Next Article in Journal
Hydrogel-Based Nitric Oxide Delivery Systems for Enhanced Wound Healing
Previous Article in Journal
Self-Healing, Electroconductive Hydrogels for Wound Healing Applications
Previous Article in Special Issue
Thermal Damping Applications of Coconut Oil–Silica Gels and Their Rheological Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tuning Nanostructure of Gels: From Structural and Functional Controls to Food Applications

Nano-Biotechnology Group, Faculty of Bioscience Engineering, Ghent University, 9000 Ghent, Belgium
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Gels 2025, 11(8), 620; https://doi.org/10.3390/gels11080620
Submission received: 14 July 2025 / Revised: 28 July 2025 / Accepted: 6 August 2025 / Published: 8 August 2025
(This article belongs to the Special Issue Thixotropic Gels: Mechanisms, Functions and Applications)

Abstract

Various gels are integral for the food industry, providing unique textural and mechanical properties essential for the quality and functions of products. These properties are fundamentally governed by the gels’ nanostructural organization. This review investigates the mechanisms of nanostructure formation in food gels, the methods for their characterization and control, and how precise tuning of these nanostructures enables targeted food applications. We examine the role of various building blocks, including biopolymers, lipids, and particles, and the gelation mechanisms leading to specific nanostructures. Advanced techniques (e.g., microscopy, scattering, spectroscopy, and rheology) are discussed for their insights into nano-/microstructures. Strategies for tuning nanostructures through chemical composition adjustments (e.g., concentration, pH, ionic strength) and physical processing controls (e.g., temperature, shear, ultrasound) are presented. Incorporating nanostructures like nanoparticles and nanofibers to enhance gel properties is also explored. The review links these nanostructures to key functional properties, including mechanical strength, water-holding capacity, optical characteristics, and bioactive delivery. By manipulating nanostructures, products can achieve tailored textures, improved stability, and controlled nutrient release. Applications enabled by nanostructure tuning include tailored sensory experiences, fat reduction, innovative food structures, and smart packaging solutions. Although significant progress has been made, precise structural control and a comprehensive understanding of complex nanoscale interactions in food gels remain challenging. This review underscores the importance of nanostructure tuning in food gels, highlighting its potential to drive future research that unlocks innovative, functional food products.

1. Introduction

Food gels are soft materials characterized by a continuous three-dimensional (3D) network that encapsulates significant volumes of liquids, such as water or oil, resulting in a viscoelastic system that exhibits properties intermediate between solids and liquids [1,2]. These networks, typically formed from biopolymers such as proteins [3], polysaccharides [4], or a combination of other biopolymers [5,6,7], are classified as hydrogels [6], oleogels [8], or bigels [9], depending on the entrapped liquid. The 3D network is critical for immobilizing liquids, which underpins the gels’ distinctive textural and mechanical properties, including firmness, elasticity, and water-holding capacity (WHC) [10]. These attributes significantly influence the texture, mouthfeel, and sensory appeal of a wide range of food products, such as yogurt, cheese, jellies, sauces, processed meats, and plant-based analogs [11,12]. Additionally, food gels enhance structural stability, prevent phase separation, extend shelf-life, and serve as carriers for bioactive compounds, enabling controlled release and improved stability of nutrients, flavors, and functional ingredients [5,10].
The adaptability of food gels arises from their tunable properties, which allow for the design of products with customized textures, ranging from smooth and soft to firm and elastic [13,14], and tailored to meet specific dietary needs, such as reduced-fat formulations or foods for individuals with swallowing difficulties [15]. This versatility supports their use in traditional foods like tofu, cheese, and jellies [2], as well as in innovative applications, including plant-based analogs, 3D-printed foods, and edible coatings [16].
The functionality of food gels is fundamentally governed by their nanoscale structural organization [17]. The arrangement and interactions of molecular building blocks (e.g., proteins [18], nanoparticles or nanocrystals [19], and low-molecular-weight gelators [20,21]) directly determine macroscopic properties like texture, mechanical strength, and performance [15]. For instance, in protein-based gels, nanoscale gelation and aggregation processes form network structures that encapsulate bioactive compounds, thereby affecting texture, stability, and nutritional functionality [18,22]. Likewise, the self-assembly of nanoparticles or nanocrystals into interconnected networks enables the transfer of size-dependent properties to the macroscale, with nanoscale connectivity influencing mechanical and transport behaviors [19,23,24]. Hierarchical self-assembly of molecular gelators into defined nanostructures enables the creation of tunable stimuli-responsive, self-healing properties gels [20,25].
Rational design of food gels depends on manipulating this nanoscale organization. By modifying building blocks through chemical functionalization, adjusting assembly conditions, or using specific linkers [26], researchers can control network connectivity, porosity, and mechanical properties to achieve targeted functionalities, such as enhanced toughness or specific release profiles [17,18]. Understanding the relationships between nanostructure, structure, and properties is thus essential for developing advanced food gels with precisely engineered textures, stability, and functional benefits.
In food gels, the nanostructure-function relationship is driven by the self-assembly of biopolymers or colloids into 3D networks. The nanoscale architecture—determined by such factors as type, density, connectivity, and intermolecular interactions (e.g., hydrogen bonding, electrostatic forces)—directly governs mechanical strength, liquid retention, and textural qualities [15,20]. By fine-tuning parameters like nanofibril morphology, concentration, and assembly conditions, scientists can customize gel properties, including elasticity, breakdown behavior, and responsiveness to external stimuli [27,28,29,30]. The stability and functionality of these gels rely on both covalent and non-covalent bonds between nano-building blocks, such as nanofibers or protein–polysaccharide complexes [30,31]. Techniques such as ionic gelation, composite systems [32], and responsive supramolecular designs [29] enable the engineering of sophisticated food gels with properties optimized for advancements in nutrition, sensory experience, and sustainability.
Spanning the disciplines of food science, materials science, colloid chemistry, and rheology, this review consolidates contemporary knowledge on the mechanisms governing nanostructure formation in food gels, the methodologies for characterizing and controlling these structures, the correlation between nanostructure and functional properties, and the ways in which precise structural control can be harnessed for targeted applications. Through this interdisciplinary lens, it aims to illustrate how tuning nanostructures can catalyze innovation in food gels’ development, thereby addressing consumer needs for quality, health, and sustainability.

2. Building Blocks and Gelation Mechanisms

2.1. Structuring Agents

Food gels are primarily structured from biopolymers, with proteins and polysaccharides serving as the main building blocks, as shown in Figure 1 [33,34].

2.1.1. Proteins

Proteins, including globular types (such as whey, soy, and egg proteins), fibrous proteins (like gelatin and collagen), and casein micelles, form gels mainly through heat-induced denaturation and aggregation or by cooling, as seen with gelatin [22,35]. Additionally, peptides, especially those capable of self-assembly, act as adaptable building blocks in food gels by organizing into nanostructures like fibrils, nanotubes, and micelles [36]. The gelation process for proteins is influenced by factors such as pH, ionic strength, and temperature, which affect their intermolecular interactions and ultimately the texture and microstructure of the gel network [3,18]. Furthermore, the interactions between different proteins (e.g., globular proteins, casein, and gelatin) during heating or cooling significantly affect gel stiffness and microstructure, with scenarios including independent aggregation, co-aggregation, and phase separation being key considerations [37,38]. This is exemplified in an ultrasound-treated soy protein isolate/egg white protein (SPI-EWP) composite gel [39] where accelerated co-aggregation kinetics (due to reduced electrostatic repulsion and exposed sulfhydryl groups) form dense yet porous networks, enhancing gel stiffness in SPI-dominated systems but increasing heterogeneity in EWP-rich gels.

2.1.2. Polysaccharides

Polysaccharides, another essential group, include helix-forming types (κ-carrageenan, gellan), charged varieties (alginate, pectin, xanthan), and neutral forms (starch, cellulose derivatives) [4]. Their gelation depends on hydration, ionic interactions, and molecular entanglement, and they are often used in combination with proteins to create composite gels with tailored properties [4,40]. Recent advances include the creation of smart hydrogels with antimicrobial and barrier properties, and the use of oleogels and emulsion gels as healthier fat substitutes [41]. Additionally, new drying technologies [42] and composite gel designs are expanding applications in food structuring, even biomedical fields, though challenges remain in optimizing gel properties and integrating sustainable raw materials [43].
The interplay between proteins and polysaccharides allows for the design of gels with diverse structures and functionalities, making them highly valuable in food applications for texture modification, stability, and delivery of bioactive substances [44]. For example, rice starch and soybean protein exhibit antagonistic interactions in composite gels, resulting in phase-separated networks with reduced gel strength and the International Dysphagia Diet Standardization Initiative (IDDSI) Level 5 suitability, enabling softer texture-modified foods for moderate dysphagia patients when blended at a 50:50 ratio [45]. Such antagonistic or synergistic effects are highly dependent on the specific biopolymer pairings and their microstructural compatibility, highlighting the need for continued research into structure–function relationships.

2.1.3. Lipids

Lipid-based food gels serve as essential carriers for lipophilic ingredients, utilizing either crystalline networks (oleogels/organogels [10,46]) or emulsified structures (emulsion gels [47,48]). Crystalline networks form through self-assembling gelators that trap oil in solid-like matrices, offering controlled release and stability for fat-soluble compounds [46]. Emulsion gels incorporate oil droplets within aqueous or lipid phases stabilized by proteins, polysaccharides, or particles (e.g., Pickering emulsions [49]), enabling encapsulation of both hydrophilic and lipophilic components [23]. Both systems can improve stability, bioavailability, and release kinetics of bioactives, with these properties influenced by factors such as gelator type, crystalline structure, droplet size, and stabilizer selection. Emulsion gels also offer additional benefits for fat replacement and texture modification in functional foods [50,51]. These tunable structures provide versatile solutions for optimizing sensory properties and health-oriented delivery systems.
Recent research has focused on developing healthier fat alternatives, improving functional properties, and expanding applications in food technology [46,48]. Glyceryl monostearate-based oleogels have been shown to form solid-like gels with sunflower and high oleic sunflower oils, with the oil type influencing gel network formation and crystal morphology, which can be tailored for improved nutritional profiles in food products [52]. Moreover, surfactant-free, plant-based double emulsion gels (O/W/O and W/O/W) have been engineered as low-calorie fat analogs, offering tunable oral perception, inhibition of lipid digestion, and effective co-delivery of bioactives like lycopene and epigallocatechin gallate (EGCG), with O/W/O gels mimicking butter’s texture and both types showing anti-inflammatory activity [53]. Comparative studies of structuring agents such as wax crystals, hydrophilic cellulose derivatives, and emulsion droplets reveal distinct rheological and thermal behaviors, guiding the selection of gel systems for specific food applications [54]. The field is also seeing rapid innovation in delivery systems, with trends moving toward complex structures like Pickering emulsions, bigels, and multiple emulsions, which improve bioactive delivery, sensory experience, and enable reduced-fat formulations [55].

2.1.4. Particles

Particles serve as building blocks in food gels because they influence their texture, stability, encapsulation, and release properties (Figure 1) [56,57]. Their incorporation into gels leads to the formation of the so-called hybrid structures [58]. Particles such as protein-based fillers can increase the elastic modulus, mechanical properties, and firmness of food gels, depending on their size, concentration, and interaction with the gel matrix [59,60]. However, gels can also be filled with emulsions [61], whose concentration can be used to tune mechanical properties. Further, gel particles can act as stabilizers in Pickering emulsion gels, forming a dense layer around oil droplets and preventing coalescence [62]. This enhances the stability and layer-by-layer interfacial structure of emulsions, which is important for products like dressings and spreads. Recent advances in particle-filled food gels have focused on enhancing mechanical properties, developing novel structures, and expanding applications in food technology [63,64]. Innovative strategies such as particle/fiber reinforcement, double-network formation, dual cross-linking, and freeze-thaw cycles have been shown to significantly improve gel strength and resilience, depending on the type of filler, gel matrix, and their interactions [63]. For example, incorporating glass microspheres into whey protein isolate/xanthan gum gels increased the elastic modulus, with effects influenced by filler size, polydispersity, and ionic strength [59]. Thereinto, glass microspheres and similar rigid particles are frequently used as model fillers in research to investigate the structural and mechanical properties of composite food gels, though they are not intended for direct consumption. Novel emulsion microgel particles, created by combining gelatinized native starch and octenyl succinic anhydride starch-stabilized emulsion droplets, can be tailored for the delivery of lipophilic molecules in food systems [65]. Additionally, the development of thermo-responsive double-interpenetrating colloidal-particle networks, such as those combining silica and lipid particles, has resulted in gels with enhanced stiffness and resilience, suitable for 3D-printed food constructs [13]. Advances in microrheology and particle tracking have further enabled detailed mapping of local viscoelastic properties and structural heterogeneities in food gels [66]. These developments collectively highlight the growing potential of particle-filled food gels for applications in personalized nutrition, fat replacement, nutrient delivery, and 3D food printing [1,15].

2.2. Gelation Mechanisms and Nanostructure Genesis

The nanostructure and functional properties of food gels are determined by their gelation mechanism, whether physical (e.g., heat/cool setting, ion-induced, phase separation, crystallization), chemical/enzymatic (e.g., covalent cross-linking via enzymes like transglutaminase), or supramolecular self-assembly (e.g., peptides, sugar derivatives [20]). The gelation mechanisms were summarized and illustrated in Figure 2.

2.2.1. Physical Gelation

  • Heat/Cool-Driven Conformational Transitions. Heat-driven conformational transition occurs when heating unfolds proteins or specific polysaccharides (like konjac glucomannan), promoting aggregation and network formation through hydrophobic interactions, hydrogen bonds, and sometimes disulfide bonds, often facilitated by additional factors like alkali treatment; conversely [18,35]. Cool-driven conformational transition occurs upon cooling, allowing polymers such as gelatin or amylose in starches to reassociate and form stable junction zones and networks, where the cooling process critically determines the resulting gel’s texture and firmness [67].
  • Ion-Induced Gelation. Addition of ions (e.g., Ca2+) to protein or polysaccharide systems (such as alginate or pectin) promotes conformational changes (e.g., α-helix to β-sheet in proteins), enhances hydrophobic interactions, and leads to the formation of stable, ordered 3D networks, as shown in Figure 3A. Excessive ion concentration also causes irregular aggregates and coarser structures [68].
  • Phase Separation. Mixing incompatible biopolymers (e.g., gelatin and polysaccharides) can induce phase separation, resulting in hierarchical or multi-domain microstructures within the gels [69,70].
  • Crystallization. In starch and polysaccharide gels, particularly starch-based systems, crystallization is a key mechanism where linear polymer chains (like amylose) align and associate during cooling, forming stable crystalline junction zones that significantly reinforce the gel network structure; the degree of this crystallization critically determines the gel’s firmness, stability, and overall textural quality in the final product [10,67].

2.2.2. Chemical and Enzymatic Gelation

  • Chemical Gelation. Food gels are mainly stabilized by non-covalent interactions such as hydrogen bonds, electrostatic forces, Van der Waals forces, and hydrophobic interactions, which enable the formation of three-dimensional networks without the need for covalent cross-linking [33,34]. However, disulfide bonds serve as an important exception, providing covalent cross-linking that enhances stability and rigidity in protein gels [15,71]. The resulting gel network’s strength and structure are strongly influenced by factors including temperature, pH, ionic strength, the concentration of gelling agents, and the presence of small molecules such as sugars, acids, and salts [33,67,72].
  • Enzymatic Gelation. The formation of covalent bonds or modify polymer structures in enzymatic gelation are catalyzed by enzymes [73], where transglutaminase and specific proteases induce protein gelation by cross-linking molecules, creating networks with unique textures [73,74], while in polysaccharide systems (e.g., carrageenan, agar, alginate), enzymes like epimerases, desulfatases, and lyases alter the carbohydrate backbone to tailor gel characteristics [74].

