Next Article in Journal
Novel NBN-Embedded Polymers and Their Application as Fluorescent Probes in Fe3+ and Cr3+ Detection
Next Article in Special Issue
Chemical Composition and Rheological Properties of Seed Mucilages of Various Yellow- and Brown-Seeded Flax (Linum usitatissimum L.) Cultivars
Previous Article in Journal
An Experimental and Numerical Investigation into the Durability of Fibre/Polymer Composites with Synthetic and Natural Fibres
Previous Article in Special Issue
Bacterial Succinoglycans: Structure, Physical Properties, and Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modified Starch-Based Adhesives: A Review

Center of Sustainable Energy and Green Materials and Department of Chemistry, Faculty of Science, Mahidol University, 999 Phuttamonthon Sai 4 Road, Salaya 73170, Thailand
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(10), 2023; https://doi.org/10.3390/polym14102023
Submission received: 31 March 2022 / Revised: 29 April 2022 / Accepted: 6 May 2022 / Published: 16 May 2022
(This article belongs to the Special Issue Polysaccharides: Structure, Properties, and Applications)

Abstract

:
Consumer trends towards environmentally friendly products are driving plastics industries to investigate more benign alternatives to petroleum-based polymers. In the case of adhesives, one possibility to achieve sustainable production is to use non-toxic, low-cost starches as biodegradable raw materials for adhesive production. While native starch contains only hydroxyl groups and has limited scope, chemically modified starch shows superior water resistance properties for adhesive applications. Esterified starches, starches with ester substituents, can be feasibly produced and utilized to prepare bio-based adhesives with improved water resistance. Syntheses of esterified starch materials can involve esterification, transesterification, alkylation, acetylation, succinylation, or enzymatic reactions. The main focus of this review is on the production of esterified starches and their utilization in adhesive applications (for paper, plywood, wood composites, fiberboard, and particleboard). The latter part of this review discusses other processes (etherification, crosslinking, grafting, oxidation, or utilizing biobased coupling agents) to prepare modified starches that can be further applied in adhesive production. Further discussion on the characteristics of modified starch materials and required processing methods for adhesive production is also included.

1. Introduction

The global adhesives and sealants market is growing rapidly and is expected to be worth USD 85.8 billion by 2026 [1]. Industries such as those manufacturing packaging, plywood, and wood composites for furniture, electronic, aerospace and other technological products are fueling this expansion. Petrochemical feedstocks are typically used as the core raw materials to produce formaldehyde, urea, and polyurethanes, which are the key components of many widely used adhesives. However, environmental concerns about pollution from petroleum spills, global warming, and the depletion of fossil fuels have prompted researchers to look at using biodegradable ‘green’ materials as alternatives to petrochemical adhesive components [2].
Natural resources are the basis of ‘green’ raw materials for producing quality adhesives. Lignin, a renewable bio-based aromatic polymer derived from wood, and its derivatives have been utilized as substrates in polyurethanes, which are major adhesive constituents [3]. Functionalized cellulose (cellulose acetate) and castor oil have also been employed to produce bio-based polyurethane adhesives, which have suitable mechanical and adhesion properties for binding to wood surfaces [4]. Plant oil derived fatty acids were used as the basis for monomers which, once polymerized, form pressure-sensitive adhesives [5]. Phenol-formaldehyde resins, which are commonly found in wood adhesives, can be made more environmentally friendly by substitution of the toxic phenol for bio-derived, lower toxicity tannin [6].
Starch is one of the most abundant polysaccharides in nature, consisting of amylose and amylopectin in various ratios depending on the botanical source. Starch has been used extensively in food, medicine, and agricultural products. Due to its low cost, renewable nature and biodegradable quality, it is attracting the attention of plastic industries as a potential building block for petroleum-free adhesives, especially for paper and wood surfaces. However, the hydrophilicity of native starch makes formulating waterproof adhesives from it challenging as the starch hydroxyl groups easily form hydrogen bonds with water [2]. The introduction of hydrophobic functional groups (e.g., esters) onto native starch chains can improve its water resistance, and this can be carried out using chemical, mechanochemical, or enzymatic methods. This work emphasizes the up-to-date starch modification processes that have been used for adhesive applications. The primary focus of this paper is a discussion of the processes used for producing esterified starches, and the properties of the bio-adhesives obtained from such components. Additional sections outline modified starches obtained from etherification, crosslinking, grafting, oxidation, or bio-based coupling agents. The relationship between the properties of modified starch (i.e., degree of substitution,) and the properties (e.g., adhesion bond strength, water resistance, anti-retrodegradation) and types (water-based, solvent-based, and hot melt adhesives) of the obtained adhesives are propose at the end of this paper.

2. Esterified Starch

Esterified starch is one of the important modified starches used in the adhesive and food industries. They are typically synthesized through esterification or transesterification reactions, in which the hydroxyl groups of starch are replaced with more bulky functional groups of free fatty acids or their derivatives (Figure 1).

2.1. Chemical Reactions

2.1.1. Esterification/Transesterification

Esterification is one of the most powerful chemical synthesis strategies for starch modification. These reactions typically utilize alcohols and carboxylic acids in the presence of basic catalysts such as potassium carbonate, sodium hydroxide, and sodium hydrogen phosphate [7], or acid catalysts such as p-toluene sulfuric acid, hydrochloric acid, and sulfuric acid. The mechanism of esterification depends on the mode of catalysis, although studies by Sakamura et al. using isotopically labelled (18O) alcohol unequivically show that in both cases the 18O label is incorporated into the product at the end of the process, indicating bonding of this atom to the carbonyl carbon [8]. The mechanism of transesterification is similar but in this case involves the reaction of an alcohol with an ester, with the end result being loss of an alcohol in conjuction with formation of the esterified starch (Figure 2).
Examples of esterified starch have been reported by many research groups, with variations in ester carbon chain length. Ratnawati et al. [9] utilized glacial acetic acid to modify “Gadung flour” starch, which involved appending acetyl groups onto the saccharide unit in the presence of base catalyst (NaOH). This process afforded a product having a degree of substitution (the moles of substituents per mole of anhydroglucose unit) between 0.16 and 0.20. Under similar conditions starch extracted from Colocasia esculenta, a perennial, tropical plant, reacts with acetic acid, providing a product with higher degree of substitution (0.25–0.65) [10]. Starch citrates, obtained from the esterification of starch with citric acid, have been intensively studied for use in adhesives and food additives due to their antibiotic activity [11]. Moreover, the structure of starch citrates provides for cross-linking to be built into adhesives containing these bio-based polymers [12]. By employing transesterification, isopropyl myristate was reacted with cassava starch under conditions of acid catalysis. In this case the degree of substitution varied between 0.2 and 0.4, depending on reaction time [13]. As shown in the last example, esterification and transesterification strategies can be also used to produce starch modified with long hydrocarbon chains [14] for use in adhesives. However, such reactions can show limitations arising from steric interactions between the starch chain and the long hydrocarbon chain starting material, which results in low degrees of substitution. Enzymatic reactions to afford esterified starch under environmentally friendly mild conditions have also been reported [7,15]. Microwave induced esterification was employed to produce stearate starches with improved thermal stability in comparison to native starch [16].

2.1.2. Acylation/Acetylation/Alkylation

Esterified starches can also be synthesized using acylation or acetylation reactions. In these cases, acylation involves appending the –C(O)R functional group to starch. Acetylation is a subset of acylation, which focuses only on the –C(O)CH3 group [17,18]. Compared to the previous section, acylation/acetylation involves the use of highly active acid chloride or anhydride starting materials, instead of carboxylic acids. Acid chlorides, due to their high relative reactivity [19,20], typically promote a higher degree of substitution than possible for carboxylic acids, and the R group on the acid chloride can be varied from short to long depending on application. Acid anhydrides can be employed in a similar manner [21,22]. However, these reactions have disadvantages in that these reagents are often more expensive and harder to handle than the corresponding carboxylic acids. In addition, the physicochemical properties of acetylated starch may also be affected by the acetylation reaction conditions, e.g., temperatures and pressure [23].
Starch modification through alkylation involves the formation of ether bonds in the product. Two common strategies for this utilize the reaction between starch and an alkyl halide [24,25], or epoxide ring-opening [26,27]. As an example of the former, etherified starch was synthesized through the reaction between allyl chloride and starch in the presence of pyridine. This etherified starch was used in a hydrogel application by reacting with methacrylic acid as a crosslinker [24]. The synthetic method of esterified starch is a major effect to determine the degree of substitution (being discussed in the following section). The comparison of the chemical reactions between esterification and acylation was studied. The acylation reactions promote the high degree of substitution product, but the esterification reactions do not give the high degree of substitution of esterified starch [28]. The reason behind the degree of substitution of product depends on the thermodynamic effect, reactivity of electrophile, and properties of leaving group. Moreover, the catalyst properties play an important role on the degree of substitution [29].

2.1.3. Succinylation

Succinylation refers to the appending of side chains to starch by the reaction with succinic anhydride. Succinic anhydride is an electrophile, which undergoes nucleophilic attack by alcohol functionalities on the starch backbone. The esterified starch product has carboxylic acid functionalities at the terminus of each chain, and this allows the product to be incorporated into the adhesive directly or be further modified or functionalized prior to use. While succinylation results in a high degree of substitution [30,31] the limitation of this reaction is the cost of succinic anhydride. Additionally, the succinylation products have a carboxylic group at the terminal chain that could further promote cross-linking reactions between starch chains [32].

2.1.4. Enzymatic Reactions with Free/Immobilized Lipase

Esterification of starch with fatty acids can be performed using enzyme catalysis, with lipases playing the key role. The use of enzymes has the advantage of not requiring extreme conditions and expensive starting materials, however the degree of substitution is often low [2,15,33] with immobilized lipase on magnetic microparticles providing a higher degree of substitution in the esterified starch than free lipase [34]. Horchani and coworkers also reported lipase immobilized calcium carbonate [35] as a catalyst in the esterification reactions between oleic acid and Maize starch. The yields of modified starch and their degree of substitution are varied, significantly depending on reaction conditions (temperature, pH, and lipase enzyme origin) [36]. Obtaining lipase from either plant of microorganisms requires multi-step isolation, purification and characterization of lipase to ensure its high purity and enzamic activity [35,37]. Additional limiations of enzymes such as low thermostability, pH sensitivity, narrow substrate scope, and low reaction yield in non-aqueous solvents [38], may result in restrictions in starch modification processes.

2.1.5. Mechanochemical Processes

These processes involve reactions taking place without the addition of solvent and utilize mechanical activation, such as ball milling. In the case of starch esterification, ball milling allows for intimate mixing of starch and the esterification agent, giving good yields of product [39]. This processing also disrupts the structure and morphology of the starch, and results in changes in physicochemical properties, compared to those of native starch, such as decreased cold-water solubility, transparency, and emulsion stabilization, along with increased gelatinization temperature and viscosity [32,40,41].

2.2. Properties of Starches

The ratio of amylose to amylopectin in starch is dependent on its botanical source, as shown in Table 1. Amylose is a linear polymer of glucopyranose units linked through α-1,4-glycosidic linkages, in contrast to amylopectin which is a branched polymer. Starch crystallinity depends on the proportion of amylopectin, as amylose is typically amorphous. Differences in amylose content and granule organization influence the gelatinization of the starch, with starch granules containing higher proportions of amylose being easily gelatinized by heating as the chains are only weakly associated through hydrogen bonding [42].

2.2.1. Degree of Substitution

The degree of substitution (DS) in esterified starch is an important factor dictating the physical properties of the material. It may be determined using 1H-NMR spectroscopy or through an acid–base titration process. Examples of the DS of selected modification are listed in Table 2.

Determination of DS via Titration

The degree of substitution (DS) is often determined by acid–base titration due to its convenience [45]. This protocol involves dissolving the esterified starch in 0.5 M NaOH solution under ambient conditions, followed by incubation with shaking (300 rpm, 30 °C) for 4 h. After that, the reaction is titrated against 0.1 M HCl solution, and the degree of substitution (DS) is determined using Equation (1).
DS = (162[HCl](V0 − V1))/(1000 × sample mass (g))
In Equation (1), [HCl] is the concentration of the HCl solution, while V0 stands for the initial volume of hydrochloric acid (V0 > 0), V1 stands for the final volume of hydrochloric acid.

Determination of DS via 1H-NMR Spectroscopy

This method requires identification of specific protons related to the ester units in the product, and integration of their relative areas in 1H NMR spectra. The DS can then be calculated using Equations (2) and (3) [7]. This method requires acetyl group substitution onto the starch unit for DS to be reliably determined. In Equation (3) I is the integration of the 1H NMR peaks corresponding to acetate or starch H-atoms.
DSfatty ester = 3 − DSacetate
DSacetate = IH-acetate/IH-starch

2.2.2. Morphology of Native and Modified Starches

The morphology of starch granules (native and modified) can be studied using scanning electron microscopy (SEM), and this technique provides information about the granule shape, and surface features. Starch granules are the vehicle of carbohydrate storage in plant cells, and the morphology of these depends on the biochemistry of the chloroplast/amyloplasts and the physiology of the plant. Normally, native starch granules are elliptical, have a smooth surface, and range from 1 to 100 μm in size [46]. Modification can result in the granules becoming irregular in shape with pronounced surface roughness, especially at high degrees of substitution (DS) [46,47,48].

2.2.3. Viscosity of Starch Pastes

In addition to the source of starch and conditions used to make the paste, functionalization of the starch can also markedly affect the viscosity of the paste. A study on the paste viscosity of OSA-esterified starch [32,49] indicated that paste obtained from this material was higher than that of native starch, as a consequence of the bulky OSA functional groups. Acylation of starch gives rise to steric interactions between starch chains, and this influences both hydrophilicity and the degree of hydrogen bonding, and thus viscosity. The viscosities of esterified starches generally decrease as the molecular weight of the ester group increases [50].

2.2.4. Solubility/Hydrophobicity

Appending larger functional groups onto the starch units results in increased hydrophobicity and lower water solubility. On the other hand, solubility in lower polarity solvents increases [2] and this impacts the potential application of these starch derivatives [47].

2.2.5. Thermal Properties of Starches

The thermal properties of starch are dependent on the amylose content or amylose/amylopectin ratio [51,52], and are also affected by functionalization. While esterified starches having low degrees of substitution (DS) do not see dramatic changes in onset temperature relative to native starch, starches having higher DS values show a decrease in gelatinization temperature [53]. Furthermore, the glass transition temperature (Tg), the temperature at which a glassy state of amorphous materials converts into a rubbery state, is affected by structure. In the case of functionalized starch, steric hindrance between starch chains increases as the functional groups become larger, resulting in higher Tg. On the other hand, the extension of free volume of chain mobility can reduce Tg [53].
The improved thermal stability of starch can be reported based on their gelatinization, retrogradation, and glass temperature. In Table 2, the properties of starch esters derived from different processes (acylation, acetylation, esterification, transesterification, enzymatic reactions) are highlighted. As indicated, the reagents employed can vastly impact the properties and degree of substitution (DS) in the esterified starch products. The starch esters can be categorized by their DS values into three levels: low (0.01–0.2) which are water soluble, and intermediate (0.2–1.5) and high (1.5–3) which are thermoplastic but more hydrophobic [51]. Oil feedstocks typically afford starch esters having lower DS values than those utilizing very reactive electrophilic substrates such as acid chlorides [47,72] and acid anhydrides [17,18]. In addition, the use of vinyl laureate and vinyl stearate afford broad DS ranges in the presence of base catalysts (Na2HPO4, K2CO3, and sodium acetate), with base strength correlating with DS [62]. When considering the same substrate and conditions, stronger base catalysts result in higher DS values (up to 2.96) [7].
Other works [50,58] also found that the transesterification of starch with fatty acid vinyl ester provides a high DS starch ester products compared to methyl laurate ester [33,40] in both catalyst-free reaction and in the presence of a catalyst. For enzymatic reactions, free enzymatic catalysts using native starch provided starch esters with lower DS values [62] than those produced using immobilized lipase as catalyst [35]. The immobilized catalyst used was composed of lipase and calcium oxide, with the synergy between the two components believed important for enhanced activity [35]. The type of acyl donor also influences the viscosity of the enzymatically modified starches, as the viscosity of the palmitic acid modified starch was found higher than that of the lauric acid modified one [62,63]. Many esterified starches reported to date show higher thermal stability or higher decomposition temperatures than native starch, and greater degrees of hydrophobicity.

2.3. Adhesives Derived from Esterified Starch

Starch-based adhesives are widely used in the production of plywood [30,73] and particleboard [74,75]. Cassava starch esterified using dodecyl succinic anhydride (DDSA) as a reactant [30] and crosslinked with polymethylene polyphenyl polyisocyanate (PAPI) results in a formaldehyde-free industrial plywood adhesive with excellent viscosity, bonding strength, and water resistance properties [30]. Natural corn starch was converted to high amylose corn starch (HACS) by solvent exchange with ethanol and toluene prior to esterification with propionic anhydride in the presence of catalytic 4-dimethylaminopyridine (DMAP). The DS of these starch propionates was in the range 0.38–2.54, with higher ratios of propionic anhydride which enhanced the adhesion strength relatively to native starch. These modified starches were mixed with glycerol and polyvinyl alcohol and hot-pressed to form hot melt adhesives on Al plates. It was found that these covered a greater surface area and exhibited higher tensile strength than adhesives formed analogously from unmodified starch [76]. As mentioned in earlier sections esterified starch from chemical modifications show enhanced hydrophobicity and water resistance, with adhesives formed from this exhibiting improved bonding strength under both dry and wet conditions relative to those from natural starch [77] (Table 3).
Figure 3a illustrates general starch modification processes that have been used in production of starch-based adhesives, while the structure of selected modified starches is given in Figure 3b. Apart from esterified starches, biobased adhesives can be obtained by using other chemically modified starches such as carboxymethyl-, chitosan modified-, and crosslinked starches as raw materials. Details some starch medication methods are given in Table 4 with additional discussion as follows.

3. Other Modified Starches for Adhesive Applications

The application of native starches in adhesives is limited due to their relatively poor shear strength, low thermal stability, and hydrophilicity [84]. Chemical modification is a successful strategy for changing the structure and properties of biopolymers to fit new applications [84] with esterification of starches discussed in the previous section being a prime example. In addition to esterification other modifications of starches can include etherification, oxidation, crosslinking, and grafted copolymerization, as shown in Figure 3 and Table 4.
Etherification can be used to modify the water resistance properties of starch through the introduction of lipophilic functional groups. This is generally accomplished by treating starch with epoxides, such as propylene or ethylene oxides. These epoxides are highly reactive, with the starch OH groups attacking the ring which results in cleavage of the C-O bonds of the epoxide, affording diol products. The lipophilic alkyl groups produced are subsequently linked to the starch chains, thus increasing the hydrophobicity of the starch and improving its water resistance. [85]. Hydroxypropylation, hydroxyethylation, and carboxymethylation are three such etherification procedures for preparation of starch ethers [75]. Carboxymethyl starch (CMS) is widely employed in adhesive manufacturing due to its favorable viscosity and stability [86]. CMS adhesives can be synthesized by wet methods using aqueous media, or dry processes without the use of solvent [87].
Crosslinking, the process of forming nonpolar covalent bonds between the hydroxyl groups of starch, is also a strategy for improving the utility of starches in adhesives. Crosslinked starch exhibits superior mechanical (tensile strength), thermal, and water stability compared to native starch. [88,89,90]. Furthermore, crosslinking acts to inhibit acid degradation, resulting in sustained viscosity in acidic media. A variety of crosslinking agents can be employed in starch modification [91]. Among these citric acid, sodium trimetaphosphate (STMP), sodium tripolyphosphate (STPP), epichlorohydrin (ECH), and phosphorus oxychloride (POCl3) are the most utilized [16]. Detduangchan et al. studied the effects of chemically crosslinking rice starch films using ECH, STMP, and STMP/STPP mixtures. After crosslinking, the tensile strength increased from 5.01 MPa (native starch) to 8.23 MPa, and the water permeability decreased, implying enhanced adhesive water resistance [88].
Another convenient method to modify starch is by grafting synthetic monomers or polymers with desirable properties onto the natural starch backbone. In these instances, it is desirable if the crystallinity and biodegradability properties of the starch are unchanged. Grafting typically occurs at the Cl–C2 end groups, and C2–C3 glycol groups on the starch glucose units. Graft copolymers produced from addition of vinyl- or other monomeric acrylates show improved water resistance and shear strength over natural starch [92]. The use of two co-monomers (vinyl acetate and butyl acrylate) resulted in starch adhesives exhibiting improved shear strength, and requiring lower film-forming temperatures, than those of natural starch [48,93]. Such adhesives retain the biodegradable profile of natural starch while being less hydrophilic, exhibit greater tensile strengths, and maintain a better appearance over time [94].
The unique properties of nanoparticles such as their small size, high surface energy, and the ability to functionalize their surfaces render them attractive for the development of high-performance composite materials. Nano-TiO2, nano-SiO2, and montmorillonite (MMT) nanoparticles have been used as crosslinking or grafting agents to improve the characteristics of biopolymer-based adhesives in recent years [95,96]. Significant improvements in water resistance and bonding strength were found when nanoTiO2, [97,98] saline coupling agents [95], vinyl acetate [93], acrylate [73,93] and polyamide [99,100] were used as crosslinking or grafting agents with the starch for adhesive purposes.
Natural polymers derived from biomass (lignin, cellulose, hemicellulose), and chitosan have the potential to be used as bridging agents in adhesive formulations. These may be useful in systems containing inorganic materials and organic components such as starch, and their use can improve the adhesive properties of the formulation [100,101,102]. Blending starch with hydrophobic biopolymers (such as lignin) could be a strategy for improving the water-resistance of the adhesive and altering its mechanical properties.
Prior to crosslinking or grafting, starch is often pretreated using acid hydrolysis or oxidation [100]. Acid hydrolysis involves breaking the native starch polymer into shorter chained polysaccharides containing amylose and amylopectin subunits through treatment with acid at low temperature. This results in active polysaccharides subunits which react readily with crosslinking agents, affording modified starch networks having superior strength and reduced retrogradation (starch recrystallization upon cooling) [85]. Acid modified starch adhesives have been found to be more water resistant and exhibit superior bonding strengths [41].
Unlike acid hydrolysis, oxidation causes partial depolymerization of native starch, creating free carboxylic or aldehyde groups from the oxidation of primary and secondary hydroxyl groups. The oxidation of native starch involves loss of crystallinity due to a reduction in the number of network hydrogen bonds. The remaining carboxyl and hydroxyl groups subsequently react with other polymers or monomers to form covalent bonds, which result in changes in adhesive properties and starch stability. While periodate, chromic acid, potassium permanganate, nitrogen dioxide, and sodium hypochlorite are often employed, sodium hypochlorite is the widely used oxidant in industry as it results in adhesives having increased tensile strengths [85]. Moreno et al. [103] prepared biodegradable active films from corn starch oxidized using sodium periodate. These films showed enhanced tensile strength and oxygen and water vapor barrier capacities, and reduced migration to acid media when compared with native starch films. Compared to oxidation alone, starch modification by oxidation and subsequent crosslinking can provide adhesives exhibiting higher shear strengths and lower tendency for retrogradation to occur [104].
In addition to the results of synthetic studies, the properties of native starches may be predicted computationally. Computational and mathematical modeling at the molecular level may allow for selection and optimization of reaction conditions for producing adhesives tailored to specific applications. For instance, sulfation of potato starch using formic acid/urea is a simple and environmentally friendly method to obtain starch sulfates [105]. Furthermore, sulfation of starches by using sulfamic acid and urea or other sulfating agents (e.g., H2SO4, HSO3CL, and SO3) could result in the modified starches having biological activity (i.e., anticoagulant, antioxidant, antiviral and anti-inflammatory activity) [105]. Sulfated starches have attracted a great attention in the area of therapeutic agents against chronic diseases [106] rather than as adhesives’ raw materials. The feasibility of utilizing sulfated starches to produce adhesives should be further explored and examined.

4. Conclusions and Perspectives

Utilization of natural resources or bio-based materials as adhesive raw materials could help future societies to become less reliant on hazardous chemicals, volatile organic compounds, and finite petroleum-based chemicals [114]. Formaldehyde-free or biobased adhesives could promote safer working conditions in related industries [115,116]. Although improved water resistance in adhesives can be achieved using modified starches as raw materials instead of native starches, the effects of starch quality on the properties of the adhesives warrants further investigation. Quality variations in natural additives (e.g., lignin, cellulosic materials, chitosan) could also create difficulties in obtaining products of uniform composition and performance, and this may be a barrier to their uptake by industry. Despite these challenges, the potential of modified starches in bio-based adhesive manufacturing is clear. Moving towards environmentally sustainable adhesives requires scaling back the use, or abolition of, volatile organic solvents in adhesive formulations. Much research needs to be conducted on the selection and compatibility of ‘green’ solvents for use alongside adhesive components. Additionally, solvent-free adhesives (such as hot-melt adhesives) derived from insoluble modified starches should be also investigated further as safe, environmentally friendly alternatives.

Author Contributions

Conceptualization, J.W., J.N. and S.M.S.; methodology, J.W. and S.M.S.; writing—original draft preparation, J.W. and J.N.; writing—review and editing, E.E.W. and S.M.S.; supervision, S.M.S.; funding acquisition, S.M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Science and Technology Development Agency (Thailand), Grant no. FDA-CO-2562-10472-TH. J.W. received a scholarship from the Development and Promotion of Science and Technology Talents Project (DPST), Thai Government. S.M.S., J.N. and E.E.W. are financially supported by Mahidol University (Basic Research Fund: fiscal year 2022; Grant no. BRF1-046/2565).

Informed Consent Statement

Not applicable.

Acknowledgments

Authors are grateful for the invaluable discussion with Christopher Smith to improve the manuscript’s quality.

Conflicts of Interest

The authors declare that there is no conflict of interest.

References

  1. Adhesives & Sealants Market by Adhesive Formulating Technology (Water-based, Solvent-based, Hot-melt, Reactive), Sealant Resin Type (Silicone, Polyurethane, Plastisol, Emulsion, Polysulfide, Butyl), Application, Region—Global Forecast to 2026. Available online: https://www.marketsandmarkets.com/Market-Reports/adhesive-sealants-market-421.html?gclid=CjwKCAjwloCSBhAeEiwA3hVo_a_Futuj6Pu555meM1rkldUxwvGxozpuIlUvFIfXlGf0ePpsb70gjhoC8C8QAvD_BwE (accessed on 30 March 2022).
  2. Heinrich, L.A. Future opportunities for bio-based adhesives—Advantages beyond renewability. Green Chem. 2019, 21, 1866–1888. [Google Scholar] [CrossRef] [Green Version]
  3. Alinejad, M.; Henry, C.; Nikafshar, S.; Gondaliya, A.; Bagheri, S.; Chen, N.; Singh, S.K.; Hodge, D.B.; Nejad, M. Lignin-Based Polyurethanes: Opportunities for Bio-Based Foams, Elastomers, Coatings and Adhesives. Polymers 2019, 11, 1202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Tenorio-Alfonso, A.; Sánchez, M.C.; Franco, J.M. Preparation, Characterization and Mechanical Properties of Bio-Based Polyurethane Adhesives from Isocyanate-Functionalized Cellulose Acetate and Castor Oil for Bonding Wood. Polymers 2017, 9, 132. [Google Scholar] [CrossRef] [Green Version]
  5. Maaßen, W.; Oelmann, S.; Peter, D.; Oswald, W.; Willenbacher, N.; Meier, M.A. Novel Insights into Pressure-Sensitive Adhesives Based on Plant Oils. Macromol. Chem. Phys. 2015, 216, 1609–1618. [Google Scholar] [CrossRef]
  6. Pizzi, A. Recent developments in eco-efficient bio-based adhesives for wood bonding: Opportunities and issues. J. Adhes. Sci. Technol. 2006, 20, 829–846. [Google Scholar] [CrossRef]
  7. Junistia, L.; Sugih, A.K.; Manurung, R.; Picchioni, F.; Janssen, L.P.B.M.; Heeres, H.J. Synthesis of Higher Fatty Acid Starch Esters using Vinyl Laurate and Stearate as Reactants. Starch-Starke 2008, 60, 667–675. [Google Scholar] [CrossRef] [Green Version]
  8. Oikawa, H.; Ichihara, A.; Sakamura, S. Biosynthetic Study of Betaenone B: Origin of the Oxygen Atoms and Accumulation of Deoxygenated Intermediate using P-450 Inhibitor. J. Chem. Soc. Chem. Commun. 1988, 600–602. [Google Scholar] [CrossRef]
  9. Kumoro, A.C.; Amalia, R.; Budiyati, C.S.; Retnowati, D.S.; Ratnawati, R. Preparation and characterization of physicochemical properties of glacial acetic acid modified Gadung (Diocorea hispida Dennst) flours. J. Food Sci. Technol. 2015, 52, 6615–6622. [Google Scholar] [CrossRef] [Green Version]
  10. Rosida, D.F.; Yuliani, R.; Djajati, S. Modification of Colocasia esculenta Starch with Acetylation Process. In Proceedings of the 4th International Seminar of Research Month, NST Proceedings, Surabaya, Indonesia, 9 October 2009. [Google Scholar]
  11. Golachowski, A.; Drożdż, W.; Golachowska, M.; Kapelko-Żeberska, M.; Raszewski, B. Production and Properties of Starch Citrates—Current Research. Foods 2020, 9, 1311. [Google Scholar] [CrossRef]
  12. Kapelko-Zeberska, M.; Buksa, K.; Szumny, A.; Zieba, T.; Gryszkin, A. Analysis of molecular structure of starch citrate obtained by a well-stablished method. LWT 2016, 69, 334–341. [Google Scholar] [CrossRef]
  13. Amini, H.W.; Masruri; Ulfa, S.M. Study on Esterification Reaction of Starch Isolated from Cassava (Manihot esculeta) with Acetic Acid and Isopropyl Myrtistate Using Ultrasonicator. In IOP Conference Series: Materials Science and Engineering, Proceedings of the International Conference on Chemistry and Material Science (IC2MS) 2017, Malang, Indonesia, 4–5 November 2017; IOP Publishing: Bristol, UK, 2017. [Google Scholar]
  14. Xie, W.; Wang, Y. Synthesis of high fatty acid starch esters with 1-butyl-3-methylimidazolium chloride as a reaction medium. Starch-Starke 2011, 63, 190–197. [Google Scholar] [CrossRef]
  15. Lukasiewicz, M.; Kowalski, S. Low power microwave-assisted enzymatic esterification of starch. Starch-Starke 2012, 64, 188–197. [Google Scholar] [CrossRef]
  16. Amaraweera, S.M.; Gunathilake, C.; Gunawardene, O.H.P.; Fernando, N.M.L.; Wanninayaka, D.B.; Dassanayake, R.S.; Rajapaksha, S.M.; Manamperi, A.; Fernando, C.A.N.; Kulatunga, A.K.; et al. Development of Starch-Based Materials Using Current Modification Techniques and Their Applications: A Review. Molecules 2021, 26, 6880. [Google Scholar] [CrossRef] [PubMed]
  17. Biswas, A.; Shogren, R.L.; Selling, G.; Salch, J.; Willett, J.L.; Buchanan, C.M. Rapid and environmentally friendly preparation of starch esters. Carbohydr. Polym. 2008, 74, 137–141. [Google Scholar] [CrossRef]
  18. Tupa, M.; Maldonado, L.; Vázquez, A.; Foresti, M.L. Simple organocatalytic route for the synthesis of starch esters. Carbohydr. Polym. 2013, 98, 349–357. [Google Scholar] [CrossRef]
  19. Thiebaud, S.; Aburto, J.; Alric, I.; Borredon, E.; Bikiaris, D.; Prinos, J.; Panayiotou, C. Properties of fatty-acid esters of starch and theirblends with LDPE. J. Appl. Polym. Sci. 1997, 65, 705–721. [Google Scholar] [CrossRef]
  20. Aburto, J.; Alric, I.; Thiebaud, S.; Borredon, E.; Bikiaris, D.; Prinos, J.; Panayiotou, C. Panayiotou Synthesis, Characterization, and Biodegradability of Fatty-Acid Esters of Amylose and Starch. J. Appl. Polym. Sci. 1999, 74, 1440–1451. [Google Scholar] [CrossRef]
  21. Tupa, M.V.; Altuna, L.; Herrera, M.L.; Foresti, M.L. Preparation and Characterization of Modified Starches Obtained in Acetic Anhydride/Tartaric Acid Medium. Starch-Starke 2020, 72, 1900300. [Google Scholar] [CrossRef]
  22. Amos, R.C.; Mesnager, J.; Kuska, M.; Gauthier, M. Production of Cyclic Anhydride-Modified Starches. Polymers 2021, 13, 1504. [Google Scholar] [CrossRef]
  23. Kim, H.S.; Choi, H.S.; Kim, B.Y.; Baik, M.Y. Characterization of Acetylated Corn Starch Prepared under Ultrahigh Pressure (UHP). J. Agric. Food Chem. 2010, 58, 3573–3579. [Google Scholar] [CrossRef]
  24. Bhuniya, S.P.; Rahman, S.; Satyanand, A.J.; Gharia, M.M.; Dave, A.M. Novel route to synthesis of allyl starch and biodegradable hydrogel by copolymerizing allyl-modified starch with methacrylic acid and acrylamide. J. Polym. Sci. A Polym. Chem. 2003, 41, 1650–1658. [Google Scholar] [CrossRef]
  25. Rosu, A.M.; Rafin, C.; Surpateanu, G.; Brabie, G.; Miron, D.N.; Veignie, E. Synthesis of alkylated potato starch derivatives and their potential in the aqueous solubilization of benzo[a]pyrene. Carbohydr. Polym. 2013, 93, 184–190. [Google Scholar] [CrossRef] [PubMed]
  26. Franssen, M.C.R.; Boeriu, C.G. Chemically Modified Starch; Allyl- and Epoxy-Starch Derivatives. In Starch Polymers; Elsevier: Amsterdam, The Netherlands, 2014; pp. 145–184. [Google Scholar]
  27. Fang, J.M.; Fowler, P.A.; Sayers, C.; Williams, P.A. The chemical modification of a range of starches under aqueous reaction conditions. Carbohydr. Polym. 2004, 55, 283–289. [Google Scholar] [CrossRef]
  28. 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. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. 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, 19554–19570. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Sun, Y.; Gu, J.; Tan, H.; Zhang, Y.; Huo, P. Physicochemical properties of starch adhesives enhanced by esterification modification with dodecenyl succinic anhydride. Int. J. Biol. Macromol. 2018, 112, 1257–1263. [Google Scholar] [CrossRef]
  31. Huang, Q.; Fu, X.; He, X.; Luo, F.; Yu, S.; Li, L. The effect of enzymatic pretreatments on subsequent octenyl succinic anhydride modifications of cornstarch. Food Hydrocoll. 2010, 24, 60–65. [Google Scholar] [CrossRef]
  32. Hu, H.; Liu, W.; Shi, J.; Huang, Z.; Zhang, Y.; Huang, A.; Yang, M.; Qin, X.; Shen, F. Structure and functional properties of octenyl succinic anhydride modified starch prepared by a non-conventional technology. Starch-Starke 2016, 68, 151–159. [Google Scholar] [CrossRef]
  33. Lu, X.; Luo, Z.; Fu, X.; Xiaos, Z. Two-Step Method of Enzymatic Synthesis of Starch Laurate in Ionic Liquids. J. Agric. Food Chem. 2013, 61, 9882–9891. [Google Scholar] [CrossRef]
  34. Prasertpornsakun, N.; Raita, M.; Laosiripojana, N.; Champreda, V. Biocatalytic synthesis of starch esters by immobilized lipase on magnetic microparticles. Biosci. Biotechnol. BioChem. 2015, 79, 1750–1758. [Google Scholar] [CrossRef] [Green Version]
  35. Horchani, H.; Chaâbounib, M.; Gargouria, Y.; Sayaria, A. Solvent-free lipase-catalyzed synthesis of long-chain starch esters using microwave heating: Optimization by response surface methodology. Carbohydr. Polym. 2010, 79, 466–474. [Google Scholar] [CrossRef]
  36. Luna, C.; Luna, D.; Calero, J.; Bautista, F.M.; Romero, A.A.; Posadillo, A.; Verdugo-Escamilla, C. 7—Biochemical catalytic production of biodiesel. In Handbook of Biofuels Production, 2nd ed.; Woodhead Publishing: Sawston, UK, 2016; pp. 165–199. [Google Scholar]
  37. Ilesanmi, O.I.; Adekunle, A.E.; Omolaiye, J.A.; Olorode, E.M.; Ogunkanmi, A.L. Isolation, optimization, and molecular characterization of lipase producing bacteria from contaminated soil. Sci. Afr. 2020, 8, e00279. [Google Scholar] [CrossRef]
  38. Osbon, Y.; Kumar, M. Biocatalysis and Strategies for Enzyme Improvement. In Biophysical Chemistry: Advance Applications, 1st ed.; Khalid, M.A.A., Ed.; IntechOpen: London, UK, 2020; Volume 7. [Google Scholar]
  39. Zhang, Y.; Gan, T.; Hu, H.; Huang, Z.; Huang, A.; Zhu, Y.; Feng, Y.; Yang, M. A Green Technology for the Preparation of High Fatty Acid Starch Esters: Solid-Phase Synthesis of Starch Laurate Assisted by Mechanical Activation with Stirring Ball Mill as Reactor. Ind. Eng. Chem. Res. 2014, 53, 2114–2120. [Google Scholar] [CrossRef]
  40. Huang, Z.Q.; Xie, X.L.; Chen, Y.; Lu, J.P.; Tong, J.P. Ball-milling treatment effect on physicochemical properties and features for cassava and maize starches. Comptes Rendus Chim. 2008, 11, 73–79. [Google Scholar] [CrossRef]
  41. Huang, Z.Q.; Lu, J.P.; Li, X.H.; Tong, Z.F. Effect of mechanical activation on physicochemical properties and structure of cassava starch. Carbohydr. Polym. 2007, 68, 128–135. [Google Scholar] [CrossRef]
  42. Singh, N.; Singh, J.; Kaur, L.; Sodhi, N.S.; Gill, B.S. Morphological, thermal, and rheological properties of starches from different botanical sources. Food Chem. 2003, 81, 219–231. [Google Scholar] [CrossRef]
  43. Marichelvam, M.K.; Jawaid, M.; Asim, M. Corn and Rice Starch-Based Bioplastics as Alternative Packaging Materials. Fibers 2019, 7, 32. [Google Scholar] [CrossRef] [Green Version]
  44. Ogunsona, E.; Ojogbo, E.; Mekonnen, T. Advanced material applications of starch and its derivatives. Eur. Polym. J. 2018, 108, 570–581. [Google Scholar] [CrossRef]
  45. Stojanović, Ž.; Jeremić, K.; Jovanović, S.; Lechner, M.D. A Comparison of Some Methods for the Determination of the Degree of Substitution of Carboxymethyl Starch. Starch-Starke 2005, 57, 79–83. [Google Scholar] [CrossRef]
  46. Wang, S.Y.; Zhang, C.; Liu, Q.Q.; Wang, Z.J.; Wan, K.X.; Qian, J.Y.; Zhang, L.; Wu, C.; Li, Q. Modification of potato starch by critical melting pretreatment combined with freeze-thawing: Preparation, morphology, structure, and functionality. LWT 2022, 158, 113109. [Google Scholar] [CrossRef]
  47. Ojogbo, E.; Blanchard, R.; Mekonnen, T. Hydrophobic and Melt Processable Starch-Laurate Esters: Synthesis, Structure–Property Correlations. J. Polym. Sci. Part A Polym. Chem. 2018, 56, 2611–2622. [Google Scholar] [CrossRef]
  48. Wang, Z.; Zhu, H.; Huang, J.; Ge, Z.; Guo, J.; Feng, X.; Xu, Q. Improvement of the bonding properties of cassava starch-based wood adhesives by using different types of acrylic ester. Int. J. Biol. Macromol. 2019, 126, 603–611. [Google Scholar] [CrossRef] [PubMed]
  49. Hui, R.; Qi-he, C.; Liang, F.M.; Qiong, X.; Guo-qing, H. Preparation and properties of octenyl succinic anhydride modified potato starch. Food Chem. 2009, 114, 81–86. [Google Scholar] [CrossRef]
  50. Muljana, H.; Knoop, S.V.D.; Keijzer, D.; Picchioni, F.; Janssen, L.P.B.M.; Heeres, H.J. Synthesis of fatty acid starch esters in supercritical carbon dioxide. Carbohydr. Polym. 2010, 82, 346–354. [Google Scholar] [CrossRef]
  51. Zhu, J.; Zhang, S.; Zhang, B.; Qiao, D.; Pu, H.; Liu, S.; Li, L. Structural features, and thermal property of propionylated starches with different amylose/amylopectin ratio. Int. J. Biol. Macromol. 2017, 97, 123–130. [Google Scholar] [CrossRef]
  52. Vamadevan, V.; Bertoft, E.; Seetharaman, K. On the importance of organization of glucan chains on thermal properties of starch. Carbohydr. Polym. 2013, 92, 1653–1659. [Google Scholar] [CrossRef]
  53. Silaket, P.; Chatakanonda, P.; Tran, T.; Wansuksri, R.; Piyachomkwan, K.; Sriroth, K. Thermal properties of esterified cassava starches and their maltodextrins in various water systems. Starch-Starke 2014, 66, 1022–1032. [Google Scholar] [CrossRef]
  54. Yang, B.Y.; Montgomery, R. Preparation and Physical Properties of Starch Mixed Esters. Starch-Starke 2008, 60, 146–158. [Google Scholar] [CrossRef]
  55. Zhao, K.; Li, B.; Xu, M.; Jing, L.; Gou, M.; Yu, Z.; Zheng, J.; Li, W. Microwave pretreated esterification improved the substitution degree, structural and physicochemical properties of potato starch esters. LWT 2018, 90, 116–123. [Google Scholar] [CrossRef]
  56. Sindhu, R.; Devi, A.; Khatkar, B.S. Morphology, structure and functionality of acetylated, oxidized and heat moisture treated amaranth starches. Food Hydrocoll. 2021, 118, 106800. [Google Scholar] [CrossRef]
  57. Neumann, U.; Wiege, B.; Warwel, S. Synthesis of Hydrophobic Starch Esters by Reaction of Starch with Various Carboxylic Acid Imidazolides. Starch-Starke 2002, 54, 449–453. [Google Scholar] [CrossRef]
  58. Gao, J.; Luo, Z.G.; Luo, F.X. Ionic liquids as solvents for dissolution of corn starch and homogeneous synthesis of fatty-acid starch esters without catalysts. Carbohydr. Polym. 2012, 89, 1215–1221. [Google Scholar] [CrossRef] [PubMed]
  59. Sun, Y.; Hu, Q.; Qian, J.; Li, T.; Ma, P.; Shi, D.; Dong, W.; Chen, M. Preparation and properties of thermoplastic poly(caprolactone) composites containing high amount of esterified starch without plasticizer. Carbohydr. Polym. 2016, 139, 28–34. [Google Scholar] [CrossRef] [PubMed]
  60. Li, P.; He, X.; Zuo, Y.; Li, X.; Wu, Y. Synthesis and characterization of lactic acid esterified starch by an in-situ solid phase method. Int. J. Biol. Macromol. 2020, 156, 1316–1322. [Google Scholar] [CrossRef]
  61. Zhong, C.; Xiong, Y.; Lu, H.; Luo, S.; Wu, J.; Ye, J.; Liu, C. Preparation and characterization of rice starch citrates by superheated steam: A new strategy of producing resistant starch. LWT 2022, 154, 112890. [Google Scholar] [CrossRef]
  62. Rajan, A.; Sudha, J.D.; Abraham, T.E. Enzymatic modification of cassava starch by fungal lipase. Ind. Crops Prod. 2008, 27, 50–59. [Google Scholar] [CrossRef]
  63. Rajan, A.; Prasad, V.S.; Abraham, T.E. Enzymatic esterification of starch using recovered coconut oil. Int. J. Biol. Macromol. 2006, 39, 265–272. [Google Scholar] [CrossRef]
  64. Zarski, A.; Bajer, K.; Zarska, S.; Kapusniak, J. From high oleic vegetable oils to hydrophobic starch derivatives: I. Development and structural studies. Carbohydr. Polym. 2019, 214, 124–130. [Google Scholar] [CrossRef]
  65. Deng, X.; Han, X.; Hu, X.; Zheng, S.; Liu, K. Enzyme-Catalyzed Starch Esterification in Deep Eutectic Solvent. ChemistrySelect 2019, 4, 565–569. [Google Scholar] [CrossRef]
  66. Söyler, Z.; Meier, M.A.R. Catalytic Transesterification of Starch with Plant Oils: A Sustainable and Efficient Route to Fatty Acid Starch Esters. ChemSusChem 2017, 10, 182–188. [Google Scholar] [CrossRef]
  67. Winkler, H.; Vorwerg, W.; Rihm, R. Thermal and mechanical properties of fatty acid starch esters. Carbohydr. Polym. 2014, 102, 941–949. [Google Scholar] [CrossRef] [PubMed]
  68. Winkler, H.; Vorwerg, W.; Wetzel, H. Synthesis, and properties of fatty acid starch esters. Carbohydr. Polym. 2013, 98, 208–216. [Google Scholar] [CrossRef] [PubMed]
  69. Aburto, J.; Alric, I.; Borredon, E. Organic Solvent-free Transesterification of Various Starches with Lauric Acid Methyl Ester and Triacyl Glycerides. Starch-Starke 2005, 57, 145–152. [Google Scholar] [CrossRef]
  70. Geng, F.; Chang, P.R.; Yu, J.; Ma, X. The fabrication and the properties of pretreated corn starch laurate. Carbohydr. Polym. 2010, 80, 360–365. [Google Scholar] [CrossRef]
  71. Hermawan, E.; Rosyanti, L.; Megasari, L.; Sugih, A.K.; Muljana, H. Transesterification of Sago Starch Using Various Fatty Acid Methyl Esters in Densified CO2. Int. J. Chem. Eng. Appl. 2015, 6, 152. [Google Scholar] [CrossRef] [Green Version]
  72. Aburto, J.; Hamaili, H.; Mouysset-Baziard, G.; Senocq, F.; Alric, I.; Borredon, E.; Toulouse. Free-solvent Synthesis and Properties of Higher Fatty Esters of Starch—Part 2. Starch-Starke 1999, 51, 302–307. [Google Scholar] [CrossRef]
  73. Xu, Q.; Wen, J.; Wang, Z. Preparation and Properties of Cassava Starch-based Wood Adhesives. BioResources 2016, 11, 6756–6767. [Google Scholar] [CrossRef] [Green Version]
  74. Owodunni, A.A.; Lamaming, J.; Hashim, R.; Abdulwahab Taiwo, O.F.; Hussin, M.H.; Mohamad Kassim, M.H.; Bustami, Y.; Sulaiman, O.; Mohamad Amini, M.H.; Hiziroglu, S. Properties of green particleboard manufactured from coconut fiber using a potato starch-based adhesive. BioResources 2020, 15, 2279–2292. [Google Scholar] [CrossRef]
  75. Lamaming, J.; Heng, N.B.; Owodunni, A.A.; Lamaming, S.Z.; Khadir, N.K.A.; Hashim, R.; Sulaiman, O.; Kassim, M.H.M.; Hussin, M.H.; Bustami, Y.; et al. Characterization of rubberwood particleboard made using carboxymethyl starch mixed with polyvinyl alcohol as adhesive. Compos. Part B Eng. 2020, 183, 107731. [Google Scholar] [CrossRef]
  76. Zhang, Z.; Macquarrie, D.; Clark, J.H.; Matharu, A.S. Chemical modification of starch and the application of expanded starch and its esters in hot melt adhesive. RSC Adv. 2014, 4, 41947–41955. [Google Scholar] [CrossRef]
  77. Wang, Z.; Li, Z.; Gu, Z.; Hong, Y.; Cheng, L. Preparation, characterization, and properties of starch-based wood adhesive. Carbohydr. Polym. 2012, 88, 699–706. [Google Scholar] [CrossRef]
  78. Xing, J.; Li, T.; Yu, Y.; Chen, C.; Chang, J. Development and characterization of a new bio-adhesive for wood using cassava starch and bio-oil. Int. J. Adhes. Adhes. 2018, 87, 91–97. [Google Scholar] [CrossRef]
  79. Qiao, Z.; Gu, J.; Lv, S.; Cao, J.; Tan, H.; Zhang, Y. Preparation and Properties of Normal Temperature Cured Starch-Based Wood Adhesive. Bioresources 2016, 11, 4839–4849. [Google Scholar] [CrossRef] [Green Version]
  80. Gu, Y.; Cheng, L.; Gu, Z.; Hong, Y.; Li, Z.; Li, C. Preparation, characterization and properties of starch-based adhesive for wood-based panels. Int. J. Biol. Macromol. 2019, 134, 247–254. [Google Scholar] [CrossRef]
  81. Zia-ud-Din; Xiong, H.; Wang, Z.; Chen, L.; Ullah, I. Effects of sucrose fatty acid ester addition on the structural, rheological and retrogradation behavior of high amylose starch-based wood adhesive. Int. J. Adhes. Adhes. 2019, 89, 51–58. [Google Scholar] [CrossRef]
  82. Yang, C.; Lin, Y.; Cheng, F.; Zhou, M.; Tan, L.; Zhu, P. Synthesis and Characterization of Corn Starch Phthalate by a Semidry Method. Starch-Starke 2019, 71, 1800315. [Google Scholar] [CrossRef]
  83. Desai, S.D.; Patel, J.V.; Sinha, V.K. Polyurethane adhesive system from biomaterial-based polyol for bonding wood. Int. J. Adhes. Adhes. 2003, 23, 393–399. [Google Scholar] [CrossRef]
  84. Huicochea, E.F. Introductory Chapter: Starch Modifications. Applications of Modified Starches; Huicochea, E.F., Villalobos, R.R., Eds.; InTechOpen: London, UK, 2018; p. 78. [Google Scholar]
  85. Egharevba, H.O. Chemical Properties of Starch and Its Application in the Food Industry. Chemical Properties of Starch; InTechOpen: London, UK, 2019; Available online: https://www.intechopen.com/chapters/68437 (accessed on 14 April 2022).
  86. Ganorkar, P.M.; Kulkarni, A.S. Studies on preparation and functional properties of carboxymethyl starch from sorghum. Int. Food Res. J. 2013, 20, 2205–2210. [Google Scholar]
  87. Li, S.; Mujyambere, J.M.V.; Liu, M. Synthesis of Carboxymethyl Starch with High Degree of Substitution by a Modified Dry Process. In Advanced Materials Research; Trans Tech Publications Ltd.: Bach, Switzerland, 2011; Volume 233–235, pp. 306–310. [Google Scholar]
  88. Detduangchan, N.; Sridach, W.; Wittaya, T. Enhancement of the properties of biodegradable rice starch films by using chemical crosslinking agents. Int. Food Res. J. 2014, 21, 1189–1199. [Google Scholar]
  89. Kim, M.; Lee, S.J. Characteristics of crosslinked potato starch and starch-filled linear low-density polyethylene films. Carbohydr. Polym. 2002, 50, 331–337. [Google Scholar] [CrossRef]
  90. Seligra, P.G.; Jaramillo, C.M.; Fama, L.; Goyanes, S. Biodegradable and non-retrogradable eco-films based on starch–glycerol with citric acid as crosslinking agent. Carbohydr. Polym. 2016, 138, 66–74. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Shah, N.; Mewada, R.; Mehta, T. Crosslinking of starch and its effect on viscosity behaviour. Rev. Chem. Eng. 2016, 32, 265–270. [Google Scholar] [CrossRef]
  92. Praveen, T.K.; Bhushanam, M.V.N.; Prasad, Y.R. Preparation, Evaluation and Characterization of Starch and Grafted Starch. IIJPSR 2021, 12, 4001–4010. [Google Scholar]
  93. Zia-ud-Din; Chen, L.; Ullah, I.; Wang, P.K.; Javaid, A.B.; Hu, C.; Zhang, M.; Ahamd, I.; Xiong, H.; Wang, Z. Synthesis, and characterization of starch-g-poly (vinyl acetate-co-butyl acrylate) bio-based adhesive for wood application. Int. J. Biol. Macromol. 2018, 114, 1186–1193. [Google Scholar] [CrossRef] [PubMed]
  94. Kramer, M.E. Structure and function of starch-based edible films and coatings. In Edible Films and Coatings for Food Applications; Springer: New York, NY, USA, 2009; pp. 113–134. [Google Scholar]
  95. Zhang, Y.; Ding, L.; Gu, J.; Tan, H.; Zhu, L. Preparation, and properties of a starch-based wood adhesive with high bonding strength and water resistance. Carbohydr. Polym. 2015, 115, 32–37. [Google Scholar] [CrossRef]
  96. Li, Z.; Wang, J.; Li, C.; Gu, Z.; Cheng, L.; Hong, Y. Effects of montmorillonite addition on the performance of starch-based wood adhesive. Carbohydr. Polym. 2015, 115, 394–400. [Google Scholar] [CrossRef]
  97. Chen, L.; Xiong, Z.; Xiong, H.; Wang, Z.; Din, Z.U.; Nawaz, A.; Wang, P.; Hu, C. Effects of nano-TiO2 on bonding performance, structure stability and film-forming properties of starch-g-VAc based wood adhesive. Carbohydr. Polym. 2018, 200, 477–486. [Google Scholar] [CrossRef]
  98. Zhao, X.F.; Peng, L.Q.; Wang, H.L.; Wang, Y.B.; Zhang, H. Environment-friendly urea-oxidized starch adhesive with zero formaldehyde-emission. Carbohydr. Polym. 2018, 181, 1112–1118. [Google Scholar] [CrossRef]
  99. Zhang, Y.; Chen, M.; Zhang, J.; Li, J.; Shi, S.Q.; Gao, Q. A High-Performance Bio-Adhesive Using Hyperbranched Aminated Soybean Polysaccharide and Bio-Based Epoxide. Adv. Mater. Interfaces 2020, 7, 2000148. [Google Scholar] [CrossRef] [Green Version]
  100. Yin, H.; Zheng, P.; Zhang, E.; Rao, J.; Lin, Q.; Fan, M.; Zhu, Z.; Zeng, Q.; Chen, N. Improved wet shear strength in eco-friendly starch-cellulosic adhesives for woody composites. Carbohydr. Polym. 2020, 250, 116884. [Google Scholar] [CrossRef]
  101. Wu, Q.; Shao, W.; Xia, N.; Wang, P.; Kong, F. A separable paper adhesive based on the starch―lignin composite. Carbohydr. Polym. 2020, 229, 115488. [Google Scholar] [CrossRef] [PubMed]
  102. Li, D.; Zhuang, B.; Wang, X.; Wu, Z.; Wei, W.; Aladejana, J.Y.; Hou, X.; Yves, K.G.; Xie, Y.; Liu, J. Chitosan used as a specific coupling agent to modify starch in preparation of adhesive film. J. Clean. Prod. 2020, 277, 123210. [Google Scholar] [CrossRef]
  103. Moreno, O.; Cárdenas, J.; Atarés, L.; Chiralt, A. Influence of starch oxidation on the functionality of starch-gelatin based active films. Carbohydr. Polym. 2017, 178, 147–158. [Google Scholar] [CrossRef] [PubMed]
  104. Woggum, T.; Sirivongpaisal, P.; Wittaya, T. Properties and characteristics of dual-modified rice starch based biodegradable films. Int. J. Biol. Macromol. 2014, 67, 490–502. [Google Scholar] [CrossRef] [PubMed]
  105. Akmana, F.; Kazachenko, A.S.; Vasilyeva, N.Y.; Malyar, Y.N. Synthesis and characterization of starch sulfates obtained by the sulfamic acid-urea complex. J. Mol. Struct. 2020, 1208, 127899. [Google Scholar] [CrossRef]
  106. Kazachenko, A.S.; Vasilyeva, N.Y.; Malyar, Y.N.; Kazachenko, A.S.; Slyusareva, E.A. Synthesis of sulfated starch-casein complex. IOP Conf. Ser. Mater. Sci. Eng. 2020, 862, 062013. [Google Scholar] [CrossRef]
  107. Lubis, M.A.R.; Park, B.D.; Hong, M.K. Tuning of Adhesion and Disintegration of Oxidized Starch Adhesives for the Recycling of Medium Density Fiberboard. BioResources 2020, 15, 5156–5178. [Google Scholar] [CrossRef]
  108. Su, M.; Wu, J.; Pan, P.; Wang, H. Preparation, and characterization of a water-resistant polyamide-oxidized starch-methyl methacrylate eco-friendly wood adhesive. Int. J. Biol. Macromol. 2022, 194, 763–769. [Google Scholar] [CrossRef]
  109. Sulaiman, N.S.; Hahim, R.; Amini, M.H.M.; Sulaiman, O.; Hiziroglu, S. Evaluation of the Properties of Particleboard Made Using Oil Palm Starch Modified with Epichlorohydrin. BioResources 2013, 8, 283–301. [Google Scholar] [CrossRef]
  110. Chen, X.; Sun, C.; Wang, Q.; Tan, H.; Zhang, Y. Preparation of glycidyl methacrylate grafted starch adhesive to apply in high-performance and environment-friendly plywood. Int. J. Biol. Macromol. 2022, 194, 954–961. [Google Scholar] [CrossRef]
  111. Chen, L.; Xiong, Z.; Li, Q.; Din, Z.U.; Xiong, H. Sodium dodecyl sulfate improves the properties of bio-based wood adhesive derived from micronized starch: Microstructure and rheological behaviors. Int. J. Biol. Macromol. 2019, 140, 1026–1036. [Google Scholar] [CrossRef] [PubMed]
  112. Jiang, Y.; Chen, Q.; Tan, H.; Gu, J.; Zhang, Y. A Low-Cost, Formaldehyde-Free, and High-Performance Starch-based Wood Adhesive. BioResources 2019, 14, 1405–1418. [Google Scholar]
  113. Nasiri, A.; Wearing, J.; Dubé, M.A. Using Lignin to Modify Starch-Based Adhesive Performance. ChemEngineering 2020, 4, 3. [Google Scholar] [CrossRef] [Green Version]
  114. Gadhave, R.V.; Mahanwar, P.; Gadekar, P. Starch-Based Adhesives for Wood/Wood Composite Bonding: Review. OJP Chem. 2017, 7, 19–32. [Google Scholar] [CrossRef] [Green Version]
  115. Eisen, A.; Bussa, M.; Röder, H. A review of environmental assessments of biobased against petrochemical adhesives. J. Clean. Prod. 2020, 277, 124277. [Google Scholar] [CrossRef]
  116. Magalhães, S.; Alves, L.; Medronho, B.; Fonseca, A.C.; Romano, A.; Coelho, J.F.J.; Norgren, M. Brief Overview on Bio-Based Adhesives and Sealants. Polymers 2019, 11, 1685. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Chemical reactions for production of esterified starch.
Figure 1. Chemical reactions for production of esterified starch.
Polymers 14 02023 g001
Figure 2. Transesterification of starch.
Figure 2. Transesterification of starch.
Polymers 14 02023 g002
Figure 3. (a) Selected examples of starch modification processes being used for adhesive applications and (b) examples of the structure of modified starches. Selected examples of starch modification reactions, with given details in Table 4.
Figure 3. (a) Selected examples of starch modification processes being used for adhesive applications and (b) examples of the structure of modified starches. Selected examples of starch modification reactions, with given details in Table 4.
Polymers 14 02023 g003
Table 1. Properties of native starch [43,44].
Table 1. Properties of native starch [43,44].
Type of StarchRatio of Amylose to AmylopectinGelatinization Temperature (°C) a% Solubility% Crystallinity b
Rice30:7068–7711–18, 95 °C38
Potato18:8258–6882, 95 °C23–53
Cassava18:8260–80-31–59
Wheat20:8058–641.55, 100 °C43–48
Corn28:7262–7222, 95 °C22–28
Differential scanning calorimetry (a) and X-ray diffraction (b) were utilized to obtain the gelatinization temperature and %Crystallinity, respectively.
Table 2. Reactions affording starch esters and their degree of substitution.
Table 2. Reactions affording starch esters and their degree of substitution.
Modification of StarchType of StarchDegree of Substitution (DS)Ref.
ReactionReagent–Catalyst
AcylationCarboxylic acid (butanoic, hexanoic, octanoic, palmitic acid)Potato2.52–3.00[54]
AcetylationAcetic anhydride-iodineCorn0.12–2.97[17]
Acetic anhydride-tartaric acidCorn0.06–1.54[21]
Acetic anhydride and glacial acetic acidPotato0.015–0.054[55]
Vinyl acetateAmaranth grain0.22[56]
EsterificationCarboxylic acid imidazolides (C8, C12, C16)—methanolic KOCH3Potato1.76[57]
Lauric acid—K2CO3Cassava0.0148–0.0412[39]
Lauroyl chlorideCorn0.45–2.92[47]
Lauric acid, palmitic acid, and stearic acidCorn0.053–0.100[58]
Stearyl chlorideCorn0.25–1.58[59]
Lactic acid—stannous octoateCorn0.015–0.12[60]
Citric acidRice0.015–0.064[61]
Enzymatic
esterification
Oleic acid—immobilized lipaseMaize2.86[35]
Hydrolyzed recovered coconut oil (lauric acid) and palmitic acid—fungal lipaseCassava1.1 and 1.04[62]
Lauric acid-lipase in ionic liquidMaize0.048–0.171[33]
Carboxylic acid (acetic, lauric, and stearic acid)-lipaseMaize0.016–0.513[15]
Recovered coconut oil-lipaseMaize1.55[63]
Hydrolyzed rapeseed oil—immobilized fungal lipasePotato0.15–1.36[64]
Palmitic, lauric, decanoic acids—lipase in deep eutectic solventNative0.07–0.19[65]
Transesterification
Transesterification
Olive oil or high oleic sunflower oil-TBDMaize1.29–1.33[66]
Fatty acid vinyl esterMaize1.40–1.73 and
2.20–2.63
[67]
Vinyl laurate—Cs2CO3Maize1.75–2.44[68]
Methyl laurate—potassium laurateMaize0.08–0.62[69]
Vinyl laurate and vinyl stearate—Na2HPO4, K2CO3 and Na acetateCorn0.24–2.96[7]
Methyl laurateCorn0.2673–0.7034[70]
Fatty acid methyl esterSago0.45[71]
Table 3. Esterified starch derived adhesives. The properties are reported relatively to those of corresponding native starches.
Table 3. Esterified starch derived adhesives. The properties are reported relatively to those of corresponding native starches.
StarchModificationPropertiesUtilizationRef.
CassavaEsterification using dodecenyl succinic anhydride (DDSA) as reactant and polymethylene polyphenyl polyisocyanate (PAPI)Improved bonding strength and water resistance of esterified starch adhesiveAdhesive film for plywood[30]
CassavaGrafted with olein monomer (vinyl acetate; VAc and butyl acrylate; BA as co-monomer)Enhanced the storage stability of starch wood adhesive and glass transition temperatureWood adhesive[73]
CassavaGrafted with different typrs of acrylic esterImproved performance and shear strength under both dry and wet conditionsWood adhesive[48]
CassavaEsterification with acid in bio-oil and grafted copolymerization with vinyl acetate and butyl acrylateLower viscosities, improved shelf life and mildew resistance of bio-adhesiveWood adhesive[78]
High amylose corn starch (HACS)Esterification using propionic anhydride as agent and catalytic 4-dimethyl aminopyridine (DMAP)Surface area of HACS was increased with improved stability, hydrophobicity, and adhesion strengthHot melt adhesive bonded Al plate[76]
CornEsterification using maleic anhydrideImproved shear strength and thermal stabilityWood adhesive[79]
CornGrafted with vinyl acetate and crosslinked with N-methylol acrylamideImproved water resistance and can be used in hot pressimg processesWood-based panel adhesive[80]
CornAddition of sucrose fatty acid esterEnhanced shear-thinning, solid-like behavior, and anti-retrodegradation of starchWood adhesive[81]
CornEsterification with phthalic anhydride in DMFReduced viscosity and thermal decompositionAdhesive[82]
Waxy cornGrafted with vinyl acetateIncreased shear strength under both dry and wet conditionsWood adhesive[77]
PotatoTransesterification with natural oil and combined with toluene 2,4-diisocyanateGood resistance to cold and hot water, moderate resistance to acid and weak resistance to akaliPolyurethane (PU) adhesive for wood[83]
Table 4. The preparation of modified starch and adhesive application from different starch.
Table 4. The preparation of modified starch and adhesive application from different starch.
Reaction of Modified StarchStarchPropertiesUtilizationRef.
Oxidation and modification with chitosanCornImproved dry and wet shear strength of plywoodAn adhesive film for plywood[102]
Oxidation using H2O2 and then crosslinking with B-pMDI and citric acidCornImproved physical properties, mechanical properties, and water resistanceMedium density fiberboard[107]
Oxidation using KMnO4, then crosslinking and copolymerization with polyamide and methyl methacrylateCornImproved wet shear strength and water resistanceAn adhesive for plywood[108]
Oxidation using KMnO4, polycondensation reaction with urea and addition of nano-TiO2CornThe nano-TiO2 effectively improves dry shear strength and viscosity of the nano-TiO2-U-OSt adhesive.An adhesive[98]
Oxidation using H2O2 and crosslinking with polyamidoamine-epichlorohydrin (PAH)RiceEnhanced thermal stability, hydrophobicity, wet-cohesion, and adhesivenessAn adhesive for wood composites[100]
Etherification with carboxymethyl and use of POCl3 as crosslinking agentWheatThe modified starch mixed with PVA improves solid content, heat and water resistance but decrease viscosity.Adhesive for particleboard[75]
Etherification with epichlorohydrinOil palmImproved mechanical strength (modulus, elasticity, and internal bond), solid content and viscosityAdhesive for particleboard[109]
Graft copolymerization with glycidyl methacrylate (GMA) and crosslinking with sodium trimetaphosphate (STMP).CassavaImproved water resistance and bonding strengthAn adhesive for plywood[110]
Graft copolymerization with sodium dodecyl sulfate (SDS)Micronized (MS)Improved shear strength and decreased viscosity of micronized starch with increasing SDS contentsWood adhesive[111]
Graft copolymerization with ligninCornImproved adhesive bond strength and moisture resistance, including extended shelf-life.An adhesive for paper[101]
Crosslinking with polyphenylene isocyanate (PAPI) with poly vinyl alcohol (PVOH) as a protective colloidCassavaImproved water resistance, shear strength, mobility and storage stability of starch adhesiveWood adhesive[112]
Crosslinking with ligninCornLignin improved the strength and water resistance of adhesiveAn adhesive for cardboard application[113]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Watcharakitti, J.; Win, E.E.; Nimnuan, J.; Smith, S.M. Modified Starch-Based Adhesives: A Review. Polymers 2022, 14, 2023. https://doi.org/10.3390/polym14102023

AMA Style

Watcharakitti J, Win EE, Nimnuan J, Smith SM. Modified Starch-Based Adhesives: A Review. Polymers. 2022; 14(10):2023. https://doi.org/10.3390/polym14102023

Chicago/Turabian Style

Watcharakitti, Jidapa, Ei Ei Win, Jaturavit Nimnuan, and Siwaporn Meejoo Smith. 2022. "Modified Starch-Based Adhesives: A Review" Polymers 14, no. 10: 2023. https://doi.org/10.3390/polym14102023

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

Article Metrics

Back to TopTop