Next Article in Journal
Experimental Investigation of Electro-Mechanical Behavior of Silver-Coated Teflon Fabric-Reinforced Nafion Ionic Polymer Metal Composite with Carbon Nanotubes and Graphene Nanoparticles
Previous Article in Journal
The Influence of Organofunctional Substituents of Spherosilicates on the Functional Properties of PLA/TiO2 Composites Used in 3D Printing (FDM/FFF)
Previous Article in Special Issue
Antimicrobial Bilayer Film Based on Chitosan/Electrospun Zein Fiber Loaded with Jaboticaba Peel Extract for Food Packaging Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biotechnological Applications of Nanoencapsulated Essential Oils: A Review

by
Patrícia Melchionna Albuquerque
1,*,
Sidney Gomes Azevedo
2,
Cleudiane Pereira de Andrade
1,
Natália Corrêa de Souza D’Ambros
1,
Maria Tereza Martins Pérez
2 and
Lizandro Manzato
2
1
Research Group on Chemistry Applied to Technology (QAT), School of Technology, Amazonas State University, Manaus 69050-020, Brazil
2
Laboratory of Synthesis and Characterization of Nanomaterials (LSCN), Federal Institute of Education, Science and Technology of Amazonas, Manaus 69075-351, Brazil
*
Author to whom correspondence should be addressed.
Polymers 2022, 14(24), 5495; https://doi.org/10.3390/polym14245495
Submission received: 15 November 2022 / Revised: 7 December 2022 / Accepted: 12 December 2022 / Published: 15 December 2022
(This article belongs to the Special Issue Biopolymer Matrices for Incorporation of Bioactive Compounds)

Abstract

:
Essential oils (EOs) are complex mixtures of volatile and semi-volatile organic compounds that originate from different plant tissues, including flowers, buds, leaves and bark. According to their chemical composition, EOs have a characteristic aroma and present a wide spectrum of applications, namely in the food, agricultural, environmental, cosmetic and pharmaceutical sectors. These applications are mainly due to their biological properties. However, EOs are unstable and easily degradable if not protected from external factors such as oxidation, heat and light. Therefore, there is growing interest in the encapsulation of EOs, since polymeric nanocarriers serve as a barrier between the oil and the environment. In this context, nanoencapsulation seems to be an interesting approach as it not only prevents the exposure and degradation of EOs and their bioactive constituents by creating a physical barrier, but it also facilitates their controlled release, thus resulting in greater bioavailability and efficiency. In this review, we focused on selecting recent articles whose objective concerned the nanoencapsulation of essential oils from different plant species and highlighted their chemical constituents and their potential biotechnological applications. We also present the fundamentals of the most commonly used encapsulation methods, and the biopolymer carriers that are suitable for encapsulating EOs.

Graphical Abstract

1. Introduction

The investigation of aromatic and medicinal plants is constantly expanding due to the current demand for natural products [1], whose benefits are directly related to the compounds produced by these plants through their secondary metabolism [2]. Essential oils (EOs) are plant secondary metabolites, and are also known as volatile oils, ethereal oils or essences [3]. They are defined as hydrophobic fluids that contain substances or compounds with “volatile aroma”, and which are extracted from plant parts, including flowers, buds, leaves and bark, and have a characteristic aroma [4]. EOs are fundamental for the survival mechanisms of plants, and play an important role in protecting them against bacteria, viruses, fungi, and herbivores, and can also attract pollinating insects and seed dispersers [5].
EOs are complex mixtures of volatile substances, including terpenes, terpenoids and phenols. This composition allows a wide spectrum of applications such as in food, agricultural/environmental, cosmetic and pharmaceutical industries [6], mainly due to their antioxidant, anxiolytic, antidepressant, anti-inflammatory, antimicrobial antibacterial, antiviral, antifungal, anti-aflatoxigenic, anticancer, antihyperglycemic and other properties [7,8,9,10]. In addition, they do not present health risks when associated with the use of synthetic pesticides [11].
However, efforts to develop new technologies, applications or functional agents based on EOs encounter challenges related to the loss of biological constituents due to high volatility, high risk of deterioration/oxidation, dependence on seasonality, limited stability and/or reduced efficacy of their biotechnological properties [12]. Nanoencapsulation, therefore, seems to be an interesting approach for minimizing these limitations, as it not only prevents the exposure and degradation of EOs and their bioactive constituents, thus creating a physical barrier, but it also facilitates their controlled release, which results in greater bioavailability and efficacy [13]. Several methods of nanoencapsulation have been developed, but those using nanoparticles or the nanoemulsions (a nanostructure system in which the encapsulation takes place) seem to be the most suitable and promising, and are therefore the most commonly explored in the current context [14].
Encapsulation processes require bioactive components to be coated within a matrix (i.e., a synthetic or natural polymer), which isolates the active ingredients from the external environment [15]. In this review, we focus on selecting studies from the last five years whose objective was to nanoencapsulate essential oils of various plant species; then, we list the main chemical constituents found in each EO and the encapsulation methods and biopolymer carriers used in the process. In addition, we highlight the biotechnological potential of nanoencapsulated EOs based on their bioactive properties and the applications developed in the selected studies.

2. Nanoencapsulation

Nanotechnology represents a revolutionary path for technological development in regards to the management of materials on a nanometric scale (one billion times smaller than one meter) [16]. The use of nanotechnology in natural products has gained prominence in many studies since it has resulted in effective alternative products for a number of purposes. Natural assets such as essential oils are metabolites that are of interest to various industries, but they have peculiar characteristics such as the volatility of their chemical constituents, which is why they appear as active ingredients of various encapsulated systems. Once their bioactive properties have been proven, it is necessary to use technological tools to circumvent the problem of high volatility and, at the same time, make the most of their bioactive power.
The technique of encapsulation of essential oils has been widely applied as a nanotechnological tool to protect these assets from the external environment, as well as modulate their release according to specific needs [17,18,19]. Encapsulation of bioactive compounds represents a viable and efficient alternative. In addition, this technique can increase the physical stability of substances, protect them from interactions with the environment, decrease their volatility, increase their bioactivity, reduce toxicity and even allow the release of assets over time in specific media [20,21].
The controlled-release mechanism consists of displacing the assets present in the nanoparticles (NPs) to the application medium. This displacement is gradual and is related to the concentration that is released over time, bringing benefits such as reduced evaporation of volatile assets, easy handling, reduction in phytotoxicity and environmental pollutants, which results in advantages for both the ecosystem and human health [22,23].
The investigation of the bioactive release profile in polymeric NPs provides important information about the mechanisms that guide this action. There are several possible mechanisms of bioactive release: release due to erosion or degradation of polymers; self-diffusion through pores; release through the erosion of the surface of the polymer and pulsed delivery initiated by the application of a magnetic field [24,25]. Studies of release mechanisms in nanoparticles conducted by Yasmin et al. [25] showed that the asset can be diffused through the polymer wall, and release occurs by diffusion or erosion of the matrix. Moreover, the asset can be released through the slow degradation of the polymer wall, or by the cleavage of the matrix through the action of enzymes. Controlled release can be schematized illustratively, as shown in Figure 1.
Several types of materials have been used as carriers of these active ingredients, principally natural and synthetic polymers [18,20,21,26]. Carriers, also known as wall materials, are the materials responsible for protecting the asset. During the development of encapsulating nano- or microparticles, they act as a protective membrane of the asset. In the case of colloidal systems, the carriers protect the asset from the aqueous medium. However, it is necessary to have a profound knowledge of the characteristics of the carriers in terms of biodegradability, capacity for surface functionalization, conjugation, complexation, encapsulation capacity and chemical affinity with the active substance [16,27].
Nanoencapsulated systems cover applications such as those that are biodefensive (being directly used in the control of pests and disease vectors) [20,28], in medicine (especially in the selective delivery of drugs) [29,30,31,32,33], in cosmetics (in the protection of substances prone to oxidation and in the delivery of active substances in deeper layers of the skin) [34], in food technology (protecting highly antioxidant substances and vitamins) [35,36], and in the most diverse applications in which it is necessary to guide the active ingredient to the site of action, as well as control its release rate. This type of technology allows one to maintain the characteristics and properties of active compounds, such as their protective capabilities, stabilization and prolonged release.
The encapsulation of natural assets, such as essential oils, in general, has been developed by a number of researchers in order to improve chemical stability, increase the activity of these substances and reduce volatilization, thus improving their biological potential. In this context, we can cite some examples of encapsulation of essential oils and their most recent applications.
Xavier et al. [26] encapsulated the essential oil from Cinnamodendron dinisii using zein as a carrier, applied in a chitosan matrix to produce an active nanocomposite film packaging for food preservation. The chitosan films obtained and functionalized with the nanoparticles demonstrated antioxidant and antimicrobial activity and were efficient in preserving ground beef. Campelo et al. [27] developed a new pharmaceutical form of semi-solid dosage based on essential oil from cloves and polysaccharides for the treatment of vaginal candidiasis. The nanoemulsions showed excellent colloidal stability and adequate pH for this specific application.
Corrado et al. [37] encapsulated essential oil from oregano in nanoparticles based on polyhydroxybutyrate and poly-3-hydroxybutyrate-co-hydroxyhexanoate using a solvent evaporation technique and achieved an encapsulation efficiency of greater than 60%. The nanoparticles obtained showed a regular distribution with a size range of 150–210 nm. When comparing the effectiveness of pure EO with encapsulated EO, it was possible to observe that encapsulated EO has greater bioactivity against microorganisms such as Micrococcus luteus. The authors emphasize the importance of nanoencapsulation of volatile bioactive compounds in biodegradable polymer matrices and conclude that the results pave the way for the effective exploitation of nanosystems developed for active packaging.
Azevedo et al. [28] developed an environmentally-friendly nanoparticle system for the encapsulation of the essential oil from Piper nigrum using gelatin and poly(ε-caprolactone) (PCL) as carriers. By evaluating the encapsulation efficiency, electrical conductivity, turbidity, pH and organoleptic properties (color and odor) after the addition of different preservatives, the authors studied the stability of the formulation generated. In this research, a particle size of between 114 ± 3 nm and 519 ± 13 nm and encapsulation efficiency of 98 ± 2% were obtained. Due to the pronounced bioactivity of the encapsulated essential oil, they concluded that the developed system has potential as a stable alternative product and as a controlling agent.

2.1. Nanoparticles

Nanoparticles can be formed from composites and are developed through the association of two materials that together give rise to a third material with properties superior to the separate forming components [16]. They can be called carriers and consist of multilayers that are superimposed according to the desired application under the study. Its formation depends on a number of factors, the main one being the chemical and physical interaction between the layers and the encapsulated material so that the stability of the system is achieved [31,38].
Nanoparticles, when they have a structured inner layer, are called nanocapsules and, when they have a continuous matrix, they are called nanospheres [29] (Figure 2). Nanocapsules are particles consisting of a polymeric wall containing a cavity inside, by which the active ingredient is adsorbed, though this may also be present in the polymeric wall; while nanospheres are formed by a polymeric matrix in which chemical components can be retained or adsorbed [30].
The development of polymer nanosystems is rapidly expanding and plays a key role in various areas, from pollution control to environmental technology, from electronics to photonics, from medicine to biotechnology, from materials to sensors, and so on. Several reports in the literature emphasize the growing interest in this area. This trend is based on the unique properties of polymer nanosystems, which meet numerous applications and market needs [31,32].
The combination of biopolymer materials represents an important alternative for encapsulating essential oils. However, understanding the nanoparticle development project is crucial for obtaining stable formulations, adequate controlled-release and surface functionalization for further conjugation with bioactive molecules or ligands. The development of these encapsulating nanoparticles involves a series of parameters that are related to the specific purpose of action of the encapsulated substances and also to the controlled release mechanisms, which can often be related to the type of carrier used [33,34,35,36]. Its formation depends on certain factors, the main one being the chemical and/or physical interaction between the layers and the encapsulated active substance.
Encapsulated systems are efficient strategies for transporting the active substance to its site of action through the choice of a carrier and the appropriate route, with the main objectives being protecting its content from environmental factors (light, moisture, oxygen and interactions with other compounds), in addition to controlled release and release under stimuli (such as changes in pH, physical disruption, swelling, dissolution, etc.). In addition, encapsulation can also mask the unpleasant taste and/or odor and increase the acting time of the active compound, thus prolonging its effect. In this case, the type of nanoparticle and the place where the active substance will be exposed (adsorbed on the surface or not) will depend on the desired final characteristics, such as application, size, size distribution, degree of biodegradability and compatibility of the polymer with the active substance [18].

