You are on page 1of 9

Food Hydrocolloids 120 (2021) 106883

Contents lists available at ScienceDirect

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

Structural, functional, and chemical properties of small starch granules:


Andean quinoa and kiwicha
Frank F. Velásquez-Barreto a, *, Hubert Arteaga Miñano a, Jose Alvarez-Ramirez b, Luis.
A. Bello-Pérez c
a
Escuela Professional de Ingeniería Agroindustrial, Facultad de Ciencias Agrarias, Universidad Nacional Autónoma de Chota, Chota, Cajamarca, 06120, Peru
b
Departamento de Ingeniería de Procesos e Hidráulica, Universidad Autónoma Metropolitana-Iztapalapa, Apartado Postal 55-534, Iztapalapa, 09340, Mexico
c
Instituto Politécnico Nacional, CEPROBI, Yautepec, Morelos, 62732, Mexico

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

Keywords: White, red, and black quinoa (Chenopodium quinoa W.) and kiwicha (Amaranthus caudatus) are increasingly
Starch consumed as specialty foods worldwide. Starch was extracted from these Andean grains and characterized by
Andean grains scanning electron microscopy, Fourier-transform infrared spectroscopy (FTIR), differential scanning calorimetry,
FTIR
and enzymatic hydrolysis. All starches showed a similar shape and small mean diameter (2–4 μm) because their
Molecular weight
physicochemical, functional, and digestibility characteristics are related to specific intrinsic factors. Kiwicha
In vitro digestibility
starches had greater swelling power and solubility at different temperatures than quinoa starches. Kiwicha starch
showed a lower final viscosity (approximately 550 mPa s) than quinoa starches (approximately 935.0–1015.0
mPa s), reflecting marked differences in molecular organization. The quinoa starches exhibited higher FTIR
1047/1022 ratios, indicating enhanced ordered crystallinity. The amylopectin molecular weight was similar for
all starches (1.95–2.37 × 108 g/mol). In contrast, significant differences in the molecular weight of amylose were
exhibited for quinoa starches (1.01–1.11 × 108 g/mol) and kiwicha starch (1.31 × 108 g/mol). The in vitro
enzymatic digestibility was similar for all starches, with marked fast and slow hydrolysis phases. The in vitro
starch digestibility was determined by the size of the starch granules rather than by their structure.

1. Introduction soups, desserts, adhesive agents, and, in recent years, in the processing
of biofilms and biodegradable packaging due to their versatility and
Andean grains, known since historical Inca times, are extensively properties (Cruz, Ribotta, Ferrero, & Iturriaga, 2016; Cruz-Tirado,
distributed in the high-altitude areas of South America (Alemayehu, Vejarano, Tapia-Blácido, Barraza-Jáuregui, & Siche, 2019; Shevkani,
Bendevis, & Jacobsen, 2015; Ruiz et al., 2014; Wang & Zhu, 2016). The Singh, Bajaj, & Kaur, 2017). Paste formation, swelling power, solubility,
adaptation, harvesting, and commercialization of quinoa (Chenopodium freezing stability and retrogradation extent are salient properties that
quinoa W.) grains has spread worldwide due to its high protein content determine starch applicability (Bashir & Aggarwal, 2019; Waterschoot,
and amino acid balance (Dakhili, Abdolalizadeh, Hosseini, Gomand, & Delcour, 2016).
Shojaee-Aliabadi, & Mirmoghtadaie, 2019; Das, 2016). Kiwicha Quinoa starches exhibit relatively small granule sizes (2–4 μm),
(Amaranthus caudatus) is another grain with characteristics similar to which has motivated their use as soft-matter stabilizers (Kierulf et al.,
quinoa. However, its production is lower due to a lack of detailed 2020; Timgren, Rayner, Dejmek, Marku, & Sjöö, 2013). Interestingly,
knowledge of its nutritional properties and chemical composition. It has kiwicha starches can offer properties similar to those of quinoa starches.
been reported that starch is the most abundant fraction in quinoa and In addition, quinoa and kiwicha starches exhibit type A crystallinite,
kiwicha grains (Martinez-Lopez, Millan-Linares, Rodriguez-Martin, which is a pattern also exhibited by most cereals (Jan, Panesar, Rana, &
Millan, & Montserrat-de la Paz, 2020; Srichuwong et al., 2017). Singh, 2017). Although the characteristics of quinoa starches have been
Novel botanical sources of starches with attractive properties for reported in recent reports, only a limited number of studies have been
food and nonfood applications have been explored in recent years conducted to explain the relationship between the structural, functional,
(Velásquez-Barreto et al., 2020). Native starches can be used to prepare nutritional properties and rheological properties of the native starches of

* Corresponding author.
E-mail addresses: frankervba@hotmail.com, fvelasquez@unach.edu.pe (F.F. Velásquez-Barreto).

https://doi.org/10.1016/j.foodhyd.2021.106883
Received 1 March 2021; Received in revised form 9 April 2021; Accepted 1 May 2021
Available online 13 May 2021
0268-005X/© 2021 Elsevier Ltd. All rights reserved.
F.F. Velásquez-Barreto et al. Food Hydrocolloids 120 (2021) 106883

quinoa and kiwicha. For instance, Li and Zhu (2017) showed that the obtained with a scanning electron microscope (Carl Zeiss EVO MA10,
amylopectin structure of quinoa starches has a determinant role in Germany) by following the procedure reported by Velásquez-Barreto
swelling power, relative crystallinity, and paste-forming properties. et al. (2020). The dried quinoa and kiwicha starches were placed onto a
Despite such results, further studies are still needed to obtain a more carbon strip and moved onto the microscope cells. The power used for
detailed relationship between the different functional and physico­ capturing the photomicrographs was 15 kV–20 kV. The photomicro­
chemical properties. graphs were processed using equipment software.
The amylose and amylopectin fractions in starch granules are not
evenly distributed. A complex pattern of interlaced amorphous and 2.5. Particle size distribution
crystalline structures is present, which in turn has a determinant effect
on functional and nutritional properties (Hoover, 2001). In particular, Laser diffraction (Mastersizer 3000, Malvern Instruments Ltd., Mal­
the relative amylose content and the fraction of long amylopectin chains vern, Worcestershire, UK) equipment was used to obtain the particle size
can affect the viscoelasticity of starch pastes (Kyung & Yoo, 2014; distribution for the starch samples. Water was used as the dispersing
Velásquez-Barreto & Velezmoro, 2018). On the other hand, in vitro agent with a refractive index of 1.33. The starch granule size was
enzymatic hydrolysis is largely affected by crystallinity, size distribu­ recorded over the range of 0.02 μm–1000 μm, and the average diameter
tion, and binding of amylose and amylopectin molecules with proteins was analyzed using equipment software.
and lipids (Ai, Hasjim, & Jane, 2013; Lin, Yang, Chi, & Ma, 2020;
Tacer-Caba & Nilufer-Erdil, 2019). 2.6. Solubility and swelling power
The extraction and purification of proteins from quinoa and kiwicha
grains has been addressed in recent years (Dakhili et al., 2019). The solubility and swelling power (SP) of quinoa and kiwicha
Although some studies have characterized the properties of quinoa and starches was assessed with the procedure reported by Jan, Panesar,
kiwicha starches, knowledge of the physicochemical, structural, func­ Rana, and Singh (2017). To this end, 1% starch suspensions (0.3 g/30
tional, and nutritional properties is still incomplete. An accurate char­ mL) were placed into plastic tubes and vortexed for 20 s. Then, the
acterization of starch properties should provide valuable insights into suspensions were placed into a 40 ◦ C water bath and heated for 30 min
the applicability of Andean grains, particularly quinoa and kiwicha. with intermittent vortexing. Subsequently, the suspension was cooled at
Therefore, this study aimed to evaluate the structural, functional and 20 ◦ C and centrifuged at 3000 rpm for 20 min. The supernatant was
chemical properties of quinoa and kiwicha starches. separated and dried at 120 ◦ C to constant weight. The dry weight of the
supernatant (P1), weight of the sediment after centrifugation (P2), and
2. Materials and methods the initial weight of the starch (P) were used to calculate the SP and
solubility of the starch (equations (1) and (2)). The same procedure was
2.1. Samples repeated for temperatures ranging from 50 to 90 ◦ C (10 ◦ C increments).

