Next Article in Journal
Visualizing the Distribution of Jujube Metabolites at Different Maturity Stages Using Matrix-Assisted Laser Desorption/Ionization Mass Spectrometry Imaging
Previous Article in Journal
Determination of Modified QuEChERS Method for Chlorothalonil Analysis in Agricultural Products Using Gas Chromatography–Mass Spectrometry (GC-MS/MS)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Non-Traditional Starches, Their Properties, and Applications

Department of Carbohydrates and Cereals, University of Chemistry and Technology, Prague, Technicka 5, 166 28 Prague, Czech Republic
*
Author to whom correspondence should be addressed.
Foods 2023, 12(20), 3794; https://doi.org/10.3390/foods12203794
Submission received: 27 September 2023 / Revised: 11 October 2023 / Accepted: 13 October 2023 / Published: 16 October 2023
(This article belongs to the Section Food Physics and (Bio)Chemistry)

Abstract

:
This review paper focuses on the recent advancements in the large-scale and laboratory-scale isolation, modification, and characterization of novel starches from accessible botanical sources and food wastes. When creating a new starch product, one should consider the different physicochemical changes that may occur. These changes include the course of gelatinization, the formation of starch–lipids and starch–protein complexes, and the origin of resistant starch (RS). This paper informs about the properties of individual starches, including their chemical structure, the size and crystallinity of starch granules, their thermal and pasting properties, their swelling power, and their digestibility; in particular, small starch granules showed unique properties. They can be utilized as fat substitutes in frozen desserts or mayonnaises, in custard due to their smooth texture, in non-food applications in biodegradable plastics, or as adsorbents. The low onset temperature of gelatinization (detected by DSC in acorn starch) is associated with the costs of the industrial processes in terms of energy and time. Starch plays a crucial role in the food industry as a thickening agent. Starches obtained from ulluco, winter squash, bean, pumpkin, quinoa, and sweet potato demonstrate a high peak viscosity (PV), while waxy rice and ginger starches have a low PV. The other analytical methods in the paper include laser diffraction, X-ray diffraction, FTIR, Raman, and NMR spectroscopies. Native, “clean-label” starches from new sources could replace chemically modified starches due to their properties being similar to common commercially modified ones. Human populations, especially in developed countries, suffer from obesity and civilization diseases, a reduction in which would be possible with the help of low-digestible starches. Starch with a high RS content was discovered in gelatinized lily (>50%) and unripe plantains (>25%), while cooked lily starch retained low levels of rapidly digestible starch (20%). Starch from gorgon nut processed at high temperatures has a high proportion of slowly digestible starch. Therefore, one can include these types of starches in a nutritious diet. Interesting industrial materials based on non-traditional starches include biodegradable composites, edible films, and nanomaterials.

Graphical Abstract

1. Introduction

Starch makes up almost 90% of the daily calorie intake from staple foods worldwide [1]. It affects food texture and is used in non-food applications.
This carbohydrate is industrially isolated mainly from corn, tapioca (cassava), wheat, potato, sago [2], pea, arrowroot [3,4], and rice. The two most common starch sources are corn (75% of global production) and tapioca (12%). In Europe, starch is manufactured mainly from corn, wheat, and potato. Additionally, pea starch is produced by Emsland Group (Germany) and by Roquette Frères (France). Oat starch is also widely used in the pharmaceutical and cosmetic industries [5]. Finland is one of the few countries in the world producing barley starch. During World War II and later in the former German Democratic Republic, ray starch was manufactured [1].
Several commercial starches originate from specific cultivars containing high-amylose starch (70% amylose) or waxy starch with a high amylopectin content. Amylose corn starch is produced by Xinfuwang New Material Technology Co., Ltd. (Henan, China) [6], and waxy corn starch by Lai Tangyuan Food Co., Ltd. (Jiangsu, China) [7] or by Samyang Genex Company (Seoul, Republic of Korea) [8].
Some starches are practically unknown in Europe. Achira starch (Canna edulis) is produced in Colombia and Venezuela [4,9], and Dua Naga, Ltd. (Indonesia) provides arenda starch. Lotus root starch is manufactured by Yangzhou Lianshun Food Co., Ltd. (Yangzhou, China) [10]. Mung bean starch can be purchased from Hengshui Fuqiao Starch Co., Ltd. (Jingxian, China) [11]. In the text below, there will be some discussion about sago and arrowroot starch.
Starch has many applications in food and non-food industries. However, the use of starch in its original form has some restrictions due to a poor flow ability, low paste transparency, susceptibility to retrogradation after gelatinization, high syneresis, low shear stress resistance, high gelatinization temperatures, rapid enzymatic degradation in the body, high gel turbidity, and high water content [12,13,14].
Therefore, there is an effort to prepare new starch derivatives or find non-traditional starches to overcome these disadvantages. In addition, there are several challenges in new industrial materials and the food industry based on customer needs that help the environment and human health. New trendy food products based on plants include snacks and food with “clean-label” starches, which respect the environment by producing no chemical by-products [15] or containing resistant starch. The hypothesis of this paper was to collect enough data from several years ago about the physicochemical properties of non-traditional starches to enhance their value-added application. The reference papers contain a detailed arrangement of the analytical methods and setting of instruments. A novelty of this research is waste utilization for isolation to obtain a new type of starch.

2. Isolation of Starch

The industrial isolation of starch depends on many factors. The first step can be the steeping of the grain (e.g., tapioca or corn), cold extraction of the mashed product (e.g., potato), or preparation of dough or slurry from flour (e.g., wheat). Most native starches have a unimodal distribution of granule size. Cereal starches are bimodal, so one needs to separate starch granules into two fractions. The botanical origin, weather conditions and environmental factors, farming techniques, and procedure used for starch isolation can all influence its chemical composition and structure [16].
Table 1 schematically summarizes some tested laboratory procedures for starch isolation from various sources. Of course, the laboratory isolation of nonconventional starches can differ from the common industrial ones.
The mentioned procedures vary greatly. The first part of the isolation follows from the character of the raw material. Some researchers have used sodium hydroxide solution (0.16–2%) or added protease for dissolving proteins from starch granules. Other scientists have applied sodium or potassium metabisulfite, sodium bisulfite, or sodium sulfite to prevent oxidation and microbial activity. In some cases, they have used organic solvents like hexane or ultrasonification.

3. Size, Shape, and Crystallinity of Starch Granules

As will be seen below, the size of starch granules affects the properties and use of the given starch, but on the other hand, the separation technique applied is also important. Small-granule starches have unique properties due to a higher water affinity, larger surface-area-to-volume ratio, and low sedimentation coefficient [21,78,79].
Table 2 summarizes the sizes and shapes of starch granules depending on their source. The smallest granules (<5 µm) with a high specific area were found in Agriophyllum squarrosum, amaranth, foxnut, and gorgon nut seeds and in the leaves of Arabidopsis thaliana. Large particles can be suitable for good separation; the diameter D [4,3] is defined in terms of the moment-ratio system as:
D 4,3 = n i D i 4 n i D i 3
where ni is the number of the measured particles and Di is their diameter. Kidney bean and canistel starches demonstrate a high value of this parameter. Ma et al. [80] measured out 200 µm in oat seeds. Chavez-Salazar et al. [65] found 65 μm long particles in unripe plantains. The largest average size (47 μm) was starch from achira rhizomes. The shape of particles highly varies and can be spherical, oval, kidney, elongated, discoid, conical, polygonal, triangular, or irregular.
Native starches include A-, B-, and C-types according to their crystallinity determined using X-ray diffraction. A- and B-type starch contains only A- and B-type crystallinity, respectively, but C-type starch contains both A- and B-type crystallinities. According to the proportion of A- and B-type crystallinity from high to low, C-type starch is usually divided into CA, CC, and CB-types [81]. Only a few starches possess B-type crystallinity, specifically those from achira rhizomes, lily bulbs, Phoenix sylvestris, and ulluco tubers, or pumpkin and winter squash fruit (Table 2).

4. Chemical Structure, Microstructure Observation

The length and configuration of the chains in amylopectin influence many properties such as the swelling power, water solubility index, crystallinity, pasting and thermal properties, main gelatinization temperatures [71], retrogradation [33], syneresis, resistant starch formation and digestibility [35,61], hardness and cohesiveness of the gels [42,61], capability for esterification and functionality of the derivative [65], and resistance to enzyme hydrolysis [77].
Hanashiro et al. [93] defined branches of amylopectin as A, B1, B2, and B3 types having degrees of polymerization of 6–12, 13–24, 25–36, and >37, resp. The chain length distributions of starch samples were determined using a high-performance anion-exchange chromatography system [27]. Small granules like those from send rice (Agriophyllum squarrosum) or quinoa starches had a high percentage of A-chains and a low percentage of B3. Similarly, the number-average degree of polymerization of amylopectin decreased with a decreasing granule size in barley starch [94]. On the contrary, the minimum number of A-chains was in canna starch and high-amylose corn starch, and Chlorella, canna, and tapioca starch with high-amylose corn starch had a maximum B3 chains content (Table 3).

5. Data of Differential Scanning Calorimetry, Pasting Properties

Differential scanning calorimetry (DSC) on starch is indicative of interfering with differences in starch structures, changes in the physical states of starch, and the interactions of starch polymers with other constituents and composite food matrices. This method has enabled the measurement of gelatinization temperatures, provided evidence that chemical modification has reduced or eliminated aggregation and the association of starch molecules during the cold storage of pastes, and evaluated the temperature and enthalpy of the formation of amylose–lipid complexes [1].
The temperature at which gelatinization occurs and the heat enthalpy are closely connected to the starch structure and molecular arrangement. Furthermore, the amylose content could also influence the enthalpy of gelatinization. The thermal data found in recent papers are in Table 4. From an industrial point of view, the onset temperature (To), which represents the initiation of the gelatinization stage in starch granules, is associated with the costs of the used processes in terms of energy and time, and thus a low temperature is required by manufacturers. This low To was found in acorn starch [19]. On the other hand, the highest To (142.9 °C) was found in sago starch isolated from pith waste fiber with a high crystallinity structure [69]. The starch melting temperature range (Tc–To) indicates the homogeneity and quality of the amylopectin crystals [96]. A narrow melting range corresponds to amylopectin crystals with a more homogeneous quality and uniform stability and vice versa [97]. The narrow range of (Tc–To) characterizes starch gained by ginger (7.4 °C), sand rice (8.4 °C), winter squash var. Heili (8.5 °C), avocado seeds (9.2–10.7 °C), Carioca beans (10.1–13.8 °C), common corn (10.4 °C), and ramon seeds (10.6 °C).
The gelatinization endotherm is associated with the gelatinization of the crystalline amylopectin lamella, but not necessarily to the proportion of the crystalline area, which corresponds to the amylose: amylopectin ratio and the distribution of amylopectin chains in the granule [37]. A higher enthalpy can also indicate a higher content of starch–lipid complexes, which requires more energy to disrupt [80]. Starches from Euryale ferox seeds and sago pith waste fiber demonstrated the highest enthalpy.
The retrogradation enthalpy (ΔHret) reflects the energy required to melt the recrystallized starch structure and starch molecular double helices formed during long-term storage. There are not enough data for a comparison of ΔHret. The highest value was found for sand rice starch (its gel exhibited significant recrystallization), while Phoenix sylvestris starch demonstrated the lowest value.
As for the pasting properties, viscosity describes the magnitude of the friction generated by the mutual movement of starch granules after gelatinization [100]. A comparison among the data from the references is very difficult, because the input concentration of the tested suspensions in aluminum vessels differed from 3% (db w/w) to 16.67% (w/w)—see Table 5.
The peak viscosity (PV) of starch indicates the maximum swelling caused by heat in the presence of water [101]. When we exclude the values for extremely high or undefined starch concentrations, the highest PV was found in samples of starches coming from ulluco, winter squash, bean, pumpkin, quinoa, and sweet potato. On the other hand, waxy rice and ginger starch had a low PV by a higher starch concentration.
Breakdown (BD) equals PV minus trough (minimum) viscosity [102]. BD represents the thermal stability of starch paste [40] and is the function of the rigidity of the swollen starch granules [103]. The authors found the maximum BD in mung bean, quinoa, and sweet potato starches. The lowest values were in ginger, foxnut, and normal corn starches. A low breakdown viscosity indicates a high paste stability to remain intact at a high temperature. This characteristic of starch will be suitable for food products requiring a high processing temperature, and alterations in viscosity are unwanted during heating and cooling [104].
High final viscosities (FV) were found in Turkish bean and winter squash starch. On the contrary, this was low in waxy amaranth. Setback (SB) equals FV minus trough (minimum) viscosity [103]. SB indicates the recrystallization degree of starch during the cooling of starch paste, especially the recrystallization and rearrangement of amylose molecules [80]. Waxy varieties and chestnuts demonstrated a reduced setback viscosity, while Turkish beans, rice, and quinoa showed the highest SB.

6. Swelling Power, Solubility and Water Absorption

The extent of granule swelling is determined by measuring the swelling power (SP; g/g), which is reported as the ratio:
S P = m g m s t m s o l
where mg— the weight of the gel (g), mst—the weight of the starch sample (g), and msol—the soluble weight (g).
SP measures the hydration capacity, because food-eating quality is often associated with the water retention in the swollen starch granules [105].
Solubility (SW; %) is the percentage of molecules leached from the starch granules after swelling under specific conditions:
S W = m s o l m s t 100
The SP and SW values provide information about the degree of interaction between the chains present in the crystalline and amorphous regions of the starch granules and are correlated with amylose content [71]. A higher amylose content in starch contributes to a lower functionality, which could limit its application in food-processing industries such as bakeries [24]. In addition, the formation of protein or lipid complexes also influences the SP [64].
The water absorption index (WAI; g/g) is defined as:
W A I = m g m s t
and had similar values to the SP. The results of the swelling power solubility and water absorption of different starches, at temperatures ranging from 60 to 80 °C, are presented in Table 6. High values of SP were found for starches of canna, quinoa, annatto, Agriophyllum squarrosum, and Phoenix sylvestris at 80 °C. All starches showed an increasing swelling power at increasing heating temperatures because of gelatinization.
Starches of canna, quinoa, black bean, and Phoenix sylvestris showed the highest SW values measured at 80 °C, while anchote, ginger, and foxnut starches demonstrated the lowest SP values.

7. Refrigeration and Freezing Stability of Starches

Frozen foods containing starch can be repeatedly freeze–thawed during storage and transportation. Thermal fluctuations in foods during transport cause water melting and recrystallization, lowering their quality and acceptability. Simultaneously, the starch recrystallizes during cooling, resulting in water separation known as syneresis. Syneresis is an undesirable phenomenon that results in a poor starch gel texture [101].
Chemical modifications such as acetylation and hydroxypropylation can positively influence freeze–thaw stability. These processes have also been tested on non-traditional starches [106].
Zarroug et al. [19] studied the stability under refrigeration and freezing in different native starches during storage times of 24, 48, 72, 96, and 120 h. According to Wu et al. [80], adzuki bean starch had a higher amylose content, swelling capacity, and freeze–thaw stability.
Hussain et al. [101] published the percent syneresis of corn, sweet potato, and Turkish bean starch gels. Vázquez-León et al. [35] dealt with the stability under refrigeration and freeze–thaw of black bean starch gel, measured as a percentage of syneresis (water exuded), determined after the 1st–5th cycles. A comparison of the cited data is not possible due to the different calculations of syneresis and low scope of the references. We can just make the following provisional conclusions:
-
the lower the temperature of the refrigeration, the higher the syneresis,
-
the longer the storage time, the higher the syneresis,
-
the syneresis of acorn starch is comparable to corn starch,
-
the syneresis of sweet potato starch and Turkish bean starch is lower than that of corn starch.
Photinam et al. [107] reported that freezing at low temperatures after gelatinization resulted in the recrystallization of starch molecules and a higher resistance to enzymatic hydrolysis. Rodboontheng et al. [108] found that canna starch granules subjected to freezing were more intact and smaller compared to those without freezing. Upon freezing, the swollen granules shrank to smaller-sized granules.
According to Liu et al. [44], native chestnut starch has characteristics similar to corn starch, such as a higher swelling power, better freeze–thaw stability, and lower pasting and gelatinization temperatures, which has made it potentially a good source of starch.

