Next Article in Journal
The Relationship between the Density of Winter Canola Stand and Weed Vegetation
Previous Article in Journal
The Impact of Digital Capabilities on Peasants’ Wage Growth: Evidence from Chinese Farmer Entrepreneurs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Applications of Plant Essential Oils in Pest Control and Their Encapsulation for Controlled Release: A Review

by
Rocío Ayllón-Gutiérrez
,
Laura Díaz-Rubio
,
Myriam Montaño-Soto
,
María del Pilar Haro-Vázquez
and
Iván Córdova-Guerrero
*
Facultad de Ciencias Químicas e Ingeniería, Universidad Autónoma de Baja California, Tijuana 22390, Mexico
*
Author to whom correspondence should be addressed.
Agriculture 2024, 14(10), 1766; https://doi.org/10.3390/agriculture14101766
Submission received: 3 September 2024 / Revised: 27 September 2024 / Accepted: 3 October 2024 / Published: 6 October 2024
(This article belongs to the Special Issue Preparation, Function and Application of Agrochemicals)

Abstract

:
Essential oils (EOs) are volatile products derived from the secondary metabolism of plants with antioxidant, antimicrobial, and pesticidal properties. They have traditionally been used in medicine, cosmetics, and food additives. In agriculture, EOs stand out as natural alternatives for pest control, as they show biocidal, repellent, and antifeedant effects. However, they are highly volatile compounds and susceptible to oxidation, which has limited their use as pesticides. This has led to exploring micro- and nano-scale encapsulation to protect these compounds, improving their stability and allowing for a controlled release. Various encapsulation techniques exist, such as emulsification, ionic gelation, and complex coacervation. Nanoemulsions are useful in the food industry, while ionic gelation and complex coacervation offer high encapsulation efficiency. Materials such as chitosan, gelatin-gum-Arabic, and cyclodextrins are promising for agricultural applications, providing stability and the controlled release of EOs. Encapsulation technology is still under development but offers sustainable alternatives to conventional agrochemicals. This article reviews the potential of EOs in pest management and encapsulation techniques that enhance their efficacy.

1. Introduction

Essential oils (EOs) are volatile and odorous phytochemical products of plants secondary metabolism. Since ancient times, they have been widely used in folk and alternative medicines to treat pain and diverse ailments [1,2,3,4,5] as insecticides and repellents [6,7]. They are highly appreciated in cosmetics and perfumery due to their pleasant aroma [8]. The food industry also takes advantage of the properties of essential oils as additives for food flavoring [9] and food preservatives to prevent oxidation, microbial spoilage [10,11,12], and the growth of food poisoning-causing bacteria [13,14].
One of the first examples of the usage of EOs for microbial and pest control comes from the mummification process in ancient Egypt, where it was the belief that existence in the afterlife required the body to be preserved in a lifelike form [15], meaning corpse decomposition by bacteria and insects must be minimized. Analysis with gas chromatography–mass spectroscopy allowed for identifying several compounds found in embalming materials, including monoterpenes, sesquiterpenes, phenols, and naphthalene; derived from the oily materials used for the preservation of the bodies and tombs. Said compounds are present in cedar oil, pine oil, and juniper cones [16,17].
Several studies have assessed the biological properties of essential oils, being particularly interested in their activities as radical-scavenger antioxidants [18,19,20], antimicrobial [21,22], antifungal [11,23,24], and antiviral agents [18,25], as well as their sedative, analgesic, and anti-inflammatory properties [1,26,27,28]. Moreover, producing EOs in plants is a stress response, mainly to, but not limited to, deterrent herbivores [29], hence their potential as natural pesticides and repellents for crop plagues.
About a third of the global crop production is lost annually due to plagues and plant diseases [30]; additionally, the primary control for pests is synthetic pesticides, which have become less effective mainly due to plagues developing resistance to their active compounds, forcing farmers to search for alternative control. On the other hand, during the last few decades, there has been an increased concern related to the environmental damage and the potential impact on human health of exposure to synthetic pesticides; consequentially, there has been a heightened interest in the search for natural alternatives to be used in the control of plagues and pests. Several EOs and their components have been studied regarding their biocide and repellent properties against insects, arthropods, nematodes, larvae, and other plagues, with promising results [31,32,33,34,35]. However, the agricultural application of EOs faces challenges such as their components’ high volatility and oxidability.
Encapsulated systems provide EOs with protection from interactions with environmental oxygen, photodamage, evaporation, and other alterations that could compromise their biological activities and can facilitate a controlled and continuous release for long-term protection, avoiding the need for constant reapplication. Encapsulation materials, also known as matrixes, are presented in an extensive range of physicochemical characteristics, from synthetic polymers to biodegradable alternatives such as carbohydrates (chitosan, dextrins, cellulose), proteins (gelatin, albumin), gums (Arabic, cashew), and lipids (paraffin, oils, beeswax) [36]. The material selection should respond to the particular need of application and the interactions between the matrix and the encapsulated substance (core).
This article aims to give an overview of EOs as prospective agrochemicals for pest management and the different encapsulation techniques and materials that improve the applications and stability of the EOs.

2. Essential Oils

Essential oils are volatile, hydrophobic liquids made up of a mixture of secondary metabolites (monoterpenes, sesquiterpenes, phenylpropanoids, etc.) (Figure 1), as well as some components that are specific to certain EOs, such as allyl sulfides, found in garlic essential oil [37], or the characteristic indole in citrus blossoms’ volatile oils [38]. However, EOs vary in composition due to genetic causes, climate, sun exposure, rainfall, geographic location, plant age, vegetative parts, and the presence of plagues [39,40].
EOs production is restricted to a limited number of families, including Myrtaceae, Lauraceae, Rutaceae, Lamiaceae, Asteraceae, Apiaceae, Zingiberaceae, and Piperaceae [41]. EOs are produced in morphologically diverse, specialized cells present in different organs in aromatic plants [42]. These secretory tissues are found in flowers (roses), leaves (eucalyptus, mint), woods (sandalwood), barks (cinnamon), roots (valerian), rhizome (ginger), peels (citrus), and seeds (nutmeg). Around 3000 EOs are known, and 300 to be commercially important, primarily destined for flavor and fragrance markets [43,44].
EOs can be obtained by different means. The extraction method can affect the composition and quality of the EO due to the exposure of the components to temperature and pressure or their interaction with the solvent [45,46,47]. Steam distillation is the most common EO extraction method. It involves placing the vegetable material in boiling water and allowing the steam to break down plant cells, releasing the oils. Around 93% of EOs are obtained through this method, and the remaining 7% are extracted via other methods [48], including the traditional methods of enfleurage (extraction with cold fat), solvent extraction (used for labile materials), and mechanical pressing, for citrus EOs [47,49]. Newer extraction techniques such as microwave- and ultrasound-assisted extraction or supercritical fluid extraction, eliminate the annoyances of traditional methods related to time and power consumption and the decomposition of EOs during the process [44,47]; nevertheless, specialized equipment is required.
Figure 1. Common monoterpenes with reported biocidal and repellent activities. The top three rows are monoterpenoids; the Fourth row is monoterpenoid phenols and sesquiterpene; the Lower row is phenylpropanoids [31,33,50,51,52,53,54,55,56,57,58].
Figure 1. Common monoterpenes with reported biocidal and repellent activities. The top three rows are monoterpenoids; the Fourth row is monoterpenoid phenols and sesquiterpene; the Lower row is phenylpropanoids [31,33,50,51,52,53,54,55,56,57,58].
Agriculture 14 01766 g001

3. Essential Oils for Pest Management

The term pest comprises any organisms that are considered undesirable—commonly, nematodes, mites, insects, and some vertebrates such as rodents [59]. Agricultural pests include those species that cause quantitative and qualitative losses in forestry, agriculture, and stored produce and grains [60], causing waste of around a third of the annual production, translating into a major toll on the global economy. To control pests, farmers usually resort to synthetic pesticides; however, in recent years, the increasing concern for developing more sustainable and safe agricultural practices has raised attention to botanically derived pesticides.

3.1. Essential Oils as Botanical-Sourced Pesticides

Essential oils are, in general, regarded as safe for vertebrates [61], and, as mentioned before, aromatic plants have been used as traditional pest control, with their properties long studied, showing EOs’ broad spectrum of biocidal (Table 1), sublethal, and repellent activities.
One of the most employed oils is Citronella oil (Cymbopogon spp.), widely used as a mosquito repellent. It also exhibits biological activities against other organisms such as nematodes [62,63] and phytopathogenic fungi [64,65]. A 2006 study showed biocidal activity against the fennel aphid and attraction to the plague’s natural predator Cycloneda sanquinea L. (ladybug) [66].
Biocidal activities against phytophagous acari have been reported for several EOs, including those from Chamomilla recuitita, Majorana hortensis [67], Rosmarinus officinalis, Salvia officinalis [68], Salvia fruticosa, Lavandula angustifolia [69], and Senecio glaucus [70]. EOs have been effective against the two-spotted spider mite Tetranychus urticae, also Zataria multiflora, Satureja hortensis agains the strawberry spider mite Tetranychus turkestani [71], and Thymbra spicata, Origanum onites, Mentha spicata and Lavandula stoechas against the carmine spider mite (Tetranychus cinnabarinus) [72].
The effects of EOs on phytopathogenic nematodes have been reported, particularly activities against the root-knot nematodes (Meloidogyne sp.), which are vulnerable to EOs from Piper hispidinervum, which suppressed egg hatching and diminished juvenile infectivity [51], and activity against Artemisia nilagirica, which reduced the root infection and tomato plants and promoted the plant’s growth in greenhouse conditions [73]. Senecio glaucus exhibited nematostatic capacity [70]. Eucalyptus globulas, Carum capticum [74], and the mixture of EOs from Haplophyllum tuberculatum and Plectranthus cylindraceus inhibited the hatching of nematode eggs [75]. Kang and colleagues have also studied the nematocidal activity from EOs-derived phytochemicals [76]. They evaluated 97 compounds (49 monoterpenes, 17 phenylpropenes, 16 sesquiterpenes, and 15 sulfides) from essential oils against the pinewood nematode (Bursaphelenchus xylophilus) via acetylcholinesterase inhibition (AChEI), finding three active monoterpenes, two phenylpropanes, and one sesquiterpene.
Table 1. Essential oils with biocidal activities against some important agricultural pests.
Table 1. Essential oils with biocidal activities against some important agricultural pests.
OrderTargetEO Botanical FamilyEOPlant PartApplication TypeReference
AcariTetranychus cinnabarinus (carmine spider mite)LamiaceaeLavandula stoechas (lavender)Leaves and stemsFumigant[72]
LamiaceaeMentha spicata (spearmint)Leaves and stemsFumigant[72]
LamiaceaeOriganum onites (oregano)Leaves and stemsFumigant[72]
LamiaceaeThymbra spicata (thyme)Leaves and stemsFumigant[72]
ZingiberaceaeCurcuma longa (turmeric)RhizomeSpray[77]
AcariTetranychus turkestani (strawberry spider mite)LamiaceaeSatureja hortensis (summer savory)LeavesFumigant[71]
LamiaceaeZataria multifloraLeavesFumigant[71]
AcariTetranychus urticae (two-spotted spider mite)AsteraceaeChamomilla recutita (chamomile)Whole plantSpray on host[67]
LamiaceaeMajorana hortensis (marjoram)Whole plantSpray on host[67]
LamiaceaeRosmarinus officinalis (rosemary)AerialContact[68]
LamiaceaeSalvia officinalis (sage)AerialContact[68]
LamiaceaeLavandula angustifolia (lavender)LeavesFumigant[69]
LamiaceaeSalvia fruticose (Greek sage)LeavesFumigant[69]
LamiaceaeMentha spicata (spearmint)Commercial, NSFumigant[78]
LamiaceaeOcimum basilicum (basil)Commercial, NSFumigant[78]
MyrtaceaeCallistemon viminalis (callistemo)Leaves and twigsContact[79]
MyrtaceaeEucalyptus bicostata (Victorian blue gum)LeavesContact[79]
MyrtaceaeEucalyptus maidenii (Maiden’s gum)LeavesContact[79]
MyrtaceaeEucalyptus sideroxylm (red ironbark)LeavesContact[79]
MyrtaceaeEucalyptus approximans (Barren Mountain mallee)LeavesContact[79]
VerbenaceaeLippia origanoides (Colombian oregano)LeavesFumigant[80]
VerbenaceaeLippia sidoides (oregano)LeavesFumigant[81]
ColeopteraCallosobruchus chinensis (adzuki bean weevil)ApiaceaeCoriandum sativum (coriander)Industrial, NSFumigant/contact[82]
MyrtaceaeEucalyptus obliqua (eucalyptus)Industrial, NSFumigant/contact[82]
PinaceaePinus langifolia (pine)Industrial, NSFumigant/contact[82]
ColeopteraCallosobruchus maculatus (cowpea weevil)LamiaceaeRosmarinus officinalis (rosemary)AerialFumigant[83]
LamiaceaeMentha piperita (peppermint)AerialFumigant[83]
RutaceaeCitrus sinensis (sweet orange)PeelFumigant[84]
ColeopteraLasioderma serricone (cigarette beetle)CruciferaeCocholeria armoracia (horseradish)Commercial, NSFumigant[85]
LauraceaeCinnamomum cassia (cinnamon)Commercial, NSContact[85]
ColeopteraSitophilus granarius
(granary weevil)
LauraceaeCinnamomum zeylanicum (cinammon)Commercial, NSContact[54]
MyrtaceaeSyzygium aromaticum (clove)Commercial, NSContact[54]
ColeopteraUlomoides dermestoides
(peanut beetle)
PoaceaeCymbopogon citratus
(lemongrass)
Commercial, NSContact[86]
DipteraCeratitis capitata (Mediterranean fruit flyLamiaceaeRosmarinus officinalis (rosemary)LeavesFumigant/Topical[87]
LamiaceaeLavandula angustifolia (lavender)LeavesFumigant/Topical[87]
DipteraDrosophila melanogaster (fruit fly)LamiaceaeMentha pulegium (pennyroyal)NSContact[31]
LamiaceaeMentha spicata (spearmint)NSContact[31]
DipteraDrosophila suzukii (spotted wing drosophila)LamiaceaeMentha piperita (peppermint)Aerial floweringFumigant[88]
LamiaceaePerilla frutescens (perilla)LeavesFumigant[88]
HemipteraAphis forbesi
(strawberry aphid)
FabaceaeTephrosia vogelii
(Vogel’s tephrosia)
FlowersSpray[89]
HemipteraBemisia tabaci (silverleaf whitefly)AmaryllidaceaeAllium sativumNSFumigant/contact[90]
LamiaceaeMicromeria fruticosa (white savory)AerialFumigant[91]
LamiaceaeNepeta racemose (dwarf catnip)AerialFumigant[91]
LamiaceaeOriganum vulgare (oregano)AerialFumigant[91]
LamiaceaeThymus vulgaris (thyme)LeavesContact[92]
RutaceaeCitrus aurantium (bitter orange)PeelFumigant[93]
HemipteraHyadaphis foeniculi (fly honeysuckle aphid, fennel aphid)LamiaceaeHyptis suaveolens (alfazema)NSTopical[66]
HemipteraMyzus persicae (green peach aphid)ApiaceaeCuminum cyminum (cumin)SchizocarpSpray on host[94]
AsteraceaeSantolina chamaecyparissus (cotton lavender)AerialSpray[95]
AsteraceaeAchillea millefolium (yarrow)AerialSpray[96]
CannabaceaeCannabis sativa (hemp)InflorescencesSpray on host[96]
LamiaceaeMentha pulegium
(pennyroyal)
AerialFumigant[97]
LamiaceaeMentha pulegium
(pennyroyal)
AerialSpray[98]
LamiaceaeOriganum majorana (marjoram)AerialSpray[98]
LamiaceaeMelissa officinalis (lemon balm)AerialSpray[98]
HemipteraRhopolasiphum maidis (corn leaf aphid)MyrtaceaeSyzygium aromaticum (clove)BudsFumigant[99]
HemipteraTrialeurodes vaporariorum (greenhouse whitefly)AsteraceaeEupatorium buniifolium (chilca)NSSpray on host[100]
HemipteraTrialeurodes vaporariorum (greenhouse whitefly)AtherospermataceaeLaurelia sempervirens (Chilean laurel)LeavesFumigant[101]
HymenopteraAcromyrmex balzani (leaf-cutter ant)LamiaceaeEplingiella fruticosaLeavesFumigant[102]
MyrtaceaeMyrcia lundianaLeavesFumigant[103]
LepidopteraMediterranean flour mothLamiaceaeOriganum onites L. (oregano)LeavesFumigant[104]
LamiaceaeSatureja thymbra L. (savory)LeavesFumigant[104]
LepidopteraIndian meal mothLamiaceaeOriganum onites L. (oregano)LeavesFumigant[104]
LamiaceaeSatureja thymbra L. (savory)LeavesFumigant[104]
LepidopteraSpodoptera littoralis (cotton leafworm)LamiaceaeThymus algeriensis
(Thyme)
AerialFumigant[35]
ApiaceaePimpinella anisum (anise)SchizocarpTopical[94]
ApiaceaeCrithmum maritimum (sea fennel)Seeds/aerealTopical[105]
EuphorbiaceaeRicinus communis (castor bean)Commercia, NSFumigant[106]
LamiaceaeOcimun gratissimum (white wild basil)AerialTopical[107]
OrthopteraSchistocerca gregaria (desert locust)AmaryllidaceaeAllium cepa (onion)LeavesTopical[108]
ApiaceaePetroselinum sativum (parsley)SeedsTopical[108]
GeraniaceaePelargonium radula (geranium)Whole plantTopical[108]
AmaryllidaceaeAllium sativum (garlic)Commercial, NSSpray[109]
NS = Not specified.