2.2.3. Self-Assembly and Supramolecular Gelation

Self-assembling peptides and low-molecular-weight gelators (including sugar derivatives) form nanofibers, nanoparticles, or hierarchical supramolecular architectures through non-covalent interactions [29,36]. Supramolecular gels exhibit diverse nanostructures (fibrils, strands, platelets) that can be modulated for specific functions, such as encapsulation or stimuli-responsiveness [20,36].
Figure 2. Summary diagram of the gelation mechanism, including: Temperature-set gelation: temperature-induced gelatin gels; Ion-induced gelation: Ca2+-induced alginate gels forming egg-box structure; Phase separation: phase separation and gelation in konjac glucomannan and gelatin system (reproduced from [70] with permission from Elsevier); Crystallization: temperature-induced starch retrogradation to form gels; Chemical cross-linking: polymers chains crosslinked by cross-linking agent; Enzymatic cross-linking: polymers chains crosslinked by enzyme; Self-assembly and supramolecular gelation: peptides or low-molecular-weight gelators form supramolecular architectures through molecular interactions.
Figure 2. Summary diagram of the gelation mechanism, including: Temperature-set gelation: temperature-induced gelatin gels; Ion-induced gelation: Ca2+-induced alginate gels forming egg-box structure; Phase separation: phase separation and gelation in konjac glucomannan and gelatin system (reproduced from [70] with permission from Elsevier); Crystallization: temperature-induced starch retrogradation to form gels; Chemical cross-linking: polymers chains crosslinked by cross-linking agent; Enzymatic cross-linking: polymers chains crosslinked by enzyme; Self-assembly and supramolecular gelation: peptides or low-molecular-weight gelators form supramolecular architectures through molecular interactions.
Gels 11 00620 g002

3. Characterization of Nano-/Microstructures in Food Gels

3.1. Microscopy Techniques

Cryo-scanning electron microscopy (Cryo-SEM), transmission electron microscopy (TEM), and atomic force microscopy (AFM) have provided 3D images of various food gels, identifying unique structural features such as fiber networks in heat-induced milk gels and thick chains in acid–heat gels (Figure 3A) [75,76]. Confocal Raman microscopy has mapped protein and water distribution in mixed whey/soy gels, detecting differences in hydration and the presence of residual lipids, with soy-rich regions showing greater auto-fluorescence (Figure 3B) [77]. Stimulated emission depletion (STED) microscopy and fluorescence lifetime imaging microscopy (FLIM) have enabled detailed visualization of protein networks and molecular dynamics and interactions in plant-dairy gels, showing that pectin incorporation increases both the thickness of protein networks and the size of voids, resulting in a more porous and cohesive gel structure (Figure 3C) [78]. STED microscopy, combined with quantitative image analysis, has distinguished between different protein aggregates in acidified milk gels, revealing that nano-particulate whey protein forms larger, more connected aggregates, which correlates with higher gel strength [79]. Light microscopy (LM) as well as polarized light microscopy (PLM), especially with iodine staining, remains a valuable technique for differentiating starch components and monitoring gelatinization, offering more detailed supramolecular insights than SEM [80,81]. Additionally, advanced image analysis and deep learning applied to microscopy images have enabled high-throughput, accurate quantification of starch gelatinization and granule swelling, achieving up to 95% accuracy in identifying birefringence loss during heating [82].
Figure 3. (A) SEM image of gelatin/chitosan/3-phenylacetic acid nanofiber film (adapted from [76] with permission from Elsevier). (B) Micrographs of whey protein isolate (WPI) and soy protein isolate (SPI) gels obtained through confocal laser scanning microscopy (adapted from [77]). High-resolution STED microscopy images of acidified fresh skimmed-milk gels (ASMG) with and without pectin: (C) The control gel without pectin; (D) Gels with high-methoxy pectin (HMP) (AMD 783); (E) Gels with low-methoxy pectin (LMP) (PSY 200) (adapted from [78] with permission from Elsevier).
Figure 3. (A) SEM image of gelatin/chitosan/3-phenylacetic acid nanofiber film (adapted from [76] with permission from Elsevier). (B) Micrographs of whey protein isolate (WPI) and soy protein isolate (SPI) gels obtained through confocal laser scanning microscopy (adapted from [77]). High-resolution STED microscopy images of acidified fresh skimmed-milk gels (ASMG) with and without pectin: (C) The control gel without pectin; (D) Gels with high-methoxy pectin (HMP) (AMD 783); (E) Gels with low-methoxy pectin (LMP) (PSY 200) (adapted from [78] with permission from Elsevier).
Gels 11 00620 g003

3.2. Scattering Techniques

Scattering techniques such as small-angle X-ray scattering (SAXS), small-angle neutron scattering (SANS), and ultra-small-angle neutron scattering (USANS) are powerful non-destructive tools for providing insights into junction zone structures, network density, and persistence length [83,84,85]. These methods enable real-time, in situ monitoring of structural changes during processes like simulated gastric digestion, revealing, for example, that gel firmness and elasticity influence disintegration behavior and digestibility [86]. Additionally, it is confirmed that physical disruption of gel particles at the macroscale does not significantly alter their internal nano- or microstructure, validating the use of blended samples for scattering analysis [87]. Scattering techniques have also elucidated complex hierarchical assembly mechanisms, such as the two-step network formation in nanodiamond hydrogels, and have been combined with imaging to provide comprehensive multi-scale characterization of supramolecular and low-molecular-weight gels [88,89].

3.3. Spectroscopy

Fourier-Transform Infrared Spectroscopy (FTIR), as a part of vibrational spectroscopy techniques, provides information on secondary structure and cross-linking, which is vital for understanding the formation and stability of protein-based and polysaccharide-based nanostructures [90]. For example, when combined with chemometric methods, FTIR effectively distinguishes gelatin sources by identifying unique spectral signatures in the amide and fingerprint regions, enabling rapid and robust authentication of bovine, porcine, and fish gelatins [91,92]. Another vibrational spectroscopy technique, Raman spectroscopy, detects vibrations of atoms in molecules but is better suited to samples containing an aqueous solution [93]. Nuclear magnetic resonance (NMR) provides high-resolution structural information on food gels, allowing for the identification of detailed analysis of composition, sequence, and substitution patterns [84]. UV/Vis also enables characterization of nano-/microstructures in food gels, such as tracking nanoparticle migration and separation in hydrogels, identifying specific particle morphologies via absorption peaks, and quantifying particle concentrations within gel matrices, as demonstrated by in situ fiber-based systems and studies on nanocomposite films and microgels [94,95]. Other spectroscopic techniques, like spectrophotometry, spectrofluorimetry, and inductively coupled plasma atomic emission spectroscopy (ICP-AES), offer high sensitivity and selectivity for detecting trace components and monitoring real-time changes in food gels, as demonstrated in the quantification of trace As(III) with low detection limits and wide linear ranges [31,96].

3.4. Rheology and Microrheology

Traditional rheology provides quantitative insights into the network assembly and viscoelastic behavior of hydrogels formed from various biopolymers, revealing how factors like cross-linking and chain architecture influence properties such as modulus, shear behavior, and thixotropy, which are critical for applications like 3D printing and ingredient delivery in foods [97,98]. Microrheology, using optical tweezers and dynamic light scattering, enables probing of local viscoelastic properties and spatial heterogeneity at the micro- and nanoscale, which is not accessible by bulk rheology [66,99]. For example, microrheological studies of iota–carrageenan gels revealed a concentration-dependent network structure, with increased entanglement density at higher concentrations limiting structural rearrangement and yielding behavior (Figure 4) [100]. In colloidal rod systems, microrheology and bulk rheology together showed that microstructure remains unchanged with increasing colloid concentration, but sol–gel transitions occur with depletant addition, highlighting the complementary nature of these methods for mapping gelation phase diagrams [101]. Additionally, microrheology can detect nanoscopic interfacial viscoelasticity, as seen in poly(ethylene oxide) hydrogels, where deviations between macro- and microrheological measurements revealed the presence of a viscoelastic interfacial shell around tracer particles [102].

3.5. Technical Challenges

Characterizing nano-/microstructures in food gels presents several key challenges, including the complexity and heterogeneity of gel networks, the need for advanced imaging and analytical techniques, and the sensitivity of gel structures to environmental and sample manipulation. The intricate interactions between biopolymer building blocks, such as proteins and polysaccharides, require precise quantification at both molecular and network levels, often necessitating the integration of super-resolution microscopy, fluorescence lifetime imaging, and nanomechanical methods to capture spatial and dynamic variations within gels.
Figure 4. Mesoscale rheology of iota–carrageenan (IC) gels at different strain rates using optical tweezers (OT). The measured force response (symbols) as a function of stage movement at different speeds. The dashed line represents the stage displacement (adapted from [100] with permission from Elsevier).
Figure 4. Mesoscale rheology of iota–carrageenan (IC) gels at different strain rates using optical tweezers (OT). The measured force response (symbols) as a function of stage movement at different speeds. The dashed line represents the stage displacement (adapted from [100] with permission from Elsevier).
Gels 11 00620 g004

4. Strategies for Tuning Nanostructure of Gels

4.1. Chemical Composition

The formulation of food gels (e.g., the type and concentration of building blocks and other additives, pH, and ionic strength) has a major impact on their nanostructure, gelation behavior, and functional properties. The common strategies for tuning nanostructures of gels, including mechanisms and impacts, are summarized in Table 1.

4.1.1. Concentration

Adjusting the concentration of key components such as nanoparticles, proteins, oils, or ions efficiently tunes the nanostructure and functional properties of food gels [15]. For example, increasing concentrations of cellulose nanocrystals (CNC) in whey protein isolate gels (up to 1.0% w/v) enhances WHC, gel strength, viscoelasticity, and thermal stability by promoting protein unfolding, cross-linking, and the formation of compact, homogeneous gel networks; however, excessive CNC can lead to agglomeration and moisture absorption effects [103]. Similarly, raising calcium ion concentration in composite gels or nanocellulose gels strengthens the gel network and viscoelastic properties by facilitating cross-linking, but too high a concentration (e.g., 0.20 M Ca2+) can cause irregular aggregation and reduced gel quality (Figure 5A) [68,104]. For Pickering emulsions stabilized by gelatin nanoparticles, higher nanoparticle concentrations (0.3–2.0 wt.%) result in smaller droplet sizes, increased viscosity, and more compact network structures, with optimal stability and viscoelasticity achieved at 1.5–2.0 wt.% [105]. In protein-based emulsion gels, increasing oil concentration (up to 20% soybean oil) improves gel strength, water-holding, and network formation, but further increases can reduce these benefits [106]. Nanofiber cellulose and faba bean protein isolate concentrations also modulate gel hardness, elasticity, and microstructure, with higher concentrations generally yielding stronger, more stable, and more printable gels, though excessive amounts may cause aggregation or inhomogeneity [107,108]. Metal ion concentration (e.g., Na+, K+, Ca2+, Al3+) similarly tunes gel microstructure and mechanical properties, with low concentrations (<0.1 M) favoring tighter, more homogeneous networks [109].

4.1.2. pH and Ionic Strength

Adjusting pH and ionic strength can control gel network density, mechanical strength, and nanostructure uniformity [113,114,115]. In kidney bean protein isolate sodium alginate gels, moderate acidification and higher Ca2+ concentrations led to improved gel properties and a honeycomb-like structure [114]. In wheat protein nanoparticle-stabilized emulsion gels, pH near the isoelectric point (~6.8) increased aggregation and viscosity, while higher salt concentrations further enhanced viscosity and elastic modulus, allowing for the design of stable, viscoelastic gels [116]. For sodium caseinate nanoparticles, higher ionic strength during enzymatic cross-linking produced larger particles, resulting in stiffer acid-induced gels with reduced syneresis [113]. Alginate–montmorillonite nanocomposite hydrogels maintained integrity across a wide pH range, with elastic moduli increasing linearly with clay content due to physical bonds, and thinner fibers forming at extreme pH values due to increased surface charge (Figure 5B) [110,117]. In pigeon pea protein gels, low pH and low ionic strength produced coarse networks with poor water-holding, while higher pH and ionic strength yielded compact, homogeneous matrices with better water retention, highlighting the importance of hydrogen bonding and electrostatic interactions in gel formation [115].

4.1.3. Co-Solutes/Sugars

In agarose-based fluid gels, increasing sucrose concentration alters the size, shape, and connectivity of microgel particles, directly impacting the gel’s viscoelasticity, texture, and lubrication by modifying the molecular interactions during gelation and shear processes [118]. For low methoxyl pectin gels, the addition of sucrose or erythritol enhances gel hardness, accelerates structuring, and increases cross-linking density, as confirmed by SEM and FTIR, with sucrose providing superior network density and physicochemical properties due to its strong hydration and chemical characteristics [119]. In gelatin-based confectionery gels, high concentrations of sweeteners raise the gelation and melting temperatures, but also weaken internal interactions and create smaller, more dispersed junction zones, while increasing gelatin concentration boosts gel hardness and tackiness [120]. For κ-carrageenan gels, sucrose as a co-solute, especially when combined with hydroxypropyl beta-cyclodextrin-curcumin composites, increases gel hardness, gumminess, and gelation temperature, and optimizes 3D printability by enhancing the gel structure formation rate [121]. Overall, sucrose’s multifaceted effects on gels stem from its ability to modulate water activity, molecular interactions, and network architecture, offering opportunities for tailored textural and functional properties in food and material applications.

4.1.4. Molecular Weight and Chain Flexibility

Low-molecular-weight gelators (LMWGs) self-assemble into diverse nanostructures through non-covalent interactions, enabling morphological diversity and responsiveness to external stimuli, but often result in gels that are fragile and difficult to manipulate due to limited structural strength and flexibility (Figure 3C) [111,122]. In contrast, increasing the molecular weight (e.g., cross-linking with high-molecular-weight agents like pullulan dialdehyde) can significantly enhance the network structure, leading to improved water vapor barrier properties, reduced solubility (from 42.3% to 18.7%), and increased tensile strength (up to 2.7 times higher than neat gelatin) [123]. High-molecular-weight fractions in gelatin also yield stronger and more stable gels, though the specific effects can vary with the source and processing method [124]. Chain flexibility, influenced by the nature of the gelator and cross-linking, allows for the tuning of gel morphology and mechanical properties, as well as the ability to respond to stimuli or external shaping methods [29,125]. Moreover, the solubility and cooperative self-assembly of LMWGs are critical in determining gelation thresholds and thermal stability, highlighting the importance of molecular design in achieving desired performance.

4.1.5. Blending of Components

Blending different components in food gels enables precise tuning of their properties through interactions such as covalent and non-covalent bonding, as well as co-assembly modes. For example, increasing κ-carrageenan concentration in gelatin gels accelerates gelation, enhances viscoelasticity, yield stress, and melting temperature, due to electrostatic complex formation and structural changes in the gel network [126]. Furthermore, incorporating dialdehyde κ-carrageenan into gelatin films via covalent cross-linking significantly reduced solubility, moisture content, and water vapor permeability, while increasing tensile strength by approximately 20-fold compared to pure gelatin films, and the addition of zein nanoparticles further enhanced antioxidant and antimicrobial activity when loaded with thymol [127]. In chitosan–gelatin composite films, the blending method improved elongation and swelling, while layer-by-layer assembly enhanced water vapor barrier, tensile strength, and thermal stability, demonstrating that the method of blending or assembly directly affects nanostructure and performance [128,129]. Multi-component organogels and bigels exhibit synergistic interactions between gelators, resulting in higher hardness and moduli than single-component systems, and the inclusion of colloidal particles further increases stability and mechanical strength (Figure 3D) [112]. The nanostructure and macroscopic properties of two-component dendritic gels can be modulated by adjusting component concentration, molecular structure, or molar ratio, leading to changes in gel morphology and transition temperatures [130,131,132]. Additionally, the choice of solvent in multi-component gels can drive different self-assembly modes and nanostructures, highlighting the importance of environmental factors in tuning gel properties [133].

4.2. Physical Control

Processing controls (e.g., temperature, shear/flow, pressure, and ultrasound) tune properties of food gels by influencing crystallization, alignment, gelation, and nano-emulsion formation.

4.2.1. Temperature Control

In starch-based emulsion-filled gels, lower gelation temperatures (around 55 °C) promote the leaching of starch molecules, forming active networks, while moderate increases (to 60 °C) enhance gel strength due to swollen granules; however, higher temperatures can weaken the gel by causing granule disintegration, as shown in Figure 6A–C [134]. In nano-emulsion gels, sequential temperature changes can restructure the gel, allowing for precise control over microstructure and rheological properties by modulating interparticle potentials and triggering self-assembly [135,136]. For protein–starch composite gels, preheating above 60 °C alters protein conformation and promotes protein–starch complex formation, resulting in softer textures and changes in digestibility, while lower preheating temperatures yield harder, bicontinuous networks [137]. These findings demonstrate that careful temperature management enables the design of food gels with tailored nanostructures and functional properties for specific applications.

4.2.2. Shear/Flow Fields

High shear rates or large strain amplitudes can fully break down gel networks, resulting in more homogeneous and stronger gels upon cessation of flow, while lower shear rates or strain amplitudes tend to create heterogeneous, weaker gels with reduced elasticity [138,139]. For example, in colloidal and flaxseed fiber gels, increasing homogenization or shear rates weakens the network structure (lower viscoelastic moduli) but can improve physical stability, making them suitable as food stabilizers [140]. Oscillatory shear is particularly effective at tuning gel properties, with large amplitudes producing single-yielding, strong gels and intermediate amplitudes leading to weak, cluster-dense structures [141]. Shear history also induces memory effects, where the microstructure and elasticity of the final gel depend on the specific shear protocol applied, such as the critical shear rate in boehmite gels that determines whether the gel exhibits glassy or gel-like viscoelastic spectra [142]. Additionally, the presence of co-solutes (e.g., sucrose) and the order of ingredient addition can further modulate the microgel network and rheological behavior under shear, impacting texture and lubrication properties [14,142].

4.2.3. Pressure

Elevated pressures can lower the minimum concentration required for gel formation, as seen with pea protein and oat β-glucan gels, where higher pressures (e.g., 600 MPa for pea protein, 500 MPa for β-glucan) enable gelation at lower concentrations and enhance gel strength and textural properties [143]. Pressure also allows for the fine-tuning of gel microstructure and mechanical properties by adjusting parameters like pH, pressure level, and treatment duration, resulting in gels with diverse textures and improved water-holding or freeze-thaw stability [144]. In nano-emulsions stabilized by plant proteins, high-pressure homogenization (HPH) controls droplet size, which inversely affects gel strength (i.e., smaller droplets yield stronger gels) [145]. For gelatin, HPH disrupts hydrogen bonds and unfolds proteins, increasing emulsifying and foaming capacities but reducing gel strength and elasticity [146]. Pressure-induced gels often differ from heat-induced ones, typically exhibiting less hardness and adhesiveness but better retention of thermolabile compounds and unique textural profiles [147]. Additionally, pressure can be used to create stable Pickering emulsion gels with enhanced encapsulation and stability of bioactive compounds like curcumin [148].