8. Hydrolysis, Digestibility

As the principal energy source of the human body, the digestibility of starch is an important parameter in the quality assessment of food. During starch digestion, glucose is released. Thus, starch digestion is closely associated with the postprandial blood glucose and insulin levels in humans [84]. It is influenced by various factors, for example, structural characteristics, the formation of starch–protein complexes (containing strong disulfide or tannin bonds [109]), interactions with polyphenolics [110], β-carotene [111] and other components, and preheating treatment. Wu et al. [80] dealt with digestibility of prepared complexes of adzuki bean starch-fatty acids.
The amorphous area of the starch molecule is more amenable to acid attacks, being more easily and quickly hydrolyzed [29]. However, a high amylose content in starch decreases its digestibility. A smooth surface of starch granules, without grooves, hinders the enzyme’s diffusivity inside the granule and delays the hydrolysis [37]. Their size can also affect their digestibility, as the surface area and starch volume affect the interaction between the starch substrate and enzyme. Therefore, the large surface area of small granules has resulted in a high rate of enzymatic hydrolysis [94,112].
The lower digestibility of starch results in lower obesity and microbial formation of short-chain fatty acids in the human gut, thus having healthy effects [110]. Englyst et al. [113] divided starch into rapidly digestible starch (RDS; starch digested from 0–20 min), slowly digestible starch (SDS; digested from 20–120 min), and resistant starch (RS; remaining starch) based on simulated digestion properties in vitro and bioavailability. RS can be used as dietary fiber, reducing calorie intake (low glycemic index), preventing fat deposition and colon cancer, and enhancing mineral absorption [22]. RS includes five groups.
The lower in vitro digestibility of legume starches than that of cereal starch might be due to their high amylose content (because of their lower susceptibility to enzymatic hydrolysis), the absence of surface pores on the granules, large amounts of B type crystallites, strong interactions between amylose chains, intact cell structures enclosing starch granules, and large amounts of soluble dietary fiber [35].
The starch molecules in freshly gelatinized starches are amorphous and susceptible to digestive enzymes [71]. Transglycosylation forming pyrodextrins is responsible for creating new glycosidic bonds, which, in turn, reduce digestibility [29].
Table 7 includes some new results of starch digestibility. Unripe plantain, pumpkin, and winter squash Yinli starches demonstrate the lowest RDS (<3%) and the highest RS (>90%). Indigenous Himalayan folk rice cultivars are good examples of the dependence RS on amylose content. For an amylose content in the range of 0.7–29.1%, RS was low, i.e., 0.4–2.3% [72], but for a higher amylose content of 31.8–40.7%, RS rapidly increased to 85.4–92.8% [73].
However, most of these starches are not eaten in their native form. A low RDS is maintained in cooked lily starch (20%) and a high RS content was found in lily starch (>50%) and unripe plantains (>25%). The highest SDS was found in the native starch of annatto and Euryale ferox (>35%). A very high SDS content was found in the thermally (80–100 °C) processed starch of gorgon nut.
Besides the evaluation of RDS, SDS, and RS, simulations of starch digestion in the small intestine were performed using various evaluation models [61,87]. The first-order rate equation was estimated for the digestibility kinetics according to the following formula:
C t = C . ( 1 e k t )
where Ct and C (mg/mL) are the digestibility of starch when the hydrolysis time is t (min) from 0– , respectively, and k indicates the kinetic coefficient of the starch digestibility (min−1). Liu et al. [87], Mahajan et al. [61], and Wu et al. [114] evaluated C at 42.6% and k at 0.029 min−1 for chestnut starch, C at 64.0% and k at 0,0264 min−1 for Kutki millet starch, and C at 66.5% and k at 0.0297 min−1 for adzuki bean starch, resp., compared to C at 65.6% and k at 0.0332 min−1 for corn starch. Mahajan et al. [61] also listed other digestion models, e.g., the Duggleby model, Paolucci-Jeanjean model, and Weibull model.
The simple magnitude is the degree of starch digestion or enzyme hydrolysis (%) defined as:
D E H = 100 . m H S m I S
where mHS (g)—the weight of hydrolyzed starch and mIS (g)—the weight of initial starch [108,115]. The rate of digestion is closely related to the glycemic index.

9. FTIR Spectral Analysis

Fourier-transform infrared spectroscopy (FTIR) is a powerful tool for characterizing the structure of starch and starch-based composite films [25,89,116]. This technique enables (i) identifying impurities in isolated starch, (ii) identifying chain conformation, single- and double-helical structures of starch, (iii) evaluating the crystallinity of starch-based products, (iv) identifying functional groups in chemically modified starches, and (v) identifying changes in starch-based biodegradable films.
The observed FTIR spectra are similar for all studied starches [25,117]. A broad and intense band of 3700–3000 cm−1 represents the stretching vibrations of O–H bonds in hydroxyl groups and overlap narrow weaker bands at 2928–2931 cm−1 attributed to the stretching vibration of the C–H bond in glucopyranosyl residues. The bands observed at 1153, 1086, 1012, and 930 cm−1 in the spectral region, often assigned as “sugar region”, are attributed to C–O–C, C–O, and C–C stretching and C–O–H bending vibrations in the glycosidic linkages and pyranoid rings [19,25]. Several bands at 921–930 cm−1 and 767 cm−1 are characteristic for 1,4-α-glycosidic bonds in starches [30,64], and the band at 860 cm−1 arises from C1α–H and CH2 deformations and is indicative of the α-anomer of glucopyranose [118]. Finally, several bands at 708 cm−1 and below arise from the skeletal modes of the pyranoid ring of α-glucopyranose [119].
Water molecules play a crucial role in starch structure [120] and contribute to the broad absorption at 400–900 cm−1 (twisting vibration of bound water), 1640–1650 cm−1 (HOH scissoring vibration), and 3500 cm−1 (HOH stretching vibration). In hydrated starches, the band near 2100 cm−1 can be attributed to the free water content [25,121] and assigned to vibrations from the scissoring and rocking vibrations of the retained water [121]. In addition, hydration causes non-linear changes in the FTIR spectra of starches depending on the contribution of the ordered structure, partially causing a shift in peak position from 1000 to 1200 cm−1 [116].
The bands in “sugar region” are sensitive to hydration and the crystalline/amorphous stay of starches. The IR band at 1022 cm−1 is related to the amorphous phase [9], while the bands at 993–995 cm−1 and 1047 cm−1 correspond to the crystalline region of the starch [25,89]. The ratio of the band intensities at 1047 cm−1/1022 cm−1 and 995 cm−1/1022 cm−1 quantitatively characterizes the degree of short-range order and the proportion of double helices in starches, respectively [89,116]. For sand rice (Agriophyllum squarrosum) and barley starches, the ratio of band intensities at 1047 cm−1/1022 cm−1 confirmed that polysaccharide chains are rearranged when heated under a low humidity, equal to or below 20%, but at a higher humidity, reaching 30%, short-range ordered structures began to be disrupted [122]. The mentioned ratios can be improved by the addition of another polysaccharide to the starch [84,115,123]. However, the possible overlap of the vibrational bands of these polysaccharides in the “sugar region” may influence the obtained ratios.
The intensities of the bands at 1200–1400 cm−1 and at 2940 cm−1 increased after the partial acidic hydrolysis of achira (Canna edulis) starch [9] and the treatment of pea starch with pullulanase [89]. These bands included CH2 vibrations and were sensitive to hydrogen bonds involving starch hydroxylic groups. The increase in these bands may indicate an enhanced number of free CH2OH groups due to the hydrolysis of 1,6-α-glycosidic bonds (unbranching).
The FTIR spectra of chemically oxidized cassava starch showed a carboxylate antisymmetric stretching band at 1600 cm−1 and some disturbances at 1060 cm−1, probably due to stretching of the C1–O5 bond in the pyranoid ring [124]. These spectral changes were consistent with the structural features of this starch suitable for baking properties. The FTIR spectrum of oxidized jackfruit seed starch demonstrated a new band at 1760 cm−1 assigned to the C=O stretching vibration of the carbonyl group [54]. The oxidation of lotus root starch led to the appearance of new IR bands at 1610, 1726, and 2820 cm−1 due to vibrations of new carbonyl and carboxylate groups [9]. The differences in the FTIR spectra of oxidized lily starches highlighted the sensitivity of crystalline and amorphous structures to specific treatments using the ratio of band intensities at 1047 cm−1/1022 cm−1 [56]. The FTIR spectra of starches esterified with fatty acids had a new IR band of C=O stretching vibration at 1749 cm−1 [125]. Two IR bands of CH2 stretching vibrations at 2860 and 2926 cm−1 also appeared and increased with the length of acyls. A new IR band of phosphorylated starches at 1417 cm−1 corresponded to the P=O stretching vibration [88].
The FTIR spectra of adzuki bean starch–fatty acid complexes demonstrated two narrow bands at 2850 cm−1 and 2917 cm−1 assigned to the antisymmetric and symmetric stretching of the CH2 groups in the long-chain fatty acids. In addition, a weak band at 1383 cm−1 assigned to the CH3 bending vibration was also found [112]. Additionally, the bands at 1635–1656 cm−1 (amide I), 1545 cm−1 (amide II), and 1240–1350 cm−1 (amide III) are attributive to the vibrations of amide groups in proteins, which are often present in starch preparations [19,25].
Composite starch-based edible films often contain proteins, lipids, aromatics, and other polysaccharides. FTIR spectroscopy is able to detect the presence of these compounds via characteristic vibration bands. Starch-based film represents specific bands at 3326 cm−1, corresponding to O–H stretching in glycerol, starch, and water [62]. The FTIR spectra of biodegradable edible film obtained from arrowroot starch and iota-carrageenan showed characteristic bands of ι-carrageenan at 1249–1269, 924–926, and 842–849 cm−1 assigned to S=O stretching, 3,6-anhydrogalactose, and galactose-4-sulfate vibrations, respectively [3]. Hussain et al. [101] observed two IR bands at 2927 and 2940 cm−1, indicating the asymmetric stretching vibration of the CH3 groups of the acetyl and methyl esters in cactus (Opuntia ficus-indica) and acacia (Acacia seyal) gums in a mixture with various starches. The IR band observed at 1721–1728 cm−1 indicates the C=O stretching vibration of the carboxylic acids and esters in polysaccharidic gums. The bands at 1415–1400 and 1310–1230 cm−1 corresponded to the symmetric stretching of the COO- of the carboxylate groups and C–O stretching of the carboxylic acids in the gums, respectively.

10. Raman and NMR Spectroscopies

In addition to FTIR, Raman spectroscopy is also widely used in the structural analyses of starches, for example in the characterization of the structural changes in retrograded [126,127], microwave-treated [119], and irradiated [118] starches. FTIR and Raman spectra were used in the analysis of biodegradable starch-based films containing saturated fatty acids [128]. Both of these vibrational spectroscopy techniques were applied to evaluate how heat treatments and extraction procedures affect the properties of ajocotte starches [129].
The Raman spectra of starches are sensitive to the presence and type of crystalline structure, degree of substitution, and relative amounts of amylose and amylopectin. The band intensity in the Raman spectra for starches depends on their structural features [130]. The typical Raman bands of starches near 2910, 1458, 1260, 1124, 864, and 475 cm−1 correspond to C–H stretching, C–H, CH2, and C–O–H bending, C–O–H, C–C–H, and O–C–H bending, C–O and C–C stretching, C–C–H and C–O–C bending, and pyranoid ring skeletal vibrations, respectively. In amorphous starch, Raman bands, except those at 2910 and 478 cm−1, were weak and almost disappeared at low frequencies. The Raman intensity varies directly with the size of the crystalline region in starch. The Raman bands for B-type starch (small crystalline part) were less intense than those for A-type starch (large crystalline part), and the most of these bands shifted to lower wavenumbers. De Gussem et al. [131] observed evident shifts in the Raman bands of amylose at 405, 481, 757, and 855 cm−1 compared to the corresponding bands of amylopectin at 410, 477, 769, and 865 cm−1. The skeletal vibration band of starch near 477 cm−1 is one of the most intense in the Raman spectrum and can be used as a marker to detect starch in mixtures with other polysaccharides [132].
Since Raman spectroscopy is less sensitive to the contribution of water and polar groups, it is more prospective in the analysis of aliphatic moieties in substituted starches. For example, no significant changes were observed for cationic quaternary ammonium starches in the FTIR spectra due to the overlap of the strong OH and H2O stretching bands [133]. In contrast, the corresponding Raman spectra showed several bands at about 3030, 970, and 761 cm−1 assigned to the vibrations of trimethylammonium substituents. The ratio of the band intensities at 761 cm−1/938 cm−1 (reference starch band) can quantify the degree of substitution. Raman spectroscopy was applied to determine the degree of succinylation in modified starches [134].
Nuclear magnetic resonance (NMR) spectroscopy is a powerful tool in the structural analysis of natural and modified polysaccharides, including starches. Applications of this method in the characterization of starch systems were reviewed by Zhu [135]. Various techniques of NMR spectroscopy can (i) characterize the structure and physical properties of native and modified starches, including the effects of moisture and the impact of physical treatments, (ii) analyze starches in gel and solid states, (iii) identify enzyme susceptibility and resistant starches, (iv) identify and quantify functional groups in chemically modified starches and starch-grafted copolymers, and (v) analyze the interaction of starch with non-starch components in food and other complex systems. For example, 1H NMR has been applied to the determination of the degree of substitution of acetylated and other acylated starches [136,137,138], and octenyl succinic anhydride-modified waxy corn starch was characterized using 1H and 13C NMR [139].
Solid-state 13C CP/MAS NMR spectra can evaluate the molecular organization of starch samples [5,63,89]. Broad resonance signals at 94–105 and 80–84 ppm correspond to the C1 and C4 carbons, respectively, involved in 1,4-α-glycosidic bonds. The former resonance signal is sensitive to the crystalline structure of starch and can split into 2–3 components, while the latter resonance signal indicates the amorphous structure in starch. For example, oat starch had a triplet peak at 101.36, 99.23, and 98.84 ppm, typical of an A-type crystalline structure [5]. After treatment at 500 MPa for 15 min, this region transformed into a single strong peak at 101.36 ppm, and the relative area of the C4 signal increased, confirming the gelatinization of oat starch. Large and overlapping signals at 68–78 ppm arose from the remaining endocyclic carbon atoms C2, C3, and C5. A signal at 58–65 ppm corresponded to the C-6 carbon atoms of exocyclic CH2OH groups [140]. For native and partially degraded pea starch, the ratio of the integral area at 58–65 ppm to the sum of areas of all the carbon signals quantify the relative proportion of double helices in starch, and the ratio of the integral area of the C4 signal to the sum of areas of all the carbon signals quantify the amount of amorphous phase [89]. Alternatively, the total 13C CP/MAS NMR spectra deconvoluted by subtracting the scaled NMR spectrum of the amorphous starch to zero intensity at 84 ppm yield the ordered and amorphous phases of the starch samples [122,141,142]. Li et al. [63] studied the cold water swelling of oat starch using subcritical ethanol–water treatment using X-ray diffraction, FTIR, and 13C CP/MAS NMR spectroscopy. The obtained results confirmed that the structure of the modified starch changed from A-type double helical to V-type single helical and the short-range crystalline structure decreased as well.
Time domain 1H NMR can evaluate the molecular mobility and water distribution in the starch–water systems [143]. This method can be applied to elucidate the molecular dynamics of starch transitions during gelatinization in dough/batter systems during heating/cooling and used in the quality control of starchy food products.

