscosity variations as a function of shear rate (flow
and viscosity curves, respectively) were recorded in a controlled rate 3.3.2. Differential scanning calorimetry (DSC)
mode up to a shear rate of 1000 s−1 using the ramp mentioned before. The results of the DSC analysis shown in Table 2 indicate a lower
Fig. S1 (supplementary material) shows the viscosity curves of 8% (w/ gelatinization (To-Tc) temperature range for melloco starch compared to
v) aqueous dispersions of potato, mashua and melloco starches. The mashua starch (46.4–57.5 °C vs 48.7–60.7 °C), and a gelatinization peak
viscosity of the dispersions was high at low shear rate and decreases (Tm) also lower for melloco starch (52.3 °C vs 54.9 °C) (p < 0.05).
with increasing shear rate, remaining constant for the three types of Additionally, these ranges were low compared to those reported for
starches at 8% (w/v) when increasing the shear rate above around potato starch (To-Tc: 58.06–67.27 °C, and Tm: 62.56 °C) (Fonseca-
50 s−1. Florido et al., 2017). The differences between the gelatinization tem-
The final viscosity was greater for melloco and mashua starches perature ranges can vary according to the botanical species. Lower
than for potato starch, which could be associated with the larger gelatinization temperatures may suggest higher crystalline perfection
average particle size of the former. Therefore, the starches extracted and/or higher amylose content (Tian et al., 2016). At the same time, a
from mashua and melloco could provide better behaviour than potato higher enthalpy value of gelatinization (ΔHgel) could be observed for
starch, as thickener of foodstuffs prepared in cold, such as soups, juices, melloco starch in comparison with mashua starch (12.3 vs. 4.3 J g−1)
sausages, films, and/or in pharmacology and cosmetics products that (p < 0.05), which indicated that melloco starch seems to possess a
require a certain and durable consistency at room or cold temperature. greater number of double helical areas in the amylose chains, and
therefore requires more energy for the breaking of the hydrogen bonds
3.3. Heat behaviour between the glucan chains (Zhang et al., 2019).
The higher Tm values for melloco and mashua flours as compared to
3.3.1. Viscosity during heating the starch samples (p < 0.05) can be due to the presence of compounds
Starches from mashua and melloco, at a shear rate of 50 s−1, dis- such as proteins, lipids and fiber that are degraded at higher tempera-
played a typical viscosity pattern based on a sharp increase of viscosity tures but with less energy consumption (ΔHgel) (Alcázar-Alay &
followed by a more gradual drop of viscosity under agitation and Meireles, 2015).
heating (Fig. 3). Melloco starch showed the highest viscosity peak,
reaching values of 17830 mPa.s at 66.1 °C, followed by mashua starch 3.3.3. Thermogravimetric analysis (TGA)
with a viscosity of 12220 mPa.s at 68.3 °C (p < 0.05). Likewise, The TGA curves (Fig. S2, supplementary material) showed the

5
M.T. Pacheco, et al. Food Chemistry 301 (2019) 125268

Fig. 3. Viscosity during heating of starches and flours from mashua and melloco.