3.2. Mechanism of Action of Essential Oils as Insecticides

Although the biocidal mechanism of essential oils and their components is still unclear, the rapid action of EOs on insects and behavioral patterns observed in treated individuals have pointed to a neurotoxic effect [110]. Moreover, due to the observation of synergic and antagonist effects of binary mixtures of essential oils and the divergence of biological activities exhibited by one essential oil and its individual components, it has been proposed that there are a handful of different mechanisms of action for insecticidal activity. Some of the most studied proposed mechanisms are inhibiting the enzyme acetylcholinesterase, modifying gamma-aminobutyric acid (GABA) receptors, and activating octopamine receptors.

3.2.1. Inhibition of Enzyme Acetylcholinesterase (AChE)

Acetylcholinesterase is an essential neuronal enzyme in the cholinergic synapses and neuromuscular junction for vertebrates and invertebrates [111,112]. The extended use of organophosphate and carbamate pesticides, both targeting the cholinergic system, has led to the rapid development of pest resistance [113,114] and increasing interest in botanical pesticides.
Several studies have been conducted to determine the potential of EOs as AChE inhibitors (AChEI); Czerniewicz and colleagues (2018) [95] reported an in vitro reduction in AChE activity after the individual application of EOs from the Asteraceae family, particularly the EOs from Santolina chamaecyparissus and Achillea millefolium. Other EOs that have exhibited AChEI activity include those from the Anthriscus nemerosa root [115], the Citrus aurantium peel [93], the Citrus sinensis peel [84], Echinacea purpurea, Thymus praecox [116], Lippia origanoides [80], Ocimum tenuiflorum [117], Salvia officinalis [118,119], Salvia lavanduleifolia [120], and Rosmarinus officinalis [121].
Individual components of EOs activity as AChE have also been assessed; monoterpenic components have displayed strong inhibitory activities against the enzyme, including (+)-, (−)-α-pinene, β-pinene, limonene, β-phellandrene, Fenchone, S-carvone, L-carvone, linalool, 2-carene, eugenol, pulegone, 1,8-cineole, and 4-terpineol [50,55,76,84,93,117,122,123].
Most studies have used non-invertebrate-sourced AChE for the in vitro assays. Consequently, whether EOs and their components perform as AChEI in insects has not yet been determined.

3.2.2. Modification of GABA Receptors

The amino acid GABA acts as a neurotransmitter in invertebrates. Its receptor comprises multiple binding sites, the most studied GABA-gated chloride channel targeted by many insecticides [124,125,126].
There is minimal information regarding the interaction of EOs with GABA receptors. Nevertheless, some studies show the activity of monoterpenic compounds on this molecular target. Thymol has been reported as an allosteric modulator of the GABA receptors in mammals and insects [127,128,129,130,131]. Tong and Coats [129] also reported carvacrol and pulegone as positive allosteric modulators at insect GABA receptors, as they significantly increased the 36Cl uptake induced by GABA in arthropods. In another study, linalool, with a synergic effect with methyl eugenol, showed good potential as a GABA agonist and allosteric modulator [58]. Abdelgaleil and colleagues [55] evaluated six natural monoterpenes (1,8-cineole, (−)-citronellal, limonene, α-pinene, pulegone, and 4-terpineol) against GABA-T, with limonene displaying the highest activity, with an IC50 of 11.37 mg/L. In an in silico evaluation conducted by Toledo and colleagues [99], the main components of clove essential oil, eugenol, and β-caryophyllene were found to bind to three molecular targets, including the GABA receptors in the phytophagous aphid Rhopalosiphum maidis. Although these are promising results, the effect of EOs on GABA receptors needs further study.

3.2.3. Interference on Octopamine Receptors

Octopamine is a biogenic amine functionally related to noradrenaline. Invertebrates act as neurotransmitters, neuromodulators, neurohormones, and neuromuscular transmitters and play a part in the regulation of physiological and biological processes, including heartbeat regulation, locomotion, reproduction, and memory [110,132,133].
In 1980, Livingston and colleagues [134] reported that octopamine injections produced behavioral and locomotory effects in lobsters, with a sustained extension of the limbs and abdomen. Enan (2001) [110] described the same effect of the hyperextension of legs and the abdomen on American cockroaches treated with the essential oil components eugenol, α-terpineol, and cinnamic alcohol, reporting immobilization and, lastly, death. Several more studies have attributed EOs and their components’ pesticidal activity to a possible competitive activation of octopaminergic receptors [111,135,136]. The role of octopaminergic interference regarding invertebrates has also been discussed, with some reports linking it to reproductive sterility and feeding behavior changes [133,137,138]. These multi-action mechanisms and octopamine’s lack of functionality in vertebrates make octopamine antagonists effective and selective pesticides [132].

3.3. Repellency and Sublethal Activities

Other widely studied biological activities of EOs that are highly important are their repellency and sublethal effects on pests; the latter are associated with their ability to survive, including locomotion, growth, and reproduction.

3.3.1. Repellency

The repellent properties of EOs against arthropods of public health importance have been widely reported, and the use of these products represents an effective and safe alternative against disease vector organisms worldwide [139,140,141,142,143,144,145]. These findings have progressively been extended to pests of agricultural interest [146,147].
Lacotte and colleagues (2023) [148] evaluated the repellent activity of forty plant essential oils against the pea aphid (Acyrthosiphon pusim). Peppermint oil showed the highest activity; these results coincide with the effect shown on Aphis craccivora [149].
Similarly, the repellent properties of Eos against domestic pest vectors of diseases have been widely described [32,150,151,152,153,154,155], suggesting their potential application as agricultural repellents. EOs from fruits from the genus Zanthoxylu, as well as the Brazilian pepper fruit (Schinus terebinthifolius), display repellency against the whitefly Bemisia tabaci [156,157]; EOs from Artemisia princeps and Cinnamomum camphora and their mixture were repellent against the store grain beetles Sitophillus oryzae and Bruchus rufimanus, with the binary mixture showing significantly higher activity [158]; EOs from red mahogany (Eucalyptus resinifera), star anise (Illicium verum) and croton (Croton anisatum), as well species from the genera Satureja and Zingiber officinale, and the Tunisian endemic plant Ferula tunetana have also exhibited repellent capacity against agricultural pests of high importance [159,160,161,162,163]. Isolated compounds have also been studied for their repellent profile; in 2017, Tak and Isman [164] tested twenty EO-occurring terpenoids against the two-spotted mite, Tetranychus urticae, including carvacrol, thymol, trans-cinnamaldehyde, and α-terpineol.

3.3.2. Antifeedant

Antifeedants deter the insect’s food consumption through feeding behavioral modifications [165]. Some of the most documented antifeedants are triterpenoids and alkaloids. However, some essential oils have displayed a deterrent capacity against agricultural pests. Sousa and colleagues (2013) [166] reported antifeedant activity from EOs of the Apiaceae members A. graveolens and P. crispum against the armyworm’s larvae (Pseudaletia unipuncta), with inhibitions above 70%. In two independent studies, Piper species were reported to be strong antifeedants against some of the most important pests in agriculture (Myzus persicae, Rhopalosiphum padi, and Spodoptera littoralis) [51,167], with P. dilatatum exhibiting the highest activity, with feeding inhibition percentages of 84.4, 98.9, and 65.3 for M. persicae, R. padi, and S. littoralis, respectively [167]. For S. littoralis, EOs from castor and camphor decreased total proteins and carbohydrates in treated larvae, changing their nutritional status [106], and another report showed that EOs from Scutellaria hastifolia have antifeedant activity at 36 ppm, displaying a good deterrent profile [168].

3.3.3. Effects on Reproductive Conduct

Reproduction control is another important tool for pest management. Some reports indicate that essential oils harm insects’ reproduction capacity, acting as oviposition-deterrents. The anti-ovipositor effect of EOs on human disease vector mosquitos has been described in a handful of reports, with EOs from Cinnamomum zeylanicum, Alpinia purpurata, Zanthoxylum limonella, and Sphaeranthus amaranthoides, among others [155,169,170,171]. As for activity on pests of agricultural importance, in a study conducted on four citrus peel oils on the cowpea weevil (Callosobruchus maculatus), the oviposition was inversely proportional to the oil concentration, with percentages of reduction from 29.74 to 71.66%, as well as a significant reduction in emergence from laid eggs [141]. Likewise, when treated with the monoterpenoids E-anethole, estragole, S-carvone, linalool, L-fenchone, geraniol, γ-terpinene, and D-, L-camphor, mated females did not lay eggs, contrary to the control females [172]. Linalool also has shown oviposition-deterrent activity on the significant Mediterranean fruit fly, Ceratitis capitata, in a concentration-independent relationship [173].
Laborda and colleagues (2013) [68] describe ovipositor-deterrent and anti-hatching activities from Salvia officinalis and Rosmarinus officinalis against Tetranychus urticae at concentrations of 0.15–0.25% and after eight days of treatment and at least 50% fewer larvae emerging from eggs in comparison with the control individuals.
Finally, Dos Santos Silva and colleagues (2016) [174] studied citronella oil’s (Cymbopogon winterianus) influence on the reproductive conduct of the fall armyworm (Spodoptera frugiperda), reporting diminished reproduction, with a 90% oviposition reduction, and the non-viability of the eggs derived from treated individuals.

3.4. Non-Target Safety

Although botanical pesticides could be considered a safer alternative, their potential effects on non-target organisms need to be evaluated. Some research groups have reported on the ecotoxicological selectivity of some EOs.
Most of the studies of the effect of EOs on non-target organisms have focused on pollinators such as honeybees, due to their biological and economic importance. Several EOs have been shown to be safe for bees with low mortality rates [175,176] or dose-dependent safe [177], positioning themselves as more environmentally friendly alternatives to non-selective pesticides, not only for agriculture but useful also as a household insecticide and even in beekeeping to control parasitic species.
A study by Sabahi and colleagues (2018) [178] evaluated the effect of lemongrass (Cymbopogon citratus) oil on one of the main parasites of honeybees in the northern United States, the Varroa destructor mite; a strong acaricidal effect on the parasite, which was safe for bees, was observed. Similar results were shown using clove oil against the Varroa jacobsoni mite. In another study, EOs of the Asteraceae family showed no acute toxicity to bees at the concentrations needed to control whiteflies and no effect on the tomato crop yield under greenhouse conditions [100].
Assessing the safety of pesticides against bees is an essential factor to consider in the quest for sustainable agriculture; however, the effect on other non-target organisms needs to be explored.
A study evaluating the effect of citronella [66] and clove oil (Syzygium aromaticum) showed high toxicity against aphid pests but showed no effect on the predatory ladybug [99]. The acute toxicity of EOs for earthworms has been evaluated in several studies, finding high tolerance by these organisms, with adverse effects sometimes shown at doses above 1000 mg [179,180]. Benelli and col. (2019) [107] also reported high earthworm tolerance to basil EO, which incidentally contained a major component of thymol, which is relatively safe for some non-target organisms, including bees and mealworm beetles [131,181,182,183].
Some studies have proposed the need to expand the evaluation of the effects of EOs on organisms used as a biological pest control (polyphagous predators) and soil invertebrates. In a study evaluating the susceptibility of the whitefly (Bemisia tabaci) and its predator (Orius albidipennis) to vapors of the EO of Artemisia sieberi Besser, Pelargonium roseum, and Ferula gummosa, it was observed that the LC50 was 10 times higher for the predator than for the target organism. Other predatory organisms that have been studied regarding the effect of EOs include Nesidiocoris tenuis, a biological control of several tomato pests.
As active as they are, the volatility of EOs discourages their use as pesticides. The constitution of essential oils makes their application very difficult; in recent years, the controlled release of EOs by microencapsulation and nanoencapsulation has been studied and tested, with interesting results.

4. Encapsulation of EOs

As pointed out, EOs comprise highly volatile and oxidable compounds, one of the main restraints against the widespread use of these highly bioactive compounds as pesticides [182].
Micro- (1–1000 µm) and nano-encapsulated (1–100 nm) systems are a way to protect EOs from oxidation and volatilization by their retention in a matrix, allowing for controlled release and the preservation of the qualitative characteristics.
When selecting the matrix, there are some parameters to be considered to achieve successful encapsulation; among others, there is the strength of the interaction between the matrix and the EO, the desired release rate, and the necessities and legislation in the field of application—for example, selecting matrix materials that are allowed in agriculture or organic agriculture could significantly reduce the options.

4.1. Materials and Techniques for the Encapsulation of EOs

Nowadays, a handful of nanomaterials suit different needs and concerns. For essential oils, the material and structure of the encapsulation matrix and their interaction with the core material are responsible for the protection of the bioactive compounds and enable a sustained release of the volatile compounds, which is also dependent on the selected encapsulation technique, the size of the particles, and their morphology and affects the release trigger [184,185]. The release rate, incidentally, can also significantly affect the bioactivity of the EOs, as reported by several authors (Table 2). In the development of EOs products for agricultural use, it is necessary to evaluate the release profile of the active compounds. It has been observed that several factors can alter this release; parameters such as pH, relative humidity, and temperature have been related to the degradation or alteration of the carrier materials [186]. The control of release has also been related to the encapsulation technique, properties, and proportions of the carrier material [187].
Some preferred techniques for generating EOs nanoparticles are emulsification, complex coacervation, and ionic gelation.