11. Starch Modification

Native starch may be modified in numerous ways, for example, physically, chemically, or via an enzyme action.
Pregelatinized starch originates when starch slurry is heated by steam [86]. It is a common ingredient in foods, as well as other industrial applications. The other tested physical methods for starch modification are microwave treatment [31], annealing (ANN), heat moisture treatment (HMT), high-pressure homogenization [5,43], ultrasound [22,69], or ultrasound combined with plasma treatment [40].
Biduski et al. [144] found that HMT increased the peak viscosity and breakdown while reducing the final viscosity and setback of the Pinhão starch. There was an increase in the gelatinization temperature and a decrease in the ΔH after treatment. HMT changed the short-range ordered structures, shifted the X-ray pattern from C-type to A-type, and reduced the relative crystallinity. Akinyosoye and Nwokocha [76] investigated the effect of HMT on the properties of Treculia africana starch. As the moisture level of HMT increased, the swelling decreased, and there was an increase in amylose leaching. Pasting temperature and retrogradation increased, and breakdown decreased with HMT. Thus, HMT strengthened the granule structure, resulting in starch which can withstand severe processing conditions without viscosity loss. Similarly, Wu et al. [74] tested the effect of heat-moisture treatment on the structural and physicochemical characteristics of sand rice starch. According to Rodboontheng et al. [108], swollen starches prepared via HMT are promising encapsulating agents for food ingredients and supplementary foods such as essential oils, nutritious oils, flavors, and pre/probiotics.
The ANN treatment applied to ulluco starch proved to be an efficient method for improving the physicochemical parameters such as solubility, water retention capacity, and the texture of the gel formed, which are valuable properties [91].
Natural starch can be modified in various ways to produce acid- or enzyme-modified, oxidized, cross-linked, esterified, or etherified starch, converted into cationic derivatives, or subjected to pyroconversion [29]. These modifications, present at a very low degree of substitution in starch preparations, produce dramatic differences in the physical and chemical properties of starch. The rate of chemical reaction depends on the starch crystallinity and is linked with the porosity and size of starch granules, temperature, pH, and the concentration of the chemical agent.
Partial hydrolysis is a way for the fabrication of acid- or enzyme-modified starches. This process is suitable for the preparation of porous starch, e.g., from banana [28] or lotus [89] starches.
Tung et al. [54] used a hydrogen peroxide agent to oxidize jackfruit seed starch. Xu et al. [9] studied the optimized process conditions of lotus root starches modified via treatment with sodium hypochlorite. Hu et al. [34] dealt with the oxidation of Tartary buckwheat starch by ozone. The product retained a relatively ordered crystal structure, lower amylose content, SP, solubility, digestibility, higher viscosity, and good viscoelastic gels than the native one, resulting from the cross-linking reaction between the aldehyde group produced by oxidation and the hydroxyl group of the adjacent starch in the amorphous region.
Phosphorylation and acetylation change the functional and physicochemical properties of starches. The hydroxypropylation of non-traditional starches enhances swelling capacity, viscosity, clarity, and freeze–thaw stability [86,106].
Also, a combination of physical and chemical modification was tested, e.g., ozone oxidation together with high-pressure homogenization [56] or high hydrostatic pressure and cationization [85].

12. Potential Application of Non-Traditional Starches

Generally, starch can be used as an additive, gelling agent, thickener, fat replacer, and emulsifier in crunchy foods, baby foods, and controlled release systems, and as a flavor enhancer and coating agent in many industrial foods, feeds, and drugs to improve their functional and sensory characteristics [19,24,98,112,145]. In the food industry, starch is used mainly to improve the properties of baking flour and bakery products such as cakes, bread, and noodles due to their high consumption. During their formulations, starch is one of the components responsible for broad technological functionality [19]. The push for plant-based food is expected to grow in new categories, e.g., snacks.
The development of new starch raw materials can be used for developing healthy foods required by consumers, thereby enhancing the value-added application [52]. They are increasingly choosing labels that help choices about the environment and human health. So-called “clean-label” starches represent a relatively new group, not marked with the E index in food (they are not chemically modified). Their advantage is that they are environmentally responsible by not producing chemical by-products [15]. Starches from new sources could replace chemically modified starches due to their properties being similar to commercial modified ones.
Resistant starch (not accessible to human digestive enzymes) has beneficial effects for health; it positively influences the digestive tract and its fermentation in the colon causes the formation of short-chain fatty acids and the lowering of the pH. It further increases fecal bulk, protects against colonic cancer, improves blood cholesterol levels, assists glucose tolerance control in diabetes, and causes lower blood lipid levels. However, the current RS/dietary fiber intake by Europeans is insufficient. Therefore, it is suitable to add it to food. According to Du et al. [146] and Yuan et al. [66], native beans, winter squash, and pumpkin starches are the RS sources in formulations with desired fiber-like benefits like a lower digestibility.
Some papers have solved the utilization of starch from wastes, e.g., coming from seeds of annatto [25], avocado, Chenopodium, and jackfruit [42,54,83,98]. Besides being used in the form of native starch from roots, tubers, seeds, and wastes, modifying non-conventional starches could provide options for extending the required engineering properties [54]. Octenyl-succinic-modified starches from various sources have been widely used to encapsulate food ingredients by forming emulsions. Moreover, nonenyl succinic quinoa-modified starch could be used for the fabrication of emulsions [99].
In recent years, the importance of high-amylose starch has grown due to its expanded range of functionalities compared to native starch. These functionalities include nutrition, food processing, medicine, and industrial use [33].
An outstanding group of starches covers starches having small granules. Cow cockle starch (0.3–1.5 μm), rice starch (2–8 μm), and amaranth starch are suitable as fat substitutes in frozen dessert and mayonnaise; similarly, acorn starch has been used in custard because the small particle starches could form a smooth creamy fat-mimetic texture [79,145,147]. Wheat B-starch with a particle size fraction below 8–10 µm has found uses in the production of alcohol, dextrin, and feed, in a mixture with bentonite for foundries [148], for hydrolysates, and for energy use in biogas units [15]. It has also been tested for use in biodegradable plastics [149]. The advantage of starches having small granules is their high specific surface.
Nanomaterials exhibit properties that are distinctive and qualitatively different from large-size particles. Starch nanoparticles (SNPs) developed from mango starch, mungbean, and water chestnut starch are comparable with wheat or potato nanoparticles. SNPs also have great potential for papermaking use, surface sizing, coating, and in paperboard as a biodegradable adhesive instead of petroleum-based adhesives. All the tested SNPs are non-cytotoxic and can be safely used in biomedical applications [150]. Nanocomposites produced with mung bean starch showed the best properties among those with different starches used as matrices. The higher amylopectin ratios in starch result in the better interaction, dispersion, and distribution of nanofillers in starch-based nanocomposites due to the branched molecular structure of amylopectin. Sago starch nanocomposites produced with zinc oxide nanorods increased the electrical conductivity [151].
Plastic-based synthetic polymers widely used in modern food packaging contribute to plastic-waste end-products, which accumulate in and pollute terrestrial and aquatic environments. Thus, there are ways of reducing the use of and dependence on them [3]. Low cost, low density, decreased tool wear, renewability, and degradability [152] are all recognized benefits of starch resources for the industry manufacturing biodegradable plastics. Involving starch in bioplastics provides in two ways, i.e., in the form of a filler with intact starch granules or incorporation directly into the composite matrix, which is so-called thermoplastic starch [153]. Waste or by-products from sugar palms, cassava, or corn are suitable as polymer composite reinforcements [154]. On the other hand, poor resistance to water and low strength are limiting factors for materials manufactured from starch [155]. Blending starch with other polymers such is PVA is an effective and convenient way of overcoming the shortcomings of starch materials [45]. The use of nanotitanium dioxide nanoparticles and essential cinnamon oil incorporated into sago starch-based bionanocomposite films show excellent antimicrobial activity against Escherichia coli, Salmonella typhimurium, and Staphylococcus aureus, which results in a high potential for use in active packaging in food industries. Nanocomposite films processed with sugar palm starch, sugar palm nanocrystalline cellulose (SPNCC) with cinnamon essential oil have similar effect [151].
Edible biodegradable films (EF), intended for primary food packaging use, are made from food-grade and biodegradable polymers. Therefore, edible films are used to wrap the food product to inhibit the moisture uptake, extend the shelf life, and are consumed together with the food. Consequently, EF can generate almost zero waste after disposal [3]. EFs fabricated from starch polysaccharides are one of the primary genuine forms of biopolymer materials for food packaging due to strength, processing, product adaptability, rheological properties, and performance. Additionally, these products are environmentally friendly and cost-effective. EF development from sesame seed gum, mung bean, cassava, and avocado seed starches has been widely studied [23]. Also, native quinoa, amaranth, and ulluco starch are ideal materials for the fabrication of EF with a good mechanical strength, thermal stability, and antimicrobial activity compared to films made from corn starch and yam starch [77,156,157]. Materials with a low opacity, like EF, can be used in food packaging for improved product visualization and consumer acceptance [144].
Several non-conventional starches have technological advantages over common starches for specific applications in the cosmetic and paper industries [145]. In addition, starch can be applied as an adsorption material in many branches [15,28,108]. The biotechnological application of duckweed starch instead of commonly used starch feedstocks via the highly efficient fermentation of glycerol was tested by Yang et al. [48].

13. Conclusions

Nowadays, the research focuses on novel starch sources in the context of food security, sustainable production, the use of by-products, e.g., coming from annatto and avocado, their regional availability, and technological advantages over traditional starches.
Published numerous data like shape, size, and crystallinity characterize non-traditional starches. Small-granule starches demonstrate unique properties due to a higher water affinity, larger surface-area-to-volume ratio, and low sedimentation coefficient. Because of their smooth texture, they can be utilized as fat substitutes in frozen desserts, mayonnaise, and custard. These starches have shown an excellent compatibility in biodegradable plastics. They also can be applied in their native or modified form as adsorbents.
The practical use of starches in the food industry requires knowledge of the thermal and pasting properties of the suggested starches. The low onset temperature (To) in acorn starch is associated with the costs of the industrial processes in terms of energy and time. The requested high retrogradation enthalpy was characteristic of sand rice starch, the smallest one for starch from Phoenix sylvestris with a high swelling power.
Starch plays a crucial role in the food industry as a thickening agent. Starches from ulluco, winter squash, bean, pumpkin, quinoa, and sweet potato demonstrate a high peak viscosity (PV); on the other hand, waxy rice and ginger starch had a low PV. Low setback (SB), indicating recrystallization during the cooling of starch paste, was found in waxy varieties and chestnuts. Turkish bean, sand rice, and quinoa showed the highest SB.
New food trendy categories comprise plant-based food e.g., snacks and food containing “clean-label” or resistant starch. Native, “clean-label” starches from new sources could replace chemically modified starches, due to their similar properties compared to common commercial modified ones. Furthermore, their production does not produce chemical by-products. These techniques for preparing “clean-label” starches include pregelatinization, microwave treatment, annealing, heat moisture treatment, high-pressure homogenization, ultrasound, or ultrasound combined with plasma treatment.
As fiber, resistant starch (RS) has beneficial effects on health. It positively influences the digestive tract and its fermentation in the colon protects against colonic cancer. RS improves blood cholesterol levels, assists glucose blood level control in diabetes, and causes lower blood lipid levels. A high RS content was found in gelatinized lily starch (>50%) and unripe plantains (>25%), and starch from indigenous Himalayan folk rice has a amylose content higher than 32%. The low content of rapidly digestible starch is maintained in cooked lily starch (20%). Thermally processed starch from gorgon nut showed a very high SDS. Therefore, these starches are suitable for a healthy diet.
However, chemical modifications of non-traditional starches are still being tested, including oxidation, phosphorylation, acetylation, hydroxypropylation, and a combination of physical and chemical approaches. Spectroscopy methods are powerful tools for the characterization of the purity, structure, and physical properties of starches, their derivatives, and composite materials based on them. In addition to X-ray diffraction, FTIR, Raman, and NMR can evaluate the ratio between the crystalline and amorphous forms of starch.
Native and modified non-traditional starches can be used also outside the food industry. Materials from sugar palm, cassava, corn, and others have been discovered to be suitable as polymer composite reinforcements. Edible foils fabricated from starch polysaccharides are one of the primary genuine forms of biopolymer materials for food packaging due to their strength, processing, product adaptability, rheological properties, and performance; they are environmentally friendly and cost-effective. A relatively new progressive modification of starches enables the creation of nanomaterials, including nanoparticles/nanofillers, nanofibers, or nanocomposites, providing perspectives for various fields of industry.