4.2.4. Ultrasound

Ultrasound treatment enables the fine-tuning of gel strength/hardness, microstructure, and WHC by influencing protein unfolding/aggregation and spatial distribution of other biopolymers (Figure 6D,E) [149,150]. In peanut protein systems, high-intensity ultrasound decreased nanoparticle size and enhanced heat-set gelling and emulsifying properties, especially at higher protein concentrations [151]. Ultrasound-assisted extraction of soy protein isolate led to a substantial increase in extraction yield (from 24.68% to 42.25%) and improved gel mechanical properties, uniformity, and water retention [152]. For starch-based gels, ultrasound induced cracks and pores in granules, increasing extraction yield by 14.55% and improving thermal stability and gel matrix strength [153]. In nanocellulose hydrogels, ultrasound reduced interfibrillar distances, resulting in stronger gel strength and improved emulsion stability (Figure 6D,E) [149].
Table 1. Summary of strategies for tuning nanostructures of gels.
Table 1. Summary of strategies for tuning nanostructures of gels.
StrategyMechanismExamplesImpact on NanostructureRef.
ConcentrationIncreased concentration enhances molecular proximityAlginate gels; κ-carrageenan gelsPromotes more frequent junction zone formation, leading to denser and stronger networks[154,155]
pH adjustmentModulates charge distribution and gelation behavior of biopolymersAcid/alkaline-induced casein, pea, or pectin gelsAlters network density, pore size, and aggregation behavior[156,157,158]
Ionic strength and Ion typeIonic cross-linking or shielding modulates gelation and structureCa2+-induced alginate gel; K+-induced κ-carrageenan gel; Na+ effect on protein gelsControls gel stiffness, porosity, and nano-fibrillar structure[159,160,161,162,163]
Solvent Quality/PolarityAffects molecular interactions and phase separationEthanol or sugar concentration to promote gelationChanges gel network compactness and aggregation state[164,165,166]
Co-gelling or composite systemsCombines multiple gelling agents or nanofibrils to form hybrid structuresAlginate–gellan gum, protein–polysaccharide blendsEnhances hierarchical structure and multi-scale network architecture[167,168,169,170,171]
Thermal treatmentInduces denaturation or conformational changes that promote gelationAlginate gels; Heat-set whey protein gels; protein–polysaccharide gels; gelatin melting and reformationModulates fibril size, network junctions, and WHC[172,173,174,175]
Shear processingAligns, disrupts, or restructures gel network during processingHomogenization, extrusion, or whipping of gelsControls anisotropy, fibrillar orientation, and nanopore structure[176,177,178]
PressureInduces protein unfolding and aggregation via non-thermal meansHigh-pressure-treated starch or protein gelsIncreases WHC and creates uniform, dense nanostructures[143,144,147,179]
UltrasoundImproves interaction between phases in composite gelsProtein, protein–polysaccharide, or emulsion-filled gelsPromotes finer emulsions and more uniform microphase distribution at the nanoscale[39,180,181,182]

4.3. Engineering Multifunctional Gel Matrices Through Nanostructure Integration

Nanostructures (e.g., cellulose nanocrystals, protein nanofibrils, and nanoparticles) interact with gelators through covalent or non-covalent bonds, restrict water mobility, facilitate protein unfolding and cross-linking, and endow food gels with multifunctionality, like antibacterial, antioxidant, and intelligent responsiveness.

4.3.1. Modulating Gel Properties via Nanoparticles as Functional Modifiers

Nanoparticles tailor the properties of food gels through various mechanisms such as interfacial adsorption, network formation, and molecular interactions. For example, combining soy protein-pectin complex nanoparticles with glycyrrhizic acid nanofibrils leads to smaller emulsion droplet sizes and stronger, more thixotropic gels due to hydrogen bonding and the formation of fibrillar hydrogel networks around droplets (Figure 7A) [183]. Shellac nanoparticle–chitosan complexes, stabilized by hydrogen bonding and electrostatic interactions, serve as both emulsifiers and gelling agents, maintaining gel strength and improving bio-accessibility of encapsulated nutrients like β-carotene [184,185]. The addition of nanoparticles like NiO, ZnO, and modified κ-carrageenan also boosts the mechanical strength, barrier properties, and thermal stability of biopolymer-based films, with tensile strength increases up to 20-fold and improved antibacterial and antioxidant activities, making them ideal for active food packaging [127,186]. Starch-based and gelatin nanoparticles further contribute by enhancing film and gel stability, enabling controlled release of active ingredients, and providing responsiveness to environmental stimuli [187,188].

4.3.2. Reinforcing Gel Networks Through Nanofiber-Based Structural Integration

Nanofibers (e.g., cellulose nanofibers and protein-based nanofibers) enhance the mechanical and rheological properties of food gels by reinforcing their 3D network structures and modulating interactions at the molecular level. The addition of nanofiber cellulose to gelatin gels increases hardness and viscosity, transforms rheological behavior from Newtonian to pseudoplastic, and raises the sol–gel transition temperature [108]. Whey protein nanofibers incorporated into protein gels boost gel strength, WHC, and storage modulus, while also shortening gelation time and creating highly porous microstructures (Figure 7B) [189]. The gelation of nanofiber-based gels is primarily driven by electrostatic interactions, fiber entanglement [190]. However, in some systems, ion-induced conformational changes also promote protein unfolding and aggregation into stable networks. For example, increasing calcium ions in cellulose nanocrystal–whey protein gels enables gel formation by promoting β-sheet formation and hydrophobic interactions [68]. Additionally, nanofiber incorporation can improve cold water solubility, emulsion stability, and foam-forming ability in gelatin gels, broadening their functional applications [191].

4.3.3. Safety and Regulatory Concerns of Nanomaterials in Food Products

Involving nanomaterials in food gels raises significant safety and regulatory concerns due to their small size and high surface area, which enable them to cross biological barriers and potentially induce toxic effects such as oxidative stress, genotoxicity, and inflammation in human cells [192]. Migration of nanoparticles from packaging or other gel products into food is a key risk, with studies showing that materials like silver, titanium dioxide, and zinc oxide nanoparticles can leach under certain conditions, leading to human exposure [193]. Regulatory frameworks remain inconsistent globally, with some countries lacking strict controls, and there is a pressing need for comprehensive legislation and standardized risk assessment methods to ensure consumer safety [194,195]. Several reports highlight that nanoparticles fully embedded in polymer matrices are less likely to migrate [196], but surface alterations or mechanical stress can increase release, and the lack of long-term toxicity data continues to hinder regulatory progress.
Figure 6. (A) Visual appearance of starch emulsion-filled gels formed by 0–90 °C gelation. Red arrows show the edges of the emulsion-filled gels under loading. (B) Storage (G’) and loss (G”) modulus, and (C) tan δ as a function of frequency for starch emulsion-filled gels formed by different gelation temperatures (55–90 °C) (adapted from [134] with permission from Elsevier). (D) The appearance of cellulose hydrogels obtained from different ultrasonic durations (left) and the reversible property of hydrogels (right). (E) The morphology of nanocellulose particles from different hydrogels characterized by atomic force microscopy (AFM) (adapted from [149] with permission from Elsevier). The scale is 1 μm.
Figure 6. (A) Visual appearance of starch emulsion-filled gels formed by 0–90 °C gelation. Red arrows show the edges of the emulsion-filled gels under loading. (B) Storage (G’) and loss (G”) modulus, and (C) tan δ as a function of frequency for starch emulsion-filled gels formed by different gelation temperatures (55–90 °C) (adapted from [134] with permission from Elsevier). (D) The appearance of cellulose hydrogels obtained from different ultrasonic durations (left) and the reversible property of hydrogels (right). (E) The morphology of nanocellulose particles from different hydrogels characterized by atomic force microscopy (AFM) (adapted from [149] with permission from Elsevier). The scale is 1 μm.
Gels 11 00620 g006

5. Nanostructural Modulation for Function Control in Food Gels

To develop food gels for multiple applications, we need to understand how the structure at the nanoscale level influences their functionalities, which could be concluded into four aspects (i.e., mechanical/textural properties, WHC, optical properties, and mass transport/release).

5.1. Texture Customization Through Controlled Mechanics of Nanostructures

The mechanical and textural properties of food gels, such as gel strength, elasticity, adhesiveness, and breakdown behavior, are crucial for food quality and sensory perception [15,197]. Nanostructures modulate these properties by reinforcing gel matrices, improving elasticity, ductility, and water resistance. For example, incorporating nanofibers or nanoparticles such as ZnO into biopolymer-based gels increases Young’s modulus, tensile strength, and elongation at break, resulting in stronger gels suitable for food packaging and high-moisture foods [198,199]. Oxidized cellulose nanofibrils, even at low concentrations, synergistically improve surimi gel whiteness, gel strength, and thermal stability through matrix reinforcement and water binding [200]. The formation of rod-like junction zones and point-like cross-links in pectin-calcium gels directly influences their mechanical strength and swelling behavior [83]. Similarly, protein-based nanostructures and nano-emulsion-based gels allow for tailored texture, such as consistency and shear-thinning behavior, by modulating the type and concentration of gelling agents [90,201]. These advances are particularly valuable for developing foods with specific textural requirements, such as those for the senior population or individuals with dysphagia, where both safety and palatability are critical.

5.2. Controlling WHC and Stability via Nanostructure-Mediated Water Confinement

Nanostructures (usually hydrophilic cellulose and protein nanofibers) can reinforce gel networks, promote water entrapment, improve gel strength and viscoelasticity, and are essential for WHC and stability of food gels. Mechanistically, these nanostructures act as active fillers, restrict water mobility, induce favorable protein conformational changes, and strengthen the gel matrix via hydrophobic, hydrogen, and disulfide bonds [103,189]. For example, adding cellulose nanocrystals to whey protein isolate gels increased WHC, gel strength, and thermal stability by facilitating protein unfolding, cross-linking, and forming compact, homogeneous structures, with WHC rising as CNC concentration increased up to 1% w/v [103]. Similarly, oxidized cellulose nanofibrils at low concentrations greatly improved surimi gel WHC through matrix reinforcement, water binding [200]. However, in Pickering emulsion gels, gelatin or xanthan gum/lysozyme nanoparticles also increased WHC by enhancing network density and droplet surface coverage [202].

5.3. Modulating Light Transmission, Scattering, and Absorption via Nanostructured Matrices

Optical properties of food gels can be influenced by various factors, like compositions, structures, and environmental conditions, enabling improvement in food packaging and sensory perception. The size and distribution of nanostructures within gels influence light scattering, affecting transparency, opacity, and color. For example, incorporation of Co-doped ZnO nanoparticles into chitosan/gelatin gels reduces the optical energy bandgap (from 5.22 to 4.90 eV direct, and 4.92 to 4.42 eV indirect), enhancing optical, dielectric, and antimicrobial properties [203]. Additionally, the supramolecular ordering and chiral nano-structuring in chitosan-aspartate glycerohydrogels result in wavelength-dependent transmittance and scattering, and the novel effect of optical clearing in the near-UV region, which is influenced by the size and shape of nanodomains [204]. Moreover, nanostructural inhomogeneity in polymer network gels, often arising from variations in polymer-segmental and cross-linking densities, affects optical clarity and performance, with more homogeneous nanostructures leading to improved transparency and functional properties [205].

5.4. Programming Bioactive Release Through Engineered Nanostructural Barriers/Diffusion

Controlling the oral delivery and bioavailability of active compounds is a promising function of food gels and also a hotspot recently, which emphasizes the health and wellness properties of products [206]. Nanostructures (e.g., cellulose nanocrystals, protein nanoparticles, and nanofibrils) enhance this function by modifying gel microstructure, increasing gel strength, and improving WHC [103,207]. However, it remains unclear how gel strength and WHC of various structures are mechanistically related to the controlled release of encapsulated compounds. For example, protein nanostructures and nanohydrogels (e.g., lactoferrin nanohydrogels) provide high encapsulation efficiency (up to 90%) and allow for controlled, matrix-dependent release of bioactives like curcumin, with release rates varying based on the hydrophilic or lipophilic nature of the surrounding medium [90,206]. Microemulsions with nanostructured matrices also show that changes in viscosity and microstructure can shift release kinetics from rapid to sustained, with systems containing thickeners like carboxymethyl cellulose exhibiting matrix-driven, slower release profiles [208].
Figure 7. (A) Schematic diagram of emulsion and emulsion gels prepared by soy protein isolate (SPI)-pectin complex nanoparticles and glycyrrhizic acid (GA) nanofibrils (adapted with permission from [183]. Copyright 2020 American Chemical Society). (B) Mechanism diagram of whey protein isolate (WPI) composite gel strengthened by whey protein isolate fibrils (WPIF) with different scales at different pH (pH 2.0 and pH 7.0) (adapted from [189] with permission from Elsevier).
Figure 7. (A) Schematic diagram of emulsion and emulsion gels prepared by soy protein isolate (SPI)-pectin complex nanoparticles and glycyrrhizic acid (GA) nanofibrils (adapted with permission from [183]. Copyright 2020 American Chemical Society). (B) Mechanism diagram of whey protein isolate (WPI) composite gel strengthened by whey protein isolate fibrils (WPIF) with different scales at different pH (pH 2.0 and pH 7.0) (adapted from [189] with permission from Elsevier).
Gels 11 00620 g007

6. Applications Enabled by Tuning Nanostructure of Food Gels

Tuning nanostructures of food gels has emerged as a powerful tool to achieve specific applications, such as tailored texture, controlled release, 3D/4D printing adaptation, fat reduction, smart packaging solutions, etc., as shown in Figure 8. The successfully developed applications of food gels are summarized in Table 2.

6.1. Tailored Texture and Sensory Perception

Recent works have demonstrated that tuning the nano-/microstructure of gels, through ingredient selection, processing, and the use of biopolymers or nanofibers, enables precise control over texture and sensory perception. For example, combining plant proteins with various polysaccharides (e.g., gellan gum, carrageenan, and pectin) allows the creation of gels with customizable firmness, spreadability, and crumbliness, suitable for applications like meat analogs and cheeses [228]. However, the addition of nanofiber cellulose to gelatin gels reduces elasticity and also alters rheological properties such as viscosity and sol–gel transition temperature [108]. Emulsion gels, which incorporate dispersed phases into gel matrices, offer further flexibility in designing breakdown behavior and sensory attributes, supporting applications in fat replacement [2]. Additionally, the interplay between texture and flavor/aroma release is significant: hydrocolloid-based texture modifications can alter flavor release profiles, while specific aromas (e.g., limonene, menthol) can modulate perceived firmness, stickiness, and creaminess [229].

6.2. Encapsulation and Delivery of Bioactives

To apply food gels as encapsulation and delivery vehicles for bioactives, recent advances have focused on optimizing gels’ nanostructures, such as nanoorganogels, nanostructured proteins, lipid-based nanocarriers, and microgels. Nanoorganogels, for example, improve the bioavailability and shelf-life of nutraceuticals by preventing degradation and enabling controlled release [230]. Nanoparticles, nanogels, and nanofibers offer pH-responsive and tunable release profiles, with encapsulation efficiency and release behavior dependent on nanostructure type and physiological conditions [231]. Furthermore, to deliver multiple bioactives, co-encapsulation systems and Pickering emulsions have been developed using protein, polysaccharide, or lipid-based assemblies to accommodate compounds with varying solubility [185,232]. Microfluidic fabrication of microgels also allows precise control over gel size and encapsulation efficiency (30–40%), enabling prolonged release and low cytotoxicity for food and pharmaceutical applications [233]. Summarily, the integration of advanced gel nanostructures, co-encapsulation strategies, and microfluidic technologies is driving the development of next-generation food delivery systems, with a focus on enhancing bioactive stability, targeted release, and functional food innovation.

6.3. Fat Reduction and Calorie Control

Food gels are effective fat replacers for calorie control, and modifying their nanostructure helps achieve optimal sensory and functional qualities. For example, composite gels made from citrus insoluble nanofibers and amylose at a 1:4 ratio closely mimic the creamy texture of fat and significantly inhibit lipid digestion [234]. Similarly, low-fat Pickering emulsion gels stabilized by zein/phytic acid nanoparticles form dense barriers around oil droplets, preventing coalescence and improving the stability of encapsulated curcumin [235]. Other innovations include composite gels made from konjac glucomannan and nano-citrus fiber, which can replicate the texture and stability of whipped animal cream [236]. Additionally, nanocellulose additives have been shown to reduce triglyceride hydrolysis by up to 48.4% in vitro and lower postprandial serum triglycerides by 36% in vivo, primarily by reducing lipase access and sequestering bile salts [237].

