Next Article in Journal
Fabrication Mechanisms of Lignin Nanoparticles and Their Ultraviolet Protection Ability in PVA Composite Film
Next Article in Special Issue
Rigid Polyurethane Biofoams Filled with Chemically Compatible Fruit Peels
Previous Article in Journal
Design Strategies for and Stability of mRNA–Lipid Nanoparticle COVID-19 Vaccines
Previous Article in Special Issue
Starch as a Matrix for Incorporation and Release of Bioactive Compounds: Fundamentals and Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Encapsulation of Bioactive Compounds for Food and Agricultural Applications

by
Giovani Leone Zabot
1,
Fabiele Schaefer Rodrigues
1,
Lissara Polano Ody
1,
Marcus Vinícius Tres
1,
Esteban Herrera
2,
Heidy Palacin
2,
Javier S. Córdova-Ramos
2,
Ivan Best
3 and
Luis Olivera-Montenegro
2,3,*
1
Laboratory of Agroindustrial Processes Engineering (LAPE), Federal University of Santa Maria (UFSM), Cachoeira do Sul 6508–010, Brazil
2
Grupo de Investigación en Bioprocesos y Conversión de la Biomasa, Universidad San Ignacio de Loyola, Lima 15024, Peru
3
Grupo de Ciencia, Tecnología e Innovación en Alimentos, Universidad San Ignacio de Loyola, Lima 15024, Peru
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(19), 4194; https://doi.org/10.3390/polym14194194
Submission received: 25 August 2022 / Revised: 21 September 2022 / Accepted: 29 September 2022 / Published: 6 October 2022
(This article belongs to the Special Issue Biopolymer Matrices for Incorporation of Bioactive Compounds)

Abstract

:
This review presents an updated scenario of findings and evolutions of encapsulation of bioactive compounds for food and agricultural applications. Many polymers have been reported as encapsulated agents, such as sodium alginate, gum Arabic, chitosan, cellulose and carboxymethylcellulose, pectin, Shellac, xanthan gum, zein, pullulan, maltodextrin, whey protein, galactomannan, modified starch, polycaprolactone, and sodium caseinate. The main encapsulation methods investigated in the study include both physical and chemical ones, such as freeze-drying, spray-drying, extrusion, coacervation, complexation, and supercritical anti-solvent drying. Consequently, in the food area, bioactive peptides, vitamins, essential oils, caffeine, plant extracts, fatty acids, flavonoids, carotenoids, and terpenes are the main compounds encapsulated. In the agricultural area, essential oils, lipids, phytotoxins, medicines, vaccines, hemoglobin, and microbial metabolites are the main compounds encapsulated. Most scientific investigations have one or more objectives, such as to improve the stability of formulated systems, increase the release time, retain and protect active properties, reduce lipid oxidation, maintain organoleptic properties, and present bioactivities even in extreme thermal, radiation, and pH conditions. Considering the increasing worldwide interest for biomolecules in modern and sustainable agriculture, encapsulation can be efficient for the formulation of biofungicides, biopesticides, bioherbicides, and biofertilizers. With this review, it is inferred that the current scenario indicates evolutions in the production methods by increasing the scales and the techno-economic feasibilities. The Technology Readiness Level (TRL) for most of the encapsulation methods is going beyond TRL 6, in which the knowledge gathered allows for having a functional prototype or a representative model of the encapsulation technologies presented in this review.

1. Introduction

The characteristics of foods depend on several factors, such as forms of presentation, nature, texture, flavor profiles, and major composition, among others [1,2]. However, many evolutions have been observed in recent years toward providing a wider range of products for consumers, with singular characteristics and properties. Minimally processed foods have shown characteristics similar to natural ones due to these developments in the aspects of engineering, nutrition, science, and technology. Accordingly, these foods have shown longer shelf life, nutritional functionality, different textures, and specific flavors, among other benefits [3,4,5].
In the same trend, the growing demand for biological products in agriculture encourages research on novel formulation techniques and especially the production of biological capsules [6]. This has occurred due to the stability of these bioproducts and higher reactivity of active ingredients, which minimizes volatility losses [7].
Based on this scenario, the encapsulation of bioactive substances (essential oils, plant extracts, fungal metabolites, etc.) in food and agricultural areas allows them to be protected against external factors and degradation. The encapsulation allows the biological integrity of the products and supports environmental conditions during storage, ensuring the viability of active ingredients for long periods [8]. For example, microbial agents are susceptible to abiotic and biotic factors that reduce the effectiveness of these living organisms and their metabolites when exposed to ultraviolet radiation and adverse temperatures, resulting in the loss of toxin integrity and spore viability [9]. In this sense, encapsulation is an alternative to these adversities. In the agricultural area, it allows for reducing losses by volatility, having better biological integrity, increasing efficiency, improving commercial viability, and increasing formulation stability, among others [10,11,12]. Therefore, the use of bioactive compounds as agents in the production of bioinsecticides, bionfungicides, bioherbicides, and biofertilizers is a promising strategy.
For encapsulation, many polymers are used as wall materials to protect the core, generally formed by bioactive compounds. Chitosan, gums (gum Arabic, Xanthan gum, gum acacia, and Shellac, for instance), maltodextrin, pectin, starch, whey protein, sodium alginate, cellulose and carboxymethylcellulose, zein, pullulan, galactomannan, and sodium caseinate, among others, are used for this purpose [13,14,15,16,17]. The polymers favor the retention of desired compounds in the systems formed during the processes and help to prolong the release of bioactive compounds for longer times or under specific conditions, such as certain pH ranges, for example [18,19].
Therefore, this review presents polymer matrices for the incorporation of bioactive compounds and discusses the characteristics of encapsulation systems. The data cited in the text were obtained from the scientific literature through the main scientific databases. The search was mainly focused on the five past years to have a recent scenario of findings and evolutions. Some figures were created by the authors to express the authors’ viewpoint and tables were compiled based on other referenced works.

2. Polymers Used for Encapsulation of Bioactive Compounds

Encapsulation is described as a process where a core material (a liquid, solid or gaseous compound) is packaged in a wall material to create capsules that are effective against chemical and environmental interactions [20,21]. Encapsulation is an alternative for problems of physical or chemical instability of compounds. It can inhibit volatilization and protect the encapsulated material against unfavorable environmental conditions, reducing the sensitivity to the degradation of plant materials and their bioactive compounds [22]. In such case, food biopolymers are used in the encapsulation process [23]. Biopolymers are classified into three classes: synthetic polymers derived from petroleum, synthetic polymers derived from renewable resources, and naturally produced renewable polymers, [24]. Many polymers (synthetic and natural) are used as a coating material for encapsulating bioactive compounds, such as polyethylene, polyethylene glycol, poly(vinylpyrrolidone), polyvinyl alcohol, polyacrylic acid, polylactic acid, polyhydroxyalkanoates, β-glucans, dextran, starch, alginate, cellulose, chitin, chitosan, pectin, collagen, gums, zein, hyaluronic acid, and gelatin, among others. Table 1 presents some examples of polymers and encapsulated bioactive compounds.

2.1. Chitosan

The deacetylation of the chitin polymer allows for producing chitosan, in which the acetyl groups of the chitin chain are removed to compose amino groups. The degree of deacetylation is directly interconnected with the performance of chitosan in its different uses [25]. Chitosan is a natural and non-toxic polysaccharide widely used due to its characteristics of biocompatibility, chemical resistance, and biodegradability. Moreover, it has the property of forming films without dependence on additives [26]. The high availability of chitosan is an important factor, which is the second most numerous natural polysaccharide biopolymer found in nature [27].
Chitosan still has antimicrobial and antioxidant activity [43]. The antimicrobial activity of chitosan depends on some parameters that affect its properties, such as pH, molecular weight, type of microorganism, source of chitosan, concentration, degree of deacetylation, complexes with certain materials, food components, and chitosan derivatives [25]. Chitosan is an excellent coating material used in the encapsulation of multiple bioactive compounds due to its characteristics and properties, having applications in food, pharmaceutical, agricultural, biomedical, environmental, and industrial segments, among others. This polymer is used in the encapsulation of essential oils, lipids, different food ingredients, vitamins, medicines, vaccines, hemoglobin, and microbial metabolites, among others [44]. In agriculture, chitosan and its encapsulated compounds have been widely used as an ecological alternative in agricultural production, such as biofertilizers, biopesticides, seed treatment agents, soil conditioners, and growth promoters [45]. Chitosan has been used as a co-encapsulating material for curcumin and resveratrol [46] and in the development of nanocomposite films to inhibit the growth of fungi, such as Penicillium chrysogenum, Aspergillus flavus, Aspergillus niger, and Aspergillus parasiticus, resulting in the control of these pathogens [27].
Table 1. Polymers used for encapsulation of bioactive compounds and the main objectives of the studies.
Table 1. Polymers used for encapsulation of bioactive compounds and the main objectives of the studies.
PolymerMaterialEncapsulated Bioactive CompoundObjectiveReferences
AlginateCa alginate hydrogel granulesReishi medicinal mushroom extract; Probiotic Lactobacillus acidophilusMask the bitter taste of extract and protect bioactive substances in oral administration to prolong cell viability under simulated gastrointestinal conditions and to protect the bioactive ingredients of Reishi mushroom along the storage[16]
AlginateMacrospheresPseudomonas sp. DN18Effective protection against diseases caused by Eclerotium rolfsii[28]
AlginateHydrogelJujube extract (Ziziphus spp.)Effect of encapsulation on antioxidant activity and protection of bioactive compounds[29]
Gum ArabicAdhesive membraneCinnamon extractPresent an active food packaging material with more control over its pungent smell and quick release[13]
Gum ArabicNanocapsuleSavory essential oilAlternative control method to the pre-emergence herbicide Metribuzin (Sencor®)[30]
Gum Arabic + ChitosanNanocapsuleSaffron (Crocus sativus L.)Increase bioavailability and protection of bioactive compounds through nanoencapsulation[31]
CellulosePowder particlesVitamin AEvaluate the emulsifying properties of cellulose particles and the ability to encapsulate with vitamin A[17]
CelluloseEdible filmsProbiotic bacteria (Lactobacillus rhamnosus GG)Search for new applications of coatings and films based on edible cellulose as carriers of various probiotics[32]
CeluloseCryogelsTebuconazole fungicideControlled release of Tebuconazole fungicide[33]
Chitosan + CelluloseNanoformulationsCitronella essential oil (Cymbopogon nardus (L.))Control of Spodoptera littoralis[34]
ChitosanNanoparticlesPeppermint oil (Mentha piperita (L.))A nanoinsecticide to control Tribolium castaneum (Herbst) and Sitophilus oryzae (L.)[35]
ChitosanNanoliposomesCaffeineRetention and release of caffeine in the digestive system[36]
PectinHydrogelsLactaseLactase encapsulated in pectin-based hydrogels[37]
PectinFilmBeetroot extract encapsulated in pectin from watermelon peelMonitor the freshness of packaged chilled beef by developing a pH indicator film[15]
PectinEdible coatingCarvacrol/2-hydroxypropyl-β-cyclodextrin (CAR/HPβCD-IC)Fungal inhibition against Botrytis cinerea and Alternaria alternata[38]
ShellacGelsRiboflavin Lactobacillus acidophilus amylaseForm shellac-based gels and oat protein at neutral pH as a carrier to entrap and deliver active substances[19]
Shellac + ChitinComposite microspheresYeast alcohol dehydrogenase (YADH)Enzymatic immobilization by adsorption[39]
Shellac + ZeinComposite capsulesCurcuminControlled release of curcumin[40]
Xanthan gum + sodium alginateGelsDebranched pea starchImprove the performance of debranched pea starch gels[41]
Xanthan gumEdible filmsXanthan Gum solutions with glycerol acidIncrease lotus root storage stability[42]

2.2. Cellulose

Cellulose is a natural polysaccharide with properties of high interest, e.g., biodegradability and biocompatibility [17]. In terms of chemical composition, cellulose is a homopolysaccharide, which is formed by many monomers of glucose [47]. Cellulose is found in plants, microorganisms, such as fungi and bacteria, and algae [48]. It is the most abundant naturally occurring polysaccharide [27]. Methylcellulose, carboxymethylcellulose, hydroxypropylmethylcellulose, and hydroxypropylcellulose are synthetic derivatives of cellulose, which are also used for the encapsulation and manufacture of edible food films. These films are usually used to protect fruits and vegetables through the use of barrier layers [26].
In addition to the synthetic derivatives of cellulose, generally carried out by etherification or esterification of hydroxyl groups, it is possible to produce cellulose derivatives from agricultural and agro-vegetable residues. New biopolymers that have unique characteristics and properties compatible with synthetic polymers can be used for their replacement with non-degradable polymers, combining their applications with environmental preservation [49]. Cellulose can be used as nanocomposites. Especially, the nanocrystals can be used as reinforcing agents in the preparation of composite materials due to their characteristics such as high reactivity, high strength and modulus, rigidity, renewability, low density, biodegradability, and no toxicity. Cellulose nanocrystals can compose films, food packaging, fibers, hydrogels, and drugs. Also, it can be used as a transport vehicle for agrochemicals and in many other segments of biotechnology [50].

2.3. Starch

Starch is a natural polymer, which is the main polysaccharide reserve material existing in photosynthetic tissues and other parts of plants such as fruits, seeds, tubers, and roots. Therefore, depending on its botanical origin, starch can have a different molecular structure, as well as distinct characteristics and properties, in addition to exhibiting diversity in the shape, size and composition of its granules [51].
Starch is characterized as a polymeric compound consisting of glucose monomer, while it is present in abundance in foods, such as grains, corn, potato, and rice [26]. This polymer has 20–25% amylose and 75–80% amylopectin, presenting hydrophilic granules. It is a material with multiple applications due to its biodegradability, availability, and abundance, in addition to presenting competitive cost compared to other polymers [49].

2.4. Alginate

Alginate is a natural hydrophilic polymer isolated from brown algae and is used in food applications due to its film-forming, gel-making, thickening, and stabilization properties [26]. Alginate exhibits antimicrobial properties, moisture absorption, gelation, biocompatibility, and application in various segments (pharmaceutical, cosmetic, food, and biomedical industries) as a consequence of its multiple materials such as films, hydrogels, microspheres, fibers, and microcapsules, among others [52].
Alginate is also widely used with other polymers to improve or obtain new properties through the integration of materials, such as the production of edible films from the composition of alginate and fish gelatin, resulting in the improvement of the mechanical characteristics of the film and increasing antioxidant capacity [53]. The manufacture of bilayer films based on sodium alginate and tea tree essential oil incorporated in TiO2 nanoparticles is reported, aiming at improving postharvest quality and reducing anthracnose in banana fruits [54]. The use of alginate for the encapsulation of vegetable oils is quite widespread, such as the use of alginate-based microparticles structured with other biopolymers such as pectin to improve the encapsulation efficiency of essential oils from olive leaves [55].

2.5. Shellac

Shellac is a natural polymer refined from lac resin that is excreted from insects, mainly from the species Kerria spp. As it is a natural resin, shellac exhibits non-toxic and biodegradable properties. It is characterized as a semi-crystalline polymer but with less regular alignment, less dense, and brittle, presenting itself as an alternative to synthetic polymers. Moreover, it has potential in the area of green technologies [56].
Shellac is composed mainly of oxyacid polyesters, being an amphiphilic biomacromolecule. This biopolymer has been widely used in several segments due to its characteristic of amphiphilicity and pH responsiveness. Examples of applications are food coating, biodegradable films, manufacturing of food waxes, preparation of food delivery systems (coated carriers, microcapsules, nanoparticles, and microparticles), and as an oil gelling agent [57].
Many studies report the use of shellac with other materials to improve the efficiency of the delivery of bioactive compounds through a synergistic effect between the substances. One example is the combination of shellac and tannic acid to prolong the shelf life of post-harvested mango and improve its quality [58]. Another example is the interaction of oat protein granules and shellac to the development of gels at near neutral pH to protect and release sensitive bioactive compounds [19]. The preparation of microspheres composed of shellac and chitin to obtain high-efficiency enzymatic immobilization with great potential in biotechnology is also highlighted [39]. Furthermore, the development of films composed of shellac and gelatin to extend the shelf life of banana fruits and improve post-harvest quality has been studied, thus presenting satisfactory results [59]. The use of zein and shellac composite particles to improve curcumin encapsulation efficiency can be cited [40], among others.

2.6. Pectin

Pectin is a natural heteropolysaccharide found in the meristematic tissue and the parenchyma. The amount of pectin in the cell wall and its quality change according to the plant source [60]. Pectin is composed of poly α−1–4 galacturonic acid residues with degrees of methylation of carboxylic acid residues. It is also composed of sugars, such as arabinose, galactose, and rhamnose [61].
The degree of esterification is a factor that influences the processing conditions and pectin characteristics. It is the percentage of the number of esterified carboxyl groups in relation to the degree of methylation. Pectin is separated into high methoxyl pectin (degree >50%) and low methoxyl pectin (degree < 50%) [62].
Pectin is a food safety active ingredient and presents several benefits, such as gelling agent, emulsion stabilizer, and binding capacity agent. It is not digested by intestinal enzymes, while it is mucoadhesive [23,61,62,63]. This biopolymer has anti-inflammatory, non-toxic, biodegradability, and biocompatibility properties. It can be found as a hydrogel, fiber, composites, and injectable cell vehicle [52].

