Starch Modifications Outside the Plant System: Comparison
Please note this is a comparison between Version 2 by Jason Zhu and Version 4 by Ardha Apriyanto.

Starch has been a convenient, economically important polymer with substantial applications in the food and processing industry. However, native starches present restricted applications, which hinder their industrial usage. Therefore, modification of starch is carried out to augment the positive characteristics and eliminate the limitations of the native starches. Modifications of starch can result in generating novel polymers with numerous functional and value-added properties that suit the needs of the industry. 

  • starch
  • starch modification
  • physical modification
  • chemical modification
  • enzymatic modification
  • combination modification

1. Chemical Modifications of Isolated Starches

In comparison to physical or enzymatic modifications, chemical techniques enable a much wider spectrum for the functionalization of starch and expand the field of applications. Chemical modifications are non-destructive, and the hydroxyl groups of starches serve as active sites for the introduction of functional groups such as, e.g., acetyl-, carboxyl-, and ethyl-groups which act as electrophilic reactants. However, the rate and efficiency of modification depends on several factors such as starch source, amylose-to-amylopectin ratio, the granule morphology, reaction conditions and the nature and amount of the modifying reagent [1]. The modifications can be categorized into the following classes: cross-linking, esterification, etherification, oxidation, grafting, and acid hydrolysis.
Cross-linking is the most frequently applied method for changing starch properties and is used in the pharmaceutical and food industries, wastewater processing, and bioplastic production. Depending on the field of application, several reagents are utilized, such as, e.g., epichlorohydrin (EPI), phosphoryl chloride, adipic acid mixed with anhydride and sodium tripolyphosphate. Cross linking involves the molecular reaction between the reactive hydroxyl groups on starch with multifunctional reagents resulting in ether or ester linkages. Such interactions not only take place between single chains but also links side by side chains as well [2]. As a result of cross-ligation in starch the degree of polymerization, molecular mass and the solubility in organic solvents are enhanced. These interactions resulting from cross-linking can also stabilize and strengthen the granules, especially by enhancing the existing hydrogen bonds through the introduction of additional covalent bonds. Therefore, modified starch exhibits a higher tolerance to acids, heat, and shear forces. Moreover, improved characteristics such as a decrease in retrogradation, a higher viscosity, and better paste clarity are obtained [2][3]. Various factors such as starch parameters, cross linking reagent concentration and composition, pH, reaction time as well as temperature have been found to affect the cross-linking efficiency [4][5].
Esterification takes place on hydroxyl groups at the C3, C2 and C6 positions (order indicates increasing reactivity) [6], whereby two types of esters can be distinguished, organic and inorganic starch esters. It involves the acid-catalyzed mediated substitution reaction of a nucleophilic acyl compound with a molecule containing an acid anhydride, acid chloride, or carboxylic acid structure. With the esterification of hydroxyl groups of available glucose units, the hydrogen bonding ability of amylose or amylopectin is compromised [7][8]. Improved hydrophobicity, swelling capacity, and lower retrogradation rate were observed [9]. Succinylation, as an example for organic esterification reduces the gelatinization temperature and retrogradation while simultaneously increasing the thickening power, viscosity, and water retention properties [10]. An example of an inorganic ester is the product of phosphorylation, as mentioned above. Besides being a natural modification, it is also used industrially to increase the viscosity, pasting transparency, and gelatinization characteristics [11]. However, the rate of esterification greatly depends on the structure of the reacting acids particularly on the steric, mesomeric and inductive effects exerted by them. Similarly other factors such as molar ratio, temperature, reaction duration is equally crucial to determine the effectiveness of the esterification reaction [8]. The degree of substitution (DS) acts as an indicator of quality and physical properties of the starch esters and determines the extent to which the starch can recrystallize or retrograde [12]. The esterification is widely used in paper production, pharmaceuticals, but also in food processing.
Improved properties regarding thermal stability, rheological behavior, and, with restrictions, ionic activity are observed for etherified starch. The etherification substitutes hydroxyl groups with anhydroglucose units using functional groups, which can be positively or negatively charged. Moreover, the generation of amphoteric or non-ionic starches are possible. One important example of a negatively charged starch ether is carboxymethyl starch, classified as a green polymer that covers a broad spectrum of applications in industry, due to its great hydrophilicity and therefore, cold-water solubility, as well as flocculation behavior, for example, in pharmaceutical approaches as a drug delivery system or in waste water purification as biosorbents to sequestrate heavy metals [7][13].