Quinoa grains (Ch. quinoa W., varieties white, red and black) were Solubilility (%) =
P1
× 100 (1)
harvested in the District of Cabana, Province of San Román, Region P
Puno, Peru in 2020. Kiwicha grains (A. caudatus) were harvested in 2020
P2 × 100
in the Yungay community, Yungay Province, Ancash Region. SP (g / g) = (2)
P × (100 − Solubility)

2.2. Starch extraction


2.7. Pasting properties
The methodology proposed by Jan, Panesar, and Singh (2017) was
used for starch extraction. Previously, quinoa and kiwicha grains were An Ar-1500ex rheometer (TA Instruments, SA) equipped with a
selected and washed under running water. The grains were soaked in a starch cell was used to determine the paste-forming properties of the
0.25% NaOH solution for 24 h in a 1/6 w/v ratio, rinsed with water and starches. To this end, an 8% w/w starch suspension sample was placed
blended with ten volumes of distilled water in a Waring blender for 3 into the starch cell and stirred at 300 s− 1. The suspension was heated to
min. The homogenate was sifted through a screen 100 US mesh (150 μm) 40 ◦ C for 2 min and then heated to 95 ◦ C (heating rate 7.5 ◦ C/min) for
until the washing water was clean. The homogenate was centrifuged at 10 min. Subsequently, the suspension was cooled to 50 ◦ C (cooling rate
10800 g for 15 min, and the sediment was resuspended in distilled water 7.5 ◦ C/min) for 2 min. The pasting properties of the starch pastes were
and centrifuged. This step was repeated three times. The white sediment recorded using TA Data32 software. Specifically, the pasting tempera­
was allowed to settle until complete sedimentation, the water was ture (Tp, ◦ C), peak viscosity (PV, Pa.s), minimum viscosity (MV, Pa.s),
decanted, and the wet starch was dried at 45 ◦ C for 72 h. final viscosity (FV, Pa.s), breakdown viscosity (BDV, Pa.s) and setback
viscosity (SBV, Pa.s) were obtained to assess the pasting characteristics
2.3. Chemical composition (Hoyos-Leyva, Bello-Pérez, Alvarez-Ramirez, & Agama-Acevedo, 2017).

The proximal chemical composition (humidity, protein, fat, ash and 2.8. Thermal properties
fiber) of grains and starches was determined using the standard AACC
methods 44–15.02, 46–13.01, 30–25.01, 08–01.01 and 32.10.01, The thermal properties of the quinoa and kiwicha starches were
respectively (AACC, 2000). The carbohydrate content was determined determined using a differential scanning calorimeter (DSC, TA Instru­
from a comparison with the other components. Phosphorous content ment, Q20, New Castle, NJ, USA). Approximately 2 mg of starch on a dry
was determined using standard AOAC methods (AOAC, 2005). The basis was placed into an aluminum cell, and then 7 μL of distilled water
amylose (%) content of the quinoa and kiwicha grains was analyzed was added. Subsequently, the cell was sealed and placed into a calo­
using Megazyme International Ireland Ltd. Kit (Megazyme, 2018). This rimeter and subjected to a heating process from 30 ◦ C to 120 ◦ C (heating
procedure consists of separating amylose from amylopectin using rate of 10 ◦ C/min). The thermal properties of the starches were obtained
concanavalin A (Con-A) (Megazyme, 2018). using calorimeter software (TA Data32).

2.4. Morphology 2.9. Fourier-transform infrared (FTIR) analysis

Photomicrographs of the quinoa and kiwicha starch granules were Starch samples were placed directly into a Vertex 70 FT-IR

2
F.F. Velásquez-Barreto et al. Food Hydrocolloids 120 (2021) 106883

spectrophotometer (Bruker Optik GmbH., Ettlingen, Germany), and Table 1


infrared absorption spectra were collected using 40 scans (spectral res­ Chemical composition of Andean grains.
olution of 1 cm− 1) in the wavenumber region from 4000 cm− 1 to 400 Parameter (%) Kiwicha White Quinoa Red Quinoa Black
cm− 1. The data were recorded using equipment software (Opus v. 7.0). Quinoa
The molecular order for the quinoa and kiwicha starches was evaluated Humidity 11.54 ± 12.09 ± 0.05b 12.34 ± 11.56 ±
by relating the absorbances at wavenumbers of 1047 cm− 1, 1022 cm− 1, 0.06c 0.11a 0.09c
and 995 cm− 1. The ratios evaluated were R1 (Abs1047/Abs1022) and R2 Protein 11.59 ± 9.66 ± 0.002d 12.81 ± 12.19 ±
(Abs995/Abs1022) (Hoyos-Leyva, Bello-Pérez, Yee-Madeira et al., 2017). 0.31c 0.17a 0.05b
Fats 6.44 ± 0.22a 4.41 ± 0.24c 5.36 ± 0.47b 4.40 ± 0.14c
Fiber 2.48 ± 0.08b 2.66 ± 0.32b 2.66 ± 0.07b 10.07 ±
2.10. Molecular weight 0.22a
Ash 2.01 ± 0.03b 1.82 ± 0.014c 1.90 ± 0.09b, 2.16 ± 0.10a
Starch (0.5 g) was solubilized with 20 mL of 95% DMSO; the solution c
Carbohydrates 65.94 ± 69.36 ± 64.93 ± 59.62 ±
was stirred for 4 days at 4 ◦ C. Then, 180 mL of 95% ethanol (4 ◦ C) was
0.04b 0.007a 0.35b 0.04c
added, and the resulting dispersion was stored overnight at room tem­
perature. Subsequently, the dispersion was centrifuged at 13000 rpm for Different letters in the same row indicate that exist significant difference (p <
10 min. The pellet was resuspended in 60 mL of ethanol. The procedure 0.05).
Values presented as mean ± standard deviation.
was carried out in triplicate. Finally, ethanol, acetone, and ethyl ether in
the correlative order were used to wash the sediment. The residue was
dried under room conditions and stored in a glass container. The mo­ digestibility characteristics. The highest fat content was presented by
lecular properties of the amylose and amylopectin molecules in the kiwicha grains. Similar results for chemical composition were reported
starch samples were determined with the procedure described by by Ballester-Sánchez, Gil, Fernández-Espinar, and Haros (2019) for
Hoyos-Leyva, Bello-Pérez, Alvarez-Ramirez, and Agama-Acevedo quinoa grains and by Martinez-Lopez et al. (2020) for kiwicha grains.
(2017). An AF4 chromatographic system (Wyatt Technology Corpora­ However, Wright, Huber, Fairbanks, and Huber (2002) reported higher
tion, Santa Barbara, USA) coupled to a multiangle laser light values of protein content (14.8–15.7%) and fiber (8.8–10.3%) in vari­
scattering-MALLS (Dawn Heleos 8, Wyatt Technology Corporation, eties of bitter quinoa grains. Additionally, Srichuwong et al. (2017) re­
Santa Barbara, USA) system and a refractive index detector-RID (1100 ported higher values for fiber content for kiwicha starches. Likewise,
Generic RI, Agilent Technologies, Santa Clara, CA, USA) was used to higher ash contents in quinoa grains (2.1–3.1%) were reported in pre­
determine these properties. The flow rate was 1 mL/min, with an in­ vious studies (Ballester-Sánchez et al., 2019; Steffolani, León, & Pérez,
jection time of 1 min and an elution time of 32 min. Water with 0.02% 2013; Wright et al., 2002).
sodium azide was used as the elutant. The average molecular weight
(Mw) and radius of gyration (Rg) were calculated using the Berry method 3.2. Chemical composition of starch
(Hoyos-Leyva, Bello-Pérez, Alvarez-Ramirez, & Agama-Acevedo, 2017)
and ASTRA® Version 5.3.1.5 software (Wyatt Technology Corporation, The starches of Andean grains exhibited low fat, protein, fiber, and
Santa Barbara, USA). ash contents (Table 2), which indicates that the starch extraction process
using NaOH was efficient and reduced the content of the nonstarch
2.11. In vitro starch digestibility fractions that were present in the whole grains (Table 1). The phos­
phorus content was higher in black quinoa grains, followed by red
The in vitro digestibility of starch was determined by the method quinoa, kiwicha, and white quinoa. Zhu and Cui (2019) showed that
described by Quintero-Castaño, Castellanos-Galeano, Álvarez-Barreto, phosphorus in starch granules may influence the pasting and rheological
Bello-Pérez, and Alvarez-Ramirez (2020). The simulated intestinal properties of pastes. White quinoa starch presented the higher fiber
digestion was carried out with a mixture of pancreatin (approximately 2 content. Jan, Panesar, and Singh (2017) and Steffolani et al. (2013)
mg mL− 1) and amyloglucosidase (approximately 28 U.mL− 1) in distilled reported fiber content values of 0.13% and 0.19–0.24% for quinoa
water. Samples were taken at increasing sampling times. Approximately starches. This difference in fiber content may be due to the starch
300 μL of a stop solution (0.3 mol. L− 1 Na2CO3) was added after the final extraction method since the extraction was more effective in starches
sampling time to avoid amylase activity. A D-glucose assay kit (GODPOD
format from Megazyme International, Bray, Ireland) was used to quan­ Table 2
tify the released sugars in the hydrolysis reaction. The results were re­ Chemical composition of Andean grain starches.
ported relative to the total starch content of the sample.
Parameter Kiwicha White Red Quinoa Black
Quinoa Quinoa
2.12. Statistical analysis
Humidity (%) 10.68 ± 11.93 ± 12.79 ± 13.15 ±
0.13d 0.08c 0.01b 0.15a
The properties of the different starch samples were compared using Protein (%) 1.03 ± 0.00c 0.86 ± 1.34 ± 0.06a 1.21 ±
the Tukey test and Microsoft Excel with the Real Statistics Resource Pack 0.000d 0.000b
2020. Data were obtained in triplicate. Fat (%) 0.14 ± 0.10a, 0.23 ± 0.17a 0.07 ± 00 ± 00c
b 0.03b
Fiber (%) 0.19 ± 0.10b 0.58 ± 0.01a 0.17 ± 0.04 ±
3. Results and discussion 0.14b 0.01b
Ash (%) 0.26 ± 0.04b 0.35 ± 0.56 ± 0.04a 0.52 ± 0.01a
3.1. Chemical composition of whole Andean grains 0.05b
Carbohydrate 87.70 ± 86.05 ± 85.07 ± 85.09 ±
(%) 0.08a 0.27b 0.23c 0.19c
The grains of quinoa and kiwicha presented similar moisture con­ Phosphorous (%) 0.09 ± 0.05a 0.05 ± 0.09 ± 0.12 ± 0.04a
tents (Table 1). The four grains showed high protein content, with red 0.01b 0.003a
quinoa presenting the highest value. Likewise, the black quinoa grains Amylose (%) 10.05 ± 19.05 ± 13.01 ± 13.08 ±
had the highest fiber content. Protein and fiber are important during the 2.38b 2.33a 1.77b 0.69b
starch isolation step because isolation is difficult and affects the purity of Different letters in the same row indicate that exist significant difference (p <
the polysaccharide. Residues of protein and fiber can be present in iso­ 0.05).
lated starch and modify the functional, physicochemical and Values presented as mean ± standard deviation.