2.2. Development of Nanoparticles

Several methodologies exist for the development of encapsulation of polymeric nanoparticles, which in their formulations generally employ materials such as polymers (synthetic or natural), surfactants, bioactive compounds, organic solvents and essential oils, depending on the formulation. For the preparation of nanoparticles with the purpose of carrying natural active ingredients, the physicochemical properties of the polymer must be taken into account. Polymers and their degradation products must be biocompatible and biodegradable, and cause no harm or impact to the environment [38,39].
According to Mora-Huerta et al. [33], the following classic methods of obtaining nanoparticles are generally used: nanoprecipitation, emulsion-diffusion, double emulsification, emulsion coacervation, polymer coating and layer-by-layer.
The nanoprecipitation method (Figure 3) is also called solvent displacement or interfacial deposition. This method requires two phases, one organic and one aqueous. The organic phase essentially consists of a solution or a mixture of solvents (ethanol, acetone, hexane, dichloromethane, etc.), of a macromolecule with a carrier role (synthetic, semisynthetic or natural polymer), the active substance, oil and a lipophilic surfactant. However, the aqueous phase consists of a mixture of surfactant in an aqueous medium. In this method, the nanoparticles are obtained as a colloidal suspension that is formed when the organic phase is slowly added to the aqueous phase with moderate agitation. The main variables of the procedure are those associated with the conditions of addition of the organic phase to the aqueous phase, such as the organic phase injection rate and the aqueous phase agitation rate [33,34,40,41].
The preparation of nanoparticles using the emulsion–diffusion method (Figure 4) allows the nanoencapsulation of lipophilic and hydrophilic active substances.
The procedure requires three phases: organic, aqueous and dilution. When the goal is the nanoencapsulation of a lipophilic active substance, the organic phase contains the polymer, the active substance, the oil and an organic solvent that is partially miscible with water. The organic phase is emulsified under vigorous stirring in the aqueous phase and, after primary emulsion formation, the organic solvent is diffused to the external aqueous phase by adding excess water (dilution), which leads to polymer precipitation and nanoparticle formation [33,39].
The nanocapsule formation mechanism suggested by Quintanar-Guerrero et al. [42] is based on the theory that each emulsion droplet produces several nanocapsules, and that these are formed by the combination of polymer precipitation and interfacial phenomena during solvent diffusion [43].
The double emulsification method consists of the formation of two emulsions that are usually prepared using two surfactants: a hydrophobic one intended to stabilize the water/oil interface of the internal emulsion, and a hydrophilic one to stabilize the external interface of the oil droplets for water/oil/water emulsions. The primary emulsion is formed with the use of ultrasound, and the hydrophobic surfactant stabilizes the water/oil interface of the internal phase. The secondary emulsion can also be formed with ultrasound, and the dispersion of the nanoparticles is stabilized by the addition of another surfactant (hydrophilic) [33,39,40].
The emulsion coacervation method involves the formation of an oil/water emulsion, in which the organic phase is composed of the solvent and the bioactive compound, and the aqueous phase is composed of the polymer, a stabilizing agent and water. The emulsion can be formed by the use of ultrasound or mechanical stirring. Then, the coacervation process is carried out by adding electrolytes, adding a water-immiscible solvent or dehydrating agent, or changing the temperature. Finally, the coacervation process is supplemented with additional measures for the formation of lattices, which makes it possible to obtain the nanoparticles. The formation of nanoparticles occurs during the coacervation phase, in which there is precipitation of the polymer from the continuous emulsion phase to form a film that agglomerates into nanoparticles [31,33,39,40].
The polymer-coating method is used for the deposition of a thin polymer layer on the surface of the nanoparticle that was previously formed by the adsorption of the polymer on uncoated nanoparticles when incubated with a polymer solution under stirring. Likewise, this polymeric layer can be added during the final phase of the methods mentioned above [33,39].
The layer-by-layer method favors the acquisition of vesicular particles, which are also called polyelectrolyte capsules. The mechanism of formation is based on irreversible electrostatic attraction that leads to the adsorption of polyelectrolytes in the formed layers. A polymer layer is adsorbed by incubation in the polymer solution, which decreases the solubility of the polymer by dropwise addition of solvent. This procedure is then repeated with a second polymer and several polymer layers are deposited sequentially [33].

3. Chemical Composition of Essential Oils

Generally speaking, constituents of EOs mainly comprise terpenes, phenylpropanoids, straight-chain compounds and diverse groups. Among these, terpenes are the most abundant compounds and comprise hydrocarbons of the class of mono, sesqui and diterpenes; and oxygenated compounds, such as alcohols, oxides, aldehydes, ketones, phenols, acids, esters and lactones [44,45].
The chemical composition of EOs varies between the different plant species that produce them. Among the studies selected for this review, 20 different botanical families were explored, with emphasis on Lamiaceae, Myrtaceae, Lauraceae, Apiaceae and Rutaceae, in which the species Thymus vulgaris and Eugenia caryophyllata are the most cited. Other plants, such as Origanum sp. and Cinnamomum sp., are also extensively used for EO extraction, and were found within the articles selected for this review. Some of the studies address EOs that present main components that correspond to more than 50% of their chemical composition, such as the EO of Aniba canelilla (1-nitro-2-phenylethane = 86.63%) [46], A. rosaeodora (linalool = 81.46%) [47], Cymbopogon citratus (citral = 67.4%) [48], Pimpinella anisum (anethole = 51.02%) [49], Mentha pulegium (pulegone = 72,18%) [50], Syzygium aromaticum (eugenol = 71.92%) [51], C. nardus (citral = 62.73%) [52], Illicium verum (anethole = 89.12%) [53], T. capitatus (carvacrol = 76.1%) [54], Kaempferia galanga (ethyl-p-metoxycinnamate = 59.4%) [55], Cinnamomum tamala (linalool = 82.64%) [56], Foeniculum vulgare (anethole = 73.27%) [57], and Coriandrum sativum (linalool = 65.18%) [58].
In addition, it is interesting to mention the case of the EO of E. caryophyllata, which is reported in the works of De Hasheminejad et al. [59]; Hadid et al. [60]; and Kujur et al. [61], and for which the main component is eugenol, with 77.2%, 89.86% and 73.6% of eugenol being found in the EO, respectively. In other cases, in the same botanical species, there may be different main components when the EO is characterized by different authors, and data indicate that the chemical composition may vary according to the season, growing conditions and the part of the plant used in the process of obtaining the EO [62].
The methods of extracting essential oils vary according to the state in which the plant is found [63,64]; thus, for each purpose of the oil, a different technique can be chosen. Among the techniques for the extraction of essential oils, hydrodistillation, pressing, solvent extraction, enfloration, supercritical gases and microwaves can be used [65,66]. In hydrodistillation, the constituents of the plant material are dragged by water vapor because they have a higher vapor pressure than water. This method is most often used for extracting EOs from fresh plants. Pressing, on the other hand, is a technique used to extract EOs from citrus fruits. The pericarps are pressed and the layer containing the EO is separated. Extraction with non-polar solvents, in turn, generates a product of low commercial value, since other lipophilic compounds are extracted along with the EO. In enfloration, on the other hand, the product has high commercial value, as it is used to obtain the EO from petals. It is carried out with the help of a fat, at room temperature, for a short period. Extraction with supercritical gases allows one to recover natural aromas of various types, and not only the essential oil. It is a very efficient method and is ideal for industrial extraction of EOs. Microwave-assisted extraction combines microwaving with traditional solvent extraction. Selective heating during extraction increases process kinetics and yield [64].
The chemical compounds found in EOs give them their biotechnological properties, which can be applied in different commercial areas. In the food industry, EOs are employed as alternative functional ingredients to extend the shelf life of food products, thus ensuring microbial safety by preventing the development of pathogens such as Salmonella spp. and Lysteria spp. [67,68]. They also act as antioxidants and preservatives in food, and can be incorporated into packaging [69,70], in addition to representing a potential natural alternative to the use of chemical preservatives [71].
The chemical composition of EOs also confers the application of EOs in the pharmaceutical industry, such as the EO of Melissa officinalis, whose in vitro cytotoxicity assay indicated that this oil can be effective against a number of human cancer cell lines (A549, MCF-7, Caco-2, HL-60, K562) [72].
Since EOs are rich in volatile bioactive substances, the nanocapsulation technique has been reported as an important ally in the biotechnological application of these metabolites, thus increasing their efficiency and preventing their degradation in the short term.

4. Biotechnological Potential of Nanoencapsulated Essential Oils

Encapsulation of EOs can be developed at micro or nano levels and presents several possibilities for biotechnological applications. However, nanoencapsulation technology has been growing exponentially and is now being used in a variety of industrial applications, such as textiles, the food industry, cell immobilization, fermentation processes, drug delivery, cell transplantation, agriculture, and cosmetics, among others [73]. In this review, we will focus on nanoencapsulated EOs with promising applications in the food, cosmetics, pharmaceutical and environmental industries. Table 1 summarizes different nanoencapsulated essential oils and highlights their chemical composition, nanoencapsulation method, the polymeric material used and their biotechnological applications.

4.1. Pharmaceutical Applications

The use of EOs in traditional systems of medicine has been practiced since ancient times in human history, as they exhibit different biological properties; but only recent advances and technologies have allowed the stabilization, prolonged release, targeted delivery and maintenance of these bioactive components. These advantages are conferred to nanoencapsulated EOs. In this context, the development of formulations that maintain the biological and physicochemical properties of EOs is an important choice when used as an active ingredient in pharmaceutical formulations.
Nanoencapsulated EOs can be used as a healing accelerator for infected wounds and dressing of diabetic ulcers. These actions are described in the work of Kreutz et al. [46] who developed and characterized a nanoemulsion using the EOs of leaves and branches of A. canelilla—an aromatic plant from the Amazon. The authors observed that the nanosystem developed is promising for the treatment of topical inflammation. This is related to its predominant chemical compound (1-nitro-2-phenylethane = 86.63%), which has already been reported as an anti-inflammatory and antinociceptive substance. The authors emphasized that nanoemulsions are the most-reported nanostructure system to encapsulate essential oils, since they allow one to incorporate high doses of these active products.
Ghodrati, Farahpour & Hamishehkar [74] also developed nanoemulsions from Mentha spp. and obtained bioproducts in the form of nanogels with promising antibacterial activity against gram-negative and gram-positive bacteria. The nanoemulsions showed adequate encapsulation efficiency and size distribution. It is interesting to note that the formulations accelerated the healing process of an infected wound model and this may be an appropriate strategy for producing topical healing formulations.
The healing of infected wounds was also reported in a study with the nanoencapsulated EOs of pennyroyal (M. pulegium) and thyme (T. vulgaris), which showed antimicrobial and antifungal potential, respectively. Nanoencapsulated pennyroyal EO decreases the duration of the inflammatory phase and thyme EO was strongly recommended for the treatment of cutaneous mycoses [50,75]. It is also interesting to mention the work of Rozman et al. [76] with Homalomena pineodora EO nanoparticles, synthesized by ion gelification, and whose pharmaceutical properties revealed a broad spectrum of activity against clinical microbial strains that infect diabetic skin lesions (Escherichia coli, Proteus mirabilis, Yersinia sp., Klebsiella pneumoniae, Shigella boydii, Salmonella typhimurium, Acinetobacter anitratus, Pseudomonas aeruginosa, Candida albicans and C. utilis). The bioactive behavior of this nanocapsule may be due to the synergistic effect of the EO with the polymeric carrier (chitosan).
Nanoencapsulated EOs have important future prospects for the treatment of various types of cancer. Recent studies have developed nanocapsules using different methods (ionic gelidification–emulsion, nanoprecipitation and high-speed homogenization) with EOs from Cynometra cauliflora [77], Morinda citrifolia [78], Citrus spp. [79]. and Origanum glandulosum [80], respectively, all encapsulated with chitosan, with the exception of the latter for which sodium alginate was used. These nanoencapsulated EOs showed anticancer action against human lung tumor cells A549, breast cancer MDA-MB-468, melanoma A-375, human hepatocellular carcinoma (HepG2) and human breast cancer cells MCF-7 and MDA-MB-231.
The use of nanoencapsulated EOs with antimicrobial activity has also been widely reported. In the pharmaceutical context, microbial resistance is a serious public health problem, and these EOs present remarkable potential against different pathogens. The following nanoencapsulated EOs exhibit antimicrobial activity: EO of O. vulgare and T. capitatus nanocarried with chitosan [81], EO of Cinnamomum spp. nanocarried with sodium alginate [82], EO of C. aurantifolia, C. hystrix and Citrofortunella microcarpa [83], EO of Poiretia latifolia [84], EO of O. vulgare and T. capitatus nanocarried with poly(ε-caprolactone) [71], EO of C. zeylanicum, T. vulgaris and Schinus molle, nanocarried with chitosan [85], EO of E. caryophyllata [86] and EO of C. commutatus, also nanocarried with chitosan [87].