Another opportunity for modifying starch is oxidation, which can cause depolymerization of starch, whereby it enhances hydrophilicity, disrupts molecule linearity and crystallinity, and reduces retrogradation as well as enthalpy by the formation of carbonyl and carboxyl groups in native starch by the use of an oxidizing agent [14]. There are various oxidants available, such as potassium permanganate, sodium hypochlorite, hydrogen peroxide, etc. Oxidation mainly happens at hydroxyl groups C2, C3, and C6, depending on the pH, temperature, and reagent, and results in the formation of dialdehyde- and dicarboxylic acid derivates [7]. Nowadays, the impact of ozone on oxidation reactions has grown, as it can react with starch without any catalyst or special reaction conditions, while additionally exhibiting minimal effects on morphology, crystallinity, and the inner structure, as it acts especially on the amorphous regions of starch [14]. This kind of modification is used for food, textile, and paper processing.
During grafting, a polymer chain (grafting onto) or monomers such as vinyl units (grafting from) is introduced into the starch, resulting in copolymerization that combines the features of both polymers. The initiation of the copolymerization process occurs in particularly at the C1–C2 end groups and the C2–C3 position in the presence of a free radical initiator [7][15]. Therefore, two types of radical initiating can be distinguished: chemical initiation with, e.g., ceric ammonium nitrate or potassium permanganate, or irradiation initiation via UV light or microwaves. It has been reported that the ceric salts form complexes with alcohols and glycols, through disproportionation reaction which is also the rate-determining step of the oxidation-reduction reaction. The key characteristic of the oxidation with ceric ions is that it advances with single electron transfer leading to generation of free radicals on the starch. The presence of free radicals on the starch causes initialization of polymerization to form a graft copolymer [16]. When manganese ions are used for the initiation process, different primary radical species are produced depending on the type and nature of the acid [17]. Only the polar bonds are “selectively excited” by microwave radiation, which results in their cleavage, generating multiple free radicals. The relatively non-polar C-C bonds however remain unaffected [18][19]. Depending on the polymer introduced, the resulting starch reveals different characteristics; mostly, it has higher biodegradability, thermal stability, hydrodynamic radius, and improved flocculation properties [20]. However, the use of microwaves and UV light for graft polymerization in starch has gained momentum in the past few decades because of the environmentally friendly nature and improved properties for commercial utilization. Grafting is mainly used in electrical engineering, production of bioplastics, agriculture as well as in pharmacy.
The acid hydrolysis of starch is performed at temperatures below the gelatinization temperature and favors properties such as solubility, pasting viscosity, gelatinization enthalpy, and swelling power, whereas the effect on granular morphology differs among species, tissues, and degrees of hydrolysis. For instance, a slight hydrolysis reveals no significant morphological alteration, which presupposes that the starch has not been boiled [21]. In general, the glycosidic bond can be attacked by various acids. The most common is the use of nitric, sulfuric, or phosphoric acid. The acid acts primarily on the starch granule surface before entering the inner structure [22]. This modification technique is mainly implemented in the food, paper, and textile industries.

2. Physical Modifications of Isolated Starches

Generally, physical modifications can be used to interrupt or alter the granular size and/or packing arrangement. Thus, physical modifications can generate starch with similar characteristics as those obtained from chemical treatments, but without any agent which is toxic or must be removed thereafter. The methods are easy, cost-effective, and largely environmentally friendly. Overall, physical modifications can be divided into two groups: thermal and non-thermal methods. The main thermal modifications include pregelatinization, extrusion, heat-moisturizing, annealing, and microwave treatments. Applications such as ultra-high pressure, ultrasonification, and pulse electric field are amongst the non-thermal methods.
Pregelatinization is a very simple modification that includes the boiling of starch in an aqueous solution followed by drum or spray drying, which then results in reduced hydrogen bonding, fragmentation, and the molecular disintegration of starch granules. The spray dried gelatinized starch finds its application in the microcapsule industry for the purpose of drug release due to its low cost and availability [23]. The pregelatinized starch is cold water soluble, shows no optical birefringence, and has enhanced viscosity and gel stability [11][24].
Extrusion is a thermomechanical method involving continuously forcing gelatinized starch under pressure out of a shaping aperture. This technique plays a role in the packaging and food industry [25]. However, small alterations in processing conditions can influence the quality and properties of extrudates. For instance, when rice flour is extruded under low-expansion condition recombined rice is obtained whereas under high-expansion conditions crispy rice is produced [26].