3
F.F. Velásquez-Barreto et al. Food Hydrocolloids 120 (2021) 106883

with lower fiber content, such as black quinoa, red quinoa, and kiwicha et al. (2020) and Contreras-Jiménez, Torres-Vargas, and Rodrí­
starches. The protein content of Andean grain starches was in the range guez-García (2019). The SEM images showed that starch extraction from
of recent reports (0.46–1.78%) (Ballester-Sánchez et al., 2019; Jan, quinoa and kiwicha was adequate because oval clusters that contain
Panesar, & Singh, 2017; Steffolani et al., 2013; Wright et al., 2002). The starch granules within the endosperm are not present (Con­
fat content of white quinoa starches was similar to the values reported treras-Jiménez et al., 2019; Kierulf et al., 2020). However, protein and
by Jan, Panesar, and Singh (2017) and Steffolani et al. (2013) fat residues were not completely removed (Table 2).
(0.40–2.56%). However, the fat content showed important differences The black quinoa starch granules had the smallest size, followed by
in kiwicha, red quinoa, and black quinoa starches. Quinoa starches red quinoa, white quinoa, and kiwicha starches (Fig. 2). The equivalent
exhibited a higher ash content than kiwicha starches. Likewise, Jan, diameter range was in the range 0.675–4.03 μm for kiwicha starch
Panesar, and Singh (2017) and Steffolani et al. (2013) reported ash granules, 0.675–3.55 μm for white quinoa, 0.461–2.75 μm for red
contents of 0.22–0.64% for quinoa starches. Recently, Kierulf et al. quinoa, and 0.594–3.12 μm for black quinoa. The average equivalent
(2020) studied the effect of successive extractions on the protein and fat diameter was higher for kiwicha starch granules (1.85 μm), followed by
content of amaranth and quinoa starch and reported that two extractions white quinoa (1.71 μm), red quinoa (1.55 μm), and black quinoa (1.29
with NaOH reduced the protein content from 2.43% to 0.67% for μm). Jan, Panesar, and Singh (2017) reported quinoa starch granule
amaranth starch and from 2.70% to 1.2% for quinoa starch. Similarly, sizes in the range of 0.77–4.05 μm and an average granule size of 1.30
the fat content was reduced from 2.2% to 0.2% for amaranth starch and
from 0.6% to 0.3% for kiwicha starch.
Quinoa starches exhibited a higher amylose content (Table 2).
Likewise, the amylose content of the white quinoa starch presented
differences from the starch obtained from the red and black varieties.
The amylose content of quinoa starches (13.01–19.05%) was within the
range (3.5–23.9%) reported in the literature (Ballester-Sánchez et al.,
2019; Jan, Panesar, & Singh, 2017; Lindeboom, Chang, Tyler, & Chib­
bar, 2005; Steffolani et al., 2013; Tang, Watanabe, & Mitsunaga, 2002;
Wright et al., 2002). The amylose content of kiwicha starch was higher
than the values reported by Srichuwong et al. (2017) (approximately
1.2%) and within the range reported by Singh et al. (2014) (4.7%–
12.5%). These differences in the amylose content in quinoa and kiwicha
starches can be attributed to the method, variety, maturity, and culture
conditions (Wright et al., 2002). However, the higher amylopectin
content and small granule size suggest that quinoa and kiwicha starches
can be used for nanocrystal production (Sanchez de la Concha et al.,
2018).

3.3. Morphology and size distribution

The quinoa and kiwicha starches exhibited a polygonal shape with a


Fig. 2. Size distribution of kiwicha and quinoa starch granules.
rough surface (Fig. 1). Similar morphologies were reported by Jiang

Fig. 1. SEM morphology of kiwicha (a), white quinoa (b), red quinoa (c), and black quinoa (d) starches.

