1,307
Views
0
CrossRef citations to date
0
Altmetric
Review Article

Recent advancements in cross-linked starches for food applications- a review

, , , &
Pages 411-430 | Received 24 Nov 2023, Accepted 04 Feb 2024, Published online: 06 Mar 2024

ABSTRACT

The industrial food applications of native starches are limited due to their limited resistance to shear, susceptibility to thermal decomposition, and high tendency for retrogradation. Cross-linking of starches adds intra and inter-molecular bonds at various locations in the starch molecule that result in the stability of the granule and enhance functional attributes. This process involves the use of various cross-linking agents such as sodium trimetaphosphate (STMP), sodium tripolyphosphate (STPP), epichlorohydrin (EPI), phosphorus chloride (POCl3), and citric acid, etc., to introduce covalent or non-covalent linkages between starch molecules. Cross-linked starch exhibits improved resistance to retrogradation, enhanced freeze-thaw stability, and increased stability during cooking, shearing, and processing, such as high or low temperatures and pH. This review paper offers an overview of cross-linking modifications, emphasizing their importance in addressing native starch limitations. It underscores the significance of selecting appropriate cross-linking agents to customize starch properties for specific applications. Furthermore, the paper discusses the properties and applications of cross-linked starches and delves into regulatory considerations regarding their use. Regulatory consideration for cross-linked starches has also been discussed. Cross-linking modifications offer a promising avenue to unlock the full potential of starch and expand its utility across a wide range of applications.

Introduction

Starch is a versatile and sustainable biopolymer with extensive food and packaging applications[Citation1,Citation2]. The abundance of hydroxyl (OH) groups in starch is a characteristic that renders it highly sensitive to water and similar liquids.[Citation3] This innate hydrophilic behavior of native starch has presented challenges in the development of starch-based materials. Therefore, researchers have directed their efforts toward employing different chemical modifications to transform the numerous OH groups present in starch into different molecular structures. Chemical modifications, including oxidization, esterification, grafting, and cross-linking of starches, have gained prominence and play a pivotal role in tailoring starch to meet the specific requirements of these diverse fields. These methods offer valuable means of tailoring starch to suit diverse applications by altering its fundamental chemical structure.[Citation4] Cross-linking of starches represents the most frequently used and transformative approach in food science and technology, unlocking a myriad of properties and applications that have revolutionized the food industry. Starch crosslinking introduces a covalent interconnection via hydrogen bonds in starch granules, providing strong bonding between the molecules and restricting the movement of the polymer chains. Cross-linking of starch granules increases its resistance to high temperatures, acidity, and shear forces, rendering it exceptionally useful in processed foods. Cross-linked starch bestows enhanced thickening capabilities, contributing to the desired consistency in products like sauces, gravies, and desserts.[Citation5] The textural improvements facilitated by cross-linking open up avenues for the creation of smoother, more appealing food items. Furthermore, cross-linked starch is categorized as a type 4 resistant starch (RS-4) due to the covalent bonding during the crosslinking process, which strengthens granule structure and maintains the starch granules integrity during cooking, leading to increased resistance to enzyme hydrolysis.[Citation6,Citation7]

So far, phosphorus oxychloride (POCl3), sodium tripolyphosphate (STPP), sodium trimetaphosphate (STMP), sodium hexametaphosphate (SHMP), citric acid, and adipic acid are commonly employed as cross-linking agents for starch. Beyond these attributes, cross-linked starches are pivotal to extending the shelf life of various food products, providing a buffer against undesirable textural changes over time. As we explore the multifaceted world of cross-linked starches in the subsequent sections of this article, their applications will come to light, revealing their indispensable role in shaping the texture, stability, and sensory appeal of a diverse range of food products.

The article highlights the necessity for manufacturers to maintain stringent quality control measures and offers insights into how expert consultation and adherence to regional regulatory nuances are vital for navigating this multifaceted regulatory landscape. In essence, this review article aims to provide a comprehensive overview of the regulatory intricacies surrounding cross-linked starches in food applications, serving as an indispensable resource for industry professionals and researchers seeking to innovate while ensuring compliance with global food regulations.

Types and mechanisms of cross-linking agents

The chemical cross-linking of starch involves the use of chemical agents that can react with the hydroxyl groups (−OH) in starch molecules, forming covalent bonds between them.[Citation8] A starch molecule has two types of hydroxyl groups: the primary (6-OH) and the secondary (2-OH and 3-OH). The reaction of these hydroxyl groups with multifunctional reagents results in crosslinked starches (). Within a starch granule, numerous chains are in close proximity, and as a result, these reactions occur not only within individual chains but also connect adjacent chains. The crosslinking agents develop either ether or ester intermolecular bonds between the hydroxyl groups on starch molecules, as detailed in studies by Yao et al.[Citation9] Remarkably, only a small quantity of multifunctional reagent is sufficient to interconnect starch molecules through crosslinking reactions.

Figure 1. Mechanism of starch crosslinking.

Figure 1. Mechanism of starch crosslinking.

Phosphorylation cross-linking

Phosphoryl chloride (POCl3), sodium trimetaphosphate (STMP), and sodium tripolyphosphate (STPP), and a mixture of STMP – STPP have been used to prepare cross-linked phosphorylated starches.[Citation1] Starch phosphates can be categorized into two groups: monostarch phosphate and distarch phosphate (cross-linked starch). Monostarch phosphate can be obtained by esterification of native starch using STPP or mixtures of sodium dihydrogen phosphate dihydrate and disodium hydrogen phosphate twehydrate. Distarch phosphate can be obtained using POCl3 and STMP. Distarch phosphate starches are used as a thickener and stabilizer, offering stability against gelling and retrogradation and high resistance to syneresis during storage. Monostarch phosphate displays improved paste clarity, viscosity, and water-binding capacity. Conversely, the formation of distarch phosphates contributes to preserving granule integrity and imparting greater resistance to retrogradation, high temperatures, and low pH compared to native starch. A mechanism of cross-linking of starches is shown in .

Figure 2. Mechanism of different cross-linking agents for starch cross-linking.

Figure 2. Mechanism of different cross-linking agents for starch cross-linking.

Cross-linking with phosphorous oxychloride (POCl3)

Phosphorous oxychloride (POCl3) exhibits a rapid reactivity with starch, resulting in the formation of a distarch phosphate. Crosslinking with POCl3 is notably influenced by the pH of the reaction environment, and the most effective crosslinking reactions occur at pH levels exceeding 11. Moreover, the presence of sodium sulfate, typically at a concentration of 2% w/w based on dry starch weight, serves to enhance further both the rate and efficiency of the crosslinking reaction, as reported by Gao et al.[Citation10] POCl3 contains three chloride ions and plays a pivotal role in this process. Upon introduction of POCl3 into the starch solution, the initial chloride ion rapidly reacts with water, forming a phosphorous dichloride with a very short half-life of approximately .01 s, as documented by Delley et al.[Citation11] This phosphorous dichloride is likely the crucial crosslinking agent. Simultaneously, the remaining two chloride ions in the phosphoryl chloride react with water over a longer half-life period of about 4 min at 25°C. It is imperative to note that phosphoryl chloride must be added to the starch solution as swiftly as possible in order to generate the phosphorous dichloride in situ. This ensures that the phosphorous dichloride has sufficient time to permeate the starch granules, where it facilitates the cross-linking process.

Cross-linking with sodium trimetaphosphate (STMP)

Sodium trimetaphosphate (STMP) serves as a traditional cross-linking agent for starch, inducing both intra and intermolecular bonds through esterification reactions within the starch. This esterification process is pivotal in creating crosslinks within the starch molecule, consequently reducing its affinity for water and reinforcing the structural integrity of the hydrogel through network formation. The alkaline environment in the crosslinking stage triggers the esterification of hydroxyl groups on starch molecules by STMP’s phosphate group, forming starch phosphate diesters.[Citation12] Crosslinking occurs between C6 and either C2 or C3 in the process. The crosslinking of starch with STMP follows a chemical mechanism in which STMP initially dissolves in water, resulting in the generation of phosphate ions. It’s important to highlight that each phosphate linkage, whether mono or distarch phosphate, results in a negatively charged group (−PO4). Consequently, the overall charge of the crosslinked starch molecule is contingent upon the extent of the reaction. The presence of mono and distarch phosphate groups in the crosslinked starches is evident from the peaks observed between 5.0–4.2 and 4.0–3.0 ppm in the pNMR spectrum. The distarch phosphates signify the presence of crosslinks in the molecule. Notably, the intensity of the distarch phosphates peak is higher when using a higher STMP concentration, indicating a more extensive crosslinking reaction.[Citation13]

Cross-linking with sodium tripolyphosphate (STPP)

STPP has two different reagents that react with starch through different mechanisms. Below a pH of 9.0, the terminal phosphate groups of STPP become protonated, generating monometaphosphates. These can swiftly react with starch hydroxyl groups, yielding monostarch phosphates.[Citation14] Conversely, at a reaction pH exceeding 10, ionized hydroxyls can initiate an attack on the central phosphate of STPP, resulting in the formation of starch pyrophosphates. Starch hydroxyl groups can subsequently engage in further attacks, leading to the production of distarch phosphate. At high pH (>10), all acidic sites on STPP become negatively charged, forming a dense shield that likely impedes the nucleophilic attack on STPP. The substantial repulsion from electronegative hydroxyls on starch, caused by tripolyphosphate, results in a relatively low degree of phosphorylation at alkaline pH. In conditions of high temperature, starch hydroxyls undergo slight ionization by alkali, leading them to attack the central phosphate carrying only one ionized hydroxyl rather than the fully ionized terminal phosphates. This bimolecular reaction forms a starch pyrophosphate with the release of orthophosphate. Subsequently, a second starch hydroxyl group can engage in an attack on the starch pyrophosphate, giving rise to distarch phosphate.