The most practiced thermal methods are heat-moisture treatment (HMT) and annealing. The physicochemical properties of starch are greatly enhanced by both of these hydrothermal processes without affecting the granular structure. Both methods differ in the moisture content used. While annealing employs an excessive amount of water (>40%), a long processing time, and temperatures over the glass transition but below the gelatinization temperature, HMT requires a high temperature and a low moisture content range (10–35%). HMT alters the crystalline and amorphous layers of granules and thus reduces swelling property, solubility, and viscosity, simultaneously raising the pasting temperature. However, the efficiency of this technique greatly depends upon the process variables and the source of starch. The lower the moisture used, the greater is the alteration on the starch granule surface and hence increased susceptibility to other types of modifications. Additionally, starch derived from cereals shows lesser sensitivity to alterations than the tuber and root starches [27].
Annealing improves the crystallinity by providing molecular rearrangement and lowering the free energy; thus, more organized structures can form [28]. Since the heat required for gelatinization of starch is inversely correlated to the area of the starch crystalline region; thus, it is a method that works better with starches that revealed increased amorphous regions. Due to rearrangement of the crystalline structure, starch presents greater resistance to digestibility after application of annealing technology [28].
Microwaves are electromagnetic waves with a wavelength between 1 mm and 1 m corresponding to the frequency range between 300 MHz and 300 GHz [29]. Polar molecules (e.g., water molecules) absorb microwave energy, rapidly align themselves in response to the electric field, and produce bulk heat by molecular friction. Microwave treatment of starch resulted in lower melting enthalpy, swelling strength, and solubility, as well as higher gelatinization temperatures [1][3], which is particularly relevant in the pharmaceutical industry.
Micronization is a significant reduction of the average particle size. It can be a thermal treatment with infrared radiation in the range of 1800 to 3400 nm. It also heats starch in a short amount of time [30]. Non-thermal treatments include the use of a vacuum ball-mill or a hammer- or pin mill, which utilize mechanical friction forces and compression to break down granules into fragments or fine powder. Crystallinity and molecular structures are destroyed in addition to granule morphology. Cryogenic micronization, a much gentler technique, has been used for cereal grains. It causes marginal damage of the starch structure but it has not yet been applied on an industrial scale, especially in the food industry [31].
Another technology used for physical modification is the γ-ray irradiation treatment of starch. Plenty of research has been done in the past few decades to assess the impact of γ-radiation on starch granules. Radiation processing via γ-irradiation includes the utilization of a radioactive isotope, either in the form of cesium 137 or cobalt-60. These isotopes can emit high energy γ-rays or photons and are capable of invading great depth into the target product. The γ-irradiation causes a decrease in crystallinity, the swelling index, and pasting properties. In addition, starch granules become internally cracked, but surface cracking is not observed [32].Factors such as water content of the samples, type of gas atmosphere, dose and rate of dose as well as structure and composition of starch affects the impact of γ-irradiation. However, this method needs safety considerations.
Furthermore, starch can be exposed to ultra-high pressures (100–1000 MPa). High pressure is a green and eco-friendly technique capable of altering non-covalent chemical linkages with very minor effects on the covalent linkages. Depending on duration, pressure intensity, temperature, and starch origin, the swelling power and solubility increase and retrogradation is delayed. Moreover, the thermal and pasting properties, as well as the amounts of resistant, fast, and slow digestible starch, are altered significantly [33]. Ultra-high pressure is applied for food processing.
The application of frequencies above 20 kHz to agitate particles is known as ultrasonication. The sonication results in acoustic cavitation which is characterized by the rapid development and dissolution of bubbles in a liquid exposed to heat and pressure. The effects of ultrasonic treatment on starch properties depend on the time, intensity, frequency, type, and structure of the starch. However, it has been demonstrated that ultra sonification generates cracks and cavities on the granule surface and thereby increases the water-solubility and rheological characteristics [34]. This kind of modification is important to produce biofuels as well as for pharmaceutical applications as preparation of nanoparticles for drug delivery.
Pulse electric field technology, wherein a field strength of around 10–80 kV/cm for a period of micro- or milliseconds is used, can result in split or deformed starch granules. With increasing electric field strength, pasting properties such as the peak, breakdown, and final viscosity drop, and the crystallinity decreases. However, product properties (such as molecular weight, pH, conductivity) and temperature conditions of the reaction system can affect the modification of starch by influencing the pulse electric field intensity. Furthermore, ohmic heating can lead to gelatinization [35][36].