4.1.1. Emulsification

The hydrophobic character of essential oils can limit their potential application in biological systems. To favor water solubility and at the same time preserve the biological activity of their active compounds, the nanoemulsion formulation represents an alternative for suitable EO delivery systems [28,188].
Nanoemulsions are nanosized isotropic systems of two immiscible liquids, oily and aqueous, combined in a single, stable formulation [22,189]. Various methods can carry out the generation of nanoemulsions; the most commonly used are (a) high-pressure homogenization, which is a high-energy mechanical method and one of the most efficient methods for producing stable o/w nanoemulsions [190,191]; (b) micro-fluidization, which uses high pressure to force coarse emulsions through microchannels to reduce the size of the particle [191,192]; and (c) the phase-inversion method (Figure 2), which refers to a phenomenon that occurs when an agitated o/w emulsion reverts to a w/o emulsion [193]. This latter method is a low-energy method that produces small particles while requiring no specialized equipment [194,195].
EO-nanoemulsions per se are mainly used in the food industry as preservatives to avoid microbial spoilage, this being one of the leading applications. They are especially useful in the food regions with higher water activities or the liquid–solid interfaces, where microbial growth is favored [196,197,198]. However, nanoemulsions can also be necessary in generating other kinds of EO nanosystems, like nanogels or nanocapsules, which can result in more suitable applications in different areas, including agriculture and medicine. As described below, ionic gelation and complex coacervation techniques observe an emulsion’s preparation stage when encapsulating lipophilic cores.

4.1.2. Ionic Gelation

Ionic gelation is a relatively facile, solvent-free technique based on ionic interactions between oppositely charged groups: a cationic polymer, such as chitosan and sodium alginate, and an anionic molecule, such as tripolyphosphate (TPP), the most widely used non-toxic, crosslinker [199,200].
The overall methodology consists of the cation generation, preparation of the emulsion, the complexation, and the precipitation. For the first of these, the polysaccharide is dissolved in an aqueous acidic solution, followed by adding the surfactant under stirring to ensure proper homogenization; then, the EOs are added, still under homogenization. The polysaccharide and crosslinker solutions are combined dropwise under continuous stirring [191,201,202,203,204], allowing for the ionic gelation and precipitation of forming particles (Figure 3). Centrifugation is then applied to separate the particles from the unreacted compounds, and finally, the particles can be heat-dried [205,206] or freeze-dried [207] and characterized.
It is essential to keep in mind that the proportion of polymer/acid in the solution [199], the selection of the surfactant [191], the polymer/crosslinker ratio [202], as well as the EO content [207] can play a part in the entrapment efficiency (EE) and particle size. Further studies are required to determine adequate conditions for specific requirements.

4.1.3. Complex Coacervation

Complex coacervation is a technique based on the electrostatic interaction between oppositely charged polymers, usually a protein and a polysaccharide [208] present in two separate immiscible phases—one that is polymer-rich, used to coat the particle, and another that is polymer-poor [208,209]. Coacervation presents a high encapsulation efficiency, requires small amounts of wall material, and can be implemented with a variety of biopolymers as matrixes [210], including gelatin-gum Arabic [161,208,211,212,213], gelatin-sodium alginate [210,214,215], gelatin-seed mucilage [209,216], and chitosan-cashew gum [217,218], along with several other proteins of animal and plant origins.
The complex coacervation involves four stages (Figure 4): (i) the preparation of the aqueous solution of the polymers; (ii) the stable emulsification of the hydrophobic phase to the aqueous solution of one polymer, which in most cases is the protein; (iii) the coacervation step, by a change in pH and temperature; and (iv) the hardening of the polymer shell through a chemical crosslinker or elevated temperature [219,220].
The complex coacervation technique is suitable for fabricating NPs or MPs with lipid cores with increased miscibility, deliverability, and the sustained release of the ingredients [221].
Table 2. Matrixes for the encapsulation of essential oils.
Table 2. Matrixes for the encapsulation of essential oils.
MaterialEncapsulated EOMethodParticle TypePart. SizeEEApplicationObservationsReferences
Biopolymers
ChitosanCitronella
(Cymbopogon spp.)
Ionic gelationMicrocapsules11–225 µm94.7–98.2%Sustained release[202]
Satureja hortensisIonic gelationMicrocapsules192 nm96.17%AcaricidalSustained release
Prolonged bioactivity
[204]
Citrus oilsIonic gelationMicrocapsules289.3–8843.2 nm61.9–68.1%Good release rates.
Particle size and EE dependent on the surfactant.
[191]
Coriandrum sativumIonic gelationNanocapsules50–80 nm26.5–75.99%Storage product preservationControlled release
Enhanced bioactivity
[222]
Eugenia caryophyllataIonic gelationNanocapsules40–100 nm31–45.77%AntifungalControlled release
Higher bioactivity
[223]
Cinnamomum zeylanicumIonic gelationMicrocapsules100–200 nm88.6%AntimicrobialProlonged stability
Improved antimicrobial capacity
[224]
Mentha piperitaIonic gelationMicrocapsules≤563.3 nm64–70%Stored food pest controlImproved AChEI and fumigant toxicity[225]
Piper nigrumIonic gelationMicrocapsulesAv. 527.5 nm35–40%Stored food pest controlImproved AChEI and insecticidal activity[226]
Eryngium campestreIonic gelationMicrocapsulesAv. 157.8 nm61–75%Storage product preservationImproved bioactivity[227]
Chitosan-Caffeic acidCuminum cyminumEO in nanogelNanogel≤100 nm85%AntifungalImproved antifungal performance by sustained release[228]
Chitosan–cashew gumLippia siodesComplex coacervationNanogel335–558 nm70%InsecticidalControlled release
Improvement in bioactivity
[217]
Chitosan-Cinnamic acidMentha piperitaEO in nanogelNanogel≤100 nmAntifungalImproved antifungal performance by sustained release[229]
Bunium persicumEO in nanogelNanogel≤100 nm41–43%Antifungal and antioxidantEnhanced antifungal and antioxidant activities[230]
β-cyclodextrinCommercial thyme oil (Thymus spp.)Complexation (kneading)Complex582.9 nm82.55 g/100 gAntimicrobialEnhanced antimicrobial activity[231]
Complexation (freeze drying)3226.7 nm71.27 g/100 g
Lippia berlanderiComplexationComplex71%AntimicrobialHigh entrapment
Increased stability
[232]
Eugenia caryophyllata61%
β-cyclodextrin (HP-β-CD)Piper nigrumComplexation (inclusion complex formation)Complex50.55%AntioxidantIncreased stability
Higher antioxidant activity
[233]
β- & γ-cyclodextrinLippia graveolens (thymol and carvacrol chemotypes)ComplexationComplex6–13.22 µm7.4–63.4%Sustained release[234]
CornstarchThymus vulgarisThermoplastic extrusionParticles0.01–0.5 cm69.07%RepellentProlonged residual effect.
Improved bioactivity.
[235]
ZeinSyzygium aromaticumAntisolvent precipitationNanoparticles125 nm91.4%InsecticideLow toxicity to a non-target organism, Caenorhaabditis. elegans
Potentialized the bioinsectide activity against Drosophila melanogaster
[236]
Gelatin-gum Arabic (Glutaraldehyde as wall hardener)Camphor oil with added oil-soluble polystyreneComplex coacervationParticles85.7–299.7 µm99.6%Sustained release[211]
Gelatin-chia mucilageOregano oil (Lippia spp.)Complex coacervation/spray dryMicrocapsules1.65–8.85 µm82.15–95.6%Enhanced bioactivity
Particle size-dependent EE
[209]
Gelatin-Persian gum, Persian gum, and gum ArabicSatureja hortensisComplex coacervationNanocapsules81–208 nm72.1–92.8%HerbicidalImprovement in herbicidal activity on encapsulated oil[161]
Gelatin-sodium alginateCymbopogon winterianusComplex coacervationComplex434.06 µm83.5%Controlled release
Usage of waste product gelatin
[215]
Piper nigrumComplex coacervationComplex49.13–82.36%Good retention capacity and preservation of composition[210]
Maltodextrin-gum ArabicSchinus molleSpray dryingMicroparticles0.2–40 µm96–100%InsecticidalSlow release, over 366 h
Prolonged insecticidal effect
[237]
Maltodextrin-caseinThymus vulgarisSpray dryingMicroparticlesAv. 0.87 µm88.9%Food preservationThermal stability[238]
Sodium alginate (crosslinker CaCl2)Satureja hortensisIonic gelationMicroparticles47–117 µm52–66%Antibacterial
Antioxidant
Controlled release
Improved bioactivity
[207]
Whey protein- mesquite gumSalvia hispanicaSpray dryingMicrocapsules2.54–3.35 µm71.4–80.7%Good encapsulation efficiency[239]
Whey protein- gum Arabic2.33–3.09 µm70.7–79.8%
Emulsions
o/w
Tween 80 as surfactant
Eucalyptus spp.SonicationDropletsAv. 3.8 nmAntimicrobial (topical)Improvement in wound-healing effectivity[240]
o/w
Tween 20/Triton x−100 as surfactant
Azadiractha indica- Cymbopogon nardusConstant stirDroplets2.8–17.8 nmAntifungalStability of components
Potent antifungal activity
[64]
Isopropanol
Tween 80 as surfactant
Curcuma longaSonicationMicroemulsionAcaricidalControlled release
Improved acaricidal activity
[77]
o/w/o
Palm oil
Sunflower oil
Soy lecithin as surfactant
Terpene mixtureHigh-pressure homogenizationDroplets75–175 nmAntimicrobialEnhanced activity[196]
o/w/ polysorbateCrithmum maritimumEmulsion phase inversionDroplets50–70 nmAgriculture and domestic pest controlIncreased anti-ovipository and toxic activity[241]
Octenyl succinic anhydride (OSA)—starchRosmarinus officinalis
Zataria multiflora
Spray dryingMicrocapsules461–854 nmInsecticidalMore effective than non-formulated in a long time[242]
Chitosan-cellulose nanofibersCitronella essential oilUltrasonicationNanoparticles426.9 nm90.8%InsecticidalNano-systems increased the insecticidal activity [243]
Nanocarriers
SiO2Crithmum maritimumLoading of hollow capsuleNanocapsules20–78 nmAgriculture and domestic pest controlIncreased anti-ovipository and toxic activity[241]
Zein sta-bilized with Tween 80Eucalyptus staigeriana
Litsea cubeba
Ultrasound-assisted nano-precipitationNanoparticles200 nmAntifungalActivity against Colletotrichum lindemuthianum[244]
Part. Size = Particle size; Av.= Average size; o/w = oil/water; o/w/o = oil, water, oil.

4.2. Prospects for Application in Agriculture

In their application in agriculture, EOs required a sustained release rate determined by their bioactivity efficiency as the minimum rate and their phytotoxicity as the maximum [187]. Moreover, environmental legislation limits the use of certain materials in farming and agriculture [245] due to concerns about their environmental and biological impact.
Biodegradable biopolymers are viable for applications in fields, greenhouses, and storage, mainly maintaining the integrity of the active compounds in EOs.
Natural carrier materials like polysaccharides have been well documented [246,247,248,249,250,251,252,253,254]. Other materials that have also shown effective controlled release include protein- and lipid-based encapsulations. Recently, various studies have researched material conjugation during the encapsulation process to enhance properties [255,256,257]. For example, electrostatic attractions in polysaccharide–protein systems provide greater protection to active ingredients [257] These systems have also shown improvements in encapsulation efficiency, solubility, and stability [139,258,259].

4.2.1. Chitosan

Chitosan (Ch) is a biopolymer product of the deacetylation of chitin from marine waste and insects; it is low-cost, highly biodegradable, and biocompatible and presents low toxicity [246,247,248].
This biopolymer was largely mentioned in Table 2 as a candidate for the encapsulation of EOs with pest management purposes. As stated by Ahmadi and colleagues [204], the obtention of ChNP via ionic gelation using TPP as a crosslinker results in a high EE, more than 96%, and a prolonged release for at least 25 days, translating into prolonged acaricidal activity with longer residual effects. Das made a similar observation [222]; he reported an initial burst (46.86% in the first eight hours) followed by a slow release of EO and an enhanced inhibition of store grain fungus. In another report, peppermint oil-chitosan nano- and microcapsules displayed higher toxicity against two stored-product pests, with almost 50% lower LC50 and a similar release behavior: an initial burst followed by a slow release rate. However, 38.5% to 62.4% of the EO was released in the initial burst, and the assays were carried out for only 72 h [225]. Rajkumar (2020) also reported the enhanced bioactivity of black pepper oil encapsulated in chitosan compared to free oil against stored-product pests, noting a decrease in the LC50 values [226].
Nanogels based on chitosan and cashew gum, obtained by complex coacervation, exhibited an EE of 70%, with a release rate dependent on the cashew gum content in the matrix: increased concentrations of the gum appear to favor the nanogel’s hydrophilic character, promoting a faster release of the EO [217].
Chitosan remains the right matrix candidate for EONPs for agricultural purposes. However, it is relevant to note that, even for different oils, release rates seem to follow a pattern of an initial burst and a posterior slow release, which may not suit the application’s specific needs and requires further experimentation to determine more adequate conditions for the desired release.

4.2.2. Gelatin-Gum Arabic

The gelatin-gum Arabic complexes profile is a promising candidate for encapsulating EO for agricultural applications. Both polymers are biodegradable, and the combination of polysaccharide and protein as well as the use of wall-hardening crosslinkers, commonly glutaraldehyde or transglutaminase, confer certain thermostability [212,249,250], a desirable quality for open field applications; high EE, up to 99.6% [198], and good release rates exist under colder and drier conditions [250,251,252].
Several factors can affect this matrix’s efficiency, as reported by Lv [250]. pH played a role in the size and thermal stability of gelatin-gum Arabic/ jasmine EO nanocapsules, obtained via complex coacervation using transglutaminase as a wall-hardening crosslinker. Incidentally, they displayed relative thermal protection when exposed to a water bath at 80 °C, where wall material keeps the core compounds’ integrity for up to 5 h.
The selection of crosslinkers was reported to affect the release control, as the chemical treatment with glutaraldehyde favors retention the most, in contrast with its enzymatic counterpart, transglutaminase [251]. Prata also reported a lower retention rate under wet conditions due to wall swelling, which was not observed under dry conditions.
Further experiments are needed to achieve a higher thermostability and controlled release; however, the protein–polysaccharide matrix remains a good prospect for EO encapsulation.

4.2.3. Cyclodextrins

Cyclodextrins (CDs) are toroidal-shaped natural oligosaccharides with hydrophobic cavities and hydrophilic exteriors, which makes them suitable for the encapsulation of lipophilic cores via complex inclusion [232,253,254]. Among CDs, the seven-sugar unit, β-cyclodextrin, is most widely used due to its simple synthesis and availability [253].
EO-CDs nanosystems have been studied and reported, positioning this matrix as an interesting alternative; some authors have observed improved thermal stability of the labile components of Eos [260], good EE [232,254,261], sustained release [234,253], and enhanced biological activity [233,261].
However, cyclodextrins are a more expensive alternative than chitosan and gelatin, which could affect the large-scale production of EO-CD NPs.

5. Conclusions

The biological activities of ECs are well documented, and there are several reports on their pest control properties. Recent efforts to implement more sustainable agricultural practices have drawn attention to EOs as potential pesticides and repellents of biological origin. However, the main challenge for their application as agrochemicals is the deterioration and high volatility of these compounds.
As nanotechnology is developed in diverse fields, the agrochemical industry is starting to rely on micro- and nanosystems as potential vehicles for bioactive compounds. In vitro, assays of nanosystems of essential oils display interesting release profiles, which translate into higher and prolonged biological activities. Encapsulating materials such as chitosan, gelatin-gum systems, and cyclodextrins are some of the most effective matrixes for EO cores. Other materials that have also shown effective controlled release include protein- and lipid-based encapsulations. Recently, to enhance properties, various studies have researched material conjugation during the encapsulation process, providing them with protection and good release kinetics while being environmentally friendly options due to their biodegradability and low toxicity.
The sub-microscopic encapsulation of EOs is still in need of further study. Several factors could influence the success of the synthesis of nanosystems, including the com-position of the oil, which varies in function of several factors and determines the strength of the interaction with the matrix; it is also important to set realistic release rates and release triggers that work for the specific application field (open field, greenhouse, food storage) and carry out the correspondent ecotoxicological evaluations.
In conclusion, although encapsulation technology is the most recent approach in essential oils, where the generation of biodegradable micro- and nanosystems of EOs is still a developing technology, it is an interesting topic for researchers for literature studies, as they appear to be a viable and more environmentally friendly alternative to conventional agrochemicals, with low soil leaching, more balanced doses, and generally low toxicity to non-target and vertebrate organisms.