2.7. Gum Arabic

Gum Arabic is a natural source of mineral salts (potassium, magnesium, calcium), fiber, and carbohydrates (arabinose and galactose). It is a heteropolysaccharide often used as a coating material due to its characteristics, such as retention of volatiles, emulsification, low cost, negatively charged properties, low viscosity, high solubility, easy use, inhibition of oxidation reactions, and colorless characteristic of solutions [64,65]. Gum Arabic is mainly used in the pharmaceutical area to concentrate syrups. In the food industry, it is used as a stabilizer, while in the agricultural industry it is mainly used to produce microcapsules containing bacteria [65,66].

2.8. Xanthan Gum

Xanthan gum is a polysaccharide biopolymer consisting of glucose, mannose, and diglucuronic acid, which is produced using some inexpensive nutrients such as sucrose, sugarcane molasses, and whey [65]. Xanthan gum is a novel generation of extracellular metabolites produced by bacteria such as Xanthomonas campestris [67].
This gum is easily dissolved in cold and hot water and produces a concentrated solution, even in a small amount [65]. Xanthan gum is used in many segments such as chemical, petroleum, cosmetic, food, and agricultural products [68]. This gum is also used in some beverages, candies, and frozen and canned foods [67]. Factors, such as temperature, pH, carbon sources, high pressure, polymer concentration, and viscosity, in the presence of galactomannan influence the production of this gum [69].

2.9. Dextran

Dextran is a linear polysaccharide biopolymer that contains hydroxyl groups used for the covalent bonding of some organic groups, such as hydrophobic compounds [70]. The modified dextran becomes insoluble or soluble in water and can form self-organized nanoscale particles by increasing the degree of substitution [71]. This biopolymer can be used for the production of nanocarriers to encapsulate active materials with different hydrophobicities. Despite its great potential, the nanoencapsulation of food bioactive compounds using modified dextrans polymers should be better studied [72].

2.10. Milk Proteins

Protein-based polymers, in addition to presenting mechanical properties of films equal or superior to other polymeric materials, have excellent gas barrier properties, which are widely used for food packaging [73]. In this context, milk proteins are used, which can be separated into caseins and whey proteins [74]. Casein is the main and high-quality milk protein [75], found naturally in milk in micellar supramolecule structures [76]. It constitutes approximately 80% of the total milk protein, while whey proteins represent approximately 20% of the total milk protein, and are normally obtained as a by-product (whey) of cheese production [77].
Milk proteins demonstrate several favorable characteristics for biotechnological, food, and agricultural applications due to their multiple functional and structural properties, benefiting the development of macro, micro or nano structures. Among its various functional properties, the emulsifying and foaming properties are the most important [74]. Milk proteins are used as nutraceutical vehicles, acting in the stabilization of bioactive compounds in foods, in the delivery of nutrients, in the improvement of food quality (sensory, color, flavor, texture), and the improvement of food safety. These proteins are found in different structural forms and levels, such as nanofilms, hydrogels, microcapsules, nanocomposites, nanocoatings, and microspheres, among others [78].
In the agricultural area, the encapsulation of an insecticide, Cydia pomonella granulovirus used in organic arboriculture, was performed with sodium caseinate, a hydrophilic milk protein. Lipophilic polyglycerol polyricinoleate and sodium caseinate concentrations larger than 1 wt% were very interesting for applications because the encapsulation remained at a high level for at least 300 days [79]. In addition, the encapsulation of neem seed oil extract within polymeric shells formed by whey protein isolate/maltodextrin was performed using spray drying. Neem seed oil contains azadirachtin as its main bioactive component, a compound having efficient insect repellent and insecticidal properties. As a result of the study, encapsulation efficiencies for neem seed oil were demonstrated higher in smaller microcapsules, with efficiency values of 60–92%. The association of botanical insecticides within biopolymer cores such as milk proteins presents a remarkable potential for increasing agricultural production levels and reducing impacts on human health and the environment [80].

3. Encapsulation of Bioactive Compounds

Bioactive compounds can be found in vegetables, fruits, cereals, legumes, roots, rhizomes, and other plant sources. In this sense, to improve their stability and applicability in both food and agriculture, encapsulation is a feasible technique. Microencapsulation and nanoencapsulation are the two most used technologies to encapsulate bioactive compounds [81].
For the encapsulation process of bioactive compounds, the selection of the appropriate encapsulating material and the encapsulation technique should be considered. Some reported techniques are emulsification, anti-solvent precipitation, electrospinning, and coacervation, among others. Coating materials influence emulsion properties, such as droplet size, viscosity, stability, and powder characteristics. Nanoencapsulation presents a particle size of less than 1 micron, microencapsulation presents a particle size between 1 to 1000 microns, and microencapsulation produces a particle size higher than 1000 microns [82].
Microencapsulation protects core bioactive compounds in a heterogeneous or homogeneous matrix. The technique and the type of coating materials play a significant role in the feed and powder properties [83]. Droplets of liquid or small particles are coated by a polymer to produce particles, such as microspheres or microcapsules. They are used in the industry as pharmaceuticals, chemical, and food agents. Microspheres are dense matrix systems, while microcapsules are hollow internally possessing a reservoir system. The microparticles can present many morphologies, which depend on the properties of the core with bioactive compounds, coating material, and microencapsulation technique [84]. The main types of encapsulation methods, physical, chemical, and physical-chemical, are presented in Figure 1.
Emulsion techniques are rather used in the food industry. Recently, Pickering emulsions have gained importance because they have higher stability to coalescence [85]. The Pickering emulsion technique for preparing microspheres, microcapsules, and foams is broadly studied because it is easy to operate and economically feasible [86]. However, there are other emulsion techniques (Figure 2).
Encapsulation allows active ingredients such as polyphenols, carotenoids, pigments, fatty acids, phytosterols, probiotics, vitamins, minerals, and bioactive peptides to be trapped within a matrix of different carriers [87]. Generally, the stability of these bioactive compounds is low, and encapsulation generates a powder with higher stability against variations in temperature, light, pH, or oxygen, increasing the release rate of these active ingredients.

3.1. Bioactive Peptides

Bioactive peptides are found in an inactive form in the structure of the original proteins and are activated after the cleavage of the proteins [88]. In most cases, the peptide chain comprises the amino acids arginine, proline and lysine together with hydrophobic residues. They can be classified as exogenous endogenous substances. Endogenous peptides are produced in different cell types, like neural, immune, or glands throughout the body. Exogenous peptides came from a variety of sources, such as food, supplements, and drugs [89].
Peptide fractions and protein hydrolysates can be used as nutraceuticals, functional foods, or ingredients in food products [90]. Some factors can affect bioactive peptides, limiting their application, e.g., pH, enzymatic degradation, low water solubility, interaction with the food matrix, hygroscopicity, and possible bitter taste. In this sense, encapsulation of peptides with some coating polymers can improve these aspects [91]. Some studies have presented that encapsulation of peptides increases their stability, bioactivity, solubility and sensory characteristics, and reduces hygroscopicity [82]. In addition, the various colloidal systems used to encapsulate peptides have presented good encapsulation efficiency and stability [92].
These bioactivities depend on the amino acid and the peptide sequence, which comprises 2 to 20 amino acids [93]. Bioactive peptides can be obtained by hydrolysis of protein of animal or vegetable origin, by chemical or enzymatic hydrolysis. Nowadays, peptides are mainly applied in the food industry to obtain functional and nutraceutical foods. Additionally, bioactive peptides have functional properties and can be used as additives, such as solubility, emulsification, foaming, and water/oil binding capacity.
Encapsulation techniques are a promising alternative to increase stability and control the release of peptides for food applications. Also, encapsulation of peptides in nanocarriers increases the sensory characteristics by masking their bitter taste, increasing the solubility, and decreasing the hygroscopicity. Some peptides can be used as natural food preservatives to prolong shelf life or incorporated into bioactive packaging in the form of coatings and edible films to improve the safety of some foods due to their antioxidant and antimicrobial properties. An overview of the bioactive peptides microencapsulation process and characterization, sources of the bioactive, coating materials, encapsulation method, process conditions and results in terms of process parameters, encapsulation efficiency, particle size, and retention of their functional properties is shown in Table 2.

3.2. Fatty Acids from Vegetable Oils

Vegetable oils rich in unsaturated fatty acids are prone to oxidation, which produces free radicals called hydroperoxides [83]. The encapsulation system reduces their degradation and/or masks certain undesirable, preserving the bioactive compounds and improving their controlled release, solubility, and bioavailability. Moreover, micro- and nano-encapsulation of extracts and essential oils can enhance their functional properties in new-generation foods [94]. Applications for the encapsulated essential oils have been emphasized, including antimicrobials, pesticides, natural insecticides, repellents, food packaging and taste preservation, and lipid oxidation [95,96,97].

3.3. Vitamins

Vitamin E is a fat-soluble vitamin from plants with health benefits [98]. This vitamin is made up of eight isoforms, including α-, β-, γ-, and δ-tocotrienol, and α-, β-, γ-, and δ-tocopherol [99]. The most prevalent vitamin E homolog in nature is α- tocopherol [100]. Tocopherols are found in high-fat foods such as nuts (almonds, peanuts, hazelnuts), seeds (sunflower, cottonseed), cereals (rice, oat, wheat), and vegetable oils (palm oil, wheat germ oil). Moreover, it is found in vegetables and fruit, such as spinach, tomatoes, green beans, broccoli, turnip, pumpkin, cranberries, carrots, avocado, kiwifruit, and raspberries [98,99,101,102].
Vitamin E is a weakly water-soluble bioactive component, which restricts its absorption in the gastrointestinal system and its total bioavailability. There is an interest in developing vitamin E fortified foods and beverages. Therefore, encapsulation is necessary to increase its absorption [103]. Furthermore, α-tocopherol is very susceptible to oxidative reactions, mediated by oxygen, light, heat, and free radicals [103,104,105].
Spray-drying is a known encapsulation technique because it allows for a continuous operating procedure. For improving vitamin E oral bioavailability, the spray freeze-drying method has shown to be effective [103]. The food industry uses encapsulated -tocopherol as antioxidants to extend the shelf life of fat-based bakery products [104]. Various studies have suggested that a vitamin E concentration beyond the advised daily allowance can have positive benefits on human health and may prevent reproductive, nervous system, inflammation, and immune system problems [106].
Recent studies on the encapsulation of bioactive peptides, fatty acids from vegetable oils, and vitamins (tocopherols) are presented in Table 2.
Table 2. Recent studies on encapsulation of bioactive peptides, fatty acids from vegetable oils, and vitamins (tocopherols).
Table 2. Recent studies on encapsulation of bioactive peptides, fatty acids from vegetable oils, and vitamins (tocopherols).
Bioactive CompoundsSourcesEncapsulation MaterialEncapsulation MethodProcess ConditionsObjectiveResultsReferences
Bioactive peptidesMilk casein hydrolysatePullulanElectrospinning (encapsulation of various bioactive compounds in the form of nanosized fibers)Pullulan at 100, 120, and 140 g kg−1Improve the stability, and low bioavailability, masking the bitter tasteProduction of clean bead-free peptides-loaded pullulan nanofibres at 120 and 140 g kg−1 with an encapsulation efficiency of 72–86% and a mean diameter of 60–133 nm[107]
Antioxidant peptidesFish hydrolyzed collagenLiposomeFreeze-dried The highest encapsulation efficiency was found in SPC-CHO-0.5% HC (p < 0.05) (85%); liposome stabilized with glycerol presented the highest efficiency (75%)Improve stability and bioactivitiesLyophilized SPC-CHO-0.5% HC presented higher stability than lyophilized SPC-GLY-0.25% HC during storage for 28 days at 25 °C [108]
Antioxidant peptidesFish protein hydrolysateNot describedSpray driedEnzymatic hydrolysis and chemical methods; spray dried at 180 °C inlet temperature to obtain powder; stored at −18 °CRetention of antioxidant properties and microstructureVisceral protein hydrolysate prepared with pepsin had better quality regarding antioxidant characteristics and papain in nutritional aspect[109]
Antihypertensive peptideWhey protein hydrolysate <3 kDaAlginate-collagen, alginate-gum Arabic, and alginate–gelatinExtrusion methodSonication for 15 minReleased antihypertensive peptides during gastrointestinal digestionThe highest efficiency was obtained in capsules of alginate—gum Arabic (95%); the released peptides incremented their ACE activity (85%)[110]
Antioxidant peptidesFlaxseed proteinMaltodextrinSpray dryingHydrolysate to maltodextrin (MD) to ratios (1:1, 1:2 and 1:3, w/w); alcalase enzyme was used to produce hydrolysatesRetention of antioxidant properties and microstructureSamples powders obtained by 1:3 ratio presented the highest radical scavenging activity for and ABTS+ (86%) and DPPH (69%); analysis of chemical structure indicated that hydrolysates were coated and dispersed within maltodextrin[91]
Antioxidant peptidesMilk Casein hydrolysateMaltodextrinSpray dryingSpraying was carried out by a pressure nozzle, compressed air flow rate of 0.54 m3 h−1, flow rate of 5 mL min−1, and inlet air temperature and outlet air temperature of 130 ± 1 °C and 70.0 ± 0.5 °CReduction of hygroscopicity and retention of antioxidant propertiesAntioxidant activities were 90–99%, 77–92%, 77–93%, 95–99%, and 77–98% after the spray-drying process; hygroscopicity was reduced by microencapsulation (p < 0.05)[111]
Antioxidant peptidesPink peach meat protein hydrolysateGum Arabic and maltodextrinEmulsions-spray dryingInlet air at 160 °C, outlet at 80 °C, nozzle diameter of 0.5 mm, air at 0.4 MPa, and spray flow feed rate of 15–20 mL min−1Retention of antioxidant properties; improvement of sensory propertiesAntioxidant activity was improved; sensory properties were improved in a concentration of up to 3%[112]
Bioactive peptidesAzocaseinNot applicableDouble emulsions water-in-oil-in-waterEnzymatic hydrolysisImproved bioavailabilityEncapsulation efficiency of casein peptides was 93%[113]
Antioxidant peptidesCasein hydrolysateGum Arabic and maltodextrinFreeze-driedEnzymatic hydrolysis, coating material (10:0, 8:2, 6:4); ultrasonication at 40 kHz, 750 W, 12 mm diameter tip and with 50% pulse for 20 minImprove the antioxidant and sensorial propertiesReduced bitterness if compared to the casein hydrolysate; maintenance of antioxidant activity (93%)[114]
Polyunsaturated fatty acidsTea oilMaltodextrin/Xanthan gum/Lysozyme nanoparticlesPickering emulsionTea oil plus composite solutions at oil-water volume ratio of 1:5; homogenization for 3 min at 18,000 rpm to obtain the tea oil Pickering emulsionReduce lipid oxidation on tea oil powderEncapsulation efficiency of 66% when 50% maltodextrin and 4% Xanthan gum/Xanthan gum/lysozyme nanoparticles was used; the surface of tea oil powder presented a relatively smooth porous microstructure[115]
Omega-3 fatty acidFlaxseed oilMaltodextrinCoacervation-Maintaining stability-[116]
Omega-3 and omega-6 fatty acidsLinum usitatissimumGum Arabic/whey protein/modified starch/sodium caseinateDrying (spray drying, freeze drying)/supercritical emulsification/emulsion/coacervation-Prevents the oxidation of fatty acidsMicrocapsules prepared with spray- and freeze-drying ranged between 10–400 and 20–5000 µm[117]
Unsaturated fatty acidsDrumstick oil (Moringa oleifera)Maltodextrin/gum Arabic (25:75); (oil to wall ratio 1:4)Spray dryingInlet air at 180 °C, outlet at 85 °C, air pressure of 0.06 MPa and air flow rate of 73 m3 h−1To evaluate the protection of the encapsulating compound in drumstick oilThe range of emulsion droplet mean diameters was 1.94 to 4.92 µm; encapsulation efficiency of 91.05% with lower water activity; good oxidative stability; peroxide value was 7.63 to 8.07 meq of peroxide/kg of oil after 30 days of storage at 45 °C; particles size was 22.56 ± 0.63 µm;showed larger smooth-surfaced particles, which may indicate that viscosity of the emulsions and emulsifying capacity are higher.[83]
Omega-3 fatty acidChia seed oilSoy protein microparticlesSupercritical CO2-assisted impregnation16 MPa impregnation pressure with ethanol as cosolvent (0.1-ethanol:oil ratio, w/w), temperature 40 °C and 4 h of contact timeProtect bioactive compounds through microencapsulation and increase bioavailabilityEncapsulation efficiency of 95% and a retention efficiency of 35%, showing excellent oxidative stability; microcapsules ranged between 1 and 10 μm, having a spherical form with occasional depressions but no pores or fissures; 95.69 ± 4.28% of the encapsulated oil is released upon exposure to gastrointestinal conditions and becomes available for absorption.[118]
Omega-3 fatty acidChia, camelina and echium oilseeds and wet microalgal lipidsSodium caseinate and lactose (oil to wall ratio 1:4 (w/w))Spray dryingInlet air at 170 °C, compressed air pressure of 0.5 MPa, air flow of 700 L min−1 and aspiration 70%Produce microencapsulated lipid extracts from sources of omega-3Particles size ranged between 1.5 and 30 μm and they presented a spherical shape and a smooth surface without cracks; the chia fatty acid ethyl esters microcapsules had the best microencapsulation efficiency of 76.9%, while the echium microcapsules had the highest payload of 142 mg/g.[14]
Omega-3 fatty acidFlaxseed oil4% gum Arabic and 16% soy protein isolateSpray dryingInlet air at 150 °C, outlet at 80–85 °C and flow rate of 4 mL min−1Evaluate the effect of flaxseed oil nano-encapsulation on stability, physical, color, rheological, textural, and organoleptic properties in egg-free cakeNanoencapsulated flaxseed oil used as an egg replacer in cakes; had the greatest percentage of omega-3 fatty acids (30%); particle size less than 100 nm; encapsulation efficiency of 72%; moisture content of 4% and peroxide value of 1.1 meq/kg.[119]
Polyunsaturated fatty acidsWalnut oilFructooligosaccharide/soybean protein isolate (20% w/w)Freeze dryingTemperature − 46 °C, pressure 4.1 Pa for 48 hEvaluate the fructooligosaccharide/soy isolate protein Microcapsules ranged between 121.51 and 162.02 µm; fructooligosaccharide reduces particle size and increases viscosity; microencapsulated walnut oil had a peroxide value of 26.84 meq/kg after 8 days of storage, compared to the walnut oil which reached a peroxide value of 74.56 meq/kg.[120]
Alpha-linolenic acidPerilla oil; (Perilla frutescens)γ-cyclodextrinInclusion complexThe formation of pseudorotaxane complexes that precipitate in aqueous mediaEvaluate the thermal stability and bioavailability α-linolenic acid from perilla oilThe complexes may serve as an effective supply of α-linolenic acid to raise plasma omega-3 fatty acid levels[121]
CinnamaldehyCinnamon essential oilCarboxymethyl cellulose and polyvinyl alcoholPickering emulsions by in situ hydrophobizationUse oleic acid as a hydrophobic compoundIncrease the shelf life of the breadNo fungal growth at 25 °C for 15 days; controlled release of cinnamon essential oil; fungal inhibition against P. digitatum in films containing 1.5 and 3% CEO.[122]
Flavonoid karanjinPongamia pinnata L. seed oilPolyuriaInterfacial polymerization400–500 rpm slow mixingEvaluate the insecticidal activity of microencapsulated P. pinnata oilHigh encapsulation efficiency of 87.41%; release kinetics was y = −0.0042 x + 6.4205; effective protection against Aphis gossypii (71.8%) and Bemisia tabaci (74.7%).[123]
Monoterpene α-pineneJuniper berry essential oilGum Arabic/maltodextrin (1:1); (oil to wall ratio 1:4, w/w)Spray dryingInlet air at 120 °C, outlet at 80 °C and 3.2 cm3 min−1 of feed flow rateEvaluate properties of microcapsulesParticle size was 10.83 µm ± 1.86; encapsulation efficiency of 70.07% and a retention efficiency of 82.66%; powder has the following characteristics: 4.92% moisture, 10.18% hygroscopicity, 63.80% solubility, 72.83% porosity, and 3.23 min of dissolving time; depending on the kind of wall material, it took between 15 and 45 min for the oil to completely discharge; presents antimicrobial and antifungal activity; it can be used as a food preservative.[124]
EugenolClove essential oilChitosan nanoparticles (oil to wall ratio 0.5:1)Emulsification (o/w) and ionic gelationHomogenize at 13,000 rpm for 10 min in ice bath conditions; for ionic gelation of the chitosan, sodium tripolyphosphate was added and agitated for 40 minImproved antioxidant and antimicrobial activity by nanoencapsulation of clove essential oilParticle size of 295.8 ± 45.6 nm; high retention rate (73.4%); high in vitro antimicrobial activity against Listeria monocytogenes, Staphylococcus aureus, Salmonella typhi and E. coli (a 4.80 to 4.78 cm inhibitory halo).[125]
Alpha-tocopherolWheat germ oil1.5% Sodium alginate and 2% pectinAir atomizationO/W emulsion dropped; 5% (w/v) calcium chloride solution agitated for 30 minIncrease antioxidant activity and thus shelf life and nutritional value of cookiesEncapsulation efficiency of 55.97%; maximum antioxidant activity of 41.1%; improved storage stability and shelf life of cookies; microencapsulated α-tocopherol can serve as an antioxidant to avoid autoxidation in fat-based bakery products.[104]
Alpha-tocopherol Palm oilMaltodextrin and sodium caseinate; (Core to wall ratio 1)Spray drying1.5 mm nozzle diameter, 10 mL min−1 of feed flow rate, 55 kgf cm2 air pressure, 20,000–25,000 rpm atomization speed, inlet air at 110 °C and outlet at 90 °CDemonstrate the encapsulating and protective capacity of the wall material for the microencapsulation of vitamin EEncapsulation efficiency of 59.9 ± 0.017 to 71.5 ± 0.027%; particle size from 13 to 29 µm; moisture content from 4.5 to 4.98%, microcapsule considered soluble due to the short solubility time of 178 to 251 s.[126]
Alpha-tocopherolPalm fatty acid distillateGalactomannan and gum acaciaSpray dryingInlet air at 180–200 °C and outlet at 90 °CEvaluate emulsion and oxidative stabilityEncapsulation efficiency between 60.68 and 70.01%; microcapsules ranged between 16 µm and 11 µm; a yield between 53.15 and 64.09%; moisture content from 3.40 to 3.08%; microencapsulation improved oxidative stability and absorption of vitamin E.[127]
Alpha-tocopherolVitamin EWhey protein isolate; (Core to wall ratio 1:3)(1) Spray drying; (2) Freeze-drying; (3) Spray freeze-drying(1) Inlet temperature 100 °C, outlet temperature 80 °C and 4 mL min−1 feed flow rate;
(2) Temperature − 25 °C for 2 h and 7.6 × 102 Torr to 0.8 Torr vacuum; shelf temperature between-25 °C and 20 °C for 16–18 h, after that 25 °C for 2 h;
Evaluate the effect of the three techniques of vitamin E microencapsulationEncapsulation efficiencies and particle size of 90% and 195.8 µm for spray dried microcapsules, 86% and 279 µm for freeze-dried microcapsules, 89% and 145.3 µm for spray freeze-dried microcapsules, respectively; the rats showed plasma vitamin E concentrations of 7.35 at 4 h, 7.69 at 4 h, 9.45 µg/mL at 3 h; area [103]
(3) The temperature was kept between −25 °C and −10 °C with a vacuum of 0.8 torr, and then brought to 10 °C with a vacuum of 0.3 torr under the curve were 109.84, 104.38 and 124.46 µg/(mL × h); spray freeze-drying microencapsulation improved the oral bioavailability by 1.13 and 1.19-fold compared to other techniques.
Vitamin EPalm oilMaltodextrin/Sodium caseinate (3:2:1)Freeze dryingTemperature −41 °C and pressure 4 × 10−4 mbarEffect on vitamin E encapsulation with selenomethionineEncapsulated vitamin E with 5.6 mg selenomethionine improves solubility and bioavailability; particle size was 3.00 µm ± 0.55; release rates of vitamin E after 30 min in simulated gastric fluid solution and simulated intestinal fluid solution were 87% and 42%, respectively.[105]
Alpha-tocopherolVitamin EPolycaprolactoneSupercritical fluid extraction of emulsionsPressure 8 MPa and temperature 40 °C; CO2 flow rate of 7.2 kg h−1 kg−1 emulsion and acetone as solventDemonstrate the feasibility of supercritical fluid technique in the nano-encapsulation of liquid lipophilic compoundsHigh encapsulation efficiency of 90%; particle size was from 8 and 276 nm; spherical, core-shell, and non-aggregated nanocapsules were formed, according to morphological analyses; higher storage stability between 6 and 12 months.[128]
Alpha-tocopherolVitamin EPolycaprolactoneNanoprecipitationAt 30 °C in an ultrasonic bath, PCL was dissolved in acetone, lecithin, acetone-methanol mixture (60 to 40%, v/v) and α-tocopherolImprove the carboxymethylcellulose film in the production of active packaging with α-tocopherol nanocapsulesHigh encapsulation efficiency of 88.43–99.66%; particle size ranged between 201.6 and 230.2 nm; alpha-tocopherol nanocapsules’ release behavior from CMC films might be best described by the Higuchi kinetic model; the maximum radical scavenging activity (68.85%) was found in films with 70% nanocapsules.[129]
Alpha-tocoferolVitamin EAcid hydrolysis-carboxymethyl starch (H-CMS) and xanthan gum (XG)Spray-dryingInlet air temperature 190 ± 5 °C and outlet air temperature 80 ± 5 °CImproved bioavailabilityMicrocapsules produced with substitution grades of 0.44 and a ratio of 1:20 (H-CMS/XG) demonstrated higher specific delivery in the small intestine, releasing 38.32% and 61.68% of vitamin E into simulated gastric and intestinal fluids, respectively[130]
Alpha-tocoferolVitamin EGelatin and gum ArabicComplex coacervationAdjust the pH with acetic acid (10% v/v) to 4–4.5 (isoelectric point of gelatin and gum arabic), at a speed of 1500 rpm, temperature 45 °C for 90 minOptimize by response surface methodology the conditions for vitamin E microencapsulationHigh encapsulation efficiency (93.42%), when the core material is 4 g and surfactant is 0.5% (%w/v); particle size was from 4 to 80 µm.[18]
Alpha-tocopherolVitamin ENano-hydroxyapatite as a Pickering stabilizerPickering emulsionsUses a mixer with continuous mode; O/W ratio of 20/80 (v/v)Improved bioavailability, bioaccessibility and stabilityParticle sizes were 7.53, 11.56 and 17.72 μm; improved bioaccessibility of vitamin E by 10.87 ± 1.04% for gelatin and 18.07 ± 2.90% for fortified milk[131]