Table 2 between the AS contents could be due to the botanical origin (Tovar,
DSC results for starch and flour from mashua and melloco. 2001). Mashua flour presented the lowest AS content (20.91%) in
Sample To (°C) Tm (°C) Tc (°C) ΔHgel (J g−1)
comparison with melloco flour (60.95%), despite its higher starch
content (p < 0.05). The presence of phenolic compounds in mashua
Mashua Starch 48.7 ± 0.3b 54.9 ± 0.2b 60.7 ± 0.6b 4.3 ± 0.1b samples (Pacheco et al., in press), which are well-known as α-amylase
Flour 58.4 ± 0.2d 63.4 ± 0.3 cd 69.3 ± 1.1 cd 1.2 ± 0.1a inhibitors (Ratseewo, Warren, & Siriamornpun, 2019), could explain
Melloco Starch 46.4 ± 0.1a 52.3 ± 0.2a 57.5 ± 0.3a 12.3 ± 0.4d
the low activity of enzymatic hydrolysis, and therefore a low amount of
Flour 56.5 ± 0.2c 62.3 ± 0.3c 67.9 ± 1.0c 9.0 ± 0.3c
AS. Melloco flour, instead, showed acceptable values for available
Different superscripts denote significant differences between values in the same starch (60.95%).
column (p < 0.05). Values expressed as means ± SD (n = 3). To: ‘onset’ initial
temperature, Tm: ‘melting’ peak temperature, Tc: ‘conclusion’ or final tem- 3.4.2. Flour and starch digestion rate in vitro
perature, ΔHgel: gelatinization enthalpy. Fig. 4 shows the in vitro digestion rate of flour and starch samples of
mashua and melloco. Either flours or starches from both Andean tubers
occurrence of three mass losses. The first loss (carried out in N2 at- in raw state presented low digestion rates because their crystalline
mosphere) was related to dehydration (20–110 °C). After that mass loss, structure was not destroyed (Villas-Boas, Yamauti, Moretti, & Franco,
the weight loss was stable, which is a characteristic behaviour for 2019). This compact structure limits the accessibility of digestive en-
starches. The second and third mass losses (occurred in O2 atmosphere) zymes, thus explaining the resistance of raw samples. Consequently,
were observed, which are related to the degradation due to carboni- this type of starch is very slowly and incompletely digested during its
zation of organic matter followed by oxidation (Hornung et al., 2017). passing through the small intestine (Shi & Maningat, 2017).
In the second step the greatest weight loss (Δm) was observed in mel- Gelatinized mashua and melloco starches presented a digestion rate
loco starch followed by mashua starch (p < 0.05); however, the weight close to 80%. In contrast, the corresponding cooked flour samples (after
loss of melloco starch occurred at 270 °C (Tp), while in mashua starch adjustment to the same amount of available starch) showed a much
was degraded at 210 °C. The mashua and melloco flours were more lesser digestion rate. This behaviour could be attributed to the presence
resistant to thermal degradation, exhibiting a greater residue (21.04% of fiber and/or protein and other food components that could prevent
and 21.64%) than their respective starches at the final of the heating effective diffusion and adsorption of α-amylase, as well as to the pre-
process (15.80% and 8.94%), in accordance with the higher degrada- sence of inhibitors of α-amylase as phenolic compounds (Pacheco et al.,
tion temperature of other components presents in mashua and melloco in press; Ratseewo et al., 2019). Remarkably, the digestion rate of
roots (eg. lipids, protein, fiber, ash) (Campos et al., 2018; Pacheco et al., cooked mashua flour was lower than that of melloco flour (p < 0.05),
in press). which is in line with its low AS content (Table 1) and it could also be
The observed thermal behaviour in the mashua and melloco star- attributed to the presence of the inhibitor compounds already men-
ches pointed out that they could be used in cooking processes requiring tioned.
high temperatures (< 210 and < 270 °C, respectively) without organic Considering that the lower the availability of starch, the lower the
matter degradation (carbonization). Mashua and melloco flours, in- digestibility and consequent conversion to glucose (Sáyago-Ayerdi,
stead, could be heated without organic degradation at lower tempera- Tovar, Osorio-Díaz, Paredes-López, & Bello-Pérez, 2005), the low starch
tures (< 125 and 130 °C, respectively). in vitro digestion rate observed in mashua cooked flour could point out
its application as a promising candidate to reduce the glycemic index
3.4. Nutritional properties (GI), as long as the ingredient has a low content of low molecular-
weight carbohydrates.
3.4.1. Starch content and available starch (AS)
Table 1 shows the starch content of flours and the available starch 3.4.3. Resistant starch (RS)
(AS) of mashua and melloco samples (flour and starch). In general, The resistant starch (RS) content was higher in melloco starch than
melloco starch and flour presented higher AS quantities than those in the mashua starch (p < 0.05) (Table 1). In contrast, the RS observed
observed in mashua starch and flour (p < 0.05). Differences observed in the raw melloco starch (5.45%) was much lower than the RS reported

6
M.T. Pacheco, et al. Food Chemistry 301 (2019) 125268

Fig. 4. Starch digestion rate in vitro of flour and starch from mashua (A) and melloco (B). A triplicate of the test was performed for all the samples. Error bars indicate
the variation of hydrolysis percentage observed between the different repetitions.