6.4. Food Packaging with Improved Shelf-Life and Smart Responsiveness

Precisely tuning the nanostructures of food gels enhances gel-based packaging performance, such as extended food shelf-life, smart freshness, and microbiological detection. This improvement is consistently achieved by incorporating nanomaterials like metal nanoparticles, nanofibers, or nanohybrids, which enhance antimicrobial activity, mechanical strength, and smart responsiveness. Key examples include: gelatin-based films incorporating zinc metal-organic frameworks (Zn-MOFs) achieved up to 100% antioxidant activity and exhibited strong antibacterial effects, significantly delaying texture deterioration and inhibiting microbial growth in cherry tomatoes during 16 days of storage [238]. Similarly, nano-emulsions incorporated into packaging systems have been shown to prevent microbial growth, reduce oxidation, and browning [239]. Moreover, tuning nanostructures not only improves mechanical and barrier properties but also enables real-time monitoring of food quality through smart features like sensors and pH-responsive tags, which can indicate spoilage or freshness [240,241]. Usually, these innovations, often based on biopolymers and sustainable materials, integrate antioxidant, antimicrobial, and freshness-indicating properties, although challenges remain regarding scalability, cost, and regulatory acceptance [21,242]. Moreover, the balance between nanostructure functionality and safety must be carefully managed, with future research needed to address long-term health and environmental impacts.

6.5. Novel Food Structures and 3D Printing

By optimizing the rheological and mechanical properties of gels (e.g., starch, protein, and Pickering emulsions), researchers have achieved improved shape accuracy, structural integrity, and sensory qualities in printed foods, as seen with sorghum protein gels at 25% concentration and optimized nozzle size [243,244]. The incorporation of innovative additives like lipids and carbohydrates further tailors gel properties, improving their printability and sensory attributes [245]. Strategies such as reinforcing emulsion gel interfaces, cross-linking, and employing 3D or coaxial printing for pH-responsive hydrogels have enabled the creation of personalized, texture-modified foods tailored for individuals with dysphagia or specific dietary needs, improving both nutritional value and swallowing safety [223,246]. Moreover, the interaction between gel components and environmental factors (e.g., ionic strength, temperature, pH) has been shown to influence the final 3D print precision, as demonstrated by the enhancement of cassava starch gel with casein protein, which increased molecular weight and elastic modulus, leading to higher printing accuracy [247]. Summarily, the predictability of print ability can be quantitatively linked to specific rheological parameters, such as phase angle and gel network density, allowing for more rational design of food inks.

7. Challenges and Perspectives

7.1. Current Challenges

Owing to their tunable properties and compatibility with natural ingredients, food gels find wide application in texture design, fat replacement, nutrient delivery, 3D printing, and food packaging. Nanostructure tuning is a crucial strategy for precisely optimizing gel properties, enabling the development of gels with tailored properties. However, challenges remain, including the complexity of multicomponent systems, difficulties in achieving precise structural control and predictability, and the need to balance mechanical performance with durability and stability. Additionally, integrating diverse functional nanomaterials without compromising gel integrity, scaling up manufacturing processes, and meeting application-specific requirements such as biocompatibility and environmental adaptability remain major hurdles. The addition of solid nanoparticles (e.g., glass, titanium dioxide) to food gels also may pose health risks by inducing toxicity, disrupting nutrient absorption, or causing oxidative stress, especially if their interactions with food components are not well understood or controlled. These issues are compounded by limited understanding of nanostructure–property relationships and evolving safety and regulatory concerns, underscoring the necessity for continued research to fully realize the potential of tuning nanostructures of food gels.

7.2. Future Directions

The future of food gels is marked by rapid advancements in nanostructure tuning technologies. Emerging trends include the use of natural, plant-based nanomaterials, double-network and hybrid gel systems, and 3D/4D printing for personalized nutrition and food design. Challenges remain in scaling up production, ensuring safety and consumer acceptance, and fully understanding the complex interactions at the nanoscale, but ongoing research is expected to yield sustainable, multifunctional food gels with tailored properties for health, sensory, and environmental benefits. Future research is expected to solve the following open research questions: (1) How can nanostructured food gels be safely scaled up for industrial production? (2) What are the long-term health and safety implications of consuming nanostructured food gels? (3) How can 3D/4D printing and plant-based nanomaterials be optimized for personalized nutrition? In summary, tuning food gels’ nanostructure is poised to revolutionize food design for multiple applications, but further research is needed to address key challenges and unlock their full potential.

8. Conclusions

This review has examined the significance of nanostructures in food gels, emphasizing how precise tuning can lead to enhanced properties and novel applications in the food industry. By manipulating the building blocks, gelation mechanisms, and processing conditions, researchers can control the nanoscale architecture of food gels, thereby tailoring their mechanical, textural, and functional characteristics. The integration of advanced characterization techniques has been pivotal in understanding and optimizing these nanostructures, enabling the development of gels with specific attributes such as improved WHC, controlled release of bioactives, and customized sensory profiles. Applications ranging from texture modification and fat reduction to 3D food printing and smart pack aging illustrate the versatility and potential of nanostructured food gels. However, several challenges persist, including the complexity of multicomponent systems, the need for precise structural control, and the scalability of production processes. Additionally, ensuring the safety and regulatory compliance of nanostructured materials in food products remains a critical concern. Future research should focus on overcoming these challenges by developing innovative methods for nanostructure manipulation, enhancing the understanding of structure-property relationships, and exploring sustainable and biocompatible materials. Interdisciplinary collaboration will be essential in driving forward the field, fostering the creation of next-generation food products that meet consumer demands for quality, health, and sustainability. In conclusion, tuning the nanostructures of food gels represents a frontier in food science and engineering, offering unprecedented opportunities for innovation and improvement in food design and functionality. Continued exploration and advancement in this area will undoubtedly contribute to the evolution of the food industry, addressing both current and future challenges.

Author Contributions

Conceptualization, T.Y., L.C. and A.G.S.; methodology, T.Y. and L.C.; software, T.Y. and L.C.; investigation, J.S.; resources, A.G.S.; data curation, T.Y. and J.S.; writing—original draft preparation, T.Y. and L.C.; writing—review and editing, J.S. and A.G.S.; visualization, L.C.; supervision, A.G.S.; project administration, A.G.S.; funding acquisition, A.G.S. All authors have read and agreed to the published version of the manuscript.

Funding

We would like to thank BOF-UGent (BOF/IOP/2022/033, BOF23/GOA/029, 01SC2621, and 01SC1420). We also acknowledge the support received from the Research Foundation Flanders (FWO) through grants G043322N, G076224N, FWO-F.N.R.S. under the Excellence of Science (EOS) program (grant number: 40007488), and FWO-SBO (S001423N). We also would like to thank the financial support from the China Scholarship Council (No. 202406150021, No. 202106910013, and No. 202006150025).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
WHCWater-Holding Capacity
3DThree-Dimensional
IDDSIInternational Dysphagia Diet Standardization Initiative
Cryo-SEMCryo-Scanning Electron Microscopy
TEMTransmission Electron Microscopy
AFMAtomic Force Microscopy
STEDStimulated Emission Depletion
FLIMFluorescence Lifetime Imaging Microscopy
LMLight Microscopy
PLMPolarized Light Microscopy
SAXSSmall-Angle X-Ray Scattering
SANSSmall-Angle Neutron Scattering
USANSUltra-Small-Angle Neutron Scattering
FTIRFourier-Transform Infrared Spectroscopy
NMRNuclear Magnetic Resonance
LMWGsLow-Molecular-Weight Gelators
ICP-ASEInductively Coupled Plasma Atomic Emission Spectroscopy
CNCCellulose Nanocrystals
HPHHigh-Pressure Homogenization