Author Contributions

Conceptualization, I.C.-G., L.D.-R. and R.A.-G.; writing—original draft preparation, R.A.-G.; writing—review and editing, L.D.-R., M.M.-S. and M.d.P.H.-V.; supervision, I.C.-G.; funding acquisition, I.C.-G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study.

Acknowledgments

We thank the Facultad de Ciencias Químicas e Ingeniería, Universidad Autónoma de Baja California, for the support provided for this project, as well as the infrastructure. Additionally, we acknowledge the Consejo Nacional de Ciencia y Tecnología (CONACYT) for the grant (CVU: 395141).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Buchbauer, G.; Jirovetz, L.; Jäger, W. Aromatherapy: Evidence for Sedative Effects of the Essential Oil of Lavender after Inhalation. Z. Für. Naturforschung C 1991, 46, 1067–1072. [Google Scholar] [CrossRef] [PubMed]
  2. Woronuk, G.; Demissie, Z.; Rheault, M.; Mahmoud, S. Biosynthesis and Therapeutic Properties of Lavandula Essential Oil Constituents. Planta Med. 2011, 77, 7–15. [Google Scholar] [CrossRef] [PubMed]
  3. Ali, B.; Al-Wabel, N.A.; Shams, S.; Ahamad, A.; Khan, S.A.; Anwar, F. Essential Oils Used in Aromatherapy: A Systemic Review. Asian Pac. J. Trop. Biomed. 2015, 5, 601–611. [Google Scholar] [CrossRef]
  4. Plant, R.M.; Dinh, L.; Argo, S.; Shah, M. The Essentials of Essential Oils. Adv. Pediatr. 2019, 66, 111–122. [Google Scholar] [CrossRef]
  5. Yuan, R.; Zhang, D.; Yang, J.; Wu, Z.; Luo, C.; Han, L.; Yang, F.; Lin, J.; Yang, M. Review of Aromatherapy Essential Oils and Their Mechanism of Action against Migraines. J. Ethnopharmacol. 2021, 265, 113326. [Google Scholar] [CrossRef] [PubMed]
  6. Burfield, T.; Reekie, S. Mosquitoes, Malaria and Essential Oils. Int. J. Aromather. 2005, 15, 30–41. [Google Scholar] [CrossRef]
  7. Tisgratog, R.; Sanguanpong, U.; Grieco, J.P.; Ngoen-Kluan, R.; Chareonviriyaphap, T. Plants Traditionally Used as Mosquito Repellents and the Implication for Their Use in Vector Control. Acta Trop. 2016, 157, 136–144. [Google Scholar] [CrossRef] [PubMed]
  8. Gallucci, S.; Neto, A.P.; Porto, C.; Barbizan, D.; Costa, I.; Marques, K.; Benevides, P.; Figueiredo, R. Essential Oil of Eugenia uniflora L.: An Industrial Perfumery Approach. J. Essent. Oil Res. 2010, 22, 176–179. [Google Scholar] [CrossRef]
  9. Smith, R.L.; Cohen, S.M.; Doull, J.; Feron, V.J.; Goodman, J.I.; Marnett, L.J.; Portoghese, P.S.; Waddell, W.J.; Wagner, B.M.; Hall, R.L.; et al. A Procedure for the Safety Evaluation of Natural Flavor Complexes Used as Ingredients in Food: Essential Oils. Food Chem. Toxicol. 2005, 43, 345–363. [Google Scholar] [CrossRef] [PubMed]
  10. Khorshidian, N.; Yousefi, M.; Khanniri, E.; Mortazavian, A.M. Potential Application of Essential Oils as Antimicrobial Preservatives in Cheese. Innov. Food Sci. Emerg. Technol. 2018, 45, 62–72. [Google Scholar] [CrossRef]
  11. Philippe, S.; Souaïbou, F.; Paulin, A.; Issaka, Y.; Dominique, S. In Vitro Antifungal Activities of Essential Oils Extracted from Fresh Leaves of Cinnamomum zeylanicum and Ocimum gratissimum against Foodborne Pathogens for Their Use as Traditional Gheese Wagashi Conservatives. Res. J. Recent Sci. 2012, 1, 67–73. [Google Scholar]
  12. 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]
  13. Bajpai, V.K.; Baek, K.-H.; Kang, S.C. Control of Salmonella in Foods by Using Essential Oils: A Review. Food Res. Int. 2012, 45, 722–734. [Google Scholar] [CrossRef]
  14. Mishra, A.P.; Devkota, H.P.; Nigam, M.; Adetunji, C.O.; Srivastava, N.; Saklani, S.; Shukla, I.; Azmi, L.; Shariati, M.A.; Melo Coutinho, H.D.; et al. Combination of Essential Oils in Dairy Products: A Review of Their Functions and Potential Benefits. LWT 2020, 133, 110116. [Google Scholar] [CrossRef]
  15. Abdel-Maksoud, G.; El-Amin, A.R. A Review on the Materials Used during the Mummification Processes in Ancient Egypt. Mediterr. Archaeol. Ar. 2011, 11, 129–150. [Google Scholar]
  16. Barnes, K.M.; Whiffin, A.L.; Bulling, M.T. A Preliminary Study on the Antibacterial Activity and Insect Repellent Properties of Embalming Fluids from the 18th Dynasty (1550–1292 BCE) in Ancient Egypt. J. Archaeol. Sci. Rep. 2019, 25, 600–609. [Google Scholar] [CrossRef]
  17. Łucejko, J.; Connan, J.; Orsini, S.; Ribechini, E.; Modugno, F. Chemical Analyses of Egyptian Mummification Balms and Organic Residues from Storage Jars Dated from the Old Kingdom to the Copto-Byzantine Period. J. Archaeol. Sci. 2017, 85, 1–12. [Google Scholar] [CrossRef]
  18. Edris, A.E. Pharmaceutical and Therapeutic Potentials of Essential Oils and Their Individual Volatile Constituents: A Review. Phytother. Res. 2007, 21, 308–323. [Google Scholar] [CrossRef]
  19. Bajalan, I.; Rouzbahani, R.; Pirbalouti, A.G.; Maggi, F. Antioxidant and Antibacterial Activities of the Essential Oils Obtained from Seven Iranian Populations of Rosmarinus officinalis. Ind. Crops Prod. 2017, 107, 305–311. [Google Scholar] [CrossRef]
  20. Kumar, V.; Mathela, C.S.; Kumar, M.; Tewari, G. Antioxidant Potential of Essential Oils from Some Himalayan Asteraceae and Lamiaceae species. Med. Drug Discov. 2019, 1, 100004. [Google Scholar] [CrossRef]
  21. Goudjil, M.B.; Zighmi, S.; Hamada, D.; Mahcene, Z.; Bencheikh, S.E.; Ladjel, S. Biological Activities of Essential Oils Extracted from Thymus capitatus (Lamiaceae). S. Afr. J. Bot. 2020, 128, 274–282. [Google Scholar] [CrossRef]
  22. Sharma, K.; Guleria, S.; Razdan, V.K.; Babu, V. Synergistic Antioxidant and Antimicrobial Activities of Essential Oils of Some Selected Medicinal Plants in Combination and with Synthetic Compounds. Ind. Crops Prod. 2020, 154, 112569. [Google Scholar] [CrossRef]
  23. Jahani, M.; Pira, M.; Aminifard, M.H. Antifungal Effects of Essential Oils against Aspergillus niger in Vitro and in Vivo on Pomegranate (Punica granatum) Fruits. Sci. Hortic. 2020, 264, 109188. [Google Scholar] [CrossRef]
  24. Nazzaro, F.; Fratianni, F.; Coppola, R.; Feo, V. De Essential Oils and Antifungal Activity. Pharmaceuticals 2017, 10, 86. [Google Scholar] [CrossRef]
  25. Schuhmacher, A.; Reichling, J.; Schnitzler, P. Virucidal Effect of Peppermint Oil on the Enveloped Viruses Herpes simplex Virus Type 1 and Type 2 in Vitro. Phytomedicine 2003, 10, 504–510. [Google Scholar] [CrossRef]
  26. Sarmento-Neto, J.; Do Nascimento, L.; Felipe, C.; De Sousa, D. Analgesic Potential of Essential Oils. Molecules 2015, 21, 20. [Google Scholar] [CrossRef]
  27. Funk, J.L.; Frye, J.B.; Oyarzo, J.N.; Chen, J.; Zhang, H.; Timmermann, B.N. Anti-Inflammatory Effects of the Essential Oils of Ginger (Zingiber officinale Roscoe) in Experimental Rheumatoid Arthritis. PharmaNutrition 2016, 4, 123–131. [Google Scholar] [CrossRef] [PubMed]
  28. Zhang, L.; Liang, X.; Wang, B.; Lin, Z.; Ye, M.; Ma, R.; Zheng, M.; Xiang, H.; Xu, P. Six Herbs Essential Oils Suppressing Inflammatory Responses via Inhibiting COX-2/TNF-α/IL-6/NF-ΚB Activation. Microchem. J. 2020, 156, 104769. [Google Scholar] [CrossRef]
  29. Akhtar, Y.; Pages, E.; Stevens, A.; Bradbury, R.; da Camara, C.A.G.; Isman, M.B. Effect of Chemical Complexity of Essential Oils on Feeding Deterrence in Larvae of the Cabbage looper. Physiol. Entomol. 2012, 37, 81–91. [Google Scholar] [CrossRef]
  30. Food and Agriculture Organization of the United Nations, FAO. Concerted Action Needed to Stop Diseases and Pests from Ravaging the Food Chain. Available online: https://www.fao.org/newsroom/detail/Concerted-action-needed-to-stop-diseases-and-pests-from-ravaging-the-food-chain/es (accessed on 10 October 2020).
  31. Franzios, G.; Mirotsou, M.; Hatziapostolou, E.; Kral, J.; Scouras, Z.G.; Mavragani-Tsipidou, P. Insecticidal and Genotoxic Activities of Mint Essential Oils. J. Agric. Food Chem. 1997, 45, 2690–2694. [Google Scholar] [CrossRef]
  32. Štefanidesová, K.; Škultéty, Ľ.; Sparagano, O.A.E.; Špitalská, E. The Repellent Efficacy of Eleven Essential Oils against Adult Dermacentor reticulatus Ticks. Ticks Tick Borne. Dis. 2017, 8, 780–786. [Google Scholar] [CrossRef] [PubMed]
  33. Katiki, L.M.; Barbieri, A.M.E.; Araujo, R.C.; Veríssimo, C.J.; Louvandini, H.; Ferreira, J.F.S. Synergistic Interaction of Ten Essential Oils against Haemonchus contortus in Vitro. Vet. Parasitol. 2017, 243, 47–51. [Google Scholar] [CrossRef] [PubMed]
  34. Pavela, R. Insecticidal Activity of Some Essential Oils against Larvae of Spodoptera littoralis. Fitoterapia 2005, 76, 691–696. [Google Scholar] [CrossRef] [PubMed]
  35. Ben El Hadj Ali, I.; Chaouachi, M.; Bahri, R.; Chaieb, I.; Boussaïd, M.; Harzallah-Skhiri, F. Chemical Composition and Antioxidant, Antibacterial, Allelopathic and Insecticidal Activities of Essential Oil of Thymus algeriensis Boiss. et Reut. Ind. Crops Prod. 2015, 77, 631–639. [Google Scholar] [CrossRef]
  36. Jafari, S.M. Nanoencapsulation Technologies for the Food and Nutraceutical Industries; Jafari, S.M., Ed.; Academic Press: Cambridge, MA, USA, 2017; pp. 1–34. [Google Scholar]
  37. Wu, C.-C.; Chung, J.G.; Tsai, S.-J.; Yang, J.H.; Sheen, L.Y. Differential Effects of Allyl Sulfides from Garlic Essential Oil on Cell Cycle Regulation in Human Liver Tumor Cells. Food Chem. Toxicol. 2004, 42, 1937–1947. [Google Scholar] [CrossRef] [PubMed]
  38. Pellizzeri, V.; Costa, R.; Grasso, E.; Dugo, G. Valuable Products from the Flowers of Lemon (Citrus limon (L.) Osbeck) and Grapefruit (Citrus paradisi Macfad.) Italian Trees. Food Bioprod. Process. 2020, 123, 123–133. [Google Scholar] [CrossRef]
  39. Ríos, J.-L. Essential Oils in Food Preservation, Flavor and Safety; Preedy, V.R., Ed.; Academic Press: Cambridge, MA, USA, 2016; pp. 3–10. [Google Scholar]
  40. dos Santos, C.P.; de Oliveira, T.C.; Pinto, J.A.O.; Fontes, S.S.; Cruz, E.M.O.; de Fátima Arrigoni-Blank, M.; Andrade, T.M.; de Matos, I.L.; Alves, P.B.; Innecco, R.; et al. Chemical Diversity and Influence of Plant Age on the Essential Oil from Lippia sidoides Cham. Germplasm. Ind. Crops Prod. 2015, 76, 416–421. [Google Scholar] [CrossRef]
  41. Tripathi, A.K.; Upadhyay, S.; Bhuiyan, M.; Bhattacharya, P.R. A Review on Prospects of Essential Oils as Biopesticide in Insect-Pest Management. J. Pharmacogn. Phytother. 2009, 1, 52–063. [Google Scholar]
  42. Rehman, R.; Asif Hanif, M. Biosynthetic Factories of Essential Oils: The Aromatic Plants. Nat. Prod. Chem. Res. 2016, 4, 1000227. [Google Scholar] [CrossRef]
  43. Burt, S. Essential Oils: Their Antibacterial Properties and Potential Applications in Foods—A Review. Int. J. Food Microbiol. 2004, 94, 223–253. [Google Scholar] [CrossRef] [PubMed]
  44. Dima, C.; Dima, S. Essential Oils in Foods: Extraction, Stabilization, and Toxicity. Curr. Opin. Food Sci. 2015, 5, 29–35. [Google Scholar] [CrossRef]
  45. Zhou, W.; Li, J.; Wang, X.; Liu, L.; Li, Y.; Song, R.; Zhang, M.; Li, X. Research Progress on Extraction, Separation, and Purification Methods of Plant Essential Oils. Separations 2023, 10, 596. [Google Scholar] [CrossRef]
  46. Khan, S.; Abdo, A.A.A.; Shu, Y.; Zhang, Z.; Liang, T. The Extraction and Impact of Essential Oils on Bioactive Films and Food Preservation, with Emphasis on Antioxidant and Antibacterial Activities—A Review. Foods 2023, 12, 4169. [Google Scholar] [CrossRef] [PubMed]
  47. Stratakos, A.C. Essential Oils in Food Preservation, Flavor and Safety; Preedy, V.R., Ed.; Academic Press: Cambridge, MA, USA, 2016; pp. 31–38. [Google Scholar]
  48. Souiy, Z. Essential Oil Extraction Process. In Book Essential Oils—Recent Advances, New Perspectives and Applications; Jonas, V., Ed.; IntechOpen: London, UK, 2024; Volume 1, Available online: https://www.intechopen.com/chapters/88433 (accessed on 22 September 2024).
  49. Kant, R.; Kumar, A. Review on essential oil extraction from aromatic and medicinal plants: Techniques, performance and economic analysis. Sustain. Chem. Pharm. 2022, 30, 100829. Available online: https://www.sciencedirect.com/science/article/pii/S2352554122002339?via%3Dihub (accessed on 22 September 2024). [CrossRef]