4.2. Cosmetic Applications

Among the problems faced by the cosmetic industry, microbial contamination stands out as one of the most important, since it negatively affects formulations and is difficult to control [15]. To circumvent such situations, cosmetic preservatives are used, which are chemical substances of the most varied classes and which prevent the proliferation of microorganisms in the formulas, thereby increasing the shelf life of these products. However, some of these preservatives may have undesirable effects, such as causing allergies, irritations, and may even have toxic effects [88]. In an attempt to reduce these problems and increase the natural commercial appeal of cosmetics, it is possible to employ nanoencapsulated EOs in cosmetic formulations.
Recent studies have involved the basic research of nanoencapsulated EOs with antimicrobial properties, as well as the exploration of the antioxidant potential of these products for cosmeceutical purposes. This is the case of the study conducted by Hadidi et al. [60], who encapsulated E. caryophyllata EO loaded with chitosan by ionic gelidification. The nanoparticles had a high antibacterial activity (47.8–48 mm inhibition halo) against L. monocytogenes and S. aureus. Huang et al. [89] and Sampaio et al. [90] evaluated the encapsulation efficiency of the EO of Cedrus deodara (loaded with modified starch), T. vulgaris and M. officinalis (without wall material) in the form of nanoemulsions. All the nanoemulsions evaluated were stable. In addition, the antioxidant and antibacterial activity of the EOs were accentuated after nanoencapsulation.
The nanoencapsulation method using nanoprecipitation seems useful for formulations that aim to maintain the antioxidant and antimicrobial activity of encapsulated EOs, with a view to use in the cosmetics industry. This method is addressed in the studies of Sheta et al. [91] involving the EO of peppermint (M. piperita) and green tea (Camellia sinensis) encapsulated with chitosan. This methodology improved the thermal stability of the EOs with a controlled release profile for 72 h; in addition, nanoencapsulation also increased the antioxidant activity for both essential oils.
The nanoprecipitation process was also useful for nanoencapsulating Cannabis sativa EO using alfalfa protein as a wall material and proved to be an efficient strategy for improving the stability and functionality of the EO. After nanoencapsulation, the increase in the antioxidant activity of the EO was observed, and a product recommended for applications in the cosmetic and food areas was obtained [92].

4.3. Food Applications

Nanoencapsulated EOs can be used for various food applications, mainly in foods that are sensitive to oxidation or degradation under certain conditions, and which decreases the final quality of the product. In this sense, some authors were able to verify the permanence of the antioxidant activity when the nanoencapsulated EOs were used, although the methods of obtaining the encapsulated product or its polymeric matrix were distinct. Karimirad et al. [93] tested nanocapsules of Citrus aurantium EO obtained by nanoprecipitation; Chaudhari et al. [94] obtained nanocapsules from Melaleuca cajuputi EO obtained by ionic gelidification; while Arabpoor et al. [95] worked with nanocapsules loaded with EO from Eryngium campestre, also obtained by ionic gelidification. The polymeric matrix used by these authors was the same, i.e., chitosan, which is a biodegradable biopolymer, but they obtained the encapsulated EOs using different methods.
For applications in the food industry, the presence of antioxidant, antifungal and anti-aflatoxigenic activity in the EO is highly desired, since it can directly influence food preservation [53,96,97]. Deepika et al. [98] used the EO of Petroselinum crispum leaves, which were nanoencapsulated using ionic gelidification with the biopolymer chitosan, and obtained promising nanoemulsions with antioxidant, antifungal and anti-aflatoxigenic activity. The authors recommended the application of this bioproduct on an industrial scale in the management of the loss that occurs during the storage of chia seeds, which is caused by aflatoxin-producing fungi [97]. Similarly, Cai et al. [99] explored the efficacy of chitosan nanocapsules loaded with EO from Ocimum basilicum using the ionic gelidification–emulsion method, and their results demonstrated the strong antibacterial and antibiofilm capacity of the nanoencapsulated EOs against the pathogenic bacteria E. coli and S. aureus.
Nanoencapsulated EOs are also highly recommended for the direct coating of fruits in order to prolong their sensory characteristics. Using ionic gelidification–emulsion, Singh et al. [56] nanoencapsulated the EO of Cinnamomum tamala in a chitosan nanoemulsion and obtained prolongation of the shelf life of stored millet by inhibiting fungi and aflatoxins. Similarly, Antonioli et al. [48] were able to perform the in vivo control of Colletotrichum acutatum in apples by using nanocapsules of the EO from C. citratus carried in poly(lactic acid) (PLA) via nanoprecipitation. The post-harvest apples showed less bitter rot lesions after the use of the nanoencapsulated EO.
Encapsulating EOs can also help control the production of mycotoxins. Wan et al. [100] studied the ability of nanoemulsions of EO of thyme (T. vulgaris), lemongrass (C. citratus), cinnamon (Cinnamomum spp.), peppermint (M. piperita) and cloves (E. caryophyllata) in inhibiting mycotoxins. The authors stated that the chemical composition of the EO directly impacts their inhibitory activity. Other studies report the anti-aflatoxigenic property of nanoencapsulated EOs, either by the formulation of nanoemulsions from EO of dried fruits of C. tamala carried by chitosan [56], or by the formulation of nanogels based on chitosan–cinnamic acid and EO of F. vulgare using the nanoprecipitation method [57].
Protecting against fungi and bacteria directly prolongs the shelf life of food. Authors, such as Sagar et al. [101], have tested nanoemulsions of EOs (cinnamon, cloves and thyme) as coating materials for breaded steamed chicken. The nanoemulsions maintained the quality and sensory attributes, and it was possible to double the storage time of the product from 10 to 20 days. Hossain et al. [102] obtained chitosan-based antifungal films reinforced with cellulose nanocrystals loaded with the EO of O. compactum, T. vulgaris, M. alternifolia and M. piperite that, when combined with radiation, were efficient against the growth of fungi (Aspergillus niger, A. flavus, A. parasiticus and Penicillium chrysogenum) during rice storage, without organoleptic changes.

4.4. Environmental Applications

In the environmental context, it is known that science has been facing important challenges. The use of synthetic agricultural pesticides, for example, used against fungi, insects and other pests, contributes not only to environmental contamination (soil and water courses) but also contamination of the food that is produced. Nanoencapsulated EOs have been described as promising alternatives to the use of these xenobiotics and the chitosan-loaded nanocapsules formulated by the ionic gelidification method seem the most appropriate for this application. Ibrahim et al. [103] evaluated the use of EOs from C. nardus carried by chitosan and cellulose in the control of cotton leafworm (Spodoptera littoralis). The nanosystems have high toxicity and cause the interruption of the development of larvae, thus not only revealing greater insecticidal activity in the mortality of larvae and pupae, but also demonstrating the great bio-insecticidal potential of encapsulated EOs.
Rajkumar et al. [104], using the same encapsulation methods as Ibrahim et al. [103], observed the insecticidal action of the EO of M. piperita against pests in grain (Tribolium castaneum and Sitophilus oryzae). The inhibition caused by the polymer nanoparticles containing the EO was more effective against Tribolium castaneum, which indicates promising potential for the establishment of a pest management program.
Sundararajan et al. [105] used the nanoencapsulation process for the formulation of nanoemulsions with the EO of Ocimum basilicum (basil) leaves, whose result was efficient in terms of antimicrobial, antioxidant and larvicidal activity against third-stage Culex quinquefasciatus larvae. The bioproduct proved to be thermodynamically stable for controlled release and effective in combating mosquitoes. The larvicidal activity of nanoencapsulated EOs was also reported by Ferreira et al. [106], who used nanoemulsions of the EO of Siparuna guianensis loaded with chitosan, and by Santos et al. [53], who used nanocomposites containing the EO of S. aromaticum, bentonite clay and polyvinylpyrrolidone (PVP), which showed effective larvicidal activity against Aedes aegypti when compared to the same EO without encapsulation.
The acaricidal activity of the EO of Satureja hortensis, nanoencapsulated in chitosan nanoparticles, was evaluated against Tetranychus urticae. The authors obtained high encapsulation efficiency (greater than 95%), with effective durability and controlled release, and maintained the acaricidal activity for a long period of time (up to 25 days), thus confirming the possibility of using the nanoencapsulated system as a suitable vehicle for other acaricidal applications [107].
The materials used in nanotechnology formulations can be selected according to important characteristics for environmental applications: biodegradability, capacity for surface functionalization, conjugation and complexation [17]. Biodegradable polymers, whether synthetic or natural, are prone to degradation through natural processes [108,109]. Biopolymers, however, have the advantage of being easily degraded in the environment. Collagen, gelatin, chitosan, gums, and starches, among other biopolymers, have peculiar characteristics, such as low toxicity, cell compatibility and are biodegradable. They therefore emerge as interesting alternatives for use as a wall material in the nanoencapsulation of essential oils [110,111].
Biodegradation can occur in the presence or absence of oxygen, through the action of microorganisms that, in turn, release enzymes capable of breaking down polymer molecules into smaller chains. Finally, these fragments can get into microbial cells where they will be decomposed into CO2, CH4, H2O, mineral salts and biomass [109].
Table 1. Nanoencapsulated essential oils and their biotechnological potential.
Table 1. Nanoencapsulated essential oils and their biotechnological potential.
Essential OilNanoencapsulation Biotechnological Potential
SpeciesCommon NamePlant Part aMain Chemical
Compounds
Encapsulation MethodPolymer
Carrier
Nanoproduct ObtainedBiological
ACTIVITY
ApplicationIndustry
Piper nigrumBlack
pepper
-β-caryophyllene (28%);
limonene (15%); sabinene (11.4%); β-pinene (11%)
Complex
Coacervation
Gelatin and sodium alginateNanocapsules--Food[13]
Aniba rosaeodoraRosewoodLeaveslinalool (81.46%); α-terpineol (7.4%); linalool oxide (1.56%)Ionic
gelidificaton-
emulsion
ChitosanNanoemulsionsAntifungal,
anti-
aflatoxigenic
Fruit coatingFood[47]
Cymbopogon
citratus
LemongrassLeavescitral (67.4%); neral (25.6%); geranial (41.8%); β-myrcene (18.1%)Nano-
Precipitation
poly(lactic acid)-PLANanocapsulesAntifungalFruit coatingFood[48]
Pimpinella anisumAniseedFruitsanethole (51.02%); estragole (24.75%); fenchone (13.22%)Ionic
gelidification-
emulsion
ChitosanNanoemulsionsAntioxidant, antifungal,
anti-
aflatoxigenic
Food
Preservative
Food[49]
Cymbopogon nardusCitronella grass-citral (62.73%);
geranyl acetate (9.53%);
geraniol (4.52%)
Ionic
gelidification-
emulsion
ChitosanNanoemulsionsAntioxidant, antifungal,
anti-
aflatoxigenic
Food
Preservative
Food[52]
Illicium verumStar aniseFruitanethole (89.12%);
estragole (4.85%)
Nano-
Precipitation
ChitosanNanocapsulesAntioxidant, antifungal,
anti-
aflatoxigenic
Food
Preservative
Food[53]
Thymus capitatusConehead thymeAerial partscarvacrol (76.1%); y-terpinene (6.7%); β-caryophyllene (2.7%)Nanoemulsion-NanoemulsionsAntibacterialFood
preservative
Food[54]
Kaempferia
galanga
Sand gingerRhizomesethyl-p-methoxycinnamate (59.4%); trans-methyl cinnamate (17.1%); pentadecane (6.9%)Nanoemulsion-NanoemulsionsAntifungalFood
preservative
Food[55]
Cinnamomum tamalaIndian bay leafFruitslinalool (82.64%);
caryophyllene oxide (3.1%); terpinen-4-ol (2.88%)
Ionic
Gelidification
ChitosanNanoemulsionsAntifungal, anti-
Aflatoxigenic
Food
preservative
Food[56]
Foeniculum vulgareCommon fennelFruitsanethole (73.27%). fenchone (6.84%); D-limonene (4.39%)Nano-
Precipitation
Chitosan-
cinnamic acid
NanogéisAntifungal, anti-
aflatoxigenic
Food
preservative
Food[57]
Coriandrum
sativum
CorianderDried seedslinalool (65.18%);
geranyl acetate (12.06%); α-pinene (4.76%)
Ionic
gelidification-
emulsion
ChitosanNanoemulsionsAntioxidant, antifungal, anti-aflatoxigenicFood
preservative
Food[58]
Eugenia
caryophyllata
ClovesGround aerial parteugenol (77.2%);
eugenyl acetate (8.31%); β-caryophyllene (7.19%)
Ionic
gelidification-
emulsion
ChitosanNanocapsulesAntifungalFood
preservative
Food[59]
Eugenia
caryophyllata
ClovesFlower budseugenol (73.6%);
caryophyllene (9.67%);
oleic acid (2.03%)
Nano-
precipitation
ChitosanNanogelsAntioxidant, antifungal,
anti-
aflatoxigenic
-Food[61]
Citrus aurantiumSeville
orange
Bark Nano-
precipitation
ChitosanNanocapsulesAntioxidantFood
preservative
Food[93]
Melaleuca cajuputiCajuputLeavesα-pinene (49.24%); bornyl acetate (21.07%); camphor (11.70%)Ionic
gelidification
ChitosanNanocapsulesAntioxidantFood
preservative
Food[94]
Eryngium
campestre
Watling Street thistleLeaves and rootsβ-sesquiphellandrene (16.44%); isophytol (12.27%);
stigmasterol (10.11%)
Ionic
gelidification
NanochitosanNanocapsulesAntioxidantFood
preservative
Food[95]
Myristica fragransMaceDried seedsmyristicin (39.43%); methyleugenol (8.15%); safrole (6.26%)Nano-
precipitation
Chitosan-
cinnamic acid
NanogelsAntioxidant, antifungal, anti-
aflatoxigenic
Food
preservative
Food[96]
Petroselinum
crispum
ParsleyLeavescarvacrol (48.45%);
D-limonene (20.80%); cuminaldehyde (15.78%)
Ionic
gelidification
ChitosanNanoemulsionsAntioxidant, antifungal, anti-
aflatoxigenic
Food
preservative
Food[98]
Ocimum basilicumBasil-eugenol (48.32%);
caryophyllene (26.26%);
methyl ester (5.78%)
Ionic
gelidification-
emulsion
ChitosanNanocapsulesAntibacterial, AntibiofilmFood
preservative
Food[99]
Thymus vulgaris Cymbopogon
citratus
Cinnamomum spp. Mentha × piperita Eugenia
caryophyllata
Thyme, lemongrass, Cinnamon, Peppermint, Cloves-Thyme: thymol (21.69%); p-cymene (21.31%); γ-terpinene (13.87%). Lemongrass: β-citral (31.33%); α-citral (14.65%). Cinnamon: eugenol (37.13%); caryophyllene (9.87%). Peppermint: menthol (29.4%); l-menthone (17.97%).
Cloves: eugenol (34.42%%);
eugenol acetate (24.53%%);
caryophyllene (21.30%%).
Nanoemulsion-NanoemulsionsAntifungal, mycotoxin
inhibitor
Food
preservative
Food[100]
Cinnamomum zeylanicum
Thymus vulgaris Syzygium
aromaticum
Cinnamon, Thyme, Cloves--Oil in water emulsionChitosanNanoemulsionsAntioxidant, antimicrobialFood
preservative
Food[101]
Origanum
compactum
Thymus vulgaris Melaleuca
alternifolia
Mentha × piperita
Compact oregano, Thyme,
Tea tree,
Peppermint
-Oregano: carvacrol (46.37%); thymol (13.70%); p-cymene (13.33%). Thyme: thymol (26.04%); p-cymene (26.36%); y-terpinene (16.69%). Tea tree: terpinen-4-ol (38.4%); γ-terpinene (22.6%).
Peppermint: menthol (33.38%); menthone (34.31%)
NanoemulsionChitosanNanocapsulesAntifungalFood storageFood[102]
Zingiber officinaleGinger--NanoemulsionCarnauba wax, hydroxypropylmethylcelluloseNanoemulsions-Food
preservative
Food[112]
Salvia rosmarinusRosemary--Nanoemulsion-Nanoemulsions--Food[113]
Origanum
majorana
Sweet
marjorum
-terpinen-4-ol (28.92%); α-terpineol (16.75%);
linalool (11.07%)
Ionic
gelidification-
emulsion
ChitosanNanocapsulesAntioxidant, antifungal,
anti-
aflatoxigenic
-Food[114]
Myristica fragransNutmegSeedselemicin (27.08%); myristicin (21.29%); thujanol (18.55%)Ionic
gelidification
ChitosanNanoemulsionsAntifungal, anti-
Aflatoxigenic
Food
preservative
Food[115]
Pelargonium
graveolens
Rose-
scented geranium
Aerial partscitronelil (19.1%); menthone (16.7%); linalool (15.1%); isomenthone (12.2%)Oil in water emulsionChitosanNanogelsAntifungal, anti
Aflatoxigenic
-Food[116]
Toddalia asiaticaOrange climber Leavescaryophyllene oxide (24.4%); 1.3-hexadiene, 3-ethyl-2,5-dimethyl (24.08%); 1,4,7-cycloundecatriene,1,5,9,9- tetramethyl-Z,Z,Z (9.46%)Ionic
gelidification
ChitosanNanocapsulesAntifungal, anti-
Aflatoxigenic
-Food[117]
Bunium persicum Seedscuminaldehyde (21.23%);
sabinene (14.66%);
γ-terpinene (12.49%)
NanoemulsionChitosan-
cinnamic acid
NanogelsAntifungal, anti-
aflatoxigenic, cytotoxic
Food
preservative
Food[118]
Myrtus communis Mentha pulegiumCommon myrtle,
Peppermint
Shoots-Nanoemulsion-NanoemulsionsAntimicrobialFood
preservative
Food[119]
Cinnamomum spp.Cinnamon--Nanoemulsion-Nano-
emulsions
--Food[120]
Satureja kermanicaSavoryLeavesthymol (46.54%); carvacrol (30.54%); γ-terpinene (6.58%)Nano-
precipitation
Chitosan-
cinnamic acid
NanogelsAntifungal-Food[121]
Cymbopogon
martinii
PalmarosaLeavesgeraniol (19.06%);
geraniol (14.84%);
geranyl propionate (12.88%)
Nano-
precipitation
ChitosanNanocapsulesAntifungalFood
preservative
Food[122]
Syzygium sp.Cloves--NanoemulsionGelatin, pullulan, inulinNanoemulsionsAntibacterialFood
preservative
Food[123]
Thymus vulgarisThyme-thymol (43.63%); p-cymene (22.86%); bornyl acetate (8.70%)Nanoemulsion-NanoemulsionsAntimicrobialFood
preservative
Food[124]
Origanum vulgare Thymus capitatusOregano ThymeAerial partsthymol (43%); γ-terpinene (15%) and p-cymene (14%)Nano-
precipitation
Poly
(ε-caprolactone)
NanocapsulesAntibacterial-Pharmaceutical, food[71]
Origanum
glandulosum
OreganoAerial partscarvacrol (26.29%);
γ-terpinene (23.43%);
thymol (19.52%)
High-speed homogenization, high-pressure homogenizationSodium alginateNanocapsules NanoemulsionsAntioxidant, anticancer-Pharmaceutical, food[80]
Cinnamomum zeylanicum
Thymus vulgaris Schinus molle
Cinnamon, Thyme, Peruvian peppertreeLeaves-Ionic
gelidification
ChitosanNanocapsulesAntimicrobial-Pharmaceutical, food[85]
Aniba canelillaPreciosaLeaves and branches1-nitro-2-phenylethane (86.63%); methyleugenol (12.7%); benzaldehyde (0.663%)Nanoemulsion-NanoemulsionsAnti-
Chemotactic
Healing of infected woundsPharmaceutical[46]
Eugenia
caryophyllata
ClovesFlower buttonseugenol (89.86%); β-caryophyllene (5.40%) Ionic
gelidification–
emulsion
ChitosanNanoparticlesAntioxidant, antibacterialPreservative, MedicinePharmaceutical, cosmetic[60]
Thymus vulgarisThyme-thymol (22.10%); p-cymene (21.31%); carvacrol (13.02%)High-pressure homogenization NanoemulsionsAntifungalHealing of infected woundsPharmaceutical[75]
Homalomena
pineodora
-Leaves-Ionic
gelidification
ChitosanNanocapsulesAntimicrobialHealing of
diabetic ulcers
Pharmaceutical[76]
Morinda citrifoliaIndian
mulberry
Seedsnordamnacanthal (22.34%); α-copaene (22.96%); α-morenone (20.45%)Nano-
precipitation
ChitosanNanocapsulesAnticancer-Pharmaceutical[78]
Citrus aurantium Citrus limon Citrus sinensisSeville
orange, Lemon,
Sweet
orange
-Seville orange: sabinene (15.6%); ɣ-terpinene (6.0%);
linalool (5.6%). Sweet orange: α-pinene (3.5%); sabinene (17%); trans-limonene oxide (3.1%). Lemon: trans-p-2,8-
menthadien-1-ol (5.0%);
cis-limonene oxide (2.6%); trans-limonene oxide (2.3%)
Ionic
gelidification
ChitosanNanocapsulesAnticancer-Pharmaceutical[79]
Origanum vulgare Thymus capitatosOregano ThymeAerial part-Ionic
gelidification
ChitosanNanocapsulesAntimicrobialMedicinePharmaceutical[81]
Cinnamomum spp.CinnamonBark-Liposomes, lipid nanoparticlesSodium alginateHybrid
composite nanoparticles
AntimicrobialMedicinePharmaceutical[82]
Citrus aurantifolia, Citrus hystrix, Citrofortunella microcarpaLime, Makrut lime Calamondin--Spontaneous emulsification-NanoemulsionsAntibacterialMedicinePharmaceutical[83]
Poiretia latifoliaErva de touroLeavestrans-dihydrocarvone (15.3–51.2%); carvone (12.3–39.0%); limonene (13.9–29.4%)Phase inversionSoy lecithinLipossomes, NanoemulsionsAntifungal, anti-
Inflammatory, antioxidant
-Pharmaceutical[84]
Eugenia
caryophyllata
ClovesAerial parts-High shear homogenization and ultrasound-NanoemulsionsAntimicrobial-Pharmaceutical[86]
Cymbopogon
commutatus
LemongrassWhole plantgeranial (38.6%);
neral (30.3%);
geranyl acetate (8.2%)
Ionic
gelidification-
emulsion
ChitosanNanocapsulesAntimicrobial-Pharmaceutical[87]
Mentha pulegiumPennyroyal-pulegone (72.18%); piperitenone (24.04%);
chrysanthenol (0.90%)
Hot melt homogenizationNanostructured lipid carriers (NLC)NanogelsAntimicrobialHealing of infected woundsPharmaceutical, cosmetic[50]
Mentha × piperitaPeppermintLeavesmenthol (39.80%); menthone (19.55%); neomenthol (8.82%)NanoemulsionNanostructured lipid carriers (NLC) and xanthan gumNanogelsAntimicrobialHealing of infected woundsPharmaceutical, cosmetic[74]
Cynometra
cauliflora
Nam NamLeaves, branches, fruits-Ionic
Gelidification–
emulsion
ChitosanNanocapsulesAntimicrobial, antioxidant, cytotoxic Pharmaceutical, cosmetic[77]
Cedrus deodaraCedarSawdustα-cedarene (32.72%); β-cedarene (12.26%);
thujopsene (24.03%)
NanoemulsionModified starchNanoemulsionsAntioxidant, antibacterialPreservative, MedicinePharmaceutical, cosmetic[89]
Thymus vulgaris Melissa officinalisThyme,
Lemon balm, Black caraway
--Phase inversionSunflower oilNanoemulsionsAntioxidant, antibacterial-Pharmaceutical, cosmetic[90]
Mentha × piperita Camellia sinensisPeppermint, Green tea--Nano-
precipitation
ChitosanNanocapsulesAntimicrobial, antioxidant-Pharmaceutical, cosmetic[91]
Syzygium
aromaticum
ClovesFlower budseugenol (71.92%); β-caryophyllene (22.80%); chavibetol acetate (2.89%)IntercalationBentonite clay and polyvinylpyrrolidone (PVP)Nano-
composites
Cytotoxic,
Larvicide
-Pharmaceutical, environmental[51]
Ocimum basilicumBasilLeavestrans-β-guaiene (16.89%); α-cadinol (15.66%);
phytol (11.68%)
Nanoemulsion-NanoemulsionsAntioxidant, antibacterial, larvicide-Pharmaceutical, environmental[105]
Satureja hortensisSummer savoryAerial partscarvacrol (35.2%); γ-terpinene (17.6%); thymol (12.1%)Ionic
gelidification
ChitosanNanocapsulesAcaricide-Environmental[107]
Cymbopogon
nardus
Citronella grass--Ionic
gelidification
Chitosan and
cellulose
NanocapsulesInsecticidePest controlEnvironmental[103]
Mentha × piperitaPeppermintDried leavesl-menthone (32.27%);
menthol (23.47%); α-phellandrene (7.71%)
Ionic
Gelidification–
emulsion
ChitosanNanocapsulesInsecticidePest controlEnvironmental[104]
Siparuna
guianensis
NegraminaWhole plant-NanoemulsionChitosanNanocapsulesLarvicidePest controlEnvironmental[106]
Cannabis sativaMarijuanaAerial parts(E)-caryophyllene (23.1%); α-pinene (15.8%);
myrcene (14.5%)
Nano-
precipitation
Alfalfa proteinNanocapsulesAntioxidant-Cosmetic, food[92]
Cymbopogon
densiflorus
LemongrassLeavestrans-p-mentha-2,8-dien-1-ol (13.13%);
cis-p-mentha-2,8- dien-1-ol (17.29%);
trans-p-mentha-1(7),8-dien-2-ol (18.99%)
Phase inversion-NanoemulsionsAntioxidant-Cosmetic[125]
a Part of the plant used in the extraction of the essential oil.