Another innovative non-thermal technology for modifying starch is cold plasma treatment. Plasma is a mixture of ions, radicals, free electrons, and neutral atoms and molecules. Accordingly, the degree of ionization can vary between 1 and 100% and is generated via an energy input through electric or magnetic fields and, radio waves, microwaves, or heat. Among these, an electric field is the most frequently used energy source. Likewise, the effect on starch properties can significantly differ depending on the exposure time, power and volt supply, and the gas composition used, so several modifications are possible, such as cross-linking, depolymerization, and etching. Thus, plasma treatment leads to either an increase or decrease in molecular weight. The same is also valid for the gelatinization temperature. Furthermore, this physical method destroys starch granules’ crystallinity. The degree of moisture content must also be considered because the bombardment with plasma can produce hydroxyl radicals that favor further breakdown of crystalline regions. This can also cause cracks and holes in granules. The use of plasma is a promising technique with numerous applications, as it allows for several types of starch modification. Importantly, it is also fast, simple, and eco-friendly, leaving no chemical and/or toxic waste products. Starch-related plasma research is in its early stages and there is an enormous energy demand that accompanies the process of utilization, so, in the end, a cost-benefit calculation is crucial [37]. This technology is currently used in food processing. It has also been shown to significantly improve seed germination, so it may have potential for industrial use in agriculture in the future [38].

3. Enzymatic Modifications of Isolated Starches

Enzymatic modification is referred to as the utilization of enzymes that alter starch parameters. This kind of modification is primarily applied in the food industry. Traditionally, starch modifications by an enzymatic reaction are generally divided into four categories: endoamylase, exoamylase, debranching enzymes, and transferases. However, further potential of various enzymes/proteins was mentioned before.
α-amylases are a well-known example of the endoamylase category. These enzymes randomly hydrolyze starch at any (1,4)-linkage within the two starch polymers, amylose, and amylopectin. As a result, they reduce the chain length [39]. Depending on the type and concentration of enzyme being employed, α-amylase can produce pockets at the surface of starch granules. Exoamylases include β-amylases [40], Amyloglucosidases [41], and α-glucosidases [42]. Exoamylases act on external glucose residues of amylose or amylopectin and thus produce β-limit dextrins (β-amylase), maltose, and glucose (glucoamylase and α-glucosidase) [43]. The enzymatic modification by β-amylases results in the development of multiple cracks on the starch granule surface. Compared to native starch, the enzymatically modified starch exhibits increased solubility but decreased swelling capacity and pasting viscosity.
The debranching enzyme category includes isoamylase [39] and pullulanase [44][45]. Debranching enzymes can hydrolyze the α-1,6 glucosidic bond. The major difference between isoamylase and pullulanase is that the pullulanase can hydrolyze the α-1,6 glycosidic bond in starch and pullulan, while isoamylase can only hydrolyze the α-1,6 bond in glycogen and starch [46]. These enzymes produce linear glucans by specifically degrading amylopectin.
The transferases include amylomaltase [39][47], cyclodextrin glycosyltransferase [39][47], and amylosucrase [48]. Transferases hydrolyze the α-1,4 glycosidic bond of a donor molecule and transmit the cleaved residue from the donor to a glycosidic acceptor with the formation of a new glycosidic bond [1].
The hydrolysis rate values for each enzyme mentioned above significantly differ depending on the starch. For native starch, the differences mainly depend on its structural properties [9][49].Since the production of enzymes on an industrial scale, such as through production of amylases by bacteria and fungi, and germination of cereals, the enzymatic modification of starch has gained commercial importance. This technology can be utilized to efficiently manage the specific properties of starch. However, a single enzyme does not substantially alter the properties of starch and therefore combination of enzymes is generally used to achieve the desired results. The specificity of the reaction, which takes place under mild temperature, pH, and agitation conditions, as well as the low effluent generation adds up to the advantages of utilizing enzymes. Furthermore, since the concentration of the enzyme used in the reaction is much lower than that of starch, introduction of even slightest of variations in the reaction medium can interrupt the reaction. These factors make enzymatic modification a very promising and attractive method for starch remodeling. However, due to the time-consuming nature of the enzyme-based method, its application prospects are hindered and therefore, there is a need of using this technology in conjunction with other modification methods. Furthermore, even though there are plenty of available enzyme resources with different properties in microbial or gene databases, the natural enzymes still possess some inevitable defects such as structural instability and low catalytic efficiency due to limitations of biological adaptation and evolution. Therefore, there is a need for improving the properties of industrially important enzymes. This could be achieved through enzyme engineering which involves altering the amino acid sequence of the enzymes and therefore plays a central role in developing efficient biocatalysts with good industrial application potential. In addition to this, attention should be paid to improving the production yields of the enzymes which demand optimization of expression elements and secretory pathways of the host cell and improvisation of high-density fermentation strategies.