Cross-linking with epichlorohydrin

The reaction between epichlorohydrin (1-chloro-2,3-epoxypropane) and starch is a multistep process, as detailed by Zhang et al.[Citation15] Epichlorohydrin, a multifunctional compound, initiates the crosslinking of starch during this reaction. As the process unfolds, a single crosslink is formed, consuming either one or two epichlorohydrin molecules in the process. This leads to the creation of diester and diglycerol molecules within the crosslinked starch structures. To perform this crosslinking reaction, a starch solution is combined with a lower concentration of epichlorohydrin. The reaction mixture is stirred, and the temperature is maintained at 40°C for 17 h. To halt the reaction, a 3% v/v hydrochloric acid solution is introduced to adjust the pH to 5.25. The resulting mixture is then filtered, washed, and subsequently dried, as described by Zhang et al. in their[Citation15]2014 study.

Cross-linking with citric acid

One compelling reason for employing citric acid as a cross-linking agent lies in the fact that unreacted citric acid (CA) is regarded as nutritionally safe and may also function as a starch plasticizer.[Citation16] Cross-linking with citric acid operates through the formation of a cyclic anhydride intermediate. The cycle opens when starch -OH functional groups catalyze esterification, resulting in a new carboxylic acid unit in citric acid (CA). This unit possesses the ability to create a new intra-molecular anhydride moiety with the adjacent carboxylic acid unit. Esterification likely occurs predominantly at primary -OH groups, recognized as the most reactive positions when compared to secondary -OH groups within the structural unity of starch.[Citation17]

Crosslinking with adipate

Adipic acid is a commonly employed reagent for starch modification. Due to its possession of two carboxylic groups, it can generate both cross-linked starch and monosubstituted derivatives. The reaction is carried out in a water suspension under alkali conditions, involving the gradual addition of an acetic anhydride/adipic acid mixture. The combination of acetic anhydride and adipic acid forms a mixed acetic adipic anhydride, which reacts with starch to produce acetylated distarch-adipate. The reaction proceeds rapidly at pH 8.0, but the reagent must be added slowly while maintaining a pH close to 8.0.[Citation18]

Properties of cross-linked starches

Functional properties

Diverse functional properties of starches, like swelling power (SP), solubility, water absorption capacity (WAC) and oil absorption capacity (OAC), pasting attributes, etc. are affected by crosslinking. A brief overview of the effect of different crosslinking agents on starch functional attributes is given in . Crosslinking decreased the SP of different starches, such as barnyard starch,[Citation19] sorghum starch,[Citation36] tiger nut starch,[Citation20] corn starch,[Citation24] talipot starch,[Citation27] etc., whereas the SP increased for Cyperus Starch,[Citation5] Cyrtosperma senegalense starch,[Citation21] etc, Similarly, the solubility also was found to be significantly changed by cross-linking. The solubility of starches like corn,[Citation24] cassava,[Citation25] lotus seed,[Citation23] Litchi Kernel,[Citation19] maize,[Citation37] etc, were decreased.

Table 1. A brief overview of the effect of different crosslinking agents on starch functional attributes.

The extent of the impact on SP and solubility of crosslinked starches varies with the starch and crosslinking method. For example, The STMP/STPP crosslinking of cassava starch decreased the SP (16.92 to 10.76 g/g) with successive increases in reagent concentration. However, the solubility was increased at a lower concentration (1%) and decreased by a further increase in concentration.[Citation25] Likewise, the STMP/STPP cross-linked corn starch exhibited a reduction in SP and solubility from 16.85 to 5.70 g/g and .46% to .18%, respectively.[Citation24] Tesfay et al.[Citation38] observed that the SP and solubility of enset starch were significantly reduced with an increase in epichlorohydrin concentration. Similarly, a reduction in SP was noted in epichlorohydrin crosslinked mandua starch at a concentration of 1.0% w/w.[Citation8]

The restricted SP and solubility of STMP/STPP crosslinked starches could be due to the presence of a phosphate group, which facilitates the development of additional covalent bonds between starch molecules. This could be beneficial for retaining the starch granule integrity and strengthening the intergranular bonds. In addition, crosslinking hinders the movement of amorphous regions and limits solubility and SP.[Citation24] Further, the crosslinking of starch augments the hydrogen bonding between the starch chains and contributes to the restricted swelling and solubility of crosslinked starches.[Citation8]

The crosslinking of starch with different crosslinking agents significantly impacts their WAC and OAC. The citric acid crosslinking of sago starch increased its WAC and reduced the OAC.[Citation39] The presence of a higher number of hydrophilic sites in citric acid crosslinked starch contributes to its higher WAC. Similarly, the STMP/STPP crosslinking of breadfruit starch significantly enhanced its WAC, whereas it did not show any significant difference in OAC.[Citation26] The researchers also stated that the presence of hydrophilic phosphorus may be responsible for increased WAC of STMP/STPP crosslinked starch, and it also contributes to the electrostatic repulsion between starch chains, thereby facilitating higher water uptake and swelling. The crosslinking of banana starch with sodium hexametaphosphate considerably increased the WAC and OAC.[Citation40] The researchers stated that the increased space between starch chains arising from the steric hindrance may contribute to increased OAC. On the contrary, the crosslinking of potato starch with phosphorous oxychloride exhibited a significantly lower WAC than its native counterpart.[Citation41] The formation of covalent bonding strengthens the granules and thus resists the mobility of the amorphous region, reducing WAC.

Crosslinking with different reagents significantly altered the pasting attributes of starches. The pasting profile of epichlorohydrin crosslinked pearl millet starch from different cultivars is given in . It was noticed that the impact of crosslinking varies among different cultivars. Sodium hexametaphosphate crosslinking decreased the pasting attributes of banana starch, including pasting temperature (PT), peak viscosity (PV), trough viscosity (TV), breakdown viscosity (BV), final viscosity (FV), and setback viscosity (SV).[Citation40] Likewise, similar changes in pasting behavior were observed in STMP/STPP crosslinked breadfruit starch[Citation26] and STMP crosslinked maize starch.[Citation37] Likewise, the crosslinking of litchi[Citation43] and lotus[Citation23] starches with STMP (1–5%) reduced the PV, BV and augmented the PT, TV, FV, and SV with an increase in STMP concentration.

Figure 3. Pasting properties of native starch and cross-linked starches analyzed using a Micro ViscoAmylo-Graph. Where: NM= normal maize, WM= Waxy maize NP=normal potato, and waxy[Citation42].

Figure 3. Pasting properties of native starch and cross-linked starches analyzed using a Micro ViscoAmylo-Graph. Where: NM= normal maize, WM= Waxy maize NP=normal potato, and waxy[Citation42].

Nevertheless, Sriprablom et al.[Citation25] observed that the impact of STMP/STPP crosslinking on pasting attributes of cassava starch varies with the concentration. All the concentrations (1, 1.5, and 2%) increased the PT and decreased the BV. However, the PV was increased at 1% and decreased at higher concentrations (1.5 and 2%). The TV, FV, and SV were increased at a concentration of 1 and 1.5 and decreased at 2% as compared to their native counterpart. Similarly, an increase in STMP concentration from 1–5% decreased the PV and BV and augmented the PT of faba bean starch. However, the TV value was higher for crosslinked starches and decreased with increased reagent concentration. The SV and FV were found to be higher at 1% STMP concentration and decreased at higher concentrations than the native counterparts.[Citation22] Similarly, STMP crosslinked barnyard starch also exhibited a concentration-dependent change in its pasting behavior.[Citation19] The cross-linking of sorghum starch with epichlorohydrin also exhibited concentration-dependent changes in pasting behavior. A lower concentration (.1 and .3%) increased the PV, TV, FV, and PT of crosslinked sorghum starch, whereas it decreased at a higher concentration (.5 and 1%).[Citation36] A is showing pasting properties of native starch and cross-linked starches of normal maize, waxy maize and normal potato.[Citation42]

The cross-linking of talipot starch with 10% STMP/STPP mixture potentially increased its pasting behavior. The authors also stated that the crosslinking enhanced the starch granules’ intra and intermolecular interaction, which resists granule degradation during heating and leads to a rise in viscosity.[Citation44] Similarly, the crosslinking of kithul starch with phosphorus oxychloride also increased the PV, TV, SV, and FV.[Citation30] Crosslinking affects the pasting behavior of starch differently depending on the starch source, crosslinking agent, its concentration, and reaction conditions. A lower reagent concentration and a lower degree of crosslinking can augment the swelling and peak viscosity while heating starch with water, whereas a higher degree of crosslinking increases the granule rigidity and reduces the swelling capacity.[Citation45]

Similar to the pasting attributes, the rheological attributes of the crosslinked starches depend on the starch, crosslinking agent concentration, and the reaction conditions. The cross-linking of sorghum starch with .1% of epichlorohydrin increased its storage modulus (G′) and loss modulus (G′′) and decreased the Tan δ value. However, a higher concentration of epichlorohydrin (.5 and 1%) exhibited a decrease in G′ and G′′, and an increase in tan δ value.[Citation36] The cross-linking of different cultivars of pearl millet with epichlorohydrin significantly decreased their G′ value. It also noted a significant reduction in yield stress and consistency index after crosslinking.[Citation46] Likewise, the crosslinking of litchi starch with various levels of STMP reduced the magnitude of yield stress and consistency index as the concentration of STMP increased from 1 to 5%.[Citation19] The reduction in the G′ value of crosslinked starch could be due to their restricted swelling, which resists the formation of intergranular interaction.[Citation46] Contrary to the above, the crosslinking of kithul starch[Citation30] and talipot starch[Citation31] with POCl3 exhibited higher G′ and G′′ values as compared to their native counterparts. The authors stated that the increased G′ and G′′ of crosslinked starch could be due to the lower degree of deformation and improved granule rigidity.