3.4. Polyphenols

Polyphenols include a heterogeneous group of bioactive compounds containing at least one phenolic ring, naturally found in fruits, rhizomes, cereals, coffee, and tea, among others [132]. Anti-inflammatory, antioxidant, and anticancer properties have been associated with them, and they also participate in the prevention of diabetes, cardiovascular diseases, and cancer [133,134]. During its encapsulation, biopolymers are used, which stand out for being biodegradable, biocompatible, and high nutritional value vehicles. Among the most widely used biopolymers are carriers based on proteins, polysaccharides, and lipids [135].
Proteins represent excellent carriers of polyphenols due to their functional properties of emulsification, amphiphilicity, gelation, and foam formation. They can form nanoemulsions, nanogels, nanoparticles, nanofibers, and nanofilms, producing hydrophilic and hydrophobic polyphenolic compounds [136]. Likewise, polysaccharides, due to their structure and physiological activity, represent very suitable nanocarriers for encapsulating and administering polyphenols. Lipids represent one of the most important vehicles for the protection and supply of fat-soluble polyphenols due to their high biodegradability and biocompatibility. Through the use of liposomes, nanoemulsions, and solid lipid nanoparticles, the bioavailability of fat-soluble polyphenols can be improved [132]. In general, for the encapsulation of plant polyphenols, different types of biobased nanocarriers can be used, such as nanomicelles, nanoemulsions, nanoparticles, nanogels, and liposomes, including soy protein, albumin, zein, cellulose, starch, and lipids, which efficiently manage and protect polyphenolic compounds, as well as improve their bioavailability [132,134].

3.5. Carotenoids

Carotenoids are isoprenoid compounds found in plants, bacteria, fungi, and algae, as well as in foods such as fruits, vegetables, and fish. There are more than 600 fat-soluble carotenoids that differ by structural changes in their polyene skeleton, subdividing them into carotenoids, xanthophylls, and lycopene [137]. Different biological actions are attributed to it, such as important antioxidant activity in the prevention of age-related macular and cardiovascular diseases, as well as participating in the strengthening of the immune system, proper functioning of the reproductive system, regression of malignant lesions, and inhibition of mutagenesis [138]. They are sensitive to degradation by heat, light, and temperature. Likewise, the presence of oxidizing agents and oxygen could cause their decomposition [139], demonstrating the necessity to encapsulate them to stabilize and properly administer these bioactive ingredients.
Among the methods used for its preservation, microencapsulation is one of the most used. It allows stabilizing and reducing the volume of the product, thus reducing storage and transport costs, as well as favoring its administration. Among the oldest and most widely used carotenoid microencapsulation methods, the spray-drying allows for obtaining a product with the appropriate characteristics by varying the drying parameters such as feed flow rate and temperature, while different agents can also be used as carriers. Currently, among the most modern methods for the encapsulation of carotenoids, supercritical micronization is used, which allows for controlling the particle size and distribution of the particles, stability, and bioavailability of the product [138,140].

3.6. Pigments

Anthocyanins, carotenoids, betalains, and chlorophylls are natural pigments widely used in the food industry [138]. Among them, anthocyanins are water-soluble pigments that compose the vacuoles of several plant tissues, mainly in flowers and fruits, leaves, stems, and storage organs. The total content and composition vary substantially between different plant species [141,142]. However, they are susceptible to degradation during processing and storage due to pH, light, heat, oxygen, and interaction with other food components, decreasing their bioavailability and biological activity [143]. Microencapsulation using biopolymers based on proteins and polysaccharides of food origin is one of the most common methods to stabilize and protect anthocyanins from degradation. However, in a physiological environment, these systems could be unstable, due to high particle size and low efficiency of encapsulation of anthocyanins and zeta potential. Currently, for the stabilization of anthocyanins, nanosystems, such as nanoemulsion and nanoliposome, are being used. Unlike conventional emulsions, nanoemulsions provide higher stability against aggregation and gravitational separation due to their high surface area. Likewise, nanoliposomes made from conventional liposomes, in which the particle size has been reduced by ultrasound, high-pressure homogenization, maintained the stability, bioavailability, and controlled administration of anthocyanins [141].
Betalains are found in leaves, flowers, fruits, roots, stems, bracts, petioles, and seeds. They are synthesized from tyrosine and are divided into two subclasses: betaxanthins and betacyanins [144]. Antioxidant and anticancer properties have been attributed to them [145]. However, they are highly unstable and can be degraded according to variations in temperature, light, oxygen, and pH [146]. It has been reported that the encapsulation of betalains allows storage stability of up to 6 months. Moreover, the encapsulated betalains could be resistant to degradation and added to liquid or solid substances [147,148].
For the encapsulation of betalains, different matrices have been used, such as polysaccharides and proteins, or a combination of both, due to the high solubility of most polysaccharides and the high capacity of these compounds to protect betalains from oxidation [146]. Among the techniques used for the encapsulation of betalains, the most common are lyophilization, emulsions, ionic gelation and spray drying. The latter is the most used due to its cheaper cost and faster processing. However, it requires that the carrier be soluble in easily evaporable solvents at high temperatures, which could degrade some bioactive compounds. The lyophilization technique presents a higher loading efficiency and stabilization of betalains compared to spray drying. Likewise, it is reported that guar gum is suitable for encapsulating betalains with the lyophilization technique. However, it is not suitable for use with the spray-drying method [147,148].

3.7. Encapsulating Efficiency

Encapsulating efficiency is the ratio of a core material that is trapped inside an encapsulate as opposed to the initial core content that was introduced to an encapsulation process. It was determined using Equation (1) [149]:
Encapsulation   efficiency   % = W a W b × 100 %
where W a represents the amount of integrated core material and W b represents the total amount of core material originally added during preparation.
Spectroscopic or chromatographic methods can be used to determine W a and W b . Encapsulation method and particle size distribution can influence encapsulation efficiency [149]. The criteria that can affect encapsulation efficiency are wall and core specifications, core/wall ratio, emulsion properties (such as viscosity and droplet size) and drying parameters [150]. Encapsulation efficiency can be improved by selecting wall materials with different functional properties, and it is preferable to use various combinations of wall or shell materials [151,152]. Low efficiency might lead to low stability since capsules are not protected from adverse storage conditions. Currently, the primary focus of food component micro/nano-encapsulation is on enhancing encapsulation efficiency and prolonging product shelf-life [153,154].

3.8. Release Characteristics and Kinetics of Microcapsules of Bioactive Components

Bioactive components are released from encapsulants in three steps: surface release, diffusion via a swelling matrix, and matrix erosion [155,156]. More mechanisms were discussed, such as matrix degradation induced by enzymes and pH, fissure development, hydrostatic pressure, and geometric changes brought on by shear forces [155,157]. Surface release may be caused by inadequate entrapment inside the matrix or by a polar bioactive compound’s tendency to partition toward the hydrophilic surface of an emulsion [158].
Concentration-time profiles are frequently used to qualitatively evaluate release kinetics, with a focus on the quantity released after a certain length of time [113]. There have been several mathematical models developed, that describe the release of bioactive components. The following kinetic models were used: The zero order (Equation (2)), first order (Equation (3)), Hixson–Crowell (Equation (4)), and Korsmeyer–Peppas (Equation (5)) [155,159,160]:
Q t = Q 0   + k 0 t
Q t = Q 0   · e k 1 t
Q 0 1 / 3 Q t 1 / 3 = Q 0   + k H C t
M t M = k t n ; ( M t / M   0.6 )
where Q t   is the quantity released after time t; Q 0   is the initial quantity which is usually zero; M t   is the cumulative release at time t; M is the cumulative release at infinite time; k 0 , k 1 , k H C , k represent the zero-order, first-order, Hixson–Crowell, and Korsmeyer–Peppas kinetic constants, respectively; n represents the diffusion exponent in the Korsmeyer–Peppas model. To characterize release from porous matrixes, the zero-order and first-order models are commonly utilized. In the Hixon-Crowell model, a linear graph of the cubic root of the unreleased fraction of capsule vs time demonstrates that the surface area of the composite changes during the release process [159]. According to R2 values, the Hixson-Crowell model was comparable to zero-order and first-order kinetics [155]. The Korsmeyer–Peppas model is typically utilized when the release mechanism is unknown or when the release process is regulated by more than one sort of release mechanism [159].
The releasing action is therefore known to be triggered by a range of internal (such as diffusion, degradation, and swelling) and external stimuli (such as changes in temperature, pH, light, ultrasound, ionic strength, and magnetic field). It is evident that numerous models have been established that may predict the release profile of active agents into the environment by taking into account the occurrence of different kinetic rates. Several researchers have attempted to forecast their experimental release data using various models such as the zero-order, first-order, Korsmeyer–Peppas model, and Hixson Crowell model. The Korsmeyer–Peppas model of bioactive compounds release is found to be used in the majority of investigations on the release of corrosion inhibitor from nanocontainers [161].
The encapsulation efficiency, release kinetics of bioactive compounds, stability of capsules, and matrix characteristics are all essential features that are affected by the technique and encapsulation materials [162]. The size of the encapsulating does not directly affect the efficiency of encapsulation, but it does affect the controlled release of bioactive compounds as well as the physical and chemical characteristics of encapsulants. This is due to the fact that smaller particle sizes result in bigger contact surfaces, resulting in more release than larger particle size encapsulates. Preferably, wall material should be insoluble, not react with bioactive components, good at producing films, and possess the appropriate protective qualities against a variety of external factors [162,163].
Finally, it has been found that most study include the encapsulation efficiency in their results, as it is one of the crucial parameters to assess the preservation of the core material (bioactive compound). However, there are few studies that cover the characteristics and release kinetics of bioactive compound microcapsules in food and agricultural applications, because kinetic models are more used in the realization of a simulation, evaluating the diffusion process that occurs through the nanocontainer, and making some assumptions appropriate for a specific application.