for raw potato starch (75.0%) (Raigond, Ezekiel, & Raigond, 2015). 24 h) was observed in all samples (Table 1), as a result of the re-
Starch resistance does not depend on any single factor but on a large arrangement of amylose chains around the remnants of amylopectin
number of factors, including the botanical origin, size of granules, during cooling, forming new crystalline structures that confer resistance
shape of granule surface, amylose content, starch crystallinity and size to digestion. The tubers may present lower resistance to enzymatic
of pores in starch granules (Tacer-Caba & Nilufer-Erdil, 2019). digestion following a loss of crystallinity in starch granules and the
In the case of flours, the value of RS determined in mashua flour was change of water distribution. With subsequent cooling, the disrupted
higher than the measured in melloco flour (p < 0.05), confirming the amylose and amylopectin chains gradually re-associated and ag-
presence of possible inhibitors of α-amylase, e.g. phenolic compounds, gregated, a process known as retrogradation (Tian et al., 2016), which
as already mentioned above. resulted in a higher content of retrograded resistant starch (RRS),
Some characteristics such as molecular weight, number and position compared to that observed in the cooked starch (RS) (Table 1). As a
of substitution and glycosylation of flavonoids seem to be related to the result, the digestibility of retrograded starch decreases because the
inhibitory effect of phenolic compounds (Martinez-Gonzalez et al., aggregation of melted amylose and amylopectin upon cooling and
2017). In a recent study, mashua flour presented a total phenolic con- storage, which make them less accessible to digestive enzymes (Xie, Hu,
tent of 1001.6 mg GAE/100 g DM vs. 260.1 mg GAE/100 g DM observed Jin, Xu, & Chen, 2014).
in melloco flour (Pacheco et al., in press). This reported difference Melloco starch showed a higher content of retrograded resistant
seems to be consistent with the observed inhibition effect. starch (RRS) as compared to mashua starch (1.72% vs. 0.51%, respec-
Overall, flours presented greater amount of RS than starches, pos- tively). Nevertheless, this content is still low as compared to that ob-
sibly due to its low rate of hydrolysis, as mentioned in the previous served, for instance, in improved genetically-modified lines of potato
section. rich in amylose and RS. The observed resistance could be improved by
increasing the cooling time or by genetic modification of the granular
structure (Chen, Singh, & Archer, 2018).
3.4.4. Retrograded resistant starch (RRS)
An increase of RS after cooking (99 °C/20 min) and cooling (4 °C/

7
M.T. Pacheco, et al. Food Chemistry 301 (2019) 125268

4. Conclusions starch blends. International Journal of Food Properties, 20, 611–622.


Genovese, D. B., & Lozano, J. E. (2001). The effect of hydrocolloids on the stability and
viscosity of cloudy apple juices. Food Hydrocolloids, 15, 1–7.
Our data highlight the potential role of two underexplored Andean Goñi, I., García-Diz, L., Mañas, E., & Saura-Calixto, F. (1996). Analysis of resistant starch:
tubers as sources of flours and starches with peculiar technological and A method for foods and food products. Food Chemistry, 56, 445–449.
nutritional properties due to the presence of proteins and/or mucilages Hammes, W. P., Brandt, M. J., Rosenheim, F. J., Seitter, M. F. H., & Vogelmann, A. S.
(2005). Microbial ecology of cereal fermentations. Trends in Food Science &
or other non-amylaceous components. According to the morphological Technology, 16, 4–11.
characterization and stability studies on cold and heated aqueous dis- Holm, J., Björck, I., Asp, N. J., Sjöberg, & Lundquist, I. (1985). Starch availability in vitro
persions, the starches extracted from mashua and melloco in the present and in vivo after flaking, steam cooking and popping of wheat. Journal of Cereal
Science, 3, 193–206.
study showed better behaviour as stabilizers than potato starch, usually Holm, B. J., Björck, I., Drews, A., & Asp, N. G. (1986). A rapid method for the analysis of
used as standard starch in this type of studies, indicating their possible starch. Starch/Stärke, 38, 224–226.
application as excellent and cost-effective thickening or gelling agents Hornung, P. S., Ávila, S., Lazzarotto, M., Lazzarotto, S. R. S., de Siqueira, G. L. A.,
Schnitzler, E., & Ribani, R. H. (2017). Enhancement of the functional properties of
not only in foods but also in a wide range of biomaterials, such as paper,
Dioscoreaceas native starches: Mixture as a green modification process. Thermochimica
textiles, plastics, cosmetic or pharmacological goods. Acta, 649, 31–40.
The nutritional analysis revealed that mashua and melloco starches Lovegrove, A., Edwards, C. H., De Noni, I., Patel, H., EI, S. N., Grassby, T., ... Shewry, P. R.
were highly available, with a high digestion rate. Moreover, the levels (2017). Role of polysaccharides in food, digestion, and health. Critical Reviews in Food
Science and Nutrition, 57, 237–253.
of resistant starch in mashua and melloco flours and starches were low, Mandala, I. G., & Bayas, E. (2004). Xanthan effect on swelling, solubility and viscosity of
regardless of the state of the sample (raw, cooked, or cooked and wheat starch dispersions. Food Hydrocolloids, 18, 191–201.
cooled). Therefore, mashua and melloco tubers could represent a novel Martinez-Gonzalez, A. I., Díaz-Sánchez, Á. G., De La Rosa, L. A., Vargas-Requena, C. L.,
Bustos-Jaimes, I., & Alvarez-Parrilla, E. (2017). Polyphenolic compounds and diges-
source of flours and starches with nutritional and physical character- tive enzymes: In vitro non-covalent interactions. Molecules, 22, 1–27.
istics different from the conventional botanical sources. Morón, C. (1999). Importancia de los cultivos andinos en la seguridad alimentaria y
nutrición. Memorias de la Reunión Técnica y Taller de Formulación de Proyecto
Regional sobre producción y nutrición humana en base a cultivos andinos. FAO.
Declaration of Competing Interest Centro Internacional de la Papa. Universidad Nacional del Altiplano. Universidad
Nacional de San Agustín. Lima, 31–53.
The authors declare no conflict of interest. Ogunsona, E., Ojogbo, E., & Mekonnen, T. (2018). Advanced material applications of
starch and its derivatives. European Polymer Journal, 108, 570–581.
Ojuederie, O. B., & Ogunsola, K. E. (2017). Impact of climate change on food security and
Acknowledgements its mitigation using modern biotechnology. Advances in Biotechnology & Microbiology,
3, 555–601.
Pacheco, M. T., Hernández-Hernández, O., Moreno, F. J., & Villamiel, M. Unexplored
The authors are grateful for the financing of MINECO-Spain
Andean tubers grown in Ecuador: New sources of functional ingredients. Food
(AGL2014-53445-R), the Community of Madrid (ALIBIRD-CM S-2013/ Bioscience. 2019. in press.
ABI-272), the support of the National Secretary of Higher Education, Raigond, P., Ezekiel, R., & Raigond, B. (2015). Resistant starch in food: A review. Journal
Science, Technology and Innovation of Ecuador (SENESCYT), and the of the Science of Food and Agriculture, 95, 1968–1978.
Ratseewo, J., Warren, F. J., & Siriamornpun, S. (2019). The influence of starch structure
collaboration of Nerea Muñoz (CIAL, CSIC-UAM) and Sergey Gorbatov and anthocyanin content on the digestibility of Thai pigmented rice. Food Chemistry.
(Vrije Universiteit Amsterdam). https://doi.org/10.1016/j.foodchem.2019.06.016.
Saura-Calixto, F., Goñi, I., Bravo, L., & Mañas, E. (1993). Resistant starch in foods:
Modified method for dietary fiber residues. Journal of Food Science, 58, 642–643.
Appendix A. Supplementary material Sáyago-Ayerdi, S. G., Tovar, J., Osorio-Díaz, P., Paredes-López, O., & Bello-Pérez, L. A.
(2005). In vitro starch digestibility and predicted glycemic index of corn tortilla, black
Supplementary data to this article can be found online at https:// beans, and tortilla-bean mixture: Effect of cold storage. Journal of Agricultural and
Food Chemistry, 53, 1281–1285.
doi.org/10.1016/j.foodchem.2019.125268. Shi, Y. -C. & Maningat, C. C. (2017). Resistant Starch. Sources, applications and health
benefits. In Y. -C. Shi & C. C. Maningat (Eds.) (pp. 1–285), 1st ed. John Wiley & Sons,
References Ltd.: Kansas, USA.
Tian, J., Chen, S., Wu, C., Chen, J., Du, X., Chen, J., ... Ye, X. (2016). Effects of pre-
paration methods on potato microstructure and digestibility: An in vitro study. Food
Alcázar-Alay, S. C., & Meireles, M. A. A. (2015). Physicochemical properties, modifica- Chemistry, 211, 564–569.
tions and applications of starches from different botanical sources. Food Science and Tacer-Caba & Nilufer-Erdil. (2019). Encyclopedia of Food Chemistry. In Laurence Melton,
Technology (Campinas), 35, 215–236. Fereidoon Shahidi, Peter Varelis (Eds.). Resistant Starch (pp. 571–575). EE.UU.:
Badui, D.S. (2006). Química de los Alimentos. Ed. Pearson educación. 4th Ed. México. Academic Press.
Cai, X., Hong, Y., Gu, Z., & Zhang, Y. (2011). The effect of electrostatic interactions on Villas-Boas, F., Yamauti, Y., Moretti, M. M. S., & Franco, C. M. L. (2019). Influence of
pasting properties of potato starch/xanthan gum combinations. Food Research molecular structure on the susceptibility of starch to α-amylase. Carbohydrate
International, 44, 3079–3086. Research, 479, 23–30.
Campos, D., Chirinos, R., Gálvez Ranilla, L., & Pedreschi, R. (2018). Bioactive Potential of Tovar, J. (2001). Fibra dietética en Iberoamérica: Tecnología y Salud. In F. M. Lajolo, F.
Andean Fruits, Seeds, and Tubers. In Advances in food and nutrition research (1st ed., Saura-Calixto, E. Witting, & E. Wenzel (Eds.). Métodos para la determinación de
Vol. 84, pp. 287–343). Elsevier Inc. almidón resistente en los alimentos (pp. 143–154). São Paulo: Varela Pres.
Chen, Y.-F., Singh, J., & Archer, R. (2018). Potato starch retrogradation in tuber: Turola Barbi, R. C., Lopes Teixeira, G., Silveira Hornung, P., Ávila, S., & Hoffmann-Ribani,
Structural changes and gastro-small intestinal digestion in vitro. Food Hydrocolloids, R. (2018). Eriobotrya japonica seed as a new source of starch: Assessment of phenolic
84, 552–560. compounds, antioxidant activity, thermal, rheological and morphological properties.
Choi, Y. J., Baik, M. Y., & Kim, B. Y. (2017). Characteristics of granular cold-water-soluble Food Hydrocolloids, 77, 646–658.
potato starch treated with alcohol and alkali. Food Science and Biotechnology, 26, Vamadevan, V., & Bertoft, E. (2015). Structure-function relationships of starch compo-
1263–1270. nents. Starch-Stärke, 67, 55–68.
de Souza Gomes, D., do Prado Cordoba, L., Rosa, L. S., Spier, M. R., Schnitzler, E., & Vilpoux, O. F., Brito, V. H., & Cereda, M. P. (2019). Starch extracted from corms, roots,
Waszczynskyj, N. (2018). Thermal, pasting properties and morphological character- rhizomes, and tubers for food application. In M. T. P. Silva Clerici (Ed.). Starches for
ization of pea starch (Pisum sativum L.), rice starch (Oryza sativa) and arracacha starch Food Application. Chapter 4 (pp. 103–165). Campinas, SP, Brazil: Academic Press.
(Arracacia xanthorrhiza) blends, established by simplex-centroid design. University of Campinas, School of Food Engineering.
Thermochimica Acta, 662, 90–99. Xie, Y. Y., Hu, X. P., Jin, Z. Y., Xu, X. M., & Chen, H. Q. (2014). Effect of repeated
FAO. (2017). Food outlook, Biannual report on global food markets. Food and Agriculture retrogradation on structural characteristics and in vitro digestibility of waxy potato
Organization of the United Nations. Rome. < http://doi.org/10.1044/leader.PPL. starch. Food Chemistry, 163, 219–225.
19102014.18 > . Accessed on December 27, 2018. Zhang, W., Wang, J., Guo, P., Dai, S., Zhang, X., Meng, M., ... Dou, H. (2019). Study on the
Fonseca-Florido, H. A., Hernández-Ávilab, J., Rodríguez-Hernández, A. I., Castro-Rosas, retrogradation behavior of starch by asymmetrical flow field-flow fractionation
J., Acevedo-Sandoval, O. A., Chavarria-Hernández, N., & Gómez-Aldapa, C. A. coupled with multiple detectors. Food Chemistry, 277, 674–681.
(2017). Thermal, rheological, and mechanical properties of normal corn and potato

You might also like