Finally, there are numerous possibilities and a great need to modify starch to change the properties in the direction required for the respective application area (see Section 4). Not all the types of modification listed here are also applied in industrial processes (so far), but there is no denying that future technologies will have to be developed/utilized which agree with sustainability, climate, and environmental protection. This involves the environmental safety of the reaction components used, process and production reliability, energy requirements, resource conservation and material recovery. Examples of such new technologies are on the one hand the material use of CO2, which is available as a quasi-inexhaustible resource, and on the other hand the use of cold plasma.
The key to success, however, is most likely to be found in the use of interdisciplinary methods; in a combination of chemical/physical methods and the use of enzymes, the reactions can occur under very mild and efficient conditions. One such example is the starch esterification, where a very high degree of substitution of the hydroxyl groups can be achieved by using lipase in combination with microwave irradiation in an ionic solution [50].

References

  1. Wang, S.; Wang, J.; Liu, Y.; Liu, X. Starch Modification and Application. In Starch Structure, Functionality and Application in Foods; Wang, S., Ed.; Springer: Singapore, 2020; pp. 131–149. ISBN 978-981-15-0621-5.
  2. Ayoub, A.S.; Rizvi, S.S.H. An Overview on the Technology of Cross-Linking of Starch for Nonfood Applications. J. Plast. Film Sheeting 2009, 25, 25–45.
  3. Wang, S. (Ed.) Starch Structure, Functionality and Application in Foods; Springer: Singapore, 2020; ISBN 978-981-15-0621-5.
  4. Koo, S.H.; Lee, K.Y.; Lee, H.G. Effect of cross-linking on the physicochemical and physiological properties of corn starch. Food Hydrocoll. 2010, 24, 619–625.
  5. Chung, H.-J.; Woo, K.-S.; Lim, S.-T. Glass transition and enthalpy relaxation of cross-linked corn starches. Carbohydr. Polym. 2004, 55, 9–15.
  6. Lehmann, A.; Volkert, B. Investigations on esterification reactions of starches in 1-N-butyl-3-methylimidazolium chloride and resulting substituent distribution. J. Appl. Polym. Sci. 2009, 114, 369–376.
  7. Chen, Q.; Yu, H.; Wang, L.; ul Abdin, Z.; Chen, Y.; Wang, J.; Zhou, W.; Yang, X.; Khan, R.U.; Zhang, H.; et al. Recent progress in chemical modification of starch and its applications. RSC Adv. 2015, 5, 67459–67474.
  8. Otache, M.A.; Duru, R.U.; Achugasim, O.; Abayeh, O.J. Advances in the Modification of Starch via Esterification for Enhanced Properties. J. Polym. Environ. 2021, 29, 1365–1379.
  9. Singh, J.; Kaur, L.; McCarthy, O.J. Factors influencing the physico-chemical, morphological, thermal and rheological properties of some chemically modified starches for food applications—A review. Food Hydrocoll. 2007, 21, 1–22.
  10. Ariyantoro, A.R.; Katsuno, N.; Nishizu, T. Effects of Dual Modification with Succinylation and Annealing on Physicochemical, Thermal and Morphological Properties of Corn Starch. Foods 2018, 7, 133.
  11. Tharanathan, R.N. Starch--value addition by modification. Crit. Rev. Food Sci. Nutr. 2005, 45, 371–384.
  12. Hong, J.; Zeng, X.-A.; Brennan, C.S.; Brennan, M.; Han, Z. Recent Advances in Techniques for Starch Esters and the Applications: A Review. Foods 2016, 5, 50.
  13. Spychaj, T.; Wilpiszewska, K.; Zdanowicz, M. Medium and high substituted carboxymethyl starch: Synthesis, characterization and application. Starch-Stärke 2013, 65, 22–33.
  14. Pandiselvam, R.; Manikantan, M.R.; Divya, V.; Ashokkumar, C.; Kaavya, R.; Kothakota, A.; Ramesh, S.V. Ozone: An Advanced Oxidation Technology for Starch Modification. Ozone Sci. Eng. 2019, 41, 491–507.
  15. Djordjevic, S.; Nikolic, L.; Kovacevic, S.; Miljkovic, M.; Djordjevic, D. Graft copolymerization of acrylic acid onto hydrolyzed potato starch using various initiators. Period. Polytech. Chem. Eng. 2013, 57, 55.
  16. Kurkuri, M.D.; Lee, J.-R.; Han, J.H.; Lee, I. Electroactive behavior of poly(acrylic acid) grafted poly(vinyl alcohol) samples, their synthesis using a Ce (IV) glucose redox system and their characterization. Smart Mater. Struct. 2006, 15, 417–423.
  17. Athawale, V.D.; Rathi, S.C. Graft Polymerization: Starch as a Model Substrate. J. Macromol. Sci. Part C Polym. Rev. 1999, 39, 445–480.
  18. Kumar, D.; Pandey, J.; Raj, V.; Kumar, P. A Review on the Modification of Polysaccharide Through Graft Copolymerization for Various Potential Applications. Open Med. Chem. J. 2017, 11, 109–126.
  19. Noordergraaf, I.-W.; Fourie, T.; Raffa, P. Free-Radical Graft Polymerization onto Starch as a Tool to Tune Properties in Relation to Potential Applications. A Review. Processes 2018, 6, 31.
  20. Jyothi, A.N. Starch Graft Copolymers: Novel Applications in Industry. Compos. Interfaces 2010, 17, 165–174.
  21. Wang, S.; Copeland, L. Effect of acid hydrolysis on starch structure and functionality: A review. Crit. Rev. Food Sci. Nutr. 2015, 55, 1081–1097.
  22. Pratiwi, M.; Faridah, D.N.; Lioe, H.N. Structural changes to starch after acid hydrolysis, debranching, autoclaving-cooling cycles, and heat moisture treatment (HMT): A review. Starch-Stärke 2018, 70, 1700028.
  23. Dang, X.; Yang, M.; Shan, Z.; Mansouri, S.; May, B.K.; Chen, X.; Chen, H.; Woo, M.W. On spray drying of oxidized corn starch cross-linked gelatin microcapsules for drug release. Mater. Sci. Eng. C Mater. Biol. Appl. 2017, 74, 493–500.
  24. Liu, Y.; Chen, J.; Luo, S.; Li, C.; Ye, J.; Liu, C.; Gilbert, R.G. Physicochemical and structural properties of pregelatinized starch prepared by improved extrusion cooking technology. Carbohydr. Polym. 2017, 175, 265–272.
  25. Moscicki, L.; Mitrus, M.; Wojtowicz, A.; Oniszczuk, T.; Rejak, A. Extrusion-Cooking of Starch. In Advances in Agrophysical Research; Grundas, S., Ed.; InTech: London, UK, 2013; ISBN 978-953-51-1184-9.
  26. Ye, J.; Hu, X.; Luo, S.; Liu, W.; Chen, J.; Zeng, Z.; Liu, C. Properties of Starch after Extrusion: A Review. Starch-Stärke 2018, 70, 1700110.
  27. Da Zavareze, E.R.; Dias, A.R.G. Impact of heat-moisture treatment and annealing in starches: A review. Carbohydr. Polym. 2011, 83, 317–328.
  28. Gunaratne, A. Heat-Moisture Treatment of Starch. In Physical Modifications of Starch; Sui, Z., Kong, X., Eds.; Springer: Singapore, 2018; pp. 15–36. ISBN 978-981-13-0724-9.
  29. Braşoveanu, M.; Nemţanu, M.R. Behaviour of starch exposed to microwave radiation treatment. Starch-Stärke 2014, 66, 3–14.
  30. Emami, S.; Meda, V.; Pickard, M.D.; Tyler, R.T. Impact of micronization on rapidly digestible, slowly digestible, and resistant starch concentrations in normal, high-amylose, and waxy barley. J. Agric. Food Chem. 2010, 58, 9793–9799.
  31. Li, E.; Dhital, S.; Hasjim, J. Effects of grain milling on starch structures and flour/starch properties. Starch-Stärke 2014, 66, 15–27.
  32. Punia, S. Barley starch modifications: Physical, chemical and enzymatic—A review. Int. J. Biol. Macromol. 2020, 144, 578–585.
  33. Castro, L.M.; Alexandre, E.M.; Saraiva, J.A.; Pintado, M. Impact of high pressure on starch properties: A review. Food Hydrocoll. 2020, 106, 105877.
  34. Bonto, A.P.; Tiozon, R.N.; Sreenivasulu, N.; Camacho, D.H. Impact of ultrasonic treatment on rice starch and grain functional properties: A review. Ultrason. Sonochem. 2021, 71, 105383.
  35. Wang, Q.; Li, Y.; Sun, D.-W.; Zhu, Z. Enhancing Food Processing by Pulsed and High Voltage Electric Fields: Principles and Applications. Crit. Rev. Food Sci. Nutr. 2018, 58, 2285–2298.
  36. Han, Z.; Zeng, X.-A.; Zhang, B.; Yu, S. Effects of pulsed electric fields (PEF) treatment on the properties of corn starch. J. Food Eng. 2009, 93, 318–323.
  37. Thirumdas, R.; Trimukhe, A.; Deshmukh, R.R.; Annapure, U.S. Functional and rheological properties of cold plasma treated rice starch. Carbohydr. Polym. 2017, 157, 1723–1731.
  38. Li, L.; Chen, J.; Wang, H.; Guo, H.; Li, D.; Li, J.; Liu, J.; Shao, H.; Zong, J. Cold plasma treatment improves seed germination and accelerates the establishment of centipedegrass. Crop Sci. 2021, 61, 2827–2836.
  39. van der Maarel, M.J.E.C.; van der Veen, B.; Uitdehaag, J.C.M.; Leemhuis, H.; Dijkhuizen, L. Properties and applications of starch-converting enzymes of the alpha-amylase family. J. Biotechnol. 2002, 94, 137–155.
  40. Hyun, H.H.; Zeikus, J.G. Regulation and genetic enhancement of beta-amylase production in Clostridium thermosulfurogenes. J. Bacteriol. 1985, 164, 1162–1170.
  41. Dura, A.; Błaszczak, W.; Rosell, C.M. Functionality of porous starch obtained by amylase or amyloglucosidase treatments. Carbohydr. Polym. 2014, 101, 837–845.
  42. Del Moral, S.; Barradas-Dermitz, D.M.; Aguilar-Uscanga, M.G. Production and biochemical characterization of α-glucosidase from Aspergillus niger ITV-01 isolated from sugar cane bagasse. 3 Biotech 2018, 8, 7.
  43. Ashok, P.; Soccol, C.R.; Soccol, V.T. Biopotential of immobilized amylases. Indian J. Microbiol. 2000, 40, 1–14.
  44. Bender, H.; Lehmann, J.; Wallenfels, K. Pullulan, an extracellular glucan from Pullularia pullulans. Biochim. Biophys. Acta 1959, 36, 309–316.
  45. Barnett, C.; Smith, A.; Scanlon, B.; Israilides, C.J. Pullulan production by Aureobasidium pullulans growing on hydrolysed potato starch waste. Carbohydr. Polym. 1999, 38, 203–209.
  46. Hii, S.L.; Tan, J.S.; Ling, T.C.; Ariff, A.B. Pullulanase: Role in starch hydrolysis and potential industrial applications. Enzym. Res. 2012, 2012, 921362.
  47. Kaur, B.; Ariffin, F.; Bhat, R.; Karim, A.A. Progress in starch modification in the last decade. Food Hydrocoll. 2012, 26, 398–404.
  48. Seo, D.-H.; Yoo, S.-H.; Choi, S.-J.; Kim, Y.-R.; Park, C.-S. Versatile biotechnological applications of amylosucrase, a novel glucosyltransferase. Food Sci. Biotechnol. 2020, 29, 1–16.
  49. Pfister, B.; Zeeman, S.C.; Rugen, M.D.; Field, R.A.; Ebenhöh, O.; Raguin, A. Theoretical and experimental approaches to understand the biosynthesis of starch granules in a physiological context. Photosynth. Res. 2020, 145, 55–70.
  50. Adak, S.; Banerjee, R. A green approach for starch modification: Esterification by lipase and novel imidazolium surfactant. Carbohydr. Polym. 2016, 150, 359–368.
More
Video Production Service