Thermal properties

The thermal attributes of crosslinked starch (onset (To), peak (Tp), and conclusion (Tc) gelatinization temperature and change in the enthalpy (∆H)) have been studied by differential scanning calorimetry (DSC). The introduction of the phosphate group to the starch molecules by crosslinking strengthens the intermolecular interaction, thus inhibiting the gelatinization and increasing gelatinization temperatures.[Citation47] The thermal attributes of citric acid crosslinked tigernut starch were studied using DSC. The To, Tp, Tc, and ∆H values increased with an increase in citric acid concentrations.[Citation5] Likewise, the crosslinking of maize starch with different levels of POCl3 and STMP slightly increased the To value (). The cross-linking of barnyard millet starch with varying levels of sodium trimetaphosphate (STMP) increased the To and Tp values, whereas it decreased the Tc and ∆H values. The increased To and Tp are due to the introduction of the phosphate group into the starch molecules, which strengthened the molecular interaction. Besides, the starch granules undergo hydrogen and covalent bonding after crosslinking, thus resulting in an internal structure that is highly amorphous and complex, which results in decreased ∆H values.[Citation19]

Figure 4. DSC thermogram of native and crosslinked maize starches. Source: Kou and Gao[Citation33].

Figure 4. DSC thermogram of native and crosslinked maize starches. Source: Kou and Gao[Citation33].

The cross-linking of corn starch with different proportions of STMP/STPP significantly augmented the To, Tp, and Tc values and reduced the ∆H values.[Citation24] Similarly, Gonenc and Us[Citation48] studied the different pH and reaction times on thermal attributes of glutaraldehyde crosslinked corn starch. The crosslinking under both acidic and alkaline mediums broadened the gelatinization temperature of corn starch. A crosslinking in an acidic medium induced an augment in To, Tp, and Tc, whereas a crosslinking in an alkaline medium caused a reduction in To and Tp. The ∆H value is also decreased in both starches crosslinked under both acidic and alkaline mediums. The cross-linking of cassava starch using STMP/STPP exhibited a concentration-dependent increase in the To and ∆H values.[Citation25] Similarly, the crosslinking of litchi kernel starch with STMP potentially increased its To, Tp, Tc, and ∆H with an increase in STMP concentration.[Citation19] The cross-linking of maize starch with STMP exhibited a shift in To, Tp, and Tc to a higher temperature, whereas it did not exhibit any significant changes in ∆H value.[Citation37] Likewise, Falsafi et al.[Citation49] also reported that crosslinking of normal maize starch with conventional and ultrasound-assisted methods elevated the To, Tp, and Tc values and reduced the ∆H value. The researchers concluded that the introduction of covalent bonds into starch molecules tightens their lattice structure and increases their stability, thus increasing the gelatinization temperature. At the same time, the reduction in ΔH might be due to the breakdown of ordered segments of the granules as a result of acute crosslinking conditions. The cross-linking of banana starch using sodium hexametabisulpates significantly increased the To value. However, it did not show any significant difference in Tp and Tc values and decreased the ∆H value.[Citation40]

The cross-linking of corn, pea, and faba bean starch with POCl3 and STMP in aqueous and semidry forms was studied by.[Citation50] The crosslinked starches exhibited potential changes in their gelatinization attributes. The crosslinking with POCl3 shifted the endothermic peak to a higher temperature for all three starch samples. Likewise, the STMP treatment in a semidry form significantly decreased the To value and increased the Tp and Tc value of all starches as compared to their native counterpart. There was a significant elevation in To, Tp, and Tc in corn starch, but only an increase in Tc was observed in pulse starch, crosslinked with STMP in aqueous form, indicating that crosslinking agents and crosslinking methods affect the crosslinking of starches differently.

The cross-linking of the waxy maize and waxy rice starches with phytic acid significantly decreased their thermal attributes as compared to control and chemically treated starches. The substitution of the phosphate group destabilizes the starch structure, thus facilitating the easy melting of starch crystals.[Citation51] The STMP/STPP crosslinked barley starch increased the To, Tp, and Tc values, whereas it decreased the ∆H values.[Citation52] The cross-linking of porous starch also potentially impacts its thermal attributes. Cyperus esculentus porous starch crosslinked using sodium phytate decreased its To and Tc and ∆H values, whereas it did not significantly impact the Tp value.[Citation53] Conversely, the crosslinking of porous corn starch with epichlorohydrin shifted its To, Tp, and Tc to a higher temperature.[Citation54] In conclusion, cross liked starches reduce granule rupture, loss of viscosity, and the formation of a stringy paste during heating, providing a starch suitable for canned foods and products.

Morphological properties

Scanning electron microscopy is one of the most effective tools for analyzing and interpreting surface morphologies of crosslinked starches. The changes in the granule morphology of crosslinked starches vary with the degree of crosslinking. The cross-linking of faba bean, pea, and corn starches with different crosslinking agents, i.e., POCl3 and STMP (semidry and aqueous form), did not cause remarkable changes in granule morphology, except for a few fissures and smooth indentations on some starch granules ().[Citation50] Similarly, STMP crosslinking of corn starch with a lower concentration (1 and 5%) did not cause noticeable changes in granule morphology; however, a higher concentration (10%) led to the agglomeration of starch granules.[Citation24]

Figure 5. (a) Micrograph of native and crosslinked (1) corn, (2) faba bean, and (3) field pea starches using different crosslinking agents. Native (A), POCl3-aqueous (B), STMP-semidry (C), and STMP-aqueous (D).

Source:[Citation50]
Figure 5. (a) Micrograph of native and crosslinked (1) corn, (2) faba bean, and (3) field pea starches using different crosslinking agents. Native (A), POCl3-aqueous (B), STMP-semidry (C), and STMP-aqueous (D).

The cross-linking of talipot starch using STMP/STPP (99:1) caused surface roughness and slight surface erosion on granule surfaces.[Citation44] Besides, the STMP crosslinking of maize starch maintained its intact shape with the formation of cracks on granule surfaces.[Citation37] The STMP/STPP crosslinking of cassava starch did not cause any remarkable changes in the granule surfaces.[Citation55] Meanwhile, the STMP/STPP crosslinking of barley starch made its granule surface rough, and smooth dents were formed.[Citation52] The cross-linking of maize starch with STMP/STPP mixtures did not cause any remarkable changes in granule shape and size.[Citation49]

The cross-linking of sorghum starch with varying levels of epichlorohydrin showed a rough surface, cavities, and cracks on their granules, and the effectiveness increased with an increase in concentration from .1 to 1% .[Citation36] Similarly, the epichlorohydrin crosslinking of mandua starch made its granule surface slightly rough, and the formation of pores or grooves was observed on some of the granules.[Citation8] Crosslinking of kithul starch with POCl3 caused the formation of deep grooves on granule surfaces.[Citation30] Similarly, the crosslinking of talipot starch using POCl3 showed roughness and erosion on the granule surface.[Citation31]

Figure 5b. Micrograph of native and epichlorohydrin (EPI) crosslinked sorghum starch: A- 0% EPI, .1% EPI, C- .3% EPI, .5% EPI, and E- 1.0% EPI. Source: Sandhu et al.[Citation36].

Figure 5b. Micrograph of native and epichlorohydrin (EPI) crosslinked sorghum starch: A- 0% EPI, .1% EPI, C- .3% EPI, .5% EPI, and E- 1.0% EPI. Source: Sandhu et al.[Citation36].

The confocal electron microscopy of STMP/STPP cross-linked corn starch showed that the granular area of modified starch became dark and increased with an increase in reagent concentration.[Citation24] Alterations in starch granule structure due to increased treatment concentration may affect light absorption and cause darkening. Likewise, the photomicrograph of citric acid crosslinked Cyperus starch revealed the formation of granule aggregates, and it increased with an increase in citric acid concentration.[Citation5]

Retrogradation

The retrogradation characteristics of crosslinked starch are generally studied in terms of paste clarity, light transmittance, and syneresis. The crosslinking of starch significantly alters its retrogradation characteristics. The STMP/STPP crosslinked barley starch exhibited a higher light transmittance and reduced retrogradation tendency than the native counterpart.[Citation56] Similarly, restricted syneresis and retrogradation were observed in STMP crosslinked tigernut starch.[Citation20] The cross-linking of cassava starch with a lower concentration of STMP/STPP considerably reduced its rate of retrogradation. However, a higher concentration of STMP/STPP did not show significant changes from the native starch. Low-concentration crosslinking limits the mobility of starch chains and could delay retrogradation. However, a higher concentration brings a more closely packed structure, enhancing the ability of hydrogen bonding between starch chains and increasing the rate of retrogradation.[Citation25] The cross-linking of lotus seed starch with different concentrations of STMP exhibited a higher light transmittance and a lower rate of retrogradation.[Citation23] Similarly, the retrogradation ability of maize starch was also decreased by crosslinking with STMP.[Citation37] Park and Lim[Citation51] reported that crosslinking of waxy rice and waxy wheat starch decreased their enthalpy of retrogradation. It could be due to the suppression of retrogradation by cross-linking.

The cross-linking of sorghum starch with epichlorohydrin enhanced its stability and diminished the retrogradation.[Citation36] Similarly, the crosslinking of Cyperus esculentus porous starch with sodium phytate exhibited higher stability during retrogradation than the native and porous starches.[Citation53] On the contrary, the POCL3 crosslinked kithul starch significantly increased the rate of retrogradation.[Citation30] It was claimed that the realignment of starch chains leached through the surface damages that occurred during modification enhanced its retrogradation.

Freeze-thaw stability

The stability of starch-containing food products during the freeze-thaw process is an important trait affecting their textural properties. During freezing, the water in the starch changes its phase to ice, and during thawing, the resultant water is easily released from the matrix, known as syneresis. The phase separation leads to the reassociation and recrystallization of starch chains. Hence, the repetitive freeze-thaw causes phase separation, causing syneresis and retrogradation.[Citation57] The crosslinking of starch with different crosslinking agents potentially affects its freeze-thaw stability. The cross-linking of turmeric starch with STMP significantly increased the freeze-thaw stability and decreased the syneresis value considerably.[Citation58] Similarly, the crosslinking of sago starch with STMP,[Citation59] tigernut starch with STMP,[Citation20] rice starch with POCl3,[Citation60] sago starch with citric acid[Citation39] breadfruit starch with STMP/STPP,[Citation26] sago starch with STMP[Citation59] also improved their freeze-thaw stability.

The treatment of chestnut starch with two hydrated sodium dihydrogen phosphate significantly increased its freeze-thaw stability.[Citation61] The introduction of the phosphate group into the starch molecule led to the formation of a three-dimensional network, which offered steric hindrance to the starch molecule and made the system more stable.[Citation20] On the contrary, the crosslinking of chestnut starch with STMP/STPP reduced its freeze-thaw stability. The higher syneresis of crosslinked starch could be due to the restricted swelling of crosslinked starch, leading to the release of a relatively large quantity of water[Citation62]

Applications of cross-linked starches

Bakery products

The crosslinked starch has gained much popularity in baked food products like biscuits, pies, bread, cakes, etc, for various reasons. It may enhance the texture, shelf life, and overall quality of the baked products. Crosslinked starches are more resistant to oven temperature; hence, it is predominantly used as ingredients in baked products[Citation63]. In the bread manufacturing process, starch can absorb ≈ 64% water, and in dough development, starch acts as an inert filler in a continuous protein matrix. It also performs as a bi-continuous network of protein and starch in dough development. Further, the rheological attributes of dough depend on the interaction of hydrated gluten and aggregated granular starch. In addition, the granule shapes play a potential part in the retention of the gas bubbles. A small granule could stabilize the gas bubbles, whereas a granule with a size larger than the gas cell wall destabilizes it. Moreover, the texture and quality of bread are affected by retrogradation during storage.[Citation64]

The substitution of flour and native starch with crosslinked starches potentially improves the quality of bread. A substitution of 10% flour with STMP/STPP crosslinked rice and corn starch potentially increased the dough stability, whereas it reduced the WAC and dough development time. Further, it exhibited a higher softness and reduced hardness. In addition, the substitution of crosslinked rice starch augmented the bread volume, evenness, and overall textural attributes, whereas crosslinked corn starch bread exhibited reduced evenness and overall textural attributes.[Citation65] Similarly, Abdul Shukri et al.[Citation66] studied the impact of the addition of crosslinked rice starch on dough rheology and characteristics of bun. It was noted that the substitution of crosslinked starch reduced the water absorption capacity, dough stability, and breakdown time. However, it enhanced the dough development time (at 15% crosslinked starch) and mixing tolerance time. Further, the substitution of crosslinked starch increased the dietary fiber content and reduced the hardness, whereas it did not affect the sensory characteristics of the bun.

The substitution of crosslinked and dual crosslinked maize starch significantly reduced the water binding capacity of control flour, whereas it did not impact the rheological attributes of the dough. Further, a 10% substitution of crosslinked and dual crosslinked maize starch potentially improved the specific volume of bread, whereas it decreased at a 20% substitution. In addition, it diminished the crumb hardness and augmented the cohesiveness. As compared to the control and crosslinked starch, dual crosslinked maize starch significantly reduced the retrogradation ability and rate of bread staling.[Citation67] Like bread, the substitution of crosslinked starches significantly impacts the properties of biscuits. Azaripour and Abbasi[Citation68] reported that the crosslinking reduced the viscosity and improved the thermal stability during baking. Further crosslinked starches reduced the density and increased stiffness of the biscuit produced. In muffins, crosslinked starch with a high swelling ability overwhelms the insoluble structural network development during the cooling process, preventing starch retrogradation and keeping muffins soft[Citation69]. The cross-linked starch can also be explored as a fat replacer in bakery products. Rodriguez-Sandoval et al.[Citation70] explored crosslinked cassava starch as a fat replacer in muffins. The study reported that the substitution of different proportions of crosslinked starch adversely affected dough development and stability. However, the substitution of crosslinked starch (8%) did not cause a significant impact on baking properties, i.e., weight loss, specific volume, crumb moisture content, or color parameters of the muffins. Further, it had a textural characteristic comparable to control and good overall acceptability. Fasuan et al.[Citation71] reported that the crosslinking of amaranth starch with sodium hexametaphosphate significantly improved its alkenyl group, and it indicates that the crosslinked amaranth starch is explored as a substituent of fat in food systems.

Milk and milk products

Starch plays a significant role in dairy-based products; it gives a creamy texture and flavor and acts as a stabilizer. It is widely explored as a thickener in yogurt manufacturing to impart a pleasing texture and reduce defects and cracks in the surface of curd milk.[Citation72] Moreover, starch reduces wheying-off during storage and transportation and boosts the number of solids in yogurt, which is crucial for its viscosity and sensory properties.[Citation73] Pang et al.[Citation74] successfully incorporated the acetylated distarch phosphate into the yogurt and reported that crosslinked starch incorporated yogurt possessed a well-organized and homogenized microstructure as compared to the control and yogurt incorporated with unmodified starch. In addition, it improved the apparent viscosity, thixotropy, pseudo-plasticity, firmness, and adhesiveness of yogurt.

The addition of crosslinked acetylated cassava starch successfully improved the structure and stability of yogurt. It was noticed that cassava starch incorporated yogurt exhibited a higher elastic and viscous modulus, conductivity, and particle size as the concentration of starch increased. Furthermore, it contributes to the structural stability of the yogurt.[Citation75] The improved rheological attributes of the yogurt with added crosslinked starches could be due to the development of starch gel and further strengthening of the starch gel network by crosslinking. It will contribute to the viscoelastic characteristic of the yogurt. Furthermore, the interaction of protein and starch will also contribute to increased rheological attributes of starch-incorporated yogurt, and this network is formed by hydrogen bonding.[Citation75]

Ice cream is a dairy-based product that contains a high proportion of fat. Surendra Babuet al.[Citation32] reported the impact of the addition of different levels of citric acid crosslinked starch (1 and 2%) to ice cream containing different levels of fats (6% and 1%). The researchers noted that the addition of 1% starch exhibited dominant sensory and quality attributes, and it improved the texture of the ice cream, which is similar to the control ice cream (containing 11% fat). However, a 2% starch adversely affected the overrun and melting resistance of the ice creams. Similarly, The substitution of different levels of POCl3 crosslinked low and high amylose rice starch (5, 10, and 15%) to the cream potentially altered its freeze-thaw stability and textural attributes. An increase in starch concentration enhanced the firmness and the freeze-thaw stability of the cream, and a 5% fat substitution maintained its creaminess.

Soup and sauce

Starch is a natural thickening agent commonly used to give body and consistency to the products; it is explored in soups and sauces. Corn and tapioca starches are the most commonly used as the thickening agent in soup and sauce. Its thickening ability is enhanced by gelatinization, and it has a number of influences, like granule shape, size, amylose content, etc.[Citation76] Nevertheless, native starch exhibits certain limits like high retrogradation, low shearing stability, etc., which cannot meet particular food processing requirements. Hence, the starch is modified to obtain the desirable properties of starch as a thickening agent.[Citation77] Similarly, the interaction or blending of starch with non-starch hydrocarbons like gums improves the rheological attributes.[Citation76] The addition of potato starch/xanthan gum in tomato ketchup exhibited dominant sensory properties, textural attributes, and viscosities when compared with the ketchup with added native and hydroxy propylated potato starch (Cai et al.[Citation78]

The crosslinking of starch enhances its rheological attributes, thermal stability, resistance to shear, paste clarity, and freeze-thaw stability, and all these collectively contribute to the thickening property of crosslinked starch in different food products like soup and sauce.[Citation79] The cross-linking of tapioca starch with different concentrations of STMP/STPP potentially altered its paste clarity, rheological properties, and thermal stability. It was noticed that the addition of 1% crosslinked cassava starch significantly enhanced the gel properties and shear stability. Further, the textural and sensory attributes were augmented by the incorporation of crosslinked cassava starch into the soup.[Citation80] Likewise, the addition of 8% di-starch phosphate barley starch enhanced the consistency of the soup with higher linear viscoelasticity, yield stress, complex viscosity, and lower tan δ value.[Citation56]

Processed meat

In meat industries, starch serves as multifunctional ingredients with diverse functionalities, including binding agent, emulsifying agent, fat replacer, gelling agent, water retention agent, etc.[Citation81] In meat emulsions, heat-induced swelling of starch granules in the protein gel matrix may be beneficial. Furthermore, starch granules will absorb water during gelatinization and augment the water retention ability of the meat products.[Citation82] Wu et al.[Citation83] observed that while heating myofibrillar protein with starch, the starch granules undergo gelatinization and expand and fill the protein network of myofibrillar protein. It leads to the binding of water and the development of the texture of the myofibrillar protein gel.

Starch can be used to increase the water-binding ability of meat patties. Lee and Chin[Citation84] studied the effect of corn starch as a water-binding agent in pork patties and noted that adding corn starch increased the water-binding ability and shear value, decreasing the cooking and thawing loss. Adding modified pea and corn starch as binding agents in low-fat bologna formulation reduced the cooking loss, expressible moisture, and purge loss. Further, it increased the bologna’s hardness, chewiness, and cohesiveness.[Citation82][Citation85] reported that crosslinked and acetylated tapioca starch enhanced the texture and gelation properties of Kung-wan and the rheological attributes of the meat batter. The modified starches were performed as a filler in a gel matrix. The gel made from crosslinked starch is denser and compact due to its restricted swelling. Similar observations were reported when the kung-wan was incorporated by acetylated distarch phosphate tapioca and corn starches.[Citation86] Likewise, The addition of crosslinked esterified potato and tapioca starch significantly increased the strength and water-holding capacity (WHC) of the myofibrillar protein starch gel.[Citation87]

It has been noted that the starch granules, having a size similar to the size of fat emulsions, could be explored as a fat substitute in low-fat food products to reduce the risk of high-fat content. Hence, meat and meat-based products can be substituted by starch and its modified forms as fat substituents.[Citation81] Garcia-Santos et al.[Citation88] successfully substituted the resistant starch as a fat substituent in beef sausage and reported that the fat substitution did not impact the centesimal composition, texture, and sensory attributes of the beef sausage. Similarly, Jairath et al.[Citation89] also successfully used corn starch as a fat substitute in buffalo calf meat sausage. It was noted that successive replacement of fat with corn starch ameliorated the sausage’s WHC, emulsion stability, texture, juiciness, etc., and 6% fat substitution exhibited optimal properties.

Regulatory considerations

Cross-linking, on the other hand, is regarded as an effective route of starch modification. It can incorporate covalent bonds and augment the naturally occurring intermolecular hydrogen bonds, enhancing the industrial applicability and versatility of starch by enhancing its mechanical properties. However, some of the chemicals used for alteration have been seen to have a detrimental impact on human health when used in an unregulated amount, i.e., in excess or beyond the limit. The United States Food and Drug Administration (US FDA) plays a crucial role in ensuring the safety and quality of food-grade starch, particularly when it comes to the application of cross-linked starch in various food products. To uphold consumer health and the transparency of food products, the US FDA framed 21 CFR 172.892 to ensure the use of chemically modified starch in foods in a controlled manner. These regulatory considerations are essential to maintaining product quality, ensuring that additives used in the cross-linking process are safe, and transparently disclosing their presence on product labels. One key aspect involves obtaining approval for the additives used in the cross-linking process, ensuring they are Generally Recognized as Safe (GRAS), particularly in the United States under FDA oversight. The USFDA specifies that not more than .1, 1, and .12% Degree of substitution (DS) (w/w of starch) of phosphoryl chloride, sodium trimetaphosphate, and adipic acetic mixed anhydride, respectively, should be used for food-grade starch. Further, the permitted maximum level of phosphorus (P) content of .4% in cross-linked starch using a mixture of STMP and STPP.[Citation90] Quality and performance standards are often in place, ensuring that cross-linked starch doesn’t compromise the texture, viscosity, or overall stability of food products. These regulatory measures collectively serve to maintain the integrity and safety of food products in which cross-linked starch is used, benefiting both manufacturers and consumers alike.

Additionally, accurate labeling is critical, requiring manufacturers to transparently disclose the presence of cross-linked starch on product labels, providing vital information to consumers, especially those with dietary restrictions or allergies.

Conclusion

In recent years, the modified starch industry has experienced substantial growth and innovation, driven by the increasing demand for its diverse applications in both the food and non-food sectors. One of the key advancements in this industry is the practice of cross-linking starch, which plays a pivotal role in enhancing its functionality and versatility. Cross-linking starch is a process that involves the introduction of covalent bonds within the starch structure, typically achieved through the use of various cross-linking agents. Cross-linking of starch with various cross-linking agents in the starch structure makes it resistant to high temperatures and shear forces, allowing it to withstand rigorous processing conditions in food manufacturing. and more suitable for applications like salad dressings, canned foods, etc. In applications like deep-fried foods, pudding, fruit filling, gravies, soups, sauces, baby foods, etc., crosslinked starches are widely applied. Further, cross-linked starches can be used to create heat-stable gels in convenience foods or as a fat replacer in low-fat products. The process of cross-linking starch represents a pivotal innovation within the modified starch industry, empowering starch to excel in demanding processing conditions while expanding its applications. Numerous food applications have solidified cross-linked starches as a cornerstone of modern food manufacturing, contributing to the development of higher quality, healthier, and more convenient food products.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Min, Y.; Yi, J.; Dai, R.; Liu, W.; Chen, H. A Novel Efficient Wet Process for Preparing Cross-Linked Starch: Impact of Urea on Cross-Linking Performance. Carbohydr. Polym. 2023, 320, 121247. DOI: 10.1016/j.carbpol.2023.121247.
  • Bangar, S. P.; Whiteside, W. S.; Dunno, K. D.; Cavender, G. A.; Dawson, P.; Love, R. Starch-Based Bio-Nanocomposites Films Reinforced with Cellulosic Nanocrystals Extracted from Kudzu (Pueraria Montana) Vine. Int. J. Biol. Macromol. 2022, 203, 350–360. DOI: 10.1016/j.ijbiomac.2022.01.133.
  • Bhatt, P.; Kumar, V.; Goel, R.; Sharma, S. K.; Kaushik, S.; Sharma, S.; Tesema, M. Structural Modifications and Strategies for Native Starch for Applications in Advanced Drug Delivery. Biomed Res. Int. 2022, 2022, 1–14. DOI: 10.1155/2022/2188940.
  • Gebresas, G. A.; Szabó, T.; Marossy, K. Effects of Acidity, Number of Hydroxyl Group, and Carbon Chain Length of Carboxylic Acids on Starch Cross-Linking. Curr. Res. Green Sustainable Chem. 2023, 6, 100354. DOI: 10.1016/j.crgsc.2022.100354.
  • Olayemi, B.; Isimi, C. Y.; Ekere, K.; Isaac, A. J.; Okoh, J. E.; Emeje, M. Green Preparation of Citric Acid Crosslinked Starch for Improvement of Physicochemical Properties of Cyperus Starch. Turk. J. Pharm. Sci. 2021, 18(1), 34. DOI: 10.4274/tjps.galenos.2019.65624.
  • Ding, L.; Huang, Q.; Xiang, W.; Fu, X.; Zhang, B.; Wu, J. Y. Chemical Cross-Linking Reduces in vitro Starch Digestibility of Cooked Potato Parenchyma Cells. Food. Hydrocol. 2022, 124, 107297. DOI: 10.1016/j.foodhyd.2021.107297.
  • Klostermann, C. E.; Buwalda, P. L.; Leemhuis, H.; de Vos, P.; Schols, H. A.; Bitter, J. H. Digestibility of Resistant Starch Type 3 is Affected by Crystal Type, Molecular Weight and Molecular Weight Distribution. Carbohydr. Polym. 2021, 265, 118069. DOI: 10.1016/j.carbpol.2021.118069.
  • Malik, M. K.; Kumar, V.; Sharma, P. P.; Singh, J.; Fuloria, S.; Subrimanyan, V.; Fuloria, N. K.; Kumar, P. Improvement in Digestion Resistibility of Mandua Starch (Eleusine Coracana) After Cross-Linking with Epichlorohydrin. ACS. Omega. 2022, 7(31), 27334–27346. DOI: 10.1021/acsomega.2c02327.
  • Yao, S.; Wang, B. J.; Weng, Y. M. Preparation and Characterization of Mung Bean Starch Edible Films Using Citric Acid as Cross-Linking Agent. Food Pack. Shelf Life. 2022, 32, 100845. DOI: 10.1016/j.fpsl.2022.100845.
  • Gao, F.; Li, D.; Bi, C. H.; Mao, Z. H.; Adhikari, B. Preparation and Characterization of Starch Crosslinked with Sodium Trimetaphosphate and Hydrolyzed by Enzymes. Carbohydr. Polym. 2014, 103, 310–318. DOI: 10.1016/j.carbpol.2013.12.028.
  • Delley, R. J.; O’Donoghue, A. C.; Hodgson, D. R. Hydrolysis Studies of Phosphodichloridate and Thiophosphodichloridate Ions. J. Org. Chem. 2012, 77(13), 5829–5831. DOI: 10.1021/jo300808m.
  • Shah, N.; Mewada, R. K.; Mehta, T. Crosslinking of Starch and Its Effect on Viscosity Behaviour. Rev. Chem. Eng. 2016, 32(2), 265–270. DOI: 10.1515/revce-2015-0047.
  • Desam, G. P.; Li, J.; Chen, G.; Campanella, O.; Narsimhan, G. Prediction of Swelling Behavior of Crosslinked Maize Starch Suspensions. Carbohydr. Polym. 2018, 199, 331–340. DOI: 10.1016/j.carbpol.2018.07.020.
  • Polnaya, F. J.; Marseno, D. W.; Cahyanto, M. N. Effects of Phosphorylation and Cross-Linking on the Pasting Properties and Molecular Structure of Sago Starch. Int. Food Res. J. 2013, 20(4), 1609–1615.
  • Zhang, B.; Tao, H.; Wei, B.; Jin, Z.; Xu, X.; Tian, Y.; Khodarahmi, R. Characterization of Different Substituted Carboxymethyl Starch Microgels and Their Interactions with Lysozyme. PloS One. 2014, 9(12), e114634. DOI: 10.1371/journal.pone.0114634.
  • Gerezgiher, A. G.; Szabó, T. Crosslinking of Starch Using Citric Acid. J. Phys.: Conf. Ser. 2022, 2315(1), 012036. IOP Publishing. DOI: 10.1088/1742-6596/2315/1/012036.
  • Uliniuc, A.; Hamaide, T.; Popa, M.; Băcăiță, S. Modified Starch-Based Hydrogels Cross-Linked with Citric Acid and Their Use as Drug Delivery Systems for Levofloxacin. Soft. Mater. 2013, 11(4), 483–493. DOI: 10.1080/1539445X.2012.710698.
  • Ačkar, Đ.; Babić, J.; Jozinović, A.; Miličević, B.; Jokić, S.; Miličević, R.; Rajič, M.; Šubarić, D. Starch Modification by Organic Acids and Their Derivatives: A Review. Molecules. 2015, 20(10), 19554–19570. DOI: 10.3390/molecules201019554.
  • Sharma, V.; Kaur, M.; Sandhu, K. S.; Kaur, S.; Nehra, M. Barnyard Millet Starch Cross-Linked at Varying Levels by Sodium Trimetaphosphate (STMP): Film Forming, Physico-Chemical, Pasting and Thermal Properties. Carbohydr. Polym. Technol. Appl. 2021, 2, 100161. DOI: 10.1016/j.carpta.2021.100161.
  • Xiaofan, L.; Chen, Y.; Zhou, W. Effect of Cross-Linking with Sodium Trimetaphosphate on Structural and Physicochemical Properties of Tigernut Starch. Food Sci. Technol. 2022, 42, 1–5. DOI: 10.1590/fst.76422.
  • Iweajunwa, S. I.; Achugasim, O.; Ogali, R. E. Effects of Acetylation, Carboxymethylation and Crosslinking on Some Physicochemical Properties of Starch from Tubers of Icacina Senegalensis and Cyrtosperma Senegalense. Scientia. Africana. 2023, 22(1), 75–84. DOI: 10.4314/sa.v22i1.8.
  • Sharma, V.; Kaur, M.; Sandhu, K. S.; Godara, S. K. Effect of Cross-Linking on Physico-Chemical, Thermal, Pasting, in vitro Digestibility and Film Forming Properties of Faba Bean (Vicia Faba L.) Starch. Int. J. Biol. Macromol. 2020, 159, 243–249. DOI: 10.1016/j.ijbiomac.2020.05.014.
  • Chandak, A.; Dhull, S. B.; Punia Bangar, S.; Rusu, A. V. Effects of Cross-Linking on Physicochemical and Film Properties of Lotus (Nelumbo Nucifera G.) Seed Starch. Foods. 2022, 11(19), 3069. DOI: 10.3390/foods11193069.
  • Ge, X.; Guo, Y.; Zhao, J.; Zhao, J.; Shen, H.; Yan, W. Dielectric Barrier Discharge Cold Plasma Combined with Cross-Linking: An Innovative Way to Modify the Multi-Scale Structure and Physicochemical Properties of Corn Starch. Int. J. Biol. Macromol. 2022, 215, 465–476. DOI: 10.1016/j.ijbiomac.2022.06.060.
  • Sriprablom, J.; Tatikunakorn, P.; Lerdpriyanun, P.; Suphantharika, M.; Wongsagonsup, R. Effect of Single and Dual Modifications with Cross-Linking and Octenylsuccinylation on Physicochemical, in-Vitro Digestibility, and Emulsifying Properties of Cassava Starch. Food Res. Int. 2023, 163, 112304. DOI: 10.1016/j.foodres.2022.112304.
  • Amorim, T. S.; Andrade, I. H. P.; Otoni, C. G.; Camilloto, G. P.; Cruz, R. S. Tailoring Breadfruit (Artocarpus Altilis) Starch: Cross‐Linking Starch from This Non‐Conventional Source Towards Improved Technologically Relevant Properties and Enabled Food Applications. Starch‐Stärke. 2021, 73(11–12), 2100058. DOI: 10.1002/star.202100058.
  • Navaf, M.; Sunooj, K. V. Impact of Different Cross-Linking Agents on Functional, Rheological, and Structural Properties of Talipot Palm Starch: A New Source of Stem Starch. Bio. & Life Sci. Forum. November 2022, 20(1): 16.
  • Li, M. N.; Xie, Y.; Chen, H. Q.; Zhang, B. Effects of Heat-Moisture Treatment After Citric Acid Esterification on Structural Properties and Digestibility of Wheat Starch, A-And B-Type Starch Granules. Food. Chem. 2019, 272, 523–529. DOI: 10.1016/j.foodchem.2018.08.079.
  • Remya, R.; Jyothi, A. N.; Sreekumar, J. Effect of Chemical Modification with Citric Acid on the Physicochemical Properties and Resistant Starch Formation in Different Starches. Carbohydr. Polym. 2018, 202, 29–38. DOI: 10.1016/j.carbpol.2018.08.128.
  • Sudheesh, C.; Sunooj, K. V.; Alom, M.; Kumar, S.; Sajeevkumar, V. A.; George, J. Effect of Dual Modification with Annealing, Heat Moisture Treatment and Cross-Linking on the Physico-Chemical, Rheological and in vitro Digestibility of Underutilised Kithul (Caryota Urens) Starch. J. Food Meas. Charact. 2020, 14(3), 1557–1567. DOI: 10.1007/s11694-020-00404-5.
  • Aaliya, B.; Sunooj, K. V.; Rajkumar, C. B. S.; Navaf, M.; Akhila, P. P.; Sudheesh, C.; Lackner, M. Effect of Thermal Pretreatments on Phosphorylation of Corypha Umbraculifera L. Stem Pith Starch: A Comparative Study Using Dry-Heat, Heat-Moisture, and Autoclave Treatments. Polymers. 2021, 13(21), 3855. DOI: 10.3390/polym13213855.
  • Surendra Babu, A.; Parimalavalli, R.; Jagan Mohan, R. Effect of Modified Starch from Sweet Potato as a Fat Replacer on the Quality of Reduced Fat Ice Creams. J. Food Meas. Charact. 2018, 12(4), 2426–2434. DOI: 10.1007/s11694-018-9859-4.
  • Kou, T.; Gao, Q. A Study on the Thermal Stability of Amylose-Amylopectin and Amylopectin-Amylopectin in Cross-Linked Starches Through Iodine Binding Capacity. Food. Hydrocoll. 2019, 88, 86–91. DOI: 10.1016/j.foodhyd.2018.09.028.
  • Wang, Y.; He, M.; Wu, Y.; Liu, Y.; Ouyang, J. Effect of Crosslinking Agents on the Physicochemical and Digestive Properties of Corn Starch Aerogel. Starch‐Stärke. 2021, 73(3–4), 2000161. DOI: 10.1002/star.202000161.
  • Kapelko, M.; Zięba, T.; Michalski, A.; Gryszkin, A. Effect of Cross-Linking Degree on Selected Properties of Retrograded Starch Adipate. Food. Chem. 2015, 167, 124–130. DOI: 10.1016/j.foodchem.2014.06.096.
  • Sandhu, K. S.; Siroha, A. K.; Punia, S.; Sangwan, L.; Nehra, M.; Purewal, S. S. Effect of Degree of Cross Linking on Physicochemical, Rheological and Morphological Properties of Sorghum Starch. Carbohydr. Polym. Technol. Appl. 2021, 2, 100073. DOI: 10.1016/j.carpta.2021.100073.
  • Jia, S.; Yu, B.; Zhao, H.; Tao, H.; Liu, P.; Cui, B. Physicochemical Properties and in vitro Digestibility of Dual‐Modified Starch by Cross‐Linking and Annealing. Starch‐Stärke. 2022, 74(1–2), 2100102. DOI: 10.1002/star.202100102.
  • Tesfay, D.; Abrha, S.; Yilma, Z.; Woldu, G.; Molla, F. Preparation, Optimization, and Evaluation of Epichlorohydrin Cross-Linked Enset (Ensete Ventricosum (Welw.) Cheeseman) Starch as Drug Release Sustaining Excipient in Microsphere Formulation. Biomed Res. Int. 2020, 2020, 1–19. BioMed research international. DOI: 10.1155/2020/2147971.
  • Marta, H.; Hasya, H. N. L.; Lestari, Z. I.; Cahyana, Y.; Arifin, H. R.; Nurhasanah, S. Study of Changes in Crystallinity and Functional Properties of Modified Sago Starch (Metroxylon Sp.) Using Physical and Chemical Treatment. Polymers. 2022, 14(22), 4845. DOI: 10.3390/polym14224845.
  • Olawoye, B.; Fagbohun, O. F.; Popoola, O. O.; Gbadamosi, S. O.; Akanbi, C. T. Understanding How Different Modification Processes Affect the Physiochemical, Functional, Thermal, Morphological Structures and Digestibility of Cardaba Banana Starch. Int. J. Biol. Macromol. 2022, 201, 158–172. DOI: 10.1016/j.ijbiomac.2021.12.134.
  • Radi, M.; Abedi, E.; Najafi, A.; Amiri, S. The Effect of Freezing-Assisted Cross-Linking on Structural and Rheological Properties of Potato Starch. Int. J. Biol. Macromol. 2022, 222, 2775–2784. DOI: 10.1016/j.ijbiomac.2022.10.057.
  • Kou, T.; Song, J.; Liu, M.; Fang, G. Effect of Amylose and Crystallinity Pattern on the Gelatinization Behavior of Cross-Linked Starches. Polymers. 2022, 14(14), 2870. DOI: 10.3390/polym14142870.
  • Sharma, V.; Kaur, M.; Sandhu, K. S.; Nain, V.; Janghu, S. Physicochemical and Rheological Properties of Cross‐Linked Litchi Kernel Starch and Its Application in Development of Bio‐Films. Starch‐Stärke. 2021, 73(7–8), 2100049. DOI: 10.1002/star.202100049.
  • Aaliya, B.; Sunooj, K. V.; John, N. E.; Navaf, M.; Akhila, P. P.; Sudheesh, C.; Sabu, S.; Sasidharan, A.; Mir, S. A.; George, J. Impact of Microwave Irradiation on Chemically Modified Talipot Starches: A Characterization Study on Heterogeneous Dual Modifications. Int. J. Biol. Macromol. 2022, 209, 1943–1955. DOI: 10.1016/j.ijbiomac.2022.04.172.
  • Dong, H.; Vasanthan, T. Effect of Phosphorylation Techniques on Structural, Thermal, and Pasting Properties of Pulse Starches in Comparison with Corn Starch. Food. Hydrocolloids. 2020b, 109, 106078. DOI: 10.1016/j.foodhyd.2020.106078.
  • Siroha, A. K.; Sandhu, K. S. Physicochemical, Rheological, Morphological, and in vitro Digestibility Properties of Cross-Linked Starch from Pearl Millet Cultivars. Int. J. Food Prop. 2018, 21(1), 1371–1385. DOI: 10.1080/10942912.2018.1489841.
  • Chakraborty, I.; N, P.; Mal, S. S.; Paul, U. C.; Rahman, M. H.; Mazumder, N. An Insight into the Gelatinization Properties Influencing the Modified Starches Used in Food Industry: A Review. Food Bioprocess. Technol. 2022, 15(6), 1195–1223. DOI: 10.1007/s11947-022-02761-z.
  • Gonenc, I.; Us, F. Effect of Glutaraldehyde Crosslinking on Degree of Substitution, Thermal, Structural, and Physicochemical Properties of Corn Starch. Starch‐Stärke. 2019, 71(3–4), 1800046. DOI: 10.1002/star.201800046.
  • Falsafi, S. R.; Maghsoudlou, Y.; Aalami, M.; Jafari, S. M.; Raeisi, M. Physicochemical and Morphological Properties of Resistant Starch Type 4 Prepared Under Ultrasound and Conventional Conditions and Their in-Vitro and in-Vivo Digestibilities. Ultrason. Sonochem. 2019, 53, 110–119. DOI: 10.1016/j.ultsonch.2018.12.039.
  • Dong, H.; Vasanthan, T. Amylase Resistance of Corn, Faba Bean, and Field Pea Starches as Influenced by Three Different Phosphorylation (Cross-Linking) Techniques. Food. Hydrocoll. 2020a, 101, 105506. DOI: 10.1016/j.foodhyd.2019.105506.
  • Park, E. Y.; Lim, S. T. Characterization of Waxy Starches Phosphorylated Using Phytic Acid. Carbohydr. Polym. 2019, 225, 115225. DOI: 10.1016/j.carbpol.2019.115225.
  • Mehfooz, T.; Ali, T. M.; Hasnain, A. Effect of Cross-Linking on Characteristics of Succinylated and Oxidized Barley Starch. J. Food Meas. Charact. 2019, 13(2), 1058–1069. DOI: 10.1007/s11694-018-00021-3.
  • Wang, J.; Zhang, D.; Xiao, J.; Wu, X. Effects of Enzymatic Modification and Cross-Linking with Sodium Phytate on the Structure and Physicochemical Properties of Cyperus esculentus Starch. Foods. 2022, 11(17), 2583. DOI: 10.3390/foods11172583.
  • Zhou, D.; Li, D.; Liu, M.; Zhong, X.; Wei, H.; Wang, Z.; Cui, D. Experimental Parameters Affecting Cross-Linking Density and Free-Thaw Stability of Cross-Linked Porous Starch. ES Food & Agroforestry. 2021, 5, 20–28.
  • Korkut, A.; Kahraman, K. Production of Cross-Linked Resistant Starch from Tapioca Starch and Effect of Reaction Conditions on the Functional Properties, Morphology, X‑Ray Pattern, FT-IR Spectra and Digestibility. J. Food Meas. Charact. 2021, 15(2), 1693–1702. DOI: 10.1007/s11694-020-00764-y.
  • Mehfooz, T.; Ali, T. M.; Shaikh, M.; Hasnain, A. Characterization of Hydroxypropylated-Distarch Phosphate Barley Starch and Its Impact on Rheological and Sensory Properties of Soup. Int. J. Biol. Macromol. 2020, 144, 410–418. DOI: 10.1016/j.ijbiomac.2019.12.142.