4
F.F. Velásquez-Barreto et al. Food Hydrocolloids 120 (2021) 106883

μm. Steffolani et al. (2013) found quinoa starch granule sizes of 1 temperatures above 50 ◦ C. Differences between the SP of quinoa and
μm–3.5 μm and an average granule size of 2.53 μm. On the other hand, kiwicha starch granules may be due to the crystallinity structure,
Bet et al. (2018) and Kierulf et al. (2020) reported sizes of 1.2 μm and amylose content, and amylopectin chain size distribution. Higher values
1.37 μm for kiwicha starches, which are smaller than those obtained in of SP (17–28.15 g/g) were determined by Wu, Morris, and Murphy
this work. Some visible differences in the size of the quinoa and kiwicha (2017), Jiang et al. (2020) and Li et al. (2016) for quinoa starch.
starch granules were found in other studies, which can be attributed to However, Jan, Panesar, Rana, and Singh (2017) reported SP values of
harvest time, growing conditions, environmental conditions and genetic 1.72–13.89 g/g for quinoa starch at temperatures ranging from 55 to
factors (Qu et al., 2018). 95 ◦ C, which are close to those obtained in this study. Kiwicha starches
presented a higher SP in comparison with the SP values (5–11.5%) re­
3.4. Solubility and swelling power ported by Fonseca-Florido, Méndez-Montealvo, Velazquez, and
Gómez-Aldapa (2016). The low values for the swelling power of quinoa
Starch solubility and swelling power (SP) provide valuable insights starch granules can be attributed to the entanglement between amylose
into the interactions between starch chains within crystalline and and amylopectin, which could be reinforced by the presence of phos­
amorphous zones (Bello-Pérez, Contreras-Ramos, Jìmenez-Aparicio, & pholipids by forming helical complexes with amylose and long amylo­
Paredes-López, 2000; Jiang et al., 2020). The solubility of quinoa and pectin chains. In turn, chain entanglement hampers the swelling of the
kiwicha starches presented a similar pattern (Table 3). Kiwicha starch granules (Srichuwong & Jane, 2007).
showed higher solubility than white, red, and black quinoa starches.
This solubility difference may be linked to differences in structural and
3.5. Pasting
molecular properties. The solubility of quinoa and kiwicha starches
increased at temperatures above 50 ◦ C, which allowed for the weak­
The pasting behavior was similar for the three quinoa starches.
ening of the hydrogen bonds between amylose and amylopectin chains.
However, the kiwicha starch exhibited a different pasting pattern
As a consequence, the water transport into the granules is improved, and
(Fig. 3). Kiwicha starches showed a more pronounced increase and
the amylose chains are leached out. Jan, Panesar, Rana, and Singh
decrease in viscosity during the heating and cooling phases than quinoa
(2017) reported solubility values of 1.32–7.11% at temperatures of
starches. Differences in solubility and swelling of the starch granules
55–95 ◦ C for quinoa starches. Jiang et al. (2020) reported solubility
could be linked to the differences in the pasting patterns. Similar
values of 6.45–17.57% at 95 ◦ C, whereas Li, Wang, and Zhu (2016)
behavior for pasting in quinoa (Contreras-Jiménez et al., 2019; Tiga,
found solubility values of 0.4–6.4% at 55 ◦ C and 41.4–86.9% at 95 ◦ C for
Kumcuoglu, Vatansever, & Tavman, 2021) and kiwicha starch (Srichu­
quinoa starches. The disparities in the solubility values obtained in this
wong et al., 2017) was reported in recent studies. The pasting properties
study relative to previous reports could be related to the quinoa varieties
of kiwicha and quinoa starch are summarized in Table 4. The highest
and harvesting conditions. On the other hand, the relatively low solu­
peak viscosity (PV) was presented by the red quinoa starch pastes, fol­
bility percentages for quinoa starches are due to the low amylose content
lowed by kiwicha starch, white quinoa, and black quinoa starch. The
and the different lengths of the amylopectin chains, which can restrict
time to reach the PV was longer in quinoa starches due to their low
the swelling of the starch and reinforce the internal structure (Bel­
swelling power (Table 3). The breakdown viscosity (BDV) of the kiwicha
lo-Pérez et al., 2000; Tang et al., 2002).
starch granules was higher than the value obtained for quinoa starches,
The SP of the quinoa and kiwicha starches increased with tempera­
which is a consequence of the higher solubility (Table 3). The highest
ture. The SP of kiwicha starches at temperatures of 40 ◦ C–90 ◦ C was
final viscosity (FV) was found for quinoa starches, probably due to the
higher than that of quinoa starches. The SP of the white, red and black
presence of granules or small disintegrated granules in the starch paste.
quinoa starches had similar tendencies, with higher values for
In the cooling phase, the black and white quinoa starch granules showed
a higher setback viscosity (SBV), possibly due to the more significant
Table 3
interaction between solubilized amylose and amylopectin molecules, as
Solubility and swelling power of Andean grain starches.
well as disintegrated starch granules compared to the red quinoa and
Temperature Kiwicha White Red Quinoa Black kiwicha starches. Higher values of PV, minimum viscosity (MV) and FV
(◦ C) Quinoa Quinoa
for quinoa starch pastes were reported by Ballester-Sánchez et al.
Solubility (%) (2019), Jan, Panesar, Rana, and Singh (2017) and Steffolani et al.
40 3.33 ± 1.15a 2.65 ± 0.28b 2.12 ± 0.11b 2.24 ± 0.14b (2013), and lower values for these properties were reported by Wright
50 5.33 ± 1.16a 3.27 ± 0.31b 2.47 ± 0.12c 2.67 ± 0.23c
60 48.67 ± 4.17 ± 0.32b 3.73 ± 0.12b 3.87 ± 0.11b
3.06a
70 71.33 ± 5.75 ± 5.77 ± 0.15c 6.33 ± 0.12b
1.16a 0.56b,c
80 79.33 ± 8.39 ± 0.02b 7.05 ± 0.22d 7.83 ± 0.25c
1.15a
90 80.46 ± 11.2 ± 0.87b 8.35 ± 0.16d 9.14 ± 0.41c
2.16a
Swelling Power (g/g)
40 4.42 ± 0.09a 3.24 ± 0.06b 3.48 ± 0.33b 3.36 ± 0.24b
50 4.69 ± 0.34b 5.49 ± 0.03a 5.13 ± 3.88 ± 0.02c
0.26a,b
60 19.76 ± 8.66 ± 0.05b 7.11 ± 0.47c 5.46 ±
1.05a 0.07d
70 21.30 ± 10.31 ± 9.54 ± 0.18c 9.31 ± 0.06c
0.84a 0.12b
80 25.76 ± 11.39 ± 12.45 ± 14.31 ±
1.37a 0.69c 0.25c 0.49b
90 39.57 ± 16.92 ± 15.01 ± 14.55 ±
3.66a 0.15b 0.24c 0.39c

Different letters in the same row indicate that exist significant difference (p <
0.05).
Values presented as mean ± standard deviation. Fig. 3. Pasting behavior of kiwicha and quinoa starches.

5
F.F. Velásquez-Barreto et al. Food Hydrocolloids 120 (2021) 106883

Table 4 starch extraction with NaOH. Ballester-Sánchez et al. (2019) and Stef­
Pasting and thermal properties of Andean grain starches. folani et al. (2013) reported gelatinization enthalpies ranging from 12.4
Pasting Properties to 13.5 J/g and 8.45–9 J/g for quinoa starches, respectively, which are
higher than those reported in the present study. However, Aluwi et al.
Starch PT (◦ C) PV (mPa. FV (mPa.s) Setback Breakdown
s) (mPa.s) (mPa.s) (2017) reported lower values for gelatinization enthalpies (3.34–5.16
J/g). On the other hand, a gelatinization enthalpy of 8.67 J/g was re­
Kiwicha 63.56 ± 821.36 548.55 ± 135.16 407.97 ±
ported by Bet et al. (2018) for kiwicha starches, which is similar to the
± ±
0.00b 10.67b 7.12c 1.76c 5.30a
White 60.37 ± 777.01 ± 1014.45 ± 312.78 ± 75.35 ± values shown in Table 4. Srichuwong et al. (2017) reported higher
Quinoa 0.00c 10.19c 13.30a 4.10a 0.99b enthalpy values (12.4 J/g). These enthalpy differences may be due to the
Red 63.46 ± 848.69 ± 953.98 ± 119.67 ± 14.39 ± starch extraction method, which can affect the internal structure of
Quinoa 0.00b 9.33a 10.48b 1.31d 0.16d
starches in their native state (Singh et al., 2006; Wang & Copeland,
Black 69.74 ± 746.95 ± 935.55 ± 227.47 ± 38.87 ± 0.62c
Quinoa 0.00a 11.93d 14.94b 3.63b 2015).
Thermal Properties
Starch Ti Tg Tc ΔH (J/g)