Author Contributions

Conceptualization, E.Š. and A.S.; methodology, E.Š. and A.S..; formal analysis, E.Š. and A.S.; resources, M.S.; data curation, E.Š., A.S. and P.S.; writing—original draft preparation, E.Š. and A.S.; writing, E.Š. and A.S.; visualization, E.Š.; supervision, A.S. and M.S.; project administration, E.Š. and P.S.; funding acquisition, E.Š. and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported from the pedagogical project of the university—grant No C1_VSCHT_2023_049 (Innovation of the textbook Chemistry and Technology of Starch).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. BeMiller, J.; Whistler, R. (Eds.) Starch: Chemistry and Technology, 3rd ed.; Academic Press: Boston, MA, USA; Elsevier: Amsterdam, The Netherlands, 2009; pp. 193–628. [Google Scholar]
  2. Tanjung, D.A.; Jamarun, N.; Arief, S.; Aziz, H.; Isfa, A.H.R.B. Influence of LLDPE-g-MA on mechanical properties, degradation performance, and water absorption of thermoplastic sago starch blends. Indones. J. Chem. 2022, 22, 171–178. [Google Scholar] [CrossRef]
  3. Abdillah, A.A.; Charles, A.L. Characterization of a natural biodegradable edible film obtained from arrowroot starch and iota-carrageenan and application in food packaging. Int. J. Biol. Macromol. 2021, 191, 618–626. [Google Scholar] [CrossRef]
  4. Perez, E.; Lares, M. Chemical composition, mineral profile, and functional properties of canna (Canna edulis) and arrowroot (Maranta spp.) starches. Plant Foods Hum. Nutr. 2005, 60, 113–116. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, J.; Zhang, M.; Zhang, Y.; Bai, X.; Wang, C. Effects of high hydrostatic pressure on the structure and retrogradation inhibition of oat starch. Int. J. Food Sci. Technol. 2022, 57, 2113–2125. [Google Scholar] [CrossRef]
  6. Chen, Y.; Liu, Y.u; Liu, H.; Gao, Y. Stabilizing emulsions using high-amylose maize starch treated by solvothermal process. Carbohydr. Polym. 2022, 284, 119190. [Google Scholar] [CrossRef] [PubMed]
  7. Yue, Y.; Ren, B.; Zhong, K.; Wu, Y.; Bu, Q.; Gao, H. Effects of konjac glucomannan on pasting, rheological, and structural properties of low-amylose rice starch. Int. J. Food Eng. 2022, 18, 291–301. [Google Scholar] [CrossRef]
  8. Zhang, C.; Wang, Z.-J.; Liu, Q.-Q.; Qian, J.-Y.; Lim, S.-T. Improvement of pasting and gelling behaviors of waxy maize starch by partial gelatinization and freeze-thawing treatment with xanthan gum. Food Chem. 2022, 375, 131656. [Google Scholar]
  9. Mendez-Montealvo, G.; Velazquez, G.; Fonseca-Florido, H.A.; Morales-Sanchez, E.; Soler, A. Insights on the acid hydrolysis of achira (Canna edulis) starch: Crystalline and double-helical structure changes impacting functionality. LWT Food Sci. Technol. 2022, 153, 112509. [Google Scholar] [CrossRef]
  10. Xu, J.; Yang, H.; Zhang, C.; Liu, C. Optimised preparation and characterisation of lotus root starch oxidised with sodium hypochlorite (NaOCl) using response surface methodology. Czech J. Food Sci. 2022, 40, 61–68. [Google Scholar] [CrossRef]
  11. Min, C.; Ma, W.; Kuang, J.; Huang, J.; Xiong, Y.L. Textural properties, microstructure and digestibility of mungbean starch–flaxseed protein composite gels. Food Hydrocoll. 2022, 126, 107482. [Google Scholar] [CrossRef]
  12. Almeida, R.L.J.; Dos Santos Pereira, T.; De Andrade Freire, V.; Santiago, A.M.; Oliveira, H.M.L.; De Sousa Conrado, L.; De Gusmao, R.P. Influence of enzymatic hydrolysis on the properties of red rice starch. Int. J. Biol. Macromol. 2019, 141, 1210–1219. [Google Scholar] [CrossRef] [PubMed]
  13. Gao, S.S.; Liu, H.; Sun, L.J.; Cao, J.W.; Yang, J.C.; Lu, M.; Wang, M. Rheological, thermal and in vitro digestibility properties on complex of plasma modified Tartary buckwheat starches with quercetin. Food Hydrocoll. 2021, 110, 106209. [Google Scholar]
  14. Vanier, N.L.; da Rosa Zavareze, E.; Pinto, V.Z.; Klein, B.; Botelho, F.T.; Dias, A.R.G.; Cardoso Elias, M. Physicochemical, crystallinity, pasting and morphological properties of bean starch oxidised by different concentrations of sodium hypochlorite. Food Chem. 2012, 131, 1255–1262. [Google Scholar] [CrossRef]
  15. Šárka, E.; Caltová, K.; Smrčková, P.; Bleha, R.; Marek, I.; Fíla, V.; Lhotka, M. Ekologické aspekty a aplikace škrobu. Chem. Listy 2023, 117, 196–207. [Google Scholar] [CrossRef]
  16. Devi, M.B.; Deka, S.C. Physicochemical properties and structure of starches of foxnut (Euryale ferox Salisb.) from India and its application. J. Food Process Preserv. 2022, 46, e16262. [Google Scholar] [CrossRef]
  17. Puncha-arnon, S.; Puttanlek, C.; Rungsardthong, V.; Pathipanawat, W.; Uttapap, D. Changes in physicochemical properties and morphology of canna starches during rhizomal development. Carbohydr. Polym. 2007, 70, 206–217. [Google Scholar] [CrossRef]
  18. Guo, K.; Liu, T.; Xu, A.; Zhang, L.; Bian, X.; Wei, C. Structural and functional properties of starches from root tubers of white, yellow, and purple sweet potatoes. Food Hydrocoll. 2019, 89, 829–836. [Google Scholar] [CrossRef]
  19. Zarroug, Y.; Boulares, M.; Sfayhi, D.; Slimi, B.; Stiti, B.; Zaieni, K.; Nefissi, S.; Kharrat, M. Structural and physicochemical properties of Tunisian Quercus suber L. starches for custard formulation: A comparative study. Polymers 2022, 14, 556. [Google Scholar] [CrossRef]
  20. Guo, C.; Han, F.; Geng, S.; Shi, Y.; Ma, H.; Liu, B. The physicochemical properties and Pickering emulsifying capacity of acorn starch. Int. J. Biol. Macromol. 2023, 239, 124289. [Google Scholar] [CrossRef]
  21. Yao, T.; Wen, Y.; Xu, Z.; Ma, M.; Li, P.; Brennan, C.; Sui, Z.; Corke, H. Octenylsuccinylation differentially modifies the physicochemical properties and digestibility of small granule starches. Int. J. Biol. Macromol. 2020, 144, 705–714. [Google Scholar] [CrossRef]
  22. Yadav, P.; Bharathi, U.; Suruthi, K.; Bosco, S.J.D. Effect of ultrasonic modification on physiochemical, structural, functional properties and in vitro starch digestibility of Amaranthus paniculatus (Rajgeera) starch. Biomass Convers. Biomass Conv. Bioref. 2023. [Google Scholar] [CrossRef]
  23. Beyan, S.M.; Amibo, T.A.; Sundramurthy, V.P. Development of anchote (Coccinia abyssinica) starch-based edible film: Response surface modeling and interactive analysis of composition for water vapor permeability. J. Food Meas. Charact. 2022, 16, 2259–2272. [Google Scholar] [CrossRef]
  24. Bikila, A.M.; Tola, Y.B.; Esho, T.B.; Forsido, S.F.; Mijena, D.F. Starch composition and functional properties of raw and pretreated anchote (Coccinia abyssinica (Lam.) Cogn.) tuber flours dried at different temperatures. Food Sci. Nutr. 2022, 10, 645–660. [Google Scholar] [CrossRef] [PubMed]
  25. Cortés-Viguri, V.; Hernández-Rodríguez, L.; Lobato-Calleros, C.; Cuevas-Bernardino, J.C.; Hernández-Rodríguez, B.E.; Alvarez-Ramirez, J.; Vernon-Carter, E.J. Annatto (Bixa orellana L.), a potential novel starch source: Antioxidant, microstructural, functional, and digestibility properties. J. Food Meas. Charact. 2022, 16, 637–651. [Google Scholar] [CrossRef]
  26. Liu, Q.; Zhou, Y.; Fettke, J. Starch granule size and morphology of Arabidopsis thaliana starch-related mutants analyzed during diurnal rhythm and development. Molecules 2021, 26, 5859. [Google Scholar] [CrossRef]
  27. Ma, M.; Wen, Y.; Zhang, C.; Xu, Z.; Li, H.; Sui, Z.; Corke, H. Extraction and characterization of starch granule-associated surface and channel lipids from small-granule starches that affect physicochemical properties. Food Hydrocoll. 2022, 126, 107370. [Google Scholar] [CrossRef]
  28. Baeza, P.; Pasten, B.; Concha, J.; Ojeda, J. Removal of thiophene and 4,6-dimethyldibenzothiophene by adsorption on different kinds of starches. Water Air Soil Pollut. 2021, 232, 395. [Google Scholar] [CrossRef]
  29. de Souza Oliveira, E.; Lovera, M.; Pires, V.R.; da Silva Mendes, F.R.; Peixoto Maia, N.V.L.; Rodrigues, J.P.V.; do Socorro Rocha Bastos, M.; Cheng, H.N.; Biswas, A.; de Azevedoreira, R.; et al. Effect of acid catalyst on pyroconversion of breadfruit (Artocarpus altilis) starch: Physicochemical and structural properties. J. Food Process. Preserv. 2022, 46, e16408. [Google Scholar] [CrossRef]
  30. Marenco-Orozco, G.A.; Rosa, M.F.; Fernandes, F.A.N. Effects of multiple-step cold plasma processing on banana (Musa sapientum) starch-based films. Packag. Technol. Sci. 2022, 35, 589–601. [Google Scholar] [CrossRef]
  31. Chen, X.; Ma, M.; Li, X.; Xu, Z.; Zhang, C.; Sui, Z.; Corke, H. Microwave treatment alters the fine molecular structure of waxy hull-less barley starch. Int. J. Biol. Macromol. Part B 2021, 193, 1086–1092. [Google Scholar] [CrossRef]
  32. Chinma, C.E.; Abu, J.O.; Afolabi, F.H.; Nwankwo, P.; Adebo, J.A.; Oyeyinka, S.A.; Njobeh, P.B.; Adebo, O.A. Structure, in vitro starch digestibility and physicochemical properties of starch isolated from germinated Bambara groundnut. J. Food Sci. Technol. 2023, 60, 190–199. [Google Scholar] [CrossRef] [PubMed]
  33. Narayanamoorthy, S.; Zhang, C.; Xu, Z.; Ma, M.; Sui, Z.; Li, K.; Corke, H. Genetic diversity and inter-relationships of common bean (Phaseolus vulgaris L.) starch traits. Starch–Stärke 2022, 74, 2100189. [Google Scholar] [CrossRef]
  34. Hu, J.; Li, X.; Cheng, Z.; Fan, X.; Ma, Z.; Hu, X.; Wu, G.; Xing, Y. Modified Tartary buckwheat (Fagopyrum tataricum Gaertn.) starch by gaseous ozone: Structural, physicochemical and in vitro digestible properties. Food Hydrocoll. 2022, 125, 107365. [Google Scholar] [CrossRef]
  35. Vázquez-León, L.A.; Aparicio-Saguilán, A.; Martínez-Medinilla, R.M.; Utrilla-Coello, R.G.; Torruco-Uco, J.G.; Carpintero-Tepole, V.; Páramo-Calderón, D.E. Physicochemical and morphological characterization of black bean (Phaseolus vulgaris L.) starch and potential application in nano-encapsulation by spray drying. J. Food Meas. Charact. 2022, 16, 547–560. [Google Scholar] [CrossRef]
  36. He, R.; Pan, Y.-G.; Shang, W.-T.; Zong, G.; Huang, W.-Y.; Xiang, D.; Pan, F.; Zhang, W.-M. Ultrasonic-assisted binding of canistel (Lucuma nervosa A.DC) seed starch with quercetin. Ultrason. Sonochem. 2023, 96, 106417. [Google Scholar] [CrossRef]
  37. Gomes Basso Los, F.; Chezini, A.; Sztoltz Piroski, C.; Lacerda, L.G.; Nogueira, A.; Mottin Demiate, I. Evaluation of physicochemical properties of starch from Brazilian Carioca beans (Phaseolus vulgaris). Starch-Stärke 2022, 74, 2000281. [Google Scholar]
  38. Su, C.; Zhang, X.; Ge, X.; Shen, H.; Zhang, Q.; Lu, Y.; Sun, X.; Sun, Z.; Li, W. Structural, physical and degradation characteristics of polyvinyl alcohol/esterified mung bean starch/gliadin ternary composite plastic. Ind. Crops Prod. 2022, 176, 114365. [Google Scholar] [CrossRef]
  39. Shahrim, N.A.; Sarifuddin, N.; Azhar, A.Z.A.; Zaki, H.H.M. Biodegradation of mango seed starch films in soil. IIUM Eng. J. 2022, 23, 1. [Google Scholar]
  40. Shen, H.; Guo, Y.; Zhao, J.; Zhao, J.; Ge, X.; Zhang, Q.; Yan, W. The multi-scale structure and physicochemical properties of mung bean starch modified by ultrasound combined with plasma treatment. Int. J. Biol. Macromol. 2021, 191, 821–831. [Google Scholar] [CrossRef]
  41. Liu, Z.; Fu, Y.; Zhang, J.; Shen, Q. Comparison on physicochemical properties of mung bean flour and isolated starch under different level of high static pressure. Cereal Chem. 2021, 98, 1203–1214. [Google Scholar] [CrossRef]
  42. Jan, R.; Saxena, D.C.; Singh, S. Physico-chemical and textural property of starch isolated from Chenopodium (Chenopodium album) grains. Cogent Food Agric. 2015, 1, 1095052. [Google Scholar] [CrossRef]
  43. Guo, Z.; Li, X.; Yang, D.; Lei, A.; Zhang, F. Structural and functional properties of chestnut starch based on high-pressure homogenization. LWT Food Sci. Technol. 2022, 154, 112647. [Google Scholar] [CrossRef]
  44. Liu, C.; Wang, S.J.; Chang, X.D.; Wang, S. Structural and functional properties of starches from Chinese chestnuts. Food Hydrocoll. 2015, 43, 568–576. [Google Scholar] [CrossRef]
  45. Zehra, K.; Nawab, A.; Alam, F.; Hadi, A.; Raza, M. Development of novel biodegradable water chestnut starch/PVA composite film. Evaluation of plasticizer effect over physical, barrier, and mechanical properties. J. Food Process Preserv. 2022, 46, e16334. [Google Scholar] [CrossRef]