  • Luo, Y.; Cheng, H.; Niu, L.; Xiao, J. Improvement in Freeze‐Thaw Stability of Rice Starch by Soybean Protein Hydrolysates‐Xanthan Gum Blends and Its Mechanism. Starch‐Stärke. 2022, 74(1–2), 2100193. DOI: 10.1002/star.202100193.
  • Nakkala, K.; Godiyal, S.; Ettaboina, S. K.; Laddha, K. S. Chemical Modifications of Turmeric Starch by Oxidation, Phosphorylation, and Succinylation. Starch Stärke. 2022, 74(9–10), 74(9–10), 2200053. DOI: 10.1002/star.202200053.
  • Sondari, D.; Restu, W. K.; Septevani, A. A.; Suryaningrum, R.; Burhani, D.; Widyaningrum, B. A.; Putri, R. Effect of Catalyst and Cross‐Linker Concentrations on the Functional and Chemical Properties of Sago Starch. Starch‐Stärke. 2022, 74(5–6), 2000266. DOI: 10.1002/star.202000266.
  • Bagheri, F.; Radi, M.; Amiri, S. Physicochemical Properties of Low and High Amylose Cross-Linked Rice Starches. Nutr. Food Sci. Res. 2018, 5(4), 31–41. DOI: 10.29252/nfsr.5.4.31.
  • Hu, N.; Li, L.; Tang, E.; Liu, X. Structural, Physicochemical, Textural, and Thermal Properties of Phosphorylated Chestnut Starches with Different Degrees of Substitution. J. Food Process. Preserv. 2020, 44(6), e14457. DOI: 10.1111/jfpp.14457.
  • Oh, S. M.; Kim, H. Y.; Bae, J. E.; Ye, S. J.; Kim, B. Y.; Choi, H. D.; Baik, M. Y. Physicochemical and Retrogradation Properties of Modified Chestnut Starches. Food Sci. Biotechnol. 2019, 28(6), 1723–1731. DOI: 10.1007/s10068-019-00622-8.
  • Egharevba, H. O. Chemical Properties of Starch and Its Application in the Food Industry. Chem. Prop. Starch. 2019, 9.
  • Hadnađev, M.; Dapčević-Hadnađev, T.; Dokić, L. Functionality of Starch Derivatives in Bakery and Confectionery Products. In Biopolymers for Food Design; Alexandru Grumezescu A., Alina-Maria Holban, A., Eds.; Academic Press, 2018; pp 279–311.
  • Akram, N.; Pasha, I.; Huma, N.; Asghar, M. Effect of Modified Cereal Starches on Dough and Bread Quality. Pak. J. Agric. Sci. 2017, 54(1), 145–151. DOI: 10.21162/PAKJAS/17.5741.
  • Shukri, F. S. A.; Refai, S. A.; Shukri, R.; Muhammad, K.; Mustapha, N. A.; Ibadullah, W. Z. W.; Ramli, N. S. Dough Rheology and Physicochemical Properties of Steamed Buns Fortified with Cross-Linked Rice Starch. Bioact Carbohydr. Dietary Fibre. 2017, 12, 1–6. DOI: 10.1016/j.bcdf.2017.10.002.
  • Roman, L.; Reguilon, M. P.; Martinez, M. M.; Gomez, M. The Effects of Starch Cross-Linking, Stabilization and Pre-Gelatinization at Reducing Gluten-Free Bread Staling. LWT. 2020, 132, 109908. DOI: 10.1016/j.lwt.2020.109908.
  • Azaripour, A.; Abbasi, H. Effect of Type and Amount of Modified Corn Starches on Qualitative Properties of Low-Protein Biscuits for Phenylketonuria. Food Science & Nutrition. 2020, 8(1), 281–290. DOI: 10.1002/fsn3.1304.
  • Subroto, E.; Indiarto, R.; Djali, M.; Rosyida, H. D. Production and Application of Crosslinking-Modified Starch as Fat Replacer: A Review. Int. J. Eng. Trends Technol. 2020, 68(12), 26–30. DOI: 10.14445/22315381/IJETT-V68I12P205.
  • Rodriguez-Sandoval, E.; Prasca-Sierra, I.; Hernandez, V. Effect of Modified Cassava Starch as a Fat Replacer on the Texture and Quality Characteristics of Muffins. J. Food Meas. Charact. 2017, 11(4), 1630–1639. DOI: 10.1007/s11694-017-9543-0.
  • Fasuan, T. O.; Gbadamosi, S. O.; Akanbi, C. T. Modification of Amaranth (Amaranthus Viridis) Starch, Identification of Functional Groups, and Its Potentials as Fat Replacer. J. Food Biochem. 2018, 42(5), e12537. DOI: 10.1111/jfbc.12537.
  • Altemimi, A. B. Extraction and Optimization of Potato Starch and Its Application as a Stabilizer in Yogurt Manufacturing. Foods. 2018, 7(2), 14. DOI: 10.3390/foods7020014.
  • Saleh, A.; Mohamed, A. A.; Alamri, M. S.; Hussain, S.; Qasem, A. A.; Ibraheem, M. A. Effect of Different Starches on the Rheological, Sensory and Storage Attributes of Non-Fat Set Yogurt. Foods. 2020, 9(1), 61. DOI: 10.3390/foods9010061.
  • Pang, Z.; Xu, R.; Luo, T.; Che, X.; Bansal, N.; Liu, X. Physiochemical Properties of Modified Starch Under Yogurt Manufacturing Conditions and Its Relation to the Properties of Yogurt. J. Food Eng. 2019, 245, 11–17. DOI: 10.1016/j.jfoodeng.2018.10.003.
  • Cui, B.; Lu, Y. M.; Tan, C. P.; Wang, G. Q.; Li, G. H. Effect of Cross-Linked Acetylated Starch Content on the Structure and Stability of Set Yoghurt. Food Hydrocolloids. 2014, 35, 576–582. DOI: 10.1016/j.foodhyd.2013.07.018.
  • Mahmood, K.; Kamilah, H.; Shang, P. L.; Sulaiman, S.; Ariffin, F.; Alias, A. K. A Review: Interaction of Starch/non-Starch Hydrocolloid Blending and the Recent Food Applications. Food Biosci. 2017, 19, 110–120. DOI: 10.1016/j.fbio.2017.05.006.
  • Himashree, P.; Sengar, A. S.; Sunil, C. K. Food Thickening Agents: Sources, Chemistry, Properties and Applications-A Review. Int. J. Gastronomy Food Sci. 2022, 27, 100468. DOI: 10.1016/j.ijgfs.2022.100468.
  • Cai, X.; Du, X.; Zhu, G.; Cai, Z.; Cao, C. The Use of Potato Starch/Xanthan Gum Combinations as a Thickening Agent in the Formulation of Tomato Ketchup. CyTA-J. Food. 2020, 18(1), 401–408. DOI: 10.1080/19476337.2020.1760943.
  • Chen, Y. F.; Kaur, L.; Singh, J. Chemical Modification of Starch. In Starch in Food; Sjoo, M., Nilsson, L., Eds.; Woodhead Publishing, 2018; pp 283–321.
  • Wongsagonsup, R.; Pujchakarn, T.; Jitrakbumrung, S.; Chaiwat, W.; Fuongfuchat, A.; Varavinit, S.; Dangtip, S.; Suphantharika, M. Effect of Cross-Linking on Physicochemical Properties of Tapioca Starch and Its Application in Soup Product. Carbohydr. Polym. 2014, 101, 656–665. DOI: 10.1016/j.carbpol.2013.09.100.
  • Kaur, R.; Sharma, M. Cereal Polysaccharides as Sources of Functional Ingredient for Reformulation of Meat Products: A Review. J. Funct. Foods. 2019, 62, 103527. DOI: 10.1016/j.jff.2019.103527.
  • Pietrasik, Z.; Soladoye, O. P. Use of Native Pea Starches as an Alternative to Modified Corn Starch in Low-Fat Bologna. Meat Sci. 2021, 171, 108283. DOI: 10.1016/j.meatsci.2020.108283.
  • Wu, M.; Wang, J.; Hu, J.; Li, Z.; Liu, R.; Liu, Y.; Cao, Y.; Ge, Q.; Yu, H. Effect of Typical Starch on the Rheological Properties and NMR Characterization of Myofibrillar Protein Gel. J. Sci. Food Agric. 2020, 100(1), 258–267. DOI: 10.1002/jsfa.10033.
  • Lee, C. H.; Chin, K. B. Changes in Physicochemical Properties of Pork Myofibrillar Protein Combined with Corn Starch and Application to Low‐Fat Pork Patties. International Journal Of Food Science & Technology. 2020, 55(1), 157–164. DOI: 10.1111/ijfs.14272.
  • Wei, S.; Liang, X.; Kong, B.; Cao, C.; Zhang, H.; Liu, Q.; Wang, H. Investigation of the Effects and Mechanism of Incorporation of Cross-Linked/acetylated Tapioca Starches on the Gel Properties and in vitro Digestibility of Kung-Wan. Meat Sci. 2023a, 204, 109265. DOI: 10.1016/j.meatsci.2023.109265.
  • Wei, S.; Liang, X.; Xu, Y.; Kong, B.; Li, X.; Zhang, H.; Liu, Q.; Wang, H. In-Depth Insight into the Effects of Tapioca or Corn Acetylated Distarch Phosphate on the Gel Properties and in vitro Digestibility of Kung-Wan. Int. J. Biol. Macromol. 2023b, 253, 126997. DOI: 10.1016/j.ijbiomac.2023.126997.
  • Wu, M.; Wang, J.; Ge, Q.; Yu, H.; Xiong, Y. L. Rheology and Microstructure of Myofibrillar Protein–Starch Composite Gels: Comparison of Native and Modified Starches. Int. J. Biol. Macromol. 2018, 118, 988–996. DOI: 10.1016/j.ijbiomac.2018.06.173.
  • Garcia-Santos, M. D. S. L.; ConceiÇÃo, F. S.; Villas Boas, F.; Salotti De Souza, B. M.; Barretto, A. C. D. S. Effect of the Addition of Resistant Starch in Sausage with Fat Reduction on the Physicochemical and Sensory Properties. Food Sci. Technol. 2019, 39(suppl 2), 491–497. DOI: 10.1590/fst.18918.
  • Jairath, G.; Sharma, D. P.; Dabur, R. S.; Singh, P. K.; Bishnoi, S. Standardization of Corn Starch as a Fat Replacer in Buffalo Calf Meat Sausages and Its Effect on the Quality Attributes. Ind. J. Animal Res. 2018, 52(10), 1521–1525. DOI: 10.18805/ijar.B-3381.
  • CFR (Code of Federal Regulations). Food Starch-Modified. Title 21, Chapter1, Part 172, Sec. 172.892. In Food Additives Permitted for Direct Addition to Food for Human Consumption; U.S. Government Printing Office: Washington, DC, 2022.