4. Novel Technologies for Encapsulation of Bioactive Compounds

The population is demanding healthy foods with higher added value, such as functional foods or foods that contain special nutrients. The food industry faces technological changes to meet the needs of consumers. For these reasons, the technology must be appropriate to process these types of foods. The micro- and nanoencapsulation of bioactive compounds, such as antioxidants, carotenoids, tocopherols, phenols, and bioactive peptides, among others, is mostly done by spraying. Currently, lyophilization is also used to microencapsulate bioactive compounds. Extrusion is another technique that has been used in this food and agricultural field.
In this sense, different techniques can be used to microencapsulate bioactive compounds. The selection of the technology depends on several factors, such as the types of bioactive compounds to be encapsulated, the sensitivity of the bioactive compound, the encapsulating agent, and the costs of the technique [164]. Each microencapsulation technique presents different results in the final product. The bioactive compound in the final product may have properties in terms of shape, structure, size, distribution, and bioavailability. Microencapsulation has some uses expressed in Table 3.

4.1. Supercritical Microencapsulation

Currently, supercritical microencapsulation is also known as “micronization”. Supercritical micronization uses mild temperatures, which avoids reducing the quality of bioactive compounds. CO2 is the most used solvent in this technique, followed by water [138]. CO2, under supercritical conditions, has become increasingly popular in the microencapsulation of bioactive compounds of interest to the food and agricultural industries [165,166,167].
Table 3. Technologies used for the microencapsulation of bioactive compounds.
Table 3. Technologies used for the microencapsulation of bioactive compounds.
Microencapsulation
Technique
MethodApplicationsReferences
PhysicalSpray-dryingPhenolic acids, carotenoids[122]
Spray chillingPigments
Spray coatingPigments
Supercritical microencapsulation—micronizationCarotenoids
Ionic gelationNutraceutical
CocrystallizationFood ingredients
Freeze-dryingCarotenoids, Pigments
Fluidized bed coatingCarotenoids
Centrifugal extrusionFood ingredients
ChemicalInterfacial polymerizationFood[138,166,168,169]
Molecular inclusionNutraceutical
In situ polymerizationNutraceutical
Physical-chemicalCoacervationVolatile flavor oils[138,166,168,169]
Complex coacervationLycopene
Emulsion-solvent evaporationFood ingredients
Solidification emulsionFood ingredients
LiposomesFood ingredients, nutraceuticals

4.2. Complex Coacervation

Complex coacervation is an encapsulation technique that consists of the interaction of two polyelectrolytes of opposite charges in an aqueous medium. The chemical complex of a protein or carbohydrate nature is formed around the bioactive compound to be encapsulated. This interaction is formed under specific conditions of ionic strength, temperature, pH, polymer concentration, the proportion of biopolymers, the molecular weight of biopolymers, and the degree of homogenization [168]. This technique allows for obtaining particles, where the core (bioactive compounds) is protected by a layer of an encapsulating agent of a protein or carbohydrate nature [169,170].

4.3. Spray Chilling

The application of vitamin C (ascorbic acid) has demonstrated beneficial effects in agriculture, such as reducing the effect of nickel and cadmium in barley [171,172], alleviating drought stress in sweet pepper [173] and Cucumber [174], and improving the qualities of apples [175]. It allows for mitigating the salinity stress and providing better growth performance in barley [176]. In this context, the encapsulation of ascorbic acid through spray chilling has been reported as an effective technique, showing retentions over 90% and controlled release of vitamin C in aqueous solutions [177,178]. These authors employed a combination of fully hydrogenated palm oil and palm oil to obtain flexible particles. The technique involves heating the fully hydrogenated palm oil to 80 °C to make it liquid and mixing it with the ascorbic acid to pass immediately through the spray dryer and be cooled to 5 °C.

4.4. Other Encapsulation Techniques

Some essential oils have the potential to be applied in agriculture as biopesticides. Rosemary oil has the potential to be used as an insecticide [179,180,181]. It has been encapsulated by spouted bed drying using a mixture of Tixosil and maltodextrin as drying carriers and Teflon beads for drying [182] and with microcrystalline cellulose cores [183]. Neem oil has shown insecticide activity [184,185], and it can be effectively nanoencapsulated to control pests [185,186,187], showing enhanced properties.

5. Applications of Encapsulated Products

Encapsulation materials consist of a variety of synthetic or natural polymers [188], the former still being the most used elsewhere. Otherwise, an increase in the use of natural polymers is seen, and non-hydrolyzed biopolymers have been widely used in agriculture due to their availability and low cost. Among them, hydrolyzed starches stand out as agents in the encapsulation of Bt pesticides (metabolites produced by Bacillus thuringiensis) because they protected environmental factors and improvement in the formulated product [7]. Moreover, encapsulation can be efficient for formulations of biofungicides, biopesticides, and/or biofertilizers in agricultural fields, becoming an economically viable technique for farmers [12]. The encapsulation can have several advantages in the agricultural sector [189], as shown in Figure 3.
Taking into account agricultural applications, [190] report Trichoderma harzianum as a biological control agent very sensitive to biotic and abiotic factors due to the presence of live spores. Consequently, encapsulation shows an improvement in activity against phytopathogens such as in the control of Sclerotinia sclerotiorum (white mold). In another study, spores of the Trichoderma species were encapsulated in biologically based lignin for the treatment of diseases of the vine trunk, demonstrated by in vitro tests that the spores remain at rest until germination is triggered by the fungus itself at the correct time [191]. Additionally, the fungus Trichoderma spp. was encapsulated in a sodium alginate matrix to guarantee good stability of the biopesticide formulations. The encapsulation strategy resulted in the survival of the fungi during the production and storage stages. After 14 months, the stored samples still had good conditions and viable cells [192].
In the case of microencapsulation by Pickering emulsion of Metarhizium conidia for the mortality of Spodoptera littoralis larvae, better cell distribution in leaves and control of the agricultural pest were observed [193]. Furthermore, in a study carried out, a formulation of Bacillus thuringiensis aizawai (BtA) encapsulated by Pickering water-in-oil emulsion was obtained. The efficiency of 92% in the mortality of Spodoptera littoralis larvae in the first instar was reported [194].
Similarly, it was shown that inoculating two strains of Pseudomonas fluorescens in potatoes using an alginate-gelatin capsule resulted in higher plant protection against harmful soil conditions and establishment in the rhizosphere [195]. Likewise, Pseudomonas spp. was encapsulated in sodium alginate and with salicylic acid containing zinc oxide nanoparticles, in which the efficiency in the antifungal activity against Sclerotium rolfsii was demonstrated [28]. Following this trend, a formulation of biopesticides with UV protection capacity in the encapsulation of fungal conidia in oil/water emulsion and stabilized by titanium dioxide (TiO2) was developed [196]. Satisfactory efficiency in their germination rate was achieved when exposed to natural UV light compared to unprotected conidia (not encapsulated).
Azadirachtin found in the neem tree (Azadirachta indica), which is a natural pesticide molecule, was encapsulated via nanoemulsification with whey protein isolate. This strategy demonstrated a positive effect on the mortality of Spodoptera frugiperda, a caterpillar that affects the soybean crop [10]. Moreover, the encapsulated essential oil extracted from orange peel effectively inhibited the growth of Staphylococcus aureus and Escherichia coli [197].
Regarding the encapsulation of biofertilizers, a study showed that formulations containing Burkholderia cepacia and Pseudomonas fluorescens encapsulated with phosphate alginate presented better growth conditions for wheat plants in semi-arid and salt-stressed areas [6]. The study demonstrated that biofertilizers consisting of Pseudomonas fluorescens and Azosprillum brasilense encapsulated with polymers of montmorillonite (clay mineral) and sodium alginate resulted in higher control of the formulation [11]. Furthermore, slow release of the active compounds was reached, thus contributing positively to the growth of the wheat crop with an increase in the biomass and aerial part of the plants.
Regarding the food industry, food encapsulation has been widely used over the years [5,115,198]. The encapsulation technique aims to protect and deliver the bioactive compounds to the target tissue of the human organism. This approach results in better stability and bioavailability of the bioactive compounds and increases their use and benefits for the human body [199]. Figure 4 shows some foods used in encapsulation technology. As presented in the scheme, the main interesting compounds for the food encapsulation technique are dyes and flavorings (candies), vitamins (vitamin A, D, E), antioxidants (citrus fruits and cereals), enzymes (lipase, invertase), bioactive peptide (milk), and polyunsaturated fatty acids (omega 3), and some mineral elements (iron, potassium), among others.
In addition to substances produced by microorganisms, substances found in herbal extracts and essential oils are also targeted to be encapsulated because they have antimicrobial and insecticidal properties, among others [21,200]. The control of the high volatility of oils and extracts is a challenge for biotechnology. Therefore, the encapsulation of essential oils and volatile extracts in the development of bioproducts for agricultural use is of great interest because the release of these compounds to the target agent is desirable to be slow and continuous [201].
Essential oil from savory leaves (Satureja hortensis L.), for example, was encapsulated with different natural polymers such as apple pectin, gum arabic, and gelatin. A higher encapsulation efficiency for all polymers was reported. Consequently, the efficiency of herbicidal activity in amaranth (Amaranthus retroflexus L.) and tomato (Lycopersicon esculentum Mill.) was increased [202]. Another active oil has been encapsulated and tested over time against bacteria oxidation: pepper oil. The encapsulation of such oil by gum arabic/maltodextrin resulted in an inhibitory effect against Pseudomonas aeruginosa, Enterococcus faecali and Staphylococcus aureus. The microencapsulation using these polymers contributed to a higher protection of pepper oil against the oxidation process during storage [203].
Curcumin extracted from turmeric was encapsulated by the Supercritical Anti-solvent (SAS) technique using various polymeric matrices to formulate a dye containing turmeric extracts. The encapsulating polymers tested were Eudragit® L100 (Evonik-Germany), Pluronic® F127 (BASF-Germany), and polyvinylpyrrolidone or mixtures of these materials. The dye formulation had a curcumin content of 4.45 µg mL−1 with a mean diameter of amorphous particles of approximately 5.7 µm. Also, the best dye solubility of 211 μg mL−1 of total curcuminoids was obtained at pH 4 [204]. Economic evaluations have also been developed in this area to provide information that can boost applications on larger scales and transfer the technology to the industrial scale. In this case, the technical-economic feasibility of the SAS technique for food applications was evaluated in terms of the precipitation of a vitamin complex containing riboflavin, δ-tocopherol, and β-carotene in zein microcapsules. The results showed that the average size of the microcapsules ranged from 8 to 18 µm, with spherical morphology, while the precipitation yield ranged from 410 to 820 g kg−1. The simulated manufacturing cost for the microcapsules ranged from USD 0.38 g−1 to USD 0.50 g−1, while these values may reduce with the possibility of reducing the mass flow of CO2 during precipitation [205].
Poly(ε-caprolactone) particles containing resveratrol were developed using the Gas Anti-solvent (GAS) technique, having heterogeneous characteristics. Encapsulation was positive because it did not change the chemical structures or the antioxidant activity of resveratrol. In addition, the microparticles maintained a constant release for 48 h and, in the thermal oxidative analysis, the difference between the samples and the control was 2.57 times smaller than the difference between pure resveratrol and the control sample [206]. Curcumin nanoparticles were also produced by encapsulation in Shellac by the GAS technique. After 2 months of storage at room temperature (25 °C), the aqueous nanodispersions of the particles with the encapsulated bioactive substance (Cur@Lac) showed long-term stability without agglomeration or sedimentation. Furthermore, the final retention rate of curcumin was over 85%. Colored drinks made from fruit syrup (pH = 4.0), fruit soda (pH = 6.0) and sparkling water (pH = 8.3) were prepared using Cur@Lac as a dye, which has a yellow-orange color [207].

6. Concluding Remarks and Future Trends

An overview of the latest research, technologies, and advances regarding the application of the encapsulation of bioactive compounds for food and agricultural applications is presented. The main forms of encapsulation are microencapsulation, followed by nanoencapsulation. A crescent interest exists in the development of micro or nanocapsules loaded with polyphenols, carotenoids, fatty acids, phytosterols, probiotics, vitamins, minerals, and bioactive peptides from natural sources, for the food industry. However, for applications in agriculture, the development of particles loaded with essential oils, lipids, phytotoxins, medicines, vaccines, hemoglobin, and microbial metabolites is the focus. The main purposes of the encapsulation strategy are to improve stability, solubility, bioavailability, sensorial properties, retention of bioactive properties and microstructure, reduction of hygroscopicity, and increase shelf life. The most widely used polymers for their high performance are gum Arabic, starch, and chitin, while the most commonly used technologies for encapsulation are emulsions-spray drying, emulsions-freeze drying, complex coacervation, followed by others, such as the emerging technology known as supercritical microencapsulation. More studies are needed on bioavailability for food applications and the effectiveness of bioactivity in agricultural applications, as well as scale-up studies at both pilot and industrial levels. However, the potential of encapsulation into polymeric matrices makes it a good strategy to protect bioactive compounds and broaden their use in both food and agriculture.

Author Contributions

Conceptualization, G.L.Z. and L.O.-M.; methodology, G.L.Z., I.B. and L.O.-M.; writing—original draft preparation, G.L.Z., F.S.R., L.P.O., M.V.T., H.P., J.S.C.-R., E.H., I.B. and L.O.-M.; writing—review and editing, G.L.Z. and L.O.-M., H.P.; visualization, G.L.Z., I.B. and L.O.-M.; supervision, G.L.Z. and L.O.-M.; project administration, G.L.Z., I.B. and L.O.-M.; funding acquisition, G.L.Z., I.B. and L.O.-M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Council for Scientific and Technological Development (CNPq), grant number 308067/2021–5” and “The APC was funded by Universidad San Ignacio de Loyola (USIL).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Genovese, A.; Balivo, A.; Salvati, A.; Sacchi, R. Functional Ice Cream Health Benefits and Sensory Implications. Food Res. Int. 2022, 161, 111858. [Google Scholar] [CrossRef]
  2. Riley, S.; Lale, A.; Nguyen, V.; Xi, H.; Wilkinson, K.; Searle, I.R.; Fisk, I. Volatile Profiles of Commercial Vetch Prepared via Different Processing Methods. Food Chem. 2022, 395, 133569. [Google Scholar] [CrossRef]
  3. Mafra, J.F.; de Santana, T.S.; Cruz, A.I.C.; Ferreira, M.A.; Miranda, F.M.; Araújo, F.M.; Ribeiro, P.R.; Evangelista-Barreto, N.S. Influence of Red Propolis on the Physicochemical, Microbiological and Sensory Characteristics of Tilapia (Oreochromis niloticus) Salami. Food Chem. 2022, 394, 133502. [Google Scholar] [CrossRef] [PubMed]
  4. Yang, Y.; Yuan, B.; Yu, P.; Jia, Y.; Zhou, Q.; Sun, J. Flavor Characteristics of Peanut Butter Pretreated by Radio Frequency Heating, Explosion Puffing, Microwave, and Oven Heating. Food Chem. 2022, 394, 133487. [Google Scholar] [CrossRef] [PubMed]
  5. Zabot, G.L.; Silva, E.K.; Emerick, L.B.; Felisberto, M.H.F.; Clerici, M.T.P.S.; Meireles, M.A.A. Physicochemical, Morphological, Thermal and Pasting Properties of a Novel Native Starch Obtained from Annatto Seeds. Food Hydrocoll. 2019, 89, 321–329. [Google Scholar] [CrossRef]
  6. Riseh, R.S.; Ebrahimi-Zarandi, M.; Tamanadar, E.; Pour, M.M.; Thakur, V.K. Salinity Stress: Toward Sustainable Plant Strategies and Using Plant Growth-Promoting Rhizobacteria Encapsulation for Reducing It. Sustainability 2021, 13, 12758. [Google Scholar] [CrossRef]
  7. Do Nascimento Junior, D.R.; Tabernero, A.; Cabral Albuquerque, E.C.d.M.; Vieira de Melo, S.A.B. Biopesticide Encapsulation Using Supercritical CO2: A Comprehensive Review and Potential Applications. Molecules 2021, 26, 4003. [Google Scholar] [CrossRef]
  8. De Oliveira, J.L.; Fraceto, L.F.; Bravo, A.; Polanczyk, R.A. Encapsulation Strategies for Bacillus thuringiensis: From Now to the Future. J. Agric. Food Chem. 2021, 69, 4564–4577. [Google Scholar] [CrossRef]
  9. Vassilev, N.; Vassileva, M.; Martos, V.; Garcia del Moral, L.F.; Kowalska, J.; Tylkowski, B.; Malusá, E. Formulation of Microbial Inoculants by Encapsulation in Natural Polysaccharides: Focus on Beneficial Properties of Carrier Additives and Derivatives. Front. Plant Sci. 2020, 11, 270. [Google Scholar] [CrossRef]
  10. Bae, M.; Lewis, A.; Liu, S.; Arcot, Y.; Lin, Y.-T.; Bernal, J.S.; Cisneros-Zevallos, L.; Akbulut, M. Novel Biopesticides Based on Nanoencapsulation of Azadirachtin with Whey Protein to Control Fall Armyworm. J. Agric. Food Chem. 2022, 70, 7900–7910. [Google Scholar] [CrossRef]
  11. Kadmiri, I.M.; El Mernissi, N.; Azaroual, S.E.; Mekhzoum, M.E.M.; Qaiss, A.E.K.; Bouhfid, R. Bioformulation of Microbial Fertilizer Based on Clay and Alginate Encapsulation. Curr. Microbiol. 2021, 78, 86–94. [Google Scholar] [CrossRef] [PubMed]
  12. Riseh, R.S.; Skorik, Y.A.; Thakur, V.K.; Pour, M.M.; Tamanadar, E.; Noghabi, S.S. Encapsulation of Plant Biocontrol Bacteria with Alginate as a Main Polymer Material. Int. J. Mol. Sci. 2021, 22, 11165. [Google Scholar] [CrossRef] [PubMed]
  13. Ali, E.A.; Nada, A.A.; Al-Moghazy, M. Self-Stick Membrane Based on Grafted Gum Arabic as Active Food Packaging for Cheese Using Cinnamon Extract. Int. J. Biol. Macromol. 2021, 189, 114–123. [Google Scholar] [CrossRef] [PubMed]
  14. Castejón, N.; Luna, P.; Señoráns, F.J. Microencapsulation by Spray Drying of Omega-3 Lipids Extracted from Oilseeds and Microalgae: Effect on Polyunsaturated Fatty Acid Composition. LWT 2021, 148, 111789. [Google Scholar] [CrossRef]
  15. Guo, Z.; Ge, X.; Li, W.; Yang, L.; Han, L.; Yu, Q. Active-Intelligent Film Based on Pectin from Watermelon Peel Containing Beetroot Extract to Monitor the Freshness of Packaged Chilled Beef. Food Hydrocoll. 2021, 119, 106751. [Google Scholar] [CrossRef]
  16. Mirmazloum, I.; Ladányi, M.; Omran, M.; Papp, V.; Ronkainen, V.-P.; Pónya, Z.; Papp, I.; Némedi, E.; Kiss, A. Co-Encapsulation of Probiotic Lactobacillus acidophilus and Reishi Medicinal Mushroom (Ganoderma lingzhi) Extract in Moist Calcium Alginate Beads. Int. J. Biol. Macromol. 2021, 192, 461–470. [Google Scholar] [CrossRef]
  17. Sodeinde, K.O.; Ojo, A.M.; Olusanya, S.O.; Ayanda, O.S.; Adeoye, A.O.; Dada, T.M.; Lawal, O.S. Cellulose Isolated from Delonixregia Pods: Characterisation and Application in the Encapsulation of Vitamin, A. Ind. Crops Prod. 2021, 160, 113138. [Google Scholar] [CrossRef]
  18. Köksal, E.; Bayram, O.; Göde, F.; Aktaş, A.H. Microencapsulation of Vitamin E: Optimization and Characterization of Complex Coacervation Conditions Using Response Surface Methodology. Sak. Univ. J. Sci. 2021, 25, 906–913. [Google Scholar] [CrossRef]
  19. Yang, C.; Wang, Y.; Lu, L.; Unsworth, L.; Guan, L.L.; Chen, L. Oat Protein-Shellac Beads: Superior Protection and Delivery Carriers for Sensitive Bioactive Compounds. Food Hydrocoll. 2018, 77, 754–763. [Google Scholar] [CrossRef]
  20. Guía-García, J.L.; Charles-Rodríguez, A.V.; Reyes-Valdés, M.H.; Ramírez-Godina, F.; Robledo-Olivo, A.; García-Osuna, H.T.; Cerqueira, M.A.; Flores-López, M.L. Micro and Nanoencapsulation of Bioactive Compounds for Agri-Food Applications: A Review. Ind. Crops Prod. 2022, 186, 115198. [Google Scholar] [CrossRef]
  21. Marcillo-Parra, V.; Tupuna-Yerovi, D.S.; González, Z.; Ruales, J. Encapsulation of Bioactive Compounds from Fruit and Vegetable By-Products for Food Application—A Review. Trends Food Sci. Technol. 2021, 116, 11–23. [Google Scholar] [CrossRef]
  22. Díaz-Torres, R.D.C.; Alonso-Castro, A.J.; Carrillo-Inungaray, M.L.; Carranza-Alvarez, C. Bioactive Compounds Obtained from Plants, Their Pharmacological Applications and Encapsulation. In Phytomedicine; Academic Press: Cambridge, MA, USA, 2021; pp. 181–205. ISBN 978-0-12-824109-7. [Google Scholar]
  23. Rehman, A.; Ahmad, T.; Aadil, R.M.; Spotti, M.J.; Bakry, A.M.; Khan, I.M.; Zhao, L.; Riaz, T.; Tong, Q. Pectin Polymers as Wall Materials for the Nano-Encapsulation of Bioactive Compounds. Trends Food Sci. Technol. 2019, 90, 35–46. [Google Scholar] [CrossRef]
  24. Ashter, S.A. Chemistry of Cellulosic Polymers. In Technology and Applications of Polymers Derived from Biomass; William Andrew Publishing: Norwich, NY, USA, 2018; pp. 57–74. ISBN 978-0-323-51115-5. [Google Scholar]
  25. Hosseinnejad, M.; Jafari, S.M. Evaluation of Different Factors Affecting Antimicrobial Properties of Chitosan. Int. J. Biol. Macromol. 2016, 85, 467–475. [Google Scholar] [CrossRef]
  26. Shah, P. Polymers in Food. In Polymer Science and Innovative Applications; Elsevier: Amsterdam, The Netherlands, 2020; pp. 567–592. ISBN 978-0-12-816808-0. [Google Scholar]
  27. Hossain, F.; Follett, P.; Salmieri, S.; Vu, K.D.; Fraschini, C.; Lacroix, M. Antifungal Activities of Combined Treatments of Irradiation and Essential Oils (EOs) Encapsulated Chitosan Nanocomposite Films in in Vitro and in Situ Conditions. Int. J. Food Microbiol. 2019, 295, 33–40. [Google Scholar] [CrossRef] [PubMed]
  28. Panichikkal, J.; Prathap, G.; Nair, R.A.; Krishnankutty, R.E. Evaluation of Plant Probiotic Performance of Pseudomonas Sp. Encapsulated in Alginate Supplemented with Salicylic Acid and Zinc Oxide Nanoparticles. Int. J. Biol. Macromol. 2021, 166, 138–143. [Google Scholar] [CrossRef] [PubMed]
  29. Farahani, Z.K.; Mousavi, M.; Ardebili, S.M.S.; Bakhoda, H. Modification of Sodium Alginate by Octenyl Succinic Anhydride to Fabricate Beads for Encapsulating Jujube Extract. Curr. Res. Food Sci. 2022, 5, 157–166. [Google Scholar] [CrossRef] [PubMed]
  30. Taban, A.; Saharkhiz, M.J.; Kavoosi, G. Development of Pre-Emergence Herbicide Based on Arabic Gum-Gelatin, Apple Pectin and Savory Essential Oil Nano-Particles: A Potential Green Alternative to Metribuzin. Int. J. Biol. Macromol. 2021, 167, 756–765. [Google Scholar] [CrossRef]
  31. Rajabi, H.; Jafari, S.M.; Rajabzadeh, G.; Sarfarazi, M.; Sedaghati, S. Chitosan-Gum Arabic Complex Nanocarriers for Encapsulation of Saffron Bioactive Components. Colloids Surf. Physicochem. Eng. Asp. 2019, 578, 123644. [Google Scholar] [CrossRef]
  32. Singh, P.; Magalhães, S.; Alves, L.; Antunes, F.; Miguel, M.; Lindman, B.; Medronho, B. Cellulose-Based Edible Films for Probiotic Entrapment. Food Hydrocoll. 2019, 88, 68–74. [Google Scholar] [CrossRef]
  33. El-Naggar, M.E.; Hasanin, M.; Youssef, A.M.; Aldalbahi, A.; El-Newehy, M.H.; Abdelhameed, R.M. Hydroxyethyl Cellulose/Bacterial Cellulose Cryogel Dopped Silver@titanium Oxide Nanoparticles: Antimicrobial Activity and Controlled Release of Tebuconazole Fungicide. Int. J. Biol. Macromol. 2020, 165, 1010–1021. [Google Scholar] [CrossRef]
  34. Ibrahim, S.S.; Abou-Elseoud, W.S.; Elbehery, H.H.; Hassan, M.L. Chitosan-Cellulose Nanoencapsulation Systems for Enhancing the Insecticidal Activity of Citronella Essential Oil against the Cotton Leafworm Spodoptera Littoralis. Ind. Crops Prod. 2022, 184, 115089. [Google Scholar] [CrossRef]
  35. Rajkumar, V.; Gunasekaran, C.; Paul, C.A.; Dharmaraj, J. Development of Encapsulated Peppermint Essential Oil in Chitosan Nanoparticles: Characterization and Biological Efficacy against Stored-Grain Pest Control. Pestic. Biochem. Physiol. 2020, 170, 104679. [Google Scholar] [CrossRef] [PubMed]
  36. Seyedabadi, M.M.; Rostami, H.; Jafari, S.M.; Fathi, M. Development and Characterization of Chitosan-Coated Nanoliposomes for Encapsulation of Caffeine. Food Biosci. 2021, 40, 100857. [Google Scholar] [CrossRef]
  37. Cargnin, M.A.; Gasparin, B.C.; dos Santos Rosa, D.; Paulino, A.T. Performance of Lactase Encapsulated in Pectin-Based Hydrogels during Lactose Hydrolysis Reactions. LWT 2021, 150, 111863. [Google Scholar] [CrossRef]
  38. Sun, C.; Cao, J.; Wang, Y.; Huang, L.; Chen, J.; Wu, J.; Zhang, H.; Chen, Y.; Sun, C. Preparation and Characterization of Pectin-Based Edible Coating Agent Encapsulating Carvacrol/HPβCD Inclusion Complex for Inhibiting Fungi. Food Hydrocoll. 2022, 125, 107374. [Google Scholar] [CrossRef]
  39. Mei, S.; Han, P.; Wu, H.; Shi, J.; Tang, L.; Jiang, Z. One-Pot Fabrication of Chitin-Shellac Composite Microspheres for Efficient Enzyme Immobilization. J. Biotechnol. 2018, 266, 1–8. [Google Scholar] [CrossRef] [PubMed]
  40. Sun, C.; Xu, C.; Mao, L.; Wang, D.; Yang, J.; Gao, Y. Preparation, Characterization and Stability of Curcumin-Loaded Zein-Shellac Composite Colloidal Particles. Food Chem. 2017, 228, 656–667. [Google Scholar] [CrossRef]
  41. Wang, K.; Sui, J.; Gao, W.; Yu, B.; Yuan, C.; Guo, L.; Cui, B.; Abd El-Aty, A.M. Effects of Xanthan Gum and Sodium Alginate on Gelatinization and Gels Structure of Debranched Pea Starch by Pullulanase. Food Hydrocoll. 2022, 130, 107733. [Google Scholar] [CrossRef]
  42. Lara, G.; Yakoubi, S.; Villacorta, C.M.; Uemura, K.; Kobayashi, I.; Takahashi, C.; Nakajima, M.; Neves, M.A. Spray Technology Applications of Xanthan Gum-Based Edible Coatings for Fresh-Cut Lotus Root (Nelumbo nucifera). Food Res. Int. 2020, 137, 109723. [Google Scholar] [CrossRef]
  43. Akbari-Alavijeh, S.; Shaddel, R.; Jafari, S.M. Encapsulation of Food Bioactives and Nutraceuticals by Various Chitosan-Based Nanocarriers. Food Hydrocoll. 2020, 105, 105774. [Google Scholar] [CrossRef]
  44. Raza, Z.A.; Khalil, S.; Ayub, A.; Banat, I.M. Recent Developments in Chitosan Encapsulation of Various Active Ingredients for Multifunctional Applications. Carbohydr. Res. 2020, 492, 108004. [Google Scholar] [CrossRef] [PubMed]
  45. Ambaye, T.G.; Vaccari, M.; Prasad, S.; van Hullebusch, E.D.; Rtimi, S. Preparation and Applications of Chitosan and Cellulose Composite Materials. J. Environ. Manage. 2022, 301, 113850. [Google Scholar] [CrossRef] [PubMed]
  46. Chen, S.; Han, Y.; Jian, L.; Liao, W.; Zhang, Y.; Gao, Y. Fabrication, Characterization, Physicochemical Stability of Zein-Chitosan Nanocomplex for Co-Encapsulating Curcumin and Resveratrol. Carbohydr. Polym. 2020, 236, 116090. [Google Scholar] [CrossRef] [PubMed]
  47. Bayer, J.; Granda, L.A.; Méndez, J.A.; Pèlach, M.A.; Vilaseca, F.; Mutjé, P. Cellulose Polymer Composites (WPC). In Advanced High Strength Natural Fibre Composites in Construction; Woodhead Publishing Series in Composites Science and Engineering; Woodhead Publishing, Ltd.: Sawston, UK, 2017; pp. 115–139. ISBN 978-0-08-100411-1. [Google Scholar]
  48. Keshk, S.M.A.S.; El-Kott, A.F. Natural Bacterial Biodegradable Medical Polymers: Bacterial Cellulose. In Science and Principles of Biodegradable and Bioresorbable Medical Polymers: Materials and Properties; Woodhead Publishing Series in Biomaterials; Woodhead Publishing, Ltd.: Sawston, UK, 2017; pp. 295–319. ISBN 978-0-08-100372-5. [Google Scholar]
  49. Moura, I.G.; Sá, A.V.; Abreu, A.S.L.; Machado, A.V.A. Bioplastics from Agro-Wastes for Food Packaging Applications. In Food Packaging; Nanotechnology in the Agri-Food Industry; Academic Press: Cambridge, MA, USA, 2017; pp. 223–263. ISBN 978-0-12-804302-8. [Google Scholar]
  50. Bagheriasl, D.; Carreau, P.J. Polymer-Cellulose Nanocrystal (CNC) Nanocomposites. In Processing of Polymer Nanocomposites; Hanser: Minhen, Germany, 2019; pp. 371–393. ISBN 978-1-56990-635-4. [Google Scholar]
  51. Avérous, L.; Halley, P.J. Starch Polymers: From the Field to Industrial Products. In Starch Polymers: From Genetic Engineering to Green Applications; Elsevier: Amsterdam, The Netherlands, 2014; pp. 3–10. ISBN 978-0-444-53730-0. [Google Scholar]
  52. Prasathkumar, M.; Sadhasivam, S. Chitosan/Hyaluronic Acid/Alginate and an Assorted Polymers Loaded with Honey, Plant, and Marine Compounds for Progressive Wound Healing—Know-How. Int. J. Biol. Macromol. 2021, 186, 656–685. [Google Scholar] [CrossRef]
  53. Park, J.; Nam, J.; Yun, H.; Jin, H.-J.; Kwak, H.W. Aquatic Polymer-Based Edible Films of Fish Gelatin Crosslinked with Alginate Dialdehyde Having Enhanced Physicochemical Properties. Carbohydr. Polym. 2021, 254, 117317. [Google Scholar] [CrossRef] [PubMed]
  54. Yang, Z.; Li, M.; Zhai, X.; Zhao, L.; Tahir, H.E.; Shi, J.; Zou, X.; Huang, X.; Li, Z.; Xiao, J. Development and Characterization of Sodium Alginate/Tea Tree Essential Oil Nanoemulsion Active Film Containing TiO2 Nanoparticles for Banana Packaging. Int. J. Biol. Macromol. 2022, 213, 145–154. [Google Scholar] [CrossRef]
  55. Flamminii, F.; Paciulli, M.; Di Michele, A.; Littardi, P.; Carini, E.; Chiavaro, E.; Pittia, P.; Di Mattia, C.D. Alginate-Based Microparticles Structured with Different Biopolymers and Enriched with a Phenolic-Rich Olive Leaves Extract: A Physico-Chemical Characterization. Curr. Res. Food Sci. 2021, 4, 698–706. [Google Scholar] [CrossRef]
  56. Thombare, N.; Kumar, S.; Kumari, U.; Sakare, P.; Yogi, R.K.; Prasad, N.; Sharma, K.K. Shellac as a Multifunctional Biopolymer: A Review on Properties, Applications and Future Potential. Int. J. Biol. Macromol. 2022, 215, 203–223. [Google Scholar] [CrossRef]
  57. Yuan, Y.; Zhang, X.; Pan, Z.; Xue, Q.; Wu, Y.; Li, Y.; Li, B.; Li, L. Improving the Properties of Chitosan Films by Incorporating Shellac Nanoparticles. Food Hydrocoll. 2021, 110, 106164. [Google Scholar] [CrossRef]
  58. Ma, J.; Zhou, Z.; Li, K.; Li, K.; Liu, L.; Zhang, W.; Xu, J.; Tu, X.; Du, L.; Zhang, H. Novel Edible Coating Based on Shellac and Tannic Acid for Prolonging Postharvest Shelf Life and Improving Overall Quality of Mango. Food Chem. 2021, 354, 129510. [Google Scholar] [CrossRef]
  59. Soradech, S.; Nunthanid, J.; Limmatvapirat, S.; Luangtana-anan, M. Utilization of Shellac and Gelatin Composite Film for Coating to Extend the Shelf Life of Banana. Food Control 2017, 73, 1310–1317. [Google Scholar] [CrossRef]
  60. Yabe, T. New Understanding of Pectin as a Bioactive Dietary Fiber: A Review. J. Food Bioact. 2018, 3, 95–100. [Google Scholar] [CrossRef] [Green Version]
  61. Lin, L.; Xu, W.; Liang, H.; He, L.; Liu, S.; Li, Y.; Li, B.; Chen, Y. Construction of PH-Sensitive Lysozyme/Pectin Nanogel for Tumor Methotrexate Delivery. Colloids Surf. B Biointerfaces 2015, 126, 459–466. [Google Scholar] [CrossRef] [PubMed]
  62. Noreen, A.; Nazli, Z.-H.; Akram, J.; Rasul, I.; Mansha, A.; Yaqoob, N.; Iqbal, R.; Tabasum, S.; Zuber, M.; Zia, K.M. Pectins Functionalized Biomaterials; a New Viable Approach for Biomedical Applications: A Review. Int. J. Biol. Macromol. 2017, 101, 254–272. [Google Scholar] [CrossRef]
  63. Neufeld, L.; Bianco-Peled, H. Pectin–Chitosan Physical Hydrogels as Potential Drug Delivery Vehicles. Int. J. Biol. Macromol. 2017, 101, 852–861. [Google Scholar] [CrossRef]
  64. Da, S.; Gulão, E.; de Souza, C.J.F.; Andrade, C.T.; Garcia-Rojas, E.E. Complex Coacervates Obtained from Peptide Leucine and Gum Arabic: Formation and Characterization. Food Chem. 2016, 194, 680–686. [Google Scholar] [CrossRef]
  65. Riseh, R.S.; Tamanadar, E.; Pour, M.M.; Thakur, V.K. Novel Approaches for Encapsulation of Plant Probiotic Bacteria with Sustainable Polymer Gums: Application in the Management of Pests and Diseases. Adv. Polym. Technol. 2022, 2022, 4419409. [Google Scholar] [CrossRef]
  66. Mirhosseini, H.; Tan, C.P.; Hamid, N.S.A.; Yusof, S. Effect of Arabic Gum, Xanthan Gum and Orange Oil Contents on ζ-Potential, Conductivity, Stability, Size Index and PH of Orange Beverage Emulsion. Colloids Surf. Physicochem. Eng. Asp. 2008, 315, 47–56. [Google Scholar] [CrossRef]
  67. Sworn, G. Xanthan Gum. In Handbook of Hydrocolloids; Woodhead Publishing Series in Food Science, Technology and Nutrition; Woodhead: Sawston, UK, 2021; pp. 833–853. ISBN 978-0-12-820104-6. [Google Scholar]
  68. Noor, I.S.M.; Majid, S.R.; Arof, A.K.; Djurado, D.; Claro Neto, S.; Pawlicka, A. Characteristics of Gellan Gum–LiCF3SO3 Polymer Electrolytes. Solid State Ion. 2012, 225, 649–653. [Google Scholar] [CrossRef]
  69. Lopes, B.D.; Lessa, V.L.; Silva, B.M.; Carvalho, M.A.D.; Schnitzler, E.; Lacerda, L.G. Xanthan Gum: Properties, Production Conditions, Quality and Economic Perspective. J. Food Nutr. Res. 2015, 54, 185–194. [Google Scholar]
  70. Kaewprapan, K.; Baros, F.; Marie, E.; Inprakhon, P.; Durand, A. Macromolecular Surfactants Synthesized by Lipase-Catalyzed Transesterification of Dextran with Vinyl Decanoate. Carbohydr. Polym. 2012, 88, 313–320. [Google Scholar] [CrossRef]
  71. Broaders, K.E.; Grandhe, S.; Fréchet, J.M.J. A Biocompatible Oxidation-Triggered Carrier Polymer with Potential in Therapeutics. J. Am. Chem. Soc. 2011, 133, 756–758. [Google Scholar] [CrossRef] [PubMed]
  72. Fathi, M.; Martín, Á.; McClements, D.J. Nanoencapsulation of Food Ingredients Using Carbohydrate Based Delivery Systems. Trends Food Sci. Technol. 2014, 39, 18–39. [Google Scholar] [CrossRef]
  73. Chen, H.; Wang, J.; Cheng, Y.; Wang, C.; Liu, H.; Bian, H.; Pan, Y.; Sun, J.; Han, W. Application of Protein-Based Films and Coatings for Food Packaging: A Review. Polymers 2019, 11, 2039. [Google Scholar] [CrossRef] [Green Version]
  74. Rashidinejad, A.; Tarhan, O.; Rezaei, A.; Capanoglu, E.; Boostani, S.; Khoshnoudi-Nia, S.; Samborska, K.; Garavand, F.; Shaddel, R.; Akbari-Alavijeh, S.; et al. Addition of Milk to Coffee Beverages; the Effect on Functional, Nutritional, and Sensorial Properties. Crit. Rev. Food Sci. Nutr. 2022, 62, 6132–6152. [Google Scholar] [CrossRef]
  75. Du, X.; Jing, H.; Wang, L.; Huang, X.; Mo, L.; Bai, X.; Wang, H. PH-Shifting Formation of Goat Milk Casein Nanoparticles from Insoluble Peptide Aggregates and Encapsulation of Curcumin for Enhanced Dispersibility and Bioactivity. LWT 2022, 154, 112753. [Google Scholar] [CrossRef]
  76. Garavand, F.; Jafarzadeh, S.; Cacciotti, I.; Vahedikia, N.; Sarlak, Z.; Tarhan, Ö.; Yousefi, S.; Rouhi, M.; Castro-Muñoz, R.; Jafari, S.M. Different Strategies to Reinforce the Milk Protein-Based Packaging Composites. Trends Food Sci. Technol. 2022, 123, 1–14. [Google Scholar] [CrossRef]
  77. Tang, C. Assembled Milk Protein Nano-Architectures as Potential Nanovehicles for Nutraceuticals. Adv. Colloid Interface Sci. 2021, 292, 102432. [Google Scholar] [CrossRef]
  78. Daniloski, D.; Petkoska, A.T.; Lee, N.A.; Bekhit, A.E.-D.; Carne, A.; Vaskoska, R.; Vasiljevic, T. Active Edible Packaging Based on Milk Proteins: A Route to Carry and Deliver Nutraceuticals. Trends Food Sci. Technol. 2021, 111, 688–705. [Google Scholar] [CrossRef]
  79. Nollet, M.; Laurichesse, E.; Besse, S.; Soubabère, O.; Schmitt, V. Determination of Formulation Conditions Allowing Double Emulsions Stabilized by PGPR and Sodium Caseinate to Be Used as Capsules. Langmuir 2018, 34, 2823–2833. [Google Scholar] [CrossRef]
  80. Sittipummongkol, K.; Chuysinuan, P.; Techasakul, S.; Pisitsak, P.; Pechyen, C. Core Shell Microcapsules of Neem Seed Oil Extract Containing Azadirachtin and Biodegradable Polymers and Their Release Characteristics. Polym. Bull. 2019, 76, 3803–3817. [Google Scholar] [CrossRef]
  81. Shishir, M.R.I.; Xie, L.; Sun, C.; Zheng, X.; Chen, W. Advances in Micro and Nano-Encapsulation of Bioactive Compounds Using Biopolymer and Lipid-Based Transporters. Trends Food Sci. Technol. 2018, 78, 34–60. [Google Scholar] [CrossRef]
  82. Daneshniya, M.; Nezhad, H.J.; Maleki, M.H.; Jalali, V.; Behrouzian, M. A Review of Encapsulation of Bioactive Peptides with Antimicrobial and Antioxidant Activity. Int. J. Acad. Eng. Res. 2020, 4, 7. [Google Scholar]
  83. Premi, M.; Sharma, H.K. Effect of Different Combinations of Maltodextrin, Gum Arabic and Whey Protein Concentrate on the Encapsulation Behavior and Oxidative Stability of Spray Dried Drumstick (Moringa oleifera) Oil. Int. J. Biol. Macromol. 2017, 105, 1232–1240. [Google Scholar] [CrossRef]
  84. Corrêa-Filho, L.; Moldão-Martins, M.; Alves, V. Advances in the Application of Microcapsules as Carriers of Functional Compounds for Food Products. Appl. Sci. 2019, 9, 571. [Google Scholar] [CrossRef] [Green Version]
  85. Yang, Y.; Fang, Z.; Chen, X.; Zhang, W.; Xie, Y.; Chen, Y.; Liu, Z.; Yuan, W. An Overview of Pickering Emulsions: Solid-Particle Materials, Classification, Morphology, and Applications. Front. Pharmacol. 2017, 8, 287. [Google Scholar] [CrossRef] [Green Version]
  86. Li, K.-Y.; Zhou, Y.; Huang, G.-Q.; Li, X.-D.; Xiao, J.-X. Preparation of Powdered Oil by Spray Drying the Pickering Emulsion Stabilized by Ovalbumin—Gum Arabic Polyelectrolyte Complex. Food Chem. 2022, 391, 133223. [Google Scholar] [CrossRef]
  87. Klettenhammer, S.; Ferrentino, G.; Morozova, K.; Scampicchio, M. Novel Technologies Based on Supercritical Fluids for the Encapsulation of Food Grade Bioactive Compounds. Foods 2020, 9, 1395. [Google Scholar] [CrossRef]
  88. Olivera-Montenegro, L.; Best, I.; Gil-Saldarriaga, A. Effect of Pretreatment by Supercritical Fluids on Antioxidant Activity of Protein Hydrolyzate from Quinoa (Chenopodium quinoa Willd.). Food Sci. Nutr. 2021, 9, 574–582. [Google Scholar] [CrossRef]
  89. Akbarian, M.; Khani, A.; Eghbalpour, S.; Uversky, V.N. Bioactive Peptides: Synthesis, Sources, Applications, and Proposed Mechanisms of Action. Int. J. Mol. Sci. 2022, 23, 1445. [Google Scholar] [CrossRef]
  90. Olivera-Montenegro, L.; Bugarin, A.; Marzano, A.; Best, I.; Zabot, G.L.; Romero, H. Production of Protein Hydrolysate from Quinoa (Chenopodium quinoa Willd.): Economic and Experimental Evaluation of Two Pretreatments Using Supercritical Fluids’ Extraction and Conventional Solvent Extraction. Foods 2022, 11, 1015. [Google Scholar] [CrossRef] [PubMed]
  91. Akbarbaglu, Z.; Mahdi Jafari, S.; Sarabandi, K.; Mohammadi, M.; Khakbaz Heshmati, M.; Pezeshki, A. Influence of Spray Drying Encapsulation on the Retention of Antioxidant Properties and Microstructure of Flaxseed Protein Hydrolysates. Colloids Surf. B Biointerfaces 2019, 178, 421–429. [Google Scholar] [CrossRef] [PubMed]
  92. Aguilar-Toalá, J.E.; Quintanar-Guerrero, D.; Liceaga, A.M.; Zambrano-Zaragoza, M.L. Encapsulation of Bioactive Peptides: A Strategy to Improve the Stability, Protect the Nutraceutical Bioactivity and Support Their Food Applications. RSC Adv. 2022, 12, 6449–6458. [Google Scholar] [CrossRef] [PubMed]
  93. Sarabandi, K.; Gharehbeglou, P.; Jafari, S.M. Spray-Drying Encapsulation of Protein Hydrolysates and Bioactive Peptides: Opportunities and Challenges. Dry. Technol. 2020, 38, 577–595. [Google Scholar] [CrossRef]
  94. Delshadi, R.; Bahrami, A.; Tafti, A.G.; Barba, F.J.; Williams, L.L. Micro and Nano-Encapsulation of Vegetable and Essential Oils to Develop Functional Food Products with Improved Nutritional Profiles. Trends Food Sci. Technol. 2020, 104, 72–83. [Google Scholar] [CrossRef]
  95. Angane, M.; Swift, S.; Huang, K.; Butts, C.A.; Quek, S.Y. Essential Oils and Their Major Components: An Updated Review on Antimicrobial Activities, Mechanism of Action and Their Potential Application in the Food Industry. Foods 2022, 11, 464. [Google Scholar] [CrossRef]
  96. Hammoud, Z.; Ben Abada, M.; Greige-Gerges, H.; Elaissari, A.; Mediouni Ben Jemâa, J. Insecticidal Effects of Natural Products in Free and Encapsulated Forms: An Overview. J. Nat. Pestic. Res. 2022, 1, 100007. [Google Scholar] [CrossRef]
  97. Ribeiro-Santos, R.; Andrade, M.; Sanches-Silva, A. Application of Encapsulated Essential Oils as Antimicrobial Agents in Food Packaging. Curr. Opin. Food Sci. 2017, 14, 78–84. [Google Scholar] [CrossRef]
  98. Niki, E.; Abe, K. Vitamin E: Structure, Properties and Functions. In Food Chemistry, Function and Analysis; Chapter 1; Niki, E., Ed.; Royal Society of Chemistry: Cambridge, UK, 2019; pp. 1–11. ISBN 978-1-78801-240-9. [Google Scholar]
  99. Samanta, S. Fat-Soluble Vitamins. In Nutrition and Functional Foods in Boosting Digestion, Metabolism and Immune Health; Elsevier: Amsterdam, The Netherlands, 2022; pp. 329–364. ISBN 978-0-12-821232-5. [Google Scholar]
  100. Ribeiro, A.M.; Estevinho, B.N.; Rocha, F. Improvement of Vitamin E Microencapsulation and Release Using Different Biopolymers as Encapsulating Agents. Food Bioprod. Process. 2021, 130, 23–33. [Google Scholar] [CrossRef]
  101. Carella, A.M.; Benvenuto, A.; Lagattolla, V.; Arbaiza, T.; De Luca, P.; Ciavarrella, G.; Modola, G.; Di Pumpo, M.; Ponziano, E.; Benvenuto, M. Vitamin Supplements in the Era of SARS-Cov2 Pandemic. GSC Biol. Pharm. Sci. 2020, 11, 7–19. [Google Scholar] [CrossRef]
  102. Trela, A.; Szymańska, R. Less Widespread Plant Oils as a Good Source of Vitamin E. Food Chem. 2019, 296, 160–166. [Google Scholar] [CrossRef] [PubMed]
  103. Parthasarathi, S.; Anandharamakrishnan, C. Enhancement of Oral Bioavailability of Vitamin E by Spray-Freeze Drying of Whey Protein Microcapsules. Food Bioprod. Process. 2016, 100, 469–476. [Google Scholar] [CrossRef]
  104. Kaur, K.; Singh, J.; Singh, V. Effect of Encapsulated Vitamin E on Physical, Storage and Retention Parameters in Cookies. J. Food Sci. Technol. 2020, 57, 3509–3517. [Google Scholar] [CrossRef] [PubMed]
  105. Lazim, N.A.M.; Muhamad, I.I. Encapsulation of Vitamin e Using Maltodextrin/Sodium Caseinate/Selenomethionine and Its Release Study. Chem. Eng. Trans. 2017, 56, 1951–1956. [Google Scholar] [CrossRef]
  106. Galli, F.; Azzi, A.; Birringer, M.; Cook-Mills, J.M.; Eggersdorfer, M.; Frank, J.; Cruciani, G.; Lorkowski, S.; Özer, N.K. Vitamin E: Emerging Aspects and New Directions. Free Radic. Biol. Med. 2017, 102, 16–36. [Google Scholar] [CrossRef]
  107. Rajanna, D.; Pushpadass, H.A.; Emerald, F.M.E.; Padaki, N.V.; Nath, B.S. Nanoencapsulation of casein-derived Peptides within Electrospun Nanofibres. J. Sci. Food Agric. 2022, 102, 1684–1698. [Google Scholar] [CrossRef]
  108. Chotphruethipong, L.; Battino, M.; Benjakul, S. Effect of Stabilizing Agents on Characteristics, Antioxidant Activities and Stability of Liposome Loaded with Hydrolyzed Collagen from Defatted Asian Sea Bass Skin. Food Chem. 2020, 328, 127127. [Google Scholar] [CrossRef]
  109. Hassan, M.A.; Xavier, M.; Gupta, S.; Nayak, B.B.; Balange, A.K. Antioxidant Properties and Instrumental Quality Characteristics of Spray Dried Pangasius Visceral Protein Hydrolysate Prepared by Chemical and Enzymatic Methods. Environ. Sci. Pollut. Res. 2019, 26, 8875–8884. [Google Scholar] [CrossRef]
  110. Alvarado, Y.; Muro, C.; Illescas, J.; Díaz, M.d.C.; Riera, F. Encapsulation of Antihypertensive Peptides from Whey Proteins and Their Releasing in Gastrointestinal Conditions. Biomolecules 2019, 9, 164. [Google Scholar] [CrossRef] [Green Version]
  111. Sarabandi, K.; Sadeghi Mahoonak, A.; Hamishekar, H.; Ghorbani, M.; Jafari, S.M. Microencapsulation of Casein Hydrolysates: Physicochemical, Antioxidant and Microstructure Properties. J. Food Eng. 2018, 237, 86–95. [Google Scholar] [CrossRef]
  112. Murthy, L.N.; Phadke, G.G.; Mohan, C.O.; Chandra, M.V.; Annamalai, J.; Visnuvinayagam, S.; Unnikrishnan, P.; Ravishankar, C.N. Characterization of Spray-Dried Hydrolyzed Proteins from Pink Perch Meat Added with Maltodextrin and Gum Arabic. J. Aquat. Food Prod. Technol. 2017, 26, 913–928. [Google Scholar] [CrossRef]
  113. Giroux, H.J.; Robitaille, G.; Britten, M. Controlled Release of Casein-Derived Peptides in the Gastrointestinal Environment by Encapsulation in Water-in-Oil-in-Water Double Emulsions. LWT-Food Sci. Technol. 2016, 69, 225–232. [Google Scholar] [CrossRef]
  114. Rao, P.S.; Bajaj, R.K.; Mann, B.; Arora, S.; Tomar, S.K. Encapsulation of Antioxidant Peptide Enriched Casein Hydrolysate Using Maltodextrin–Gum Arabic Blend. J. Food Sci. Technol. 2016, 53, 3834–3843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Xu, W.; Sun, H.; Li, H.; Li, Z.; Zheng, S.; Luo, D.; Ning, Y.; Wang, Y.; Shah, B.R. Preparation and Characterization of Tea Oil Powder with High Water Solubility Using Pickering Emulsion Template and Vacuum Freeze-Drying. LWT 2022, 160, 113330. [Google Scholar] [CrossRef]
  116. Kouamé, K.J.E.-P.; Bora, A.F.M.; Li, X.; Sun, Y.; Liu, L. Novel Trends and Opportunities for Microencapsulation of Flaxseed Oil in Foods: A Review. J. Funct. Foods 2021, 87, 104812. [Google Scholar] [CrossRef]
  117. Yakdhane, A.; Labidi, S.; Chaabane, D.; Tolnay, A.; Nath, A.; Koris, A.; Vatai, G. Microencapsulation of Flaxseed Oil—State of Art. Processes 2021, 9, 295. [Google Scholar] [CrossRef]
  118. Gañan, N.; Bordón, M.G.; Ribotta, P.D.; González, A. Study of Chia Oil Microencapsulation in Soy Protein Microparticles Using Supercritical CO2-Assisted Impregnation. J. CO2 Util. 2020, 40, 101221. [Google Scholar] [CrossRef]
  119. Murugkar, D.A.; Zanwar, A.A.; Shrivastava, A. Effect of Nano-Encapsulation of Flaxseed Oil on the Stability, Characterization and Incorporation on the Quality of Eggless Cake. Appl. Food Res. 2021, 1, 100025. [Google Scholar] [CrossRef]
  120. Lin, D.; Xiao, L.; Li, S.; Qin, W.; Loy, D.A.; Chen, H.; Zhang, Q. Effects of Fructooligosaccharide and Soybean Protein Isolate in the Microencapsulation of Walnut Oil. Ind. Crops Prod. 2022, 177, 114431. [Google Scholar] [CrossRef]
  121. Yoshikiyo, K.; Yoshioka, Y.; Narumiya, Y.; Oe, S.; Kawahara, H.; Kurata, K.; Shimizu, H.; Yamamoto, T. Thermal Stability and Bioavailability of Inclusion Complexes of Perilla Oil with γ-Cyclodextrin. Food Chem. 2019, 294, 56–59. [Google Scholar] [CrossRef] [PubMed]
  122. Fasihi, H.; Noshirvani, N.; Hashemi, M.; Fazilati, M.; Salavati, H.; Coma, V. Antioxidant and Antimicrobial Properties of Carbohydrate-Based Films Enriched with Cinnamon Essential Oil by Pickering Emulsion Method. Food Packag. Shelf Life 2019, 19, 147–154. [Google Scholar] [CrossRef]
  123. Purkait, A.; Mukherjee, A.; Hazra, D.K.; Roy, K.; Biswas, P.K.; Kole, R.K. Encapsulation, Release and Insecticidal Activity of Pongamia pinnata (L.) Seed Oil. Heliyon 2021, 7, e06557. [Google Scholar] [CrossRef] [PubMed]
  124. Bajac, J.; Nikolovski, B.; Lončarević, I.; Petrović, J.; Bajac, B.; Đurović, S.; Petrović, L. Microencapsulation of Juniper Berry Essential Oil (Juniperus communis L.) by Spray Drying: Microcapsule Characterization and Release Kinetics of the Oil. Food Hydrocoll. 2022, 125, 107430. [Google Scholar] [CrossRef]
  125. Hadidi, M.; Pouramin, S.; Adinepour, F.; Haghani, S.; Jafari, S.M. Chitosan Nanoparticles Loaded with Clove Essential Oil: Characterization, Antioxidant and Antibacterial Activities. Carbohydr. Polym. 2020, 236, 116075. [Google Scholar] [CrossRef]
  126. Selamat, S.N.; Mohamad, S.N.H.; Muhamad, I.I.; Khairuddin, N.; Md Lazim, N.A. Characterization of Spray-Dried Palm Oil Vitamin E Concentrate. Arab. J. Sci. Eng. 2018, 43, 6165–6169. [Google Scholar] [CrossRef]
  127. Tarigan, J.B.; Kaban, J.; Zulmi, R. Microencapsulation of Vitamin e from Palm Fatty Acid Distillate with Galactomannan and Gum Acacia Using Spray Drying Method. IOP Conf. Ser. Mater. Sci. Eng. 2018, 309, 12095. [Google Scholar] [CrossRef]
  128. Prieto, C.; Calvo, L. Supercritical Fluid Extraction of Emulsions to Nanoencapsulate Vitamin E in Polycaprolactone. J. Supercrit. Fluids 2017, 119, 274–282. [Google Scholar] [CrossRef]
  129. Mirzaei-Mohkam, A.; Garavand, F.; Dehnad, D.; Keramat, J.; Nasirpour, A. Optimisation, Antioxidant Attributes, Stability and Release Behaviour of Carboxymethyl Cellulose Films Incorporated with Nanoencapsulated Vitamin E. Prog. Org. Coat. 2019, 134, 333–341. [Google Scholar] [CrossRef]
  130. Jiang, M.; Hong, Y.; Gu, Z.; Cheng, L.; Li, Z.; Li, C. Preparation of a Starch-Based Carrier for Oral Delivery of Vitamin E to the Small Intestine. Food Hydrocoll. 2019, 91, 26–33. [Google Scholar] [CrossRef]
  131. Ribeiro, A.; Gonçalves, R.F.S.; Pinheiro, A.C.; Manrique, Y.A.; Barreiro, M.F.; Lopes, J.C.B.; Dias, M.M. In Vitro Digestion and Bioaccessibility Studies of Vitamin E-Loaded Nanohydroxyapatite Pickering Emulsions and Derived Fortified Foods. LWT 2022, 154, 112706. [Google Scholar] [CrossRef]
  132. Zhang, Z.; Qiu, C.; Li, X.; McClements, D.J.; Jiao, A.; Wang, J.; Jin, Z. Advances in Research on Interactions between Polyphenols and Biology-Based Nano-Delivery Systems and Their Applications in Improving the Bioavailability of Polyphenols. Trends Food Sci. Technol. 2021, 116, 492–500. [Google Scholar] [CrossRef]
  133. Costa, M.; Sezgin-Bayindir, Z.; Losada-Barreiro, S.; Paiva-Martins, F.; Saso, L.; Bravo-Díaz, C. Polyphenols as Antioxidants for Extending Food Shelf-Life and in the Prevention of Health Diseases: Encapsulation and Interfacial Phenomena. Biomedicines 2021, 9, 1909. [Google Scholar] [CrossRef]
  134. Zhang, Z.; Li, X.; Sang, S.; McClements, D.J.; Chen, L.; Long, J.; Jiao, A.; Jin, Z.; Qiu, C. Polyphenols as Plant-Based Nutraceuticals: Health Effects, Encapsulation, Nano-Delivery, and Application. Foods 2022, 11, 2189. [Google Scholar] [CrossRef] [PubMed]
  135. Akbarbaglu, Z.; Peighambardoust, S.H.; Sarabandi, K.; Jafari, S.M. Spray Drying Encapsulation of Bioactive Compounds within Protein-Based Carriers; Different Options and Applications. Food Chem. 2021, 359, 129965. [Google Scholar] [CrossRef]
  136. Guan, T.; Zhang, Z.; Li, X.; Cui, S.; McClements, D.J.; Wu, X.; Chen, L.; Long, J.; Jiao, A.; Qiu, C.; et al. Preparation, Characteristics, and Advantages of Plant Protein-Based Bioactive Molecule Delivery Systems. Foods 2022, 11, 1562. [Google Scholar] [CrossRef] [PubMed]
  137. Milani, A.; Basirnejad, M.; Shahbazi, S.; Bolhassani, A. Carotenoids: Biochemistry, Pharmacology and Treatment. Br. J. Pharmacol. 2017, 174, 1290–1324. [Google Scholar] [CrossRef] [Green Version]
  138. Janiszewska-Turak, E. Carotenoids Microencapsulation by Spray Drying Method and Supercritical Micronization. Food Res. Int. 2017, 99, 891–901. [Google Scholar] [CrossRef]
  139. Gul, K.; Tak, A.; Singh, A.K.; Singh, P.; Yousuf, B.; Wani, A.A. Chemistry, Encapsulation, and Health Benefits of β-Carotene—A Review. Cogent Food Agric. 2015, 1, 1018696. [Google Scholar] [CrossRef]
  140. Eun, J.-B.; Maruf, A.; Das, P.R.; Nam, S.-H. A Review of Encapsulation of Carotenoids Using Spray Drying and Freeze Drying. Crit. Rev. Food Sci. Nutr. 2020, 60, 3547–3572. [Google Scholar] [CrossRef]
  141. Chen, B.-H.; Stephen Inbaraj, B. Nanoemulsion and Nanoliposome Based Strategies for Improving Anthocyanin Stability and Bioavailability. Nutrients 2019, 11, 1052. [Google Scholar] [CrossRef] [Green Version]
  142. Mattioli, R.; Francioso, A.; Mosca, L.; Silva, P. Anthocyanins: A Comprehensive Review of Their Chemical Properties and Health Effects on Cardiovascular and Neurodegenerative Diseases. Molecules 2020, 25, 3809. [Google Scholar] [CrossRef] [PubMed]
  143. Nayak, B.; Dahmoune, F.; Moussi, K.; Remini, H.; Dairi, S.; Aoun, O.; Khodir, M. Comparison of Microwave, Ultrasound and Accelerated-Assisted Solvent Extraction for Recovery of Polyphenols from Citrus Sinensis Peels. Food Chem. 2015, 187, 507–516. [Google Scholar] [CrossRef] [PubMed]
  144. Sadowska-Bartosz, I.; Bartosz, G. Biological Properties and Applications of Betalains. Molecules 2021, 26, 2520. [Google Scholar] [CrossRef] [PubMed]
  145. Madadi, E.; Mazloum-Ravasan, S.; Yu, J.S.; Ha, J.W.; Hamishehkar, H.; Kim, K.H. Therapeutic Application of Betalains: A Review. Plants 2020, 9, 1219. [Google Scholar] [CrossRef] [PubMed]
  146. Castro-Enríquez, D.D.; Montaño-Leyva, B.; Del Toro-Sánchez, C.L.; Juaréz-Onofre, J.E.; Carvajal-Millan, E.; Burruel-Ibarra, S.E.; Tapia-Hernández, J.A.; Barreras-Urbina, C.G.; Rodríguez-Félix, F. Stabilization of Betalains by Encapsulation—A Review. J. Food Sci. Technol. 2020, 57, 1587–1600. [Google Scholar] [CrossRef]
  147. Gandía-Herrero, F.; García-Carmona, F. Biosynthesis of Betalains: Yellow and Violet Plant Pigments. Trends Plant Sci. 2013, 18, 334–343. [Google Scholar] [CrossRef]
  148. Ravichandran, K.; Palaniraj, R.; Saw, N.M.M.T.; Gabr, A.M.M.; Ahmed, A.R.; Knorr, D.; Smetanska, I. Effects of Different Encapsulation Agents and Drying Process on Stability of Betalains Extract. J. Food Sci. Technol. 2014, 51, 2216–2221. [Google Scholar] [CrossRef] [Green Version]
  149. Piacentini, E. Encapsulation Efficiency. In Encyclopedia of Membranes; Drioli, E., Giorno, L., Eds.; Springer: Berlin/Heidelberg, Germany, 2016; pp. 706–707. ISBN 978-3-662-44323-1. [Google Scholar]
  150. Mehran, M.; Masoum, S.; Memarzadeh, M. Improvement of Thermal Stability and Antioxidant Activity of Anthocyanins of Echium Amoenum Petal Using Maltodextrin/Modified Starch Combination as Wall Material. Int. J. Biol. Macromol. 2020, 148, 768–776. [Google Scholar] [CrossRef]
  151. Alvarenga, D.; Victória de Barros, R.; Vilela, S. Chapter 12: Microencapsulation of Essential Oils Using Spray Drying Technology. In Microencapsulation and Microspheres for Food Applications; Sagis, L.M.C., Ed.; Academic Press: Cambridge, MA, USA, 2015; p. 245. ISBN 978-0-12-800350-3. [Google Scholar]
  152. Potdar, S.B.; Landge, V.K.; Barkade, S.S.; Potoroko, I.; Sonawane, S.H. Flavor Encapsulation and Release Studies in Food. In Encapsulation of Active Molecules and Their Delivery System; Elsevier: Waltham, MA, USA, 2020; p. 296. ISBN 978-0-12-819363-1. [Google Scholar]
  153. Akhavan Mahdavi, S.; Jafari, S.M.; Assadpour, E.; Ghorbani, M. Storage Stability of Encapsulated Barberry’s Anthocyanin and Its Application in Jelly Formulation. J. Food Eng. 2016, 181, 59–66. [Google Scholar] [CrossRef]
  154. Samborska, K.; Jedlińska, A. Spray Drying Encapsulation of Anthocyanins. In Spray Drying Encapsulation of Bioactive Materials; Advances in Drying Sciences and Technology; CRC Press: Boca Raton, FL, USA, 2021; p. 107. ISBN 978-0-367-36646-9. [Google Scholar]
  155. Flores, F.P.; Kong, F. In Vitro Release Kinetics of Microencapsulated Materials and the Effect of the Food Matrix. Annu. Rev. Food Sci. Technol. 2017, 8, 237–259. [Google Scholar] [CrossRef]
  156. Vinceković, M.; Jurić, S.; Đermić, E.; Topolovec-Pintarić, S. Kinetics and Mechanisms of Chemical and Biological Agents Release from Biopolymeric Microcapsules. J. Agric. Food Chem. 2017, 65, 9608–9617. [Google Scholar] [CrossRef] [PubMed]
  157. Siepmann, J.; Siepmann, F. Mathematical Modeling of Drug Delivery. Int. J. Pharm. 2008, 364, 328–343. [Google Scholar] [CrossRef] [PubMed]
  158. Rawat, A.; Burgess, D.J. USP Apparatus 4 Method for in Vitro Release Testing of Protein Loaded Microspheres. Int. J. Pharm. 2011, 409, 178–184. [Google Scholar] [CrossRef] [PubMed]
  159. Dolinina, E.S.; Akimsheva, E.Y.; Parfenyuk, E.V. Silica Microcapsules as Containers for Protein Drugs: Direct and Indirect Encapsulation. J. Mol. Liq. 2019, 287, 110938. [Google Scholar] [CrossRef]
  160. Sonawane, S.H.; Bhanvase, B.A.; Sivakumar, M. (Eds.) Encapsulation of Active Molecules and Their Delivery System, 1st ed.; Elsevier: Waltham, MA, USA, 2020; ISBN 978-0-12-819363-1. [Google Scholar]
  161. Pradhane, A.; Barai, D.; Bhanvase, B.; Sonawane, S. Mathematical Modeling and Simulation of the Release of Active Agents from Nanocontainers/Microspheres. In Encapsulation of Active Molecules and Their Delivery System; Elsevier: Waltham, MA, USA, 2020; pp. 257–291. ISBN 978-0-12-819363-1. [Google Scholar]
  162. Mehta, N.; Kumar, P.; Verma, A.K.; Umaraw, P.; Kumar, Y.; Malav, O.P.; Sazili, A.Q.; Domínguez, R.; Lorenzo, J.M. Microencapsulation as a Noble Technique for the Application of Bioactive Compounds in the Food Industry: A Comprehensive Review. Appl. Sci. 2022, 12, 1424. [Google Scholar] [CrossRef]
  163. Yun, P.; Devahastin, S.; Chiewchan, N. Microstructures of Encapsulates and Their Relations with Encapsulation Efficiency and Controlled Release of Bioactive Constituents: A Review. Compr. Rev. Food Sci. Food Saf. 2021, 20, 1768–1799. [Google Scholar] [CrossRef]
  164. Mahdavi, S.A.; Jafari, S.M.; Ghorbani, M.; Assadpoor, E. Spray-Drying Microencapsulation of Anthocyanins by Natural Biopolymers: A Review. Dry. Technol. 2014, 32, 509–518. [Google Scholar] [CrossRef]
  165. Santos, D.T.; Albarelli, J.Q.; Beppu, M.M.; Meireles, M.A.A. Stabilization of Anthocyanin Extract from Jabuticaba Skins by Encapsulation Using Supercritical CO2 as Solvent. Food Res. Int. 2013, 50, 617–624. [Google Scholar] [CrossRef] [Green Version]
  166. De Freitas Santos, P.D.; Rubio, F.T.V.; Da Silva, M.P.; Pinho, L.S.; Favaro-Trindade, C.S. Microencapsulation of Carotenoid-Rich Materials: A Review. Food Res. Int. 2021, 147, 110571. [Google Scholar] [CrossRef]
  167. Tabernero, A.; Martín del Valle, E.M.; Galán, M.A. Supercritical Fluids for Pharmaceutical Particle Engineering: Methods, Basic Fundamentals and Modelling. Chem. Eng. Process. Process Intensif. 2012, 60, 9–25. [Google Scholar] [CrossRef]
  168. Gheonea, I.; Aprodu, I.; Cîrciumaru, A.; Râpeanu, G.; Bahrim, G.E.; Stănciuc, N. Microencapsulation of Lycopene from Tomatoes Peels by Complex Coacervation and Freeze-Drying: Evidences on Phytochemical Profile, Stability and Food Applications. J. Food Eng. 2021, 288, 110166. [Google Scholar] [CrossRef]
  169. Timilsena, Y.P.; Adhikari, R.; Barrow, C.J.; Adhikari, B. Physicochemical and Functional Properties of Protein Isolate Produced from Australian Chia Seeds. Food Chem. 2016, 212, 648–656. [Google Scholar] [CrossRef] [PubMed]
  170. Marques da Silva, T.; Jacob Lopes, E.; Codevilla, C.F.; Cichoski, A.J.; de Moraes Flores, É.M.; Motta, M.H.; de Bonada Silva, C.; Grosso, C.R.F.; de Menezes, C.R. Development and Characterization of Microcapsules Containing Bifidobacterium Bb-12 Produced by Complex Coacervation Followed by Freeze Drying. LWT 2018, 90, 412–417. [Google Scholar] [CrossRef]
  171. Sabir, M.; Naseem, Z.; Ahmad, W.; Usman, M.; Nadeem, F.; Saifullah; Ahmad, H.R. Alleviation of Adverse Effects of Nickel on Growth and Concentration of Copper and Manganese in Wheat through Foliar Application of Ascorbic Acid. Int. J. Phytoremediation 2022, 24, 695–703. [Google Scholar] [CrossRef] [PubMed]
  172. Yaseen, S.; Amjad, S.F.; Mansoora, N.; Kausar, S.; Shahid, H.; Alamri, S.A.M.; Alrumman, S.A.; Eid, E.M.; Ansari, M.J.; Danish, S.; et al. Supplemental Effects of Biochar and Foliar Application of Ascorbic Acid on Physio-Biochemical Attributes of Barley (Hordeum vulgare L.) under Cadmium-Contaminated Soil. Sustainability 2021, 13, 9128. [Google Scholar] [CrossRef]
  173. Khazaei, Z.; Estaji, A. Effect of Foliar Application of Ascorbic Acid on Sweet Pepper (Capsicum annuum) Plants under Drought Stress. Acta Physiol. Plant. 2020, 42, 118. [Google Scholar] [CrossRef]
  174. Ghahremani, Z.; Mikaealzadeh, M.; Barzegar, T.; Ranjbar, M.E. Foliar Application of Ascorbic Acid and Gamma Aminobutyric Acid Can Improve Important Properties of Deficit Irrigated Cucumber Plants (Cucumis sativus Cv. Us). Gesunde Pflanz. 2021, 73, 77–84. [Google Scholar] [CrossRef]
  175. Allahveran, A.; Farokhzad, A.; Asghari, M.; Sarkhosh, A. Foliar Application of Ascorbic and Citric Acids Enhanced ‘Red Spur’ Apple Fruit Quality, Bioactive Compounds and Antioxidant Activity. Physiol. Mol. Biol. Plants 2018, 24, 433–440. [Google Scholar] [CrossRef]
  176. Hassan, A.; Amjad, S.F.; Saleem, M.H.; Yasmin, H.; Imran, M.; Riaz, M.; Ali, Q.; Joyia, F.A.; Mobeen; Ahmed, S.; et al. Foliar Application of Ascorbic Acid Enhances Salinity Stress Tolerance in Barley (Hordeum vulgare L.) through Modulation of Morpho-Physio-Biochemical Attributes, Ions Uptake, Osmo-Protectants and Stress Response Genes Expression. Saudi J. Biol. Sci. 2021, 28, 4276–4290. [Google Scholar] [CrossRef]
  177. Dos Santos Carvalho, J.D.; Oriani, V.B.; de Oliveira, G.M.; Hubinger, M.D. Characterization of Ascorbic Acid Microencapsulated by the Spray Chilling Technique Using Palm Oil and Fully Hydrogenated Palm Oil. LWT 2019, 101, 306–314. [Google Scholar] [CrossRef]
  178. Dos Santos, J.D.; Oriani, V.B.; Oliveira, G.M.; Hubinger, M.D. Solid Lipid Microparticles Loaded with Ascorbic Acid: Release Kinetic Profile during Thermal Stability. J. Food Process. Preserv. 2021, 45, e15557. [Google Scholar] [CrossRef]
  179. Jahanian, H.; Kahkeshani, N.; Sanei-Dehkordi, A.; Isman, M.B.; Saeedi, M.; Khanavi, M. Rosmarinus Officinalis as a Natural Insecticide: A Review. Int. J. Pest Manag. 2022. [Google Scholar] [CrossRef]
  180. Shawer, R.; El-Shazly, M.M.; Khider, A.M.; Baeshen, R.S.; Hikal, W.M.; Kordy, A.M. Botanical Oils Isolated from Simmondsia Chinensis and Rosmarinus Officinalis Cultivated in Northern Egypt: Chemical Composition and Insecticidal Activity against Sitophilus oryzae (L.) and Tribolium castaneum (Herbst). Molecules 2022, 27, 4383. [Google Scholar] [CrossRef]
  181. Trombin de Souza, M.; Trombin de Souza, M.; Bernardi, D.; da Oliveira, D.C.; Morais, M.C.; de Melo, D.J.; Richardi, V.S.; Zarbin, P.H.G.; Zawadneak, M.A.C. Essential Oil of Rosmarinus officinalis Ecotypes and Their Major Compounds: Insecticidal and Histological Assessment Against Drosophila suzukii and Their Impact on a Nontarget Parasitoid. J. Econ. Entomol. 2022, 115, 955–966. [Google Scholar] [CrossRef]
  182. Souza, C.R.F.; Baldim, I.; Bankole, V.O.; da Ana, R.; Durazzo, A.; Lucarini, M.; Cicero, N.; Santini, A.; Souto, E.B.; Oliveira, W.P. Spouted Bed Dried Rosmarinus Officinalis Extract: A Novel Approach for Physicochemical Properties and Antioxidant Activity. Agriculture 2020, 10, 349. [Google Scholar] [CrossRef]
  183. Benelli, L.; Oliveira, W.P. Fluidized Bed Coating of Inert Cores with a Lipid-Based System Loaded with a Polyphenol-Rich Rosmarinus Officinalis Extract. Food Bioprod. Process. 2019, 114, 216–226. [Google Scholar] [CrossRef]
  184. Dono, D.; Widayani, N.S.; Ishmayana, S.; Hidayat, Y.; Widiantini, F.; Nasahi, C. Resistance of Nilaparvata Lugens to Fenobucarb and Imidacloprid and Susceptibility to Neem Oil Insecticides. HAYATI J. Biosci. 2022, 29, 234–244. [Google Scholar] [CrossRef]
  185. Pascoli, M.; de Albuquerque, F.P.; Calzavara, A.K.; Tinoco-Nunes, B.; Oliveira, W.H.C.; Gonçalves, K.C.; Polanczyk, R.A.; Vechia, J.F.D.; de Matos, S.T.S.; de Andrade, D.J.; et al. The Potential of Nanobiopesticide Based on Zein Nanoparticles and Neem Oil for Enhanced Control of Agricultural Pests. J. Pest Sci. 2020, 93, 793–806. [Google Scholar] [CrossRef]
  186. Bagle, A.V.; Jadhav, R.S.; Gite, V.V.; Hundiwale, D.G.; Mahulikar, P.P. Controlled Release Study of Phenol Formaldehyde Microcapsules Containing Neem Oil as an Insecticide. Int. J. Polym. Mater. 2013, 62, 421–425. [Google Scholar] [CrossRef]
  187. Pascoli, M.; Jacques, M.T.; Agarrayua, D.A.; Avila, D.S.; Lima, R.; Fraceto, L.F. Neem Oil Based Nanopesticide as an Environmentally-Friendly Formulation for Applications in Sustainable Agriculture: An Ecotoxicological Perspective. Sci. Total Environ. 2019, 677, 57–67. [Google Scholar] [CrossRef] [Green Version]
  188. Luiz De Oliveira, J.; Ramos Campos, E.V.; Fraceto, L.F. Recent Developments and Challenges for Nanoscale Formulation of Botanical Pesticides for Use in Sustainable Agriculture. J. Agric. Food Chem. 2018, 66, 8898–8913. [Google Scholar] [CrossRef] [PubMed]
  189. De Souza, M.T.; Porsani, M.V.; Bach, R.P.; De Souza, M.T. Encapsulamento de Moléculas Como Oportunidade Emergente Na Agricultura. Pesqui. Agropecuária Pernambucana 2021, 26, 1–5. [Google Scholar]
  190. Maruyama, C.R.; Bilesky-José, N.; de Lima, R.; Fraceto, L.F. Encapsulation of Trichoderma Harzianum Preserves Enzymatic Activity and Enhances the Potential for Biological Control. Front. Bioeng. Biotechnol. 2020, 8, 225. [Google Scholar] [CrossRef] [PubMed]
  191. Peil, S.; Beckers, S.J.; Fischer, J.; Wurm, F.R. Biodegradable, Lignin-Based Encapsulation Enables Delivery of Trichoderma Reesei with Programmed Enzymatic Release against Grapevine Trunk Diseases. Mater. Today Bio 2020, 7, 100061. [Google Scholar] [CrossRef] [PubMed]
  192. Locatelli, G.O.; dos Santos, G.F.; Botelho, P.S.; Finkler, C.L.L.; Bueno, L.A. Development of Trichoderma Sp. Formulations in Encapsulated Granules (CG) and Evaluation of Conidia Shelf-Life. Biol. Control 2018, 117, 21–29. [Google Scholar] [CrossRef]
  193. Yaakov, N.; Ananth Mani, K.; Felfbaum, R.; Lahat, M.; Da Costa, N.; Belausov, E.; Ment, D.; Mechrez, G. Single Cell Encapsulation via Pickering Emulsion for Biopesticide Applications. ACS Omega 2018, 3, 14294–14301. [Google Scholar] [CrossRef] [Green Version]
  194. Yaakov, N.; Kottakota, C.; Mani, K.A.; Naftali, S.M.; Zelinger, E.; Davidovitz, M.; Ment, D.; Mechrez, G. Encapsulation of Bacillus Thuringiensis in an Inverse Pickering Emulsion for Pest Control Applications. Colloids Surf. B Biointerfaces 2022, 213, 112427. [Google Scholar] [CrossRef]
  195. Pour, M.M.; Saberi-Riseh, R.; Mohammadinejad, R.; Hosseini, A. Investigating the Formulation of Alginate- Gelatin Encapsulated Pseudomonas Fluorescens (VUPF5 and T17-4 Strains) for Controlling Fusarium Solani on Potato. Int. J. Biol. Macromol. 2019, 133, 603–613. [Google Scholar] [CrossRef]
  196. Amar Feldbaum, R.; Yaakov, N.; Ananth Mani, K.; Yossef, E.; Metbeev, S.; Zelinger, E.; Belausov, E.; Koltai, H.; Ment, D.; Mechrez, G. Single Cell Encapsulation in a Pickering Emulsion Stabilized by TiO2 Nanoparticles Provides Protection against UV Radiation for a Biopesticide. Colloids Surf. B Biointerfaces 2021, 206, 111958. [Google Scholar] [CrossRef]
  197. De Araújo, J.S.F.; de Souza, E.L.; Oliveira, J.R.; Gomes, A.C.A.; Kotzebue, L.R.V.; da Silva Agostini, D.L.; de Oliveira, D.L.V.; Mazzetto, S.E.; da Silva, A.L.; Cavalcanti, M.T. Microencapsulation of Sweet Orange Essential Oil (Citrus aurantium Var. Dulcis) by Liophylization Using Maltodextrin and Maltodextrin/Gelatin Mixtures: Preparation, Characterization, Antimicrobial and Antioxidant Activities. Int. J. Biol. Macromol. 2020, 143, 991–999. [Google Scholar] [CrossRef]
  198. Nyari, N.; Paulazzi, A.; Zamadei, R.; Steffens, C.; Zabot, G.L.; Tres, M.V.; Zeni, J.; Venquiaruto, L.; Dallago, R.M. Synthesis of Isoamyl Acetate by Ultrasonic System Using Candida antarctica Lipase B Immobilized in Polyurethane. J. Food Process Eng. 2018, 41, e12812. [Google Scholar] [CrossRef]
  199. Timilsena, Y.P.; Haque, M.A.; Adhikari, B. Encapsulation in the Food Industry: A Brief Historical Overview to Recent Developments. Food Nutr. Sci. 2020, 11, 481–508. [Google Scholar] [CrossRef]
  200. Pavoni, L.; Benelli, G.; Maggi, F.; Bonacucina, G. Green Nanoemulsion Interventions for Biopesticide Formulations; Elsevier: Amsterdam, The Netherlands, 2019; ISBN 9780128158296. [Google Scholar]
  201. Maes, C.; Bouquillon, S.; Fauconnier, M.L. Encapsulation of Essential Oils for the Development of Biosourced Pesticides with Controlled Release: A Review. Molecules 2019, 24, 2539. [Google Scholar] [CrossRef] [PubMed]
  202. Taban, A.; Saharkhiz, M.J.; Naderi, R. A Natural Post-Emergence Herbicide Based on Essential Oil Encapsulation by Cross-Linked Biopolymers: Characterization and Herbicidal Activity. Environ. Sci. Pollut. Res. 2020, 27, 45844–45858. [Google Scholar] [CrossRef]
  203. Karaaslan, M.; Şengün, F.; Cansu, Ü.; Başyiğit, B.; Sağlam, H.; Karaaslan, A. Gum Arabic/Maltodextrin Microencapsulation Confers Peroxidation Stability and Antimicrobial Ability to Pepper Seed Oil. Food Chem. 2021, 337. [Google Scholar] [CrossRef]
  204. Arango-Ruiz, Á.; Martin, Á.; Cosero, M.J.; Jiménez, C.; Londoño, J. Encapsulation of Curcumin Using Supercritical Antisolvent (SAS) Technology to Improve Its Stability and Solubility in Water. Food Chem. 2018, 258, 156–163. [Google Scholar] [CrossRef]
  205. Rosa, M.T.M.G.; Alvarez, V.H.; Albarelli, J.Q.; Santos, D.T.; Meireles, M.A.A.; Saldaña, M.D.A. Supercritical Anti-Solvent Process as an Alternative Technology for Vitamin Complex Encapsulation Using Zein as Wall Material: Technical-Economic Evaluation. J. Supercrit. Fluids 2020, 159, 104499. [Google Scholar] [CrossRef]
  206. Sakata, G.S.B.; Ribas, M.M.; Dal Magro, C.; Santos, A.E.; Aguiar, G.P.S.; Oliveira, J.V.; Lanza, M. Encapsulation of Trans-Resveratrol in Poly(ε-Caprolactone) by GAS Antisolvent. J. Supercrit. Fluids 2021, 171, 105164. [Google Scholar] [CrossRef]
  207. Zou, Y.; Wang, F.; Li, A.; Wang, J.; Wang, D.; Chen, J. Synthesis of Curcumin-loaded Shellac Nanoparticles via Co-precipitation in a Rotating Packed Bed for Food Engineering. J. Appl. Polym. Sci. 2022, 139, e52421. [Google Scholar] [CrossRef]
Figure 1. Encapsulation and morphology of microcapsule.
Figure 1. Encapsulation and morphology of microcapsule.
Polymers 14 04194 g001
Figure 2. Schematic illustration of the emulsions.
Figure 2. Schematic illustration of the emulsions.
Polymers 14 04194 g002
Figure 3. Advantages and applications of encapsulation in the agricultural sector.
Figure 3. Advantages and applications of encapsulation in the agricultural sector.
Polymers 14 04194 g003
Figure 4. Advantages and applications of encapsulation in the food area.
Figure 4. Advantages and applications of encapsulation in the food area.
Polymers 14 04194 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zabot, G.L.; Schaefer Rodrigues, F.; Polano Ody, L.; Vinícius Tres, M.; Herrera, E.; Palacin, H.; Córdova-Ramos, J.S.; Best, I.; Olivera-Montenegro, L. Encapsulation of Bioactive Compounds for Food and Agricultural Applications. Polymers 2022, 14, 4194. https://doi.org/10.3390/polym14194194

AMA Style

Zabot GL, Schaefer Rodrigues F, Polano Ody L, Vinícius Tres M, Herrera E, Palacin H, Córdova-Ramos JS, Best I, Olivera-Montenegro L. Encapsulation of Bioactive Compounds for Food and Agricultural Applications. Polymers. 2022; 14(19):4194. https://doi.org/10.3390/polym14194194

Chicago/Turabian Style

Zabot, Giovani Leone, Fabiele Schaefer Rodrigues, Lissara Polano Ody, Marcus Vinícius Tres, Esteban Herrera, Heidy Palacin, Javier S. Córdova-Ramos, Ivan Best, and Luis Olivera-Montenegro. 2022. "Encapsulation of Bioactive Compounds for Food and Agricultural Applications" Polymers 14, no. 19: 4194. https://doi.org/10.3390/polym14194194

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