  46. Ch, C.; He, Y.; Xiao, X.; Chen, B.; Zhou, Y.; Tan, X.; Ji, Z.; Zhang, Y.; Liu, P. A novel very small granular starch from Chlorella sp. MBFJNU-17. Int. J. Biol. Macromol. 2023, 225, 557–564. [Google Scholar] [CrossRef] [PubMed]
  47. Adewale Oderinde, A.; Ibikunle, A.A.; Bakre, L.G.; Babarinde, N.A. Effects of acetylation, acid-thinning and oxidation on Chrysophyllum albidum (African Star Apple) kernel native starch. Chem. Papers 2023, 77, 5385–5394. [Google Scholar] [CrossRef]
  48. Yang, Y.; Chen, Y.; Jin, Y.; Liu, J.; Qin, X.; Liu, W.; Guo, L. Highly efficient fermentation of glycerol and 1,3-propanediol using a novel starch as feedstock. Food Biosci. 2022, 46, 101521. [Google Scholar] [CrossRef]
  49. Singh, R.; Sharanagat, V.S. Physico-functional and structural characterization of ultrasonic-assisted chemically modified elephant foot yam starch. Int. J. Biol. Macromol. 2020, 164, 1061–1069. [Google Scholar] [CrossRef]
  50. Zhang, L.; Chen, Y.; Zeng, J.; Zang, J.; Liang, Q.; Tang, D.; Wang, Z.; Yin, Z. Digestive and physicochemical properties of small granular starch from Euryale ferox seeds growing in Yugan of China. Food Biophys. 2022, 17, 126–135. [Google Scholar] [CrossRef]
  51. Ni’mah, Y.L.; Suprapto, S.; Harmami; Ulfin, I.; Fauziyah, P.A. The fabrication of water-soluble chitosan capsule shell modified by alginate and gembili starch (Dioscorea esculenta L.). J. Renew. Mater. 2022, 10, 2365–2376. [Google Scholar] [CrossRef]
  52. Wang, Y.-C.; Liang, Y.-C.; Huang, F.-L.; Chang, W.-C. Effect of freeze–thaw cycles on physicochemical and functional properties of ginger starch. Processes 2023, 11, 1828. [Google Scholar] [CrossRef]
  53. Zeng, J.-C.; Xiao, P.-J.; Ling, L.-J.; Zhang, L.; Tang, D.-B.; Zhang, Q.-F.; Chen, J.-G.; Li, J.-E.; Yin, Z.-P. Processing, digestion property and structure characterization of slowly digestible gorgon nut starch. Food Sci. Biotechnol. 2022, 31, 49–59. [Google Scholar] [CrossRef] [PubMed]
  54. Tung, N.T.; Thuy, L.T.H.; Luong, N.T.; Khoi, N.V.; Ha, P.T.T.; Thang, N.H. The molecular structural transformation of jackfruit seed starch in hydrogen peroxide oxidation condition. J. Indian Chem. Soc. 2021, 98, 100192. [Google Scholar] [CrossRef]
  55. Kushwaha, R.; Kaur, S.; Kaur, D. Potential of jackfruit (Artocarpus Heterophyllus Lam.) seed starch as an alternative to the commercial starch source–A review. Food Rev. Int. 2023, 39, 2635–2654. [Google Scholar] [CrossRef]
  56. Xu, H.; Fu, X.; Ding, Z.; Kong, H.; Ding, S. Effect of ozone and high-pressure homogenization on the physicochemical, functional, and in vitro digestibility properties of lily starch. J. Food Process Preserv. 2021, 45, e16076. [Google Scholar] [CrossRef]
  57. Zhang, Y.; Zeng, H.; Wang, Y.; Zeng, S.; Zheng, B. Structural characteristics and crystalline properties of lotus seed resistant starch and its prebiotic effects. Food Chem. 2014, 155, 311–318. [Google Scholar] [CrossRef]
  58. Rather, J.A.; Makroo, H.A.; Showkat, Q.A.; Majid, D.; Dar, B.N. Recovery of gelatin from poultry waste: Characteristics of the gelatin and lotus starch-based coating material and its application in shelf-life enhancement of fresh cherry tomato. Food Packag. Shelf Life 2022, 31, 100775. [Google Scholar] [CrossRef]
  59. Lagunes-Delgado, C.; Agama-Acevedo, E.; Patino-Rodríguez, O.; Martinez, M.M.; Bello-Pérez, L.A. Recovery of mango starch from unripe mango juice. LWT-Food Sci. Technol. 2022, 153, 112514. [Google Scholar] [CrossRef]
  60. Nayak, P.; Rayaguru, K.; Brahma, S.; Routray, W.; Dash, S.K. Standardization of process protocol for isolation of starch from mango kernel and its characterization. J. Sci. Food Agric. 2021, 102, 2813–2825. [Google Scholar] [CrossRef]
  61. Mahajan, P.; Bera, M.B.; Panesar, P.S. Structural, functional, textural characterization and in vitro digestibility of underutilized Kutki millet (Panicum sumatrense) starch. LWT-Food Sci. Technol. 2022, 154, 112831. [Google Scholar] [CrossRef]
  62. Bangar, S.P.; Sandhu, K.S.; Rusu, A.V.; Kaur, P.; Purewal, S.S.; Kaur, M.; Kaur, N.; Trif, M. Proso-millet-starch-based edible films: An innovative approach for food industries. Coatings 2021, 11, 1167. [Google Scholar] [CrossRef]
  63. Li, Y.; Qi, Y.; Li, H.; Chen, Z.; Xu, B. Improving the cold water swelling properties of oat starch by subcritical ethanol-water treatment. Int. J. Biol. Macromol. 2022, 194, 594–601. [Google Scholar] [CrossRef] [PubMed]
  64. Biswal, A.K.; Mishra, S.; Bhavy, M.B.; Samal, A.K.; Merugu, R.; Singh, M.K.; Misra, P.K. Identifcation of starch with assorted shapes derived from the fleshy root tuber of Phoenix sylvestris: Extraction, morphological and techno-functional characterization. J. Food Meas. Charact. 2022, 16, 1688–1701. [Google Scholar] [CrossRef]
  65. Chavez-Salazar, A.; Alvarez-Barreto, C.I.; Hoyos-Leyva, J.D.; Bello-Pérez, L.A.; Castellanos-Galeano, F.J. Drying processes of OSA-modified plantain starch trigger changes in its functional properties and digestibility. LWT-Food Sci. Technol. 2022, 154, 112846. [Google Scholar] [CrossRef]
  66. Yuan, T.; Ye, F.; Chen, T.; Li, M.; Zhao, G. Structural characteristics and physicochemical properties of starches from winter squash (Cucurbita maxima Duch.) and pumpkin (Cucurbita moschata Duch. ex Poir.). Food Hydrocoll. 2022, 122, 107115. [Google Scholar] [CrossRef]
  67. Li, G.; Wang, S.; Zhu, F. Physicochemical properties of quinoa starch. Carbohydr. Polym. 2016, 137, 328–338. [Google Scholar] [CrossRef]
  68. Pech-Cohuo, S.C.; Hernandez-Colula, J.; Gonzalez-Canche, N.G.; Salgado-Transito, I.; Uribe-Calderon, J.; Cervantes-Uc, J.M.; Cuevas-Bernardino, J.C.; Ayora -Talavera, T.; Pacheco, N. Starch from Ramon seed (Brosimum alicastrum) obtained by two extraction methods. MRS Adv. 2021, 6, 875–880. [Google Scholar] [CrossRef]
  69. Noor, M.H.M.; Ngadi, N.; Suhaidi, A.N.; Inuwa, I.M.; Opotu, L.A. Response surface optimization of ultrasound-assisted extraction of sago starch from sago pith waste. Starch-Stärke 2022, 74, 2100012. [Google Scholar] [CrossRef]
  70. Sudheesh, C.; Sunooj, K.V.; George, J.; Kumar, S.; Sajeevkumar, V.A. Physico-chemical, morphological, pasting and thermal properties of stem flour and starch isolated from kithul palm (Caryota urens) grown in valley of Western Ghats of India. J. Food Meas. Charact. 2019, 13, 1030. [Google Scholar] [CrossRef]
  71. Han, L.; Qiu, S.; Cao, S.; Yu, Y.; Yu, S.; Liu, Y. Molecular characteristics and physicochemical properties of very small granule starch isolated from Agriophyllum squarrosum seeds. Carbohydr. Polym. 2021, 273, 118583. [Google Scholar] [CrossRef]
  72. Mondal, D.; Kantamraju, P.; Jha, S.; Sundarrao, G.S.; Bhowmik, A.; Chakdar, H.; Mandal, S.; Sahana, N.; Roy, B.; Madhab Bhattacharya, P.; et al. Evaluation of indigenous aromatic rice cultivars from sub-Himalayan Terai region of India for nutritional attributes and blast resistance. Sci. Rep. 2021, 11, 4786. [Google Scholar] [CrossRef] [PubMed]
  73. Gani, A.; Ashwar, B.A.; Akhter, G.; Gani, A.; Shah, A.; Masoodi, F.A.; Wani, A. Resistant starch from five Himalayan rice cultivars and Horse chestnut: Extraction method optimization and characterization. Sci. Rep. 2020, 10, 4097. [Google Scholar] [CrossRef]
  74. Wu, C.; Ji, G.; Gao, F.; Qian, J.-Y.; Zhang, L.; Li, Q.; Zhang, C. Effect of heat-moisture treatment on the structural and physicochemical characteristics of sand rice (Agriophyllum squarrosum) starch. Food Sci. Nutr. 2021, 9, 6720–6727. [Google Scholar] [CrossRef]
  75. Navaf, M.; Sunooj, K.V.; Krishna, N.U.; Aaliya, B.; Sudheesh, C.; Akhila, P.P.; Sabu, S.; Sasidharan, A.; Mir, S.A.; George, J. Effect of different hydrothermal treatments on pasting, textural, and rheological properties of single and dual modified Corypha Umbraculifera L. starch. Starch-Stärke 2022, 74, 2100236. [Google Scholar] [CrossRef]
  76. Akinyosoye, T.S.; Nwokocha, L.M. Morphological, structural, and paste characteristics of native and heat-moisture treated Treculia africana starch. Starch-Stärke 2022, 74, 2100196. [Google Scholar] [CrossRef]
  77. Galindez, A.; Daz, L.D.; Homez-Jara, A.; Eim, V.S.; Váquiro, H.A. Characterization of ulluco starch and its potential for use in edible films prepared at low drying temperature. Carbohydr. Polym. 2019, 215, 143–150. [Google Scholar] [CrossRef] [PubMed]
  78. Chiotelli, E.; Le Meste, M. Effect of small and large wheat starch granules on thermomechanical behavior of starch. Cereal Chem. 2002, 79, 286–293. [Google Scholar] [CrossRef]
  79. Lindeboom, N.; Chang, P.R.; Tyler, R.T. Analytical, biochemical and physicochemical aspects of starch granule size, with emphasis on small granule starches: A review. Starch-Starke 2004, 56, 89–99. [Google Scholar] [CrossRef]
  80. Ma, M.; Zhu, H.; Liu, Z.; Sui, Z.; Corke, H. Removal of starch granule-associated proteins alters the physicochemical properties of diverse small granule starches. Food Hydrocoll. Part B 2022, 124, 107318. [Google Scholar] [CrossRef]
  81. He, W.; Wei, C. Progress in C-type starches from different plant sources. Food Hydrocoll. 2017, 73, 162–175. [Google Scholar] [CrossRef]
  82. Cardoso, G.J.; Kipp, S.D.M.; Garcia, V.A.S.; Carvalho, R.A.; Vanin, F.M. Arrowroot starch (Maranta arundinacea) as a bread ingredient for product development. J. Food Process Preserv. 2022, 46, e16251. [Google Scholar] [CrossRef]
  83. Esquivel-Fajardo, E.A.; Martinez-Ascencio, E.U.; Oseguera-Toledo, M.E.; Londono-Restrepo, S.M.; Rodriguez-García, M.E. Influence of physicochemical changes of the avocado starch throughout its pasting profile: Combined extraction. Carbohydr. Polym. 2022, 281, 119048. [Google Scholar] [CrossRef] [PubMed]
  84. Zhou, R.; Wang, Y.; Wang, Z.; Liu, K.; Wang, Q.; Bao, H. Effects of Auricularia auricula-judae polysaccharide on pasting, gelatinization, rheology, structural properties and in vitro digestibility of kidney bean starch. Int. J. Biol. Macromol. 2021, 191, 1105–1113. [Google Scholar] [CrossRef] [PubMed]
  85. Wang, N.; Dong, Y.; Dai, Y.; Zhang, H.; Hou, H.; Wang, W.; Ding, X.; Zhang, H.; Li, C. Influences of high hydrostatic pressure on structures and properties of mung bean starch and quality of cationic starch. Food Res. Int. 2023, 165, 112532. [Google Scholar] [CrossRef] [PubMed]
  86. Jacinto Almeida, R.L.; Santos, N.C.; da Silva, G.M.; Fonseca Feitoza, J.V.; de Alcântara Silva, V.M.; de Alcântara Ribeiro, V.H.; da Silva Eduardo, R.; de Sousa Muniz, C.E. Effects of hydrothermal pretreatments on thermodynamic and technological properties of red bean starch. J. Food Process. Eng. 2022, 45, e13994. [Google Scholar] [CrossRef]
  87. Liu, W.; Zhang, Y.; Wang, R.; Li, J.; Pan, W.; Zhang, X.; Xiao, W.; Wen, H.; Xie, J. Chestnut starch modification with dry heat treatment and addition of xanthan gum: Gelatinization, structural and functional properties. Food Hydrocoll. 2022, 124, 107205. [Google Scholar] [CrossRef]
  88. Liu, C.; Yan, H.; Liu, S.; Chang, X. Influence of phosphorylation and acetylation on structural, physicochemical and functional properties of chestnut starch. Polymers 2022, 14, 172. [Google Scholar] [CrossRef]
  89. Lin, Y.; Liu, L.; Li, L.; Xu, Y.; Zhang, Y.; Zeng, H. Properties and digestibility of a novel porous starch from lotus seed prepared via synergistic enzymatic treatment. Int. J. Biol. Macromol. 2022, 194, 144–152. [Google Scholar] [CrossRef]
  90. Shi, L.; Li, Y.; Lin, L.; Bian, X.; Wei, C. Effects of variety and growing location on physicochemical properties of starch from sweet potato root tuber. Molecules 2021, 26, 7137. [Google Scholar] [CrossRef]
  91. Parra, D.O.; Daza Ramírez, L.D.; Sandoval-Aldana, A.; Eim, V.S.; Váquiro, H.A. Annealing treatment of ulluco starch: Effect of moisture content and time on the physicochemical properties. J. Food Process. Preserv. 2022, 46, e16353. [Google Scholar] [CrossRef]
  92. Parra, D.O.; Eim, V.S.; Váquiro, H.A. Influence of ulluco starch concentration on the physicochemical properties of starch–chitosan biocomposite films. Polymers 2021, 13, 4232. [Google Scholar]
  93. Hanashiro, I.; Abe, J.; Huzukuri, S. A periodic distribution of the chain length of amylopectin as revealed by high-performance anion-exchange chromatography. Carbohydr. Res. 1996, 283, 151–159. [Google Scholar] [CrossRef]