  50. Kim, S.-W.; Kang, J.; Park, I.-K. Fumigant Toxicity of Apiaceae Essential Oils and Their Constituents against Sitophilus oryzae and Their Acetylcholinesterase Inhibitory Activity. J. Asia Pac. Entomol. 2013, 16, 443–448. [Google Scholar] [CrossRef]
  51. Andrés, M.F.; Rossa, G.E.; Cassel, E.; Vargas, R.M.F.; Santana, O.; Díaz, C.E.; González-Coloma, A. Biocidal Effects of Piper hispidinervum (Piperaceae) Essential Oil and Synergism among Its Main Components. Food Chem. Toxicol. 2017, 109, 1086–1092. [Google Scholar] [CrossRef]
  52. Cao, J.-Q.; Guo, S.-S.; Wang, Y.; Pang, X.; Geng, Z.-F.; Du, S.-S. Toxicity and Repellency of Essential Oil from Evodia lenticellata Huang Fruits and Its Major Monoterpenes against Three Stored-Product Insects. Ecotoxicol. Environ. Saf. 2018, 160, 342–348. [Google Scholar] [CrossRef]
  53. Lerdau, M.; Litvak, M.; Monson, R. Plant Chemical Defense: Monoterpenes and the Growth-Differentiation Balance Hypothesis. Trends Ecol. Evol. 1994, 9, 58–61. [Google Scholar] [CrossRef] [PubMed]
  54. Plata-Rueda, A.; Campos, J.M.; da Silva Rolim, G.; Martínez, L.C.; Dos Santos, M.H.; Fernandes, F.L.; Serrão, J.E.; Zanuncio, J.C. Terpenoid Constituents of Cinnamon and Clove Essential Oils Cause Toxic Effects and Behavior Repellency Response on Granary Weevil, Sitophilus Granarius. Ecotoxicol. Environ. Saf. 2018, 156, 263–270. [Google Scholar] [CrossRef]
  55. Abdelgaleil, S.A.M.; Badawy, M.E.I.; Mahmoud, N.F.; Marei, A.E.-S.M. Acaricidal Activity, Biochemical Effects and Molecular Docking of Some Monoterpenes against Two-Spotted Spider Mite (Tetranychus urticae Koch). Pestic. Biochem. Physiol. 2019, 156, 105–115. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, Y.; Zhang, L.-T.; Feng, Y.-X.; Zhang, D.; Guo, S.-S.; Pang, X.; Geng, Z.-F.; Xi, C.; Du, S.-S. Comparative Evaluation of the Chemical Composition and Bioactivities of Essential Oils from Four Spice Plants (Lauraceae) against Stored-Product Insects. Ind. Crops Prod. 2019, 140, 111640. [Google Scholar] [CrossRef]
  57. de Andrade Brito, F.; Bacci, L.; da Silva Santana, A.; da Silva, J.E.; de Castro Nizio, D.A.; de Lima Nogueira, P.C.; de Fatima Arrigoni-Blank, M.; Melo, C.R.; de Melo, J.O.; Blank, A.F. Toxicity and Behavioral Alterations Caused by Essential Oils of Croton tetradenius and Their Major Compounds on Acromyrmex balzani. Crop Prot. 2020, 137, 105259. [Google Scholar] [CrossRef]
  58. Li, A.S.; Iijima, A.; Huang, J.; Li, Q.X.; Chen, Y. Putative Mode of Action of the Monoterpenoids Linalool, Methyl Eugenol, Estragole, and Citronellal on Ligand-Gated Ion Channels. Engineering 2020, 6, 541–545. [Google Scholar] [CrossRef]
  59. Walker, K.; Frederick, R. Encyclopedia of Environmental Health; Nriagu, J.O., Ed.; Elsevier Science: Philadelphia, PA, USA, 2011; pp. 306–314. ISBN 0444522735. [Google Scholar]
  60. Sawicka, B.; Egbuna, C. Pests of Agricultural Crops and Control Measures. In Natural Remedies for Pest, Disease and Weed Control; Elsevier: Amsterdam, The Netherlands, 2020; pp. 1–16. [Google Scholar]
  61. U.S. Food and Drug Administration, FDA. Substances Generally Recognized as Safe (GRAS). Available online: https://www.accessdata.fda.gov/scripts/cdrh/cfdocs/cfcfr/CFRSearch.cfm?FR=182.20 (accessed on 22 September 2024).
  62. da Silva, J.C.P.; Campos, V.P.; Barros, A.F.; Pedroso, L.A.; de Freitas Silva, M.; de Souza, J.T.; Pedroso, M.P.; de Medeiros, F.H.V. Performance of Volatiles Emitted from Different Plant Species against Juveniles and Eggs of Meloidogyne incognita. Crop Prot. 2019, 116, 196–203. [Google Scholar] [CrossRef]
  63. de Freitas Silva, M.; Paulo Campos, V.; Barros, A.F.; da Silva, J.C.P.; Pedroso, M.P.; de Jesus Silva, F.; Gomes, V.A.; Justino, J.C. Medicinal Plant Volatiles Applied against the Root-Knot Nematode Meloidogyne incognita. Crop Prot. 2020, 130, 105057. [Google Scholar] [CrossRef]
  64. Osman Mohamed Ali, E.; Shakil, N.A.; Rana, V.S.; Sarkar, D.J.; Majumder, S.; Kaushik, P.; Singh, B.B.; Kumar, J. Antifungal Activity of Nano Emulsions of Neem and Citronella Oils against Phytopathogenic Fungi, Rhizoctonia solani and Sclerotium rolfsii. Ind. Crops Prod. 2017, 108, 379–387. [Google Scholar] [CrossRef]
  65. Regnier, T.; Combrinck, S.; Veldman, W.; Du Plooy, W. Application of Essential Oils as Multi-Target Fungicides for the Control of Geotrichum citriaurantii and Other Postharvest Pathogens of Citrus. Ind. Crops Prod. 2014, 61, 151–159. [Google Scholar] [CrossRef]
  66. Abramson, C.I.; Wanderley, P.A.; Wanderley, M.J.A.; Mina, A.J.S.; Souza, O.B. de Effect of Essential Oil from Citronella and Alfazema on Fennel Aphids Hyadaphis foeniculi Passerini (Hemiptera: Aphididae) and Its Predator Cycloneda sanguinea L. (Coleoptera: Coccinelidae). Am. J. Environ. Sci. 2007, 3, 9–10. [Google Scholar] [CrossRef]
  67. Afify, A.E.-M.M.; Ali, F.S.; Turky, A. Control of Tetranychus urticae Koch by Extracts of Three Essential Oils of Chamomile, Marjoram and Eucalyptus. Asian Pac. J. Trop. Biomed. 2012, 2, 24–30. [Google Scholar] [CrossRef]
  68. Laborda, R.; Manzano, I.; Gamón, M.; Gavidia, I.; Pérez-Bermúdez, P.; Boluda, R. Effects of Rosmarinus officinalis and Salvia officinalis Essential Oils on Tetranychus urticae Koch (Acari: Tetranychidae). Ind. Crops Prod. 2013, 48, 106–110. [Google Scholar] [CrossRef]
  69. Chrysargyris, A.; Laoutari, S.; Litskas, V.D.; Stavrinides, M.C.; Tzortzakis, N. Effects of Water Stress on Lavender and Sage Biomass Production, Essential Oil Composition and Biocidal Properties against Tetranychus urticae (Koch). Sci. Hortic. 2016, 213, 96–103. [Google Scholar] [CrossRef]
  70. Basaid, K.; Mayad, E.H.; Bouharroud, R.; Furze, J.N.; Benjlil, H.; de Oliveira, A.L.; Chebli, B. Biopesticidal Value of Senecio glaucus Subsp. Coronopifolius Essential Oil against Pathogenic Fungi, Nematodes, and Mites. Mater. Today Proc. 2020, 27, 3082–3090. [Google Scholar] [CrossRef]
  71. Zandi-Sohani, N.; Ramezani, L. Evaluation of Five Essential Oils as Botanical Acaricides against the Strawberry Spider Mite Tetranychus turkestani Ugarov and Nikolskii. Int. Biodeterior. Biodegrad. 2015, 98, 101–106. [Google Scholar] [CrossRef]
  72. Sertkaya, E.; Kaya, K.; Soylu, S. Acaricidal Activities of the Essential Oils from Several Medicinal Plants against the Carmine Spider Mite (Tetranychus Cinnabarinus Boisd.) (Acarina: Tetranychidae). Ind. Crops Prod. 2010, 31, 107–112. [Google Scholar] [CrossRef]
  73. Kalaiselvi, D.; Mohankumar, A.; Shanmugam, G.; Thiruppathi, G.; Nivitha, S.; Sundararaj, P. Altitude-Related Changes in the Phytochemical Profile of Essential Oils Extracted from Artemisia nilagirica and Their Nematicidal Activity against Meloidogyne incognita. Ind. Crops Prod. 2019, 139, 111472. [Google Scholar] [CrossRef]
  74. Gupta, A.; Sharma, S.; Naik, S.N. Biopesticidal Value of Selected Essential Oils against Pathogenic Fungus, Termites, and Nematodes. Int. Biodeterior. Biodegrad. 2011, 65, 703–707. [Google Scholar] [CrossRef]
  75. Onifade, A.K.; Fatope, M.O.; Deadman, M.L.; Al-Kindy, S.M.Z. Nematicidal Activity of Haplophyllum tuberculatum and Plectranthus cylindraceus Oils against Meloidogyne javanica. Biochem. Syst. Ecol. 2008, 36, 679–683. [Google Scholar] [CrossRef]
  76. Kang, J.S.; Kim, E.; Lee, S.H.; Park, I.-K. Inhibition of Acetylcholinesterases of the Pinewood Nematode, Bursaphelenchus xylophilus, by Phytochemicals from Plant Essential Oils. Pestic. Biochem. Physiol. 2013, 105, 50–56. [Google Scholar] [CrossRef]
  77. Cheng, Z.-H.; Fan, F.-F.; Zhao, J.-Z.; Li, R.; Li, S.-C.; Zhang, E.-J.; Liu, Y.-K.; Wang, J.-Y.; Zhu, X.-R.; Tian, Y.-M. Optimization of the Microemulsion Formulation of Curcuma Oil and Evaluation of Its Acaricidal Efficacy against Tetranychus cinnabarinus (Boisduval) (Acari: Tetranychidae). J. Asia Pac. Entomol. 2020, 23, 1014–1022. [Google Scholar] [CrossRef]
  78. Pavela, R.; Stepanycheva, E.; Shchenikova, A.; Chermenskaya, T.; Petrova, M. Essential Oils as Prospective Fumigants against Tetranychus urticae Koch. Ind. Crops Prod. 2016, 94, 755–761. [Google Scholar] [CrossRef]
  79. Roh, H.S.; Lee, B.H.; Park, C.G. Acaricidal and Repellent Effects of Myrtacean Essential Oils and Their Major Constituents against Tetranychus urticae (Tetranychidae). J. Asia Pac. Entomol. 2013, 16, 245–249. [Google Scholar] [CrossRef]
  80. Mar, J.M.; Silva, L.S.; Azevedo, S.G.; França, L.P.; Goes, A.F.F.; dos Santos, A.L.; Bezerra, J.d.A.; Nunomura, R.d.C.S.; Machado, M.B.; Sanches, E.A. Lippia origanoides Essential Oil: An Efficient Alternative to Control Aedes aegypti, Tetranychus urticae and Cerataphis lataniae. Ind. Crops Prod. 2018, 111, 292–297. [Google Scholar] [CrossRef]
  81. Cavalcanti, S.C.H.; dos, S. Niculau, E.; Blank, A.F.; Câmara, C.A.G.; Araújo, I.N.; Alves, P.B. Composition and Acaricidal Activity of Lippia sidoides Essential Oil against Two-Spotted Spider Mite (Tetranychus urticae Koch). Bioresour. Technol. 2010, 101, 829–832. [Google Scholar] [CrossRef] [PubMed]
  82. Rani, P.U. Fumigant and Contact Toxic Potential of Essential Oils from Plant Extracts against Stored Product Pests. J. Biopestic. 2011, 5, 120–128. [Google Scholar] [CrossRef]
  83. Mahmoudvand, M.; Abbasipour, H.; Hosseinpour, M.H.; Rastegar, F.; Basij, M. Using Some Plant Essential Oils as Natural Fumigants against Adults of Callosobruchus maculatus (F.) (Coleoptera: Bruchidae). Munis Entomol. Zool. 2011, 6, 150–154. [Google Scholar]
  84. Oyedeji, A.O.; Okunowo, W.O.; Osuntoki, A.A.; Olabode, T.B.; Ayo-folorunso, F. Insecticidal and Biochemical Activity of Essential Oil from Citrus sinensis Peel and Constituents on Callosobrunchus maculatus and Sitophilus zeamais. Pestic. Biochem. Physiol. 2020, 168, 104643. [Google Scholar] [CrossRef] [PubMed]
  85. Kim, S.-I.; Park, C.; Ohh, M.-H.; Cho, H.-C.; Ahn, Y.-J. Contact and Fumigant Activities of Aromatic Plant Extracts and Essential Oils against Lasioderma serricorne (Coleoptera: Anobiidae). J. Stored Prod. Res. 2003, 39, 11–19. [Google Scholar] [CrossRef]
  86. Plata-Rueda, A.; Martínez, L.C.; da Silva Rolim, G.; Coelho, R.P.; Santos, M.H.; de Souza Tavares, W.; Zanuncio, J.C.; Serrão, J.E. Insecticidal and Repellent Activities of Cymbopogon citratus (Poaceae) Essential Oil and Its Terpenoids (Citral and Geranyl Acetate) against Ulomoides dermestoides. Crop Prot. 2020, 137, 105299. [Google Scholar] [CrossRef]
  87. Benelli, G.; Flamini, G.; Canale, A.; Cioni, P.L.; Conti, B. Toxicity of Some Essential Oil Formulations against the Mediterranean Fruit Fly Ceratitis capitata (Wiedemann) (Diptera tephritidae). Crop Prot. 2012, 42, 223–229. [Google Scholar] [CrossRef]
  88. Park, C.G.; Jang, M.; Yoon, K.A.; Kim, J. Insecticidal and Acetylcholinesterase Inhibitory Activities of Lamiaceae Plant Essential Oils and Their Major Components against Drosophila suzukii (Diptera: Drosophilidae). Ind. Crops Prod. 2016, 89, 507–513. [Google Scholar] [CrossRef]
  89. Bravim dos Santos, A.T.; Zanuncio Junior, J.S.; Parreira, L.A.; Pedra de Abreu, K.M.; de Oliveira Bernardes, C.; Romário de Carvalho, J.; Menini, L. Chemical Identification and Insecticidal Effect of Tephrosia vogelii Essential Oil against Cerosipha forbesi in Strawberry Crop. Crop Prot. 2021, 139, 105405. [Google Scholar] [CrossRef]
  90. Kim, S.; Chae, S.; Youn, H.; Yeon, S.; Ahn, Y. Contact and Fumigant Toxicity of Plant Essential Oils and Efficacy of Spray Formulations Containing the Oils against B- and Q-biotypes of Bemisia tabaci. Pest Manag. Sci. 2011, 67, 1093–1099. [Google Scholar] [CrossRef] [PubMed]
  91. Çalmaşur, Ö.; Aslan, İ.; Şahin, F. Insecticidal and Acaricidal Effect of Three Lamiaceae Plant Essential Oils against Tetranychus urticae Koch and Bemisia tabaci Genn. Ind. Crops Prod. 2006, 23, 140–146. [Google Scholar] [CrossRef]
  92. Yang, N.-W.; Li, A.-L.; Wan, F.-H.; Liu, W.-X.; Johnson, D. Effects of Plant Essential Oils on Immature and Adult Sweetpotato Whitefly, Bemisia tabaci Biotype B. Crop Prot. 2010, 29, 1200–1207. [Google Scholar] [CrossRef]