References

  1. Liu, B.; Yang, H.; Zhu, C.; Xiao, J.; Cao, H.; Simal-Gandara, J.; Li, Y.; Fan, D.; Deng, J. A comprehensive review of food gels: Formation mechanisms, functions, applications, and challenges. Crit. Rev. Food Sci. Nutr. 2024, 64, 760–782. [Google Scholar] [CrossRef]
  2. Abdullah; Liu, L.; Javed, H.U.; Xiao, J. Engineering Emulsion Gels as Functional Colloids Emphasizing Food Applications: A Review. Front. Nutr. 2022, 9, 890188. [Google Scholar] [CrossRef]
  3. Dickinson, E. Emulsion gels: The structuring of soft solids with protein-stabilized oil droplets. Food Hydrocoll. 2012, 28, 224–241. [Google Scholar] [CrossRef]
  4. Fu, K.; Wang, H.; Pan, T.; Cai, Z.; Yang, Z.; Liu, D.; Wang, W. Gel-forming polysaccharides of traditional gel-like foods: Sources, structure, gelling mechanism, and advanced applications. Food Res. Int. 2024, 198, 115329. [Google Scholar] [CrossRef]
  5. Yan, J.; Zhang, Z.; Lai, B.; Wang, C.; Wu, H. Recent advances in marine-derived protein/polysaccharide hydrogels: Classification, fabrication, characterization, mechanism and food applications. Trends Food Sci. Technol. 2024, 151, 104637. [Google Scholar] [CrossRef]
  6. Klein, M.; Poverenov, E. Natural biopolymer-based hydrogels for use in food and agriculture. J. Sci. Food Agric. 2020, 100, 2337–2347. [Google Scholar] [CrossRef]
  7. Miao, W.; Jiang, H.; Li, X.; Sang, S.; Jiang, L.; Lin, Q.; Zhang, Z.; Chen, L.; Long, J.; Jiao, A.; et al. Recent advances in natural gums as additives to help the construction and application of edible biopolymer gels: The example of hydrogels and oleogels. Crit. Rev. Food Sci. Nutr. 2024, 64, 12702–12719. [Google Scholar] [CrossRef]
  8. Chen, L.; Lin, S.; Sun, N. Food gel-based systems for efficient delivery of bioactive ingredients: Design to application. Crit. Rev. Food Sci. Nutr. 2023, 64, 13193–13211. [Google Scholar] [CrossRef]
  9. Francavilla, A.; Corradini, M.G.; Joye, I.J. Bigels as Delivery Systems: Potential Uses and Applicability in Food. Gels 2023, 9, 648. [Google Scholar] [CrossRef] [PubMed]
  10. Mao, L.; Lu, Y.; Cui, M.; Miao, S.; Gao, Y. Design of gel structures in water and oil phases for improved delivery of bioactive food ingredients. Crit. Rev. Food Sci. Nutr. 2020, 60, 1651–1666. [Google Scholar] [CrossRef] [PubMed]
  11. Faber, T.J.; Van Breemen, L.C.A.; McKinley, G.H. From firm to fluid—Structure-texture relations of filled gels probed under Large Amplitude Oscillatory Shear. J. Food Eng. 2017, 210, 1–18. [Google Scholar] [CrossRef]
  12. Min, C.; Ma, W.; Kuang, J.; Huang, J.; Xiong, Y.L. Textural properties, microstructure and digestibility of mungbean starch–flaxseed protein composite gels. Food Hydrocoll. 2022, 126, 107482. [Google Scholar] [CrossRef]
  13. Macias-Rodriguez, B.A.; Gouzy, R.; Coulais, C.; Velikov, K.P. Thermoresponsive oil-continuous gels based on double-interpenetrating colloidal-particle networks. Soft Matter 2024, 20, 3033–3043. [Google Scholar] [CrossRef]
  14. Li, A.; Gong, T.; Yang, X.; Guo, Y. Interpenetrating network gels with tunable physical properties: Glucono-δ-lactone induced gelation of mixed Alg/gellan sol systems. Int. J. Biol. Macromol. 2020, 151, 257–267. [Google Scholar] [CrossRef]
  15. Cao, Y.; Mezzenga, R. Design principles of food gels. Nat. Food 2020, 1, 106–118. [Google Scholar] [CrossRef]
  16. Wang, C.; Yan, R.; Li, X.; Sang, S.; McClements, D.J.; Chen, L.; Long, J.; Jiao, A.; Wang, J.; Qiu, C.; et al. Development of emulsion-based edible inks for 3D printing applications: Pickering emulsion gels. Food Hydrocoll. 2023, 138, 108482. [Google Scholar] [CrossRef]
  17. Edwards, C.E.R.; Mai, D.J.; Tang, S.; Olsen, B.D. Molecular anisotropy and rearrangement as mechanisms of toughness and extensibility in entangled physical gels. Phys. Rev. Mater. 2020, 4, 015602. [Google Scholar] [CrossRef]
  18. Foegeding, E.A. Food Biophysics of Protein Gels: A Challenge of Nano and Macroscopic Proportions. Food Biophys. 2006, 1, 41–50. [Google Scholar] [CrossRef]
  19. Vecchio, D.A.; Hammig, M.D.; Xiao, X.; Saha, A.; Bogdan, P.; Kotov, N.A. Spanning Network Gels from Nanoparticles and Graph Theoretical Analysis of Their Structure and Properties. Adv. Mater. 2022, 34, 2201313. [Google Scholar] [CrossRef]
  20. Datta, S.; Bhattacharya, S. Multifarious facets of sugar-derived molecular gels: Molecular features, mechanisms of self-assembly and emerging applications. Chem. Soc. Rev. 2015, 44, 5596–5637. [Google Scholar] [CrossRef]
  21. Yang, T.; Skirtach, A.G. Nanoarchitectonics of Sustainable Food Packaging: Materials, Methods, and Environmental Factors, Materials. Materials 2025, 18, 1167. [Google Scholar] [CrossRef]
  22. Martins, J.T.; Bourbon, A.I.; Pinheiro, A.C.; Fasolin, L.H.; Vicente, A.A. Protein-Based Structures for Food Applications: From Macro to Nanoscale. Food Syst. 2018, 2, 77. [Google Scholar] [CrossRef]
  23. Zeng, Y.; Yang, T.; Liu, Y.; Li, B.; Li, L.; Zhang, X. Enhanced foamability and stability of aqueous foams through novel pickering fat globules formulated with solid lipid particles. Food Hydrocoll. 2024, 150, 109684. [Google Scholar] [CrossRef]
  24. Rusch, P.; Zámbó, D.; Bigall, N.C. Control over Structure and Properties in Nanocrystal Aerogels at the Nano-, Micro-, and Macroscale. Accounts Chem. Res. 2020, 53, 2414–2424. [Google Scholar] [CrossRef]
  25. Cao, L.; Huang, Y.; Parakhonskiy, B.; Skirtach, A.G. Nanoarchitectonics beyond perfect order—Not quite perfect but quite useful. Nanoscale 2022, 14, 15964–16002. [Google Scholar] [CrossRef] [PubMed]
  26. Sherman, Z.M.; Green, A.M.; Howard, M.P.; Anslyn, E.V.; Truskett, T.M.; Milliron, D.J. Colloidal Nanocrystal Gels from Thermodynamic Principles. Accounts Chem. Res. 2021, 54, 798–807. [Google Scholar] [CrossRef] [PubMed]
  27. Ossai, C.I.; Raghavan, N. Nanostructure and nanomaterial characterization, growth mechanisms, and applications, Nanotechnology Reviews. Nanotechnol. Rev. 2018, 7, 209–231. [Google Scholar] [CrossRef]
  28. Li, X.; Lou, L.; Song, W.; Zhang, Q.; Huang, G.; Hua, Y.; Zhang, H.-T.; Xiao, J.; Wen, B.; Zhang, X. Controllably Manipulating Three-Dimensional Hybrid Nanostructures for Bulk Nanocomposites with Large Energy Products. Nano Lett. 2017, 17, 2985–2993. [Google Scholar] [CrossRef]
  29. Zhang, L.; Wang, X.; Wang, T.; Liu, M. Tuning Soft Nanostructures in Self-assembled Supramolecular Gels: From Morphology Control to Morphology-Dependent Functions. Small 2015, 11, 1025–1038. [Google Scholar] [CrossRef] [PubMed]
  30. Mao, X.; Liu, Y.; Qiao, C.; Sun, Y.; Zhao, Z.; Liu, J.; Zhu, L.; Zeng, H. Nano-fibrous biopolymers as building blocks for gel networks: Interactions, characterization, and applications. Adv. Colloid Interface Sci. 2025, 338, 103398. [Google Scholar] [CrossRef]
  31. Zhang, Q.; Zhou, Y.; Yue, W.; Qin, W.; Dong, H.; Vasanthan, T. Nanostructures of protein-polysaccharide complexes or conjugates for encapsulation of bioactive compounds. Trends Food Sci. Technol. 2021, 109, 169–196. [Google Scholar] [CrossRef]
  32. Tan, Y.; Zi, Y.; Peng, J.; Shi, C.; Zheng, Y.; Zhong, J. Gelatin as a bioactive nanodelivery system for functional food applications. Food Chem. 2023, 423, 136265. [Google Scholar] [CrossRef] [PubMed]
  33. Nishinari, K. Gelling Properties. In Food Hydrocolloids: Functionalities and Applications; Fang, Y., Zhang, H., Nishinari, K., Eds.; Springer: Singapore, 2021; pp. 119–170. [Google Scholar]
  34. Banerjee, S.; Bhattacharya, S. Food Gels: Gelling Process and New Applications. Crit. Rev. Food Sci. Nutr. 2012, 52, 334–346. [Google Scholar] [CrossRef] [PubMed]
  35. Nicolai, T. Gelation of food protein-protein mixtures. Adv. Colloid Interface Sci. 2019, 270, 147–164. [Google Scholar] [CrossRef]
  36. Chen, H.; Zhang, T.; Tian, Y.; You, L.; Huang, Y.; Wang, S. Novel self-assembling peptide hydrogel with pH-tunable assembly microstructure, gel mechanics and the entrapment of curcumin. Food Hydrocoll. 2022, 124, 107338. [Google Scholar] [CrossRef]
  37. Wang, Y.; Zhang, S.; Zheng, L. Soy protein-based protein composite system: Gelation, application, and challenges—A review. Eur. Food Res. Technol. 2025, 251, 311–325. [Google Scholar] [CrossRef]
  38. Lv, Y.; Xu, L.; Su, Y.; Chang, C.; Gu, L.; Yang, Y.; Li, J. Effect of soybean protein isolate and egg white mixture on gelation of chicken myofibrillar proteins under salt/-free conditions. LWT 2021, 149, 111871. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Liu, J.; Yan, Z.; Zhang, R.; Du, Z.; Shang, X.; Zhang, T.; Liu, X. Mechanism of ultrasound-induced soybean/egg white composite gelation: Gel properties, morphological structure and co-aggregation kinetics. Int. J. Biol. Macromol. 2024, 266, 131267. [Google Scholar] [CrossRef]
  40. Manzoor, A.; Dar, A.H.; Pandey, V.K.; Shams, R.; Khan, S.; Panesar, P.S.; Kennedy, J.F.; Fayaz, U.; Khan, S.A. Recent insights into polysaccharide-based hydrogels and their potential applications in food sector: A review. Int. J. Biol. Macromol. 2022, 213, 987–1006. [Google Scholar] [CrossRef]
  41. Hu, X.; Jiang, Q.; Du, L.; Meng, Z. Edible polysaccharide-based oleogels and novel emulsion gels as fat analogues: A review. Carbohydr. Polym. 2023, 322, 121328. [Google Scholar] [CrossRef]
  42. Kong, D.; Zhang, M.; Mujumdar, A.S.; Yu, D. New drying technologies for animal/plant origin polysaccharide-based future food processing: Research progress, application prospects and challenges. Food Biosci. 2023, 56, 103315. [Google Scholar] [CrossRef]
  43. Xiang, Q.; Hao, Y.; Xia, Z.; Liao, M.; Rao, X.; Lao, S.; He, Q.; Ma, C.; Liao, W. Biomedical Applications and Nutritional Value of Specific Food-Derived Polysaccharide-Based Hydrogels. Adv. Nutr. Int. Rev. J. 2024, 15, 100309. [Google Scholar] [CrossRef]
  44. Yang, X.; Li, A.; Li, D.; Guo, Y.; Sun, L. Applications of mixed polysaccharide-protein systems in fabricating multi-structures of binary food gels—A review. Trends Food Sci. Technol. 2021, 109, 197–210. [Google Scholar] [CrossRef]
  45. Ren, S.; Zhang, G.; Wang, Z.; Sun, F.; Cheng, T.; Wang, D.; Yang, H.; Wang, Z.; Guo, Z. Potentially texture-modified food for dysphagia: Gelling, rheological, and water fixation properties of rice starch–soybean protein composite gels in various ratios. Food Hydrocoll. 2024, 153, 110025. [Google Scholar] [CrossRef]
  46. Du, L.; Guo, Y.; Meng, Z. Organogels, O/W and W/O emulsion gels structured by monoglycerides: The study on the gelation behavior and crystal network. Eur. Food Res. Technol. 2025, 251, 165–177. [Google Scholar] [CrossRef]
  47. Shi, Z.; Shi, Z.; Wu, M.; Shen, Y.; Li, G.; Ma, T. Fabrication of emulsion gel based on polymer sanxan and its potential as a sustained-release delivery system for β-carotene. Int. J. Biol. Macromol. 2020, 164, 597–605. [Google Scholar] [CrossRef] [PubMed]
  48. Funami, T.; Ishihara, S.; Maeda, K.; Nakauma, M. Review paper: Recent development in Pickering emulsion gel technology for food and beverage applications. Food Hydrocoll. 2025, 162, 110901. [Google Scholar] [CrossRef]
  49. Yang, T.; Zan, S.; Li, B.; Li, L.; Zhang, X. Interfacial adsorption dynamics of solid lipid particles at oil/water interfaces through QCM-D technique. Food Hydrocoll. 2024, 148, 109431. [Google Scholar] [CrossRef]
  50. Gomes, A.; Furtado, G.d.F.; Cunha, R.L. Bioaccessibility of Lipophilic Compounds Vehiculated in Emulsions: Choice of Lipids and Emulsifiers. J. Agric. Food Chem. 2018, 67, 13–18. [Google Scholar] [CrossRef]
  51. Zhang, R.; Li, B.; Song, Y.; Li, L.; Zhang, X. Tailoring stability in oil-in-water high internal phase Pickering emulsions (HIPPEs) through surface modification of beeswax-based solid lipid particles (SLPs) with various surfactants. Food Hydrocoll. 2024, 156, 110264. [Google Scholar] [CrossRef]
  52. Ferro, A.C.; Okuro, P.K.; Badan, A.P.; Cunha, R.L. Role of the oil on glyceryl monostearate based oleogels. Food Res. Int. 2019, 120, 610–619. [Google Scholar] [CrossRef] [PubMed]
  53. Du, L.; Meng, Z. Engineering surfactant-free pickering double emulsions gels with different structures as low-calorie fat analogues: Tunable oral perception, inhibiting lipid digestion, and potent co-delivery for lycopene and epigallocatechin gallate. Food Chem. 2025, 463, 141378. [Google Scholar] [CrossRef]
  54. Patel, A.R.; Dewettinck, K. Comparative evaluation of structured oil systems: Shellac oleogel, HPMC oleogel, and HIPE gel. Eur. J. Lipid Sci. Technol. 2015, 117, 1772–1781. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, X.; Song, Z.; Tian, W.; Abdullah; Huang, Q.; Chen, M.; Huang, Y.; Xiao, H.; Xiao, J. Advancements in lipid-based delivery systems for functional foods: A comprehensive review of literature and patent trends. Crit. Rev. Food Sci. Nutr. 2025, 65, 2456–2472. [Google Scholar] [CrossRef]
  56. Dille, M.; Draget, K.; Hattrem, M. 9—The effect of filler particles on the texture of food gels. In Modifying Food Texture; Chen, J., Rosenthal, A., Eds.; Woodhead Publishing: London, UK, 2015; pp. 183–200. [Google Scholar]
  57. Ghebremedhin, M.; Seiffert, S.; Vilgis, T.A. Molecular behavior of fluid gels—The crucial role of edges and particle surface in macroscopic properties. Food Funct. 2022, 13, 6902–6922. [Google Scholar] [CrossRef]
  58. Song, J.; Vikulina, A.S.; Parakhonskiy, B.V.; Skirtach, A.G. Hierarchy of hybrid materials. Part-II: The place of organics-on-inorganics in it, their composition and applications. Front. Chem. 2023, 11, 1078840. [Google Scholar] [CrossRef]
  59. Gravelle, A.J.; Nicholson, R.A.; Barbut, S.; Marangoni, A.G. Considerations for readdressing theoretical descriptions of particle-reinforced composite food gels. Food Res. Int. 2019, 122, 209–221. [Google Scholar] [CrossRef]
  60. Abalymov, A.; Van der Meeren, L.; Saveleva, M.; Prikhozhdenko, E.; Dewettinck, K.; Parakhonskiy, B.V.; Skirtach, A.G. Cells-Grab-on Particles: A Novel Approach to Control Cell Focal Adhesion on Hybrid Thermally Annealed Hydrogels. ACS Biomater. Sci. Eng. 2020, 6, 3933–3944. [Google Scholar] [CrossRef]
  61. Cao, L.; Lewille, B.; Dewettinck, K.; Willaert, R.G.; Tan, M.; Skirtach, A.G.; Parakhonskiy, B.V. Emulsion-filled bulk gels based on alginate and gellan gum: The fabrication, characterization, curcumin delivery, and antioxidative properties. Chem. Eng. J. 2024, 501, 157649. [Google Scholar] [CrossRef]
  62. Zou, Y.; Guo, J.; Yin, S.-W.; Wang, J.-M.; Yang, X.-Q. Pickering Emulsion Gels Prepared by Hydrogen-Bonded Zein/Tannic Acid Complex Colloidal Particles. J. Agric. Food Chem. 2015, 63, 7405–7414. [Google Scholar] [CrossRef] [PubMed]
  63. Khalesi, H.; Lu, W.; Nishinari, K.; Fang, Y. New insights into food hydrogels with reinforced mechanical properties: A review on innovative strategies. Adv. Colloid Interface Sci. 2020, 285, 102278. [Google Scholar] [CrossRef]
  64. Huang, R.; Liu, L.; Cai, M.; Sun, X.; Feng, G.; Zeng, M. Fabrication of emulsion gels with oyster protein particles through depletion attraction for 3D printing. Food Hydrocoll. 2024, 155, 110150. [Google Scholar] [CrossRef]
  65. Torres, O.; Tena, N.M.; Murray, B.; Sarkar, A. Novel starch based emulsion gels and emulsion microgel particles: Design, structure and rheology. Carbohydr. Polym. 2017, 178, 86–94. [Google Scholar] [CrossRef]
  66. Moschakis, T. Microrheology and particle tracking in food gels and emulsions. Curr. Opin. Colloid Interface Sci. 2013, 18, 311–323. [Google Scholar] [CrossRef]
  67. Wang, X.; Liu, S.; Ai, Y. Gelation mechanisms of granular and non-granular starches with variations in molecular structures. Food Hydrocoll. 2022, 129, 107658. [Google Scholar] [CrossRef]
  68. Xiao, Y.; Kang, S.; Liu, Y.; Guo, X.; Li, M.; Xu, H. Effect and mechanism of calcium ions on the gelation properties of cellulose nanocrystals-whey protein isolate composite gels. Food Hydrocoll. 2021, 111, 106401. [Google Scholar] [CrossRef]
  69. Hou, X.; Lin, L.; Li, K.; Jiang, F.; Qiao, D.; Zhang, B.; Xie, F. Towards superior biopolymer gels by enabling interpenetrating network structures: A review on types, applications, and gelation strategies. Adv. Colloid Interface Sci. 2024, 325, 103113. [Google Scholar] [CrossRef]