5. Encapsulated EOs versus Non-Encapsulated EOs

The advantages of using encapsulated EOs have recently been demonstrated in studies comparing the use of encapsulated and non-encapsulated oils. Through this evidence, the benefits of using encapsulation methods are clear, since they protect these bioproducts from environmental influences (decomposition by heat, humidity, light and oxygen), reduce volatility, improve stability, thus promoting a longer life, in addition to providing controlled release, which prolongs the biological effect of the compounds [46].
Mukurumbira et al. [126] conducted an extensive review of the effects of in natura and encapsulated essential oils on food contamination by harmful and pathogenic microorganisms. According to the authors, although essential oils are potent antimicrobials, they are chemically and biologically unstable and have strong aromas that limit their application as additives in food. Various encapsulation methods are increasingly being explored as a way to stabilize essential oils, mask their aromas, and possibly enhance their antimicrobial activity with a more sustained release of antimicrobials. The authors also mention the evidence of the greater effectiveness of encapsulated essential oils compared to non-encapsulated essential oils.
Barros et al. [127] evaluated the effect of the application of en-capsulated and non-encapsulated thyme essential oil on the mortality and persistence of the pest Sitophilus zeamais on the quality of corn grains during storage. The study showed that insect mortality was dependent on the concentration and time of exposure to the EO, and that the encapsulated essential oil was the most efficient in combating the insect. In addition, encapsulated EOs did not alter the quality characteristics of corn grains.
Singh et al. [128] conducted an extensive review of the action of essential oils as inhibitors of fungal infestation, their mode of action against fungal growth and the production of mycotoxins. The authors cited the use of nanoencapsulation as a promising new technology for protection of plant-based raw materials (herbal raw materials—HRMs). The use of formulations based on essential oils has been recommended as a green alternative to synthetic preservatives, since they are safer and more environmentally friendly. Nanoencapsulation maintains the stability of EOs and facilitates controlled delivery with improved maintenance of HRMs bioactive ingredients, which can boost the pharmaceutical, food and cosmetic sectors.
Milagres de Almeida et al. [129] studied the bacteriostatic and bactericidal effects of the essential oils of oregano, thyme, cloves, cinnamon and black pepper against strains of Staphylococcus aureus, Listeria sp., Escherichia coli and Salmonella sp., which are agents that contaminate food and cause foodborne illness. The study was conducted with encapsulated EOs and unencapsulated EOs and evaluated the synergistic effect between them. The encapsulation of the EOs of oregano, thyme and cloves was performed with different wall materials obtained by complex coacervation between three different polymers (chitosan, gelatin and gum arabic). The authors observed that encapsulation positively influenced the inhibitory power of the oils by resulting in minimal inhibitory concentrations (MICs), which were lower than those of the oil in an unencapsulated form. It was also proven that the encapsulation potentiated the effect of the EOs evaluated.

6. Conclusions

Currently, EOs are being investigated with prominent intensity, which brings about discoveries of relevant biological properties and the development of environmentally friendly bioproducts. In addition, this demand leverages the need for new technologies to preserve the stability, bioactivity and bioavailability of these substances, and the nanoencapsulation of EOs has been explored as an efficient approach to address such constraints.
In this review, an overview of the techniques, chemical composition and applications of polymer nanoparticles loaded with EOs, prepared via different encapsulation methods and different wall materials, as well as different bioactive elements, was emphasized.
In light of these data, the scenario points to a significant biotechnological potential of nanoencapsulated EOs. However, for future prospects of industrial applications more tests are needed, especially in vivo, in order to provide safe and unquestionable results. In addition, it is essential to develop studies that seek the production of nanoencapsulated EOs on a large scale, in order to meet the demands of the sectors in which they are applied.

Author Contributions

Conceptualization, P.M.A., S.G.A. and L.M.; methodology, C.P.d.A., N.C.d.S.D. and M.T.M.P.; formal analysis, P.M.A., S.G.A. and C.P.d.A.; data curation, S.G.A., C.P.d.A., N.C.d.S.D. and M.T.M.P.; writing—original draft preparation, S.G.A., C.P.d.A., N.C.d.S.D. and M.T.M.P.; writing—review and editing, P.M.A.; project administration, P.M.A. and L.M.; funding acquisition, P.M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Amazonas State Research Foundation-FAPEAM (grant number 01.02.016301.00568/2021-05) and by Coordination of Higher Level Personnel Improvement-CAPES (finance code 001).

Acknowledgments

The authors gratefully acknowledge SisNANO, FAPEAM, UEA, IFAM, CNPq and CAPES for supporting this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chikezie, P.C.; Ibegbulem, C.O.; Mbagwu, F.N. Bioactive principles from medicinal plants. Res. J. Phytochem. 2015, 9, 88–115. [Google Scholar] [CrossRef] [Green Version]
  2. Lammari, N.; Louaer, O.; Meniai, A.H.; Fessi, H.; Elaissari, A. Plant oils, from chemical composition to encapsulated form use. Int. J. Pharm. 2021, 601, 120538. [Google Scholar] [CrossRef] [PubMed]
  3. Maleck, M.; Dutra Dias, T.; Cruz, I.L.S.; Teixeira Serdeiro, M.; Eiras Nascimento, N.; Marins Carraro, V. Essential oils—A brief report. Rev. Eletrônica Teccen. 2021, 14, 43–49. [Google Scholar] [CrossRef]
  4. Bhavaniramya, S.; Vishnupriya, S.; Al-Aboody, M.S. Role of essential oils in food safety, antimicrobial and antioxidant applications. Grain Oil Sci. Technol. 2019, 2, 49–55. [Google Scholar] [CrossRef]
  5. Lin, H.J.; Lin, Y.L.; Huang, B.B.; Lin, Y.T.; Li, H.K.; Lu, W.J.; Lin, T.C.; Tsui, Y.C.; Lin, H.T.V. Solid- and vapour-phase antifungal activities of six essential oils and their applications in postharvest fungal control of peach (Prunus persica L. Batsch). LWT 2022, 156, 113031. [Google Scholar] [CrossRef]
  6. Turek, C.; Stintzing, F.C. Stability of essential oils: A review. Compr. Rev. Food Sci. Food Saf. 2013, 12, 40–53. [Google Scholar] [CrossRef]
  7. Bluma, R.V.; Etcheverry, M.G. Application of essential oils in maize grain: Impact on Aspergillus section Flavi growth parameters and aflatoxin accumulation. Food Microbiol. 2008, 25, 324–334. [Google Scholar] [CrossRef]
  8. Vergis, J.; Gokulakrishnan, P.; Agarwal, R.K.; Kumar, A. Essential oils as natural food antimicrobial agents: A review. Crit. Rev. Food Sci. Nutr. 2015, 55, 1320–1323. [Google Scholar] [CrossRef]
  9. Hasanein, P.; Riahi, H. Antinociceptive and antihyperglycemic effects of Melissa officinalis essential oil in an experimental model of diabetes. Med. Princ. Pract. 2015, 24, 47–52. [Google Scholar] [CrossRef]
  10. Baptista-Silva, S.; Borges, S.; Ramos, O.L.; Pintado, M.; Sarmento, B. The progress of essential oils as potential therapeutic agents: A review. J. Essent. Oil Res. 2020, 32, 279–295. [Google Scholar] [CrossRef]
  11. Lammari, N.; Louaer, O.; Meniai, A.H.; Elaissari, A. Encapsulation of essential oils via nanoprecipitation process: Overview, progress, challenges and prospects. Pharmaceutics 2020, 12, 431. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, W.; Jiang, H.; Rhim, J.W.; Cao, J.; Jiang, W. Effective strategies of sustained release and retention enhancement of essential oils in active food packaging films/coatings. Food Chem. 2022, 367, 130671. [Google Scholar] [CrossRef]
  13. Bastos, L.P.H.; Vicente, J.; Santos, C.H.C.; Carvalho, M.G.; Garcia-Rojas, E.E. Encapsulation of black pepper (Piper nigrum L.) essential oil with gelatin and sodium alginate by complex coacervation. Food Hydrocoll. 2020, 102, 105605. [Google Scholar] [CrossRef]
  14. Giunti, G.; Palermo, D.; Laudani, F.; Algeri, G.M.; Campolo, O.; Palmeri, V. Repellence and acute toxicity of a nano-emulsion of sweet orange essential oil toward two major stored grain insect pests. Ind. Crops Prod. 2019, 142, 41–48. [Google Scholar] [CrossRef]
  15. Gonçalves, N.D.; Lima Pena, F.; Sartoratto, A.; Derlamelina, C.; Duarte, M.C.T.; Antunes, A.E.C.; Prata, A.S. Encapsulated thyme (Thymus vulgaris) essential oil used as a natural preservative in bakery product. Food Res. Int. 2017, 96, 154–160. [Google Scholar] [CrossRef] [PubMed]
  16. Nasrollahzadeh, M.; Sajadi, S.M.; Sajjadi, M.; Issaabadi, Z. Chapter 1—An Introduction to nanotechnology. In An Introduction to Green Nanotechnology; Nasrollahzadeh, M., Sajadi, S.M., Sajjadi, M., Issaabadi, Z., Atarod, M., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; Volume 28, pp. 1–27. [Google Scholar]
  17. Kumari, A.; Yadav, S.K.; Pakade, Y.B.; Singh, B.; Yadav, S.C. Development of biodegradable nanoparticles for delivery of quercetin. Colloid Surf. B. 2010, 80, 184–192. [Google Scholar] [CrossRef] [PubMed]
  18. Mahapatro, A.; Singh, D.K. Biodegradable nanoparticles are excellent vehicle for site directed in-vivo delivery of drugs and vaccines. J. Nanobiotechnol. 2011, 9, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Silva, J.S.M.; Rabelo, M.S.; Lima, S.X.; Analf, R.; Tadei, W.P.; Chaves, F.C.M.; Bezerra, J.A.; Biondo, M.M.; Campelo, P.H.; Sanches, E.A. Biodegradable nanoparticles loaded with Lippia alba essential oil, a sustainable alternative for Aedes aegypti larvae control. Eur. Acad. Res. 2020, VII, 6237–6258. [Google Scholar]
  20. Bilia, A.R.; Guccione, C.; Isacchi, B.; Righeschi, C.; Firenzuoli, F.; Bergonzi, M.C. Essential oils loaded in nanosystems: A developing strategy for a successful therapeutic approach. Evid.-Based Complement. Altern. Med. 2014, 14, 651593. [Google Scholar] [CrossRef] [Green Version]
  21. Donsì, F.; Annunziata, M.; Sessa, M.; Ferrari, G. Nanoencapsulation of essential oils to enhance their antimicrobial activity in foods. LWT 2011, 44, 1908–1914. [Google Scholar] [CrossRef]
  22. Bizerra, A.; Silva, V. Drug delivery systems: Mechanisms and applications. Rev. Saúde Meio Ambiente 2016, 3, 1–12. [Google Scholar]
  23. Roy, A.; Singh, S.K.; Bajpai, J.; Bajpai, A.K. Controlled pesticide release from biodegradable polymers. Cent. Eur. J. Chem. 2014, 12, 453–469. [Google Scholar] [CrossRef]
  24. Neri-Badang, M.C.; Chakraborty, S. Carbohydrate polymers as controlled release devices for pesticides. J. Carbohydr. Chem. 2019, 38, 67–85. [Google Scholar] [CrossRef]
  25. Yasmin, R.; Shah, M.; Khan, S.A.; Ali, R. Gelatin nanoparticles: A potential candidate for medical applications. Nanotechnol. Rev. 2017, 6, 191–207. [Google Scholar] [CrossRef]
  26. Xavier, L.O.; Sganzerla, W.G.; Rosa, G.B.; Rosa, C.G.; Agostinetto, L.; Veeck, A.P.L.; Bretanha, L.C.; Micke, G.A.; Costa, M.D.; Bertoldi, F.C.; et al. Chitosan packaging functionalized with Cinnamodendron dinisii essential oil loaded zein, a proposal for meat conservation. Int. J. Biol. Macromol. 2021, 169, 183–193. [Google Scholar] [CrossRef] [PubMed]
  27. Campelo, M.S.; Melo, E.O.; Arrais, S.P.; Nascimento, F.B.S.A.; Gramosa, N.V.; Soares, S.A.; Ribeiro, M.E.N.P.; Silva, C.R.; Nobre Júnior, H.V.; Ricardo, N.M.P.S. Clove essential oil encapsulated on nanocarrier based on polysaccharide: A strategy for the treatment of vaginal candidiasis. Colloid Surf. A. 2021, 610, 125732. [Google Scholar] [CrossRef]
  28. Azevedo, S.G.; Rocha, A.L.F.; Nunes, R.Z.A.; Pinto, C.C.; Ţălu, Ş.; Fonseca Filho, H.D.; Bezerra, J.A.; Lima, A.R.; Guimarães, F.E.G.; Campelo, P.H.; et al. Pulsatile controlled release and stability evaluation of polymeric particles containing Piper nigrum essential oil and preservatives. Materials 2022, 15, 5415. [Google Scholar] [CrossRef]
  29. Nimesh, S. 2- Methods of nanoparticle preparation. In Woodhead Publishing Series in Biomedicine, Gene Therapy; Ninmesh, S., Ed.; Woodhead Publishing: Sawston, UK, 2013; pp. 13–42. ISBN 9781907568404. [Google Scholar] [CrossRef]
  30. Schaffazick, S.R.; Pohlmann, A.R. Characterization and stability study of suspensions of nanocá capsules and polymeric nanospheres containing diclofenac. Acta Farm. Bonaer. 2002, 21, 99–106. [Google Scholar]
  31. Rao, J.P.; Geckeler, K.E. Polymer nanoparticles: Preparation techniques and size-control parameters. Prog. Polym. Sci. 2011, 36, 887–913. [Google Scholar] [CrossRef]
  32. Puhl, A.C.; Fagundes, M.; Dos Santos, K.C.; Polikarpov, I.; Silva, M.G.F.; Fernandes, J.B.; Vieira, P.C.; Forim, M.R. Preparation and characterization of polymeric nanoparticles loaded with the flavonoid luteolin, by using factorial design. Int. J. Drug Deliv. 2011, 3, 683–698. [Google Scholar] [CrossRef]
  33. Mora-Huertas, C.E.; Fessi, H.; Elaissari, A. Polymer-based nanocapsules for drug delivery. Int. J. Pharm. 2010, 385, 113–142. [Google Scholar] [CrossRef] [PubMed]
  34. Fessi, H.; Puisieux, F.; Devissaguet, J.P.; Ammoury, N.; Benita, S. Nanocapsule formation by interfacial polymer deposition following solvent displacement. Int. J. Pharm. 1989, 55, R1–R4. [Google Scholar] [CrossRef]
  35. Martínez Rivas, C.J.; Tarhini, M.; Badri, W.; Miladi, K.; Greige-Borges, H.; Nazari, Q.A.; Rodrígues, S.A.G.; Román, R.A.; Fessi, H.; Elaissari, A. Nanoprecipitation process: From encapsulation to drug delivery. Int. J. Pharm. 2017, 532, 66–81. [Google Scholar] [CrossRef] [PubMed]
  36. Risch, S.J.; Reineccius, G. Encapsulation release of food and controlled ingredients. In ACS Symposium Series 590; Comstock, M.J., Ed.; American Chemical Society: New York, NY, USA, 1995; p. 214. [Google Scholar] [CrossRef] [Green Version]
  37. Corrado, I.; Di Girolamo, R.; Regalado-González, C.; Pezzella, C. Polyhydroxyalkanoates-based nanoparticles as essential oil carriers. Polymers 2022, 14, 166. [Google Scholar] [CrossRef] [PubMed]
  38. Schaffazick, S.R.; Guterres, S.S.; De Lucca Freitas, L.; Pohlmann, A.R. Physicochemical characterization and stability of the polymeric nanoparticle systems for drug administration. Quim. Nova 2003, 26, 726–737. [Google Scholar] [CrossRef]
  39. Grillo, R.; Souza, P.M.S.; Rosa, A.H.; Fraceto, L.F. Nanopartículas poliméricas como sistemas de liberação para herbicidas. In Nanotecnologia, Ciência e Engenharia; Graeff, C., Ed.; Cultura Academica: São Paulo, Brazil, 2012; pp. 83–124. ISBN 9788579833779. [Google Scholar]
  40. Sahoo, N.; Sahoo, R.K.; Biswas, N.; Guha, A.; Kuotsu, K. Recent advancement of gelatin nanoparticles in drug and vaccine delivery. Int. J. Biol. Macromol. 2015, 81, 317–331. [Google Scholar] [CrossRef]
  41. Elzoghby, A.O. Gelatin-based nanoparticles as drug and gene delivery systems: Reviewing three decades of research. J. Control Release 2013, 172, 1075–1091. [Google Scholar] [CrossRef]
  42. Quintanar-Guerrero, D.; Allémann, E.; Fessi, H.; Doelker, E. Preparation techniques and mechanisms of formation of biodegradable nanoparticles from preformed polymers. Drug Dev. Ind. Pharm. 1998, 24, 1113–1128. [Google Scholar] [CrossRef]
  43. Quintanar-Guerrero, D.; Allémann, E.; Doelker, E.; Fessi, H. Preparation and characterization of nanocapsnles from preformed polymers by a new process based on emulsification-diffusion technique. Pharm. Res. 1998, 15, 1056–1062. [Google Scholar] [CrossRef]
  44. Calo, J.R.; Crandall, P.G.; O’Bryan, C.A.; Ricke, S.C. Essential oils as antimicrobials in food systems: A review. Food Control 2015, 54, 111–119. [Google Scholar] [CrossRef]
  45. Dhifi, W.; Bellili, S.; Jazi, S.; Bahloul, N.; Mnif, W. Essential oils chemical characterization and investigation of some biological activities: A critical review. Medicines 2016, 3, 25. [Google Scholar] [CrossRef] [PubMed]
  46. Kreutz, T.; Carneiro, S.B.; Soares, K.D.; Limberger, R.P.; Apel, M.A.; Veiga-Junior, V.F.; Koester, L.S. Aniba canelilla (Kunth) Mez essential oil-loaded nanoemulsion: Improved stability of the main constituents and in vitro antichemotactic activity. Ind. Crops Prod. 2021, 171, 113949. [Google Scholar] [CrossRef]
  47. Singh, B.K.; Tiwari, S.; Chaudhari, A.K.; Maurya, A.; Das, S.; Singh, V.K.; Dubey, N.K. Chitosan encompassed Aniba rosaeodora essential oil as innovative green candidate for antifungal and antiaflatoxigenic activity in millets with emphasis on cellular and its mode of action. Front. Microbiol. 2022, 13, 970670. [Google Scholar] [CrossRef] [PubMed]
  48. Antonioli, G.; Fontanella, G.; Echeverrigaray, S.; Longaray Delamare, A.P.; Fernandes Pauletti, G.; Barcellos, T. Poly(lactic acid) nanocapsules containing lemongrass essential oil for postharvest decay control: In vitro and in vivo evaluation against phytopathogenic fungi. Food Chem. 2020, 326, 126997. [Google Scholar] [CrossRef] [PubMed]
  49. Das, S.; Singh, V.K.; Dwivedy, A.K.; Chaudhari, A.K.; Dubey, N.K. Nanostructured Pimpinella anisum essential oil as novel green food preservative against fungal infestation, aflatoxin B1 contamination and deterioration of nutritional qualities. Food Chem. 2021, 344, 128574. [Google Scholar] [CrossRef] [PubMed]
  50. Khezri, K.; Farahpour, M.R.; Mounesi Rad, S. Efficacy of Mentha pulegium essential oil encapsulated into nanostructured lipid carriers as an in vitro antibacterial and infected wound healing agent. Colloid Surf. A 2020, 589, 124414. [Google Scholar] [CrossRef]
  51. Nunes, R.K.V.; Santos, J.S.; Santos, C.P.; Gonsalves, J.K.M.C.; Lira, A.A.M.; Cavalcanti, S.C.H.; Santos, R.L.C.; Sarmento, V.H.V.; Nunes, R.G. Clay/PVP nanocomposites enriched with Syzygium aromaticum essential oil as a safe formulation against Aedes aegypti larvae. Appl. Clay Sci. 2020, 185, 105394. [Google Scholar] [CrossRef]
  52. Prasad, J.; Das, S.; Maurya, A.; Jain, S.K.; Dwivedy, A.K. Synthesis, characterization and in situ bioefficacy evaluation of Cymbopogon nardus essential oil impregnated chitosan nanoemulsion against fungal infestation and aflatoxin B1 contamination in food system. Int. J. Biol. Macromol. 2022, 205, 240–252. [Google Scholar] [CrossRef]