  93. Zarrad, K.; Hamouda, A.B.; Chaieb, I.; Laarif, A.; Jemâa, J.M.-B. Chemical Composition, Fumigant and Anti-Acetylcholinesterase Activity of the Tunisian Citrus aurantium L. Essential Oils. Ind. Crops Prod. 2015, 76, 121–127. [Google Scholar] [CrossRef]
  94. Benelli, G.; Pavela, R.; Petrelli, R.; Cappellacci, L.; Canale, A.; Senthil-Nathan, S.; Maggi, F. Not Just Popular Spices! Essential Oils from Cuminum cyminum and Pimpinella anisum Are Toxic to Insect Pests and Vectors without Affecting Non-Target Invertebrates. Ind. Crops Prod. 2018, 124, 236–243. [Google Scholar] [CrossRef]
  95. Czerniewicz, P.; Chrzanowski, G.; Sprawka, I.; Sytykiewicz, H. Aphicidal Activity of Selected Asteraceae Essential Oils and Their Effect on Enzyme Activities of the Green Peach Aphid, Myzus persicae (Sulzer). Pestic. Biochem. Physiol. 2018, 145, 84–92. [Google Scholar] [CrossRef]
  96. Benelli, G.; Pavela, R.; Petrelli, R.; Cappellacci, L.; Santini, G.; Fiorini, D.; Sut, S.; Dall’Acqua, S.; Canale, A.; Maggi, F. The Essential Oil from Industrial Hemp (Cannabis Sativa L.) by-Products as an Effective Tool for Insect Pest Management in Organic Crops. Ind. Crops Prod. 2018, 122, 308–315. [Google Scholar] [CrossRef]
  97. Kimbaris, A.C.; Papachristos, D.P.; Michaelakis, A.; Martinou, A.F.; Polissiou, M.G. Toxicity of Plant Essential Oil Vapours to Aphid Pests and Their Coccinellid Predators. Biocontrol. Sci. Technol. 2010, 20, 411–422. [Google Scholar] [CrossRef]
  98. Petrakis, E.A.; Kimbaris, A.C.; Perdikis, D.C.; Lykouressis, D.P.; Tarantilis, P.A.; Polissiou, M.G. Responses of Myzus persicae (Sulzer) to Three Lamiaceae Essential Oils Obtained by Microwave-Assisted and Conventional Hydrodistillation. Ind. Crops Prod. 2014, 62, 272–279. [Google Scholar] [CrossRef]
  99. Toledo, P.F.S.; Viteri Jumbo, L.O.; Rezende, S.M.; Haddi, K.; Silva, B.A.; Mello, T.S.; Della Lucia, T.M.C.; Aguiar, R.W.S.; Smagghe, G.; Oliveira, E.E. Disentangling the Ecotoxicological Selectivity of Clove Essential Oil against Aphids and Non-Target Ladybeetles. Sci. Total Environ. 2020, 718, 137328. [Google Scholar] [CrossRef] [PubMed]
  100. Umpiérrez, M.L.; Paullier, J.; Porrini, M.; Garrido, M.; Santos, E.; Rossini, C. Potential Botanical Pesticides from Asteraceae Essential Oils for Tomato Production: Activity against Whiteflies, Plants and Bees. Ind. Crops Prod. 2017, 109, 686–692. [Google Scholar] [CrossRef]
  101. Zapata, N.; Vargas, M.; Latorre, E.; Roudergue, X.; Ceballos, R. The Essential Oil of Laurelia sempervirens Is Toxic to Trialeurodes vaporariorum and Encarsia formosa. Ind. Crops Prod. 2016, 84, 418–422. [Google Scholar] [CrossRef]
  102. Silva, D.C.; de Fátima Arrigoni-Blank, M.; Bacci, L.; Blank, A.F.; Nunes Faro, R.R.; Oliveira Pinto, J.A.; Garcia Pereira, K.L. Toxicity and Behavioral Alterations of Essential Oils of Eplingiella fruticosa Genotypes and Their Major Compounds to Acromyrmex balzani. Crop Prot. 2019, 116, 181–187. [Google Scholar] [CrossRef]
  103. Melo, C.R.; Blank, A.F.; Oliveira, B.M.S.; Santos, A.C.C.; Cristaldo, P.F.; Araújo, A.P.A.; Bacci, L. Formicidal Activity of Essential Oils of Myrcia lundiana Chemotypes on Acromyrmex balzani. Crop Prot. 2021, 139, 105343. [Google Scholar] [CrossRef]
  104. Ayvaz, A.; Sagdic, O.; Karaborklu, S.; Ozturk, I. Insecticidal Activity of the Essential Oils from Different Plants Against Three Stored-Product Insects. J. Insect Sci. 2010, 10, 1–13. [Google Scholar] [CrossRef]
  105. Pavela, R.; Maggi, F.; Lupidi, G.; Cianfaglione, K.; Dauvergne, X.; Bruno, M.; Benelli, G. Efficacy of Sea Fennel (Crithmum maritimum L., Apiaceae) Essential Oils against Culex quinquefasciatus Say and Spodoptera littoralis (Boisd.). Ind. Crops Prod. 2017, 109, 603–610. [Google Scholar] [CrossRef]
  106. Ali, A.M.; Ibrahim, A.M.A. Castor and Camphor Essential Oils Alter Hemocyte Populations and Induce Biochemical Changes in Larvae of Spodoptera littoralis (Boisduval) (Lepidoptera: Noctuidae). J. Asia Pac. Entomol. 2018, 21, 631–637. [Google Scholar] [CrossRef]
  107. Benelli, G.; Pavela, R.; Maggi, F.; Nkuimi Wandjou, J.G.; Yvette Fofie, N.G.B.; Koné-Bamba, D.; Sagratini, G.; Vittori, S.; Caprioli, G. Insecticidal Activity of the Essential Oil and Polar Extracts from Ocimum gratissimum Grown in Ivory Coast: Efficacy on Insect Pests and Vectors and Impact on Non-Target Species. Ind. Crops Prod. 2019, 132, 377–385. [Google Scholar] [CrossRef]
  108. Mansour, S.A.; El-Sharkawy, A.Z.; Abdel-Hamid, N.A. Toxicity of Essential Plant Oils, in Comparison with Conventional Insecticides, against the Desert Locust, Schistocerca gregaria (Forskål). Ind. Crops Prod. 2015, 63, 92–99. [Google Scholar] [CrossRef]
  109. Abdelatti, Z.A.S.; Hartbauer, M. Plant Oil Mixtures as a Novel Botanical Pesticide to Control Gregarious locusts. J. Pest Sci. 2020, 93, 341–353. [Google Scholar] [CrossRef]
  110. Enan, E. Insecticidal Activity of Essential Oils: Octopaminergic Sites of Action. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2001, 130, 325–337. [Google Scholar] [CrossRef]
  111. Jankowska, M.; Rogalska, J.; Wyszkowska, J.; Stankiewicz, M. Molecular Targets for Components of Essential Oils in the Insect Nervous System—A Review. Molecules 2017, 23, 34. [Google Scholar] [CrossRef] [PubMed]
  112. Kim, Y.H.; Lee, S.H. Invertebrate Acetylcholinesterases: Insights into Their Evolution and Non-Classical Functions. J. Asia Pac. Entomol. 2018, 21, 186–195. [Google Scholar] [CrossRef]
  113. Fukuto, T.R. Mechanism of Action of Organophosphorus and Carbamate Insecticides. Environ. Health Perspect. 1990, 87, 245–254. [Google Scholar] [CrossRef] [PubMed]
  114. Lee, S.H.; Kim, Y.H.; Kwon, D.H.; Cha, D.J.; Kim, J.H. Mutation and Duplication of Arthropod Acetylcholinesterase: Implications for Pesticide Resistance and Tolerance. Pestic. Biochem. Physiol. 2015, 120, 118–124. [Google Scholar] [CrossRef] [PubMed]
  115. Karakaya, S.; Yılmaz, S.V.; Koca, M.; Demirci, B.; Sytar, O. Screening of Non-Alkaloid Acetylcholinesterase Inhibitors from Extracts and Essential Oils of Anthriscus nemorosa (M.Bieb.) Spreng. (Apiaceae). S. Afr. J. Bot. 2019, 125, 261–269. [Google Scholar] [CrossRef]
  116. Orhan, I.; Şenol, F.S.; Gülpinar, A.R.; Kartal, M.; Şekeroglu, N.; Deveci, M.; Kan, Y.; Şener, B. Acetylcholinesterase Inhibitory and Antioxidant Properties of Cyclotrichium niveum, Thymus praecox Subsp. caucasicus Var. caucasicus, Echinacea purpurea and E. pallida. Food Chem. Toxicol. 2009, 47, 1304–1310. [Google Scholar] [CrossRef]
  117. Bhavya, M.L.; Chandu, A.G.S.; Devi, S.S. Ocimum tenuiflorum Oil, a Potential Insecticide against Rice Weevil with Anti-Acetylcholinesterase Activity. Ind. Crops Prod. 2018, 126, 434–439. [Google Scholar] [CrossRef]
  118. Castillo-Morales, R.M.; Carreño Otero, A.L.; Mendez-Sanchez, S.C.; Da Silva, M.A.N.; Stashenko, E.E.; Duque, J.E. Mitochondrial Affectation, DNA Damage and AChE Inhibition Induced by Salvia officinalis Essential Oil on Aedes aegypti Larvae. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2019, 221, 29–37. [Google Scholar] [CrossRef]
  119. El Euch, S.K.; Hassine, D.B.; Cazaux, S.; Bouzouita, N.; Bouajila, J. Salvia officinalis Essential Oil: Chemical Analysis and Evaluation of Anti-Enzymatic and Antioxidant Bioactivities. S. Afr. J. Bot. 2019, 120, 253–260. [Google Scholar] [CrossRef]
  120. Perry, N.S.L.; Houghton, P.J.; Jenner, P.; Keith, A.; Perry, E.K. Salvia lavandulaefolia Essential Oil Inhibits Cholinesterase in Vivo. Phytomedicine 2002, 9, 48–51. [Google Scholar] [CrossRef] [PubMed]
  121. Kiran, S.; Prakash, B. Toxicity and Biochemical Efficacy of Chemically Characterized Rosmarinus Officinalis Essential Oil against Sitophilus oryzae and Oryzaephilus surinamensis. Ind. Crops Prod. 2015, 74, 817–823. [Google Scholar] [CrossRef]
  122. Seo, S.-M.; Kim, J.; Kang, J.; Koh, S.-H.; Ahn, Y.-J.; Kang, K.-S.; Park, I.-K. Fumigant Toxicity and Acetylcholinesterase Inhibitory Activity of 4 Asteraceae Plant Essential Oils and Their Constituents against Japanese Termite (Reticulitermes speratus Kolbe). Pestic. Biochem. Physiol. 2014, 113, 55–61. [Google Scholar] [CrossRef]
  123. López, M.D.; Pascual-Villalobos, M.J. Mode of Inhibition of Acetylcholinesterase by Monoterpenoids and Implications for Pest Control. Ind. Crops Prod. 2010, 31, 284–288. [Google Scholar] [CrossRef]
  124. Lummis, S.C.R.; Sattelle, D.B. Insect Central Nervous System γ-Aminobutyric Acid. Neurosci. Lett. 1985, 60, 13–18. [Google Scholar] [CrossRef]
  125. Robinson, T.; MacAllan, D.; Lunt, G.; Battersby, M. Γ-Aminobutyric Acid Receptor Complex of Insect CNS: Characterization of a Benzodiazepine Binding Site. J. Neurochem. 1986, 47, 1955–1962. [Google Scholar] [CrossRef]
  126. Ashby, J.A.; McGonigle, I.V.; Price, K.L.; Cohen, N.; Comitani, F.; Dougherty, D.A.; Molteni, C.; Lummis, S.C.R. GABA Binding to an Insect GABA Receptor: A Molecular Dynamics and Mutagenesis Study. Biophys. J. 2012, 103, 2071–2081. [Google Scholar] [CrossRef] [PubMed]
  127. Priestley, C.M.; Williamson, E.M.; Wafford, K.A.; Sattelle, D.B. Thymol, a Constituent of Thyme Essential Oil, Is a Positive Allosteric Modulator of Human GABA A Receptors and a Homo-oligomeric GABA Receptor from Drosophila melanogaster. Br. J. Pharmacol. 2003, 140, 1363–1372. [Google Scholar] [CrossRef] [PubMed]
  128. García, D.A.; Bujons, J.; Vale, C.; Suñol, C. Allosteric Positive Interaction of Thymol with the GABAA Receptor in Primary Cultures of Mouse Cortical Neurons. Neuropharmacology 2006, 50, 25–35. [Google Scholar] [CrossRef]
  129. Tong, F.; Coats, J.R. Effects of Monoterpenoid Insecticides on [3H]-TBOB Binding in House Fly GABA Receptor and 36Cl− Uptake in American Cockroach Ventral Nerve Cord. Pestic. Biochem. Physiol. 2010, 98, 317–324. [Google Scholar] [CrossRef]
  130. Waliwitiya, R.; Belton, P.; Nicholson, R.A.; Lowenberger, C.A. Effects of the Essential Oil Constituent Thymol and Other Neuroactive Chemicals on Flight Motor Activity and Wing Beat Frequency in the Blowfly Phaenicia sericata. Pest Manag. Sci. 2010, 66, 277–289. [Google Scholar] [CrossRef]
  131. Gashout, H.A.; Guzman-Novoa, E.; Goodwin, P.H.; Correa-Benítez, A. Impact of Sublethal Exposure to Synthetic and Natural Acaricides on Honey Bee (Apis mellifera) Memory and Expression of Genes Related to Memory. J. Insect Physiol. 2020, 121, 104014. [Google Scholar] [CrossRef] [PubMed]
  132. Atwood, H.L.; Klose, M.K. Neuromuscular Transmission Modulation at Invertebrate Neuromuscular Junctions. In Encyclopedia of Neuroscience; Elsevier: Amsterdam, The Netherlands, 2009; pp. 671–690. [Google Scholar]
  133. Hana, S.; Lange, A.B. Octopamine and Tyramine Regulate the Activity of Reproductive Visceral Muscles in the Adult Female Blood-Feeding Bug, Rhodnius prolixus. J. Exp. Biol. 2017, 220, 1830–1836. [Google Scholar] [CrossRef]
  134. Livingstone, M.S.; Harris-Warrick, R.M.; Kravitz, E.A. Serotonin and Octopamine Produce Opposite Postures in Lobsters. Science 1980, 208, 76–79. [Google Scholar] [CrossRef] [PubMed]
  135. Kostyukovsky, M.; Rafaeli, A.; Gileadi, C.; Demchenko, N.; Shaaya, E. Activation of Octopaminergic Receptors by Essential Oil Constituents Isolated from Aromatic Plants: Possible Mode of Action against Insect Pests. Pest Manag. Sci. 2002, 58, 1101–1106. [Google Scholar] [CrossRef]
  136. Price, D.N.; Berry, M.S. Comparison of Effects of Octopamine and Insecticidal Essential Oils on Activity in the Nerve Cord, Foregut, and Dorsal Unpaired Median Neurons of Cockroaches. J. Insect Physiol. 2006, 52, 309–319. [Google Scholar] [CrossRef] [PubMed]
  137. Neckameyer, W.S.; Leal, S.M. Non-Mammalian Hormone-Behavior Systems. In Hormones, Brain and Behavior; Pfaff, D.W., Joëls, M., Eds.; Academic Press: Cambridge, MA, USA, 2017; Volume 1, pp. 3–365. [Google Scholar]
  138. Mizunami, M.; Matsumoto, Y.; Watanabe, H.; Nishino, H. Olfactory and Visual Learning in Cockroaches and Crickets; Elsevier: Amsterdam, The Netherlands, 2013; pp. 549–560. [Google Scholar]
  139. Lee, K.H.; Lee, J.-S.; Kim, E.S.; Lee, H.G. Preparation, Characterization, and Food Application of Rosemary Extract-Loaded Antimicrobial Nanoparticle Dispersions. LWT 2019, 101, 138–144. [Google Scholar] [CrossRef]