  70. Jin, W.; Xu, W.; Ge, H.; Li, J.; Li, B. Coupling process of phase separation and gelation in konjac glucomannan and gelatin system. Food Hydrocoll. 2015, 51, 188–192. [Google Scholar] [CrossRef]
  71. Totosaus, A.; Montejano, J.G.; A Salazar, J.; Guerrero, I. A review of physical and chemical protein-gel induction. Int. J. Food Sci. Technol. 2002, 37, 589–601. [Google Scholar] [CrossRef]
  72. Liu, Z.; Ren, X.; Cheng, Y.; Zhao, G.; Zhou, Y. Gelation mechanism of alkali induced heat-set konjac glucomannan gel. Trends Food Sci. Technol. 2021, 116, 244–254. [Google Scholar] [CrossRef]
  73. Huyst, A.M.; Deleu, L.J.; Luyckx, T.; Van der Meeren, L.; Housmans, J.A.; Grootaert, C.; Monge-Morera, M.; Delcour, J.A.; Skirtach, A.G.; Rousseau, F.; et al. Impact of heat and enzymatic treatment on ovalbumin amyloid-like fibril formation and enzyme-induced gelation. Food Hydrocoll. 2022, 131, 107784. [Google Scholar] [CrossRef]
  74. Rhein-Knudsen, N.; Meyer, A.S. Chemistry, gelation, and enzymatic modification of seaweed food hydrocolloids. Trends Food Sci. Technol. 2021, 109, 608–621. [Google Scholar] [CrossRef]
  75. Kalab, M.; Harwalkar, V. Milk Gel Structure. I. Application of Scanning Electron Microscopy to Milk and Other Food Gels. J. Dairy Sci. 1973, 56, 835–842. [Google Scholar] [CrossRef]
  76. Liu, Y.; Wang, D.; Sun, Z.; Liu, F.; Du, L.; Wang, D. Preparation and characterization of gelatin/chitosan/3-phenylacetic acid food-packaging nanofiber antibacterial films by electrospinning. Int. J. Biol. Macromol. 2021, 169, 161–170. [Google Scholar] [CrossRef] [PubMed]
  77. Gómez-Mascaraque, L.G.; Pinho, S.C. Microstructural Analysis of Whey/Soy Protein Isolate Mixed Gels Using Confocal Raman Microscopy. Foods 2021, 10, 2179. [Google Scholar] [CrossRef]
  78. Mishra, V.S.; Via, M.A.; Andersen, U.; Møller, F.; Brewer, J.R.; Simonsen, A.C. Quantifying microstructure and molecular dynamics in plant-dairy based model gels using FLIM and STED microscopy. Food Hydrocoll. 2025, 163, 111086. [Google Scholar] [CrossRef]
  79. Li, R.; Ebbesen, M.F.; Glover, Z.J.; Jæger, T.C.; Rovers, T.A.; Svensson, B.; Brewer, J.R.; Simonsen, A.C.; Ipsen, R.; Hougaard, A.B. Discriminating between different proteins in the microstructure of acidified milk gels by super-resolution microscopy. Food Hydrocoll. 2023, 138, 108468. [Google Scholar] [CrossRef]
  80. Błaszczak, W.; Lewandowicz, G. Light Microscopy as a Tool to Evaluate the Functionality of Starch in Food. Foods 2020, 9, 670. [Google Scholar] [CrossRef] [PubMed]
  81. De Witte, F.; Penagos, I.A.; Moens, K.; Skirtach, A.G.; Van Bockstaele, F.; Dewettinck, K. Multiscale assessment of the effect of a stearic-palmitic sucrose ester on the crystallization of anhydrous milk fat. Food Res. Int. 2024, 197, 115243. [Google Scholar] [CrossRef] [PubMed]
  82. Niu, Y.; Zheng, Y.; Fu, X.; Zeng, D.; Liu, H. A novel characterization of starch gelatinization using microscopy observation with deep learning methodology. J. Food Eng. 2022, 327, 111057. [Google Scholar] [CrossRef]
  83. Ventura, I.; Jammal, J.; Bianco-Peled, H. Insights into the nanostructure of low-methoxyl pectin–calcium gels. Carbohydr. Polym. 2013, 97, 650–658. [Google Scholar] [CrossRef] [PubMed]
  84. Di Lorenzo, F.; Seiffert, S. Nanostructural heterogeneity in polymer networks and gels. Polym. Chem. 2015, 6, 5515–5528. [Google Scholar] [CrossRef]
  85. Olakanmi, S.; Karunakaran, C.; Jayas, D. Applications of X-ray micro-computed tomography and small-angle X-ray scattering techniques in food systems: A concise review. J. Food Eng. 2023, 342, 111355. [Google Scholar] [CrossRef]
  86. Bayrak, M.; Whitten, A.E.; Mata, J.P.; Conn, C.E.; Floury, J.; Logan, A. Logan, Real-time monitoring of casein gel microstructure during simulated gastric digestion monitored by small-angle neutron scattering. Food Hydrocoll. 2023, 144, 108919. [Google Scholar] [CrossRef]
  87. Bayrak, M.; Mata, J.P.; Whitten, A.E.; Conn, C.E.; Floury, J.; Logan, A. Logan, Physical disruption of gel particles on the macroscale does not affect the study of protein gel structure on the micro or nanoscale. Colloid Interface Sci. Commun. 2022, 46, 100574. [Google Scholar] [CrossRef]
  88. McDowall, D.; Adams, D.J.; Seddon, A.M. Using small angle scattering to understand low molecular weight gels. Soft Matter 2022, 18, 1577–1590. [Google Scholar] [CrossRef]
  89. Tomchuk, O.V.; Avdeev, M.V.; Aleksenskii, A.E.; Vul, A.Y.; Ivankov, O.I.; Ryukhtin, V.V.; Füzi, J.; Garamus, V.M.; Bulavin, L.A. Sol–Gel Transition in Nanodiamond Aqueous Dispersions by Small-Angle Scattering. J. Phys. Chem. C 2019, 123, 18028–18036. [Google Scholar] [CrossRef]
  90. Bourbon, A.I.; Pereira, R.N.; Pastrana, L.M.; Vicente, A.A.; Cerqueira, M.A. Protein-Based Nanostructures for Food Applications. Gels 2019, 5, 9. [Google Scholar] [CrossRef]
  91. Hassan, N.; Ahmad, T.; Zain, N.M.; Awang, S.R. Identification of bovine, porcine and fish gelatin signatures using chemometrics fuzzy graph method. Sci. Rep. 2021, 11, 9793. [Google Scholar] [CrossRef] [PubMed]
  92. Lshaq, A.; Rahman, U.U.; Sahar, A.; Perveen, R.; Deering, A.J.; Khalil, A.A.; Aadil, R.M.; Hafeez, M.A.; Khaliq, A.; Siddique, U. Potentiality of analytical approaches to determine gelatin authenticity in food systems: A review. LWT 2020, 121, 108968. [Google Scholar] [CrossRef]
  93. Martin, B.; Skirtach, A.G.; Boon, N.; De Troch, M. Lipid Quantification and Determination of Astaxanthin Conjugation by Raman Microscopy on Individual Copepod Eggs. J. Raman Spectrosc. 2025, 56, 472–480. [Google Scholar] [CrossRef]
  94. Marvizadeh, M.M.; Oladzadabbasabadi, N.; Nafchi, A.M.; Jokar, M. Preparation and characterization of bionanocomposite film based on tapioca starch/bovine gelatin/nanorod zinc oxide. Int. J. Biol. Macromol. 2017, 99, 1–7. [Google Scholar] [CrossRef] [PubMed]
  95. Barasinski, M.; Jasper, V.; Görke, M.; Garnweitner, G. In Situ Tracking of Nanoparticles During Electrophoresis in Hydrogels Using a Fiber-Based UV-Vis System. Powders 2025, 4, 3. [Google Scholar] [CrossRef]
  96. Rahman, N.; Ahmad, I. Coordination polymer gel mediated spectrophotometric, ICP-AES and spectrofluorimetric methods for trace As(III) determination in water and food samples. Chemosphere 2024, 351, 141272. [Google Scholar] [CrossRef]
  97. Fan, Z.; Cheng, P.; Zhang, P.; Zhang, G.; Han, J. Rheological insight of polysaccharide/protein based hydrogels in recent food and biomedical fields: A review. Int. J. Biol. Macromol. 2022, 222, 1642–1664. [Google Scholar] [CrossRef] [PubMed]
  98. Ilyin, S.O. Structural Rheology in the Development and Study of Complex Polymer Materials. Polymers 2024, 16, 2458. [Google Scholar] [CrossRef]
  99. Krajina, B.A.; Tropini, C.; Zhu, A.; DiGiacomo, P.; Sonnenburg, J.L.; Heilshorn, S.C.; Spakowitz, A.J. Dynamic Light Scattering Microrheology Reveals Multiscale Viscoelasticity of Polymer Gels and Precious Biological Materials. ACS Central Sci. 2017, 3, 1294–1303. [Google Scholar] [CrossRef]
  100. Geonzon, L.C.; Kobayashi, M.; Tassieri, M.; Bacabac, R.G.; Adachi, Y.; Matsukawa, S. Microrheological properties and local structure of ι-carrageenan gels probed by using optical tweezers. Food Hydrocoll. 2023, 137, 108325. [Google Scholar] [CrossRef]
  101. He, S.; Caggioni, M.; Lindberg, S.; Schultz, K.M. Gelation phase diagrams of colloidal rod systems measured over a large composition space. RSC Adv. 2022, 12, 12902–12912. [Google Scholar] [CrossRef]
  102. Schmidt, R.F.; Kiefer, H.; Dalgliesh, R.; Gradzielski, M.; Netz, R.R. Nanoscopic Interfacial Hydrogel Viscoelasticity Revealed from Comparison of Macroscopic and Microscopic Rheology. Nano Lett. 2024, 24, 4758–4765. [Google Scholar] [CrossRef]
  103. Xiao, Y.; Liu, Y.; Wang, Y.; Jin, Y.; Guo, X.; Liu, Y.; Qi, X.; Lei, H.; Xu, H. Heat-induced whey protein isolate gels improved by cellulose nanocrystals: Gelling properties and microstructure. Carbohydr. Polym. 2020, 231, 115749. [Google Scholar] [CrossRef] [PubMed]
  104. Qu, R.-J.; Wang, Y.; Li, D.; Wang, L.-J. Rheological behavior of nanocellulose gels at various calcium chloride concentrations. Carbohydr. Polym. 2021, 274, 118660. [Google Scholar] [CrossRef]
  105. Feng, X.; Dai, H.; Ma, L.; Fu, Y.; Yu, Y.; Zhou, H.; Guo, T.; Zhu, H.; Wang, H.; Zhang, Y. Properties of Pickering emulsion stabilized by food-grade gelatin nanoparticles: Influence of the nanoparticles concentration. Colloids Surf. B Biointerfaces 2020, 196, 111294. [Google Scholar] [CrossRef]
  106. Yao, W.; Huang, X.; Li, C.; Kong, B.; Xia, X.; Sun, F.; Liu, Q.; Cao, C. Underlying the effect of soybean oil concentration on the gelling properties of myofibrillar protein-based emulsion gels: Perspective on interfacial adsorption, rheological properties and protein conformation. Food Hydrocoll. 2025, 162, 110935. [Google Scholar] [CrossRef]
  107. Hu, Y.; Cheng, L.; Lee, S.J.; Yang, Z. Formation and characterisation of concentrated emulsion gels stabilised by faba bean protein isolate and its applications for 3D food printing. Colloids Surf. A Physicochem. Eng. Asp. 2023, 671, 131622. [Google Scholar] [CrossRef]
  108. Wang, W.; Zhang, X.; Teng, A.; Liu, A. Mechanical reinforcement of gelatin hydrogel with nanofiber cellulose as a function of percolation concentration. Int. J. Biol. Macromol. 2017, 103, 226–233. [Google Scholar] [CrossRef]
  109. Liu, H.; Wu, Y.; Li, Y.; Xie, X.-A.; Li, P.; Du, B.; Li, L. Effect of different valence metal cations on the gel characteristics and microstructure of Inca peanut albumin gels. Food Hydrocoll. 2024, 151, 109783. [Google Scholar] [CrossRef]
  110. Kaur, H.; Sharma, P.; Patel, N.; Pal, V.K.; Roy, S. Accessing Highly Tunable Nanostructured Hydrogels in a Short Ionic Complementary Peptide Sequence via pH Trigger. Langmuir 2020, 36, 12107–12120. [Google Scholar] [CrossRef]
  111. Adams, D.J. Personal Perspective on Understanding Low Molecular Weight Gels. J. Am. Chem. Soc. 2022, 144, 11047–11053. [Google Scholar] [CrossRef]
  112. Shakeel, A.; Farooq, U.; Gabriele, D.; Marangoni, A.G.; Lupi, F.R. Bigels and multi-component organogels: An overview from rheological perspective. Food Hydrocoll. 2021, 111, 106190. [Google Scholar] [CrossRef]
  113. Raak, N.; Leonhardt, L.; Rohm, H.; Jaros, D. Size Modulation of Enzymatically Cross-Linked Sodium Caseinate Nanoparticles via Ionic Strength Variation Affects the Properties of Acid-Induced Gels. Dairy 2021, 2, 148–164. [Google Scholar] [CrossRef]
  114. Yan, X.; Jia, Y.; Man, H.; Liu, L.; Sun, S.; Qi, B.; Li, Y. Intermolecular interactions and gel properties of composite agglomerative networks based on oppositely charged polymers: Effects of pH and ionic strength. Food Hydrocoll. 2023, 139, 108557. [Google Scholar] [CrossRef]
  115. Sosa, E.I.F.; Chaves, M.G.; Peyrano, F.; Quiroga, A.V.; Avanza, M.V. Thermal Gelation of Proteins from Cajanus cajan Influenced by pH and Ionic Strength. Plant Foods Hum. Nutr. 2023, 78, 574–583. [Google Scholar] [CrossRef] [PubMed]
  116. Zhu, Y.Q.; Chen, X.; McClements, D.J.; Zou, L.; Liu, W. Pickering-stabilized emulsion gels fabricated from wheat protein nanoparticles: Effect of pH, NaCl and oil content. J. Dispers. Sci. Technol. 2018, 39, 826–835. [Google Scholar] [CrossRef]
  117. Malektaj, H.; Drozdov, A.D.; Fini, E.; Christiansen, J.D. The Effect of pH on the Viscoelastic Response of Alginate–Montmorillonite Nanocomposite Hydrogels. Molecules 2024, 29, 244. [Google Scholar] [CrossRef] [PubMed]
  118. Ghebremedhin, M.; Seiffert, S.; Vilgis, T.A. Effects of sugar molecules on the rheological and tribological properties and on the microstructure of agarose-based fluid gels. Front. Soft Matter 2024, 4, 1363898. [Google Scholar] [CrossRef]
  119. Wu, C.-L.; Liao, J.-S.; Wang, J.-M.; Qi, J.-R. Gelation behavior and mechanism of low methoxyl pectin in the presence of erythritol and sucrose: The role of co-solutes. Int. J. Biol. Macromol. 2024, 271, 132261. [Google Scholar] [CrossRef]
  120. Wang, R.; Hartel, R.W. Confectionery gels: Gelling behavior and gel properties of gelatin in concentrated sugar solutions. Food Hydrocoll. 2022, 124, 107132. [Google Scholar] [CrossRef]
  121. Liu, Z.; Ruan, M.; Ha, S.; Chitrakar, B.; Li, H.; Hu, L.; Mo, H. Effect of curcumin/cyclodextrin composite on the gelation and 3D printability of k-carrageenan in the sucrose co-solute field. J. Food Eng. 2025, 391, 112441. [Google Scholar] [CrossRef]
  122. Chivers, P.R.A.; Smith, D.K. Shaping and structuring supramolecular gels. Nat. Rev. Mater. 2019, 4, 463–478. [Google Scholar] [CrossRef]
  123. Liu, J.; Zhang, L.; Liu, C.; Zheng, X.; Tang, K. Tuning structure and properties of gelatin edible films through pullulan dialdehyde crosslinking. LWT 2021, 138, 110607. [Google Scholar] [CrossRef]
  124. Montero, M.P.; Tovar, C.A.; Alemán, A.; Gómez-Guillén, M.C. Characterisation of desolvation-produced high-molecular-weight gelatin fractions and their use for nanoparticles synthesis. J. Food Eng. 2024, 370, 111965. [Google Scholar] [CrossRef]
  125. Yue, B.; Shi, Y.; Yang, F.; Ye, D.; Jin, J.; Zhang, P.; Song, S.; Xu, Y.; Lin, H.; Zhu, S.; et al. Smartly tuning supramolecular chirality in polymer gels via photoexcitation-induced cooperative self-assembly. Sci. China Chem. 2025, 68, 1126–1135. [Google Scholar] [CrossRef]
  126. Derkach, S.R.; Ilyin, S.O.; Maklakova, A.A.; Kulichikhin, V.G.; Malkin, A.Y. The rheology of gelatin hydrogels modified by κ-carrageenan. LWT 2015, 63, 612–619. [Google Scholar] [CrossRef]
  127. Maroufi, L.Y.; Ghorbani, M.; Tabibiazar, M.; Mohammadi, M.; Pezeshki, A. Advanced properties of gelatin film by incorporating modified kappa-carrageenan and zein nanoparticles for active food packaging. Int. J. Biol. Macromol. 2021, 183, 753–759. [Google Scholar] [CrossRef]
  128. Luangapai, F.; Iwamoto, S. Influence of blending and layer-by-layer assembly methods on chitosan–gelatin composite films enriched with curcumin nanoemulsion. Int. J. Biol. Macromol. 2023, 249, 126061. [Google Scholar] [CrossRef]
  129. Li, J.; Parakhonskiy, B.V.; Skirtach, A.G. A decade of developing applications exploiting the properties of polyelectrolyte multilayer capsules. Chem. Commun. 2022, 59, 807–835. [Google Scholar] [CrossRef]
  130. Hirst, A.R.; Smith, D.K.; Feiters, M.C.; Geurts, H.P.M.; Wright, A.C. Two-Component Dendritic Gels:  Easily Tunable Materials. J. Am. Chem. Soc. 2003, 125, 9010–9011. [Google Scholar] [CrossRef] [PubMed]
  131. Hardy, J.G.; Hirst, A.R.; Smith, D.K.; Brennan, C.; Ashworth, I. Controlling the materials properties and nanostructure of a single-component dendritic gel by adding a second component. Chem. Commun. 2004, 3, 385–387. [Google Scholar] [CrossRef] [PubMed]
  132. Li, Y.; Xiang, Y.; Chen, Y.; Wang, Y.; Dong, W.; Liu, Y.; Qi, X.; Shen, J. A Natural Eumelanin-Assisted Pullulan/Chitosan Hydrogel for the Management of Diabetic Oral Ulcers. Macromol. Biosci. 2024, 25, 2400526. [Google Scholar] [CrossRef]
  133. Gudmundsson, T.A.; Kuppadakkath, G.; Ghosh, D.; Ruether, M.; Seddon, A.; Ginesi, R.E.; Doutch, J.; Adams, D.J.; Gunnlaugsson, T.; Damodaran, K.K. Nanoscale assembly of enantiomeric supramolecular gels driven by the nature of solvents. Nanoscale 2024, 16, 8922–8930. [Google Scholar] [CrossRef]
  134. Zhao, X.; Li, D.; Wang, L.-J.; Wang, Y. Role of gelation temperature in rheological behavior and microstructure of high elastic starch-based emulsion-filled gel. Food Hydrocoll. 2023, 135, 108208. [Google Scholar] [CrossRef]
  135. Cheng, L.-C.; Vehusheia, S.L.K.; Doyle, P.S. Tuning Material Properties of Nanoemulsion Gels by Sequentially Screening Electrostatic Repulsions and Then Thermally Inducing Droplet Bridging. Langmuir 2020, 36, 3346–3355. [Google Scholar] [CrossRef] [PubMed]
  136. Cheng, L.-C.; Hashemnejad, S.M.; Zarket, B.; Muthukrishnan, S.; Doyle, P.S. Thermally and pH-responsive gelation of nanoemulsions stabilized by weak acid surfactants. J. Colloid Interface Sci. 2020, 563, 229–240. [Google Scholar] [CrossRef]