  53. Dwivedy, A.K.; Singh, V.K.; Prakash, B.; Dubey, N.K. Nanoencapsulated Illicium verum Hook.f. essential oil as an effective novel plant-based preservative against aflatoxin B1 production and free radical generation. Food Chem. Toxicol. 2018, 111, 102–113. [Google Scholar] [CrossRef]
  54. Benjemaa, M.; Neves, M.A.; Falleh, H.; Isoda, H.; Ksouri, R.; Nakajima, M. Nanoencapsulation of Thymus capitatus essential oil: Formulation process, physical stability, characterization and antibacterial efficiency monitoring. Ind. Crops Prod. 2018, 113, 414–421. [Google Scholar] [CrossRef]
  55. Kundu, A.; Mandal, A.; Dutta, A.; Saha, S.; Raina, A.P.; Kumar, R.; Ghosh, A. Nanoemulsification of Kaempferia galanga essential oil: Characterizations and molecular interactions explaining fungal growth suppression. Process Biochem. 2022, 121, 228–239. [Google Scholar] [CrossRef]
  56. Singh, B.K.; Chaudhari, A.K.; Das, S.; Tiwari, S.; Dubey, N.K. Preparation and characterization of a novel nanoemulsion consisting of chitosan and Cinnamomum tamala essential oil and its effect on shelf-life lengthening of stored millets. Pestic. Biochem. Physiol. 2022, 187, 105214. [Google Scholar] [CrossRef] [PubMed]
  57. Kumar, A.; Pratap Singh, P.; Prakash, B. Unravelling the antifungal and anti-aflatoxin B1 mechanism of chitosan nanocomposite incorporated with Foeniculum vulgare essential oil. Carbohydr. Polym. 2020, 236, 116050. [Google Scholar] [CrossRef] [PubMed]
  58. Das, S.; Singh, V.K.; Dwivedy, A.K.; Chaudhari, A.K.; Upadhyay, N.; Singh, P.; Sharma, S.; Dubey, N.K. Encapsulation in chitosan-based nanomatrix as an efficient green technology to boost the antimicrobial, antioxidant and in situ efficacy of Coriandrum sativum essential oil. Int. J. Biolog. Macromol. 2019, 133, 294–305. [Google Scholar] [CrossRef]
  59. Hasheminejad, N.; Khodaiyan, F.; Safari, M. Improving the antifungal activity of clove essential oil encapsulated by chitosan nanoparticles. Food Chem. 2019, 275, 113–122. [Google Scholar] [CrossRef]
  60. 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, 117410. [Google Scholar] [CrossRef]
  61. Kujur, A.; Kumar, A.; Prakash, B. Elucidation of antifungal and aflatoxin B1 inhibitory mode of action of Eugenia caryophyllata L. essential oil loaded chitosan nanomatrix against Aspergillus flavus. Pestic. Biochem. Physiol. 2021, 172, 104755. [Google Scholar] [CrossRef]
  62. Fornari, T.; Vicente, G.; Vázquez, E.; García-Risco, M.R.; Reglero, G. Isolation of essential oil from different plants and herbs by supercritical fluid extraction. J. Chromatogr. A. 2012, 1250, 34–48. [Google Scholar] [CrossRef] [Green Version]
  63. Craveiro, A.A.; Alencar, J.W.; Matos, F.J.A.; Andrade, C.H.S.; Machado, M.I.L. Essential oils from Brazilian Verbenaceae genus Lippia. J. Nat. Prod. 1981, 44, 598–601. [Google Scholar] [CrossRef]
  64. Chaar, J.D.S. Analytical and Chemical Studies by Acetylation of Linalool Contained in the Essential Oil of the Species Aniba duckei Kostermans. Ph.D. Thesis, University of São Paulo, São Paulo, Brazil, 2000; 150p. [Google Scholar]
  65. Guenther, E. The production of essential oils. In The Essential Oils; Guenther, E., Ed.; Krieger Publishing: New York, NY, USA, 1972; pp. 453–454. [Google Scholar]
  66. Santos, F.J.; Lopes, J.A.D.; Cito, A.G.L.; Oliveira, E.H.; Lima, S.G.; Reis, F.D.A. Composition and biological activity of essential oils from Lippia origanoides HBK. J. Essent. Oil Res. 2004, 16, 504–506. [Google Scholar] [CrossRef]
  67. Hayouni, E.A.; Chraief, I.; Abedrabba, M.; Bouix, M.; Leveau, J.Y.; Mohammed, H.; Hamdi, M. Tunisian Salvia officinalis L. and Schinus molle L. essential oils: Their chemical compositions and their preservative effects against Salmonella inoculated in minced beef meat. Int. J. Food Microbiol. 2008, 125, 242–251. [Google Scholar] [CrossRef] [PubMed]
  68. Prakash, B.; Kujur, A.; Yadav, A.; Kumar, A.; Singh, P.P.; Dubey, N.K. Nanoencapsulation, an efficient technology to boost the antimicrobial potential of plant essential oils in food system. Food Control 2018, 89, 1–11. [Google Scholar] [CrossRef]
  69. Kuorwel, K.K.; Cran, M.J.; Sonneveld, K.; Miltz, J.; Bigger, S.W. Essential oils and their principal constituents as antimicrobial agents for synthetic packaging films. J. Food Sci. 2011, 76, 164–177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Kumar, S.; Mendiratta, S.K.; Agrawal, R.K.; Sharma, H.; Singh, B.P. Anti-oxidant and anti-microbial properties of mutton nuggets incorporated with blends of essential oils. J. Food Sci. Technol. 2018, 55, 821–832. [Google Scholar] [CrossRef] [PubMed]
  71. Granata, G.; Stracquadanio, S.; Leonardi, M.; Napoli, E.; Consoli, G.M.L.; Cafiso, V.; Stefani, S.; Geraci, C. Essential oils encapsulated in polymer-based nanocapsules as potential candidates for application in food preservation. Food Chem. 2018, 269, 286–292. [Google Scholar] [CrossRef]
  72. De Sousa, A.C.; Gattass, C.R.; Alviano, D.S.; Alviano, C.S.; Blank, A.F.; Alves, P.B. Melissa officinalis L. essential oil, antitumoral and antioxidant activities. J. Pharm. Pharmacol. 2004, 56, 677–681. [Google Scholar] [CrossRef]
  73. Stan, M.S.; Chirila, L.; Popescu, A.; Radulescu, D.M.; Radulescu, D.E.; Dinischiotu, A. Materials essential oil microcapsules immobilized on textiles and certain induced effects. Materials 2019, 12, 2029. [Google Scholar] [CrossRef] [Green Version]
  74. Ghodrati, M.; Farahpour, M.R.; Hamishehkar, H. Encapsulation of Peppermint essential oil in nanostructured lipid carriers: In-vitro antibacterial activity and accelerative effect on infected wound healing. Colloid Surf. A 2019, 564, 161–169. [Google Scholar] [CrossRef]
  75. Moazeni, M.; Davari, A.; Shabanzadeh, S.; Akhtari, J.; Saeed, M.; Mortyeza-Semnani, K.; Abastabar, M.; Nabili, M.; Moghadam, F.H.; Roohi, B.; et al. In vitro antifungal activity of Thymus vulgaris essential oil nanoemulsion. J. Herb. Med. 2021, 28, 100452. [Google Scholar] [CrossRef]
  76. Rozman, N.A.S.; Yenn, T.W.; Ring, L.C.; Anuar, M.R.; Karim, S.; Kooi, O.S.; Yusof, F.A.M.; Wen-Nee, T.; Sulaiman, B.; Lee, O.M.; et al. Homalomena pineodora essential oil nanoparticle inhibits diabetic wound pathogens. Sci. Rep. 2020, 10, 3307. [Google Scholar] [CrossRef] [Green Version]
  77. Samling, B.A.; Assim, Z.; Tong, W.Y.; Leong, C.R.; Rashid, S.A.; Kamal, N.N.S.N.M.; Muhamad, M.; Tan, W.M. Cynometra cauliflora essential oils loaded-chitosan nanoparticles: Evaluations of their antioxidant, antimicrobial and cytotoxic activities. Int. J. Biol. Macromol. 2022, 210, 742–751. [Google Scholar] [CrossRef] [PubMed]
  78. Rajivgandhi, G.; Saravanan, K.; Ramachandran, G.; Li, J.L.; Yin, L.; Quero, F.; Alharbi, N.S.; Kadaikunnan, S.; Khaled, J.M.; Manoharan, N.; et al. Enhanced anti-cancer activity of chitosan loaded Morinda citrifolia essential oil against A549 human lung cancer cells. Int. J. Biol. Macromol. 2020, 164, 4010–4021. [Google Scholar] [CrossRef] [PubMed]
  79. Alipanah, H.; Farjam, M.; Zarenezhad, E.; Roozitalab, G.; Osanloo, M. Chitosan nanoparticles containing limonene and limonene-rich essential oils, potential phytotherapy agents for the treatment of melanoma and breast cancers. BMC Complement. Med. Ther. 2021, 21, 186. [Google Scholar] [CrossRef] [PubMed]
  80. Ali, H.; Al-Khalifa, A.R.; Aouf, A.; Boukhebti, H.; Farouk, A. Effect of nanoencapsulation on volatile constituents, and antioxidant and anticancer activities of Algerian Origanum glandulosum Desf. essential oil. Sci. Rep. 2020, 10, 2812. [Google Scholar] [CrossRef] [Green Version]
  81. Granata, G.; Stracquadanio, S.; Leonardi, M.; Napoli, E.; Malandrino, G.; Cafiso, V.; Stefani, S.; Geraci, C. Molecules oregano and thyme essential oils encapsulated in chitosan nanoparticles as effective antimicrobial agents against foodborne pathogens. Molecules 2021, 26, 4055. [Google Scholar] [CrossRef]
  82. Yostawonkul, J.; Nittayasut, N.; Phasuk, A.; Junchay, R.; Boonrungsiman, S.; Temisak, S.; Kongsema, M.; Phoolcharoen, W.; Yata, T. Nano/microstructured hybrid composite particles containing Cinnamon oil as an antibiotic alternative against food-borne pathogens. J. Food Eng. 2021, 290, 110209. [Google Scholar] [CrossRef]
  83. Liew, S.N.; Utra, U.; Alias, A.K.; Tan, T.B.; Tan, C.P.; Yussof, N.S. Physical, morphological and antibacterial properties of lime essential oil nanoemulsions prepared via spontaneous emulsification method. LWT 2020, 128, 162–169. [Google Scholar] [CrossRef]
  84. Paim, L.F.N.A.; Dalla Lana, D.F.; Giaretta, M.; Danielli, L.J.; Fuentefria, A.M.; Apel, M.A.; Külkamp-Guerreiro, I.C. Poiretia latifolia essential oil as a promising antifungal and anti-inflammatory agent, chemical composition, biological screening, and development of a nanoemulsion formulation. Ind. Crops Prod. 2018, 126, 280–286. [Google Scholar] [CrossRef]
  85. Barrera-Ruiz, D.G.; Cuestas-Rosas, G.C.; Sánchez-Mariñez, R.I.; Álvarez-Ainza, M.L.; Moreno-Ibarra, G.M.; López-Meneses, A.K.; Plascencia-Jatomea, M.; Cortez-Rocha, M.O. Antibacterial activity of essential oils encapsulated in chitosan nanoparticles. Food Sci. Technol. 2020, 40, 568–573. [Google Scholar] [CrossRef]
  86. Fazly, B.S.B.; Khameneh, B.; Namazi, N.; Iranshahi, M.; Davoodi, D.; Golmohammadzadeh, S. Solid lipid nanoparticles carrying Eugenia caryophyllata essential oil, the novel nanoparticulate systems with broad-spectrum antimicrobial activity. Lett. Appl. Microbiol. 2018, 66, 506–513. [Google Scholar] [CrossRef]
  87. Soltanzadeh, M.; Peighambardoust, S.H.; Ghanbarzadeh, B.; Mohammadi, M.; Lorenzo, J.M. Chitosan nanoparticles encapsulating lemongrass (Cymbopogon commutatus) essential oil, physicochemical, structural, antimicrobial and in-vitro release properties. Int. J. Biol. Macromol. 2021, 192, 1084–1097. [Google Scholar] [CrossRef] [PubMed]
  88. Oliveira, C.; Coelho, C.; Teixeira, J.A.; Ferreira-Santos, P.; Botelho, C.M. Nanocarriers as active ingredients enhancers in the cosmetic industry—The European and North America regulation challenges. Molecules 2022, 27, 1669. [Google Scholar] [CrossRef] [PubMed]
  89. Huang, K.; Liu, R.; Zhang, Y.; Guan, X. Characteristics of two cedarwood essential oil emulsions and their antioxidant and antibacterial activities. Food Chem. 2021, 346, 128970. [Google Scholar] [CrossRef]
  90. Sampaio, C.I.; Bourbon, A.I.; Gonçalves, C.; Pastrana, L.M.; Dias, A.M.; Cerqueira, M.A. Low energy nanoemulsions as carriers of thyme and lemon balm essential oils. LWT 2022, 154, 112748. [Google Scholar] [CrossRef]
  91. Shetta, A.; Kegere, J.; Mamdouh, W. Comparative study of encapsulated peppermint and green tea essential oils in chitosan nanoparticles: Encapsulation, thermal stability, in-vitro release, antioxidant and antibacterial activities. Int. J. Biol. Macromol. 2019, 126, 731–742. [Google Scholar] [CrossRef]
  92. Hadidi, M.; Rostamabadi, H.; Moreno, A.; Jafari, S.M. Nanoencapsulation of essential oils from industrial hemp (Cannabis sativa L.) by-products into alfalfa protein nanoparticles. Food Chem. 2022, 386, 132765. [Google Scholar] [CrossRef]
  93. Karimirad, R.; Behnamian, M.; Dezhsetan, S.; Sonnenberg, A. Chitosan nanoparticles-loaded Citrus aurantium essential oil, a novel delivery system for preserving the postharvest quality of Agaricus bisporus. J. Sci. Food Agric. 2018, 98, 5112–5119. [Google Scholar] [CrossRef] [PubMed]
  94. Chaudhari, A.K.; Das, S.; Singh, B.K.; Kishore Dubey, N. Green facile synthesis of cajuput (Melaleuca cajuputi Powell.) essential oil loaded chitosan film and evaluation of its effectiveness on shelf-life extension of white button mushroom. Food Chem. 2023, 401, 134114. [Google Scholar] [CrossRef]
  95. Arabpoor, B.; Yousefi, S.; Weisany, W.; Ghasemlou, M. Multifunctional coating composed of Eryngium campestre L. essential oil encapsulated in nano-chitosan to prolong the shelf-life of fresh cherry fruits. Food Hydrocoll. 2021, 111, 106394. [Google Scholar] [CrossRef]
  96. Yadav, A.; Kujur, A.; Kumar, A.; Singh, P.P.; Prakash, B.; Dubey, N.K. Assessing the preservative efficacy of nanoencapsulated mace essential oil against food borne molds, aflatoxin B1 contamination, and free radical generation. LWT 2019, 108, 429–436. [Google Scholar] [CrossRef]
  97. Das, S.; Singh, V.K.; Dwivedy, A.K.; Chaudhari, A.K.; Upadhyay, N.; Singh, A.; Dubey, N.K. Antimicrobial activity, anti-aflatoxigenic potential and in situ efficacy of novel formulation comprising of Apium graveolens essential oil and its major component. Pestic. Biochem. Physiol. 2019, 160, 102–111. [Google Scholar] [CrossRef] [PubMed]
  98. Deepika; Chaudhari, A.K.; Singh, A.; Das, S.; Dubey, N.K. Nanoencapsulated Petroselinum crispum essential oil: Characterization and practical efficacy against fungal and aflatoxin contamination of stored chia seeds. Food Biosci. 2021, 42, 101117. [Google Scholar] [CrossRef]