  94. Tang, H.J.; Ando, H.; Watanabe, K.; Takeda, Y.; Mitsunaga, T. Physicochemical properties and structure of large, medium and small granule starches in fractions of normal barley endosperm. Carbohydr. Res. 2001, 330, 241–248. [Google Scholar] [CrossRef] [PubMed]
  95. Jane, J.; Chen, Y.Y.; Lee, L.F.; McPherson, A.E.; Wong, K.S.; Radosavljevic, M.; Kasemsuwan, T. Effects of amylopectin branch chain length and amylose content on the gelatinization and pasting properties of starch. Cereal Chem. 1999, 78, 629–637. [Google Scholar] [CrossRef]
  96. Annor, G.A.; Marcone, M.; Bertoft, E.; Seetharaman, K. Unit and internal chain profile of millet amylopectin. Cereal Chem. 2014, 91, 29–34. [Google Scholar] [CrossRef]
  97. Ratnayake, W.S.; Hoover, R.; Shahidi, F.; Perera, C.; Jane, J. Composition, molecular structure, and physicochemical properties of starches from four field pea (Pisum sativum L.) cultivars. Food Chem. 2001, 74, 189–202. [Google Scholar] [CrossRef]
  98. de Dios-Avila, N.; Tirado-Gallegos, J.M.; Rios-Velasco, C.; Luna-Esquivel, G.; Isiordia-Aquino, N.; Zamudio-Flores, P.B.; Estrada-Virgen, M.O.; Cambero-Campos, O.J. Physicochemical, structural, thermal and rheological properties of flour and starch isolated from avocado seeds of Landrace and Hass cultivars. Molecules 2022, 27, 910. [Google Scholar] [CrossRef]
  99. Li, G.; Zhu, F. Physicochemical, rheological, and emulsification properties of nonenyl succinic anhydride (NSA) modified quinoa starch. Int. J. Biol. Macromol. 2021, 193, 1371–1378. [Google Scholar] [CrossRef]
  100. Hongsprabhas, P.; Israkarn, K.; Rattanawattanaprakit, C. Architectural changes of heated mungbean, rice and cassava starch granules: Effects of hydrocolloids and protein-containing envelope. Carbohydr. Polym. 2007, 67, 614–622. [Google Scholar] [CrossRef]
  101. Hussain, S.; Mohamed, A.A.; Alamri, M.S.; Ibraheem, M.A.; Qasem, A.A.A.; Alsulami, T.; Ababtain, I.A. Effect of cactus (Opuntia ficus-indica) and acacia (Acacia seyal) gums on the pasting, thermal, textural, and rheological properties of corn, sweet potato, and Turkish bean starches. Molecules 2022, 27, 701. [Google Scholar] [CrossRef]
  102. Crosbie, G.B.; Ross, A.S. (Eds.) The RVA Handbook, 5th ed.; AACC International: St. Paul, MN, USA, 2015; pp. 138–140. [Google Scholar]
  103. Asmeda, R.; Noorlaila, A.; Norziah, M.H. Relationships of damaged starch granules and particle size distribution with pasting and thermal profiles of milled MR263 rice flour. Food Chem. 2016, 191, 45–51. [Google Scholar] [CrossRef] [PubMed]
  104. Sit, N.; Misra, S.; Deka, S.C. Physicochemical, functional, textural and colour characteristics of starches isolated from four taro cultivars of North-East India. Starch-Stärke 2013, 65, 1011–1021. [Google Scholar] [CrossRef]
  105. Vengaiah, P.C.; Murthy, G.N.; Prasad, K.R.; Kumari, K.U.; Raj, S.A. Physico-chemical and functional characteristics of palmyrah (Borassus flabellifer L.) spongy haustorium flour. J. Plantation Crops 2013, 41, 437–440. [Google Scholar]
  106. Mehfooz, T.; Ali, T.M.; Ahsan, M.; Abdullah, S.; Hasnain, A. Use of hydroxypropylated barley starch as partial casein replacer in imitation mozzarella cheese. J. Food Process Preserv. 2021, 45, e16094. [Google Scholar] [CrossRef]
  107. Photinam, R.; Moongngarm, A.; Paseephol, T. Process optimization to increase resistant starch in vermicelli prepared from mung bean and cowpea starch. Emir. J. Food Agric. 2016, 28, 449–458. [Google Scholar]
  108. Rodboontheng, W.; Uttapap, D.; Wandee, Y.; Udchumpisa, W.; Kotatha, D.; Puttanlek, C.; Rungsardthong, V. Simple thermal and freezing treatments to improve absorption capacity and alter digestibility of canna starch granules. Int. J. Biol. Macromol. 2022, 194, 861–869. [Google Scholar] [CrossRef]
  109. Mohamed, H.I.; Fawzi, E.M.; Basit, A.; Kaleemullah; Lone, R.; Sofy, M.R. Sorghum: Nutritional factors, bioactive compounds, pharmaceutical and application in food systems: A review. Phyton 2022, 91, 1303–1325. [Google Scholar] [CrossRef]
  110. Šárka, E.; Sluková, M.; Henke, S. Changes of phenolics during cooking extrusion: A review. Foods 2021, 10, 2100. [Google Scholar] [CrossRef]
  111. Escobar-Puentes, A.A.; Reyes-Lopez, S.Y.; de Jesús Ruíz Baltazar, A.; Lopez-Teros, V.; Wall-Medrano, A. Molecular interaction of β-carotene with sweet potato starch: A bleaching-restitution assay. Food Hydrocoll. 2022, 127, 107522. [Google Scholar] [CrossRef]
  112. Kulp, K. Characteristics of small-granule starch of flour and wheat. Cereal Chem. 1973, 50, 666–679. [Google Scholar]
  113. Englyst, H.N.; Anderson, V.; Cummings, J.H. Starch and non-starch polysaccharides in some cereal foods. J. Sci. Food Agric. 1983, 34, 1434–1440. [Google Scholar] [CrossRef] [PubMed]
  114. Wu, X.; Yu, H.; Bao, G.; Luan, M.; Wang, C. Preparation of adzuki bean starch-lipid complexes and their anti-digestion mechanism. J. Food Meas. Charact. 2022, 16, 945–956. [Google Scholar] [CrossRef]
  115. Warren, F.J.; Gidley, M.J.; Flanagan, B.M. Infrared spectroscopy as a tool to characterise starch ordered structure—A joint FTIR–ATR, NMR, XRD and DSC study. Carbohydr. Polym. 2016, 139, 35–42. [Google Scholar] [CrossRef] [PubMed]
  116. Pelissari, F.M.; Andrade-Mahecha, M.M.; Sobral, P.J.d.A.; Menegalli, F.C. Isolation and characterization of the flour and starch of plantain bananas (Musa paradisiaca). Starch-Stärke 2012, 64, 382–391. [Google Scholar] [CrossRef]
  117. Kizil, R.; Irudayaraj, J.; Seetharaman, K. Characterization of irradiated starches by using FT-Raman and FTIR spectroscopy. J. Agric. Food Chem. 2002, 50, 3912–3918. [Google Scholar] [CrossRef]
  118. Fan, D.; Ma, W.; Wang, L.; Huang, J.; Zhao, J.; Zhang, H.; Chen, W. Determination of structural changes in microwaved rice starch using Fourier transform infrared and Raman spectroscopy. Starch-Stärke 2012, 64, 598–606. [Google Scholar] [CrossRef]
  119. Dankar, I.; Haddarah, A.; Omar, F.E.; Pujolà, M.; Sepulcre, F. Characterization of food additive-potato starch complexes by FTIR and X-ray diffraction. Food Chem. 2018, 260, 7–12. [Google Scholar] [CrossRef]
  120. Olsson, A.-M.; Salmén, L. The association of water to cellulose and hemicellulose in paper examined by FTIR spectroscopy. Carbohydr. Res. 2004, 339, 813–818. [Google Scholar] [CrossRef]
  121. Liu, K.; Zhang, B.J.; Chen, L.; Li, X.X.; Zheng, B. Hierarchical structure and physicochemical properties of highland barley starch following heat moisture treatment. Food Chem. 2019, 271, 102–108. [Google Scholar] [CrossRef]
  122. Ren, Y.; Jiang, L.; Wang, W.; Xiao, Y.; Liu, S.; Luo, Y.; Shen, M.; Xie, J. Effects of Mesona chinensis Benth polysaccharide on physicochemical and rheological properties of sweet potato starch and its interactions. Food Hydrocoll. 2020, 99, 105371. [Google Scholar] [CrossRef]
  123. Xiao, Y.; Liu, S.; Shen, M.; Jiang, L.; Ren, Y.; Luo, Y.; Xie, J. Effect of different Mesona chinensis polysaccharides on pasting, gelation, structural properties and in vitro digestibility of tapioca starch-Mesona chinensis polysaccharides gels. Food Hydrocoll. 2020, 99, 105327. [Google Scholar] [CrossRef]
  124. Demiate, I.M.; Dupuy, N.; Huvenne, J.P.; Cereda, M.P.; Wosiacki, G. Relationship between baking behavior of modified cassava starches and starch chemical structure determined by FTIR spectroscopy. Carbohydr. Polym. 2000, 42, 149–158. [Google Scholar] [CrossRef]
  125. Fang, J.M.; Fowler, P.A.; Sayers, C.; Williams, P.A. The chemical modification of a range of starches under aqueous reaction conditions. Carbohydr. Polym. 2004, 55, 283–289. [Google Scholar] [CrossRef]
  126. Hu, X.; Shi, J.; Zhang, F.; Zou, X.; Holmes, M.; Zhang, W.; Huang, X.; Cui, X.; Xue, J. Determination of retrogradation degree in starch by Mid-infrared and Raman spectroscopy during storage. Food Anal. Methods 2017, 10, 3694–3705. [Google Scholar] [CrossRef]
  127. Flores-Morales, A.; Jiménez-Estrada, M.; Mora-Escobedo, R. Determination of the structural changes by FT-IR, Raman, and CP/MAS 13C NMR spectroscopy on retrograded starch of maize tortillas. Carbohydr. Polym. 2012, 87, 61–68. [Google Scholar] [CrossRef] [PubMed]
  128. Nobrega, M.M.; Olivato, J.B.; Müller, C.M.; Yamashita, F. Biodegradable starch-based films containing saturated fatty acids: Thermal, infrared and Raman spectroscopic characterization. Polímeros 2012, 22, 475–480. [Google Scholar] [CrossRef]
  129. Bernardino-Nicanor, A.; Acosta-García, G.; Güemes-Vera, N.; Montañez-Soto, J.L.; de los Ángeles Vivar-Vera, M.; González-Cruz, L. Fourier transform infrared and Raman spectroscopic study of the effect of the thermal treatment and extraction methods on the characteristics of ayocote bean starches. J. Food Sci. Technol. 2017, 54, 933–943. [Google Scholar] [CrossRef]
  130. Liu, Y.; Xu, Y.; Yan, Y.; Hu, D.; Yang, L.; Shen, R. Application of Raman spectroscopy in structure analysis and crystallinity calculation of corn starch. Starch-Stärke 2015, 67, 612–619. [Google Scholar] [CrossRef]
  131. De Gussem, K.; Vandenabeele, P.; Verbeken, A.; Moens, L. Raman spectroscopic study of Lactarius spores (Russulales, Fungi). Spectrochim. Acta A Mol. Biomol. Spectrosc. 2005, 61, 2896–2908. [Google Scholar] [CrossRef]
  132. Vandenabeele, P.; Wehling, B.; Moens, L.; Edwards, H.; De Reu, M.; Van Hooydonk, G. Analysis with micro-Raman spectroscopy of natural organic binding media and varnishes used in art. Anal. Chim. Acta 2000, 407, 261–274. [Google Scholar] [CrossRef]
  133. Pigorsch, E. Spectroscopic characterisation of cationic quaternary ammonium starches. Starch-Stärke 2009, 61, 129–138. [Google Scholar] [CrossRef]
  134. Phillips, D.L.; Xing, J.; Chong, C.K.; Liu, H.; Corke, H. Determination of the degree of succinylation in diverse modified starches by Raman spectroscopy. J. Agric. Food Chem. 2000, 48, 5105–5108. [Google Scholar] [CrossRef] [PubMed]
  135. Zhu, F. NMR spectroscopy of starch systems. Food Hydrocoll. 2017, 63, 611–624. [Google Scholar] [CrossRef]
  136. Abdul Hadi, N.; Wiege, B.; Stabenau, S.; Marefati, A.; Rayner, M. Comparison of three methods to determine the degree of substitution of quinoa and rice starch acetates, propionates, and butyrates: Direct stoichiometry, FTIR, and 1H-NMR. Foods 2020, 9, 83. [Google Scholar] [CrossRef]
  137. Cuenca, P.; Ferrero, S.; Albani, O. Preparation and characterization of cassava starch acetate with high substitution degree. Food Hydrocoll. 2020, 100, 105430. [Google Scholar] [CrossRef]
  138. Elomaa, M.; Asplund, T.; Soininen, P.; Laatikainen, R.; Peltonen, S.; Hyvärinen, S.; Urtti, A. Determination of the degree of substitution of acetylated starch by hydrolysis, 1H NMR and TGA/IR. Carbohydr. Polym. 2004, 57, 261–267. [Google Scholar] [CrossRef]
  139. Bai, Y.; Shi, Y.C.; Herrera, A.; Prakash, O.M. Study of octenyl succinic anhydride-modified waxy maize starch by nuclear magnetic resonance spectroscopy. Carbohydr. Polym. 2011, 83, 407–413. [Google Scholar] [CrossRef]
  140. Fan, D.; Ma, W.; Wang, L.; Huang, J.; Zhang, F.; Zhao, J.; Zhang, H.; Chen, W. Determining the effects of microwave heating on the ordered structures of rice starch by NMR. Carbohydr. Polym. 2013, 92, 1395–1401. [Google Scholar] [CrossRef]
  141. Lopez-Rubio, A.; Flanagan, B.M.; Gilbert, E.P.; Gidley, M.J. A novel approach for calculating starch crystallinity and its correlation with double helix content: A combined XRD and NMR study. Biopolymers 2008, 89, 761–768. [Google Scholar] [CrossRef]
  142. Tan, I.; Flanagan, B.M.; Halley, P.J.; Whittaker, A.K.; Gidley, M.J. A method for estimating the nature and relative proportions of amorphous, single, and double-helical components in starch granules by 13C CP/MAS NMR. Biomacromolecules 2007, 81, 885–891. [Google Scholar] [CrossRef]
  143. Riley, I.M.; Nivelle, M.A.; Ooms, N.; Delcour, J.A. The use of time domain 1H NMR to study proton dynamics in starch-rich foods: A review. Compr. Rev. Food Sci. Food Saf. 2022, 21, 4738–4775. [Google Scholar] [CrossRef] [PubMed]
  144. Biduski, B.; Amaral do Evangelho, J.; Oliveira Lima, K.; de Mello El Halal, S.L.; da Rosa Zavareze, E.; Guerra Dias, A.R.; Pinto, V.Z. Heat–moisture treatment on pinhão starch impacts the properties of the biodegradable films. Starch-Stärke 2023, 75, 2200209. [Google Scholar] [CrossRef]