  137. Wang, Z.; Zeng, J.; Deng, Y.; Zhou, P.; Li, P.; Zhao, Z.; Liu, G.; Zhang, M. Regulating heat-induced fibrous whey protein-wheat starch composite emulsion gels as dysphagia food by preheating temperature: Insights from protein-starch interactions. Food Hydrocoll. 2025, 159, 110621. [Google Scholar] [CrossRef]
  138. Moghimi, E.; Jacob, A.R.; Koumakis, N.; Petekidis, G. Colloidal gels tuned by oscillatory shear. Soft Matter 2017, 13, 2371–2383. [Google Scholar] [CrossRef]
  139. Koumakis, N.; Moghimi, E.; Besseling, R.; Poon, W.C.K.; Brady, J.F.; Petekidis, G. Tuning colloidal gels by shear. Soft Matter 2015, 11, 4640–4648. [Google Scholar] [CrossRef]
  140. Alfaro-Rodríguez, M.-C.; García, M.C.; Prieto-Vargas, P.; Muñoz, J. Rheological Properties and Physical Stability of Aqueous Dispersions of Flaxseed Fibers. Gels 2024, 10, 787. [Google Scholar] [CrossRef] [PubMed]
  141. Das, M.; Petekidis, G. Shear induced tuning and memory effects in colloidal gels of rods and spheres. J. Chem. Phys. 2022, 157, 234902. [Google Scholar] [CrossRef]
  142. Sudreau, I.; Manneville, S.; Servel, M.; Divoux, T. Shear-induced memory effects in boehmite gels. J. Rheol. 2021, 66, 91–104. [Google Scholar] [CrossRef]
  143. Zhang, S.; Han, J.; Chen, L. Fabrication of pea protein gels with modulated rheological properties using high pressure processing. Food Hydrocoll. 2023, 144, 109002. [Google Scholar] [CrossRef]
  144. Peyrano, F.; de Lamballerie, M.; Avanza, M.V.; Speroni, F. Gelation of cowpea proteins induced by high hydrostatic pressure. Food Hydrocoll. 2021, 111, 106191. [Google Scholar] [CrossRef]
  145. Renggli, D.; Doyle, P.S. Thermogelation of nanoemulsions stabilized by a commercial pea protein isolate: High-pressure homogenization defines gel strength. Soft Matter 2025, 21, 652–669. [Google Scholar] [CrossRef]
  146. Heidary, A.; Soltanizadeh, N. The Effects of High-Pressure Homogenization on Physicochemical and Functional Properties of Gelatin. Food Bioprocess Technol. 2024, 17, 100–122. [Google Scholar] [CrossRef]
  147. Liu, Y.; Chao, C.; Yu, J.; Wang, S.; Wang, S.; Copeland, L. New insights into starch gelatinization by high pressure: Comparison with heat-gelatinization. Food Chem. 2020, 318, 126493. [Google Scholar] [CrossRef]
  148. Lv, P.; Wang, D.; Dai, L.; Wu, X.; Gao, Y.; Yuan, F. Pickering emulsion gels stabilized by high hydrostatic pressure-induced whey protein isolate gel particles: Characterization and encapsulation of curcumin. Food Res. Int. 2020, 132, 109032. [Google Scholar] [CrossRef]
  149. Ni, Y.; Wu, J.; Jiang, Y.; Li, J.; Fan, L.; Huang, S. High-internal-phase pickering emulsions stabilized by ultrasound-induced nanocellulose hydrogels. Food Hydrocoll. 2022, 125, 107395. [Google Scholar] [CrossRef]
  150. Song, Y.; Xiao, J.; Li, L.; Wan, L.; Li, B.; Zhang, X. Ultrasound treatment of crystalline oil-in-water emulsions stabilized by sodium caseinate: Impact on emulsion stability through altered crystallization behavior in the oil globules. Ultrason. Sonochem. 2024, 106, 106897. [Google Scholar] [CrossRef] [PubMed]
  151. Chen, L.; Zhang, S.-B. Structural and functional properties of self-assembled peanut protein nanoparticles prepared by ultrasonic treatment: Effects of ultrasound intensity and protein concentration. Food Chem. 2023, 413, 135626. [Google Scholar] [CrossRef] [PubMed]
  152. Li, G.; Tao, R.; Sun, Y.; Wang, L.; Li, Y.; Fan, B.; Wang, F. Enhancing the Gelation Behavior of Transglutaminase-Induced Soy Protein Isolate(SPI) through Ultrasound-Assisted Extraction. Foods 2024, 13, 738. [Google Scholar] [CrossRef] [PubMed]
  153. Nguyen, H.V.; Huynh, P.X.; Kha, T.C. Ultrasound-Induced Modification of Durian Starch (Durio zibethinus) for Gel-Based Applications: Physicochemical and Thermal Properties. Gels 2025, 11, 296. [Google Scholar] [CrossRef] [PubMed]
  154. Li, P.; Guo, C.; Li, X.; Yuan, K.; Yang, X.; Guo, Y.; Yang, X. Preparation and structural characteristics of composite alginate/casein emulsion gels: A microscopy and rheology study. Food Hydrocoll. 2021, 118, 106792. [Google Scholar] [CrossRef]
  155. Wang, X.; Zhou, D.; Guo, Q.; Liu, C. Textural and structural properties of a κ-carrageenan–konjac gum mixed gel: Effects of κ-carrageenan concentration, mixing ratio, sucrose and Ca2+ concentrations and its application in milk pudding. J. Sci. Food Agric. 2020, 101, 3021–3029. [Google Scholar] [CrossRef]
  156. Zhang, T.; Wang, Z.; Yu, S.; Guo, X.; Ai, C.; Tang, X.; Chen, H.; Lin, J.; Zhang, X.; Meng, H. Effects of pH and temperature on the structure, rheological and gel-forming properties of sugar beet pectins. Food Hydrocoll. 2021, 116, 106646. [Google Scholar] [CrossRef]
  157. Yang, Y.; Xu, Q.; Wang, X.; Bai, Z.; Xu, X.; Ma, J. Casein-based hydrogels: Advances and prospects. Food Chem. 2024, 447, 138956. [Google Scholar] [CrossRef] [PubMed]
  158. Tanger, C.; Müller, M.; Andlinger, D.; Kulozik, U. Influence of pH and ionic strength on the thermal gelation behaviour of pea protein. Food Hydrocoll. 2022, 123, 106903. [Google Scholar] [CrossRef]
  159. 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]
  160. Mousavi, S.M.R.; Rafe, A.; Yeganehzad, S. Structure-rheology relationships of composite gels: Alginate and Basil seed gum/guar gum. Carbohydr. Polym. 2020, 232, 115809. [Google Scholar] [CrossRef] [PubMed]
  161. Patel, P.; Mujmer, K.; Aswal, V.K.; Gupta, S.; Thareja, P. Structure, rheology, and 3D printing of salt-induced κ-carrageenan gels. Mater. Today Commun. 2023, 35, 105807. [Google Scholar] [CrossRef]
  162. Guo, R.; Liu, L.; Huang, Y.; Lv, M.; Zhu, Y.; Wang, Z.; Zhu, X.; Sun, B. Effect of Na+ and Ca2+ on the texture, structure and microstructure of composite protein gel of mung bean protein and wheat gluten. Food Res. Int. 2023, 172, 113124. [Google Scholar] [CrossRef]
  163. Lei, Y.; Ouyang, H.; Peng, W.; Yu, X.; Jin, L.; Li, S. Effect of NaCl on the Rheological, Structural, and Gelling Properties of Walnut Protein Isolate-κ-Carrageenan Composite Gels. Gels 2022, 8, 259. [Google Scholar] [CrossRef] [PubMed]
  164. Baek, M.; Yoo, B.; Lim, S.-T. Effects of sugars and sugar alcohols on thermal transition and cold stability of corn starch gel. Food Hydrocoll. 2004, 18, 133–142. [Google Scholar] [CrossRef]
  165. Andreadis, M.; Moschakis, T. Effect of ethanol on gelation and microstructure of whey protein gels in the presence of NaCl. Food Hydrocoll. 2023, 134, 107985. [Google Scholar] [CrossRef]
  166. Jiang, W.-X.; Qi, J.-R.; Liao, J.-S.; Yang, X.-Q. Acid/ethanol induced pectin gelling and its application in emulsion gel. Food Hydrocoll. 2021, 118, 106774. [Google Scholar] [CrossRef]
  167. Wang, Y.; Yang, X.; Li, L. Formation of pH-responsive hydrogel beads and their gel properties: Soybean protein nanofibers and sodium alginate. Carbohydr. Polym. 2024, 329, 121748. [Google Scholar] [CrossRef] [PubMed]
  168. Cao, L.; Li, J.; Parakhonskiy, B.; Skirtach, A.G. Intestinal-specific oral delivery of lactoferrin with alginate-based composite and hybrid CaCO3-hydrogel beads. Food Chem. 2024, 451, 139205. [Google Scholar] [CrossRef]
  169. Le, X.T.; Rioux, L.-E.; Turgeon, S.L. Formation and functional properties of protein–polysaccharide electrostatic hydrogels in comparison to protein or polysaccharide hydrogels. Adv. Colloid Interface Sci. 2017, 239, 127–135. [Google Scholar] [CrossRef] [PubMed]
  170. Deng, W.; Tang, Y.; Mao, J.; Zhou, Y.; Chen, T.; Zhu, X. Cellulose nanofibril as a crosslinker to reinforce the sodium alginate/chitosan hydrogels. Int. J. Biol. Macromol. 2021, 189, 890–899. [Google Scholar] [CrossRef]
  171. Liu, Q.; Liu, J.; Qin, S.; Pei, Y.; Zheng, X.; Tang, K. High mechanical strength gelatin composite hydrogels reinforced by cellulose nanofibrils with unique beads-on-a-string morphology. Int. J. Biol. Macromol. 2020, 164, 1776–1784. [Google Scholar] [CrossRef]
  172. Jeong, C.; Kim, S.; Lee, C.; Cho, S.; Kim, S.-B. Changes in the Physical Properties of Calcium Alginate Gel Beads under a Wide Range of Gelation Temperature Conditions. Foods 2020, 9, 180. [Google Scholar] [CrossRef]
  173. Yan, J.; Li, S.; Chen, G.; Ma, C.; McClements, D.J.; Liu, X.; Liu, F. Formation, physicochemical properties, and comparison of heat- and enzyme-induced whey protein-gelatin composite hydrogels. Food Hydrocoll. 2023, 137, 108384. [Google Scholar] [CrossRef]
  174. Ryu, J.; McClements, D.J. Impact of heat-set and cold-set gelling polysaccharides on potato protein gelation: Gellan gum, agar, and methylcellulose. Food Hydrocoll. 2024, 149, 109535. [Google Scholar] [CrossRef]
  175. Obas, F.-L.; Thomas, L.C.; Terban, M.W.; Schmidt, S.J. Characterization of the thermal behavior and structural properties of a commercial high-solids confectionary gel made with gelatin. Food Hydrocoll. 2024, 148, 109432. [Google Scholar] [CrossRef]
  176. Jia, R.; He, X.; McClements, D.J.; Qin, Y.; Xiong, L.; Dai, L.; Sun, Q. Control of starch gel properties by shear-induced disruption of gelatinized swollen starch granule integrity. Food Hydrocoll. 2025, 168, 111528. [Google Scholar] [CrossRef]
  177. Fang, F.; Martinez, M.M.; Campanella, O.H.; Hamaker, B.R. Long-term low shear-induced highly viscous waxy potato starch gel formed through intermolecular double helices. Carbohydr. Polym. 2020, 232, 115815. [Google Scholar] [CrossRef]
  178. Toprakcioglu, Z.; Knowles, T.P.J. Shear-mediated sol-gel transition of regenerated silk allows the formation of Janus-like microgels. Sci. Rep. 2021, 11, 6673. [Google Scholar] [CrossRef]
  179. Larrea-Wachtendorff, D.; Sousa, I.; Ferrari, G. Starch-Based Hydrogels Produced by High-Pressure Processing (HPP): Effect of the Starch Source and Processing Time. Food Eng. Rev. 2021, 13, 622–633. [Google Scholar] [CrossRef]
  180. Sun, Y.; Wang, L.; Wang, H.; Zhou, B.; Jiang, L.; Zhu, X. Effect of pH-shifting and ultrasound on soy/potato protein structure and gelation. Food Hydrocoll. 2025, 159, 110672. [Google Scholar] [CrossRef]
  181. Mozafarpour, R.; Koocheki, A.; Sani, M.A.; McClements, D.J.; Mehr, H.M. Ultrasound-modified protein-based colloidal particles: Interfacial activity, gelation properties, and encapsulation efficiency. Adv. Colloid Interface Sci. 2022, 309, 102768. [Google Scholar] [CrossRef]
  182. Li, R.; Ma, C.; Wang, N.; Wang, J.; Yang, X. Exploring the potential of ultrasound to improve the physicochemical properties of protein-based emulsion gels: By ultrasonically treating emulsion (pre-emulsification, post-emulsification) and substrate solution respectively. Food Hydrocoll. 2025, 168, 111492. [Google Scholar] [CrossRef]
  183. Li, Q.; He, Q.; Xu, M.; Li, J.; Liu, X.; Wan, Z.; Yang, X. Food-Grade Emulsions and Emulsion Gels Prepared by Soy Protein–Pectin Complex Nanoparticles and Glycyrrhizic Acid Nanofibrils. J. Agric. Food Chem. 2020, 68, 1051–1063. [Google Scholar] [CrossRef]
  184. Hao, Y.; Li, S.; Guo, X.; Fang, M.; Liu, X.; Gong, Z. Preparation of shellac nanoparticles-chitosan complexes stabilized Pickering emulsion gels and its application in β-carotene delivery. Int. J. Biol. Macromol. 2024, 281, 136583. [Google Scholar] [CrossRef]
  185. Yang, T.; Jia, H.; Song, Y.; Xu, D.; Li, B.; Li, L.; Skirtach, A.; Zhang, X. Natural shellac nanoparticles embedded porous chitosan microgel as a stability-enhanced Pickering interfacial biocatalyst. Int. J. Biol. Macromol. 2025, 321, 146213. [Google Scholar] [CrossRef]
  186. Momtaz, M.; Momtaz, E.; Mehrgardi, M.A.; Momtaz, F.; Narimani, T.; Poursina, F. Preparation and characterization of gelatin/chitosan nanocomposite reinforced by NiO nanoparticles as an active food packaging. Sci. Rep. 2024, 14, 519. [Google Scholar] [CrossRef] [PubMed]
  187. Yu, M.; Ji, N.; Wang, Y.; Dai, L.; Xiong, L.; Sun, Q. Starch-based nanoparticles: Stimuli responsiveness, toxicity, and interactions with food components. Compr. Rev. Food Sci. Food Saf. 2021, 20, 1075–1100. [Google Scholar] [CrossRef]
  188. Gong, H.; Zi, Y.; Kan, G.; Li, L.; Shi, C.; Wang, X.; Zhong, J. Preparation of food-grade EDC/NHS-crosslinked gelatin nanoparticles and their application for Pickering emulsion stabilization. Food Chem. 2024, 436, 137700. [Google Scholar] [CrossRef]
  189. Liu, H.; Liu, C.; McClements, D.J.; Xu, X.; Bai, C.; Sun, Q.; Xu, F.; Dai, L. Reinforcement of heat-set whey protein gels using whey protein nanofibers: Impact of nanofiber morphology and pH values. Food Hydrocoll. 2024, 153, 109954. [Google Scholar] [CrossRef]
  190. Mendoza, L.; Batchelor, W.; Tabor, R.F.; Garnier, G. Gelation mechanism of cellulose nanofibre gels: A colloids and interfacial perspective. J. Colloid Interface Sci. 2018, 509, 39–46. [Google Scholar] [CrossRef]
  191. Ghorani, B.; Emadzadeh, B.; Rezaeinia, H.; Russell, S. Improvements in gelatin cold water solubility after electrospinning and associated physicochemical, functional and rheological properties. Food Hydrocoll. 2020, 104, 105740. [Google Scholar] [CrossRef]
  192. Onyeaka, H.; Passaretti, P.; Miri, T.; Al-Sharify, Z.T. The safety of nanomaterials in food production and packaging. Curr. Res. Food Sci. 2022, 5, 763–774. [Google Scholar] [CrossRef] [PubMed]
  193. Störmer, A.; Bott, J.; Kemmer, D.; Franz, R. Critical review of the migration potential of nanoparticles in food contact plastics. Trends Food Sci. Technol. 2017, 63, 39–50. [Google Scholar] [CrossRef]
  194. Kaphle, A.; Navya, P.N.; Umapathi, A.; Daima, H.K. Nanomaterials for agriculture, food and environment: Applications, toxicity and regulation. Environ. Chem. Lett. 2018, 16, 43–58. [Google Scholar] [CrossRef]
  195. Qi, X.; Li, Y.; Xiang, Y.; Chen, Y.; Shi, Y.; Ge, X.; Zeng, B.; Shen, J. Hyperthermia-enhanced immunoregulation hydrogel for oxygenation and ROS neutralization in diabetic foot ulcers. Cell Biomater. 2025, 1, 100020. [Google Scholar] [CrossRef]
  196. Gupta, R.K.; Guha, P.; Srivastav, P.P. Investigating the toxicological effects of nanomaterials in food packaging associated with human health and the environment. J. Hazard. Mater. Lett. 2024, 5, 100125. [Google Scholar] [CrossRef]
  197. Rahman, J.M.H.; Shiblee, N.I.; Ahmed, K.; Khosla, A.; Kawakami, M.; Furukawa, H. Rheological and mechanical properties of edible gel materials for 3D food printing technology. Heliyon 2020, 6, e05859. [Google Scholar] [CrossRef] [PubMed]
  198. Beikzadeh, S.; Hosseini, S.M.; Mofid, V.; Ramezani, S.; Ghorbani, M.; Ehsani, A.; Mortazavian, A.M. Electrospun ethyl cellulose/poly caprolactone/gelatin nanofibers: The investigation of mechanical, antioxidant, and antifungal properties for food packaging. Int. J. Biol. Macromol. 2021, 191, 457–464. [Google Scholar] [CrossRef]
  199. Chen, J.; Luo, L.; Cen, C.; Liu, Y.; Li, H.; Wang, Y. The nano antibacterial composite film carboxymethyl chitosan/gelatin/nano ZnO improves the mechanical strength of food packaging. Int. J. Biol. Macromol. 2022, 220, 462–471. [Google Scholar] [CrossRef]
  200. Piao, X.; Li, J.; Zhao, Y.; Guo, L.; Zheng, B.; Zhou, R.; Ostrikov, K. Oxidized cellulose nanofibrils-based surimi gel enhancing additives: Interactions, performance and mechanisms. Food Hydrocoll. 2022, 133, 107893. [Google Scholar] [CrossRef]
  201. Riquelme, N.; Savignones, C.; López, A.; Zúñiga, R.N.; Arancibia, C. Effect of Gelling Agent Type on the Physical Properties of Nanoemulsion-Based Gels. Colloids Interfaces 2023, 7, 49. [Google Scholar] [CrossRef]
  202. Xu, W.; Ning, Y.; Wang, M.; Zhang, S.; Sun, H.; Yin, Y.; Li, N.; Li, P.; Luo, D. Construction of astaxanthin loaded Pickering emulsions gel stabilized by xanthan gum/lysozyme nanoparticles with konjac glucomannan from structure, protection and gastrointestinal digestion perspective. Int. J. Biol. Macromol. 2023, 252, 126421. [Google Scholar] [CrossRef]
  203. Alqarni, L.S.; Alghamdi, A.M.; Elamin, N.Y.; Rajeh, A. Enhancing the optical; electrical, dielectric properties and antimicrobial activity of chitosan/gelatin incorporated with Co-doped ZnO nanoparticles: Nanocomposites for use in energy storage and food packaging. J. Mol. Struct. 2024, 1297, 137011. [Google Scholar] [CrossRef]
  204. Shipovskaya, A.B.; Ushakova, O.S.; Volchkov, S.S.; Shipenok, X.M.; Shmakov, S.L.; Gegel, N.O.; Burov, A.M. Chiral Nanostructured Glycerohydrogel Sol–Gel Plates of Chitosan L- and D-Aspartate: Supramolecular Ordering and Optical Properties. Gels 2024, 10, 427. [Google Scholar] [CrossRef]
  205. Seiffert, S. Origin of nanostructural inhomogeneity in polymer-network gels. Polym. Chem. 2017, 8, 4472–4487. [Google Scholar] [CrossRef]
  206. Araújo, J.F.; Bourbon, A.I.; Simões, L.S.; Vicente, A.A.; Coutinho, P.J.G.; Ramos, O.L. Physicochemical characterisation and release behaviour of curcumin-loaded lactoferrin nanohydrogels into food simulants. Food Funct. 2020, 11, 305–317. [Google Scholar] [CrossRef]
  207. Zhao, Y.; Han, X.; Hu, N.; Zhao, C.; Wu, Y.; Liu, J. Study on properties of TGase-induced pea protein–zein complex gels. J. Food Eng. 2023, 354, 111578. [Google Scholar] [CrossRef]
  208. Cortés, N.M.; Montoto, S.S.; Ruiz, M.E.; Califano, A.N.; Zaritzky, N.; Lorenzo, G. Rheological properties and microstructure of thermodynamically stable microemulsions as factors influencing the release rate of liposoluble vitamins. Food Hydrocoll. 2023, 141, 108699. [Google Scholar] [CrossRef]
  209. Zhou, H.; Hu, X.; Xiang, X.; McClements, D.J. Modification of textural attributes of potato protein gels using salts, polysaccharides, and transglutaminase: Development of plant-based foods. Food Hydrocoll. 2023, 144, 108909. [Google Scholar] [CrossRef]
  210. Cheng, Y.; Meng, Y.; Liu, S. Diversified Techniques for Restructuring Meat Protein-Derived Products and Analogues. Foods 2024, 13, 1950. [Google Scholar] [CrossRef]
  211. Xu, Q.; Wang, H.; Ren, Y.; Sun, M.; Zhang, T.; Li, H.; Liu, X. Functionality and application of emulsion gels in fat replacement strategies for dairy products. Trends Food Sci. Technol. 2024, 152, 104673. [Google Scholar] [CrossRef]
  212. Ren, Y.; Huang, L.; Zhang, Y.; Li, H.; Zhao, D.; Cao, J.; Liu, X. Application of Emulsion Gels as Fat Substitutes in Meat Products. Foods 2022, 11, 1950. [Google Scholar] [CrossRef] [PubMed]
  213. Kwon, H.C.; Shin, D.-M.; Yune, J.H.; Jeong, C.H.; Han, S.G. Evaluation of gels formulated with whey proteins and sodium dodecyl sulfate as a fat replacer in low-fat sausage. Food Chem. 2021, 337, 127682. [Google Scholar] [CrossRef]
  214. Milićević, N.; Sakač, M.; Hadnađev, M.; Škrobot, D.; Šarić, B.; Hadnađev, T.D.; Jovanov, P.; Pezo, L. Physico-chemical properties of low-fat cookies containing wheat and oat bran gels as fat replacers. J. Cereal Sci. 2020, 95, 103056. [Google Scholar] [CrossRef]
  215. Li, S.; Yan, J.; Yang, J.; Chen, G.; McClements, D.J.; Ma, C.; Liu, X.; Liu, F. Modulating peppermint oil flavor release properties of emulsion-filled protein gels: Impact of cross-linking method and matrix composition. Food Res. Int. 2024, 185, 114277. [Google Scholar] [CrossRef]
  216. Kim, Y.M.; Lee, K.; Lee, Y.; Yang, K.; Choe, D.; Roh, Y.H. Thermoresponsive semi-interpenetrating gelatin-alginate networks for encapsulation and controlled release of scent molecules. Int. J. Biol. Macromol. 2022, 208, 1096–1105. [Google Scholar] [CrossRef]
  217. Keum, D.H.; Han, J.H.; Kwon, H.C.; Park, S.M.; Kim, H.Y.; Han, S.G. Enhancing the flavor of plant-based meat analogues using flavor-capturing alginate/β-cyclodextrin hydrogel beads. Int. J. Biol. Macromol. 2025, 309, 142930. [Google Scholar] [CrossRef]
  218. Liao, P.; Dai, S.; Lian, Z.; Tong, X.; Yang, S.; Chen, Y.; Qi, W.; Peng, X.; Wang, H.; Jiang, L. The Layered Encapsulation of Vitamin B2 and β-Carotene in Multilayer Alginate/Chitosan Gel Microspheres: Improving the Bioaccessibility of Vitamin B2 and β-Carotene. Foods 2022, 11, 20. [Google Scholar] [CrossRef] [PubMed]
  219. Yan, B.; Davachi, S.M.; Ravanfar, R.; Dadmohammadi, Y.; Deisenroth, T.W.; Pho, T.V.; Odorisio, P.A.; Darji, R.H.; Abbaspourrad, A. Improvement of vitamin C stability in vitamin gummies by encapsulation in casein gel. Food Hydrocoll. 2021, 113, 106414. [Google Scholar] [CrossRef]
  220. Kuo, C.-C.; Clark, S.; Qin, H.; Shi, X. Development of a shelf-stable, gel-based delivery system for probiotics by encapsulation, 3D printing, and freeze-drying. LWT 2022, 157, 113075. [Google Scholar] [CrossRef]
  221. Zhang, S.; Leidy, H.J.; Vardhanabhuti, B. Protein Beverage vs. Protein Gel on Appetite Control and Subsequent Food Intake in Healthy Adults. Nutrients 2015, 7, 8700–8711. [Google Scholar] [CrossRef]
  222. Jiang, Y.; Zhou, T.; Zhang, S.; Leng, J.; Li, L.; Zhao, W. β-Glucan-based superabsorbent hydrogel ameliorates obesity-associated metabolic disorders via delaying gastric emptying, improving intestinal barrier function, and modulating gut microbiota. Int. J. Biol. Macromol. 2025, 304, 140846. [Google Scholar] [CrossRef] [PubMed]
  223. Lenie, M.D.R.; Ahmadzadeh, S.; Van Bockstaele, F.; Ubeyitogullari, A. Development of a pH-responsive system based on starch and alginate-pectin hydrogels using coaxial 3D food printing. Food Hydrocoll. 2024, 153, 109989. [Google Scholar] [CrossRef]
  224. Yu, J.; Wang, X.-Y.; Li, D.; Wang, L.-J.; Wang, Y. Development of soy protein isolate emulsion gels as extrusion-based 3D food printing inks: Effect of polysaccharides incorporation. Food Hydrocoll. 2022, 131, 107824. [Google Scholar] [CrossRef]
  225. Chen, J.; Wu, A.; Yang, M.; Ge, Y.; Pristijono, P.; Li, J.; Xu, B.; Mi, H. Characterization of sodium alginate-based films incorporated with thymol for fresh-cut apple packaging. Food Control. 2021, 126, 108063. [Google Scholar] [CrossRef]
  226. Gheorghita, R.; Gutt, G.; Amariei, S. The Use of Edible Films Based on Sodium Alginate in Meat Product Packaging: An Eco-Friendly Alternative to Conventional Plastic Materials. Coatings 2020, 10, 166. [Google Scholar] [CrossRef]
  227. Sagiri, S.S.; Poverenov, E. Oleogel-Based Nanoemulsions for Beverages: Effect of Self-Assembled Fibrillar Networks on Stability and Release Properties of Emulsions. Foods 2024, 13, 680. [Google Scholar] [CrossRef]
  228. Ozorio, L.; Passerini, A.B.; Silva, A.P.; Braga, A.R.; Perrechil, F. Designing Plant-Based Foods: Biopolymer Gelation for Enhanced Texture and Functionality. Foods 2025, 14, 1645. [Google Scholar] [CrossRef]
  229. Salahi, M.R.; Mohebbi, M.; Razavi, S.M.A. Analyzing the effects of aroma and texture interactions on oral processing behavior and dynamic sensory perception: A case study on cold-set emulsion-filled gels containing limonene and menthol. Food Hydrocoll. 2024, 154, 110128. [Google Scholar] [CrossRef]
  230. Raj, A.S.; Rahul, R.; Karthik, P. Nanoorganogels for Encapsulating Food Bioactive Compounds. Food Bioprocess Technol. 2025, 18, 129–149. [Google Scholar] [CrossRef]
  231. Mohammadian, M.; Waly, M.I.; Moghadam, M.; Emam-Djomeh, Z.; Salami, M.; Moosavi-Movahedi, A.A. Nanostructured food proteins as efficient systems for the encapsulation of bioactive compounds. Food Sci. Hum. Wellness 2020, 9, 199–213. [Google Scholar] [CrossRef]
  232. Chen, K.; Zhou, F.; Chen, Y.; Shen, Q.; Feng, S.; Liang, L. Co-encapsulation of bioactive components using protein-based various assemblies: Necessary, assembling structure, location and partition. Food Hydrocoll. 2024, 148, 109492. [Google Scholar] [CrossRef]
  233. Lai, W.-F.; Wong, W.-T. Property-Tuneable Microgels Fabricated by Using Flow-Focusing Microfluidic Geometry for Bioactive Agent Delivery. Pharmaceutics 2021, 13, 787. [Google Scholar] [CrossRef]
  234. Chen, H.; Pan, S.; Regenstein, J.M.; Huang, J.; Wang, L. Characteristics of composite gels composed of citrus insoluble nanofiber and amylose and their potential to be used as fat replacers. Food Chem. 2023, 409, 135269. [Google Scholar] [CrossRef]
  235. Cui, S.; McClements, D.J.; Shi, J.; Xu, X.; Ning, F.; Liu, C.; Zhou, L.; Sun, Q.; Dai, L. Fabrication and characterization of low-fat Pickering emulsion gels stabilized by zein/phytic acid complex nanoparticles. Food Chem. 2023, 402, 134179. [Google Scholar] [CrossRef] [PubMed]
  236. He, W.; Huang, Y.; Zhou, S.; Regenstein, J.M.; Wang, L. A composite gel formed by konjac glucomannan together with Nano-CF obtained by FeCl3-citric acid hydrolysis as a potential fat substitute. Int. J. Biol. Macromol. 2024, 268, 131618. [Google Scholar] [CrossRef] [PubMed]
  237. DeLoid, G.M.; Sohal, I.S.; Lorente, L.R.; Molina, R.M.; Pyrgiotakis, G.; Stevanovic, A.; Zhang, R.; McClements, D.J.; Geitner, N.K.; Bousfield, D.W.; et al. Reducing Intestinal Digestion and Absorption of Fat Using a Nature-Derived Biopolymer: Interference of Triglyceride Hydrolysis by Nanocellulose. ACS Nano 2018, 12, 6469–6479. [Google Scholar] [CrossRef] [PubMed]
  238. Riahi, Z.; Hong, S.J.; Rhim, J.-W.; Shin, G.H.; Kim, J.T. High-performance multifunctional gelatin-based films engineered with metal-organic frameworks for active food packaging applications. Food Hydrocoll. 2023, 144, 108984. [Google Scholar] [CrossRef]
  239. Ahari, H.; Naeimabadi, M. Employing Nanoemulsions in Food Packaging: Shelf Life Enhancement. Food Eng. Rev. 2021, 13, 858–883. [Google Scholar] [CrossRef]
  240. Babu, P.J. Nanotechnology mediated intelligent and improved food packaging. Int. Nano Lett. 2021, 12, 1–14. [Google Scholar] [CrossRef]
  241. Moustafa, H.; Hemida, M.H.; Nour, M.A.; Abou-Kandil, A.I. Intelligent packaging films based on two-dimensional nanomaterials for food safety and quality monitoring: Future insights and roadblocks. J. Thermoplast. Compos. Mater. 2024, 38, 1208–1230. [Google Scholar] [CrossRef]
  242. Koirala, P.; Sagar, N.A.; Thuanthong, A.; Al-Asmari, F.; Jagtap, S.; Nirmal, N. Revolutionizing seafood packaging: Advancements in biopolymer smart nano-packaging for extended shelf-life and quality assurance. Food Res. Int. 2025, 203, 115826. [Google Scholar] [CrossRef] [PubMed]
  243. Qin, Z.; Li, Z.; Huang, X.; Du, L.; Li, W.; Gao, P.; Chen, Z.; Zhang, J.; Guo, Z.; Li, Z.; et al. Advances in 3D and 4D Printing of Gel-Based Foods: Mechanisms, Applications, and Future Directions. Gels 2025, 11, 94. [Google Scholar] [CrossRef]
  244. Barekat, S.; Ubeyitogullari, A. Maximizing sorghum proteins printability: Optimizing gel formulation and 3D-printing parameters to develop a novel bioink. Int. J. Biol. Macromol. 2025, 300, 140245. [Google Scholar] [CrossRef]
  245. Chen, Y.; Zhang, M.; Sun, Y.; Phuhongsung, P. Improving 3D/4D printing characteristics of natural food gels by novel additives: A review. Food Hydrocoll. 2022, 123, 107160. [Google Scholar] [CrossRef]
  246. Li, X.; Liuping, F.; Yuanfa, L. New insights into food O/W emulsion gels: Strategies of reinforcing mechanical properties and outlook of being applied to food 3D printing. Crit. Rev. Food Sci. Nutr. 2021, 63, 1564–1586. [Google Scholar] [CrossRef] [PubMed]
  247. Ji, S.; Xu, T.; Liu, Y.; Li, H.; Luo, J.; Zou, Y.; Zhong, Y.; Li, Y.; Lu, B. Investigation of the mechanism of casein protein to enhance 3D printing accuracy of cassava starch gel. Carbohydr. Polym. 2022, 295, 119827. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of building steps and structures of food gels, including hydrogels, aerogels, oleogels, and bigels, and their application fields. Created in BioRender. group, n. (2025) https://BioRender.com/orxhurf (accessed on 10 July 2025).
Figure 1. Overview of building steps and structures of food gels, including hydrogels, aerogels, oleogels, and bigels, and their application fields. Created in BioRender. group, n. (2025) https://BioRender.com/orxhurf (accessed on 10 July 2025).
Gels 11 00620 g001
Figure 5. (A) Visual appearance, WHC, and gel strength of cellulose nanocrystals-whey protein isolate composite gels with different concentrations of Ca2+ (adapted from [68] with permission from Elsevier). Different lowercase letters (a–d) indicate significant differences among treatments (p < 0.05). (B) Chemical structure of the ionic complementary sequence, naphthoxyacetic acid-FEFK peptide, and schematic representation of the self-assembling behavior of the designed peptide sequence depending on the pH of the system (adapted with permission from [110]. Copyright 2020 American Chemical Society). (C) Schematic description of a low-molecular-weight gel assembling to form fibrous structures that entangle to form a network: Single-component and two-component systems (adapted with permission from [111]. Copyright 2022 American Chemical Society). (D) Different conventional bigels obtained by varying organogel/hydrogel ratio (adapted from [112] with permission from Elsevier).
Figure 5. (A) Visual appearance, WHC, and gel strength of cellulose nanocrystals-whey protein isolate composite gels with different concentrations of Ca2+ (adapted from [68] with permission from Elsevier). Different lowercase letters (a–d) indicate significant differences among treatments (p < 0.05). (B) Chemical structure of the ionic complementary sequence, naphthoxyacetic acid-FEFK peptide, and schematic representation of the self-assembling behavior of the designed peptide sequence depending on the pH of the system (adapted with permission from [110]. Copyright 2020 American Chemical Society). (C) Schematic description of a low-molecular-weight gel assembling to form fibrous structures that entangle to form a network: Single-component and two-component systems (adapted with permission from [111]. Copyright 2022 American Chemical Society). (D) Different conventional bigels obtained by varying organogel/hydrogel ratio (adapted from [112] with permission from Elsevier).
Gels 11 00620 g005
Figure 8. Schematic diagram of various applications of gels in food areas.
Figure 8. Schematic diagram of various applications of gels in food areas.
Gels 11 00620 g008
Table 2. Summary of gels’ food applications.
Table 2. Summary of gels’ food applications.
ApplicationFunctionExamplesKey BenefitsRef.
Textural modificationModify mouthfeel, texture, and rheologyRestructured meat, plant-based foodsImproves sensory appeal and consumer acceptability[45,209,210]
Fat replacementMimic fat mouthfeel with lower calorie contentProtein or polysaccharide-based gels fat mimeticsReduces calorie content while maintaining creamy texture[211,212,213,214]
Flavor encapsulationTrap and release volatile aroma/flavor compoundsEmulsion-filled gels, hydrogels with flavor compoundsControlled flavor release and protection from oxidation[215,216,217]
Nutrient deliveryControlled release and protection of bioactivesGels encapsulating vitamins, probiotics, polyphenolsEnhances stability, bioavailability, and targeted release[218,219,220]
Satiety enhancementInduce gastric retention or swelling to promote fullnessβ-Glucan, or protein gels for appetite regulationSupports weight management and satiety[221,222]
3D food printingServe as printable bio-ink or scaffold for customized food shapesAlginate, pectin, starch, or protein-based printable gelsEnables designable textures and personalized nutrition[223,224]
Edible coatings/filmsServe as sustainable food packaging materialsGelatin or alginate-based films on fruits or meatEnhances shelf-life and appearance[225,226]
Water or oil structuringStructure liquids into gels for functional or sensory improvementOleogels, hydrogel particles in beveragesStabilizes emulsions and improves mouthfeel in reduced-fat products[227]
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

Yang, T.; Cao, L.; Song, J.; Skirtach, A.G. Tuning Nanostructure of Gels: From Structural and Functional Controls to Food Applications. Gels 2025, 11, 620. https://doi.org/10.3390/gels11080620

AMA Style

Yang T, Cao L, Song J, Skirtach AG. Tuning Nanostructure of Gels: From Structural and Functional Controls to Food Applications. Gels. 2025; 11(8):620. https://doi.org/10.3390/gels11080620

Chicago/Turabian Style

Yang, Tangyu, Lin Cao, Junnan Song, and Andre G. Skirtach. 2025. "Tuning Nanostructure of Gels: From Structural and Functional Controls to Food Applications" Gels 11, no. 8: 620. https://doi.org/10.3390/gels11080620

APA Style

Yang, T., Cao, L., Song, J., & Skirtach, A. G. (2025). Tuning Nanostructure of Gels: From Structural and Functional Controls to Food Applications. Gels, 11(8), 620. https://doi.org/10.3390/gels11080620

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