  99. Cai, M.; Wang, Y.; Wang, R.; Li, M.; Zhang, W.; Yu, J.; Hua, R. Antibacterial and antibiofilm activities of chitosan nanoparticles loaded with Ocimum basilicum L. essential oil. Int. J. Biol. Macromol. 2022, 202, 122–129. [Google Scholar] [CrossRef] [PubMed]
  100. Wan, J.; Zhong, S.; Schwarz, P.; Chen, B.; Rao, J. Physical properties, antifungal and mycotoxin inhibitory activities of five essential oil nanoemulsions, Impact of oil compositions and processing parameters. Food Chem. 2019, 291, 199–206. [Google Scholar] [CrossRef]
  101. Sagar, N.A.; Agrawal, R.K.; Singh, R.; Talukder, S.; Kumar, R.R.; Mendiratta, S.K. Quality and shelf life assessment of steam-cooked chicken fingers coated with essential oil nanoemulsions. Food Biosci. 2022, 49, 101902. [Google Scholar] [CrossRef]
  102. 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]
  103. 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]
  104. 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]
  105. Sundararajan, B.; Moola, A.K.; Vivek, K.; Kumari, B.D.R. Formulation of nanoemulsion from leaves essential oil of Ocimum basilicum L. and its antibacterial, antioxidant and larvicidal activities (Culex quinquefasciatus). Microb. Pathog. 2018, 125, 475–485. [Google Scholar] [CrossRef]
  106. Ferreira, T.P.; Haddi, K.; Corrêa, R.F.T.; Zapata, V.L.B.; Piau, T.B.; Souza, L.F.N.; Santos, S.M.G.; Oliveira, E.E.; Jumbo, L.O.V.; Ribeiro, B.M.; et al. Prolonged mosquitocidal activity of Siparuna guianensis essential oil encapsulated in chitosan nanoparticles. PLoS Negl. Trop. Dis. 2019, 13, e0007624. [Google Scholar] [CrossRef] [Green Version]
  107. Ahmadi, Z.; Saber, M.; Akbari, A.; Mahdavinia, G.R. Encapsulation of Satureja hortensis L. (Lamiaceae) in chitosan/TPP nanoparticles with enhanced acaricide activity against Tetranychus urticae Koch (Acari, Tetranychidae). Ecotoxicol. Environ. Saf. 2018, 161, 111–119. [Google Scholar] [CrossRef] [PubMed]
  108. Clarinval, A.M.; Halleux, J. Classification of biodegradable polymers. In Biodegradable Polymers for Industrial Applications; Smith, R., Ed.; Elsevier: Chennai, India, 2005; pp. 3–31. ISBN 9781855739345. [Google Scholar]
  109. Brito, G.F.; Agrawal, P.; Araújo, E.M.; Mélo, T.J.A. Biopolymers, biodegradable polymers and green polymers. Rev. Eletrônica Mater. Process. 2011, 2, 127–139. [Google Scholar]
  110. Polman, E.M.N.; Gruter, G.J.M.; Parsons, J.R.; Tietema, A. Comparison of the aerobic biodegradation of biopolymers and the corresponding bioplastics: A review. Sci. Total Environ. 2021, 753, 141953. [Google Scholar] [CrossRef] [PubMed]
  111. Awasthi, M.K.; Sarsaiya, S.; Patel, A.; Juneja, A.; Singh, R.P.; Yan, B.; Awasthi, S.K.; Jain, A.; Liu, T.; Duan, Y.; et al. Refining biomass residues for sustainable energy and bio-products: An assessment of technology, its importance, and strategic applications in circular bio-economy. Renew. Sustain. Energy Rev. 2020, 127, 1098. [Google Scholar] [CrossRef]
  112. Miranda, M.; Sun, X.; Marín, A.; Santos, L.C.; Plotto, A.; Bai, J.; Assis, O.B.G.; Ferreira, M.D.; Baldwin, E. Nano-and micro-sized carnauba wax emulsions-based coatings incorporated with ginger essential oil and hydroxypropyl methylcellulose on papaya: Preservation of quality and delay of post-harvest fruit decay. Food Chem. X 2022, 13, 100249. [Google Scholar] [CrossRef]
  113. Martin-Piñero, M.J.; Muñoz, J.; Alfaro-Rodriguez, M.C. Improvement of the rheological properties of rosemary oil nanoemulsions prepared by microfluidization and vacuum evaporation. J. Ind. Eng. Chem. 2020, 91, 340–346. [Google Scholar] [CrossRef]
  114. Chaudhari, A.K.; Singh, V.K.; Das, S.; Prasad, J.; Dwivedy, A.K.; Dubey, N.K. Improvement of in vitro and in situ antifungal, AFB1 inhibitory and antioxidant activity of Origanum majorana L. essential oil through nanoemulsion and recommending as novel food preservative. Food Chem. Toxicol. 2020, 143, 111536. [Google Scholar] [CrossRef]
  115. Das, S.; Singh, V.K.; Dwivedy, A.K.; Chaudhari, A.K.; Upadhyay, N.; Singh, A.; Dubey, N.K. Fabrication, characterization and practical efficacy of Myristica fragrans essential oil nanoemulsion delivery system against postharvest biodeterioration. Ecotoxicol. Environ. Saf. 2020, 189, 110000. [Google Scholar] [CrossRef]
  116. Kujur, A.; Kumar, A.; Yadav, A.; Prakash, B. Antifungal and aflatoxin B1 inhibitory efficacy of nanoencapsulated Pelargonium graveolens L. essential oil and its mode of action. LWT 2020, 130, 109619. [Google Scholar] [CrossRef]
  117. Roshan, A.B.; Venkatesh, H.N.; Dubey, N.K.; Mohana, D.C. Chitosan-based nanoencapsulation of Toddalia asiatica (L.) Lam. essential oil to enhance antifungal and aflatoxin B1 inhibitory activities for safe storage of maize. Int. J. Biol. Macromol. 2022, 204, 476–484. [Google Scholar] [CrossRef]
  118. Yadav, A.; Kujur, A.; Kumar, A.; Singh, P.P.; Gupta, V.; Prakash, B. Encapsulation of Bunium persicum essential oil using chitosan nanopolymer: Preparation, characterization, antifungal assessment, and thermal stability. Int. J. Biol. Macromol. 2020, 142, 172–180. [Google Scholar] [CrossRef] [PubMed]
  119. Falleh, H.; Ben Jemaa, M.; Neves, M.A.; Isoda, H.; Nakajima, M.; Ksouri, R. Peppermint and Myrtle nanoemulsions: Formulation, stability, and antimicrobial activity. LWT 2021, 152, 112377. [Google Scholar] [CrossRef]
  120. Chuesiang, P.; Siripatrawan, U.; Sanguandeekul, R.; McLandsborough, L.; Julian McClements, D. Optimization of Cinnamon oil nanoemulsions using phase inversion temperature method, impact of oil phase composition and surfactant concentration. J. Colloid Interface Sci. 2018, 514, 208–216. [Google Scholar] [CrossRef] [PubMed]
  121. Payandeh, M.; Ahmadyousefi, M.; Alizadeh, H.; Zahedifar, M. Chitosan nanocomposite incorporated Satureja kermanica essential oil and extract: Synthesis, characterization and antifungal assay. Int. J. Biol. Macromol. 2022, 221, 1356–1364. [Google Scholar] [CrossRef]
  122. Kalagatur, N.K.; Nirmal Ghosh, O.S.; Sundararaj, N.; Mudili, V. Antifungal activity of chitosan nanoparticles encapsulated with Cymbopogon martinii essential oil on plant pathogenic fungi Fusarium graminearum. Front. Pharmacol. 2018, 9, 610. [Google Scholar] [CrossRef] [Green Version]
  123. Shen, Y.; Ni, Z.J.; Thakur, K.; Zhang, J.G.; Hu, F.; Wei, Z.J. Preparation and characterization of clove essential oil loaded nanoemulsion and pickering emulsion activated pullulan-gelatin based edible film. Int. J. Biol. Macromol. 2021, 181, 528–539. [Google Scholar] [CrossRef]
  124. El-Sayed, S.M.; El-Sayed, H.S. Antimicrobial nanoemulsion formulation based on thyme (Thymus vulgaris) essential oil for UF labneh preservation. J. Mater. Res. Technol. 2021, 10, 1029–1041. [Google Scholar] [CrossRef]
  125. Seibert, J.B.; Rodrigues, I.V.; Carneiro, S.P.; Amparo, T.R.; Lanza, J.S.; Frézard, F.J.G.; Souza, G.H.B.; Santos, O.D.H. Seasonality study of essential oil from leaves of Cymbopogon densiflorus and nanoemulsion development with antioxidant activity. Flavour Fragr. J. 2019, 34, 5–14. [Google Scholar] [CrossRef] [Green Version]
  126. Mukurumbira, A.R.; Shellie, R.A.; Keast, R.; Palombo, E.A.; Jadhav, S.R. Encapsulation of essential oils and their application in antimicrobial active packaging. Food Control 2022, 136, 108883. [Google Scholar] [CrossRef]
  127. Barros, F.A.P.; Radünz, M.; Scariot, M.A.; Camargo, T.M.; Nunes, C.F.P.; de Souza, R.R.; Gilson, I.K.; Hackbart, H.C.S.; Radünz, L.L.; Oliveira, J.V.; et al. Efficacy of encapsulated and non-encapsulated thyme essential oil (Thymus vulgaris L.) in the control of Sitophilus zeamais and its effects on the quality of corn grains throughout storage. Crop Prot. 2022, 153, e10812. [Google Scholar] [CrossRef]
  128. Singh, B.K.; Tiwari, S.; Maurya, A.; Kumar, S.; Dubey, N.K. Fungal and mycotoxin contamination of herbal raw materials and their protection by nanoencapsulated essential oils: An overview. Biocatal. Agric. Biotechnol. 2022, 39, 102257. [Google Scholar] [CrossRef]
  129. Milagres de Almeida, J.; Crippa, B.L.; Martins Alencar de Souza, V.V.; Perez Alonso, V.P.; da Motta Santos Júnior, E.; Picone, C.F.S.; Prata, A.S.; Silva, N.C.C. Antimicrobial action of Oregano, Thyme, Clove, Cinnamon and Black pepper essential oils free and encapsulated against foodborne pathogens. Food Control 2023, 144, 109356. [Google Scholar] [CrossRef]
Figure 1. Scheme of the controlled release of essential oil and the degradation of polymeric nanoparticles. The nanosystem comprises the active ingredient encapsulated by a biodegradable polymer. Controlled release occurs due to the degradation of the polymer. When the polymer interacts with the environment in which the release takes place, there is a breakdown of the wall material, which can be due to thermal degradation or photodegradation, and then the release of the active ingredient occurs.
Figure 1. Scheme of the controlled release of essential oil and the degradation of polymeric nanoparticles. The nanosystem comprises the active ingredient encapsulated by a biodegradable polymer. Controlled release occurs due to the degradation of the polymer. When the polymer interacts with the environment in which the release takes place, there is a breakdown of the wall material, which can be due to thermal degradation or photodegradation, and then the release of the active ingredient occurs.
Polymers 14 05495 g001
Figure 2. Polymer nanoparticles. (a) Nanocapsule: active ingredients dissolved in the matrix core and (b) Nanospheres: active ingredients dissolved throughout the polymer matrix. In the nanocapsules, the active substance is in the nucleus and is surrounded by a polymeric membrane. In the nanosphere, the active substance is dispersed in the polymeric matrix, and therefore it does not have a defined nucleus.
Figure 2. Polymer nanoparticles. (a) Nanocapsule: active ingredients dissolved in the matrix core and (b) Nanospheres: active ingredients dissolved throughout the polymer matrix. In the nanocapsules, the active substance is in the nucleus and is surrounded by a polymeric membrane. In the nanosphere, the active substance is dispersed in the polymeric matrix, and therefore it does not have a defined nucleus.
Polymers 14 05495 g002
Figure 3. Formation of polymeric nanoparticles using the nanoprecipitation method. The organic phase consists of a solution—or a mixture—of a polymer (carrier), the organic solvent, the active substance (natural active ingredient), oil and a lipophilic surfactant. The aqueous phase consists of a mixture of surfactant (stabilizer) and ultrapure water. Using a burette and a magnetic stirrer, the organic phase is slowly added to the aqueous phase.
Figure 3. Formation of polymeric nanoparticles using the nanoprecipitation method. The organic phase consists of a solution—or a mixture—of a polymer (carrier), the organic solvent, the active substance (natural active ingredient), oil and a lipophilic surfactant. The aqueous phase consists of a mixture of surfactant (stabilizer) and ultrapure water. Using a burette and a magnetic stirrer, the organic phase is slowly added to the aqueous phase.
Polymers 14 05495 g003
Figure 4. Formation of carrier systems using the emulsion–diffusion method. The organic phase contains the polymer, the organic solvent, the natural active ingredient, and the oil. The aqueous phase contains a different polymer, a stabilizer and the ultrapure water. The particles are formed by polymer precipitation along with the encapsulation of the natural active ingredient.
Figure 4. Formation of carrier systems using the emulsion–diffusion method. The organic phase contains the polymer, the organic solvent, the natural active ingredient, and the oil. The aqueous phase contains a different polymer, a stabilizer and the ultrapure water. The particles are formed by polymer precipitation along with the encapsulation of the natural active ingredient.
Polymers 14 05495 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

Albuquerque, P.M.; Azevedo, S.G.; de Andrade, C.P.; D’Ambros, N.C.d.S.; Pérez, M.T.M.; Manzato, L. Biotechnological Applications of Nanoencapsulated Essential Oils: A Review. Polymers 2022, 14, 5495. https://doi.org/10.3390/polym14245495

AMA Style

Albuquerque PM, Azevedo SG, de Andrade CP, D’Ambros NCdS, Pérez MTM, Manzato L. Biotechnological Applications of Nanoencapsulated Essential Oils: A Review. Polymers. 2022; 14(24):5495. https://doi.org/10.3390/polym14245495

Chicago/Turabian Style

Albuquerque, Patrícia Melchionna, Sidney Gomes Azevedo, Cleudiane Pereira de Andrade, Natália Corrêa de Souza D’Ambros, Maria Tereza Martins Pérez, and Lizandro Manzato. 2022. "Biotechnological Applications of Nanoencapsulated Essential Oils: A Review" Polymers 14, no. 24: 5495. https://doi.org/10.3390/polym14245495

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