  145. Zhu, F. Underutilized and unconventional starches: Why should we care? Trends Food Sci. Technol. 2020, 100, 363–373. [Google Scholar] [CrossRef]
  146. Du, S.K.; Jiang, H.; Ai, Y.; Jane, J.L. Physicochemical and functional properties of whole legume flour. Carbohydr. Polym. 2014, 108, 308–313. [Google Scholar] [CrossRef]
  147. Malinski, E.; Daniel, J.R.; Zhang, X.X.; Whistler, R.L. Isolation of small starch granules and determination of their fat mimic characteristics. Cereal Chem. 2003, 80, 1–4. [Google Scholar] [CrossRef]
  148. Šárka, E.; Bubník, Z. Morfologie, chemická struktura, vlastnosti a možnost využití pšeničného B-škrobu. Chem. Listy 2010, 104, 318–325. [Google Scholar]
  149. Šárka, E.; Krulis, Z.; Kotek, J.; Ruzek, L.; Korbarova, A.; Bubnik, Z.; Ruzkova, M. Application of wheat B-starch in biodegradable plastic materials. Czech J. Food Sci. 2011, 29, 232–242. [Google Scholar] [CrossRef]
  150. Nain, V.; Kaur, M.; Sandhu, K.S.; Thory, R.; Sinhmar, A. Development of starch nanoparticle from mango kernel in comparison with cereal, tuber, and legume starch nanoparticles: Characterization and cytotoxicity. Starch-Stärke 2022, 74, 2100252. [Google Scholar] [CrossRef]
  151. de Sousa Martinez de Freitas, A.; Bernardo da Silva, A.P.; Stieven Montagna, L.; Araujo Nogueira, I.; Kevin Carvalho, N.; Siqueira de Faria, V.; Bomfim dos Santos, N.; Lemes, A.P. Thermoplastic starch nanocomposites: Sources, production and applications—A review. J. Biomater. Sci. Polymer Ed. 2022, 33, 900–945. [Google Scholar] [CrossRef]
  152. Sahoo, P.K.; Rana, P.K. Synthesis and biodegradability of starch-g-ethyl methacrylate/sodium acrylate/sodium silicate superabsorbing composite. J. Mater. Sci. 2006, 41, 6470–6475. [Google Scholar] [CrossRef]
  153. Šárka, E.; Kruliš, Z.; Kotek, J.; Růžek, L.; Koláček, J.; Hrušková, K.; Bubník, Z. Biodegradabilní kompozitní materiály na bázi pšeničného B-škrobu s upotřebením v zemědělství. Listy Cukrovar. Řep. 2011, 127, 402–405. [Google Scholar]
  154. Tarique, J.; Zainudin, E.S.; Sapuan, S.M.; Ilyas, R.A.; Khalina, A. Physical, mechanical, and morphological performances of arrowroot (Maranta arundinacea) fiber reinforced arrowroot starch biopolymer composites. Polymers 2022, 14, 388. [Google Scholar] [CrossRef] [PubMed]
  155. Wang, X.L.; Yang, K.K.; Wang, Y.Y. Properties of starch blends with biodegradable biopolymers. J. Macromol. Sci. Part C 2003, 43, 385–409. [Google Scholar] [CrossRef]
  156. Kierulf, A.; Whaley, J.; Liu, W.C.; Enayati, M.; Tan, C.; Perez-Herrera, M.; You, Z.; Abbaspourrad, A. Protein content of amaranth and quinoa starch plays a key role in their ability as Pickering emulsifiers. Food Chem. 2020, 315, 126246. [Google Scholar] [CrossRef]
  157. Pagno, C.H.; Costa, T.M.; de Menezes, E.W.; Benvenutti, E.V.; Hertz, P.F.; Matte, C.R.; Tosati, J.V.; Monteiro, A.R.; Rios, A.O.; Flores, S.H. Development of active biofilms of quinoa (Chenopodium quinoa W.) starch containing gold nanoparticles and evaluation of antimicrobial activity. Food Chem. 2015, 173, 755–762. [Google Scholar] [CrossRef]
Table 1. Laboratory isolation of non-traditional starches.
Table 1. Laboratory isolation of non-traditional starches.
PlantStep 1Step 2Reference
Achira (Canna edulis Ker.)/rhizomes; sweet potato/tubersCutting, extraction in waterSieving, washing[17,18]
Acorn (Quercus suber L.)/kernelsAcorn flourWater and alkaline extraction[19]
Acorn/kernelsShelling, drying, grindingSoaking in n-hexane 6 h, in ethanol 30 min, centrifugation, washing[20]
Amaranth/seeds; oat/seedsSoaking water 24 hSieving, centrifugation[21]
Amaranth/grainsSoaking 87.5 mM NaOH, washing, grindingSieving, centrifugation[22]
Anchote/seedsGrindingExtraction 0.075% (w/v) Na2S2O5[23]
Anchote/tubersPeeling, slicingDrying, grinding[24]
Annatto seedsDepigmented with KOHDrying, grinding[25]
Arabidopsis thaliana/leavesFreezing, grindingExtraction pH 7.4 + EDTA, centrifugation[26]
Avocado/fruitPeeling, cuttingGrinding, pasting[27]
Banana/fruit; breadfruit (Artocarpus altilis)Peeling, cuttingMacerated, sieving[28,29]
Banana/fruitExtraction in 4% (w/v) NaHSO3 4 hCentrifugation, drying[30]
Barley/hull-less kernelsMicrowave oven, extraction in 0.1 M NaOH 12 hWashing, grinding[31]
Bambara groundnut/germinated seedsSteeping in water 6 h, hydration-germination 24, 48, 72 h, grindingSieving, extraction in 0.3% (w/v) NaOH 4 h, decantation, sieving[32]
Bean (Phaseolus vulgaris L.)/seeds; Tartary buckwheat/flourExtraction in 0.15 or 0.20% NaOH 16 hSieving, washing[33,34]
Black bean/seedsExtraction in water 12 hGrinding, sieving[35]
Canistel/seedsDrying, dehulled, grindingAddition of n-hexane, washing, centrifugation[36]
Carioca beansExtraction in 0.16% (w/v) Na2S2O5 8 hSieving, centrifugation[37]
Mung bean/seeds; mango/flourSoaking 6 h, mixingSieving, washing[38,39]
Mung bean/seedsSoaking 4 h, mixingSieving, washing[40]
Mung bean/seedsWet millingWashing 0.2% NaOH, centrifugation[41]
Chenopodium album/grainsDrying at 40 °C 24 h, grindingExtraction 0.25% (w/v) NaOH 24 h, wet grinding, filtration[42]
Chinese chestnut (Castanea mollissima BL.)/kernelsPeeling, wet-millingFiltration, decantation[43,44]
Water chestnut (Trapanatans L. var. bispinosa)Drying 24 h, grinding, extraction in 0.2% NaOH, pH 9Filtration, decantation[45]
Chlorella sp. MBFJNU-17; quinoa/seedsSteeping in 0.45% (w/v) Na2S2O5 12 hGrinding, sieving, centrifugation[46]
Chrysophyllum albidum (African Star Apple)/kernelsSoaking in n-hexane 72 h, extraction in 0.05 M NaOH, pH 8Centrifugation, washing[47]
Duckweed (Landoltia punctata)Wet milling 0.1% NaOH, incubation 4 hDecantation, neutralization, washing[48]
Elephant foot yam starch (Amorphophallus paeoniifolius)/tubers Peeling, milling, Washing, filtration[49]
Euryale ferox/seedsMilling, extraction in water 5 hSieving, centrifugation[50]
Foxnut (Euryale ferox Salisb.)/kernelsDrying 60 °C, 24 h, extraction in water 5 h,Filtration, washing [16]
Gembili (Dioscorea esculenta L.)/tubersPeeling, cutting, mixingDecantation[51]
Ginger (Zingiber officinale Roscoe.) starch/waste productDispersed in water, stirringWashing[52]
Gorgon nut/seedExtraction in 0.20% Na2SO3 5 h, washingCrushing, filtration, centrifugation[53]
Jackfruit (Artocarpus heterophyllus Lam.)/seedSoaking 2% NaOH 0.5 h, peeling, Wet milling 0.1% NaHSO3, filtration, decantation[54]
Jackfruit/seedsSoaking water (6 h and 8 h), stirringFiltration, washing[55]
Lily (Lilium spp.)/bulbHomogenization Filtration, washing[56]
Lotus/seedHomogenization with waterFiltration, washing[57]
Lotus/stemCutting, extraction in 0.12% K2S2O5 + 0.25% citric acid, 1 hGrinding, filtration, washing[58]
Unripe mango (Mangifera indica L.) of Haden and Palmer cultivars/pulpCutting, drying, millingSieving, washing[59]
Mango/kernelsWet millingSieving, washing[60]
Dehusked Kutki millet/grainsGrinding, extraction in 0.25% NaOH 24 h Sieving, washing, centrifugation[61]
Proso millet (Panicum miliaceum L.)/grainsExtraction in 0.15% Na2S2O5 15 hWashing, grinding, sieving [62]
Oat/seedsMechanically flaked, defatted, grindingIncubation in water + protease 24 h, 40 °C, centrifugation[63]
Phoenix sylvestris/root tuberCutting, wet grindingFiltration, decantation, washing[64]
Unripe plantains (Musa sp. AAB sub-group Plantain cv. Dominico
Harton)
Peeling, cutting, extraction in 0.30% (w/v) Na2S2O5Mixing, sieving, washing[65]
Pumpkin (Cucurbita moschata
Duch. ex Poir.)/fruit; winter squash (Cucurbita maxima Duch.)/fruit
Cutting, mixingSieving, centrifugation[66]
Quinoa/seedsSoaking water 24 hMixing with 62.5 mM NaOH, centrifugation[21]
Quinoa/seedsFreezing by liquid nitrogen, grinding, mixing with 12.5 mM borate buffer (pH 10) + 0.5% SDS + 0.5% Na2S2O5Centrifugation, washing, filtration[67]
Ramon (Brosimum alicastrum)/seedsSoaking water 24 h, mixingFiltration, centrifugation[68]
Ramon (Brosimum alicastrum)/seedsSoaking 1% (w/w) NaOH 24 h, washing in 0.1% (w/v) NaOHNeutralization, centrifuging, washing[68]
Sago pith waste fiberUltrasonicationFiltration, washing[69]
Sago-kithul palm (Caryota urens)/flourSteeping in 0.50 M NaOH 30 min, centrifugationWashing, neutralization, filtration[70]
Sand rice (Agriophyllum
squarrosum)/seeds
Dehulled seedsAlkaline extraction 12 h [71,72,73,74]
Talipot palm (Corypha umbraculifera L.)/flourAlkali (0.05 M NaOH) method?[75]
Treculia africana/seedsMilling, dispersion in 0.2% NaOHSieving, washing[76]
Ulluco (Ullucus tuberosus Caldas) tubersCutting, homogenization in waterFiltration, washing[77]
Table 2. Shape, size, and crystallinity of starch granules.
Table 2. Shape, size, and crystallinity of starch granules.
Starch fromShapeSizeCrystallinity TypeReflections of 2θReference
Acorn (Quercus suber L.)/kernelsSpherical, oval5–12 µmC15.3° and 23°, 17.4° and 19.4°[19]
Acorn/kernelsUnimodal modeMedian
8.4 μm
A15°, 17°, 23°[20]
Adzuki beans (Vigna angularis L.)Spherical, oval13–80 μm, average 40.8 μmC-[81]
AchiraSmooth oval or elliptical -B15°, 17°, 20°, 22°, 24°[9]
Achira (Canna edulis Ker.)/rhizomesOval, disc10–110 μm (average size 47.1 μm)B5.6°, 17°,
22°, 24°
[17]
Amaranth/seedsPolygonal with sharp edges1–5 µm, d50 1.7 µmA15.1°, 17.2°, 18.1°, 23.2°[21]
Amaranth/seeds-1–4 µm--[80]
Amaranth/seeds-Bimodal, 1–5 µm; d50 1.7 µm--[27]
Amaranth/grains--A15.5°, 17°, 18.2°, 23.2°, 20.4°, 27.1°.[22]
Annatto/seedsOval (predominant), spherical, triangular, and irregular Bimodal, 13–197 µm, D3,2 27–31 µmA15.1°, 17–18°, 19.9°, 22.7°[25]
Arabidopsis thaliana/leavesDiscoid0.8–5.0 µm--[26]
Arrowroot starch--A15.15°, 17°, 17.95°[3]
Arrowroot starch-22 × 27 µm--[4]
Arrowroot starchRound and polygonal particles, with jagged edges10–20 µmA15.5°, 17°, 18°, 23°[82]
Avocado/fruitLenticular, spherical, irregular 5–30 µm × 4–18 µmC [83]
Bambara groundnut/germinated seedsSmooth oval shaped, some granules irregular and kidney shaped7–38 µm, average 18–24 µmA15°, 17°, 18°, 24° [32]
Bean (Phaseolus vulgaris L.)/seeds-D [4,3] 23.8–32.3 µm--[33]
Black bean/seedsPolyhedral, rounded Bimodal, peaks at 6.3 µm and 1 µmA15°, 17°, 23°[35]
Carioca beans/seedsOval---[37]
Kidney bean-D [4,3] 56.2 µm--[84]
Mung bean/seedsSpherical, oval, kidney -A-[40]
Mung bean/seedsKidney, round10–27 µmC15.20°, 17.21°, 17.92°, 23.20°[41]
Mung bean/starch--C15.3°, 17.3°, 23.5°[85]
Red bean/seeds--A15°, 16.89°, 17.74°, 19.29°, 22.45°[86]
Breadfruit (Artocarpus altilis)Spherical5.7–8.5 μm--[29]
Tartary buckwheat/flourMost polygonal, some spherical and irregular-A15.1°, 17.2°, 18.0°, 23.0°[34]
Canistel/seeds D [4,3] 69.1 µm, d50 38.4 µmA15.1°, 17.0°, 18.0°, 23.0°[36]
Chenopodium album/grainsPolygonal, angular0.6–182.4 μm, median 13.2 μm--[42]
Chestnut (Castanea mollissima BL.)/kernelsRound-C5.6°, 15.1°, 17.0°, 23.0°[43]
Chestnut--C5°, 15°, 17°, 23°[87]
Chinese chestnut (Castanea mollissima BL.) var. Yanshan Zaofeng/fruitRound, irregular, triangula-C5.6°, 15.1°, 17.0°, 23.0°[88]
Chlorella sp. MBFJNU-17 0.4–1.3 μm, average
1.0 μm
A15°, 17°, 18°, 23°[46]
Duckweed (Landoltia punctata)-1–10 μm, average 5 μm--[48]
Euryale ferox/seedsIrregular polyhedron0.5–5.6 μm, mean 2.3 μmA15.4°, 17.4°, 18.5°, 23.3°[50]
Foxnut (Euryale ferox Salisb.)/kernelsPolyhedral, angular, oval1–3 μmA15°, 23°, 17°, 18°[16]
Ginger (Zingiber officinale Roscoe.) starch/waste productOblong shape with a smooth surfaceAverage
14 μm
C10°, 11°, 15°, 17°, 18°, 23°[52]
Gorgon nut/seedRegular polyhedrons1–3 μmA15.4°, 17.2°, 18.2°, 23.3°[53]
Jackfruit (Artocarpus heterophyllus Lam.)/seed--A15.4°, 17.1°, 18.2°, 23.3°[54]
Jackfruit/seedsRound, bell4.2–13 μmA15°, 17°, 23°[55]
Lily (Lilium spp.)/bulbRound and ellipticalAverage 20.5 µmB17.0°, 19.5°, 22.2°, 23.6°,[56]
Lotus/seed-7.4–18.3 µmC15.10°, 17.10°, 18.00°, 23.10°[89]
Unripe mango (Mangifera indica L.) of Haden and Palmer cultivars/pulpBimodalMain peaks 5 μm and 17 μmC5°, 11°, 15°, 17°, 18°, 23° [59]
Mango/kernelsOval, spherical, elliptical and oblong A14.6°, 23.4°, 18.1° [60]
Dehusked Kutki millet/grainsLarge polygonal and sphericalMedian 8.1 µmA15°, 17° 18°, 23°[61]
Proso millet (Panicum miliaceum L.)/grains--A15.2°, 17.4°, 18.1°, 20.2°, 23.0°[62]
Oat/seeds Bimodal, 3–25 µm, d50 8.6 µm--[27]
Oat/seedsOval, irregular5 µmA15.1°, 17.2°, 17.6°, 20.1°, 22.9 [63]
Oat/seeds Bimodal, 1–200 µm--[80]