  140. Soares de Oliveira, M.A.; Melo Coutinho, H.D.; Jardelino de Lacerda Neto, L.; Castro de Oliveira, L.C.; Bezerra da Cunha, F.A. Repellent Activity of Essential Oils against Culicids: A Review. Sustain. Chem. Pharm. 2020, 18, 100328. [Google Scholar] [CrossRef]
  141. de Andrade Dutra, K.; de Oliveira, J.V.; Navarro, D.M. do A.F.; Barbosa, D.R. e S.; Santos, J.P.O. Control of Callosobruchus maculatus (FABR.) (Coleoptera: Chrysomelidae: Bruchinae) in Vigna Unguiculata (L.) WALP. with Essential Oils from Four Citrus Spp. Plants. J. Stored Prod. Res. 2016, 68, 25–32. [Google Scholar] [CrossRef]
  142. Al-Sarar, A.S.; Hussein, H.I.; Abobakr, Y.; Al-Zabib, A.A.S.; Bazeyad, A.Y. Mosquitocidal and Repellent Activities of Essential Oils against Culex pipiens L. Entomol. Res. 2020, 50, 182–188. [Google Scholar] [CrossRef]
  143. Sheikh, Z.; Amani, A.; Basseri, H.R.; Kazemi, S.H.M.; Sedaghat, M.M.; Azam, K.; Azizi, M.; Amirmohammadi, F. Repellent Efficacy of Eucalyptus globulus and Syzygium aromaticum Essential Oils against Malaria Vector, Anopheles stephensi (Diptera: Culicidae). Iran. J. Public Health 2021, 50, 1668. [Google Scholar] [CrossRef]
  144. Akeumbiwo Tchumkam, C.; Kojom Foko, L.P.; Ndo, C.; Essangui Same, E.; Cheteug Nguetsa, G.; Eya’Ane Meva, F.; Ayong, L.; Eboumbou Moukoko, C.E. Chemical Composition and Repellent Activity of Essential Oils of Tithonia Diversifolia (Asteraceae) Leaves against the Bites of Anopheles coluzzii. Sci. Rep. 2023, 13, 6001. [Google Scholar] [CrossRef] [PubMed]
  145. Haris, A.; Azeem, M.; Abbas, M.G.; Mumtaz, M.; Mozūratis, R.; Binyameen, M. Prolonged Repellent Activity of Plant Essential Oils against Dengue Vector, Aedes aegypti. Molecules 2023, 28, 1351. [Google Scholar] [CrossRef] [PubMed]
  146. Catani, L.; Grassi, E.; Cocozza di Montanara, A.; Guidi, L.; Sandulli, R.; Manachini, B.; Semprucci, F. Essential Oils and Their Applications in Agriculture and Agricultural Products: A Literature Analysis through VOSviewer. Biocatal. Agric. Biotechnol. 2022, 45, 102502. [Google Scholar] [CrossRef]
  147. Danna, C.; Malaspina, P.; Cornara, L.; Smeriglio, A.; Trombetta, D.; De Feo, V.; Vanin, S. Eucalyptus Essential Oils in Pest Control: A Review of Chemical Composition and Applications against Insects and Mites. Crop Prot. 2024, 176, 106319. [Google Scholar] [CrossRef]
  148. Lacotte, V.; Rey, M.; Peignier, S.; Mercier, P.-E.; Rahioui, I.; Sivignon, C.; Razy, L.; Benhamou, S.; Livi, S.; da Silva, P. Bioactivity and Chemical Composition of Forty Plant Essential Oils against the Pea Aphid Acyrthosiphon pisum Revealed Peppermint Oil as a Promising Biorepellent. Ind. Crops Prod. 2023, 197, 116610. [Google Scholar] [CrossRef]
  149. Saıfı, R.; Saıfı, H.; Akca, İ.; Benabadelkader, M.; Askın, A.K.; Belghoul, M. Insecticidal and Repellent Effects of Mentha longifolia L. Essential Oil against Aphis craccivora Koch (Hemiptera: Aphididae). Chem. Biol. Technol. Agric. 2023, 10, 18. [Google Scholar] [CrossRef]
  150. Singh, D.; Singh, A.K. Repellent and Insecticidal Properties of Essential Oils against Housefly, Musca domestica L. Int. J. Trop. Insect Sci. 1991, 12, 487–491. [Google Scholar] [CrossRef]
  151. Choi, W.S.; Park, B.S.; Ku, S.K.; Lee, S.E. Repellent Activities of Essential Oils and Monoterpenes against Culex pipiens pallens. J. Am. Mosq. Control. Assoc. 2002, 18, 348–351. [Google Scholar] [PubMed]
  152. Park, B.-S.; Choi, W.-S.; Kim, J.-H.; Kim, K.-H.; Lee, S.-E. Monoterpenes from Thyme (Thymus vulgaris) as Potential Mosquito Repellents. J. Am. Mosq. Control. Assoc. 2005, 21, 80–83. [Google Scholar] [CrossRef]
  153. Odalo, J.O.; Omolo, M.O.; Malebo, H.; Angira, J.; Njeru, P.M.; Ndiege, I.O.; Hassanali, A. Repellency of Essential Oils of Some Plants from the Kenyan Coast against Anopheles gambiae. Acta Trop. 2005, 95, 210–218. [Google Scholar] [CrossRef]
  154. Erler, F.; Ulug, I.; Yalcinkaya, B. Repellent Activity of Five Essential Oils against Culex pipiens. Fitoterapia 2006, 77, 491–494. [Google Scholar] [CrossRef] [PubMed]
  155. Prajapati, V.; Tripathi, A.; Aggarwal, K.; Khanuja, S. Insecticidal, Repellent and Oviposition-Deterrent Activity of Selected Essential Oils against Anopheles stephensi, Aedes aegypti and Culex quinquefasciatus. Bioresour. Technol. 2005, 96, 1749–1757. [Google Scholar] [CrossRef] [PubMed]
  156. Hussein, H.S.; Salem, M.Z.M.; Soliman, A.M. Repellent, Attractive, and Insecticidal Effects of Essential Oils from Schinus terebinthifolius Fruits and Corymbia citriodora Leaves on Two Whitefly Species, Bemisia tabaci, and Trialeurodes ricini. Sci. Hortic. 2017, 216, 111–119. [Google Scholar] [CrossRef]
  157. Costa, E.C.C.; Christofoli, M.; de Souza Costa, G.C.; Peixoto, M.F.; Fernandes, J.B.; Forim, M.R.; de Castro Pereira, K.; Silva, F.G.; de Melo Cazal, C. Essential Oil Repellent Action of Plants of the Genus Zanthoxylum against Bemisia tabaci Biotype B (Homoptera: Aleyrodidae). Sci. Hortic. 2017, 226, 327–332. [Google Scholar] [CrossRef]
  158. Liu, C.H.; Mishra, A.K.; Tan, R.X.; Tang, C.; Yang, H.; Shen, Y.F. Repellent and Insecticidal Activities of Essential Oils from Artemisia princeps and Cinnamomum camphora and Their Effect on Seed Germination of Wheat and Broad Bean. Bioresour. Technol. 2006, 97, 1969–1973. [Google Scholar] [CrossRef]
  159. Reyes, E.I.M.; Farias, E.S.; Silva, E.M.P.; Filomeno, C.A.; Plata, M.A.B.; Picanço, M.C.; Barbosa, L.C.A. Eucalyptus resinifera Essential Oils Have Fumigant and Repellent Action against Hypothenemus hampei. Crop Prot. 2019, 116, 49–55. [Google Scholar] [CrossRef]
  160. Chiluwal, K.; Kim, J.; Bae, S.D.; Park, C.G. Essential Oils from Selected Wooden Species and Their Major Components as Repellents and Oviposition Deterrents of Callosobruchus chinensis (L.). J. Asia Pac. Entomol. 2017, 20, 1447–1453. [Google Scholar] [CrossRef]
  161. Taban, A.; Saharkhiz, M.J.; Khorram, M. Formulation and Assessment of Nano-Encapsulated Bioherbicides Based on Biopolymers and Essential Oil. Ind. Crops Prod. 2020, 149, 112348. [Google Scholar] [CrossRef]
  162. Parreira, D.S.; Alcántara-de la Cruz, R.; Dimaté, F.A.R.; Batista, L.D.; Ribeiro, R.C.; Ferreira, G.A.R.; Zanuncio, J.C. Bioactivity of Ten Essential Oils on the Biological Parameters of Trichogramma pretiosum (Hymenoptera: Trichogrammatidae) Adults. Ind. Crops Prod. 2019, 127, 11–15. [Google Scholar] [CrossRef]
  163. Baccari, W.; Znati, M.; Zardi-Bergaoui, A.; Chaieb, I.; Flamini, G.; Ascrizzi, R.; Ben Jannet, H. Composition and Insecticide Potential against Tribolium castaneum of the Fractionated Essential Oil from the Flowers of the Tunisian Endemic Plant Ferula Tunetana nomel Ex Batt. Ind. Crops Prod. 2020, 143, 111888. [Google Scholar] [CrossRef]
  164. Tak, J.-H.; Isman, M.B. Acaricidal and Repellent Activity of Plant Essential Oil-Derived Terpenes and the Effect of Binary Mixtures against Tetranychus urticae Koch (Acari: Tetranychidae). Ind. Crops Prod. 2017, 108, 786–792. [Google Scholar] [CrossRef]
  165. Isman, M.B. Plant Essential Oils for Pest and Disease Management. Crop Prot. 2000, 19, 603–608. [Google Scholar] [CrossRef]
  166. Sousa, R.M.O.F.; Rosa, J.S.; Oliveira, L.; Cunha, A.; Fernandes-Ferreira, M. Activities of Apiaceae Essential Oils against Armyworm, Pseudaletia unipuncta (Lepidoptera: Noctuidae). J. Agric. Food Chem. 2013, 61, 7661–7672. [Google Scholar] [CrossRef] [PubMed]
  167. Jaramillo-Colorado, B.E.; Pino-Benitez, N.; González-Coloma, A. Volatile Composition and Biocidal (Antifeedant and Phytotoxic) Activity of the Essential Oils of Four Piperaceae species from Choco-Colombia. Ind. Crops Prod. 2019, 138, 111463. [Google Scholar] [CrossRef]
  168. Formisano, C.; Rigano, D.; Senatore, F.; Arnold, N.A.; Simmonds, M.S.J.; Rosselli, S.; Bruno, M.; Loziene, K. Essential Oils of Three Species of Scutellaria and Their Influence on Spodoptera littoralis. Biochem. Syst. Ecol. 2013, 48, 206–210. [Google Scholar] [CrossRef]
  169. Santos, G.K.N.; Dutra, K.A.; Barros, R.A.; da Câmara, C.A.G.; Lira, D.D.; Gusmão, N.B.; Navarro, D.M.A.F. Essential Oils from Alpinia purpurata (Zingiberaceae): Chemical Composition, Oviposition Deterrence, Larvicidal and Antibacterial Activity. Ind. Crops Prod. 2012, 40, 254–260. [Google Scholar] [CrossRef]
  170. Soonwera, M.; Phasomkusolsil, S. Adulticidal, Larvicidal, Pupicidal and Oviposition Deterrent Activities of Essential Oil from Zanthoxylum limonella Alston (Rutaceae) against Aedes aegypti (L.) and Culex quinquefasciatus (Say). Asian Pac. J. Trop. Biomed. 2017, 7, 967–978. [Google Scholar] [CrossRef]
  171. Thanigaivel, A.; Chanthini, K.M.-P.; Karthi, S.; Vasantha-Srinivasan, P.; Ponsankar, A.; Sivanesh, H.; Stanley-Raja, V.; Shyam-Sundar, N.; Narayanan, K.R.; Senthil-Nathan, S. Toxic Effect of Essential Oil and Its Compounds Isolated from Sphaeranthus amaranthoides Burm. f. against Dengue Mosquito Vector Aedes aegypti Linn. Pestic. Biochem. Physiol. 2019, 160, 163–170. [Google Scholar] [CrossRef]
  172. Mbata, G.N.; Payton, M.E. Effect of Monoterpenoids on Oviposition and Mortality of Callosobruchus maculatus (F.) (Coleoptera: Bruchidae) under Hermetic Conditions. J. Stored Prod. Res. 2013, 53, 43–47. [Google Scholar] [CrossRef]
  173. Papanastasiou, S.A.; Ioannou, C.S.; Papadopoulos, N.T. Oviposition-deterrent Effect of Linalool—A Compound of Citrus Essential Oils—On Female Mediterranean Fruit Flies, Ceratitis capitata (Diptera: Tephritidae). Pest Manag. Sci. 2020, 76, 3066–3077. [Google Scholar] [CrossRef] [PubMed]
  174. dos Santos Silva, C.T.; Wanderley-Teixeira, V.; da Cunha, F.M.; de Oliveira, J.V.; de Andrade Dutra, K.; do Amaral Ferraz Navarro, D.M.; Teixeira, Á.A.C. Biochemical Parameters of Spodoptera frugiperda (J. E. Smith) Treated with Citronella Oil (Cymbopogon winterianus Jowitt Ex Bor) and Its Influence on Reproduction. Acta Histochem. 2016, 118, 347–352. [Google Scholar] [CrossRef] [PubMed]
  175. Ferreira, T.P.; Oliveira, E.E.; Tschoeke, P.H.; Pinheiro, R.G.; Maia, A.M.S.; Aguiar, R.W.S. Potential Use of Negramina (Siparuna guianensis Aubl.) Essential Oil to Control Wax Moths and Its Selectivity in Relation to Honey Bees. Ind. Crops Prod. 2017, 109, 151–157. [Google Scholar] [CrossRef]
  176. Iglesias, A.; Mitton, G.; Szawarski, N.; Cooley, H.; Ramos, F.; Meroi Arcerito, F.; Brasesco, C.; Ramirez, C.; Gende, L.; Eguaras, M.; et al. Essential Oils from Humulus lupulus as Novel Control Agents against Varroa destructor. Ind. Crops Prod. 2020, 158, 113043. [Google Scholar] [CrossRef]
  177. Rossini, C.; Rodrigo, F.; Davyt, B.; Umpiérrez, M.L.; González, A.; Garrido, P.M.; Cuniolo, A.; Porrini, L.P.; Eguaras, M.J.; Porrini, M.P. Sub-Lethal Effects of the Consumption of Eupatorium buniifolium Essential Oil in Honeybees. PLoS ONE 2020, 15, e0241666. [Google Scholar] [CrossRef]
  178. Sabahi, Q.; Hamiduzzaman, M.M.d.; Barajas-Pérez, J.S.; Tapia-Gonzalez, J.M.; Guzman-Novoa, E. Toxicity of Anethole and the Essential Oils of Lemongrass and Sweet Marigold to the Parasitic Mite Varroa destructor and Their Selectivity for Honey Bee (Apis mellifera) Workers and Larvae. Psyche A J. Entomol. 2018, 2018, 6196289. [Google Scholar] [CrossRef]
  179. Vasantha-Srinivasan, P.; Senthil-Nathan, S.; Ponsankar, A.; Thanigaivel, A.; Chellappandian, M.; Edwin, E.-S.; Selin-Rani, S.; Kalaivani, K.; Hunter, W.B.; Duraipandiyan, V.; et al. Acute Toxicity of Chemical Pesticides and Plant-Derived Essential Oil on the Behavior and Development of Earthworms, Eudrilus eugeniae (Kinberg) and Eisenia fetida (Savigny). Environ. Sci. Pollut. Res. 2018, 25, 10371–10382. [Google Scholar] [CrossRef]
  180. Pavela, R. Essential Oils from Foeniculum Vulgare Miller as a Safe Environmental Insecticide against the Aphid Myzus persicae Sulzer. Environ. Sci. Pollut. Res. 2018, 25, 10904–10910. [Google Scholar] [CrossRef]
  181. George, D.R.; Sparagano, O.A.E.; Port, G.; Okello, E.; Shiel, R.S.; Guy, J.H. Repellence of Plant Essential Oils to Dermanyssus gallinae and Toxicity to the Non-Target Invertebrate Tenebrio Molitor. Vet. Parasitol. 2009, 162, 129–134. [Google Scholar] [CrossRef] [PubMed]