Kiwicha 60.61 ± 66.38 ± 74.14 ± 8.46 ± 3.7. FTIR analysis


0.25b 0.18b 0.13b 0.15a
White 56.2 ± 62.84 ± 71.79 ± 7.44 ± Starches of kiwicha and quinoa presented similar infrared spectra in
Quinoa 0.06d 0.03c 0.22c 0.07b
the MID-IR region (Fig. 4.a). The wide band at 3900-3300 cm− 1 in
Red 58.45 ± 65.14 ± 74.82 ± 7.73 ±
Quinoa 0.26c 1.37b.c 0.09b.c 0.12b quinoa and kiwicha starches is related to the presence of hydrogen
Black 65.66 ± 70.97 ± 78.44 ± 7.82 ± bonds with intramolecular and intermolecular tension vibrations of
Quinoa 0.09a 0.26a 0.20a 0.02b hydroxyl groups (O–H) (Jan, Panesar, Rana, & Singh, 2017). Pro­
PT (pasting temperature), PV (peak viscosity) and FV (final viscosity). Ti: Initial nounced peaks between 2000 cm− 1 and 800 cm− 1 are characteristic
temperature; Tg: Gelatinization temperature; Tc: conclusion temperature; ΔH: peaks for starches isolated from other sources. Banded peaks in the range
gelatinization temperature. 3000-2800 cm− 1 and 3500-3000 cm− 1 are characteristic of C–H (i.e.,
Different letters in the same column for pasting properties and thermal prop­ aliphatic) and O–H bonds stretching (Hoyos-Leyva, Bello-Pérez,
erties indicate that exist significant difference (p < 0.05). Yee-Madeira et al., 2017). Characteristic peaks at 1641 and 1647 cm− 1
Values presented as mean ± standard deviation. for quinoa and kiwicha starches are attributed to the presence of slightly
bound or absorbed water and can be attributed to H–O–H bending vi­
et al. (2002) in comparison with the data obtained in this study. Like­ brations (Jan, Panesar, Rana, & Singh, 2017). In particular, the finger­
wise, the kiwicha starch pasting properties were lower than those re­ print regions show similar peaks for starch from other plant sources at
ported by Bet et al. (2018), except for the pasting temperature (PT),
which was slightly higher than that in this study (64.32 ◦ C). These dif­
ferences may be due to the starch concentration, amylose contents,
amylopectin chain size, and molecular weight of the amylose and
amylopectin molecules (Singh, Kaur, Sandhu, Kaur, & Nishinari, 2006).

3.6. Thermal properties

The lowest gelatinization temperature was presented by white


quinoa starch, followed by red quinoa, kiwicha and black quinoa starch
(Table 4). Kiwicha starches had gelatinization temperatures (Tg) that
were similar to those for red quinoa starches. The white, red, and black
quinoa starches presented different Tg values, which were similar to the
PT obtained from the pasting properties. The differences in the Tg of the
quinoa starches are due to the internal structure of the starch granules
and the amylose content. Differences in the gelatinization temperature
Tg have been reported for different varieties of quinoa starches. Bal­
lester-Sánchez et al. (2019) reported Tg values in the range
58.66–61.1 ◦ C, while Steffolani et al. (2013) documented values of
61.66–63.01 ◦ C and Aluwi, Murphy, and Ganjyal (2017) reported values
ranging from 68.8 to 74 ◦ C for quinoa starches. Bet et al. (2018) reported
a Tg value of 66.27 ◦ C for kiwicha starch, which is similar to that ob­
tained in this study. The differences between the enthalpies reported by
other authors and the results in this study can be attributed to the
different varieties of quinoa, starch extraction methods, and harvesting
conditions (Wang & Copeland, 2015).
The gelatinization enthalpy of the kiwicha starch was higher than the
gelatinization enthalpy of the quinoa starches from the three varieties.
Different plant species can present different enthalpies of gelatinization
because genetic factors can influence the internal architecture of
amylose and amylopectin within the starch granule. The white, red and
black quinoa starches exhibited similar gelatinization enthalpies, which
indicates that the internal structures of amylopectin and amylose were Fig. 4. (a) FTIR spectra for kiwicha and quinoa starches. (b) FTIR 995/1022
similar in the three varieties. However, the enthalpies can be affected by and 1047/1022 ratios were obtained by deconvolution of the 1080-980 cm− 1
the quinoa starch extraction process. Ballester-Sánchez et al. (2019) region. Letters over the same column pattern denote significant (p < 0.05)
pointed out that some modifications of starch can occur during quinoa differences.

6
F.F. Velásquez-Barreto et al. Food Hydrocolloids 120 (2021) 106883

1155 cm− 1, 1080 cm− 1, 1021 cm− 1 and 930 cm− 1, which are related to Paredes-López, 1998; Fu, Luo, BeMiller, Liu, & Liu, 2015). The Rg of
C–O–C and C–O–H tension vibrations of glycosidic bonds that occur amylopectin in red quinoa starches was higher than that in kiwicha,
between amylose and amylopectin. The peak at 1638 cm− 1 is possibly white quinoa, and black quinoa starches, which may indicate that such
related to the presence of bound water in amylose and amylopectin (Hui, starch had a more compact structure. Moreover, the amylose of white
Qi-he, Ming-liang, Qiong, & Guo-qing, 2009). quinoa and red quinoa presented a higher Rg than kiwicha and black
The short molecular order for the starches was evaluated at typical quinoa amylose. The amylopectin in starch from red quinoa and black
absorbances at 1047 cm− 1 and 1022 cm− 1, where the first band is quinoa exhibited a higher dispersion (Mn/Mw > 1) than the amylopectin
related to the crystalline structure and the second band to the amor­ in kiwicha and white quinoa starch. The dispersion of the starch amylose
phous regions of the starch granules. The absorbance band at 995 cm− 1 in the three varieties of quinoa was similar, which indicates that the
is related to water-starch interactions (Hoyos-Leyva, Bello-Pérez, amylose molecules were more uniform than the amylopectin molecules
Yee-Madeira et al., 2017; van Soest, Tournois, de Wit, & Vliegenthart, due to the branched structure of amylopectin (Çőrüşlü & Pekin, 1983).
1995). Fig. 4.b shows the absorbance ratios at 1047/1022 and 995/1022
for the different starches, which reflect the molecular order of the starch
chains. The 995/1022 ratio showed only slight variations in the range 3.9. In vitro starch digestibility
0.92–0.97, meaning that the starch samples contain similar levels of
hydrated structures relative to disordered structures. Values for the The enzymatic hydrolysis of the kiwicha and quinoa starches is
absorption ratio of 1.19–1.22 have been reported for mashua, oca and presented in Fig. 5.a. In general, only slight differences were exhibited
olluco starches (Velásquez-Barreto et al., 2019, 2020), from 1.02 to 1.11 for the different starch samples. Interestingly, the hydrolysis rate of
for taro starches (Hoyos-Leyva, Bello-Pérez, Yee-Madeira et al., 2017),
and from 1.11 to 1.19 for wheat starches (Karwasra, Gill, & Kaur, 2017).
The 1047/1022 ratio exhibited more visible differences, mainly between
kiwicha and quinoa starches. In fact, the kiwicha starch exhibited a
value of approximately 0.59, whereas the quinoa starches showed values
in the range of 0.67–0.77. The highest ratio was shown by the red quinoa
starch. This means that the crystalline regions of quinoa starches are
more ordered than the crystalline regions of kiwicha starch (Warren,
Gidley, & Flanagan, 2016). Similar values for this ratio have been re­
ported by Velásquez-Barreto et al. (2020) for mashua, oca, and olluco
starches (0.68–0.71) and by Hoyos-Leyva, Bello-Pérez, Yee-Madeira
et al. (2017) for taro starches (0.68–0.69). Wheat starch exhibits
higher values in the range of 0.83–0.87 (Karwasra et al., 2017).

3.8. Molecular weight

The molecular weight (Mw) of amylose and amylopectin chains is a


key property with marked influence on the functional properties, paste
formation and rheological properties of starch during processing (Cor­
nejo-Ramírez et al., 2018; Yang, Xia, Junejo, & Zhou, 2018). The Mw of
amylopectin from quinoa and kiwicha starches was higher than the Mw
of amylose (Table 5), and this behavior is similar to that found for other
sources of starch (Hoyos-Leyva, Bello-Pérez, Alvarez-Ramirez, &
Agama-Acevedo, 2017; Yoo & Jane, 2002). Such similarity can be
attributed to the fact that amylopectin contains higher glucose units
linked to the branched structure. Red and black quinoa starches had a
higher amylopectin molecular weight than kiwicha and white quinoa
starches. The Mw of amylose in the quinoa starches of the three varieties
was higher than that of kiwicha starch. This behavior is related to the
pasting properties of the starches, mainly to the final viscosities
(Table 4). For instance, the quinoa starch that presented the highest Mw
also exhibited a higher FV.
Fig. 5. (a) Hydrolysis kinetics of kiwicha and quinoa starches. (b) log-of-slope
A low value for the gyration radius (Rg) indicates that the structure is
(LOS) plot of the hydrolysis kinetics. The dotted lines denote two different (fast
more compact and is related to differences in the branching of amylose
and slow) time scales.
and amylopectin molecules (Bello-Pérez, Colonna, Roger, &

Table 5
Molecular weight and radius of gyration of kiwicha and quinoa starches.
Starch Amylopectin Amylose

Molecular Weight (g/mol x 108) Mn/Mw Rg Molecular Weight (g/mol x 107) Mn/Mw Rg

Kiwicha 2.11 ± 0.01b 1.10 ± 0.001c 240.3 ± 0.00b 1.58 ± 0.02d 1.31 ± 0.004a 127.15 ± 0.35c
White Quinoa 1.94 ± 0.02c 1.34 ± 0.00b 239.05 ± 0.92b 3.44 ± 0.08b 1.01 ± 0.00b 182.50 ± 0.85a
Red Quinoa 2.34 ± 0.01a 1.73 ± 0.04a 272.65 ± 3.32a 6.55 ± 0.02a 1.01 ± 0.001b 164.6 ± 0.42b
Black Quinoa 2.37 ± 0.07a 1.11 ± 0.007c 211.6 ± 2.12c 1.94 ± 0.01c 1.11 ± 0.014b 105.65 ± 0.35d

Rg: gyration radius; Mw: molecular weight: Mn: number average molecular weight.
Different letters in the same column indicate that exist significant difference (p < 0.05).
Values presented as mean ± standard deviation.

7
F.F. Velásquez-Barreto et al. Food Hydrocolloids 120 (2021) 106883

kiwicha starch was similar to that of quinoa starch. Fig. 5.b presents the Alemayehu, F. R., Bendevis, M. A., & Jacobsen, S.-E. (2015). The potential for utilizing
the seed crop amaranth (Amaranthus spp.) in east africa as an alternative crop to
corresponding log-of-slope (LOS) plot for the hydrolysis kinetics (Patel,
support food security and climate change mitigation. Journal of Agronomy and Crop
Day, Butterworth, & Ellis, 2014). Two hydrolysis phases in terms of the Science, 201(5), 321–329. https://doi.org/10.1111/jac.12108
slope of the plot can be observed. The first phase reflects fast hydrolysis Aluwi, N. A., Murphy, K. M., & Ganjyal, G. M. (2017). Physicochemical characterization
for times up to approximately 20 min. The second phase is linked to of different varieties of quinoa. Cereal Chemistry Journal, 94(5), 847–856. https://
doi.org/10.1094/CCHEM-10-16-0251-R
slower hydrolysis, as reflected by the smaller slope of the LOS plot. On AOAC. (2005). Official methods of Analysis of AOAC international. In association of official
average, for the four cases, the fast and slow hydrolysis rate constants (i. Analysis chemists international.
e., the slope of the LOS plot) were estimated to be 0.11 min− 1 and Ballester-Sánchez, J., Gil, J. V., Fernández-Espinar, M. T., & Haros, C. M. (2019). Quinoa
wet-milling: Effect of steeping conditions on starch recovery and quality. Food
0.0024 min− 1, respectively. Patel et al. (2014) reported values of Hydrocolloids, 89, 837–843. https://doi.org/10.1016/j.foodhyd.2018.11.053
approximately 0.04–0.07 min− 1 for the fast phase and 0.006–0.012 Bashir, K., & Aggarwal, M. (2019). Physicochemical, structural and functional properties
min− 1 for the slow phase for the different starches (e.g., maize, potato, of native and irradiated starch: A review. Journal of Food Science & Technology, 56(2),
513–523. https://doi.org/10.1007/s13197-018-3530-2
rice and wheat). The results for the Andean grains showed that the initial Bello-Pérez, L. A., Colonna, P., Roger, P., & Paredes-López, O. (1998). Macromolecular
hydrolysis phase was approximately 40–50 times faster than the hy­ features of amaranth starch. Cereal Chemistry Journal, 75(4), 395–402. https://doi.
drolysis phase at longer times. In terms of Englyst’s classification for the org/10.1094/CCHEM.1998.75.4.395
Bello-Pérez, L. A., Contreras-Ramos, S. M., Jìmenez-Aparicio, A., & Paredes-López, O.
digestible starch fractions, the rapidly digestible starch (RDS) fraction (2000). Acetylation and characterization of banana (Musa paradisiaca) starch. Acta
was approximately 29.5%, and the slowly digestible starch (SDS) frac­ Cientifica Venezolana, 51(3), 143–149. http://www.ncbi.nlm.nih.gov/pubmed
tion was approximately 26.5%. The differences exhibited by the kiwicha /11265448.
Bet, C. D., de Oliveira, C. S., Colman, T. A. D., Marinho, M. T., Lacerda, L. G.,
and quinoa starches were not statistically significant (p > 0.05).
Ramos, A. P., et al. (2018). Organic amaranth starch: A study of its technological
properties after heat-moisture treatment. Food Chemistry, 264, 435–442. https://doi.
4. Conclusions org/10.1016/j.foodchem.2018.05.021
Contreras-Jiménez, B., Torres-Vargas, O. L., & Rodríguez-García, M. E. (2019).
Physicochemical characterization of quinoa (Chenopodium quinoa) flour and isolated
Quinoa and kiwicha grains and starches showed different chemical starch. Food Chemistry, 298. https://doi.org/10.1016/j.foodchem.2019.124982,
compositions. Similar polygonal shapes were exhibited by quinoa and 124982.
Cornejo-Ramírez, Y. I., Martínez-Cruz, O., Del Toro-Sánchez, C. L., Wong-Corral, F. J.,
kiwicha starches, whereas kiwicha starches were larger in size than Borboa-Flores, J., & Cinco-Moroyoqui, F. J. (2018). The structural characteristics of
quinoa starches. Kiwicha starches had a higher swelling power and starches and their functional properties. CyTA - Journal of Food, 16(1), 1003–1017.
solubility at different temperatures than quinoa starches. Kiwicha starch https://doi.org/10.1080/19476337.2018.1518343
Cruz-Tirado, J. P., Vejarano, R., Tapia-Blácido, D. R., Barraza-Jáuregui, G., & Siche, R.
showed different pasting properties and gelatinization enthalpies than
(2019). Biodegradable foam tray based on starches isolated from different Peruvian
quinoa starches. Quinoa and kiwicha starches showed similar absor­ species. International Journal of Biological Macromolecules, 125, 800–807. https://doi.
bance ratios at 995/1022, which indicates that these starches have org/10.1016/j.ijbiomac.2018.12.111
similar hydrated molecular structures. In contrast, quinoa starches Cruz, G., Ribotta, P., Ferrero, C., & Iturriaga, L. (2016). Physicochemical and rheological
characterization of andean tuber starches: Potato (Solanum tuberosum ssp.
exhibited higher values for the 1047/1022 ratio, indicating a higher andigenum), oca (Oxalis tuberosa molina) and papalisa (Ullucus tuberosus caldas).
organization of crystalline structures in quinoa starches. Red and black Starch - Stärke, 68(11–12), 1084–1094. https://doi.org/10.1002/star.201600103
quinoa starches had a higher amylopectin molecular weight than kiwi­ Çőrüşlü, E., & Pekin, B. (1983). Starch hydrolysates with very low degrees of
polydispersity. Starch - Stärke, 35(3), 98–100. https://doi.org/10.1002/
cha and white quinoa starches. The amylopectin in red quinoa starches star.19830350309
presented a more compact structure than in kiwicha, white quinoa and Dakhili, S., Abdolalizadeh, L., Hosseini, S. M., Shojaee-Aliabadi, S., & Mirmoghtadaie, L.
black quinoa starches. The amylopectin in starch from red quinoa and (2019). Quinoa protein: Composition, structure and functional properties. Food
Chemistry, 299. https://doi.org/10.1016/j.foodchem.2019.125161, 125161.
black quinoa exhibited a higher dispersion than amylopectin in kiwicha Das, S. (2016). Amaranthus: A promising crop of future. In Amaranthus: A promising crop of
and white quinoa starch. Future studies should be oriented to clarify the future. Springer Singapore. https://doi.org/10.1007/978-981-10-1469-7
relationship between the structural characteristics, functional proper­ Fonseca-Florido, H. A., Méndez-Montealvo, G., Velazquez, G., & Gómez-Aldapa, C. A.
(2016). Thermal study in the interactions of starches blends: Amaranth and achira.
ties, and digestibility of quinoa and kiwicha starches. Food Hydrocolloids, 61, 640–648. https://doi.org/10.1016/j.foodhyd.2016.06.027
Fu, Z., Luo, S.-J., BeMiller, J. N., Liu, W., & Liu, C.-M. (2015). Effect of high-speed jet on
Credit author statement flow behavior, retrogradation, and molecular weight of rice starch. Carbohydrate
Polymers, 133, 61–66. https://doi.org/10.1016/j.carbpol.2015.07.006
Hoover, R. (2001). Composition, molecular structure, and physicochemical properties of
Frank F. Velásquez-Barreto, Hubert Arteaga Miñano, J. Alvarez- tuber and root starches: A review. Carbohydrate Polymers, 45(3), 253–267. https://
Ramirez and Luis. A. Bello-Pérez: Conceptualization, Methodology, doi.org/10.1016/S0144-8617(00)00260-5
Hoyos-Leyva, J. D., Bello-Pérez, L. A., Alvarez-Ramirez, J., & Agama-Acevedo, E. (2017).
Validation, Formal Analysis, Formal Analysis, Writing - Original Draft, Structural characterization of aroid starches by means of chromatographic
And Writing - Review & Editing. techniques. Food Hydrocolloids, 69, 97–102. https://doi.org/10.1016/j.
foodhyd.2017.01.034
Hoyos-Leyva, J. D., Bello-Pérez, L. A., Yee-Madeira, H., Rodríguez-García, M. E., &
Declaration of competing interest Aguirre-cruz, A. (2017). Characterization of the flour and starch of aroid cultivars
grown in Mexico. Starch - Stärke, 69(9–10), 1600370.
Hui, R., Qi-he, C., Ming-liang, F., Qiong, X., & Guo-qing, H. (2009). Preparation and
The authors declare that they have no known competing financial properties of octenyl succinic anhydride modified potato starch. Food Chemistry, 114
interests or personal relationships that could have appeared to influence (1), 81–86. https://doi.org/10.1016/j.foodchem.2008.09.019
Jan, K. N., Panesar, P. S., Rana, J. C., & Singh, S. (2017). Structural, thermal and
the work reported in this paper. rheological properties of starches isolated from Indian quinoa varieties. International
Journal of Biological Macromolecules, 102, 315–322. https://doi.org/10.1016/j.
ijbiomac.2017.04.027
Acknowledgments
Jan, K. N., Panesar, P. S., & Singh, S. (2017). Process standardization for isolation of
quinoa starch and its characterization in comparison with other starches. Journal of
To Universidad Nacional Autonoma de Chota, Cajamarca, Peru, for Food Measurement and Characterization, 11(4), 1919–1927. https://doi.org/10.1007/
financing this research with CANON funds. s11694-017-9574-6
Jiang, F., Du, C., Guo, Y., Fu, J., Jiang, W., & Du, S. (2020). Physicochemical and
structural properties of starches isolated from quinoa varieties. Food Hydrocolloids,
References 101. https://doi.org/10.1016/j.foodhyd.2019.105515, 105515.
Karwasra, B. L., Gill, B. S., & Kaur, M. (2017). Rheological and structural properties of
starches from different Indian wheat cultivars and their relationships. International
AACC, American Association of Cereal Chemists. (2000). Approved methods of the AACC
Journal of Food Properties, 20(sup1), S1093–S1106. https://doi.org/10.1080/
(10th Methods). St Paul, MN: AACC.
10942912.2017.1328439
Ai, Y., Hasjim, J., & Jane, J. (2013). Effects of lipids on enzymatic hydrolysis and
Kierulf, A., Whaley, J., Liu, W., Enayati, M., Tan, C., Perez-Herrera, M., et al. (2020).
physical properties of starch. Carbohydrate Polymers, 92(1), 120–127. https://doi.
Protein content of amaranth and quinoa starch plays a key role in their ability as
org/10.1016/j.carbpol.2012.08.092

8
F.F. Velásquez-Barreto et al. Food Hydrocolloids 120 (2021) 106883

Pickering emulsifiers. Food Chemistry, 315. https://doi.org/10.1016/j. constituents. Food Chemistry, 233, 1–10. https://doi.org/10.1016/j.
foodchem.2020.126246, 126246. foodchem.2017.04.019
Kyung, J.-S., & Yoo, B. (2014). Rheological properties of azuki bean starch pastes in Srichuwong, S., & Jane, J. L. (2007). Physicochemical properties of starch affected by
steady and dynamic shear. Starch - Stärke, 66(9–10), 802–808. https://doi.org/ molecular composition and structures: A review. Food Science and Biotechnology, 16
10.1002/star.201400024 (5), 663–674.
Lindeboom, N., Chang, P. R., Tyler, R. T., & Chibbar, R. N. (2005). Granule-bound starch Steffolani, M. E., León, A. E., & Pérez, G. T. (2013). Study of the physicochemical and
synthase I (GBSSI) in quinoa (Chenopodium quinoa willd.) and its relationship to functional characterization of quinoa and kañiwa starches. Starch - Stärke, 65
amylose content. Cereal Chemistry Journal, 82(3), 246–250. https://doi.org/ (11–12), 976–983. https://doi.org/10.1002/star.201200286
10.1094/CC-82-0246 Tacer-Caba, Z., & Nilufer-Erdil, D. (2019). Resistant starch. In encyclopedia of food
Lin, L., Yang, H., Chi, C., & Ma, X. (2020). Effect of protein types on structure and chemistry (pp. 571–575). Elsevier. https://doi.org/10.1016/B978-0-08-100596-
digestibility of starch-protein-lipids complexes. Lebensmittel-Wissenschaft & 5.22407-4
Technologie, 134, 110175. https://doi.org/10.1016/j.lwt.2020.110175 Tang, H., Watanabe, K., & Mitsunaga, T. (2002). Characterization of storage starches
Li, G., Wang, S., & Zhu, F. (2016). Physicochemical properties of quinoa starch. from quinoa, barley and adzuki seeds. Carbohydrate Polymers, 49(1), 13–22. https://
Carbohydrate Polymers, 137, 328–338. https://doi.org/10.1016/j. doi.org/10.1016/S0144-8617(01)00292-2
carbpol.2015.10.064 Tiga, B. H., Kumcuoglu, S., Vatansever, M., & Tavman, S. (2021). Thermal and pasting
Li, G., & Zhu, F. (2017). Amylopectin molecular structure in relation to physicochemical properties of Quinoa—wheat flour blends and their effects on production of extruded
properties of quinoa starch. Carbohydrate Polymers, 164, 396–402. https://doi.org/ instant noodles. Journal of Cereal Science, 97. https://doi.org/10.1016/j.
10.1016/j.carbpol.2017.02.014 jcs.2020.103120, 103120.
Martinez-Lopez, A., Millan-Linares, M. C., Rodriguez-Martin, N. M., Millan, F., & Timgren, A., Rayner, M., Dejmek, P., Marku, D., & Sjöö, M. (2013). Emulsion stabilizing
Montserrat-de la Paz, S. (2020). Nutraceutical value of kiwicha (Amaranthus capacity of intact starch granules modified by heat treatment or octenyl succinic
caudatus L.). Journal of Functional Foods, 65. https://doi.org/10.1016/j. anhydride. Food Sciences and Nutrition, 1(2), 157–171. https://doi.org/10.1002/
jff.2019.103735, 103735. fsn3.17
Megazyme. (2018). Amylose/Amylopectin: Assay procedure K-amyl 06/18. In Megazyme Velásquez-Barreto, F. F., Bello-Pérez, L. A., Yee-Madeira, H., & Velezmoro Sánchez, C. E.
data booklet (Vol. 6). (2019). Esterification and characterization of starch from andean tubers. Starch -
Patel, H., Day, R., Butterworth, P. J., & Ellis, P. R. (2014). A mechanistic approach to Stärke, 71(1–2). https://doi.org/10.1002/star.201800101, 1800101.
studies of the possible digestion of retrograded starch by α-amylase revealed using a Velásquez-Barreto, F., & Velezmoro, C. (2018). Rheological and viscoelastic properties of
log of slope (LOS) plot. Carbohydrate Polymers, 113, 182–188. https://doi.org/ Andean tubers starches. Scientia Agropecuaria, 9(2), 189–197. https://doi.org/
10.1016/j.carbpol.2014.06.089 10.17268/sci.agropecu.2018.02.03
Quintero-Castaño, V. D., Castellanos-Galeano, F. J., Álvarez-Barreto, C. I., Bello- Velásquez-Barreto, F. F., Bello-Pérez, L. A., Yee-Madeira, H., Alvarez-Ramirez, J., &
Pérez, L. A., & Alvarez-Ramirez, J. (2020). In vitro digestibility of octenyl succinic Velezmoro-Sánchez, C. E. (2020). Effect of the OSA esterification of Oxalis tuberosa
anhydride-starch from the fruit of three Colombian Musa. Food Hydrocolloids, 101. starch on the physicochemical, molecular, and emulsification properties. Starch -
https://doi.org/10.1016/j.foodhyd.2019.105566, 105566. Stärke, 72(5–6). https://doi.org/10.1002/star.201900305, 1900305.
Qu, J., Xu, S., Zhang, Z., Chen, G., Zhong, Y., Liu, L., et al. (2018). Evolutionary, Wang, S., & Copeland, L. (2015). Effect of acid hydrolysis on starch structure and
structural and expression analysis of core genes involved in starch synthesis. functionality: A review. Critical Reviews in Food Science and Nutrition, 55(8),
Scientific Reports, 8(1), 1–16. https://doi.org/10.1038/s41598-018-30411-y 1081–1097. https://doi.org/10.1080/10408398.2012.684551
Ruiz, K. B., Biondi, S., Oses, R., Acuña-Rodríguez, I. S., Antognoni, F., Martinez- Wang, S., & Zhu, F. (2016). Formulation and quality attributes of quinoa food products.
Mosqueira, E. A., et al. (2014). Quinoa biodiversity and sustainability for food Food and Bioprocess Technology, 9(1), 49–68. https://doi.org/10.1007/s11947-015-
security under climate change. A review. Agronomy for Sustainable Development, 34 1584-y
(2), 349–359. https://doi.org/10.1007/s13593-013-0195-0 Warren, F. J., Gidley, M. J., & Flanagan, B. M. (2016). Infrared spectroscopy as a tool to
Sanchez de la Concha, B. B., Agama-Acevedo, E., Nuñez-Santiago, M. C., Bello- characterise starch ordered structure - a joint FTIR-ATR, NMR, XRD and DSC study.
Perez, L. A., Garcia, H. S., & Alvarez-Ramirez, J. (2018). Acid hydrolysis of waxy Carbohydrate Polymers, 139, 35–42. https://doi.org/10.1016/j.carbpol.2015.11.066
starches with different granule size for nanocrystal production. Journal of Cereal Waterschoot, J., Gomand, S. V., & Delcour, J. A. (2016). Impact of swelling power and
Science, 79, 193–200. https://doi.org/10.1016/j.jcs.2017.10.018 granule size on pasting of blends of potato, waxy rice and maize starches. Food
Shevkani, K., Singh, N., Bajaj, R., & Kaur, A. (2017). Wheat starch production, structure, Hydrocolloids, 52, 69–77. https://doi.org/10.1016/j.foodhyd.2015.06.012
functionality and applications-a review. International Journal of Food Science and Wright, K. H., Huber, K. C., Fairbanks, D. J., & Huber, C. S. (2002). Isolation and
Technology, 52(1), 38–58. https://doi.org/10.1111/ijfs.13266 characterization of Atriplex hortensis and sweet Chenopodium quinoa starches. Cereal
Singh, N., Kaur, S., Kaur, A., Isono, N., Ichihashi, Y., Noda, T., et al. (2014). Structural, Chemistry Journal, 79(5), 715–719. https://doi.org/10.1094/CCHEM.2002.79.5.715
thermal, and rheological properties of Amaranthus hypochondriacus and Amaranthus Wu, G., Morris, C. F., & Murphy, K. M. (2017). Quinoa starch characteristics and their
caudatus starches. Starch - Stärke, 66(5–6), 457–467. https://doi.org/10.1002/ correlations with the texture profile Analysis (TPA) of cooked quinoa. Journal of Food
star.201300157 Science, 82(10), 2387–2395. https://doi.org/10.1111/1750-3841.13848
Singh, N., Kaur, L., Sandhu, K. S., Kaur, J., & Nishinari, K. (2006). Relationships between Yang, L., Xia, Y., Junejo, S. A., & Zhou, Y. (2018). Composition, structure and
physicochemical, morphological, thermal, rheological properties of rice starches. physicochemical properties of three coloured potato starches. International Journal of
Food Hydrocolloids, 20(4), 532–542. https://doi.org/10.1016/j. Food Science and Technology, 53(10), 2325–2334. https://doi.org/10.1111/
foodhyd.2005.05.003 ijfs.13824
van Soest, J. J. G., Tournois, H., de Wit, D., & Vliegenthart, J. F. G. (1995). Short-range Yoo, S., & Jane, J. (2002). Structural and physical characteristics of waxy and other
structure in (partially) crystalline potato starch determined with attenuated total wheat starches. Carbohydrate Polymers, 49(3), 297–305. https://doi.org/10.1016/
reflectance Fourier-transform IR spectroscopy. Carbohydrate Research, 279, 201–214. S0144-8617(01)00338-1
https://doi.org/10.1016/0008-6215(95)00270-7 Zhu, F., & Cui, R. (2019). Comparison of molecular structure of oca (Oxalis tuberosa),
Srichuwong, S., Curti, D., Austin, S., King, R., Lamothe, L., & Gloria-Hernandez, H. potato, and maize starches. Food Chemistry, 296, 116–122. https://doi.org/10.1016/
(2017). Physicochemical properties and starch digestibility of whole grain sorghums, j.foodchem.2019.05.192
millet, quinoa and amaranth flours, as affected by starch and non-starch

You might also like