Oat/seeds A15.0°, 17.2°, 17.9°, 19.8°, 23.2[5]
Phoenix sylvestris/root tubersOval, spherical, flower-shaped-1 to 10 µmB15.09°, 17.36°, 23.32°[64]
Unripe plantains (Musa sp. AAB sub-group Plantain cv. Dominico
Harton)
Elongated20–30 μm wide, 50–65 μm long,CB5.4°, 15°, 17.1°, 20.4°, 23.1°[65]
Pumpkin (Cucurbita moschata
Duch. ex Poir.)/fruit
Dome-shapedD [4,3] 22 μmB5.6°, 15°, 17.2°, 19.8°, 22.3°, 24.0°[66]
Quinoa/seedsPolyhedral and angular1–2 µm, D [4,3] 1.3 µmA15.8°, 17.1°, 18.3°, 23.3°[71]
Quinoa/seedsPolygonal with sharp edges1–36 µm, d50 3 µmA15.1°, 17.2°, 18.1°, 23.2°[21]
Quinoa/seeds 0.5–1.2 µm average
1.1 µm
A15°, 17°, 18°, 23°[46]
Quinoa/seeds-Bimodal, 1–40 µm, d50 4.5 µm--[27]
Quinoa/seeds-Monomodal 0.7–8 μm--[80]
ramon (Brosimum alicastrum)/seedsSphericalAverage diameter 7–8 μmC15°, 17°, 18°, 20°, 23°, 26°,[68]
Sago pith waste fibreTemple bell-like oval 13–36 μm, mean diameter 28 μmC15.1°, 17.1°, 18.0°, 23.0°[69]
Sago-kithul palm (Caryota urens)/flourOval12–62 µm A15.1°, 17.1°, 17.9°, 23°[70]
Sand rice (Agriophyllum
squarrosum)/seeds
SphericalMostly < 1 μmA15.1°, 17.2°, 18.0°, 23.1° [74]
Sand rice (Agriophyllum
squarrosum)/seeds
Smooth spherical1–2 µm, D [4,3] 1.1–1.3 µmA15.8°, 17.1°, 18.3°, 23.3°[71]
Sweet potato/tubers-Bimodal; 0.8–3 µm and 6–40 µm--[90]
Sweet potato/tubers-Bimodal; 0.5–3 μm and 4–50 μmCA-[18]
Talipot palm (Corypha umbraculifera L.)/flour--A15.1°, 23.2°, doublet at 17.2°, 18.1°, 15.1°[75]
Treculia africana/seedsIrregularAverage diameter 20.6 µmA15.5°, 17.4°, 18.0°, 23.0° [76]
Ulluco (Ullucus tuberosus Caldas)/tubersOval and conical-B5.6°, 15°, 17°, 20°, 22°, 24°[77]
Ulluco (Ullucus tuberosus Caldas)/tubersEllipsoid, oval, smooth, irregular---[91]
Ulluco (Ullucus tuberosus Caldas)/tubersElongated and rounded-B5.6°, 15°, 17°, 20°, 22°, 24°[92]
Winter squash (Cucurbita maxima Duch.)/fruitPolyhedral, irregularD [4,3] 33–45 µmB5.6°, 15°, 17.2°, 19.8°, 22.3°, 24.0°[66]
Table 3. Chain lengths in amylopectin.
Table 3. Chain lengths in amylopectin.
Starch fromPolydispersity Index of Starch (Mw/Mn)Chain Length Distributions in Amylopectin (%)Reference
DP 6–12 DP 13–24 DP 25–36 DP > 37
Normal amaranth/seeds -28.649.912.69.0[27]
Waxy amaranth-25.147.712.514.7[95]
Waxy hull-less barley/kernel-27.251.414.86.6[31]
Normal barley-20.848.917.712.6[95]
Canna/green leaf-11.745.316.226.8[95]
Chlorella sp. MBFJNU-17 23.426.013.631.9[46]
Green banana-16.846.312.924.0[95]
Lotus/root-16.447.215.421.0[95]
High-amylose corn starch-8.5–9.740.7–43.920.3–21.326.1–29.5[95]
Normal corn-20.150.217.612.1[80]
Normal corn-17.947.914.919.3[95]
Waxy corn-21.350.517.111.1[27]
Waxy corn-17.049.417.116.5[95]
Cat tail millet-20.253.812.713.3[95]
Mung bean/seeds-22.748.717.511.7[40]
Mung bean-15.647.618.318.5[95]
Oat/seeds-26.847.416.98.9[27]
Pumpkin, var. Miben2.121.447.315.314.1[66]
Quinoa/seeds18.230.049.213.87.0[71]
Quinoa/seeds-24.622.050.014.7[27]
Quinoa/seeds 26.524.912.432.3[46]
Normal rice/seeds-19.052.212.316.5[95]
Sweet rice/seeds-23.548.713.714.0[95]
Waxy rice/seeds-20.452.714.612.3[27]
Waxy rice/seeds-27.453.412.66.6[95]
Sand rice (Agriophyllum
squarrosum)/seeds
1.3–5.830.8–39.845.9–53.19.4–11.82.0–4.9[71]
Tapioca-17.340.415.626.7[95]
Chinese taro-18.848.714.817.7[95]
Water chestnut-17.843.715.323.2[95]
Winter squash, var. Yinli2.223.248.514.913.5[66]
Winter squash, var. Heili2.024.149.314.012.6[66]
Table 4. Thermal data of non-traditional starches.
Table 4. Thermal data of non-traditional starches.
Starch fromTo (°C) Tp (°C)Tc (°C)Tc–To (°C)ΔHgel (J/g)ΔHret (J/g)Reference
Acorn/fruits 41.7–45.278.4–84.7117.4–118.973.7–76.317.7–23.4 [19]
Adzuki bean51.473.390.138.78.1 [80]
Avocado/seeds66.7–69.370.3–73.676.6–79.59.2–10.76.6–13.4 [98]
Achira starch59.2 65.273.814.610.9 [9]
Amaranth/grains59.468.482.322.917.4 [22]
Arrowroot starch67.873.681.413.610.3 [4]
Waxy amaranth/seeds62.469.375.913.510.2 [80]
Normal barley52.0–61.458.1–65.362.7–74.4-6.8–14.2 [1]
Waxy barley54.5–61.361.8–65.573.8–81.8-9.9–13.1 [1]
Amylose-barley53.0–61.062.0–68.474.3–79.5-6.6–12.2 [1]
Bean (Phaseolus vulgaris L.)/seeds54.6–69.464.7–76.874.1–81.3-3.8–16.35.8–9.0[33]
Canna starch59.063.672.113.19.4 [4]
Chestnut (Castanea mollissima BL.)/kernels56.0 66.781.325.317.3 [43]
Chinese chestnut (Castanea mollissima BL.) var. Yanshan Zaofeng/fruit57.4 65.675.117.79.6 [88]
Chlorella sp. MBFJNU-1763.066.374.111.1- [46]
Chrysophyllum albidum (African Star Apple)/kernel75.880.986.110.315.6 [47]
Carioca beans69.7–70.775.0–76.679.9–81.010.1–13.87.8–9.14.5–4.7[37]
Euryale ferox/seeds84.7110.3118.133.41576.0 [50]
Ginger (Zingiber officinale Roscoe.) starch/waste product70.7-78.17.415.2 [52]
Bambara groundnut/germinated seeds70.972.885.614.73.6 [32]
Jackfruit/seeds72.6–84.977.7–88.085.7–96.6-7.9–16.6 [55]
Lily (Lilium spp.)/bulb61.3 64.572.711.415.7 [56]
Normal corn starch64.069.774.410.412.8 [80]
Waxy corn starch63.971.476.812.912.9 [80]
Mango/kernels74.0 117.5182.1108.1148.6 [60]
Mung bean/seeds57.6 67.777.720.17.8 [40]
Mung bean/seeds58.963.585.626.710.9 [41]
Mung bean/starch58.664.673.014.410.5 [85]
Oat/seeds55.560.665.710.28.0 [80]
Oat/seeds51.1–69.557.8–66.8--8.6–14.6 [1]
Quinoa/seeds56.3 64.574.017.716.16.2[71]
Quinoa/seeds52.761.269.516.89.4 [80]
Quinoa/seeds61.1 66.373.912.814.6 [99]
Quinoa/seeds54.860.167.312.58.1 [46]
Phoenix sylvestris/root tuber68.2 82.395.026.811.31.6–2.9[64]
Ramon (Brosimum alicastrum)/seeds70.876.281.410.612.6 [68]
Sago pith waste fiber142.9149.6167.124.2270.2 [69]
Sand rice (Agriophyllum
squarrosum)/seeds
70.975.179.38.410.9 [74]
Sand rice (Agriophyllum
squarrosum)/seeds
67.5–68.172.8–74.281.4–84.113.8–16.319.8–24.711.6–11.9[71]
Sweet potato/tubers51.1–62.961.2–77.277.3–84.017.9–28.111.7–14.3 [90]
Pumpkin (Cucurbita moschata Duch. ex Poir.) cultivar Miben/fruit66.971.678.012.8–32.216.89 [66]
Waxy rice/seeds74.880.088.811.123.9 [80]
Winter squash. var. Heili/fruit61.364.871.08.510.4 [66]
Winter squash. var. Yinli/fruit65.367.973.81413.9 [66]
To—onset temperature, Tp—peak temperature, Tc—conclusion temperature, Tc–To—gelatinization range, ΔHgel—enthalpy change during gelatinization, and ΔHret—enthalpy change during gelatinization.
Table 5. Pasting properties of starches.
Table 5. Pasting properties of starches.
Starch fromStarch ConcentrationPV (mPa.s)BD (mPa.s)FV (mPa.s)SB (mPa.s)Peak Time (min)Pasting Temperature (°C)Reference
Acorn/kernels12% (w/v)3410116035571307-74.6[20]
Waxy amaranth/seeds7% (db w/w)84229563992--[80]
Avocado/fruit16.67% (w/w)21820-6791---[83]
Black bean/seeds10% (w/v)1981-1753---[35]
Turkish beans10.71% (w/w)4430-52592592-71.8[99]
Bean (Phaseolus vulgaris L.)/seeds3% (db w/w)667–47301–21651144–5170437–2667--[33]
Carioca beans8% (w/w) 1231–221880–3461790–3556640–1465-78.4–82.6[37]
Tartary buckwheat/flour10.71% (w/w)339595241291671 75.9[34]
Chestnut-52819651142--[87]
Chinese chestnut (Castanea mollissima BL.) var. Yanshan Zaofeng/fruit12% db (w/v)17964452260909 74.2[88]
Chrysophyllum albidum (African Star Apple)/kernel12% (w/w)2959526363712045.981.6[47]
Euryale ferox/seeds-54481656527414823.982.3[50]
Foxnut (Euryale ferox Salisb.)/kernels10.71% (w/w)1953–273452–2492422–3078521–593 81.8–88.7[16]
Ginger (Zingiber officinale Roscoe.) starch/waste product10% (w/w)15501031101560-89.6[52]
Bambara groundnut/germinated seeds 4538–4813446–4855512–58911420–15637.1–7.487.1–88.2[32]
Jackfruit/seeds-840–3800110–13001350–5200550–2400-64.2–91.3[55]
Normal corn starch7% (db w/w)11661621137234--[80]
Normal corn starch10.71% (w/w)2872-26971038-75.5[99]
Normal corn starch10.71% (w/w)2289-8482763--[61]
Waxy corn/starch7% (db w/w)191210401011139--[80]
Mango/kernels8% (w/w)11988521391026-84.5[60]
Kutki millet/grain10.71% (w/w)2580142630471893 [61]
Mung bean/seeds10.71% (w/w)5167226043551425-75.1[40]
Oat/seeds7% (db w/w)120748820391319--[80]
Phoenix sylvestris/root tuber10% db (w/v)1824-2887- 71.1[64]
Pumpkin (Cucurbita moschata Duch. ex Poir.) cultivar Miben/fruit10% db (w/v)4468135240759594.475.2[66]
Quinoa/seeds7% (db w/w)909137965193--[80]
Quinoa/seeds10.71% (w/w)30612125303821024.068.7[71]
Quinoa/seeds10% (db w/v)4800150055002100 69.0[99]
Sand rice (Agriophyllum
squarrosum)/seeds
6% (db w/w) 245888538642853 75.5[74]
Sand rice (Agriophyllum
squarrosum)/seeds
10.71% (w/w)1474–1601200–5223033–34261868–23384.7–5.578.3–79.2[71]
Sweet potato/tubers10% (w/v)4047-2977853-76.7[99]
Sweet potato/tubers(10% w/v)3918–49291479–25912769–3579541–7694.1–4.670.1–78.7[90]
Ulluco (Ullucus tuberosus Caldas)/tubers7% (db w/w)8600-3470570-62.8[77]
Waxy rice/seeds10% (db w/w)14075991018210--[80]
Winter squash. var. Heili/fruit10% (db w/w)4690430607318145.870.6[66]
Winter squash. Var. Yinli/fruit10% (w/v)6266239046647874.372.3[66]
PV—peak viscosity, FV—final viscosity, BD—breakdown, and SB—setback.
Table 6. Swelling power, solubility, and water absorption.
Table 6. Swelling power, solubility, and water absorption.
Starch fromP (g/g)SW (%)WAI (g/g)Reference
60 °C80 °C60 °C80 °C60 °C80 °C
Acorn/fruits3.010.10.25.4--[19]
Acorn/kernels31226--[20]
Anchote tuber5.04.613.213.83.93.9[24]
Annatto seeds6.218.03.815.0--[25]
Avocado/seeds3.5–6.57.0–10.50.5–17.55.5–17.03.5–4.86.0–9.8[98]
Arrowroot starch-7.5-7.0-2.3[82]
Black bean/seeds412448--[35]
Canistel/seeds2.7102.69.0--[36]
Canna starch14.432.213.320.013.431.2[4]
Euryale ferox/seeds--19--[50]
Foxnut (Euryale ferox Salisb.)/kernels0.84.3–5.80.32–2.7--[16]
Ginger (Zingiber officinale Roscoe.) starch/waste product-4.5-3.7--[52]
Jackfruit/seeds3–67.5–134.2–6.28.2–13.8--[55]
Lily (Lilium spp.)/bulb-8.6-6.6--[56]
Normal corn starch4.19.7----[61]
Mango/kernels8.6-6.6---[60]
Kutki millet/grain3.88.3----[61]
Mung bean/seeds2.99.31.08.1--[40]
Phoenix sylvestris/root tuber8152232--[64]
Unripe plantains (Musa sp. AAB)4.4-1.6-4.2-[65]
Quinoa/seeds1021422--[71]
Sand rice (Agriophyllum
squarrosum)/seeds
2.513–1517–8--[71]
Treculia africana/seeds3.912----[76]
Table 7. Rapidly digestible, slowly digestible, and resistant starch.
Table 7. Rapidly digestible, slowly digestible, and resistant starch.
Starch fromRDS (%)SDS (%)RS (%)Reference
Acorn/kernels16.2, 69.6 *37.6, 14.5 *46.2, 15.9 *[20]
Adzuki beans (Vigna angularis L.)27.630.640.8[80]
Amaranth/grains57.332.89.9[22]
Annatto seeds233714[25]
Waxy hull-less barley55.02.043.0[31]
Kidney bean *82.24.513.3[84]
Carioca beans4.2–5.26.5–11.883.0–88.7[37]
Tartary buckwheat/flour57.330.112.6[34]
Chestnut21.121.357.6[87]
Chinese chestnut (Castanea mollissima BL.) var. Yanshan Zaofeng/fruit8.233.658.2[88]
Chlorella sp. MBFJNU-1748.8, 88.0 *8.8, 3.5 *42.4, 8.5 *[46]
Euryale ferox/seeds26.0 37.536.5[50]
Foxnut (Euryale ferox Salisb.)/kernels8.4–14.0, 79.1–83.5 *22.8–26.4,10.6–11.6 *59.6–68.8, 4.9–10.3 *[16]
Ginger (Zingiber officinale Roscoe.) starch/waste product1.8, 39.85 *11.7, 23.01 *63.1, 8.50 *[52]
Gorgon nut/seed-47.0–60.1 *-[53]
Bambara groundnut/germinated seeds12.1–12.914.7–15.872.0–73.2[32]
Lily (Lilium spp.)/bulb1.8, 20.3 *8.2, 27.5 *90.0, 52.2 *[56]
Lotus/seed18.516.065.5[89]
Unripe plantains (Musa sp. AAB)1.2, 58.8 *8.7, 15.9 *90.0, 25.2 *[65]
Pumpkin, var. Miben2.5, 78.1 *4.9,3.1 *92.7,18.8 *[66]
Quinoa/seeds72 *20 *8 *[71]
Quinoa/seeds65.6, 85.0 *1.9, 6.5 *32.5, 8.5 *[46]
Himalayan rice/seeds--0.4–2.3[72]
Himalayan rice/seeds 85.4–92.8[73]
Sand rice (Agriophyllum
squarrosum)/seeds *
66–70248–12[71]
Sand rice (Agriophyllum
squarrosum)/seeds
89.34.18.6[74]
Sweet potato/tubers52.96.638[111]
Sweet potato/tubers2.2–1.9, 70.5–81.4 *8.8–30.9, 2.1–11.0 *58.2–89.1, 10.8–23.3 *[90]
Winter squash. var. Heili/fruit4.5, 90.5 *9.6, 1.3 *85.9, 8.2 *[66]
Winter squash. var. Yinli/fruit2.6, 84.2 *4.9, 4.1 *92.8, 11.6 *[66]
* cooked.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Šárka, E.; Sinica, A.; Smrčková, P.; Sluková, M. Non-Traditional Starches, Their Properties, and Applications. Foods 2023, 12, 3794. https://doi.org/10.3390/foods12203794

AMA Style

Šárka E, Sinica A, Smrčková P, Sluková M. Non-Traditional Starches, Their Properties, and Applications. Foods. 2023; 12(20):3794. https://doi.org/10.3390/foods12203794

Chicago/Turabian Style

Šárka, Evžen, Andrej Sinica, Petra Smrčková, and Marcela Sluková. 2023. "Non-Traditional Starches, Their Properties, and Applications" Foods 12, no. 20: 3794. https://doi.org/10.3390/foods12203794

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop