Next Article in Journal
Chemical Mechanisms Underlying Sweetness Enhancement During Processing of Rehmanniae Radix: Carbohydrate Hydrolysis, Degradation of Bitter Compounds, and Interaction with Taste Receptors
Previous Article in Journal
The Impact of Yeast Strains and Oenological Procedures on the Chemical Composition, Antioxidant Potential, and Aromatic Profile of Blueberry Wines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Sodium Alginate Modifications: A Critical Review of Current Strategies and Emerging Applications

1
School of Food and Biological Engineering, Jiangsu University, Zhenjiang 212013, China
2
Department of Agricultural and Biosystems Engineering, Faculty of Agriculture, Benha University, Moshtohor, Qaluobia P.O. Box 13736, Egypt
3
Institute of Food Physical Processing, Jiangsu University, Zhenjiang 212013, China
*
Author to whom correspondence should be addressed.
The authors contributed equally to this work.
Foods 2025, 14(22), 3931; https://doi.org/10.3390/foods14223931
Submission received: 3 October 2025 / Revised: 12 November 2025 / Accepted: 14 November 2025 / Published: 17 November 2025
(This article belongs to the Section Food Physics and (Bio)Chemistry)

Abstract

Sodium alginate, a natural anionic polysaccharide, exhibits broad potential applications in food, biomedicine, and environmental engineering due to its favorable biocompatibility, degradability, and functional tunability. This review systematically summarizes its chemical structure, physicochemical characteristics, sources, and extraction methods. It also focused on modification strategies, including chemical approaches (e.g., esterification, oxidation, sulfation, graft copolymerization), physical methods (composite modification, irradiation cross-linking, ultrasound treatment), and biological (e.g., enzyme regulation), and elucidated their underlying mechanisms. In the context of food science, special emphasis is placed on food-compatible chemistries and mild modification routes (such as phenolic crosslinking, enzyme-assisted coupling, and other green reactions) that enable the development of edible films, coatings, and functional carriers, while distinguishing these from non-food-oriented chemical strategies. The review further highlights novel applications of modified sodium alginate in areas including food packaging, functional delivery systems, drug release, tissue engineering, and environmental remediation (heavy metal and dye removal). Overall, this work provides a comprehensive perspective linking modification pathways to food-relevant applications and clarifies how chemical tailoring of alginate contributes to the design of safe, sustainable, and high-performance bio-based materials.

1. Introduction

Sodium alginate (SA) is a linear copolymer composed of β-D-mannuronic acid (M) and α-L-guluronic acid (G) units [1]. Its unique egg-box gel structure and M/G ratio-dependent properties make it a popular study topic in natural polymer materials. With the growing emphasis on green chemistry and sustainable development, SA has received much attention in food engineering, biomedicine, and environmental remediation due to its wide availability, high biocompatibility, and strong modifiability [2]. Furthermore, SA is officially recognized as a safe and approved additive by major regulatory authorities. In the European Union, it is listed as food additive E401 under Regulation (EC) No 1333/2008 and specified in Commission Regulation (EU) No 231/2012 [3,4]. In the United States, sodium alginate is classified as a “Generally Recognized as Safe” (GRAS) substance under 21 CFR § 184.1724 [5] by the Food and Drug Administration (FDA), where it is permitted for use as a stabilizer, thickener, and emulsifier. These regulatory approvals further validate its broad and safe applications in food, pharmaceutical, and biomedical fields. It should be noted that the chemical tailoring discussed herein is directed to the design of edible and food-contact materials, rather than the authorization of new food additives per se. This clarification ensures that the review focuses on food-safe modification routes and materials relevant to food science, rather than legislative approval processes.
Despite its advantages, SA often exhibits inherent limitations in mechanical strength, environmental stability, bioactivity, and controlled release [6]. To overcome these shortcomings, three major modification approaches—chemical, physical, and enzymatic—have been developed. Previous reviews have typically addressed only individual modification pathways or specific properties, lacking a comprehensive and application-oriented perspective. This review, therefore, integrates current progress across all three modification domains, examining molecular mechanisms, process routes, and resulting functionalities. Particular emphasis is placed on food-compatible and green chemistries (e.g., phenolic crosslinking and enzyme-assisted coupling) that support the development of edible films, smart/active packaging, 3D-printed constructs, and functional food carriers, while stronger synthetic chemistries are discussed as forward-looking strategies for non-food fields.
This synthesis establishes a coherent framework for understanding structure–property relationships and guides engineering of multifunctional, sustainable SA-based materials for food systems (see Figure 1).

2. Preparation of Sodium Alginate

Alginate is a naturally abundant polysaccharide found in brown algae and plays a key structural role. The primary production sources include Macrocystis pyrifera (giant kelp), Laminaria digitata (oarweed), Laminaria japonica (kombu), Ascophyllum nodosum (knotted wrack), and other macroalgae [7]. The traditional extraction process [8] of SA from brown algae generally involves several key steps, including acid pretreatment [9], alkaline extraction [10], filtration, and drying [11], as illustrated in Figure 2. In this process, hydrochloric acid pretreatment helps remove impurities and enhance viscosity, followed by sodium carbonate extraction and purification to obtain high-purity SA. Recently, ultrasound-assisted extraction [12] has been introduced as an efficient alternative method to improve extraction efficiency, shorten processing time, and maintain the structural integrity of polysaccharides. Youssouf et al. [13] developed an ultrasound-assisted extraction (UAE) [14,15,16] procedure for SA preparation from brown algae (Sargassum binderi and Turbinaria ornata), while simultaneously enabling the extraction of carrageenan from red algae (Kappaphycus alvarezii and Euchema denticulatum). The UAE technique can effectively maintain the conformational integrity of target polysaccharides and significantly improves the extraction efficiency, providing a new technological pathway for algal polysaccharides production. Recently, Ummat et al. [17] used ultrasound-assisted sodium bicarbonate [18,19] to extract alginate from solid-phase by-products obtained after conventional extraction of fucoidan from brown algae [20], resulting in improved extraction yields.

3. Physicochemical Properties of Sodium Alginate

3.1. Chemical Structure and Molecular Weight

SA is a natural linear copolymer obtained from seaweed cell walls, consisting of (1→4)-linked β-D-mannuronic acid (M unit with a 4C1 ring conformation) and α-L-guluronic acid (G unit with a 1C4 ring conformation) residues [21]. Its structure is characterized by homopolymeric blocks (continuous M or G sequences) and heteropolymeric blocks (alternating M and G units), as shown in Figure 3A [22]. More specifically, SA exhibits an irregular block architecture with varying proportions of MG, GM, GG, and MM blocks, as illustrated in Figure 3B [23,24]. SA from various sources varies in the M/G ratio, the length of the polymer chains, and the degree of blockiness or randomness in the distribution of M and G residues, which predominantly determine its structural characteristics as well as physicochemical properties. The proportion and distribution of GulA and ManA units vary with seaweed species. The relative abundance and distribution of GulA and ManA units, as well as their chemical modification, can significantly influence the physicochemical properties of SA. SA enriched with GulA units shows a rigid molecular structure, whereas SA rich in ManA units possesses a flexible structure [25]. The ratio of MM, GG and MG blocks determines the physical properties of SA [8]. For example, SA with a high G content has a higher gelling capacity, whereas SA with a high M level exhibits a higher viscosity [26]. The assessment of the M/G ratio is also important since a high M/G ratio causes SA to form an elastic gel, whereas a low M/G ratio produces brittle gels [27,28]. Only the G-block of SA is thought to contribute to the formation of hydrogels by intermolecular cross-linking with divalent cations (e.g., Ca2+) [29].
The molecular weight of SA varies considerably, as it is not genetically encoded but enzymatically synthesized by alginate polymerases and can undergo partial depolymerization during extraction and purification [27]. Consequently, the molecular weight of SA is conventionally expressed as an average value, typically represented by the number-average (Mn) and weight-average (Mw) molecular weights [28]. Reported Mw values span a broad range—approximately 60,000 to 700,000 g/mol for commercial alginates [30], and in some cases reaching up to ~986,000 g/mol depending on algal species and extraction conditions [31]. Commercial-grade materials from suppliers such as Sigma-Aldrich (Product PHR1471) are generally specified within the 12,000–40,000 g/mol range, reflecting controlled depolymerization used to standardize viscosity for analytical and industrial purposes [32]. The molecular weight distribution of SA is influenced by several factors, including biological source, extraction chemistry (acidic vs. alkaline), temperature, oxidative conditions, and mechanical shear during processing [33]. The polydispersity index (PDI = Mw/Mn), which characterizes the breadth of chain-length distribution, typically falls between 1.5 and 3.0, though values up to 6 have been reported for heterogeneous or partially degraded alginate preparations [34]. Collectively, these findings indicate that the molecular weight of SA should be regarded as a variable distribution dependent on biological origin and processing history, rather than a single fixed value.

3.2. Solubility

The solubility of SA in aqueous media is influenced by several factors, including solution pH, the presence of co-solvents, ionic strength, gel-promoting ions, and the intrinsic structure of the biopolymer. Dissolution of SA requires a pH above a critical threshold necessary for the deprotonation of its carboxyl groups [35]. The solubility of SA is fundamentally determined by its molecular structure and the ionization state of the carboxyl group in the polymer backbone. When the carboxyl groups are protonated to form alginic acid, the polymer exhibits minimal solubility and becomes insoluble in most solvents, including water. In contrast, salt forms of SA are water-soluble but remain insoluble in organic solvents, such as alcohol, hydroalcoholic solutions containing at least 30% alcohol, chloroform, and ether. An exception is the tetrabutylammonium (TBA) salt of SA, which is soluble in water, ethylene glycol, and polar aprotic solvents containing tetrabutylammonium fluoride (TBAF) [36]. Under acidic conditions, the solubility of SA is significantly reduced due to limited hydration capacity [22]. SA dissolves slowly in cold water, forming viscous solutions. Furthermore, the ability of alginate to form hydrogels via cross-linking with divalent ions such as Ca2+, Ba2+, and Zn2+ considerably influences its solubility behavior [27].

3.3. Gel Formation Ability

Gelation of SA can be induced by cations such as H+, Ca2+, Cu2+, Ba2+ [37], primarily through electrostatic interactions between negatively charged carboxyl groups on the polymer chain and positively charged cations, leading to the formation of ionic cross-links and resultant polyelectrolyte complexes [38]. Ca2+-induced gelation represents one of the most important functional properties of SA [39]. This process is facilitated by ion exchange between sodium ions associated with G units of SA and calcium ions, which promotes the aggregation of G-blocks into a characteristic egg-box gel structure (Figure 4) [40]. There are two key methods of forming calcium ionic crosslinked hydrogels: internal gelation and external gelation. Internal gelation involves adding Ca2+ directly to the polymer dispersion and stirring for a specific duration to activate the ionic cross-linking reaction [41]. The properties of gel formed by internal gelation usually depend on the calcium salt used. For example, calcium chloride cross-links very rapidly with SA, forming a heterogeneous gel structure [42]. External gelation of Ca2+ usually involves immersing the pregel into a Ca2+ solution, where the Ca2+ diffuses inward to form the gel and replaces the Na+ in the SA structure [43]. External gelation enables rapid gel formation (within seconds) without disrupting the continuity of the pre-gel structure, and the crosslink density of the final gel network correlates with the Ca2+ concentration [44]. Alginates also form gels with a variety of other mono-, di-, or trivalent cations. The gelation mechanisms [45] for these cations often differ from Ca2+-induced gelation, resulting in gels with diverse structural characteristics [46] and potential applications [47].

3.4. Biocompatibility

Alginate is a biocompatible and biodegradable polymer whose properties vary significantly based on its G/M ratio and overall chemical composition [48]. The immunogenicity of SA is influenced by its structural features and the presence of specific functional groups. For example, alginates rich in M units are more immunogenic and effective in triggering the generation of cytokines compared to those with a high G content [49]. Therefore, structural modifications represent a viable strategy to modulate its immunogenic potential [50]. SA hydrogel [51] is widely used in the biomedical field due to its inherent biocompatibility and non-toxicity. For instance, a novel hydrogel loaded with Ag-Metal–organic frameworks has been developed to enhance its antibacterial efficacy [52]. SA-based hydrogels demonstrate adequate biological safety profiles, making them suitable for wound healing applications [53]. Moreover, these hydrogels can be tailored to release drugs in a controlled manner, making them ideal for controlled drug delivery systems [27].

4. Modification Methods of Sodium Alginate

In recent years, research on SA modification has shifted from single property optimization toward multifunctional synergistic design. To meet diverse material requirements in different fields, researchers have developed proposed multidimensional modification strategies, typically classified into three major approaches: chemical modification, physical modification, and biological modification.

4.1. Chemical Modification

SA is widely used due to its superior biocompatibility and processability. Researchers have developed various chemical modification methods, including esterification, oxidation, sulfation, Ugi reaction, aldehyde cross-linking, phosphorylation, amidation, and grafting, to enhance its functional attributes. These methods precisely control the molecular structure of SA, imparting hydrophobicity, enhancing adsorption/binding capacity, introducing bioactivity, optimizing drug delivery profiles, improving stability, regulating dissolution/mineralization properties, enhancing processability, and conferring smart responsive characteristics to the material. As a result, the potential of SA in the design and application of functional materials has been significantly expanded. To align the discussion with food science, we classify chemical modifications into two practical categories: (i) food-compatible or green chemistries potentially suitable for edible/food-contact uses (e.g., phenolic crosslinking with tannic/cinnamic acids, genipin crosslinking, enzyme-assisted coupling), and (ii) non-food chemistries (e.g., strong sulfation or periodate oxidation, multi-component Ugi reactions, certain vinyl-graft routes) that are mainly forward-looking material strategies. Throughout this section and the Applications part (Section 5), we explicitly indicate when a given chemistry is relevant to food systems versus non-food domains.

4.1.1. Esterification

Esterification of SA typically involves an acid-catalyzed dehydration and condensation reaction between carboxyl and hydroxyl groups to form ester bonds (Figure 5). This reaction also provides a straightforward method for attaching alkyl groups to the main chain of SA [54]. Esterification procedures can be categorized into two methods: (i) surface modification of alginate products. For instance, esterification of maleic anhydride on the Ca2+–alginate hydrogel beads has been shown to improve their oil adsorption capacity [23]; (ii) molecular modification prior to product preparation. Esterification can improve the mechanical properties of SA by introducing ester groups into its polymer chains. This modification enhances both rigidity and tensile strength, making it more suitable for applications requiring higher mechanical stability, such as hydrogels [55]. Specifically, maleic anhydride esterification significantly increases the hydrophobicity and oil adsorption capacity of SA, which forms a hydrogel that can efficiently capture oil droplets in water, suitable for oil cleanup and wastewater treatment [56]. Similarly, thioacetic acid esterification considerably improves the mechanical properties and bioadhesion of SA, showing promising potential for biomedical applications such as drug delivery and wound healing [57].
Food relevance: Certain esterification reactions of sodium alginate can be food-compatible when employing food-grade or naturally derived acids (e.g., fatty, phenolic, or cinnamic acids). These mild routes are useful for tuning hydrophobicity and barrier properties in edible films and food-contact materials, whereas strong acid-catalyzed reactions remain non-food.

4.1.2. Oxidization

SA can be oxidized by sodium periodate to introduce aldehyde groups (Figure 6). This reaction selectively targets the C-2 and C-3 positions of the hyaluronic acid units, converting them into aldehyde groups. This modification enhances the binding affinity of alginate toward heavy metal ions [58]. Furthermore, the process of oxidation significantly alters the molecular structure and properties of alginate, including its aldehyde content and molecular weight [23]. The gravimetric molar mass of alginate decreases rapidly with increasing degree of oxidation, even at low levels (e.g., 5 mol%) [58]. Notably, when the molar ratio of sodium periodate to alginate is held constant, the amount of aldehyde groups produced on the oxidized polymer backbone increases with increasing alginate concentration. Concurrently, the molecular weight decreases as alginate concentration increases, probably due to enhanced chain breakage. This occurs because higher molecular collision frequencies in concentrated solutions accelerate oxidation of adjacent hyaluronic acids within chains [59]. To minimize the possibility of side reactions, the oxidation procedure should be performed in complete darkness.
Food relevance: Mild oxidation using controlled periodate or enzymatic routes can be applied for structural modulation in edible coatings and hydrogels, but high-degree oxidation is mainly non-food and used for biomedical/environmental purposes.

4.1.3. Sulfation

Sulfation is a prominent modification strategy for polysaccharides in biomedical research due to its broad range of bioactivity, including anticoagulant, anticancer, antimicrobial, antiviral, and immunomodulatory effects [60]. This chemical process involves introducing sulfate groups onto the hydroxyl residues of alginate, resulting in the formation of sulfated alginate derivatives (Figure 7). Common methods involve reacting alginate with chlorosulfonic acid or 1,3-benzene sulfonyl chloride in formamide [23]. Among various sulfating agents used for sulfation of alginate (e.g., sulfuric acid, chlorosulfonic acid, sulfuryl chloride, sulfamic acid, and sulfur trioxide), chlorosulfonic acid is particularly suitable for alginate functionalization due to its reaction controllability and reproducibility [61]. The introduction of sulfate groups induces conformational changes in the three-dimensional structure of the polysaccharide chain, thereby enabling these polysaccharides with weak or no bioactivity to enhance or acquire some sulfate group-related activities [62]. Sulfated polysaccharides have demonstrated a wide range of biological activities, such as anticoagulant, anti-inflammatory, antiviral, and immunomodulatory properties, making them beneficial for biomedicine, functional food, and biomaterial applications [63].
Food relevance: Sulfation commonly involves non-food reagents (e.g., chlorosulfonic acid) and thus remains non-food; its relevance is confined to biomedical studies rather than direct food-contact applications.

4.1.4. Ugi Reaction

The Ugi reaction is a four-component coupling process involving a carboxylic acid, a carbonyl compound, an amine, and an isocyanide. This reaction allows the efficient synthesis of amino acid and peptide derivative libraries [64]. All reaction components are in a dynamic equilibrium through several intermediates, until an irreversible intramolecular 1,4-O→N acyl transfer yields an N-acylamino acid amide [65] (Figure 8). Amphiphilic alginate derivatives with tunable organic solubility and thermal properties have been successfully synthesized via the Ugi multicomponent reaction [66]. This approach incorporates hydrophobic components, such as octylamine and oleylamine, into the SA backbone, significantly enhancing its hydrophobicity and improving its self-assembly behavior [67]. The resultant amphiphilic SA derivatives demonstrate enhanced ability to encapsulate hydrophobic drugs and exhibit improved drug release profiles, highlighting their potential for hydrophobic drug delivery systems [68].
Food relevance: Because the Ugi reaction uses isocyanides and organic solvents, it is not food-compatible and is considered a non-food modification, summarized here only for its structural design value.

4.1.5. Aldehyde Cross-Linking

Aldehyde cross-linking forms stable covalent linkages between aldehyde and amino groups, thereby enhancing the structural stability of biomolecules [69] (Figure 9). In collagen, lysyl oxidase generates crucial intermolecular cross-links through telopeptide lysine aldehydes and telopeptide hydroxylysine aldehydes [70,71]. However, in hydrogels, aldehyde cross-linking occurs via both covalent and Schiff-base pathways. Covalent cross-linking involves the formation of stable bonds between aldehyde and amino groups, whereas Schiff base cross-linking induces the formation of reversible linkages [72]. Aldehyde cross-linking offers an alternative strategy to reinforce the stability and functionality of alginate-based materials [73]. For instance, dialdehyde starch has been employed to crosslink gelatin and SA hydrogels, enabling precise modulation of their mechanical properties. This approach significantly improves the stability and functionality of the hydrogels [74].
Food relevance: Naturally derived aldehyde donors (e.g., genipin or reducing sugars) enable safe crosslinking for food-contact or edible materials, whereas strong aldehydes like glutaraldehyde are non-food and restricted to biomedical/industrial uses.

4.1.6. Phosphorylation

Phosphorylated alginates are commonly used to promote nucleation and growth of hydroxyapatite (HAP). Phosphorylation is typically carried out using a mixture of urea and phosphoric acid (Figure 10) [35]. Coleman et al. [75] successfully functionalized alginate hydroxyl groups through heterogeneous urea/phosphate reactions. Multidimensional NMR studies indicate that phosphorylation preferentially targets the C3 equatorial carbon of mannuronate (M) residues, although the reactivity of guluronate (G) units remains unclear [66]. The modification can considerably alter the physicochemical properties of alginate, including its solubility, viscosity, and gelation behavior, while also enhancing its bioavailability [76] via improved biomolecular interactions [30]. Wang et al. [77] established a straightforward method for fabricating macroporous alginate hydrogels, which involves dissolving phosphorylated SA in water and using Ca2+ released from acid-induced CaCO3 dissolution to trigger gelation, during which CO2 bubble formation creates pores. The resulting hydrogels display dynamic and reversible viscoelasticity, making them highly suitable for applications requiring tunable mechanical performance [78].
Food relevance: Due to the use of non-food reagents and migration concerns, phosphorylated alginates have limited direct food applicability and are primarily used in biomedical and nutrient-delivery models.

4.1.7. Amidation

Amidation is a reaction that involves the formation of amide bonds between carboxyl groups and amine groups, facilitating the incorporation of hydrophobic tails into alginate and other polysaccharides (Figure 11). This modification enhances their ability to self-assemble into micelles and vesicles in aqueous environments [79]. Using 4-(4,6-dimethoxy-1,3,5-triazin-2-yl)-4-methylmorpholinium chloride (DMTMM) as a condensing agent, alginate can be efficiently functionalized with various amines such as furanamine, adipic acid dihydrazide, amino acids, and peptides. The reaction proceeds through a pH-dependent mechanism, with optimal efficiency observed at pH 5–6, leading to amide bond formation between the carboxyl groups of alginate and the amino groups of the modifiers [80]. DMTMM-mediated amidation with amino acids, including alanine, leucine, and serine, significantly enhances the enzymatic resistance of alginate oligosaccharides against alginate lyase [81]. Moreover, amidation disrupts intramolecular hydrogen bonding, resulting in reduced surface tension and conductivity of electrospun solutions, while enhancing spinnability and molecular flexibility [82]. The modified alginate also exhibits greater binding affinity and selectivity toward both anionic and cationic dyes, underscoring its potential as an effective adsorbent material [83].
Food relevance: Amidation can be food-compatible when the amine donors are natural biomolecules (e.g., amino acids, peptides), providing enhanced digestibility and biocompatibility for functional food carriers; synthetic amines remain non-food.

4.1.8. Graft Modification

The limitations of SA in terms of its strong hydrophilicity, poor mechanical properties, and limited stability have driven significant interest in graft modification as a means to enhance its functionality [84]. The grafting technique is widely used to modify SA that generates active sites (e.g., free radicals or functional groups) on the polymer backbone [2]. The introduction of the desired components can effectively improve mechanical strength, thermal responsiveness, and compatibility of SA, while preserving its natural properties and biocompatibility [85]. Grafting methods are mainly classified into free radical grafting and ionic grafting. In free radical grafting, the active site exists in the form of a functional group or free radical, triggering the reaction of the monomer with the polymer backbone. In contrast, ionic grafting utilizes ionic active sites to attach monomers through ionic interactions. For example, Sand et al. [86] modified alginate by grafting N-vinyl-2-pyrrolidone via free radical initiation, resulting in improved swelling behavior, metal-ion adsorption, and flocculation properties. Similarly, free radical-induced grafting has been employed to synthesize diethylmalonic acid-grafted chitosan and alginate-g-poly (1-carboxy-4-acrylamidobenzenesulfonamide) copolymer [87,88], whose freeze-dried forms show great potential in water treatment. Shehzad et al. [89] developed calcium carbamate-partially grafted alginate hydrogel beads by reacting SA with calcium chloride and 4-phenylcarbamyl urea (Figure 12). These beads enable efficient and economical recovery of the silver ion (Ag+) from aqueous solutions. Ionic cross-linking is a major strategy for functionalizing grafted sodium alginate (NaAlg). For instance, Fe3⁺-crosslinked acrylamide (AAm)-grafted poly (vinyl alcohol) (PVA)/NaAlg microspheres achieved controlled release of the anticancer drug 5-fluorouracil (5-FU), with release behavior modulated by NaAlg content [90]. Similarly, Sr2+-crosslinked alginate hydrogels have been shown to significantly enhance mechanical strength and bioactivity, supporting their potential for biomedical applications [91]. Additionally, ionically crosslinked AAm-grafted PVA/NaAlg/sodium carboxymethylcellulose pH-sensitive microspheres have been developed for the delivery of donepezil hydrochloride, a drug used in Alzheimer’s disease treatment [92]. These examples demonstrate how ionic crosslinking synergized with grafting (e.g., Fe3⁺, Sr2⁺) enhances drug-controlled release capability, mechanical properties, and environmental responsiveness in SA-based materials.
Food relevance: Graft modification can also be adapted for food-contact or edible applications when natural or biocompatible monomers—such as polysaccharides, polyphenols, or amino acids—are employed, leading to functional coatings and smart packaging materials. In contrast, grafting with synthetic vinyl or acrylate monomers remains outside the scope of food use and is primarily suited for biomedical and environmental applications.
Overall, not all chemical routes are directly applicable to food systems; however, distinguishing food-compatible from non-food chemistries clarifies the relevance of alginate functionalization to food science. Mild esterification, amino-acid amidation, and natural crosslinking (e.g., phenolic or genipin routes) represent safe and effective strategies to enhance edible and food-contact materials, whereas stronger or synthetic chemical reactions are retained as forward-looking, non-food-oriented approaches.

4.2. Physical Modification

4.2.1. Composite Modification

Alginate-based composite material systems have attracted growing attention in functional materials research due to their high tunability and wide range of performance control. By employing multi-component synergistic strategies, the benefits of different materials can effectively be integrated, overcoming the limitations of single alginate systems in terms of mechanical strength and functional responsiveness. Current research in this area focuses primarily on three technical pathways: polymer blending, inorganic materials composites, and nanomaterial hybridization. The latter refers to the incorporation of nanostructured components (e.g., graphene, nanoclay, carbon nanotubes, metal or oxide nanoparticles) into the alginate network to generate synergistic interfacial effects, such as enhanced mechanical strength, conductivity, or barrier performance. Common fabrication techniques, such as homogeneous blending and in situ synthesis, frequently utilize the sol–gel phase transition of SA to immobilize functional components within three-dimensional network structures. This achieves synergistic enhancement of mechanical properties, biocompatibility, and environmental responsiveness. These multi-scale composite strategies provide a valuable approach for developing advanced functional materials like smart drug carriers and flexible sensors.
Polymer Material Composite
Modification of SA with polyethylene glycol diacrylate (PEGDA) has been widely investigated to improve material performance, showing significant potential in biomedical applications. Incorporation of PEGDA chains effectively reduces viscosity and enhances rheological behavior of the alginate solution, improving spinning performance and enabling the fabrication of diverse dual-network structures, such as microspheres and thin films, via modulated molding processes [27,93]. Further, Zhou et al. [94] have developed a “one-pot” synthesis method to produce high-strength semi-interpenetrating SA/PEGDA dual-network fibers, and their rheological schematic diagram is shown in Figure 13. For wound repair, integrated SA/PEGDA hydrogel systems (integrated polyethylene glycol/alginate-based hydrogel) leverage structural designs for functional synergy [95]. In tissue engineering scaffold design, sodium alginate (ALG)/PEGDA composite systems utilize “primary-secondary” dual network structures (PABC scaffolds) to accomplish multifunctional integration and enhanced performance [96]. The system combines the bioactivity and degradability of SA (by ionic crosslinking) with the mechanical stability and antioxidant function of PEGDA (by covalent crosslinking). This integration not only optimizes the rheological properties and processing adaptability of the material but also achieves multi-functional integration of antimicrobial, pro-restorative, and self-restorative functions through the spatial structure modulation.
Chitosan (CS) and its derivatives, as polycationic backbones, interact electrostatically with the polyanionic alginate, forming stable polyelectrolyte complexes [97]. The literature shows that SA/CS composites demonstrate remarkable synergistic effects in biomedical materials through polyelectrolyte complexation, dynamic cross-linking, and multi-scale structural design. The SA/CS composite systems provide many functional advantages in wound repair materials due to the synergistic effect of polyelectrolyte self-assembly and cross-linking, which endows the material with optimized mechanical strength (adapted to the dynamic mechanical environment of the wound) and interconnected porous architecture to provide physical support for cell migration and neovascularization [98]. Guan et al. [99] successfully developed a CS-SA (GTA0.3) gel system with a semi-decomposed chain network structure by glutaraldehyde (GTA)-mediated polyelectrolyte composite. Among them, the moderate cross-linking of GTA weakened the rigidity of the conventional dense polyionic network of CS-SA, forming a loose and porous amorphous structure. Additionally, Li et al. [100] developed biomedical composites with a bilayer porous structure by synergizing SA with chitosan-based materials. Both of them successfully fabricated an ordered laminar porous skeleton using high-speed homogeneous foaming and two-step freeze-drying. Building on this foundation, Zhao et al. [101] developed a composite system of carboxymethyl chitosan (CMCS) and sodium alginate (OSA). Using dynamic chemical bonding and in situ metal nanoparticle integration, they fabricated functional electroactive hydrogels (OSA/CMCS/AgNPs) without external conductive additives, establishing an innovative paradigm for smart biomaterial design. Song et al. [102] further demonstrated that this system exhibited outstanding hemostatic performance due to polyelectrolyte synergy and multiscale architecture. The flexible SA chains and the stiff chains of CMCS in the composite system synergize to enhance the mechanical properties, preventing the secondary damage caused by the insufficient mechanical strength of traditional hemostatic materials. In summary, SA/CS interpenetrating networks formed by electrostatic self-assembly not only optimize the mechanical properties and porous structure of the materials, but also confer antimicrobial, pro-restorative, adsorptive, and electroactive functions through the synergistic effect of functional groups (amino, carboxyl, etc.). These structurally innovative systems, driven by intermolecular interactions, when combined with freeze-drying and in situ nanoparticle integration, provide multifunctional platforms for the development of smart hemostatic materials, dynamically responsive wound dressings, and environmental remediation materials, significantly expanding the boundaries of biomedical applications of natural polysaccharide-based composite systems.
Polyvinyl alcohol (PVA), a water-soluble polymer with pronounced film-forming properties, has emerged as a promising material for combinations with SA due to its low density, processability, and microbial compatibility [103]. SA and PVA demonstrate multidimensional synergistic benefits in biomedical and environmental engineering via physical cross-linking, dynamic bonding, and functional composite strategies. These polymers form three-dimensional networks [104] and antibacterial composite gel beads (C/PVA/SA) [105] through physical cross-linking. On the other hand, the flexible cross-linking network of SA/PVA protects microorganisms from environmental stresses (e.g., extreme pH or temperature), strengthens material robustness, and provides a highly efficient carrier design strategy for pesticide contamination bioremediation. Furthermore, PVA, as a flexible reinforcing agent, significantly improves the mechanical strength and deformation resistance of the material through the formation of a hydrogen bonding between hydroxyl and carboxyl groups of SA. SA and PVA can also be used to develop dynamically reversible hydrogels [106] and physically cross-linked scaffolds [107], where the anionic properties of SA and hydrogen bonding of PVA synergistically create three-dimensional porous networks. These confer the scaffolds with good hydrophilicity and tunable degradation, and also endow the composite hydrogels with robust mechanical toughness and self-recovery properties. In summary, the three-dimensional porous structure fabricated by SA and PVA combines the biological activity and degradability of SA with the mechanical stability of PVA. Functional integration of biochar, nanosilver, or conductive materials further imparts synergistic adsorption–degradation, broad-spectrum antibacterial activity, and intelligent sensing properties.
Inorganic Material Composite
SA and calcium carbonate (CaCO3) exhibit strong potential for innovation in environmental remediation, smart materials, and flexible electronics through dynamic ionic cross-linking (Ca2+-COO-) and multifunctional composite strategies. For example, they can function as anticorrosive pigments [108] or enhance the mechanical and electrochemical performance of smart hydrogel [109]. The carboxyl groups of SA interact with Ca2+ released from CaCO3 to form dynamic networks, which improve the densification and corrosion inhibition of the coating, stabilize the dispersion of pigment particles, and endow hydrogels with high tensile strength and environmental adaptability. Fu et al. [110] fabricated composite hydrogel beads by incorporating CaCO3 and bentonite (Be) into a carboxymethyl cellulose (CMC)/SA matrix. In this system, carboxyl groups from SA dynamically cross-linked with Ca2+, and when combined with the lamellar structure of Be, produced a homogeneous, dense, and porous surface. Fourier transform infrared (FTIR) and X-ray diffraction (XRD) analyses confirmed the stable embedding of CaCO3/Be.
Composite systems constructed by organic-inorganic hybridization strategies between SA and silica (SiO2) exhibit unique synergistic advantages in functional materials. Organic-inorganic interpenetrating polymer networks (IPNs) can be constructed by in situ hydrolytic condensation of SA and SiO2, typically using tetraethoxysilane (TEOS) as a precursor. Molecular-level interpenetration occurs through hydrogen bonding and electrostatic interactions between carboxyl groups of SA and silicone hydroxyl groups (Si-OH), with FT-IR confirming the chemical compatibility between the two phases [111]. SA and SiO2 are cross-linked with the assistance of CaCl2 to form CS/SA/SiO2 composites. Carboxyl groups of SA create an “egg-box” gel network with Ca2+, establishing a porous anionic skeleton and a nitrogen/nitrogen-containing/corporate network. The interfacial compatibility between SA and SiO2, achieved through hydrogen bonding and electrostatic interactions, forms folded surfaces and through-pore structures, which maintain the structural integrity of the material under thermal and aqueous stress, providing a composite design strategy for biomedical scaffolds or adsorbent materials that combines bioactivity and durability [106]. Based on the multiple interactions between SA and SiO2, researchers have constructed IPNs with multi-level pore structures. This dual-phase integration combines the bioactivity of SA with the rigidity of SiO2, improving both mechanical strength and environmental tolerance.
Nanomaterials
Nanomaterials [112] have distinct physicochemical properties compared to macroscopic materials due to their size effects, surface effects, and quantum effects. They typically possess high specific surface area, superb dispersibility, and tunable interfacial characteristics [113], demonstrating broad application potential in fields such as drug delivery, food packaging, antimicrobial materials, tissue engineering, and environmental remediation [114]. Recent advances in green synthesis have promoted the use of natural polymers as templates or stabilizers [115]. Among these, polysaccharides have garnered significant attention due to their intrinsic biocompatibility [116], renewability, and functionalization potential. SA serves as an effective biotemplate for the green synthesis and functional assembly of nanomaterials, utilizing its unique carboxylic acid moieties, porous network, and regulated ionic coordination [117]. For example, granular SA carriers formed via a high-shear wet granulation process can control the synthesis of ferrite oxide (NixFeyO4) nanoparticles. This method significantly reduces the high water consumption compared with conventional liquid-phase approaches, while simplifying the operational procedures and enhancing the process scalability [118]. SA hydrogel templates can direct the synthesis of Cu2O/Cu/@carbon heterostructures. The carboxyl groups of SA coordinate with metal ions to form a three-dimensional network, guiding the in situ assembly of Cu2O/Cu heterojunctions incorporated with reduced graphene oxide (rGO). The resulting hierarchical interface synergizes with the Schottky barrier of Cu2O/Cu through the electron conduction channel of rGO, significantly improving the efficiency of photogenerated carrier separation [119]. Moreover, SA can complex with copper-alumina nanoparticles to create ternary magnetoresponsive nanofluidic systems, where electrostatic interactions between anionic SA chains and nanoparticle surfaces control the viscoelastic and interfacial behavior [120]. Through solid–liquid exchange, hydrogel templates, or electrostatic dispersion, SA can effectively control the nucleation development and interfacial architecture of nanoparticles to augment material properties. Successful implementations in ferrite synthesis, heterojunction photocatalysts, and magnetic fluids illustrate the potential of SA in structure orientation and property optimization and provide a new way for the design and macro-preparation of environmentally friendly nanocomposites.

4.2.2. Physical Processing Technique

Ultrasonication
Ultrasonication is a non-thermal physical processing technique widely applied to modify the physicochemical and functional properties of food biopolymers such as SA. [121,122] As illustrated in Figure 14, ultrasound waves induce mechanical vibration and cavitation in the liquid medium, generating localized high temperature and pressure zones that enhance molecular motion, mixing, and mass transfer [123,124,125,126,127]. These cavitation effects disrupt polymer aggregates, promote chain disentanglement, and facilitate structural rearrangements, resulting in smoother, denser, and more homogeneous films [128,129,130,131]. Ultrasonication has been shown to significantly improve encapsulation efficiency and stability of bioactive compounds in SA-based films [132,133,134,135], and to modulate mechanical and barrier properties by tailoring polymer chain interactions. Moreover, ultrasound can induce conformational unfolding of proteins or enzymes (e.g., papain), exposing active sites and enhancing their binding affinity to SA matrices [136,137].
Overall, ultrasound-assisted processing provides a controllable and energy-efficient approach to enhance the structural and functional performance of SA-based materials.
Irradiance
Irradiation technology, including electron beam and γ-ray irradiation, provides an efficient and chemical-free route for modifying SA and expanding its applications in sustainable material design and agricultural preservation. As illustrated in Figure 15, irradiation induces the formation of free radicals that initiate crosslinking or grafting reactions along SA molecular chains, enabling the development of biodegradable hydrogels and coating systems with enhanced structural integrity and functionality. Through this mechanism, SA-based materials can be precisely tailored for specific applications, offering improved mechanical properties, water absorption, and environmental stability. For example, a biodegradable SA-g-acrylamide/acrylic acid hydrogel has been developed via electron beam-induced graft copolymerization, where high-energy electrons initiate the formation of a three-dimensional (3D) crosslinked network on the SA chains, resulting in materials with superior water absorption and controllable degradation behavior [138]. Similarly, γ-irradiation-induced copolymerization enables tunable crosslinking density and improved swelling capacity of hydrogels in saline media [139]. The irradiation dose directly influences the balance between crosslinking and degradation, allowing fine modulation of gel strength and swelling ratio.
In agricultural preservation, SA coating combined with 60Co-γ irradiation has been shown to significantly extend the postharvest life of fruits such as jujube. The irradiation treatment enhances the semi-permeable nature of the SA film, reducing moisture loss and oxygen permeability while delaying nutrient degradation and microbial growth [140]. Furthermore, irradiation-assisted SA systems have been used to construct pH-responsive and superabsorbent hydrogels, offering sustainable solutions for soil moisture retention and crop protection. This residue-free and energy-efficient physical technique, when integrated with the inherent biocompatibility of SA, holds great promise for eco-friendly material development, sustainable agriculture, and biomedical engineering.

4.2.3. Physical Crosslinking

SA synergistically constructs multifunctional hydrogel systems with tunable properties via dynamically reversible physical crosslinking mechanisms (e.g., ionic bonding, hydrogen bonding, and photochemical/chemical responsive interactions). For the hydrogel delivery system, SA was modified with glycerol methacrylate to introduce photosensitive groups and combined with photocrosslinking technology to create an NO-responsive hydrogel. In this system, carboxyl groups from SA and the NO-responsive agent form a dual crosslinked network through covalent and hydrogen bonds, resulting in low swelling rates and high temperature stability [141]. Regarding gels, SA was synergistically constructed with polyacrylamide (PAAm)/pozzolanic (Pal) nanorods [142] and gelatin (Gel) [143] to form high-strength composite hydrogels (SA-PAAm/Pal and Gel-Alg) via dynamic metal ion cross-linking. Additionally, dynamic hydrogels were also constructed with starch through the pH-responsive cross-linking mechanism induced by glucolactone (GDL) [144]. Further, oxidatively modified SA (OSA) formed physically crosslinked nanogels with neutral proteins (e.g., hemoglobin), assisted by divalent cations [145]. This design endowed the materials with self-recovery capability, enhanced matrix stiffness, and high flexibility and ionic conductivity, which synergistically reinforced the mechanical strength and thermal stability of gels. Alternatively, SA constructed mechanically enhanced biosponges (SA-BG) by ionic cross-linking with Ca2+ released from bioactive glass particles (RFNP-BG). This crosslinked network maximized the porous structure and provided enhanced water resistance, while the continuous release of Ca2+ imparted bioactivity to the material [146]. To end with, the carboxyl groups in SA are coordinated with metal ions, bound to reactive substances, and modulated in response to the network, collectively conferring self-healing, energy dissipation, and environmental adaptation properties.

4.3. Biological Modification

The biomodification primarily refers to the controlled degradation and structural alteration of SA through biological methods to obtain target products with specific molecular weight, sequence structure, and desired bioactivity. Enzymatic modification [147] of alginate serves as the core strategy due to its high efficiency and substrate specificity. Compared to chemical methods, enzymatic modification provides several advantages [148], such as milder reaction conditions, higher specificity, and the production of well-defined bioactive oligosaccharides by cleaving 1,4-glycoside bonds through the β-elimination reaction. Sodium alginate lyase, as a key enzyme to realize this process, is produced by various organisms, including seaweeds, marine mollusks, fungi, viruses, as well as terrestrial and marine bacteria, with bacterial source being the most predominant [149]. These enzymes are categorized into different families based on their catalytic mechanisms and structures [150]. According to the Carbohydrate-Active enZYmes (CAZy) database, alginate lyase sequences are distributed across 12 polysaccharide lyase (PL) families (PL5, PL6, PL7, PL14, PL15, PL17, PL18, PL31, PL32, PL34, PL36, and PL39) [151]. The catalytic mechanism of alginate lyase acting on sodium alginate involves β-elimination of the 1,4-glycosidic linkage between uronic acid residues, leading to the formation of unsaturated alginate oligosaccharide (AOSs), as illustrated in Figure 16. This β-elimination process represents the general catalytic pathway of alginate lyases. According to the action mode, alginate lyases are classified as either endolytic or exolytic enzymes [152]. Endolytic lyases cleave internal bonds within the alginate polymer to generate oligosaccharides [153], while exolytic lyases further degrade oligosaccharides into monomers and/or dimers from the non-reducing ends [154]. Furthermore, alginate lyases are categorized into three groups based on their substrate specificities: poly(M)-specific lyases, poly(G)-specific lyases, and bifunctional lyases that can degrade both poly(M) and poly(G) [155]. These enzymes are widely employed to produce alginate oligosaccharides (AOSs) [156] with tailored bioactivities, demonstrating considerable potential for applications in the pharmaceutical and functional food industries [157].

4.4. Comparison of Modification Methods

Chemical modification involves the change in the chemical structure of SA through chemical reactions (such as esterification, oxidation, grafting, etc.). Physical modification, on the other hand, alters the properties of materials through physical actions (such as composite modification, physical cross-linking, etc.) without damaging the chemical structure. Biological modification is the process of regulating the structure of SA or complex bioactive compounds by biological approaches (e.g., enzymatic treatment) under relatively mild conditions. These three modification approaches—chemical, physical, and biological—are compared and analyzed in terms of modification effect, cost, process difficulty, toxicity risk, and application field, as summarized in Table 1.

5. Application of Modified SA

5.1. Food Industry

5.1.1. Food Packaging

Food packaging materials [158] protect food from external contamination, inhibit deterioration [159], and extend shelf life [160]. High-quality packaging materials [161] can also facilitate storage, transport, and sale of food, safeguard food safety, enhance product value, and meet the needs of consumers for food hygiene, esthetics, and convenience [162]. Alginate excels in film-forming, creating films characterized by high tensile strength, flexibility, tear resistance, oil resistance, stiffness, and superb gloss [163]. For instance, tannic-acid-modified SA edible films show a concrete barrier/antioxidant gain: WVP drops from 1.24 × 10−6 to 0.54 × 10−6 g m/(h Pa), with ≈89% DPPH scavenging and ≈98% UV blocking at 280 nm [164]. Furthermore, alginate is odorless and has a neutral taste; its coatings possess antimicrobial properties that inhibit bacterial growth and oxidative odor formation [165]. In food applications, alginate coatings can significantly improve the organoleptic acceptability of products and reduce cooking losses [85]. As a representative case, an electrospun PVA/SA/PVDF bilayer indicator for pork gives a strong NH3 response (ΔE ≈ 48) and prolongs shelf life by ~24 h at 25 °C [166].
For biodegradable food packaging materials, SA can be compounded with cinnamic acid-modified pectin [167] or tannic acid (TA) [164]. Quantitatively, the SA–pectin–cinnamate film shows ≈43.26% soil mass loss at 15 d while maintaining plastic-like mechanics, and the TA–SA system retains mechanical integrity while lowering WVP as above. Both compounds form biodegradable films with indicators and performance comparable to polyethylene plastic films, indicating the potential to replace traditional plastics. In smart indication packaging [168], SA was composited with agar (AG) [169], PVA/alizarin [166], PVA/rosemary anthocyanins (RAs) [170,171], and casein carboxymethylcellulose nanocomplexes (CAS-CMC-ACNs) [172]. These films enabled real-time freshness monitoring [173], pH visualization [165], and dynamic spoilage tracking [174]. Zhang et al. [175] composited SA with PVA and metal–organic framework (ZIF-8) to create a colorimetric sensor (PA-SA-ZA) for high-precision visualization of beef freshness with ΔE < 5 under light-aging, contact angle ≈ 52°, and a color–TVB-N correlation R2 ≈ 0.91 [176].
For active packaging, SA can be laminated with cyclobutanedicarboxylic acid (CBDA-10) [177], gelatin (GEL)/waste green tea extract (GTE) [178], and curdlan (CD) [163]. For example, CBDA-crosslinked SA films report TS ≈ 148 MPa, Td ≈ 249 °C, and ~60% soil mass loss at 4 weeks; curdlan–SA films improve mushroom (Volvariella volvacea) shelf life with reduced microbial load and firmer texture. These bioactive films can strengthen the mechanical strength and stability of the film while providing highly efficient dual-functionality (antimicrobial freshness preservation). To better illustrate the diversity of food-compatible SA-based packaging systems, Table 2 summarizes representative formulations, modification routes, and their corresponding performance indices (e.g., barrier, antioxidant, and antimicrobial properties). Figure 17 further visualizes their practical applications in meat, mushroom, and produce preservation, highlighting both active and intelligent functionalities. Concerning nanomaterials [179], SA can be blended with polyethylene oxide (PEO)/leaf skin tannin (Ph) [180] or complexed with pullulan polysaccharide [181] to produce nanofiber films for targeted antimicrobial protection and enhanced mechanical strength [182] and moisture barriers [183] of the films. Furthermore, sodium alginate (NaAlg) was used as a nanoencapsulation matrix for bovine lactoferrin (LFb) into nanoparticles (LFNP) and microcapsules via high-voltage electrohydrodynamic atomization (EHDA) technology. The EHDA process yields ~100–200 nm particles with |ζ| ~ 20 mV, enabling controlled stabilization/release for iron delivery in functional packaging.
Regarding other materials, SA was blended with guar gum to form a composite film (SG), where mechanical strength and water resistance were synergistically enhanced by incorporating a β-cyclodextrin/persimmon pectin-stabilized baobab oil-peeled Kleenex emulsion (BOPE) [184]. Based on this, SA was synergistically combined with guar gum and agar to construct a bilayer composite film. Within this structure, the carboxylic acid groups of SA enhanced the densification of polysaccharide network through hydrogen bonding, improving the mechanical strength and moisture resistance of the film [185], thereby achieving dual optimization of “structure-function” for food packaging [186]. Notably, several chemical modification routes relevant to food-contact applications rely on food-compatible reagents or bio-based crosslinkers (e.g., phenolic acids such as tannic/cinnamic acids, genipin or enzyme-mediated coupling), which improve barrier/mechanical/antimicrobial performance while maintaining compliance pathways typical for edible films and food-contact materials. These uses are distinguished from non-food biomedical/environmental chemistries summarized elsewhere in this review. In the ultrasound-assisted preparation of soybean protein isolate-based packaging materials, SA effectively inhibits the formation of lysinoalanine (LAL) by promoting protein conformational crosslinking [187].
In summary, SA significantly improves the antibacterial, antioxidant, and UV-barrier properties of biodegradable food packaging materials, thereby extending the shelf life of fresh foods and enabling visualization and monitoring of spoilage. These integrated environmentally friendly–functional–intelligent characteristics provide a technological foundation for the green and intelligent transformation of the food packaging industry.
Table 2. Representative modified SA-based films for food packaging.
Table 2. Representative modified SA-based films for food packaging.
SystemModificationFood MatrixKey KPl/ValueCondition/NoteReference
Curdlan–SA active filmPolysaccharide blendVolvariella volvacea (mushroom)Shelf-life ↑; microbial load ↓; firmness retainedCold storage[163]
SA–Pectin + Cinnamic AcidPhenolic ester (active, biodegradable)General≈43.26% soil mass loss at 15 d; plastic-like mechanicsSoil burial vs. PE[167]
SA + Tannic Acid (TA) edible filmPhenolic crosslinkingProduce/Meat (general)WVP 1.24 × 10−6 → 0.54 × 10−6 g·m/(h·Pa); DPPH ≈ 89%; UV-block ~ 98%@280 nmLab films; edible/food-contact[164]
PVA/SA/PVDF bilayer (alizarin sensor + antibacterial top)Electrospun bilayer indicatorPorkΔE ≈ 48 (NH3); shelf-life +~24 h @25 °CPack test[166]
PVA–SA + ZIF-8@alizarin (PA-SA-ZA)MOF-stabilized dye sensorBeefΔE < 5 under light aging; R2
0.91 (TVB-N vs. color); contact angle ~ 52°
Pack test[175]
PEO/SA nanofiber + phlorotanninElectrospun antimicrobialChickenSalmonella counts ↓; shelf-life ↑Cold storage[180]
SA/Guar Gum + BOPE Pickering filmβ-CD/persimmon pectin-stabilized oil emulsionMushroomsBrowning/shrinkage ↓; water/oxygen ingress ↓Postharvest[184]
GG/AG/SA
bilayer + TiO2
Bilayer + Pickering + nanofillersHigh-moisture produceBarrier ↑; antifungal ↑Postharvest[186]
Abbreviations: SA, sodium alginate; TA, tannic acid; PVA, poly(vinyl alcohol); PVDF, poly(vinylidene fluoride); ZIF-8, zeolitic imidazolate framework-8; GG, guar gum; AG, agar; BOPE, β-cyclodextrin/persimmon-pectin-stabilized baobab-oil Pickering emulsion; WVP, water vapor permeability; TVB-N, total volatile basic nitrogen; SGF/SIF, simulated gastric/intestinal fluid; EE%, encapsulation efficiency; arrows (↑/↓) indicate direction of change.
Figure 17. Representative applications of modified SA-based packaging for food preservation and freshness indication. (A) Active film incorporating carvacrol via a dialdehyde β-cyclodextrin/gelatin–carrageenan network for ready-to-eat meat preservation (quality retention over 0–7 days) [160]; (B) Ammonia-responsive colorimetric indicator (PA-SA-ZA; PVA/SA with ZIF-8-alizarin) tracking beef freshness over storage days [175]; (C) SA/guar-gum (SG) film combined with BOPE Pickering emulsion preserving mushrooms during cold storage (0–30 days) [184]; (D) GG/AG/SA bilayer incorporating Pickering emulsion and TiO2, exhibiting antifungal and barrier enhancement at 4, 10, and 25 °C (comparison with and without film) [186].
Figure 17. Representative applications of modified SA-based packaging for food preservation and freshness indication. (A) Active film incorporating carvacrol via a dialdehyde β-cyclodextrin/gelatin–carrageenan network for ready-to-eat meat preservation (quality retention over 0–7 days) [160]; (B) Ammonia-responsive colorimetric indicator (PA-SA-ZA; PVA/SA with ZIF-8-alizarin) tracking beef freshness over storage days [175]; (C) SA/guar-gum (SG) film combined with BOPE Pickering emulsion preserving mushrooms during cold storage (0–30 days) [184]; (D) GG/AG/SA bilayer incorporating Pickering emulsion and TiO2, exhibiting antifungal and barrier enhancement at 4, 10, and 25 °C (comparison with and without film) [186].
Foods 14 03931 g017

5.1.2. Functional Food Carrier

Functional food ingredients [188] require suitable delivery systems (or carriers) [189] to accommodate their diverse physicochemical and biological characteristics. Such carriers are essential for protecting bioactive molecules during processing and gastrointestinal transit, ensuring that they reach their intended target sites. SA is a core carrier material for functional active ingredient delivery systems, owing to its biocompatibility, controllable gelation, and designability of multi-scale structural design [190]. For lipophilic antioxidants (e.g., α-tocopherol), SA microcapsules provide gastric protection with intestinal release: cumulative release is ~29% in SGF but ~82% in SIF, with T50% ≈ 3.8 h and T70% ≈ 12.3 h (SIF) [191].
For microencapsulation, SA has been combined with whey protein isolate (WPI) [192], pectin [191], chitosan [193], and gelatin [194], as wall materials. In SA–WPI ionic-gel systems, formulation knobs (polymer ratio and Ca2+ dosage) tune encapsulation efficiency (EE%) and bead size; typical working windows used for food-grade beads are SA ≈ 0.8–1.5% (w/w), WPI ≈ 1–3%, and CaCl2 ≈ 20–60 mM, which can be optimized through response surface design to balance an increase in EE%, a reduction in particle size, and a decrease in burst release [192]. These SA-based composites demonstrate efficient encapsulation and structural control of active ingredients, optimizing both microcapsule architecture and encapsulation performance while providing high stability.
Moreover, SA plays a critical protective role in probiotic targeting delivery by forming a dual network gel (FSDN) with fish gelatin (FG) [195]. Through ionic gelation (internal or external) [196], SA forms stable networks capable of withstanding harsh gastrointestinal conditions (e.g., gastric acid, bile salts), thereby ensuring probiotic viability and controlled intestinal release. Quantitatively, FG/SA-DN capsules increase encapsulation efficiency from ~16% to ~92% as FG increases [195] and can raise acid-stage survival to ~83.6% in SGF compared with significantly lower survival of free cells [196]. SA-based microcapsules formed by combining proteins, polysaccharides, or dual-network systems exhibit notable advantages in encapsulation efficiency, targeted release, and environmental tolerance, effectively preserving essential oils, fat-soluble nutrients, probiotics, and other sensitive bioactives. To summarize representative SA carrier systems and their key physicochemical–functional readouts, Table 3 outlines typical encapsulation matrices, mechanisms, and food-oriented applications.
Beyond ionic gelation, mild chemical tailoring using food-compatible reagents—such as phenolic crosslinking or peptide/amino acid coupling—can further stabilize alginate carriers while maintaining their suitability for ingestion or food-contact use under standard migration and safety assessments. Practically, phenolic (e.g., tannic/cinnamic) crosslinking can be performed at ≤45 °C in aqueous media with pH 5–7, followed by Ca2⁺ setting to co-lock the network; enzyme-assisted (laccase/tyrosinase) routes allow low-temperature curing for heat-sensitive cargos while limiting residuals to food-compatible species. These food-oriented chemistries are explicitly distinguished from stronger non-food reactions described in biomedical and environmental contexts.

5.2. Biomedical Application

5.2.1. D Printing

SA is a low-cost and biocompatible biomaterial characterized by rapid and mild cross-linking, which has led to its widespread use in biological soft tissue repair and regeneration. Especially with the advent of 3D bioprinting technology, SA hydrogels have become increasingly popular in tissue engineering as a result of their exceptional printability [198]. Compared with traditional 2D approaches, 3D printing overcomes the restrictions of planar structure through a layer-by-layer stacking strategy, offering significant advantages in constructing bone tissue engineering scaffolds. Based on CT/MRI image data from patients with bone defects, 3D printing enables precise bionic design of personalized scaffolds, ensuring that both the macroscopic morphology and microscopic pore structure closely match natural bone tissues. Accordingly, it addresses the problems inherent to 2D printing, such as uneven distribution of cells and poor mechanical suitability resulting from the inability to construct complex 3D interconnected pore channels. The multilevel structural controllability of 3D printing not only promotes the directional migration of osteoclasts and vascularization, but also facilitates the mechanical gradient in scaffolds. By mimicking the stress-transferring properties of bone trabeculae through these designed gradients, 3D printing significantly improves the efficiency of bone integration [199]. In layered mesoporous bioactive glass/sodium alginate–sodium alginate (MBG/SA–SA) scaffolds, the printed architecture delivers porosity of ≈78% and a compressive strength of ≈4.2 MPa, matching cancellous-bone-level requirements and enabling fast release of BSA from the SA layer while maintaining sustained release of ibuprofen from the MBG/SA layer [200].
Liu et al. [201] utilized SA as a core component of 3D printed bioink, composited with nacreous layer powder (NP) to construct bionic bone scaffolds, demonstrating the benefit of precise structure-function synergy in bone tissue engineering. Similarly, Song et al. [202] constructed a drug-carrying bilayer scaffold (SG-rhEGF) by compositing SA with gelatin, achieving precise structure-function regulation in the field of skin regeneration. Quantitatively, this bilayer shows elongation at break 102.09 ± 6.74% and tensile modulus 206.83 ± 32.10 kPa, with a hydrophobic outer layer (water contact angle 112.09 ± 4.67°) and hydrophilic inner hydrogel layer (48.87 ± 5.52°), supporting barrier function and moist-wound healing. Additionally, in bone repair, SA was incorporate with MBG to form a layered double network scaffold (MBG/SA-SA), exhibiting an integrated structural-functional design (see the above porosity/strength and dual-release metrics for process targeting and benchmarking in future studies) [200]. An illustrative schematic of the 3D printing process and scaffold design is shown in Figure 18, highlighting the layer-by-layer stacking strategy, hierarchy structure, and integration of SA-based materials with functional bioactive components.

5.2.2. Drug Delivery Systems

SA is one of the most widely used natural polymers in drug delivery and biomedical applications. It functions as a gelling, thickening, and tablet disintegrating agent and has gained prominence as a cross-linking agent and for pH-controlled delivery. Due to its pH/ionic responsiveness and programmable crosslinking mechanism, SA is an important vehicle for multifunctional drug delivery systems. Alginate acts as a rate-controlling polymer in drug delivery, forming gels when hydrated with water under mild conditions without organic solvents. The resulting hydrogels formed were relatively inert, containing only distilled water or sucrose solutions [203]. Veronica et al. [204] employed SA as a natural pH-responsive carrier in oral slow-release formulations through ionic cross-linking and gelation mechanisms for controlled drug release. SA, as a polyanionic polysaccharide, can be complexed with cellulose nanocrystals (CNW). As a polyanionic polysaccharide, SA can form polyelectrolyte complexes (PECs). It is complexed with cellulose nanocrystals (CNCs) [205] to create PEC hydrogels for vaginal drug delivery, demonstrating precise environmental suitability. Similarly, SA complexed with carboxymethyl chitosan (CMC) [206] to form PEC sponges, which exhibited efficient synergistic effects in hemostatic materials. These applications highlight the smart responsiveness and multifunctional biomedical potential of SA-based PECs.
Reddy et al. [207] used SA as an ion-responsive carrier to develop a double crosslinked microbead with montmorillonite (MMT) for efficient loading and controlled release of hydrophobic drugs (e.g., curcumin CUR). The beads exhibited strong pH-gated behavior: at pH 7.4 the swelling degree rose to ~670% (CaMg formulation) within ~30–40 min versus ~110% at pH 1.2 (120 min), and CUR release reached ~68% at pH 7.4 vs. ~30% at pH 1.2 by ~1000 min; release followed a Korsmeyer–Peppas mechanism. Lin et al. [208] employed SA as a nanogel matrix to form a core–shell composite carrier by embedding liposomes, achieving a synergistic stability enhancement and controlled-release functionality. Cross-linking an internal alginate network increased liposomal particle rigidity by ~3 times, doubled blood circulation time, and enhanced drug accumulation in arthritic joints without altering surface properties. SA formed a 3D network in the core of liposomes by cross-linking with Ca2+ ions, providing mechanical support and reducing drug leakage rate, and preserving the surface properties of the liposomes to achieve target recognition.
In summary, SA-based carriers efficiently regulate drug release kinetics via dynamic ionic bonding, polyelectrolyte complexation, or dual-network construction. This enables gastric acid barrier breakthrough, intestinal-targeted slow release, and focal microenvironmental response. An illustrative schematic of SA-based multifunctional drug delivery systems is presented in Figure 19, highlighting the pH/ion-responsive hydrogel formation, the construction of PEC networks, and the liposome–SA composite carriers for controlled drug release and enhanced stability.

5.3. Environmental Engineering

Sodium alginate and its derivatives exhibit significant potential in environmental remediation. Their surfaces are rich in hydroxyl (-OH) and carboxyl (-COOH) functional groups that can efficiently capture dye molecules, heavy metal ions, and organic pollutants [209] through ion exchange, electrostatic interactions, and coordination mechanisms. To improve adsorption performance, SA is commonly compounded with functional materials (e.g., activated carbon, graphene oxide, biochar, carbon nanotubes, etc.), and either physically embedded or chemically crosslinked. This creates a porous structure while introducing specific adsorption sites. Furthermore, cross-linking modification augments the spatial distribution of functional groups, significantly improving the stability and recyclability of materials. The environmentally friendly properties of these materials are derived from the biodegradability and renewability of their natural sources, offering innovative solutions for green water treatment technologies.

5.3.1. Metal Ion Adsorption

With rapid industrialization, heavy metal pollution [210] has become a major global environmental challenge [211]. Common heavy metal ion removal processes [212] include membrane separation, electrochemical recovery, chemical precipitation, and ion-exchange, as summarized in Table 4. In addition to these conventional techniques [213], functionalized SA-based adsorbents have recently attracted significant attention owing to their tunable surface chemistry, biocompatibility, and environmental sustainability. The key advantages and limitations of both traditional methods and SA-based functionalized systems are compared in Table 4. As a naturally occurring anionic polysaccharide, the carboxyl (-COOH) and hydroxyl (-OH) functional groups in the SA molecular chain can specifically bind to heavy metal ions via coordination, ion exchange, and electrostatic adsorption [214]. SA-based adsorbents exhibit high adsorption capacities for heavy metal ions, and their adsorption mechanism [215] is depicted in Figure 20. Composites of SA and PVA form Fe0-Fe3O4 nanocomposite beads [216], which demonstrate enhanced physical properties and catalytic reactivity in Cr(VI) removal. Under an optimized bead formulation (5.0 wt% PVA, 1.5 wt% SA with acidification/reduction), only 0.075 wt% Fe0 with 0.30 wt% Fe3O4 was sufficient to completely remove 20 mg·L−1 Cr(VI); the removal efficiency decreased from 100% to 79.5% as the initial Cr(VI) rose from 5 to 40 mg·L−1, and from 99.3% to 76.3% as pH increased 3.0 to 11.0; after four reuse cycles the beads retained 69.8% efficiency. Therefore, researchers incorporated graphene oxide (GO) into this system construct to develop 3D hydrogel microspheres (SPGs) [217], exhibiting efficient synergistic adsorption for heavy metal ions. Encapsulating magnetite-GO composites with SA produced core–shell structure adsorption microbeads (mGO/beads) [218], providing dual advantages of efficient adsorption and convenient recovery. Furthermore, SA is loaded onto melamine sponge (MS) via an in situ gelation process, creating composite adsorbents (alginate-MS) [219], significantly enhancing mechanical strength and cyclic stability while achieving a balance between adsorption efficiency and structural robustness.
Overall, the incorporation of sodium alginate not only enhances the adsorption efficiency of metal ions in wastewater but also provides a scalable and renewable solution for green and efficient heavy metal removal.

5.3.2. Dye Wastewater Treatment

Synthetic dyes are among the most persistent water pollutants, posing dual threats to aquatic ecosystems due to their distinct physicochemical properties. Even at low concentrations, they can cause severe sensory pollution and disrupt aquatic food chains via bioaccumulation. These color-revealing compounds (cationic, anionic, and non-ionic) are extensively used in industries such as textiles, paper, plastics, and leather manufacturing. According to the United Nations Environment Programme (UNEP), globally, approximately 280,000 tons of dyestuffs enter the aquatic environments annually through industrial wastewater. A small portion of these dyestuffs resists biodegradation due to their stable molecular structure, forming persistent pollutants. Importantly, industrial dyes like triphenylmethane and azo types reduce water light transmittance, and their metabolic intermediates (e.g., carcinogenic aromatic amines) exhibit teratogenicity, highlighting the urgency of pollution control [230]. SA can be incorporated with TEMPO oxidized cellulose (TOC) [231] and PVA/starch [232] to construct composite gels. SA enables a novel photocatalytic-biological synergistic system (ICPB) by coupling R. palustris with carbon nanotubes-silver-modified titanium dioxide (CNT-Ag-TiO2) photocatalysts [233], overcoming the limitations of the stability of traditional adsorbents. Additionally, a composite photocatalytic material (Cu-BTC @Alg/Fe3O4) was synthesized by integrating copper-based MOF (Cu-BTC) with magnetic Fe3O4 [234], addressing photocatalyst deactivation and secondary pollution, and providing an environmentally friendly treatment for highly toxic dyes. Quantitatively, Cu-BTC @Alg/Fe3O4 exhibited a BET surface area of ~160 m2·g−1, an adsorption capacity of ~200 mg·g−1 for Rhodamine B with ~97% removal, achieving equilibrium within ~100 min and following pseudo-second-order kinetics (R2 ≈ 0.999) [233].
Overall, the incorporation of SA not only improves the adsorption efficiency of metal ions in wastewater but also provides a scalable and renewable solution for green and efficient heavy metal removal.

5.4. Other Applications

5.4.1. Smart Textiles

The rapid development of flexible electronics is propelling smart fiber research into a new phase, focusing on the construction of wearable electronic platforms with inherent flexibility and system integration through textile-based components. The global smart textiles market is predicted to increase remarkably, thanks to breakthroughs in woven smart fibers in thermal management, electromagnetic protection, and bioelectrical signal coupling. Such fiber-level systems achieve conformal contact with the curved surfaces of the human body, providing superior breathability and biocompatibility than traditional rigid electronics [235]. SA builds multifunctional fibers by compositing with liquid metal (LM) micro/nanodroplets through a wet-spinning process [236], and also modifies cotton nonwoven fabrics in synergy with silver nitrate via the sol–gel method [237]. Practically, an SA–Ag system using 1 wt% sodium alginate and 15 wt% AgNO3 achieves surface resistivity < 100 Ω sq−1 on cotton nonwovens, meeting conductive-textile requirements while retaining textile handle [237]. These innovations deliver breakthrough material properties in smart textiles. The composite system maintains the high air permeability and flexibility of the material, while the ionic cross-linking network of SA imparts self-repairing properties, providing a solution for the human–machine interfaces in extreme environments.

5.4.2. Microbial Fuel Cell

Alginates and their derivatives have emerged as promising alternative materials for microbial fuel cells. Due to their unique physicochemical properties, these natural polymers are particularly suitable for various fuel cell systems, including polymer biofuel cells, electrolyte fuel cells, polymer electrolyte fuel cells, and direct methanol fuel cells. In these applications, alginates and their compounds can effectively replace conventional proton exchange membranes, demonstrating strong performance and application prospects [85]. SA can be composited with polyaniline (PANI) [238], agar-activated carbon (AC) [239], or super-activated carbon (SAC) [240] to fabricate self-supporting anodes, achieving synergistic enhancement in electron storage and energy conversion in microbial fuel cells (MFCs). This induced dual optimization of electrode performance and biocompatibility, notably optimizing the microbial immobilization and electron transfer efficiency. Additionally, SA maintains stable power generation in complex wastewater environments through a two-layer hydrogel structure [241], i.e., the inner layer of SA and Fe3O4 to form a 3D porous network, and the outer layer of SA covalently cross-linking with PVA to form a dense protective layer. Under high-salinity waste-leachate feeding, this double-layer SA bioanode delivered an open-circuit voltage (OCV) ~ 1.17 V and an operating voltage ~ 781 mV, evidencing salt-resistant, long-term stable output. This “functional gradient” bioelectrode development is for high salt/high bacterial load wastewater. Additionally, SA can be compounded with PVA/chitosan to form a functionalized ion-conducting layers that optimize electrode interfaces and electrical output performance in MFCs [239].

5.4.3. Soil Conditioner

Improper application of chemical fertilizers (e.g., unsuitable timing or dosage) leads to nutrient loss, soil degradation [242], and water pollution [243], contradicting modern environmental protection principles. Although various soil amendments are widely employed in agricultural, natural polymers have recently gained popularity for improving the utilization of fertilizers [244]. Among these, SA has become a key functional material in soil remediation and sustainable agriculture due to its biodegradability, dynamic cross-linking ability, and multiscale structural designability [245]. SA can be compounded with sulfide-modified iron nanoparticles [246], biochar [247], and gelatin [248] to construct a soil conditioner or hydrogel, achieving the synergistic effect of “degradation-passivation-structural” improvement in soil pollution control, and the multifunctional integration of “carrier-nutrient-improvement” for intelligent slow-release fertilizers. Specifically, for SA/sulfide-coated Fe nanoparticles activating persulfate in contaminated soil, response-surface optimization indicated a theoretical 99.79% TBBPA degradation at 34.28 °C using 3.57 g·kg−1 SA@S-Fe and 36.35 mM persulfate, while heavy metals (Fe, Cu, Zn) remained in stable residual fractions and soil SOM/TN/C/N/TOC only slightly decreased [246]. Furthermore, SA can capture organic pesticides and inhibit their migration and diffusion via electron beam irradiation grafting [138] and amphiphilic modification technology [249]. This addresses the residual pollution from traditional improvers, while providing environmentally friendly solutions for water-saving agriculture and precision fertilization. In summary, SA-based system, through nanoparticle compositing, biochar integration, or amphiphilic modification, achieve multi-effect synergy by targeted degradation of pollutants, intelligent slow release of nutrients, and optimization of soil microstructure. This approach overcomes the limitation of traditional single-function and residual pollution of soil conditioners, providing effective solutions for green agriculture and ecological restoration.

6. Conclusions and Outlook

This review highlights that the modification of sodium alginate should be viewed within a food-oriented and food-contact framework, rather than as the development of new food additives. Various modification strategies—including chemical tailoring, physical processing, and enzymatic regulation—have effectively expanded the functional potential of SA while maintaining its safety and biocompatibility for edible and food-related applications.
Food-compatible chemical routes (e.g., phenolic, enzymatic, and genipin crosslinking) and green physical techniques (e.g., ultrasound treatment, irradiation, and biopolymer blending) enable the fabrication of biodegradable films, coatings, and delivery systems that meet both functional and regulatory requirements. These mild modification approaches enhance mechanical strength, barrier performance, and antimicrobial activity, supporting the development of active and intelligent food packaging and functional ingredient carriers.
Beyond food applications, insights derived from non-food-oriented modification studies (such as graft copolymerization or nanoparticle incorporation) can guide the design of next-generation bio-based materials with tailored structure–function relationships. Future research should focus on integrating molecular modification mechanisms with performance evaluation in real food systems, while advancing scalable, low-residue, and life cycle-optimized production processes. Through interdisciplinary efforts linking food science, materials chemistry, and sustainable engineering, sodium alginate is expected to serve as a core platform for safe, multifunctional, and eco-friendly food systems.

Author Contributions

W.W.: Writing—original draft, Investigation, Formal analysis, Visualization, Conceptualization. Y.H.: Conceptualization, Methodology, Supervision, Writing—review and editing. Y.P.: Resources, Supervision. M.D.: Conceptualization; Methodology; Supervision; Writing—review and editing. C.D.: Conceptualization, Funding Acquisition, Resources, Supervision, Writing—review and editing. M.Z.: Supervision, Methodology. R.H.: Supervision, Methodology. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by Agricultural Science and Technology Support Program of Danyang in China (SNY202304).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

Thank you to King Draw software for structural diagramming and PPT software for creating diagrams.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Akbar, M.; Yaqoob, A.; Ahmad, A.; Luque, R. Chapter 1—Sodium Alginate: An Overview. In Sodium Alginate-Based Nanomaterials for Wastewater Treatment; Ahmad, A., Ahmad, I., Kamal, T., Asiri, A.M., Tabassum, S., Eds.; Micro and Nano Technologies; Elsevier: Amsterdam, The Netherlands, 2023; pp. 1–17. ISBN 978-0-12-823551-5. [Google Scholar]
  2. Wang, K. Research Progress and Application of Sodium Alginate Grafted Copolymer Composites. Int. J. Res. Eng. Sci. IJRES 2024, 12, 238–257. [Google Scholar]
  3. European Parliament and Council of the European Union. Regulation (EC) No 1333/2008 of the European Parliament and of the Council of 16 December 2008 on Food Additives. Off. J. Eur. Union 2008, L 354, 16–33. [Google Scholar]
  4. Commission of the European Union. Commission Regulation (EU) No 231/2012 of 9 March 2012 Laying down Specifications for Food Additives Listed in Annexes II and III to Regulation (EC) No 1333/2008. Off. J. Eur. Union 2012, L 83, 1–295. [Google Scholar]
  5. U.S. Food and Drug Administration (FDA). Code of Federal Regulations, Title 21, §184.1724—Sodium Alginate. Available online: https://www.ecfr.gov/current/title-21/part-184/section-184.1724 (accessed on 2 November 2025).
  6. Zhang, J.; Zhang, S.; Liu, C.; Lu, Z.; Li, M.; Hurren, C.; Wang, D. Photopolymerized Multifunctional Sodium Alginate-Based Hydrogel for Antibacterial and Coagulation Dressings. Int. J. Biol. Macromol. 2024, 260, 129428. [Google Scholar] [CrossRef]
  7. Chowdhury, S.; Chakraborty, S.; Maity, M.; Hasnain, M.S.; Nayak, A.K. Chapter 7—Biocomposites of Alginates in Drug Delivery. In Alginates in Drug Delivery; Nayak, A.K., Hasnain, M.S., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 153–185. ISBN 978-0-12-817640-5. [Google Scholar]
  8. Łabowska, M.B.; Michalak, I.; Detyna, J. Methods of Extraction, Physicochemical Properties of Alginates and Their Applications in Biomedical Field—A Review. Open Chem. 2019, 17, 738–762. [Google Scholar] [CrossRef]
  9. Hernández-carmona, G.; McHugh, D.J.; Arvizu-Higuera, D.L.; Rodríguez-montesinos, Y.E. Pilot Plant Scale Extraction of Alginate from Macrocystis pyrifera. 1. Effect of Pre-Extraction Treatments on Yield and Quality of Alginate. J. Appl. Phycol. 1998, 10, 507–513. [Google Scholar] [CrossRef]
  10. Vauchel, P.; Arhaliass, A.; Legrand, J.; Kaas, R.; Baron, R. Decrease in dynamic viscosity and average molecular weight of alginate from Laminaria digitata during alkaline extraction. J. Phycol. 2008, 44, 515–517. [Google Scholar] [CrossRef]
  11. Silva, M.; Gomes, F.; Oliveira, F.; Morais, S.; Delerue-Matos, C. Microwave-Assisted Alginate Extraction from Portuguese Saccorhiza polyschides—Influence of Acid Pretreatment. World Acad. Sci. Eng. Technol. Int. J. Biotechnol. Bioeng. 2015, 9, 804–808. [Google Scholar] [CrossRef]
  12. Wang, F.; Yu, X.; Cui, Y.; Xu, L.; Huo, S.; Ding, Z.; Hu, Q.; Xie, W.; Xiao, H.; Zhang, D. Efficient Extraction of Phycobiliproteins from Dry Biomass of Spirulina platensis Using Sodium Chloride as Extraction Enhancer. Food Chem. 2023, 406, 135005. [Google Scholar] [CrossRef]
  13. Youssouf, L.; Lallemand, L.; Giraud, P.; Soulé, F.; Bhaw-Luximon, A.; Meilhac, O.; D’Hellencourt, C.L.; Jhurry, D.; Couprie, J. Ultrasound-Assisted Extraction and Structural Characterization by NMR of Alginates and Carrageenans from Seaweeds. Carbohydr. Polym. 2017, 166, 55–63. [Google Scholar] [CrossRef]
  14. Dabbour, M.; Hamoda, A.; Mintah, B.K.; Wahia, H.; Betchem, G.; Yolandani; Xu, H.; He, R.; Ma, H. Ultrasonic-Aided Extraction and Degossypolization of Cottonseed Meal Protein: Optimization and Characterization of Functional Traits and Molecular Structure. Ind. Crops Prod. 2023, 204, 117261. [Google Scholar] [CrossRef]
  15. Ding, Q.; Li, Z.; Wu, W.; Su, Y.; Sun, N.; Luo, L.; Ma, H.; He, R. Physicochemical and Functional Properties of Dietary Fiber from Nannochloropsis Oceanica: A Comparison of Alkaline and Ultrasonic-Assisted Alkaline Extractions. LWT Food Sci. Technol. 2020, 133, 110080. [Google Scholar] [CrossRef]
  16. Ayim, I.; Ma, H.; Alenyorege, E.A.; Ali, Z.; Zhou, C.; Donkor, P.O. Effect of Alkali Concentration on Functionality, Lysinoalanine Formation, and Structural Characteristics of Tea Residue Proteins. J. Food Process Eng. 2018, 41, e12877. [Google Scholar] [CrossRef]
  17. Ummat, V.; Zhao, M.; Sivagnanam, S.P.; Karuppusamy, S.; Lyons, H.; Fitzpatrick, S.; Noore, S.; Rai, D.K.; Gómez-Mascaraque, L.G.; O’Donnell, C.; et al. Ultrasound-Assisted Extraction of Alginate from Fucus Vesiculosus Seaweed By-Product Post-Fucoidan Extraction. Mar. Drugs 2024, 22, 516. [Google Scholar] [CrossRef]
  18. Ayim, I.; Ma, H.; Alenyorege, E.A. Optimizing and Predicting Degree of Hydrolysis of Ultrasound Assisted Sodium Hydroxide Extraction of Protein from Tea (Camellia sinensis L.) Residue Using Response Surface Methodology. J. Food Sci. Technol. 2018, 55, 5166–5174. [Google Scholar] [CrossRef]
  19. Ayim, I.; Ma, H.; Alenyorege, E.A.; Ali, Z.; Donkor, P.O. Influence of Ultrasound Pretreatment on Enzymolysis Kinetics and Thermodynamics of Sodium Hydroxide Extracted Proteins from Tea Residue. J. Food Sci. Technol. 2018, 55, 1037–1046. [Google Scholar] [CrossRef]
  20. Iqbal, M.W.; Riaz, T.; Mahmood, S.; Bilal, M.; Manzoor, M.F.; Qamar, S.A.; Qi, X. Fucoidan-Based Nanomaterial and Its Multifunctional Role for Pharmaceutical and Biomedical Applications. Crit. Rev. Food Sci. Nutr. 2024, 64, 354–380. [Google Scholar] [CrossRef]
  21. Mohammed, A.; Rivers, A.; Stuckey, D.C.; Ward, K. Alginate Extraction from Sargassum Seaweed in the Caribbean Region: Optimization Using Response Surface Methodology. Carbohydr. Polym. 2020, 245, 116419. [Google Scholar] [CrossRef]
  22. Hasnain, M.S.; Jameel, E.; Mohanta, B.; Dhara, A.K.; Alkahtani, S.; Nayak, A.K. Chapter 1—Alginates: Sources, Structure, and Properties. In Alginates in Drug Delivery; Nayak, A.K., Hasnain, M.S., Eds.; Academic Press: Cambridge, MA, USA, 2020; pp. 1–17. ISBN 978-0-12-817640-5. [Google Scholar]
  23. Tao, L.; Shi, C.; Zi, Y.; Zhang, H.; Wang, X.; Zhong, J. A Review on the Chemical Modification of Alginates for Food Research: Chemical Nature, Modification Methods, Product Types, and Application. Food Hydrocoll. 2024, 147, 109338. [Google Scholar] [CrossRef]
  24. Yang, J.; Xie, Y.; He, W. Research Progress on Chemical Modification of Alginate: A Review. Carbohydr. Polym. 2011, 84, 33–39. [Google Scholar] [CrossRef]
  25. Cao, L.; Lu, W.; Mata, A.; Nishinari, K.; Fang, Y. Egg-Box Model-Based Gelation of Alginate and Pectin: A Review. Carbohydr. Polym. 2020, 242, 116389. [Google Scholar] [CrossRef]
  26. Hernández-Carmona, G.; Freile-Pelegrín, Y.; Hernández-Garibay, E. 14—Conventional and Alternative Technologies for the Extraction of Algal Polysaccharides. In Functional Ingredients from Algae for Foods and Nutraceuticals; Domínguez, H., Ed.; Woodhead Publishing Series in Food Science, Technology and Nutrition; Woodhead Publishing: Cambridge, UK, 2013; pp. 475–516. ISBN 978-0-85709-512-1. [Google Scholar]
  27. Li, Q.-Q.; Xu, D.; Dong, Q.-W.; Song, X.-J.; Chen, Y.-B.; Cui, Y.-L. Biomedical Potentials of Alginate via Physical, Chemical, and Biological Modifications. Int. J. Biol. Macromol. 2024, 277, 134409. [Google Scholar] [CrossRef] [PubMed]
  28. Hu, C.; Lu, W.; Sun, C.; Zhao, Y.; Zhang, Y.; Fang, Y. Gelation Behavior and Mechanism of Alginate with Calcium: Dependence on Monovalent Counterions. Carbohydr. Polym. 2022, 294, 119788. [Google Scholar] [CrossRef] [PubMed]
  29. George, M.; Abraham, T.E. Polyionic Hydrocolloids for the Intestinal Delivery of Protein Drugs: Alginate and Chitosan—A Review. J. Control. Release 2006, 114, 1–14. [Google Scholar] [CrossRef] [PubMed]
  30. Abka-khajouei, R.; Tounsi, L.; Shahabi, N.; Patel, A.K.; Abdelkafi, S.; Michaud, P. Structures, Properties and Applications of Alginates. Mar. Drugs 2022, 20, 364. [Google Scholar] [CrossRef]
  31. Saji, S.; Hebden, A.; Goswami, P.; Du, C. A Brief Review on the Development of Alginate Extraction Process and Its Sustainability. Sustainability 2022, 14, 5181. [Google Scholar] [CrossRef]
  32. Merck KGaA (Sigma-Aldrich). Sodium Alginate—Pharmaceutical Secondary Standard; Certified Reference Material (PHR1471), CAS 9005-38-3. Available online: https://www.sigmaaldrich.com/HK/en/product/sial/phr1471 (accessed on 2 November 2025).
  33. Rostami, Z.; Tabarsa, M.; You, S.; Rezaei, M. Relationship between Molecular Weights and Biological Properties of Alginates Extracted under Different Methods from Colpomenia Peregrina. Process Biochem. 2017, 58, 289–297. [Google Scholar] [CrossRef]
  34. Xiao, Q.; Tong, Q.; Zhou, Y.; Deng, F. Rheological Properties of Pullulan-Sodium Alginate Based Solutions during Film Formation. Carbohydr. Polym. 2015, 130, 49–56. [Google Scholar] [CrossRef]
  35. Rosiak, P.; Latanska, I.; Paul, P.; Sujka, W.; Kolesinska, B. Modification of Alginates to Modulate Their Physic-Chemical Properties and Obtain Biomaterials with Different Functional Properties. Molecules 2021, 26, 7264. [Google Scholar] [CrossRef]
  36. Pawar, S.N.; Edgar, K.J. Alginate Derivatization: A Review of Chemistry, Properties and Applications. Biomaterials 2012, 33, 3279–3305. [Google Scholar] [CrossRef]
  37. Sun, Q.; Zhang, Z.; Zhang, R.; Gao, R.; McClements, D.J. Development of Functional or Medical Foods for Oral Administration of Insulin for Diabetes Treatment: Gastroprotective Edible Microgels. J. Agric. Food Chem. 2018, 66, 4820–4826. [Google Scholar] [CrossRef] [PubMed]
  38. Hu, C.; Lu, W.; Mata, A.; Nishinari, K.; Fang, Y. Ions-Induced Gelation of Alginate: Mechanisms and Applications. Int. J. Biol. Macromol. 2021, 177, 578–588. [Google Scholar] [CrossRef] [PubMed]
  39. Shi, H.; Zhou, T.; Wang, X.; Zou, Y.; Wang, D.; Xu, W. Effects of the Structure and Gel Properties of Myofibrillar Protein on Chicken Breast Quality Treated with Ultrasound-Assisted Potassium Alginate. Food Chem. 2021, 358, 129873. [Google Scholar] [CrossRef] [PubMed]
  40. Sikorski, P.; Mo, F.; Skjåk-Bræk, G.; Stokke, B. Evidence for Egg-Box-Compatible Interactions in Calcium−Alginate Gels from Fiber X-Ray Diffraction. Biomacromolecules 2007, 8, 2098–2103. [Google Scholar] [CrossRef]
  41. Ching, S.H.; Bansal, N.; Bhandari, B. Alginate Gel Particles-A Review of Production Techniques and Physical Properties. Crit. Rev. Food Sci. Nutr. 2017, 57, 1133–1152. [Google Scholar] [CrossRef]
  42. Shu, J.; McClements, D.J.; Luo, S.; Ye, J.; Liu, C. Effect of Internal and External Gelation on the Physical Properties, Water Distribution, and Lycopene Encapsulation Properties of Alginate-Based Emulsion Gels. Food Hydrocoll. 2023, 139, 108499. [Google Scholar] [CrossRef]
  43. Reis, C.P.; Neufeld, R.J.; Vilela, S.; Ribeiro, A.J.; Veiga, F. Review and Current Status of Emulsion/Dispersion Technology Using an Internal Gelation Process for the Design of Alginate Particles. J. Microencapsul. 2006, 23, 245–257. [Google Scholar] [CrossRef]
  44. Houben, S.; Pitet, L.M. Ionic Crosslinking Strategies for Poly(Acrylamide)/Alginate Hybrid Hydrogels. React. Funct. Polym. 2023, 191, 105676. [Google Scholar] [CrossRef]
  45. Leng, W.; Li, Y.; Niu, J.; Yuan, L.; Li, X.; Gao, R. Protein-Glutaminase-Mediated Functional Modification of Fish Myofibrillar Protein and Its Gelation Mechanism. LWT Food Sci. Technol. 2025, 223, 117782. [Google Scholar] [CrossRef]
  46. Monto, A.R.; Li, M.; Wang, X.; Wijaya, G.Y.A.; Shi, T.; Xiong, Z.; Yuan, L.; Jin, W.; Li, J.; Gao, R. Recent Developments in Maintaining Gel Properties of Surimi Products under Reduced Salt Conditions and Use of Additives. Crit. Rev. Food Sci. Nutr. 2022, 62, 8518–8533. [Google Scholar] [CrossRef]
  47. Gao, R.; Wang, X.; Shi, T.; Wijaya, G.Y.A.; Bai, F.; Wang, J.; Jin, W.; Yuan, L. Enhanced Physical Properties of Reduced-Salt Surimi Gels from Amur Sturgeon (Acipenser schrenckii) by l-Arginine and l-Histidine. J. Food Process. Preserv. 2021, 45, e15887. [Google Scholar] [CrossRef]
  48. Ren, Y.; Wang, Q.; Xu, W.; Yang, M.; Guo, W.; He, S.; Liu, W. Alginate-Based Hydrogels Mediated Biomedical Applications: A Review. Int. J. Biol. Macromol. 2024, 279, 135019. [Google Scholar] [CrossRef] [PubMed]
  49. Rahman, M.; Shahid, A.; Hossain, T.; Sheikh, S.; Rahman, S.; Uddin, N.; Rahim, A.; Khan, R.A.; Hossain, I. Sources, Extractions, and Applications of Alginate: A Review. Discov. Appl. Sci. 2024, 6, 443. [Google Scholar] [CrossRef]
  50. Zhang, S.; Dong, J.; Pan, R.; Xu, Z.; Li, M.; Zang, R. Structures, Properties, and Bioengineering Applications of Alginates and Hyaluronic Acid. Polymers 2023, 15, 2149. [Google Scholar] [CrossRef]
  51. Hao, M.; Li, Z.; Huang, X.; Wang, Y.; Wei, X.; Zou, X.; Shi, J.; Huang, Z.; Yin, L.; Gao, L.; et al. A Cell-Based Electrochemical Taste Sensor for Detection of Hydroxy-α-Sanshool. Food Chem. 2023, 418, 135941. [Google Scholar] [CrossRef]
  52. Wang, J.; Li, K.; Yuan, H. Preparation of Ag-Metal Organic Frameworks-Loaded Sodium Alginate Hydrogel for the Treatment of Periodontitis. Sci. Rep. 2025, 15, 800. [Google Scholar] [CrossRef]
  53. Bao, Y.; Yan, D.; Xu, G.; Hong, H.; Gao, R. Effects of Chopping Temperature on the Gel Quality of Silver Carp (Hypophthalmichthys molitrix) Surimi: Insight from Gel-Based Proteomics. J. Sci. Food Agric. 2024, 104, 8212–8218. [Google Scholar] [CrossRef]
  54. Yan, P.; Lan, W.; Xie, J. Modification on Sodium Alginate for Food Preservation: A Review. Trends Food Sci. Technol. 2024, 143, 104217. [Google Scholar] [CrossRef]
  55. Saad, M.A.; Sadik, E.R.; Eldakiky, B.M.; Moustafa, H.; Fadl, E.; He, Z.; Elashtoukhy, E.Z.; Khalifa, R.E.; Zewail, T.M.M. Synthesis and Characterization of an Innovative Sodium Alginate/Polyvinyl Alcohol Bioartificial Hydrogel for Forward-Osmosis Desalination. Sci. Rep. 2024, 14, 8225. [Google Scholar] [CrossRef]
  56. Eskhan, A.; Banat, F. Removal of Oil from Water by Calcium Alginate Hydrogel Modified with Maleic Anhydride. J. Polym. Environ. 2018, 26, 2901–2916. [Google Scholar] [CrossRef]
  57. Behrooznia, Z.; Kouhanestani, D.J.N. Preparation and Characterization of Thiolated Alginate Using Esterification Method for Biomedical Application. Conference Presentation, Iran, November 2023. Available online: https://www.researchgate.net/publication/381861230 (accessed on 3 June 2025).
  58. Gomez, C.G.; Rinaudo, M.; Villar, M.A. Oxidation of Sodium Alginate and Characterization of the Oxidized Derivatives. Carbohydr. Polym. 2007, 67, 296–304. [Google Scholar] [CrossRef]
  59. Emami, Z.; Ehsani, M.; Zandi, M.; Foudazi, R. Controlling Alginate Oxidation Conditions for Making Alginate-Gelatin Hydrogels. Carbohydr. Polym. 2018, 198, 509–517. [Google Scholar] [CrossRef] [PubMed]
  60. Ingerma, K.M.; Reile, I.; Tuvikene, R. Regioselective Sulfation of Alginate at 2-O-Position of Mannuronic Acid Unit with Py∙SO3 in DMSO. Carbohydr. Res. 2024, 545, 109276. [Google Scholar] [CrossRef] [PubMed]
  61. Zare-Gachi, M.; Sadeghi, A.; Choshali, M.A.; Ghadimi, T.; Forghani, S.F.; Pezeshki-Modaress, M.; Daemi, H. Degree of Sulfation of Freeze-Dried Calcium Alginate Sulfate Scaffolds Dramatically Influence Healing Rate of Full-Thickness Diabetic Wounds. Int. J. Biol. Macromol. 2024, 283, 137557. [Google Scholar] [CrossRef]
  62. Yao, Z.; Xu, L.; Wang, B.; Ye, T.; Li, Y.; Wu, H. Optimization of Preparation Conditions, Molecular Structure Analysis and Antitumor Activity of Sulfated Sodium Alginate Oligosaccharides. Eur. Polym. J. 2023, 201, 112571. [Google Scholar] [CrossRef]
  63. Wijesinghe, W.A.J.P.; Jeon, Y.-J. Biological Activities and Potential Industrial Applications of Fucose Rich Sulfated Polysaccharides and Fucoidans Isolated from Brown Seaweeds: A Review. Carbohydr. Polym. 2012, 88, 13–20. [Google Scholar] [CrossRef]
  64. Tomohara, K.; Ohashi, N.; Uchida, T.; Nose, T. Synthesis of Natural Product Hybrids by the Ugi Reaction in Complex Media Containing Plant Extracts. Sci. Rep. 2022, 12, 15568. [Google Scholar] [CrossRef]
  65. Ullrich, A.; Kazmaier, U. A Half Century of the Ugi Reaction: Classic Variant. In Organic Reactions; Wiley: Hoboken, NJ, USA, 2023; pp. 1–560. ISBN 978-0-471-26418-7. [Google Scholar]
  66. Putri, A.P.; Picchioni, F.; Harjanto, S.; Chalid, M. Alginate Modification and Lectin-Conjugation Approach to Synthesize the Mucoadhesive Matrix. Appl. Sci. 2021, 11, 11818. [Google Scholar] [CrossRef]
  67. Liu, Z.; Chen, X.; Wen, Y.; Bao, C.; Liu, C.; Cao, S.; Yan, H.; Lin, Q. Chemical Modification of Alginate with Tosylmethyl Isocyanide, Propionaldehyde and Octylamine via the Ugi Reaction for Hydrophobic Drug Delivery. Polym. Bull. 2022, 79, 7809–7826. [Google Scholar] [CrossRef]
  68. Liu, Z.; Chen, X.; Huang, Z.; Shi, J.; Liu, C.; Cao, S.; Yan, H.; Lin, Q. Self-Assembled Oleylamine Grafted Alginate Aggregates for Hydrophobic Drugs Loading and Controlled Release. Int. J. Polym. Mater. Polym. Biomater. 2023, 72, 212–223. [Google Scholar] [CrossRef]
  69. Velema, W.A.; Lu, Z. Chemical RNA Cross-Linking: Mechanisms, Computational Analysis, and Biological Applications. JACS Au 2023, 3, 316–332. [Google Scholar] [CrossRef]
  70. Eyre, D.R.; Weis, M.; Rai, J. Analyses of Lysine Aldehyde Cross-Linking in Collagen Reveal That the Mature Cross-Link Histidinohydroxylysinonorleucine Is an Artifact. J. Biol. Chem. 2019, 294, 6578–6590. [Google Scholar] [CrossRef]
  71. Acharya, A.S.; Manning, J.M. Reaction of Glycolaldehyde with Proteins: Latent Crosslinking Potential of Alpha-Hydroxyaldehydes. Proc. Natl. Acad. Sci. USA 1983, 80, 3590–3594. [Google Scholar] [CrossRef]
  72. Rumon, M.H.; Rahman, S.; Akib, A.A.; Sohag, S.; Rakib, R.A.; Khan, A.R.; Yesmin, F.; Shakil, M.S.; Rahman Khan, M.M. Progress in Hydrogel Toughening: Addressing Structural and Crosslinking Challenges for Biomedical Applications. Discov. Mater. 2025, 5, 5. [Google Scholar] [CrossRef]
  73. Alavarse, A.C.; Frachini, E.C.G.; da Silva, R.L.C.G.; Lima, V.H.; Shavandi, A.; Petri, D.F.S. Crosslinkers for Polysaccharides and Proteins: Synthesis Conditions, Mechanisms, and Crosslinking Efficiency, a Review. Int. J. Biol. Macromol. 2022, 202, 558–596. [Google Scholar] [CrossRef]
  74. Skopinska-Wisniewska, J.; Tuszynska, M.; Kaźmierski, Ł.; Bartniak, M.; Bajek, A. Gelatin-Sodium Alginate Hydrogels Cross-Linked by Squaric Acid and Dialdehyde Starch as a Potential Bio-Ink. Polymers 2024, 16, 2560. [Google Scholar] [CrossRef]
  75. Coleman, R.J.; Lawrie, G.; Lambert, L.K.; Whittaker, M.; Jack, K.S.; Grøndahl, L. Phosphorylation of Alginate: Synthesis, Characterization, and Evaluation of in Vitro Mineralization Capacity. Biomacromolecules 2011, 12, 889–897. [Google Scholar] [CrossRef]
  76. Chen, C.; Wang, X.; Chen, W.; Wang, L. Encapsulation of Caffeic Acid into Sodium Caseinate Using pH-Driven Method: Fabrication, Characterization, and Bioavailability. Food Bioprocess Technol. 2024, 17, 544–553. [Google Scholar] [CrossRef]
  77. Wang, X.; Liu, C.; Liu, C.; Shi, Z.; Huang, F. Development of Alginate Macroporous Hydrogels Using Sacrificial CaCO3 Particles for Enhanced Hemostasis. Int. J. Biol. Macromol. 2024, 259, 129141. [Google Scholar] [CrossRef] [PubMed]
  78. Scott, S.; Villiou, M.; Colombo, F.; La Cruz-García, A.D.; Tydecks, L.; Toelke, L.; Siemsen, K.; Selhuber-Unkel, C. Dynamic and Reversible Tuning of Hydrogel Viscoelasticity by Transient Polymer Interactions for Controlling Cell Adhesion. Adv. Mater. 2025, 37, 2408616. [Google Scholar] [CrossRef]
  79. Wong, T.; Brault, L.; Gasparotto, E.; Vallée, R.; Morvan, P.-Y.; Ferrières, V.; Nugier-Chauvin, C. Formation of Amphiphilic Molecules from the Most Common Marine Polysaccharides, toward a Sustainable Alternative? Molecules 2021, 26, 4445. [Google Scholar] [CrossRef] [PubMed]
  80. Heydari, A.; Borazjani, N.; Kazemi-Aghdam, F.; Filo, J.; Lacík, I. DMTMM-Mediated Amidation of Sodium Alginate in Aqueous Solutions: pH-Dependent Efficiency of Conjugation. Carbohydr. Polym. 2025, 348, 122893. [Google Scholar] [CrossRef] [PubMed]
  81. Labre, F.; Mathieu, S.; Chaud, P.; Morvan, P.-Y.; Vallée, R.; Helbert, W.; Fort, S. DMTMM-Mediated Amidation of Alginate Oligosaccharides Aimed at Modulating Their Interaction with Proteins. Carbohydr. Polym. 2018, 184, 427–434. [Google Scholar] [CrossRef] [PubMed]
  82. Chen, X.; Zhu, Q.; Wen, Y.; Li, Z.; Cao, S.; Yan, H.; Lin, Q. Chemical Modification of Alginate via the Oxidation-Reductive Amination Reaction for the Development of Alginate Derivative Electrospun Composite Nanofibers. J. Drug Deliv. Sci. Technol. 2022, 68, 103113. [Google Scholar] [CrossRef]
  83. Alenezi, H.; Gad, E.S.; Albassami, N.A.; Alatawi, I.S.; Alshareef, S.A.; Aljowni, M.A.; Jame, R.; Abdelaziz, M.A.; El-din, A.S.B.; Saleh, A.K. Development of Oxidized Sodium Alginate/Silica Hybrid as Efficient Adsorbent for Anionic and Cationic Dyes: Mechanism and Thermodynamic Studies. Biomass Convers. Biorefinery 2025, 15, 18247–18261. [Google Scholar] [CrossRef]
  84. Qu, W.; Zhang, X.; Han, X.; Wang, Z.; He, R.; Ma, H. Structure and Functional Characteristics of Rapeseed Protein Isolate-Dextran Conjugates. Food Hydrocoll. 2018, 82, 329–337. [Google Scholar] [CrossRef]
  85. Kumar, B.; Singh, N.; Kumar, P. A Review on Sources, Modification Techniques, Properties and Potential Applications of Alginate-Based Modified Polymers. Eur. Polym. J. 2024, 213, 113078. [Google Scholar] [CrossRef]
  86. Sand, A.; Yadav, M.; Mishra, D.K.; Behari, K. Modification of Alginate by Grafting of N-Vinyl-2-Pyrrolidone and Studies of Physicochemical Properties in Terms of Swelling Capacity, Metal-Ion Uptake and Flocculation. Carbohydr. Polym. 2010, 80, 1147–1154. [Google Scholar] [CrossRef]
  87. Zhang, H.; Zhou, L.; Shehzad, H.; Farooqi, Z.H.; Sharif, A.; Ahmed, E.; Habiba, U.; Qaisar, F.; Fatima, N.-E.; Begum, R.; et al. Innovative Free Radical Induced Synthesis of WO3-Doped Diethyl Malonate Grafted Chitosan Encapsulated with Phosphorylated Alginate Matrix for UO22+ Adsorption: Parameters Optimisation through Response Surface Methodology. Sep. Purif. Technol. 2025, 353, 128455. [Google Scholar] [CrossRef]
  88. Darwesh, H.; Mohamed, R.R.; Soliman, S.M.A. Synthesis of Grafted Copolymer Alginate-g-Poly(1-Carboxylic 4-Acrylamidobenzenesulfonamide) and Its Application in Water Treatment. Desalination Water Treat. 2022, 252, 210–218. [Google Scholar] [CrossRef]
  89. Shehzad, H.; Ahmed, E.; Sharif, A.; Din, M.I.; Farooqi, Z.H.; Nawaz, I.; Bano, R.; Iftikhar, M. Amino-Carbamate Moiety Grafted Calcium Alginate Hydrogel Beads for Effective Biosorption of Ag(I) from Aqueous Solution: Economically-Competitive Recovery. Int. J. Biol. Macromol. 2020, 144, 362–372. [Google Scholar] [CrossRef]
  90. Şanlı, O.; Olukman, M. Preparation of Ferric Ion Crosslinked Acrylamide Grafted Poly (Vinyl Alcohol)/Sodium Alginate Microspheres and Application in Controlled Release of Anticancer Drug 5-Fluorouracil. Drug Deliv. 2014, 21, 213–220. [Google Scholar] [CrossRef]
  91. Place, E.S.; Rojo, L.; Gentleman, E.; Sardinha, J.P.; Stevens, M.M. Strontium- and Zinc-Alginate Hydrogels for Bone Tissue Engineering. Tissue Eng. Part A 2011, 17, 2713–2722. [Google Scholar] [CrossRef] [PubMed]
  92. Bulut, E.; Şanlı, O. Novel Ionically Crosslinked Acrylamide-Grafted Poly(Vinyl Alcohol)/Sodium Alginate/Sodium Carboxymethyl Cellulose pH-Sensitive Microspheres for Delivery of Alzheimer’s Drug Donepezil Hydrochloride: Preparation and Optimization of Release Conditions. Artif. Cells Nanomedicine Biotechnol. 2016, 44, 431–442. [Google Scholar] [CrossRef] [PubMed]
  93. Zhou, W.; Zhang, H.; Liu, Y.; Zou, X.; Shi, J.; Zhao, Y.; Ye, Y.; Yu, Y.; Guo, J. Preparation of Calcium Alginate/Polyethylene Glycol Acrylate Double Network Fiber with Excellent Properties by Dynamic Molding Method. Carbohydr. Polym. 2019, 226, 115277. [Google Scholar] [CrossRef]
  94. Zhou, W.; Zhang, H.; Liu, Y.; Zou, X.; Shi, J.; Zhao, Y.; Ye, Y.; Yu, Y.; Guo, J. Sodium Alginate-Polyethylene Glycol Diacrylate Based Double Network Fiber: Rheological Properties of Fiber Forming Solution with Semi-Interpenetrating Network Structure. Int. J. Biol. Macromol. 2020, 142, 535–544. [Google Scholar] [CrossRef]
  95. Wang, N.; Yu, K.-K.; Li, K.; Yu, X.-Q. A Biocompatible Polyethylene Glycol/Alginate Composite Hydrogel with Significant Reactive Oxygen Species Consumption for Promoting Wound Healing. J. Mater. Chem. B 2023, 11, 6934–6942. [Google Scholar] [CrossRef]
  96. Li, Y.; Xu, T.; Tu, Z.; Dai, W.; Xue, Y.; Tang, C.; Gao, W.; Mao, C.; Lei, B.; Lin, C. Bioactive Antibacterial Silica-Based Nanocomposites Hydrogel Scaffolds with High Angiogenesis for Promoting Diabetic Wound Healing and Skin Repair. Theranostics 2020, 10, 4929–4943. [Google Scholar] [CrossRef]
  97. Saberian, M.; Safari Roudsari, R.; Haghshenas, N.; Rousta, A.; Alizadeh, S. How the Combination of Alginate and Chitosan Can Fabricate a Hydrogel with Favorable Properties for Wound Healing. Heliyon 2024, 10, e32040. [Google Scholar] [CrossRef]
  98. Shi, X.; Xu, S.; Xu, J.; He, J. Preparation and Properties of a Multi-Crosslinked Chitosan/Sodium Alginate Composite Hydrogel. Mater. Lett. 2024, 354, 135414. [Google Scholar] [CrossRef]
  99. Guan, X.; Zhang, B.; Li, D.; Ren, J.; Zhu, Y.; Sun, Z.; Chen, Y. Semi-Unzipping of Chitosan-Sodium Alginate Polyelectrolyte Gel for Efficient Capture of Metallic Mineral Ions from Tannery Effluent. Chem. Eng. J. 2023, 452, 139532. [Google Scholar] [CrossRef]
  100. Li, H.; Shen, S.; Yu, K.; Wang, H.; Fu, J. Construction of Porous Structure-Based Carboxymethyl Chitosan/Sodium Alginate/Tea Polyphenols for Wound Dressing. Int. J. Biol. Macromol. 2023, 233, 123404. [Google Scholar] [CrossRef]
  101. Zhao, L.; Feng, Z.; Lyu, Y.; Yang, J.; Lin, L.; Bai, H.; Li, Y.; Feng, Y.; Chen, Y. Electroactive Injectable Hydrogel Based on Oxidized Sodium Alginate and Carboxymethyl Chitosan for Wound Healing. Int. J. Biol. Macromol. 2023, 230, 123231. [Google Scholar] [CrossRef]
  102. Song, Y.; Li, S.; Chen, H.; Han, X.; Duns, G.J.; Dessie, W.; Tang, W.; Tan, Y.; Qin, Z.; Luo, X. Kaolin-Loaded Carboxymethyl Chitosan/Sodium Alginate Composite Sponges for Rapid Hemostasis. Int. J. Biol. Macromol. 2023, 233, 123532. [Google Scholar] [CrossRef]
  103. Hasan, H.A.; Mohd Saharuddin, S.N.D.; Muhamad, M.H. Unlocking the Potential of Polyvinyl Alcohol (PVA) as a Biocarrier for Enhanced Wastewater Treatment: A Comprehensive Review. J. Water Process Eng. 2025, 74, 107780. [Google Scholar] [CrossRef]
  104. Xiang, X.; Yi, X.; Zheng, W.; Li, Y.; Zhang, C.; Wang, X.; Chen, Z.; Huang, M.; Ying, G.-G. Enhanced Biodegradation of Thiamethoxam with a Novel Polyvinyl Alcohol (PVA)/Sodium Alginate (SA)/Biochar Immobilized Chryseobacterium Sp H5. J. Hazard. Mater. 2023, 443, 130247. [Google Scholar] [CrossRef] [PubMed]
  105. Xie, L.; Zhang, Z.; He, Y. Antibacterial Effect of Polyvinyl Alcohol/Biochar–Nano Silver/Sodium Alginate Gel Beads. Processes 2023, 11, 2330. [Google Scholar] [CrossRef]
  106. Tang, L.; Wu, P.; Zhuang, H.; Qin, Z.; Yu, P.; Fu, K.; Qiu, P.; Liu, Y.; Zhou, Y. Nitric Oxide Releasing Polyvinyl Alcohol and Sodium Alginate Hydrogels as Antibacterial and Conductive Strain Sensors. Int. J. Biol. Macromol. 2023, 241, 124564. [Google Scholar] [CrossRef] [PubMed]
  107. Kamel, S.; Dacrory, S.; Hesemann, P.; Bettache, N.; Ali, L.M.A.; Postel, L.; Akl, E.M.; El-Sakhawy, M. Wound Dressings Based on Sodium Alginate-Polyvinyl Alcohol-Moringa Oleifera Extracts. Pharmaceutics 2023, 15, 1270. [Google Scholar] [CrossRef]
  108. Nawaz, M.; Shakoor, R.A.; Al-Qahtani, N.; Bhadra, J.; Al-Thani, N.J.; Kahraman, R. Polyolefin-Based Smart Self-Healing Composite Coatings Modified with Calcium Carbonate and Sodium Alginate. Polymers 2024, 16, 636. [Google Scholar] [CrossRef]
  109. Wang, Y.; Liu, Y.; Yang, H.; Fu, Y.; Huan, L.; Zhu, F.; Wang, D.; Liu, C.; Han, D. Thermal Responsive Sodium Alginate/Polyacrylamide/Poly (N-Isopropylacrylamide) Ionic Hydrogel Composite via Seeding Calcium Carbonate Microparticles for the Engineering of Ultrasensitive Wearable Sensors. Int. J. Biol. Macromol. 2024, 280, 135909. [Google Scholar] [CrossRef]
  110. Fu, J.; Yap, J.X.; Leo, C.P.; Chang, C.K. Carboxymethyl Cellulose/Sodium Alginate Beads Incorporated with Calcium Carbonate Nanoparticles and Bentonite for Phosphate Recovery. Int. J. Biol. Macromol. 2023, 234, 123642. [Google Scholar] [CrossRef] [PubMed]
  111. Chen, L.-L.; Quan, F.-Y.; Kong, Q.-S. Preparation and Characterization of Interpenetrating Networks of Sodium Alginate-Silicon Dioxide. Polym. Mater. Sci. Eng. 2009, 25, 152–154. Available online: https://www.researchgate.net/publication/288973139_Preparation_and_characterization_of_interpenetrating_networks_of_sodium_alginate-silicon_dioxide (accessed on 9 March 2025). [CrossRef]
  112. Li, Y.; Zhang, Y.; Chai, Z.; Cui, L.; Li, C.; Ma, K.; Hu, X.; Feng, J. Entrapment of an ACE Inhibitory Peptide into Ferritin Nanoparticles Coated with Sodium Deoxycholate: Improved Chemical Stability and Intestinal Absorption. LWT Food Sci. Technol. 2021, 147, 111547. [Google Scholar] [CrossRef]
  113. Zhou, C.; Zhao, T.; Chen, L.; Yagoub, A.E.A.; Chen, H.; Yu, X. Effect of Dialysate Type on Ultrasound-Assisted Self-Assembly Zein Nanocomplexes: Fabrication, Characterization, and Physicochemical Stability. Food Res. Int. 2022, 162, 111812. [Google Scholar] [CrossRef]
  114. Dai, J.; Bai, M.; Li, C.; Cui, H.; Lin, L. The Improvement of Sodium Dodecyl Sulfate on the Electrospinning of Gelatin O/W Emulsions for Production of Core-Shell Nanofibers. Food Hydrocoll. 2023, 145, 109092. [Google Scholar] [CrossRef]
  115. Wang, Y.-Y.; Qiu, W.-Y.; Sun, L.; Ding, Z.-C.; Yan, J.-K. Preparation, Characterization, and Antioxidant Capacities of Selenium Nanoparticles Stabilized Using Polysaccharide–Protein Complexes from Corbicula Fluminea. Food Biosci. 2018, 26, 177–184. [Google Scholar] [CrossRef]
  116. Gan, C.; Liu, Q.; Zhang, Y.; Shi, T.; He, W.-S.; Jia, C. A Novel Phytosterols Delivery System Based on Sodium Caseinate-Pectin Soluble Complexes: Improving Stability and Bioaccessibility. Food Hydrocoll. 2022, 124, 107295. [Google Scholar] [CrossRef]
  117. Mu, R.; Hong, X.; Ni, Y.; Li, Y.; Pang, J.; Wang, Q.; Xiao, J.; Zheng, Y. Recent Trends and Applications of Cellulose Nanocrystals in Food Industry. Trends Food Sci. Technol. 2019, 93, 136–144. [Google Scholar] [CrossRef]
  118. Liu, L.; Li, C.; Wang, Z.; Wang, X.Z. Structurally Related Electromagnetic Properties of NixFeyO4 Nanomaterials Synthesized by Granulated Sodium Alginate. J. Alloys Compd. 2021, 858, 157641. [Google Scholar] [CrossRef]
  119. Su, R.; Ge, S.; Li, H.; Su, Y.; Li, Q.; Zhou, W.; Gao, B.; Yue, Q. Synchronous Synthesis of Cu2O/Cu/rGO@carbon Nanomaterials Photocatalysts via the Sodium Alginate Hydrogel Template Method for Visible Light Photocatalytic Degradation. Sci. Total Environ. 2019, 693, 133657. [Google Scholar] [CrossRef]
  120. Dawar, A.; Said, N.M.; Islam, S.; Shah, Z.; Mahmuod, S.R.; Wakif, A. A Semi-Analytical Passive Strategy to Examine a Magnetized Heterogeneous Mixture Having Sodium Alginate Liquid with Alumina and Copper Nanomaterials near a Convectively Heated Surface of a Stretching Curved Geometry. Int. Commun. Heat Mass Transf. 2022, 139, 106452. [Google Scholar] [CrossRef]
  121. Li, N.; Chen, F.; Cui, F.; Sun, W.; Zhang, J.; Qian, L.; Yang, Y.; Wu, D.; Dong, Y.; Jiang, J.; et al. Improved Postharvest Quality and Respiratory Activity of Straw Mushroom (Volvariella volvacea) with Ultrasound Treatment and Controlled Relative Humidity. Sci. Hortic. 2017, 225, 56–64. [Google Scholar] [CrossRef]
  122. Huang, L.; Ding, X.; Dai, C.; Ma, H. Changes in the Structure and Dissociation of Soybean Protein Isolate Induced by Ultrasound-Assisted Acid Pretreatment. Food Chem. 2017, 232, 727–732. [Google Scholar] [CrossRef] [PubMed]
  123. Guo, L.; Hong, C.; Wang, W.; Zhang, X.; Chen, J.; Chen, Z.; Ashokkumar, M.; Ma, H. Evaluation of Low-Temperature Ultrasonic Marination of Pork Meat at Various Frequencies on Physicochemical Properties, Myoglobin Levels, and Volatile Compounds. Meat Sci. 2024, 217, 109606. [Google Scholar] [CrossRef]
  124. Guo, L.; Zhang, X.; Guo, Y.; Chen, Z.; Ma, H. Evaluation of Ultrasonic-Assisted Pickling with Different Frequencies on NaCl Transport, Impedance Properties, and Microstructure in Pork. Food Chem. 2024, 430, 137003. [Google Scholar] [CrossRef]
  125. Chen, H.; Xu, B.; Zhou, C.; Yagoub, A.E.-G.A.; Cai, Z.; Yu, X. Multi-Frequency Ultrasound-Assisted Dialysis Modulates the Self-Assembly of Alcohol-Free Zein-Sodium Caseinate to Encapsulate Curcumin and Fabricate Composite Nanoparticles. Food Hydrocoll. 2022, 122, 107110. [Google Scholar] [CrossRef]
  126. Alenyorege, E.A.; Ma, H.; Ayim, I.; Aheto, J.H.; Hong, C.; Zhou, C. Reduction of Listeria Innocua in Fresh-Cut Chinese Cabbage by a Combined Washing Treatment of Sweeping Frequency Ultrasound and Sodium Hypochlorite. LWT Food Sci. Technol. 2019, 101, 410–418. [Google Scholar] [CrossRef]
  127. Virk, M.S.; Virk, M.A.; Liang, Q.; Sun, Y.; Zhong, M.; Tufail, T.; Rashid, A.; Qayum, A.; Rehman, A.; Ekumah, J.-N.; et al. Enhancing Storage and Gastroprotective Viability of Lactiplantibacillus plantarum Encapsulated by Sodium Caseinate-Inulin-Soy Protein Isolates Composites Carried within Carboxymethyl Cellulose Hydrogel. Food Res. Int. 2024, 187, 114432. [Google Scholar] [CrossRef]
  128. Musa, A.; Ma, H.; Gasmalla, M.A.A.; Sarpong, F.; Awad, F.N.; Duan, Y. Effect of Multi-Frequency Counter-Current S Type Ultrasound Pretreatment on the Enzymatic Hydrolysis of Defatted Corn Germ Protein: Kinetics and Thermodynamics. Process Biochem. 2019, 87, 112–118. [Google Scholar] [CrossRef]
  129. Li, S.; Yang, X.; Zhang, Y.; Ma, H.; Liang, Q.; Qu, W.; He, R.; Zhou, C.; Mahunu, G.K. Effects of Ultrasound and Ultrasound Assisted Alkaline Pretreatments on the Enzymolysis and Structural Characteristics of Rice Protein. Ultrason. Sonochem. 2016, 31, 20–28. [Google Scholar] [CrossRef] [PubMed]
  130. Oladejo, A.O.; Ma, H.; Qu, W.; Zhou, C.; Wu, B.; Yang, X.; Onwude, D.I. Effects of Ultrasound Pretreatments on the Kinetics of Moisture Loss and Oil Uptake during Deep Fat Frying of Sweet Potato (Ipomea batatas). Innov. Food Sci. Emerg. Technol. 2017, 43, 7–17. [Google Scholar] [CrossRef]
  131. Liu, Y.; Liang, Q.; Liu, Y.; Rashid, A.; Qayum, A.; Tuly, J.A.; Ma, H.; Miao, S.; Ren, X. Sodium Caseinate/Pectin Complex-Stabilized Emulsion: Multi-Frequency Ultrasound Regulation, Characterization and Its Application in Quercetin Delivery. Food Hydrocoll. 2024, 156, 110316. [Google Scholar] [CrossRef]
  132. Guo, J.; Yu, X.; Zhou, C.; Wang, B.; Zhang, L.; Otu, P.; Chen, L.; Niu, Y.; Yao, D.; Hua, C.; et al. Preparation of Umami Peptides from Chicken Breast by Ultrasound-Assisted Gradient Dilution Feeding Substrate and Study of Their Formation Mechanism. Food Biosci. 2024, 62, 105176. [Google Scholar] [CrossRef]
  133. Mustapha, A.T.; Zhou, C.; Amanor-Atiemoh, R.; Ali, T.A.A.; Wahia, H.; Ma, H.; Sun, Y. Efficacy of Dual-Frequency Ultrasound and Sanitizers Washing Treatments on Quality Retention of Cherry Tomato. Innov. Food Sci. Emerg. Technol. 2020, 62, 102348. [Google Scholar] [CrossRef]
  134. Luo, F.; Zhang, Z.; Lu, F.; Li, D.; Zhou, C.; Niu, L.; Xu, Y.; Feng, L.; Dai, Z.; He, W. Ultrasound Modification of Pectin and the Mechanism of Its Interaction with Cyanidin-3-O-Glucoside. Food Hydrocoll. 2024, 152, 109898. [Google Scholar] [CrossRef]
  135. Karim, A.; Rehman, A.; Jafari, S.M.; Miao, S.; Dabbour, M.; Ashraf, W.; Rasheed, H.A.; Assadpour, E.; Hussain, A.; Suleria, H.A.R.; et al. Fabrication and Characterization of Sonicated Peach Gum-Sodium Caseinate Nanocomplexes: Physicochemical, Spectroscopic, Morphological, and Correlation Analyses. Food Bioprocess Technol. 2025, 18, 2462–2481. [Google Scholar] [CrossRef]
  136. Yi, Y.; Song, J.; Zhou, P.; Shu, Y.; Liang, P.; Liang, H.; Liu, Y.; Yuan, X.; Shan, X.; Wu, X. An Ultrasound-Triggered Injectable Sodium Alginate Scaffold Loaded with Electrospun Microspheres for on-Demand Drug Delivery to Accelerate Bone Defect Regeneration. Carbohydr. Polym. 2024, 334, 122039. [Google Scholar] [CrossRef]
  137. Feng, L.; Cao, Y.; Xu, D.; You, S.; Han, F. Influence of Sodium Alginate Pretreated by Ultrasound on Papain Properties: Activity, Structure, Conformation and Molecular Weight and Distribution. Ultrason. Sonochem. 2016, 32, 224–230. [Google Scholar] [CrossRef]
  138. Manaila, E.; Craciun, G.; Calina, I.C. Sodium Alginate-g-Acrylamide/Acrylic Acid Hydrogels Obtained by Electron Beam Irradiation for Soil Conditioning. Int. J. Mol. Sci. 2023, 24, 104. [Google Scholar] [CrossRef]
  139. Hua, S.; Wang, A. Synthesis, Characterization and Swelling Behaviors of Sodium Alginate-g-Poly(Acrylic Acid)/Sodium Humate Superabsorbent. Carbohydr. Polym. 2009, 75, 79–84. [Google Scholar] [CrossRef]
  140. Shou-kui, S. Effect of Sodium Alginate Coating and ~(60)Co-γ Irradiation Treatment on Fresh-Keeping of Golden Silk Jujube. Sci. Technol. Food Ind. 2012, 33, 213–216. [Google Scholar]
  141. Chen, P.; Cheng, H.; Tian, J.; Pan, H.; Chen, S.; Ye, X.; Chen, J. Photo-Crosslinking Modified Sodium Alginate Hydrogel for Targeting Delivery Potential by NO Response. Int. J. Biol. Macromol. 2023, 253, 126454. [Google Scholar] [CrossRef] [PubMed]
  142. Ding, J.; Zhang, H.; Wang, W.; Zhu, Y.; Wang, Q.; Wang, A. Synergistic Effect of Palygorskite Nanorods and Ion Crosslinking to Enhance Sodium Alginate-Based Hydrogels. Eur. Polym. J. 2021, 147, 110306. [Google Scholar] [CrossRef]
  143. Liu, Y.; Shen, S.; Wu, Y.; Wang, M.; Cheng, Y.; Xia, H.; Jia, R.; Liu, C.; Wang, Y.; Xia, Y.; et al. Percutaneous Electroosmosis of Berberine-Loaded Ca2+ Crosslinked Gelatin/Alginate Mixed Hydrogel. Polymers 2022, 14, 5101. [Google Scholar] [CrossRef] [PubMed]
  144. Su, C.; Li, D.; Sun, W.; Wang, L.; Wang, Y. Green, Tough, and Heat-Resistant: A GDL-Induced Strategy for Starch-Alginate Hydrogels. Food Chem. 2024, 449, 139188. [Google Scholar] [CrossRef]
  145. Wei, X.; Xiong, H.; Zhou, D.; Jing, X.; Huang, Y. Ion-Assisted Fabrication of Neutral Protein Crosslinked Sodium Alginate Nanogels. Carbohydr. Polym. 2018, 186, 45–53. [Google Scholar] [CrossRef]
  146. Gong, W.; Liu, L.; Luo, L.; Ji, L. Preparation and Characterization of a Self-Crosslinking Sodium Alginate-Bioactive Glass Sponge. J. Biomed. Mater. Res. B Appl. Biomater. 2023, 111, 173–183. [Google Scholar] [CrossRef]
  147. Abdualrahman, M.A.Y.; Ma, H.; Zhou, C.; Yagoub, A.E.A.; Hu, J.; Yang, X. Thermal and Single Frequency Counter-current Ultrasound Pretreatments of Sodium Caseinate: Enzymolysis Kinetics and Thermodynamics, Amino Acids Composition, Molecular Weight Distribution and Antioxidant Peptides. J. Sci. Food Agric. 2016, 96, 4861–4873. [Google Scholar] [CrossRef]
  148. Abdualrahman, M.A.Y.; Zhou, C.; Zhang, Y.; Yagoub, A.E.A.; Ma, H.; Mao, L.; Wang, K. Effects of Ultrasound Pretreatment on Enzymolysis of Sodium Caseinate Protein: Kinetic Study, Angiotensin-converting Enzyme Inhibitory Activity, and the Structural Characteristics of the Hydrolysates. J. Food Process. Preserv. 2017, 41, e13276. [Google Scholar] [CrossRef]
  149. Noach, M.; Mampana, R.; Van Rensburg, E.; Goosen, N.; Pott, R. Chemical and Enzymatic Hydrolysis of Alginate: A Review. Bot. Mar. 2024, 67, 487–511. [Google Scholar] [CrossRef]
  150. Xu, F.; Cha, Q.-Q.; Zhang, Y.-Z.; Chen, X.-L. Degradation and Utilization of Alginate by Marine Pseudoalteromonas: A Review. Appl. Environ. Microbiol. 2021, 87, e0036821. [Google Scholar] [CrossRef] [PubMed]
  151. Cantarel, B.L.; Coutinho, P.M.; Rancurel, C.; Bernard, T.; Lombard, V.; Henrissat, B. The Carbohydrate-Active EnZymes Database (CAZy): An Expert Resource for Glycogenomics. Nucleic Acids Res. 2009, 37, D233–238. [Google Scholar] [CrossRef] [PubMed]
  152. Kim, H.T.; Ko, H.-J.; Kim, N.; Kim, D.; Lee, D.; Choi, I.-G.; Woo, H.C.; Kim, M.D.; Kim, K.H. Characterization of a Recombinant Endo-Type Alginate Lyase (Alg7D) from Saccharophagus Degradans. Biotechnol. Lett. 2012, 34, 1087–1092. [Google Scholar] [CrossRef]
  153. Yang, Q.; Yin, D.; Zhang, X.; Solairaj, D.; Xi, Y.; Chen, H.; Li, Y.; Zhang, H. Alginate Oligosaccharide-Driven Resistance in Debaryomyces Hansenii Y3: A Dual Omics Perspective. New Zealand J. Crop Hortic. Sci. 2025, 53, 563–586. [Google Scholar] [CrossRef]
  154. Ochiai, A.; Yamasaki, M.; Mikami, B.; Hashimoto, W.; Murata, K. Crystal Structure of Exotype Alginate Lyase Atu3025 from Agrobacterium tumefaciens*. J. Biol. Chem. 2010, 285, 24519–24528. [Google Scholar] [CrossRef]
  155. Xu, F.; Wang, P.; Zhang, Y.-Z.; Chen, X.-L. Diversity of Three-Dimensional Structures and Catalytic Mechanisms of Alginate Lyases. Appl. Environ. Microbiol. 2018, 84, e02040-17. [Google Scholar] [CrossRef]
  156. Song, M.; Chen, L.; Dong, C.; Tang, M.; Wei, Y.; Lv, D.; Li, Q.; Chen, Z. Alginate Oligosaccharide and Gut Microbiota: Exploring the Key to Health. Nutrients 2025, 17, 1977. [Google Scholar] [CrossRef]
  157. Kaur, S.; Abraham, R.E.; Franco, C.M.M.; Puri, M. Production of Alginate Oligosaccharides (AOSs) Using Enhanced Physicochemical Properties of Immobilized Alginate Lyase for Industrial Application. Mar. Drugs 2024, 22, 120. [Google Scholar] [CrossRef]
  158. Aheto, J.H.; Huang, X.; Xiaoyu, T.; Bonah, E.; Ren, Y.; Alenyorege, E.A.; Chunxia, D. Investigation into Crystal Size Effect on Sodium Chloride Uptake and Water Activity of Pork Meat Using Hyperspectral Imaging. J. Food Process. Preserv. 2019, 43, e14197. [Google Scholar] [CrossRef]
  159. Ding, F.; Fu, L.; Huang, X.; Shi, J.; Povey, M.; Zou, X. Self-Healing Carboxymethyl Chitosan Hydrogel with Anthocyanin for Monitoring the Spoilage of Flesh Foods. Food Hydrocoll. 2025, 165, 111270. [Google Scholar] [CrossRef]
  160. Cui, H.; Cheng, Q.; Li, C.; Khin, M.N.; Lin, L. Schiff Base Cross-Linked Dialdehyde β-Cyclodextrin/Gelatin-Carrageenan Active Packaging Film for the Application of Carvacrol on Ready-to-Eat Foods. Food Hydrocoll. 2023, 141, 108744. [Google Scholar] [CrossRef]
  161. Cui, H.; Yang, X.; Li, C.; Ye, Y.; Chen, X.; Lin, L. Enhancing Anti-E. Coli O157:H7 Activity of Composite Phage Nanofiber Film by D-Phenylalanine for Food Packaging. Int. J. Food Microbiol. 2022, 376, 109762. [Google Scholar] [CrossRef]
  162. Zhang, J.; Zhang, J.; Huang, X.; Zhai, X.; Li, Z.; Shi, J.; Sobhy, R.; Khalifa, I.; Zou, X. Lemon-Derived Carbon Quantum Dots Incorporated Guar Gum/Sodium Alginate Films with Enhanced the Preservability for Blanched Asparagus Active Packaging. Food Res. Int. 2025, 202, 115736. [Google Scholar] [CrossRef] [PubMed]
  163. Khan, A.A.; Cui, F.-J.; Ullah, M.W.; Qayum, A.; Khalifa, I.; Bacha, S.A.S.; Ying, Z.-Z.; Khan, I.; Zeb, U.; Alarfaj, A.A.; et al. Fabrication and Characterization of Bioactive Curdlan and Sodium Alginate Films for Enhancing the Shelf Life of Volvariella volvacea. Food Biosci. 2024, 62, 105137. [Google Scholar] [CrossRef]
  164. Li, H.; Liu, C.; Sun, J.; Lv, S. Bioactive Edible Sodium Alginate Films Incorporated with Tannic Acid as Antimicrobial and Antioxidative Food Packaging. Foods 2022, 11, 3044. [Google Scholar] [CrossRef]
  165. Ding, F.; Wu, R.; Huang, X.; Shi, J.; Zou, X. Anthocyanin Loaded Composite Gelatin Films Crosslinked with Oxidized Alginate for Monitoring Spoilage of Flesh Foods. Food Packag. Shelf Life 2024, 42, 101255. [Google Scholar] [CrossRef]
  166. Zhang, J.; Zhang, J.; Guan, Y.; Huang, X.; Arslan, M.; Shi, J.; Li, Z.; Gong, Y.; Holmes, M.; Zou, X. High- Sensitivity Bilayer Nanofiber Film Based on Polyvinyl Alcohol/Sodium Alginate/Polyvinylidene Fluoride for Pork Spoilage Visual Monitoring and Preservation. Food Chem. 2022, 394, 133439. [Google Scholar] [CrossRef]
  167. Tong, W.Y.; Ahmad Rafiee, A.R.; Leong, C.R.; Tan, W.-N.; Dailin, D.J.; Almarhoon, Z.M.; Shelkh, M.; Nawaz, A.; Chuah, L.F. Development of Sodium Alginate-Pectin Biodegradable Active Food Packaging Film Containing Cinnamic Acid. Chemosphere 2023, 336, 139212. [Google Scholar] [CrossRef]
  168. Zhang, Z.; Zhang, Y.; Wang, C.; Liu, X.; El-Seedi, H.R.; Gómez, P.L.; Al-Zamora, S.M.; Zou, X.; Guo, Z. Enhanced Composite Co-MOF-Derived Sodium Carboxymethyl Cellulose Visual Films for Real-Time and in Situ Monitoring Fresh-Cut Apple Freshness. Food Hydrocoll. 2024, 157, 110475. [Google Scholar] [CrossRef]
  169. An, N.; Zhou, W. Sodium Alginate/Ager Colourimetric Film on Porous Substrate Layer: Potential in Intelligent Food Packaging. Food Chem. 2024, 445, 138790. [Google Scholar] [CrossRef] [PubMed]
  170. Huang, X.; Zhao, W.; Li, Z.; Zhang, N.; Wang, S.; Shi, J.; Zhai, X.; Zhang, J.; Shen, T. Preparation of a Dual-Functional Active Film Based on Bilayer Hydrogel and Red Cabbage Anthocyanin for Maintaining and Monitoring Pork Freshness. Foods 2023, 12, 4520. [Google Scholar] [CrossRef] [PubMed]
  171. Zhang, J.; Huang, X.; Shi, J.; Liu, L.; Zhang, X.; Zou, X.; Xiao, J.; Zhai, X.; Zhang, D.; Li, Y.; et al. A Visual Bi-Layer Indicator Based on Roselle Anthocyanins with High Hydrophobic Property for Monitoring Griskin Freshness. Food Chem. 2021, 355, 129573. [Google Scholar] [CrossRef] [PubMed]
  172. Rashid, A.; Qayum, A.; Shah Bacha, S.A.; Liang, Q.; Liu, Y.; Kang, L.; Chi, Z.; Chi, R.; Han, X.; Ekumah, J.-N.; et al. Novel Pullulan-Sodium Alginate Film Incorporated with Anthocyanin-Loaded Casein-Carboxy Methyl Cellulose Nanocomplex for Real-Time Fish and Shrimp Freshness Monitoring. Food Hydrocoll. 2024, 156, 110356. [Google Scholar] [CrossRef]
  173. Zhang, Z.; Wang, C.; Jayan, H.; Gao, M.; Hesham, R.; El-Seedi, H.R.; Zou, X.; Guo, Z. Novel pH-Sensitive Organic Ligand-Based Luminescent MOFs Modified CMC-Na/SA Films for Real-Time Monitoring of Fruit Freshness. Food Packag. Shelf Life 2025, 49, 101521. [Google Scholar] [CrossRef]
  174. Lin, L.; Agyemang, K.; Abdel-Samie, M.A.-S.; Cui, H. Antibacterial Mechanism of Tetrapleura Tetraptera Extract against Escherichia Coli and Staphylococcus Aureus and Its Application in Pork. J. Food Saf. 2019, 39, e12693. [Google Scholar] [CrossRef]
  175. Zhang, J.; Zhang, J.; Zhang, X.; Huang, X.; Shi, J.; Sobhy, R.; Khalifa, I.; Zou, X. Ammonia-Responsive Colorimetric Film of Phytochemical Formulation (Alizarin) Grafted onto ZIF-8 Carrier with Poly(Vinyl Alcohol) and Sodium Alginate for Beef Freshness Monitoring. J. Agric. Food Chem. 2024, 72, 11706–11715. [Google Scholar] [CrossRef]
  176. Herrera-Balandrano, D.D.; Chai, Z.; Li, C.; Zhao, X.; Zhao, X.; Li, B.; Yang, Y.; Huang, W. Gastrointestinal Fate of Blueberry Anthocyanins in Ferritin-Based Nanocarriers. Food Res. Int. 2024, 176, 113811. [Google Scholar] [CrossRef]
  177. Ruan, P.; Zhang, K.; Zhang, W.; Kong, Y.; Zhou, Y.; Yao, B.; Wang, Y.; Wang, Z. Polyphenolic Truxillic Acid Crosslinked Sodium Alginate Film with Notable Antimicrobial and Biodegradable Properties for Food Packaging. Int. J. Biol. Macromol. 2024, 279, 135184. [Google Scholar] [CrossRef]
  178. Shan, P.; Wang, K.; Yu, F.; Yi, L.; Sun, L.; Li, H. Gelatin/Sodium Alginate Multilayer Composite Film Crosslinked with Green Tea Extract for Active Food Packaging Application. Colloids Surf. Physicochem. Eng. Asp. 2023, 662, 131013. [Google Scholar] [CrossRef]
  179. Lin, L.; Mei, C.; Shi, C.; Li, C.; Abdel-Samie, M.A.; Cui, H. Preparation and Characterization of Gelatin Active Packaging Film Loaded with Eugenol Nanoparticles and Its Application in Chicken Preservation. Food Biosci. 2023, 53, 102778. [Google Scholar] [CrossRef]
  180. Surendhiran, D.; Cui, H.; Lin, L. Encapsulation of Phlorotannin in Alginate/PEO Blended Nanofibers to Preserve Chicken Meat from Salmonella Contaminations. Food Packag. Shelf Life 2019, 21, 100346. [Google Scholar] [CrossRef]
  181. Rashid, A.; Qayum, A.; Bacha, S.A.S.; Liang, Q.; Liu, Y.; Kang, L.; Chi, Z.; Chi, R.; Han, X.; Ekumah, J.-N.; et al. Preparation and Functional Characterization of Pullulan-Sodium Alginate Composite Film Enhanced with Ultrasound-Assisted Clove Essential Oil Nanoemulsions for Effective Preservation of Cherries and Mushrooms. Food Chem. 2024, 457, 140048. [Google Scholar] [CrossRef]
  182. Gao, L.; Li, Z.; Wei, X.; Hao, M.; Song, W.; Zou, X.; Huang, X. A Cell-Based Electrochemical Biosensor for the Detection of Capsaicin. J. Food Meas. Charact. 2024, 18, 9341–9352. [Google Scholar] [CrossRef]
  183. Huang, X.; Du, L.; Li, Z.; Xue, J.; Shi, J.; Tahir, H.E.; Zhai, X.; Zhang, J.; Zhang, N.; Sun, W.; et al. A Visual Bi-Layer Indicator Based on Mulberry Anthocyanins with High Stability for Monitoring Chinese Mitten Crab Freshness. Food Chem. 2023, 411, 135497. [Google Scholar] [CrossRef] [PubMed]
  184. Yang, Z.; Li, M.; Li, Y.; Huang, X.; Li, Z.; Zhai, X.; Shi, J.; Zou, X.; Xiao, J.; Sun, Y.; et al. Sodium Alginate/Guar Gum Based Nanocomposite Film Incorporating β-Cyclodextrin/Persimmon Pectin-Stabilized Baobab Seed Oil Pickering Emulsion for Mushroom Preservation. Food Chem. 2024, 437, 137891. [Google Scholar] [CrossRef] [PubMed]
  185. Chen, C.; Chen, Z.; Zhong, Q. Caseinate Nanoparticles Co-Loaded with Quercetin and Avenanthramide 2c Using a Novel Two-Step pH-Driven Method: Formation, Characterization, and Bioavailability. Food Hydrocoll. 2022, 129, 107669. [Google Scholar] [CrossRef]
  186. Yang, Z.; Li, M.; Li, Z.; Li, Y.; Shi, J.; Huang, X.; Sun, Y.; Zhai, X.; Zou, X.; Xiao, J. Incorporation of Hawthorn Pectin/β-Cyclodextrin-Stabilized Pickering Emulsion and Titanium Dioxide Nanoparticles for Improving the Physical, Biological, and Release Properties of Guar Gum/Agar/Sodium Alginate-Based Bilayer Films. Ind. Crops Prod. 2024, 212, 118302. [Google Scholar] [CrossRef]
  187. Xu, H.; Pan, J.; Ma, C.; Mintah, B.K.; Dabbour, M.; Huang, L.; Dai, C.; Ma, H.; He, R. Stereo-Hindrance Effect and Oxidation Cross-Linking Induced by Ultrasound-Assisted Sodium Alginate-Glycation Inhibit Lysinoalanine Formation in Silkworm Pupa Protein. Food Chem. 2025, 463, 141284. [Google Scholar] [CrossRef]
  188. Karim, A.; Rehman, A.; Khalifa, I.; Hussain, A.; Ashraf, W.; Miao, S.; Lianfu, Z. Encapsulation of Lutein Within Ultrasonicated Peach Gum-Sodium Caseinate Complex Nanoparticles Via Electrostatic Complexation: Physiochemical Properties, Structural Interaction Mechanisms, and In Vitro Release Analyses. Food Bioprocess Technol. 2025, 18, 4392–4409. [Google Scholar] [CrossRef]
  189. Li, Y.; Wang, Z.; Xia, W.; Chai, Z.; Feng, J.; Teng, C.; Ma, K.; Hu, X.; Xu, L. Sodium Alginate Coated Ferritin as ACE Inhibitory Peptide Carrier: Prolonged Release Property and Enhanced Transepithelial Transport. Food Sci. Biotechnol. 2025, 34, 1867–1878. [Google Scholar] [CrossRef]
  190. Yerramathi, B.B.; Muniraj, B.A.; Kola, M.; Konidala, K.K.; Arthala, P.K.; Sharma, T.S.K. Alginate Biopolymeric Structures: Versatile Carriers for Bioactive Compounds in Functional Foods and Nutraceutical Formulations: A Review. Int. J. Biol. Macromol. 2023, 253, 127067. [Google Scholar] [CrossRef]
  191. Yoo, S.-H.; Song, Y.-B.; Chang, P.-S.; Lee, H.G. Microencapsulation of α-Tocopherol Using Sodium Alginate and Its Controlled Release Properties. Int. J. Biol. Macromol. 2006, 38, 25–30. [Google Scholar] [CrossRef] [PubMed]
  192. Mazza, K.E.L.; Costa, A.M.M.; da Silva, J.P.L.; Alviano, D.S.; Bizzo, H.R.; Tonon, R.V. Microencapsulation of Marjoram Essential Oil as a Food Additive Using Sodium Alginate and Whey Protein Isolate. Int. J. Biol. Macromol. 2023, 233, 123478. [Google Scholar] [CrossRef] [PubMed]
  193. Jin, Y.; Lv, H.; Wang, M.; Cho, C.-S.; Shin, J.; Cui, L.; Yan, C. Effect of Microencapsulation of Egg Yolk Immunoglobulin Y by Sodium Alginate/Chitosan/Sodium Alginate on the Growth Performance, Serum Parameters, and Intestinal Health of Broiler Chickens. Anim. Biosci. 2023, 36, 1241–1251. [Google Scholar] [CrossRef] [PubMed]
  194. de Matos, E.F.; Scopel, B.S.; Dettmer, A. Citronella Essential Oil Microencapsulation by Complex Coacervation with Leather Waste Gelatin and Sodium Alginate. J. Environ. Chem. Eng. 2018, 6, 1989–1994. [Google Scholar] [CrossRef]
  195. Liu, J.; Liu, F.; Ren, T.; Wang, J.; Yang, M.; Yao, Y.; Chen, H. Fabrication of Fish Gelatin/Sodium Alginate Double Network Gels for Encapsulation of Probiotics. J. Sci. Food Agric. 2021, 101, 4398–4408. [Google Scholar] [CrossRef]
  196. Etchepare, M.D.A.; Barin, J.S.; Cichoski, A.J.; Jacob-Lopes, E.; Wagner, R.; Fries, L.L.M.; Menezes, C.R.D. Microencapsulation of Probiotics Using Sodium Alginate. Ciênc. Rural 2015, 45, 1319–1326. [Google Scholar] [CrossRef]
  197. Man, Y.; Zhou, C.; Adhikari, B.; Wang, Y.; Xu, T.; Wang, B. High Voltage Electrohydrodynamic Atomization of Bovine Lactoferrin and Its Encapsulation Behaviors in Sodium Alginate. J. Food Eng. 2022, 317, 110842. [Google Scholar] [CrossRef]
  198. Wei, Q.; Zhou, J.; An, Y.; Li, M.; Zhang, J.; Yang, S. Modification, 3D Printing Process and Application of Sodium Alginate Based Hydrogels in Soft Tissue Engineering: A Review. Int. J. Biol. Macromol. 2023, 232, 123450. [Google Scholar] [CrossRef]
  199. Liu, N.; Zhang, X.; Guo, Q.; Wu, T.; Wang, Y. 3D Bioprinted Scaffolds for Tissue Repair and Regeneration. Front. Mater. 2022, 9, 925321. [Google Scholar] [CrossRef]
  200. Fu, S.; Du, X.; Zhu, M.; Tian, Z.; Wei, D.; Zhu, Y. 3D Printing of Layered Mesoporous Bioactive Glass/Sodium Alginate-Sodium Alginate Scaffolds with Controllable Dual-Drug Release Behaviors. Biomed. Mater. Bristol Engl. 2019, 14, 065011. [Google Scholar] [CrossRef] [PubMed]
  201. Liu, B.; Hu, C.; Huang, X.; Qin, K.; Wang, L.; Wang, Z.; Liang, J.; Xie, F.; Fan, Z. 3D Printing Nacre Powder/Sodium Alginate Scaffold Loaded with PRF Promotes Bone Tissue Repair and Regeneration. Biomater. Sci. 2024, 12, 2418–2433. [Google Scholar] [CrossRef] [PubMed]
  202. Song, Y.; Hu, Q.; Liu, S.; Wang, Y.; Zhang, H.; Chen, J.; Yao, G. Electrospinning/3D Printing Drug-Loaded Antibacterial Polycaprolactone Nanofiber/Sodium Alginate-Gelatin Hydrogel Bilayer Scaffold for Skin Wound Repair. Int. J. Biol. Macromol. 2024, 275, 129705. [Google Scholar] [CrossRef]
  203. Chaturvedi, K.; Ganguly, K.; More, U.A.; Reddy, K.R.; Dugge, T.; Naik, B.; Aminabhavi, T.M.; Noolvi, M.N. Chapter 3—Sodium Alginate in Drug Delivery and Biomedical Areas. In Natural Polysaccharides in Drug Delivery and Biomedical Applications; Hasnain, M.S., Nayak, A.K., Eds.; Academic Press: Cambridge, MA, USA, 2019; pp. 59–100. ISBN 978-0-12-817055-7. [Google Scholar]
  204. Veronica, N.; Heng, P.W.S.; Liew, C.V. Alginate-Based Matrix Tablets for Drug Delivery. Expert Opin. Drug Deliv. 2023, 20, 115–130. [Google Scholar] [CrossRef]
  205. Dubashynskaya, N.V.; Petrova, V.A.; Romanov, D.P.; Skorik, Y.A. pH-Sensitive Drug Delivery System Based on Chitin Nanowhiskers-Sodium Alginate Polyelectrolyte Complex. Molecules 2022, 15, 5860. [Google Scholar] [CrossRef]
  206. Peng, D.; Deng, D.; Lv, J.; Zhang, W.; Tian, H.; Zhang, X.; Wu, M.; Zhao, Y. A Novel Macroporous Carboxymethyl Chitosan/Sodium Alginate Sponge Dressing Capable of Rapid Hemostasis and Drug Delivery. Int. J. Biol. Macromol. 2024, 278, 134943. [Google Scholar] [CrossRef]
  207. Sreekanth Reddy, O.; Subha, M.C.S.; Jithendra, T.; Madhavi, C.; Chowdoji Rao, K. Curcumin Encapsulated Dual Cross Linked Sodium Alginate/Montmorillonite Polymeric Composite Beads for Controlled Drug Delivery. J. Pharm. Anal. 2021, 11, 191–199. [Google Scholar] [CrossRef]
  208. Lin, X.; Li, Y.; Zhang, B.; Li, J.; Ren, J.; Tang, Y.; Wu, S.; Yang, J.; Wang, Q. Alginate Nanogel-Embedded Liposomal Drug Carriers Facilitate Drug Delivery Efficiency in Arthritis Treatment. Int. J. Biol. Macromol. 2024, 273, 133065. [Google Scholar] [CrossRef]
  209. Lu, R.; Liu, Z.; Shao, Y.; Su, J.; Li, X.; Sun, F.; Zhang, Y.; Li, S.; Zhang, Y.; Cui, J.; et al. Nitric Oxide Enhances Rice Resistance to Rice Black-Streaked Dwarf Virus Infection. Rice 2020, 13, 24. [Google Scholar] [CrossRef]
  210. Sun, L.; Zhang, X.; Zhou, Y.; Peng, Z.; Cui, F.; Zhou, Q.; Man, Z.; Guo, J.; Sun, W. Can Cadmium-Contaminated Rice Be Used to Produce Food Additive Sodium Erythorbate? Food Chem. 2025, 462, 140923. [Google Scholar] [CrossRef]
  211. Chen, A.; Liang, H.; Chen, T.; Yang, W.; Ding, C. Influence of Long-Term Irrigation with Treated Papermaking Wastewater on Soil Ecosystem of a Full-Scale Managed Reed Wetland. J. Soils Sediments 2016, 16, 1352–1359. [Google Scholar] [CrossRef]
  212. Jiang, C.; Wang, X.; Hou, B.; Hao, C.; Li, X.; Wu, J. Construction of a Lignosulfonate–Lysine Hydrogel for the Adsorption of Heavy Metal Ions. J. Agric. Food Chem. 2020, 68, 3050–3060. [Google Scholar] [CrossRef]
  213. Chen, P.; Yin, L.; El-Seedi, H.R.; Zou, X.; Guo, Z. Green Reduction of Silver Nanoparticles for Cadmium Detection in Food Using Surface-Enhanced Raman Spectroscopy Coupled Multivariate Calibration. Food Chem. 2022, 394, 133481. [Google Scholar] [CrossRef]
  214. Zhang, Y.; Feng, T.; Ni, X.; Xia, J.; Suo, H.; Yan, L.; Zou, B. Immobilized Lipase Based on SBA-15 Adsorption and Gel Embedding for Catalytic Synthesis of Isoamyl Acetate. Food Biosci. 2024, 60, 104427. [Google Scholar] [CrossRef]
  215. Gao, X.; Guo, C.; Hao, J.; Zhao, Z.; Long, H.; Li, M. Adsorption of Heavy Metal Ions by Sodium Alginate Based Adsorbent-a Review and New Perspectives. Int. J. Biol. Macromol. 2020, 164, 4423–4434. [Google Scholar] [CrossRef] [PubMed]
  216. Lv, X.; Jiang, G.; Xue, X.; Wu, D.; Sheng, T.; Sun, C.; Xu, X. Fe0-Fe3O4 Nanocomposites Embedded Polyvinyl Alcohol/Sodium Alginate Beads for Chromium (VI) Removal. J. Hazard. Mater. 2013, 262, 748–758. [Google Scholar] [CrossRef] [PubMed]
  217. Yi, X.; Sun, F.; Han, Z.; Han, F.; He, J.; Ou, M.; Gu, J.; Xu, X. Graphene Oxide Encapsulated Polyvinyl Alcohol/Sodium Alginate Hydrogel Microspheres for Cu (II) and U (VI) Removal. Ecotoxicol. Environ. Saf. 2018, 158, 309–318. [Google Scholar] [CrossRef]
  218. Vu, H.C.; Dwivedi, A.D.; Le, T.T.; Seo, S.-H.; Kim, E.-J.; Chang, Y.-S. Magnetite Graphene Oxide Encapsulated in Alginate Beads for Enhanced Adsorption of Cr(VI) and As(V) from Aqueous Solutions: Role of Crosslinking Metal Cations in pH Control. Chem. Eng. J. 2017, 307, 220–229. [Google Scholar] [CrossRef]
  219. Feng, Y.; Wang, Y.; Wang, Y.; Zhang, X.-F.; Yao, J. In-Situ Gelation of Sodium Alginate Supported on Melamine Sponge for Efficient Removal of Copper Ions. J. Colloid Interface Sci. 2018, 512, 7–13. [Google Scholar] [CrossRef]
  220. Foong, C.Y.; Wirzal, M.D.H.; Bustam, M.A. A Review on Nanofibers Membrane with Amino-Based Ionic Liquid for Heavy Metal Removal. J. Mol. Liq. 2020, 297, 111793. [Google Scholar] [CrossRef]
  221. Karatutlu, A.; Barhoum, A.; Sapelkin, A. Chapter 1—Liquid-Phase Synthesis of Nanoparticles and Nanostructured Materials. In Emerging Applications of Nanoparticles and Architecture Nanostructures; Barhoum, A., Makhlouf, A.S.H., Eds.; Micro and Nano Technologies; Elsevier: Cambridge, MA, USA, 2018; pp. 1–28. ISBN 978-0-323-51254-1. [Google Scholar]
  222. Wu, H.; Wang, W.; Huang, Y.; Han, G.; Yang, S.; Su, S.; Sana, H.; Peng, W.; Cao, Y.; Liu, J. Comprehensive Evaluation on a Prospective Precipitation-Flotation Process for Metal-Ions Removal from Wastewater Simulants. J. Hazard. Mater. 2019, 371, 592–602. [Google Scholar] [CrossRef] [PubMed]
  223. Hosseini, S.M.; Alibakhshi, H.; Jashni, E.; Parvizian, F.; Shen, J.N.; Taheri, M.; Ebrahimi, M.; Rafiei, N. A Novel Layer-by-Layer Heterogeneous Cation Exchange Membrane for Heavy Metal Ions Removal from Water. J. Hazard. Mater. 2020, 381, 120884. [Google Scholar] [CrossRef] [PubMed]
  224. Papageorgiou, S.K.; Katsaros, F.K.; Kouvelos, E.P.; Nolan, J.W.; Le Deit, H.; Kanellopoulos, N.K. Heavy Metal Sorption by Calcium Alginate Beads from Laminaria Digitata. J. Hazard. Mater. 2006, 137, 1765–1772. [Google Scholar] [CrossRef]
  225. Li, X.; Qi, Y.; Li, Y.; Zhang, Y.; He, X.; Wang, Y. Novel Magnetic Beads Based on Sodium Alginate Gel Crosslinked by Zirconium(IV) and Their Effective Removal for Pb2+ in Aqueous Solutions by Using a Batch and Continuous Systems. Bioresour. Technol. 2013, 142, 611–619. [Google Scholar] [CrossRef]
  226. Wang, B.; Gao, B.; Wan, Y. Entrapment of Ball-Milled Biochar in Ca-Alginate Beads for the Removal of Aqueous Cd(II). J. Ind. Eng. Chem. 2018, 61, 161–168. [Google Scholar] [CrossRef]
  227. Patiño-Ruiz, D.; Bonfante, H.; De Ávila, G.; Herrera, A. Adsorption Kinetics, Isotherms and Desorption Studies of Mercury from Aqueous Solution at Different Temperatures on Magnetic Sodium Alginate-Thiourea Microbeads. Environ. Nanotechnol. Monit. Manag. 2019, 12, 100243. [Google Scholar] [CrossRef]
  228. Godiya, C.B.; Xiao, Y.; Lu, X. Amine Functionalized Sodium Alginate Hydrogel for Efficient and Rapid Removal of Methyl Blue in Water. Int. J. Biol. Macromol. 2020, 144, 671–681. [Google Scholar] [CrossRef]
  229. Zhong, Y.; Ning, S.; Wu, K.; Li, Z.; Wang, X.; He, C.; Fujita, T.; Wang, J.; Chen, L.; Yin, X.; et al. Novel Phosphate Functionalized Sodium Alginate Hydrogel for Efficient Adsorption and Separation of Nd and Dy from Co. J. Environ. Manag. 2024, 353, 120283. [Google Scholar] [CrossRef]
  230. Zhao, X.; Wang, X.; Song, G.; Lou, T. Microwave Assisted Copolymerization of Sodium Alginate and Dimethyl Diallyl Ammonium Chloride as Flocculant for Dye Removal. Int. J. Biol. Macromol. 2020, 156, 585–590. [Google Scholar] [CrossRef]
  231. Zhang, Z.; Li, K.; Dong, W.; Wang, Z.; Zhang, X.; Wang, J. An Ingenious Construction of Porous Sodium Alginate/TEMPO-Oxidized Cellulose Composite Aerogels for Efficient Adsorption of Crystal Violet Dyes in Wastewater. J. Sol-Gel Sci. Technol. 2024, 110, 391–405. [Google Scholar] [CrossRef]
  232. Shil, D.C.; Rahman, N.; Sultana, S.; Sardar, M.N.; Majumder, P.; Robel, F.N. Preparation & Characterization of Polyvinyl Alcohol-Sodium Alginate-Starch Based Hydrogel by Gamma Radiation and Its Application for the Treatment of Dye Containing Water. Adv. Environ. Eng. Res. 2023, 4, 1–17. [Google Scholar] [CrossRef]
  233. Hemdan, M.; Ragab, A.H.; Gumaah, N.F.; Mubarak, M.F. Sodium Alginate-Encapsulated Nano-Iron Oxide Coupled with Copper-Based MOFs (Cu-BTC@Alg/Fe3O4): Versatile Composites for Eco-Friendly and Effective Elimination of Rhodamine B Dye in Wastewater Purification. Int. J. Biol. Macromol. 2024, 274, 133498. [Google Scholar] [CrossRef]
  234. Jiang, X.; Xiang, N.; Zhang, H.; Sun, Y.; Lin, Z.; Hou, L. Preparation and Characterization of Poly(Vinyl Alcohol)/Sodium Alginate Hydrogel with High Toughness and Electric Conductivity. Carbohydr. Polym. 2018, 186, 377–383. [Google Scholar] [CrossRef]
  235. Ma, W.; Zhang, Y.; Pan, S.; Cheng, Y.; Shao, Z.; Xiang, H.; Chen, G.; Zhu, L.; Weng, W.; Bai, H.; et al. Smart Fibers for Energy Conversion and Storage. Chem. Soc. Rev. 2021, 50, 7009–7061. [Google Scholar] [CrossRef]
  236. Qi, X.; Liu, Y.; Yu, L.; Yu, Z.; Chen, L.; Li, X.; Xia, Y. Versatile Liquid Metal/Alginate Composite Fibers with Enhanced Flame Retardancy and Triboelectric Performance for Smart Wearable Textiles. Adv. Sci. Weinh. Baden-Wurtt. Ger. 2023, 10, e2303406. [Google Scholar] [CrossRef]
  237. Zaman, S.U.; Mushtaq, B.; Ahmad, F.; Ahmad, S.; Rasheed, A.; Nawab, Y. Development of Conductive Cotton Non-Woven Alginate Hydrogel Composite for Smart Textiles. J. Polym. Environ. 2023, 31, 3998–4006. [Google Scholar] [CrossRef]
  238. Wang, Y.; Wen, Q.; Chen, Y.; Zheng, H.; Wang, S. Enhanced Performance of Microbial Fuel Cell with Polyaniline/Sodium Alginate/Carbon Brush Hydrogel Bioanode and Removal of COD. Energy 2020, 202, 117780. [Google Scholar] [CrossRef]
  239. Nugraha, I.A.; Supriyanto, A.; Pauzi, G.A. The Microbial Fuel Cell Characteristics of the PVA/Chitosan Membrane with Variations of Phosphate Acid and Sodium Alginate Derived from Vegetable Waste. J. Energy, Mater. Instrum. Technol. 2023, 4, 136–143. [Google Scholar] [CrossRef]
  240. Cheng, L.; Jiang, L.; Yang, X.; Gao, Y.; Gai, R.; Wang, M.; Chen, L. The Performance of Microbial Fuel Cell with Sodium Alginate and Super Activated Carbon Composite Gel Modified Anode. AMB Express 2024, 14, 67. [Google Scholar] [CrossRef]
  241. Zhao, C.; Chen, H.; Song, Y.; Zhu, L.; Ai, T.; Wang, X.; Liu, Z.; Wei, X. Electricity Production Performance Enhancement of Microbial Fuel Cells with Double-Layer Sodium Alginate Hydrogel Bioanodes Driven by High-Salinity Waste Leachate. Water Res. 2023, 242, 120281. [Google Scholar] [CrossRef]
  242. Solangi, K.A.; Siyal, A.A.; Wu, Y.; Abbasi, B.; Solangi, F.; Lakhiar, I.A.; Zhou, G. An Assessment of the Spatial and Temporal Distribution of Soil Salinity in Combination with Field and Satellite Data: A Case Study in Sujawal District. Agronomy 2019, 9, 869. [Google Scholar] [CrossRef]
  243. Li, Z.; Wu, Y.; Xing, D.; Zhang, K.; Xie, J.; Yu, R.; Chen, T.; Duan, R. Effects of Foliage Spraying with Sodium Bisulfite on the Photosynthesis of Orychophragmus Violaceus. Horticulturae 2021, 7, 137. [Google Scholar] [CrossRef]
  244. Yang, Q.; Yang, X.; Zhang, Z.; Wang, J.; Fu, W.; Li, Y. Investigating the Resistance Levels and Mechanisms to Penoxsulam and Cyhalofop-Butyl in Barnyardgrass (Echinochloa Crus-Galli) from Ningxia Province, China. Weed Sci. 2021, 69, 422–429. [Google Scholar] [CrossRef]
  245. Wang, N.; Wang, B.; Wan, Y.; Gao, B.; Rajput, V.D. Alginate-Based Composites as Novel Soil Conditioners for Sustainable Applications in Agriculture: A Critical Review. J. Environ. Manag. 2023, 348, 119133. [Google Scholar] [CrossRef]
  246. Yuan, X.; Yu, S.; Xue, N.; Li, T.; Sun, M. Persulfate Activation with Sodium Alginate/Sulfide Coated Iron Nanoparticles for Degradation of Tetrabromobisphenol a in Soil. Environ. Res. 2023, 221, 114820. [Google Scholar] [CrossRef]
  247. Feng, Q.; Chen, M.; Wu, P.; Zhang, X.; Wang, S.; Yu, Z.; Wang, B. Calcium Alginate-Biochar Composite as a Novel Amendment for the Retention and Slow-Release of Nutrients in Karst Soil. Soil Tillage Res. 2022, 223, 105495. [Google Scholar] [CrossRef]
  248. Lu, W.; Tang, H.; Li, Y.; Wang, J.; Sun, J. Preparation and Application of a Natural Microspheric Soil Conditioner Based on Gelatin, Sodium Alginate, and Zeolite. ACS Appl. Polym. Mater. 2023, 5, 5211–5220. [Google Scholar] [CrossRef]
  249. Zhang, S.; He, F.; Fang, X.; Zhao, X.; Liu, Y.; Yu, G.; Zhou, Y.; Feng, Y.; Li, J. Enhancing Soil Aggregation and Acetamiprid Adsorption by Ecofriendly Polysaccharides Hydrogel Based on Ca2+-Amphiphilic Sodium Alginate. J. Environ. Sci. 2022, 113, 55–63. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Conceptual roadmap illustrating the main modification strategies of sodium alginate (chemical, physical, and enzymatic) and their corresponding application fields. In the drug-delivery schematic, A and B represent two inlet channels introducing different drug formulations, while C denotes the outlet channel. The arrows indicate the flow direction during microfluidic-assisted mixing.
Figure 1. Conceptual roadmap illustrating the main modification strategies of sodium alginate (chemical, physical, and enzymatic) and their corresponding application fields. In the drug-delivery schematic, A and B represent two inlet channels introducing different drug formulations, while C denotes the outlet channel. The arrows indicate the flow direction during microfluidic-assisted mixing.
Foods 14 03931 g001
Figure 2. Schematic diagram of the traditional extraction of sodium alginate.
Figure 2. Schematic diagram of the traditional extraction of sodium alginate.
Foods 14 03931 g002
Figure 3. Chain conformation (A) and block distribution (B) of sodium alginate, where M and G represent β-D-mannuronic acid and α-L-guluronic acid residues, respectively.
Figure 3. Chain conformation (A) and block distribution (B) of sodium alginate, where M and G represent β-D-mannuronic acid and α-L-guluronic acid residues, respectively.
Foods 14 03931 g003
Figure 4. “Egg-box” gel structure of sodium alginate, illustrating the coordination of Ca2+ ions within G-block regions.
Figure 4. “Egg-box” gel structure of sodium alginate, illustrating the coordination of Ca2+ ions within G-block regions.
Foods 14 03931 g004
Figure 5. Esterification of sodium alginate: formation of ester linkages between hydroxyl and acyl groups.
Figure 5. Esterification of sodium alginate: formation of ester linkages between hydroxyl and acyl groups.
Foods 14 03931 g005
Figure 6. Oxidation of alginate chains via periodate cleavage of vicinal diols to generate dialdehyde sites.
Figure 6. Oxidation of alginate chains via periodate cleavage of vicinal diols to generate dialdehyde sites.
Foods 14 03931 g006
Figure 7. Sulfation of alginate through substitution of hydroxyl groups with sulfate esters (–OSO3), enhancing hydrophilicity.
Figure 7. Sulfation of alginate through substitution of hydroxyl groups with sulfate esters (–OSO3), enhancing hydrophilicity.
Foods 14 03931 g007
Figure 8. Ugi multi-component reaction forming amide linkages between amino, aldehyde, carboxyl, and isocyanide groups.
Figure 8. Ugi multi-component reaction forming amide linkages between amino, aldehyde, carboxyl, and isocyanide groups.
Foods 14 03931 g008
Figure 9. Aldehyde cross-linking through Schiff-base formation between aldehyde and hydroxyl or amino groups.
Figure 9. Aldehyde cross-linking through Schiff-base formation between aldehyde and hydroxyl or amino groups.
Foods 14 03931 g009
Figure 10. Phosphorylation introducing phosphate groups that improve metal chelation and biological functionality.
Figure 10. Phosphorylation introducing phosphate groups that improve metal chelation and biological functionality.
Foods 14 03931 g010
Figure 11. Amidation reaction producing amide bonds between carboxyl and amino groups via carbodiimide coupling.
Figure 11. Amidation reaction producing amide bonds between carboxyl and amino groups via carbodiimide coupling.
Foods 14 03931 g011
Figure 12. Grafted calcium alginate (PSC–CA) hydrogel beads functionalized with amino-carbamate moieties for enhanced strength and adsorption.
Figure 12. Grafted calcium alginate (PSC–CA) hydrogel beads functionalized with amino-carbamate moieties for enhanced strength and adsorption.
Foods 14 03931 g012
Figure 13. Simplified schematic illustration of the structural evolution and rheological behavior of SA/PEGDA systems under increasing PEGDA content.
Figure 13. Simplified schematic illustration of the structural evolution and rheological behavior of SA/PEGDA systems under increasing PEGDA content.
Foods 14 03931 g013
Figure 14. Schematic illustration of the ultrasonication-assisted modification of SA-based systems. (a) Ultrasound wave and cavitation process induced by Ca2+; (b) localized high temperature and pressure leading to rearrangement of SA polymer chains; (c) chain disentanglement and partial depolymerization of SA.
Figure 14. Schematic illustration of the ultrasonication-assisted modification of SA-based systems. (a) Ultrasound wave and cavitation process induced by Ca2+; (b) localized high temperature and pressure leading to rearrangement of SA polymer chains; (c) chain disentanglement and partial depolymerization of SA.
Foods 14 03931 g014
Figure 15. Schematic illustration of the irradiation-assisted modification of SA-based systems. (a) Formation of reactive free radicals through electron beam or γ-ray irradiation; (b) crosslinking and grafting reactions between SA molecular chains induced by the generated radicals; (c) formation of SA hydrogel or coating materials applied for fruit preservation and other functional uses.
Figure 15. Schematic illustration of the irradiation-assisted modification of SA-based systems. (a) Formation of reactive free radicals through electron beam or γ-ray irradiation; (b) crosslinking and grafting reactions between SA molecular chains induced by the generated radicals; (c) formation of SA hydrogel or coating materials applied for fruit preservation and other functional uses.
Foods 14 03931 g015
Figure 16. Schematic representation of the β-elimination catalytic mechanism of alginate lyase acting on SA, showing the enzymatic cleavage of 1,4-glycosidic linkages between uronic acid residues and the formation of unsaturated alginate oligosaccharides (AOSs).
Figure 16. Schematic representation of the β-elimination catalytic mechanism of alginate lyase acting on SA, showing the enzymatic cleavage of 1,4-glycosidic linkages between uronic acid residues and the formation of unsaturated alginate oligosaccharides (AOSs).
Foods 14 03931 g016
Figure 18. Schematic illustration of the 3D printing process using SA-based bioinks for tissue engineering scaffolds. (a) Layer-by-layer 3D printing of SA hydrogel guided by patient CT/MRI data; (b) formation of SA hydrogel scaffolds with interconnected porous microstructures; (c) incorporation of bioactive components such as mesoporous bioactive glass (MBG) or nacreous powder (NP) within the printed scaffold to enhance biological functionality.
Figure 18. Schematic illustration of the 3D printing process using SA-based bioinks for tissue engineering scaffolds. (a) Layer-by-layer 3D printing of SA hydrogel guided by patient CT/MRI data; (b) formation of SA hydrogel scaffolds with interconnected porous microstructures; (c) incorporation of bioactive components such as mesoporous bioactive glass (MBG) or nacreous powder (NP) within the printed scaffold to enhance biological functionality.
Foods 14 03931 g018
Figure 19. Schematic illustration of SA-based multifunctional drug delivery systems. (a) Formation of pH/ion-responsive hydrogels through Ca2+ crosslinking under acidic or basic conditions, enabling controlled drug release; (b) Construction of polyelectrolyte complex (PEC) networks between SA and oppositely charged polymers for enhanced drug encapsulation and stability; (c) Development of liposome–SA composite systems, where liposomes are coated with SA hydrogels to improve mechanical strength and achieve sustained release performance.
Figure 19. Schematic illustration of SA-based multifunctional drug delivery systems. (a) Formation of pH/ion-responsive hydrogels through Ca2+ crosslinking under acidic or basic conditions, enabling controlled drug release; (b) Construction of polyelectrolyte complex (PEC) networks between SA and oppositely charged polymers for enhanced drug encapsulation and stability; (c) Development of liposome–SA composite systems, where liposomes are coated with SA hydrogels to improve mechanical strength and achieve sustained release performance.
Foods 14 03931 g019
Figure 20. Adsorption mechanism of metal ions by SA-based adsorbent [215].
Figure 20. Adsorption mechanism of metal ions by SA-based adsorbent [215].
Foods 14 03931 g020
Table 1. Comparison of chemical, physical, and biological modification methods for SA.
Table 1. Comparison of chemical, physical, and biological modification methods for SA.
DimensionChemical ModificationPhysical ModificationBiological Modification
Modification effectStrong functionalization and high stabilitySimple, fast, and environmentally friendlyHigh biocompatibility and strong specificity
CostHigh (reagents, purification)Low (no need for complex equipment)Extremely high (enzyme/genetic engineering)
Technological difficultyComplex (requiring precise control of reaction conditions)Simple (easy to industrialize)Complex (requiring biotechnological conditions)
Toxic riskHarmful substances may remainNoneNone
ApplicationIndustrial adsorbents and functional materialsFood packaging, sustained-release carriersBiomedical, tissue engineering
Table 3. SA-based functional carriers: methods and readouts.
Table 3. SA-based functional carriers: methods and readouts.
Cargo/SystemMatrix and MethodKey ReadoutsUse-CaseReference
Bovine lactoferrin (LFNP)EHDA nanoparticles in NaAlg matrixSize ~ 100–200 nm; |ζ| ~ 20 mV; stable dispersionsIron-delivery/antioxidant[197]
Marjoram essential oil (EO)SA + WPI (ionic gel)EE/size tuned by SA/WPI/Ca2+; stable aroma retentionAntimicrobial flavor delivery[192]
α-TocopherolSA beads (ionic gel)Release ~ 29% (SGF) vs. ~82% (SIF); T50% ~ 3.8 h; T70% ~ 12.3 h (SIF)Gastric protection; intestinal delivery[191]
Probiotics (Lactobacillus spp.)Fish-gelatin/SA double-network (FG/SA-DN)Encapsulation efficiency ~16%→~92% (FG ↑); GI/thermal survival ↑Fermented/baked foods[195]
Abbreviations: arrows (↑/↓) indicate the direction of increase or decrease in the measured parameter.
Table 4. Comparative summary of conventional heavy metal ion removal techniques and functionalized SA-based adsorption systems.
Table 4. Comparative summary of conventional heavy metal ion removal techniques and functionalized SA-based adsorption systems.
TechnologyKey AdvantagesMajor LimitationsReference
Membrane separationMinimal chemical usage, compact system footprint, selective metal recoveryHigh membrane procurement cost, frequent fouling issues, restricted throughput capacity[220]
Electrochemical recoveryHigh-purity metal recovery, ambient condition compatibilityIntensive energy demand, slow reaction kinetics, potential electrolyte contamination[221]
Chemical precipitationSimplified operational workflow, low infrastructure costExcessive sludge yield (high disposal burden), non-selective removal, risk of secondary contamination[222]
Ion-exchangeTargeted metal binding capability, high regeneration efficiencyElevated upfront investment, narrow pH operating range, recurrent maintenance expenses[223]
SA (Ca2+-crosslinked) hydrogel beadsAbundant carboxylate groups for chelation; low cost; biocompatible; easy beadization; regenerable with mild eluentsGel swelling/softening; dissolution at low pH or chelating eluents; limited selectivity; intraparticle diffusion limits[224]
Magnetic SA/Fe3O4 beadsRapid magnetic separation; easy recovery and reuse; good dispersionFe3O4 oxidation/leaching; capacity decay across cycles; acid instability; added material cost[225]
SA–biochar/zeolite/clay hybridsLow-cost supports; improved permeability and strength; resilience to turbidityBatch-to-batch variability; competing ions; fines shedding[226]
Thiol-functionalized SA (–SH, dithiocarbamate)High selectivity for soft metal ions (e.g., Hg2+, Ag+, Pb2+, Cd2+)Thiol oxidation; odor; multi-step synthesis; cost[227]
Amine/EDA/PEI-functionalized SAStrong complexation with Cu2+/Ni2+/Cr (VI); rapid kineticsAmine protonation at low pH reduces capacity; polymer leaching; fouling[228]
Phosphate/phosphonate-modified SAHigh affinity for Pb2+, rare earths; improved selectivity in competing electrolytesSynthesis complexity; potential ligand leaching; cost[229]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, W.; Huang, Y.; Pan, Y.; Dabbour, M.; Dai, C.; Zhou, M.; He, R. Sodium Alginate Modifications: A Critical Review of Current Strategies and Emerging Applications. Foods 2025, 14, 3931. https://doi.org/10.3390/foods14223931

AMA Style

Wang W, Huang Y, Pan Y, Dabbour M, Dai C, Zhou M, He R. Sodium Alginate Modifications: A Critical Review of Current Strategies and Emerging Applications. Foods. 2025; 14(22):3931. https://doi.org/10.3390/foods14223931

Chicago/Turabian Style

Wang, Wenning, Yuanyuan Huang, Yun Pan, Mokhtar Dabbour, Chunhua Dai, Man Zhou, and Ronghai He. 2025. "Sodium Alginate Modifications: A Critical Review of Current Strategies and Emerging Applications" Foods 14, no. 22: 3931. https://doi.org/10.3390/foods14223931

APA Style

Wang, W., Huang, Y., Pan, Y., Dabbour, M., Dai, C., Zhou, M., & He, R. (2025). Sodium Alginate Modifications: A Critical Review of Current Strategies and Emerging Applications. Foods, 14(22), 3931. https://doi.org/10.3390/foods14223931

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