  182. Pavela, R.; Benelli, G. Essential Oils as Ecofriendly Biopesticides? Challenges and Constraints. Trends Plant Sci. 2016, 21, 1000–1007. [Google Scholar] [CrossRef]
  183. Tabari, M.A.; Youssefi, M.R.; Maggi, F.; Benelli, G. Toxic and Repellent Activity of Selected Monoterpenoids (Thymol, Carvacrol and Linalool) against the Castor Bean Tick, Ixodes ricinus (Acari: Ixodidae). Vet. Parasitol. 2017, 245, 86–91. [Google Scholar] [CrossRef]
  184. Corrêa-Filho, L.C.; Lourenço, M.M.; Moldão-Martins, M.; Alves, V.D. Microencapsulation of β-Carotene by Spray Drying: Effect of Wall Material Concentration and Drying Inlet Temperature. Int. J. Food Sci. 2019, 2019, 8914852. [Google Scholar] [CrossRef] [PubMed]
  185. Locali-Pereira, A.R.; Lopes, N.A.; Menis-Henrique, M.E.C.; Janzantti, N.S.; Nicoletti, V.R. Modulation of Volatile Release and Antimicrobial Properties of Pink Pepper Essential Oil by Microencapsulation in Single- and Double-Layer Structured Matrices. Int. J. Food Microbiol. 2020, 335, 108890. [Google Scholar] [CrossRef]
  186. Yammine, J.; Chihib, N.-E.; Gharsallaoui, A.; Ismail, A.; Karam, L. Advances in Essential Oils Encapsulation: Development, Characterization and Release Mechanisms. Polym. Bull. 2024, 81, 3837–3882. [Google Scholar] [CrossRef]
  187. Maes, C.; Bouquillon, S.; Fauconnier, M.-L. Encapsulation of Essential Oils for the Development of Biosourced Pesticides with Controlled Release: A Review. Molecules 2019, 24, 2539. [Google Scholar] [CrossRef]
  188. Naseema, A.; Kovooru, L.; Behera, A.K.; Kumar, K.P.P.; Srivastava, P. A Critical Review of Synthesis Procedures, Applications and Future Potential of Nanoemulsions. Adv. Colloid Interface Sci. 2021, 287, 102318. [Google Scholar] [CrossRef]
  189. Simonazzi, A.; Cid, A.G.; Villegas, M.; Romero, A.I.; Palma, S.D.; Bermúdez, J.M. Nanotechnology Applications in Drug Controlled Release. In Drug Targeting and Stimuli Sensitive Drug Delivery Systems; Elsevier: Amsterdam, The Netherlands, 2018; pp. 81–116. [Google Scholar]
  190. Ruiz-Montañez, G.; Ragazzo-Sanchez, J.A.; Picart-Palmade, L.; Calderón-Santoyo, M.; Chevalier-Lucia, D. Optimization of Nanoemulsions Processed by High-Pressure Homogenization to Protect a Bioactive Extract of Jackfruit (Artocarpus heterophyllus Lam). Innov. Food Sci. Emerg. Technol. 2017, 40, 35–41. [Google Scholar] [CrossRef]
  191. Li, Y.; Wu, C.-L.; Liu, J.; Zhu, Y.; Zhang, X.-Y.; Jiang, L.-Z.; Qi, B.-K.; Zhang, X.-N.; Wang, Z.-J.; Teng, F. Soy Protein Isolate-Phosphatidylcholine Nanoemulsions Prepared Using High-Pressure Homogenization. Nanomaterials 2018, 8, 307. [Google Scholar] [CrossRef] [PubMed]
  192. Bai, L.; Huan, S.; Gu, J.; McClements, D.J. Fabrication of Oil-in-Water Nanoemulsions by Dual-Channel Microfluidization Using Natural Emulsifiers: Saponins, Phospholipids, Proteins, and Polysaccharides. Food Hydrocoll. 2016, 61, 703–711. [Google Scholar] [CrossRef]
  193. Perazzo, A.; Preziosi, V.; Guido, S. Phase Inversion Emulsification: Current Understanding and Applications. Adv. Colloid Interface Sci. 2015, 222, 581–599. [Google Scholar] [CrossRef] [PubMed]
  194. Chircov, C.; Grumezescu, A.M. Nanoemulsion Preparation, Characterization, and Application in the Field of Biomedicine. In Nanoarchitectonics in Biomedicine; Elsevier: Amsterdam, The Netherlands, 2019; pp. 169–188. [Google Scholar]
  195. Ren, G.; Sun, Z.; Wang, Z.; Zheng, X.; Xu, Z.; Sun, D. Nanoemulsion Formation by the Phase Inversion Temperature Method Using Polyoxypropylene Surfactants. J. Colloid Interface Sci. 2019, 540, 177–184. [Google Scholar] [CrossRef]
  196. Donsì, F.; Annunziata, M.; Sessa, M.; Ferrari, G. Nanoencapsulation of Essential Oils to Enhance Their Antimicrobial Activity in Foods. LWT-Food Sci. Technol. 2011, 44, 1908–1914. [Google Scholar] [CrossRef]
  197. Al-Tayyar, N.A.; Youssef, A.M.; Al-Hindi, R.R. Edible Coatings and Antimicrobial Nanoemulsions for Enhancing Shelf Life and Reducing Foodborne Pathogens of Fruits and Vegetables: A review. Sustain. Mater. Technol. 2020, 26, e00215. [Google Scholar] [CrossRef]
  198. Xiong, Y.; Li, S.; Warner, R.D.; Fang, Z. Effect of Oregano Essential Oil and Resveratrol Nanoemulsion Loaded Pectin Edible Coating on the Preservation of Pork Loin in Modified Atmosphere Packaging. Food Control. 2020, 114, 107226. [Google Scholar] [CrossRef]
  199. Fan, W.; Yan, W.; Xu, Z.; Ni, H. Formation Mechanism of Monodisperse, Low Molecular Weight Chitosan Nanoparticles by Ionic Gelation Technique. Colloids Surf. B Biointerfaces 2012, 90, 21–27. [Google Scholar] [CrossRef] [PubMed]
  200. Pant, A.; Negi, J.S. Novel Controlled Ionic Gelation Strategy for Chitosan Nanoparticles Preparation Using TPP-β-CD Inclusion Complex. Eur. J. Pharm. Sci. 2018, 112, 180–185. [Google Scholar] [CrossRef] [PubMed]
  201. Calvo, P.; Remuñán-López, C.; Vila-Jato, J.L.; Alonso, M.J. Novel Hydrophilic Chitosan-Polyethylene Oxide Nanoparticles as Protein Carriers. J. Appl. Polym. Sci. 1997, 63, 125–132. [Google Scholar] [CrossRef]
  202. Hsieh, W.-C.; Chang, C.-P.; Gao, Y.-L. Controlled Release Properties of Chitosan Encapsulated Volatile Citronella Oil Microcapsules by Thermal Treatments. Colloids Surf. B Biointerfaces 2006, 53, 209–214. [Google Scholar] [CrossRef]
  203. Giri, T.K. Nanoarchitectured Polysaccharide-Based Drug Carrier for Ocular Therapeutics. In Nanoarchitectonics for Smart De-livery and Drug Targeting; Holban, A.M., Grumezescu, A.M., Eds.; William Andrew Publishing: New York, NY, USA, 2016; Volume 1, pp. 119–141. [Google Scholar]
  204. 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]
  205. Carvalho, E.L.S.; Grenha, A.; Remuñán-López, C.; Alonso, M.J.; Seijo, B. Chapter 15 Mucosal Delivery of Liposome–Chitosan Nanoparticle Complexes. Methods Enzymol. 2009, 465, 289–312. [Google Scholar] [PubMed]
  206. Vaezifar, S.; Razavi, S.; Golozar, M.A.; Karbasi, S.; Morshed, M.; Kamali, M. Effects of Some Parameters on Particle Size Distribution of Chitosan Nanoparticles Prepared by Ionic Gelation Method. J. Clust. Sci. 2013, 24, 891–903. [Google Scholar] [CrossRef]
  207. Hosseini, S.M.; Hosseini, H.; Mohammadifar, M.A.; Mortazavian, A.M.; Mohammadi, A.; Khosravi-Darani, K.; Shojaee-Alibadi, S.; Dehghan, S.; Khaksar, R. Incorporation of Essential Oil in Alginate Microparticles by Multiple Emulsion/Ionic Gelation Process. Int. J. Biol. Macromol. 2013, 62, 582–588. [Google Scholar] [CrossRef]
  208. Zhang, K.; Zhang, H.; Hu, X.; Bao, S.; Huang, H. Synthesis and Release Studies of Microalgal Oil-Containing Microcapsules Prepared by Complex Coacervation. Colloids Surf. B Biointerfaces 2012, 89, 61–66. [Google Scholar] [CrossRef] [PubMed]
  209. Hernández-Nava, R.; López-Malo, A.; Palou, E.; Ramírez-Corona, N.; Jiménez-Munguía, M.T. Encapsulation of Oregano Essential Oil (Origanum vulgare) by Complex Coacervation between Gelatin and Chia Mucilage and Its Properties after Spray Drying. Food Hydrocoll. 2020, 109, 106077. [Google Scholar] [CrossRef]
  210. Bastos, L.P.H.; Vicente, J.; dos Santos, C.H.C.; de 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]
  211. Chang, C.-P.; Leung, T.-K.; Lin, S.-M.; Hsu, C.-C. Release Properties on Gelatin-Gum Arabic Microcapsules Containing Camphor Oil with Added Polystyrene. Colloids Surf. B Biointerfaces 2006, 50, 136–140. [Google Scholar] [CrossRef] [PubMed]
  212. Shaddel, R.; Hesari, J.; Azadmard-Damirchi, S.; Hamishehkar, H.; Fathi-Achachlouei, B.; Huang, Q. Use of Gelatin and Gum Arabic for Encapsulation of Black Raspberry Anthocyanins by Complex Coacervation. Int. J. Biol. Macromol. 2018, 107, 1800–1810. [Google Scholar] [CrossRef] [PubMed]
  213. Ferreira, S.; Nicoletti, V.R. Complex Coacervation Assisted by a Two-Fluid Nozzle for Microencapsulation of Ginger Oil: Effect of Atomization Parameters. Food Res. Int. 2020, 138, 109828. [Google Scholar] [CrossRef] [PubMed]
  214. Saravanan, M.; Rao, K.P. Pectin–Gelatin and Alginate–Gelatin Complex Coacervation for Controlled Drug Delivery: Influence of Anionic Polysaccharides and Drugs Being Encapsulated on Physicochemical Properties of Microcapsules. Carbohydr. Polym. 2010, 80, 808–816. [Google Scholar] [CrossRef]
  215. de Matos, E.F.; Scopel, B.S.; Dettmer, A. Citronella Essential Oil Microencapsulation by Complex Coacervation with Leather Waste Gelatin and Sodium Alginate. J. Environ. Chem. Eng. 2018, 6, 1989–1994. [Google Scholar] [CrossRef]
  216. Mohseni, F.; Goli, S.A.H. Encapsulation of Flaxseed Oil in the Tertiary Conjugate of Oxidized Tannic Acid-Gelatin and Flaxseed (Linum usitatissimum) Mucilage. Int. J. Biol. Macromol. 2019, 140, 959–964. [Google Scholar] [CrossRef] [PubMed]
  217. Abreu, F.O.M.S.; de Oliveira, E.F.; Paula, H.C.B.; de Paula, R.C.M. Chitosan/Cashew Gum Nanogels for Essential Oil Encapsulation. Carbohydr. Polym. 2012, 89, 1277–1282. [Google Scholar] [CrossRef]
  218. de Oliveira, W.Q.; Wurlitzer, N.J.; de Oliveira Araújo, A.W.; Comunian, T.A.; do Socorro Rocha Bastos, M.; de Oliveira, A.L.; Magalhães, H.C.R.; Ribeiro, H.L.; de Figueiredo, R.W.; de Sousa, P.H.M. Complex Coacervates of Cashew Gum and Gelatin as Carriers of Green Coffee Oil: The Effect of Microcapsule Application on the Rheological and Sensorial Quality of a Fruit Juice. Food Res. Int. 2020, 131, 109047. [Google Scholar] [CrossRef] [PubMed]
  219. Nakagawa, K.; Nagao, H. Microencapsulation of Oil Droplets Using Freezing-Induced Gelatin–Acacia Complex Coacervation. Colloids Surf. A Physicochem. Eng. Asp. 2012, 411, 129–139. [Google Scholar] [CrossRef]
  220. Aloys, H.; Korma, S.A.; Alice, T.M.; Chantal, N.; Ali, A.H.; Abed, S.M. Microencapsulation by Complex Coacervation: Methods, Techniques, Benefits, and Applications—A Review. Am. J. Food Sci. Nut. Res. 2016, 1, 188–192. [Google Scholar]
  221. Timilsena, Y.P.; Akanbi, T.O.; Khalid, N.; Adhikari, B.; Barrow, C.J. Complex Coacervation: Principles, Mechanisms and Applications in Microencapsulation. Int. J. Biol. Macromol. 2019, 121, 1276–1286. [Google Scholar] [CrossRef] [PubMed]
  222. 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. Biol. Macromol. 2019, 133, 294–305. [Google Scholar] [CrossRef] [PubMed]
  223. 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] [PubMed]
  224. Mohammadi, A.; Hosseini, S.M.; Hashemi, M. Emerging Chitosan Nanoparticles Loading-System Boosted the Antibacterial Activity of Cinnamomum zeylanicum Essential Oil. Ind. Crops Prod. 2020, 155, 112824. [Google Scholar] [CrossRef]
  225. Rajkumar, V.; Gunasekaran, C.; Paul, C.A.; Dharmaraj, J. Development of Encapsulated Peppermint Essential Oil in Chitosan Nanoparticles: Characterization and Biological Efficacy against Stored-Grain Pest Control. Pestic. Biochem. Physiol. 2020, 170, 104679. [Google Scholar] [CrossRef] [PubMed]
  226. Rajkumar, V.; Gunasekaran, C.; Dharmaraj, J.; Chinnaraj, P.; Paul, C.A.; Kanithachristy, I. Structural Characterization of Chitosan Nanoparticle Loaded with Piper nigrum Essential Oil for Biological Efficacy against the Stored Grain Pest Control. Pestic. Biochem. Physiol. 2020, 166, 104566. [Google Scholar] [CrossRef] [PubMed]
  227. 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]
  228. Zhaveh, S.; Mohsenifar, A.; Beiki, M.; Khalili, S.T.; Abdollahi, A.; Rahmani-Cherati, T.; Tabatabaei, M. Encapsulation of Cuminum cyminum Essential Oils in Chitosan-Caffeic Acid Nanogel with Enhanced Antimicrobial Activity against Aspergillus flavus. Ind. Crops Prod. 2015, 69, 251–256. [Google Scholar] [CrossRef]
  229. Beyki, M.; Zhaveh, S.; Khalili, S.T.; Rahmani-Cherati, T.; Abollahi, A.; Bayat, M.; Tabatabaei, M.; Mohsenifar, A. Encapsulation of Mentha Piperita Essential Oils in Chitosan–Cinnamic Acid Nanogel with Enhanced Antimicrobial Activity against Aspergillus flavus. Ind. Crops Prod. 2014, 54, 310–319. [Google Scholar] [CrossRef]
  230. 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]
  231. Tao, F.; Hill, L.E.; Peng, Y.; Gomes, C.L. Synthesis and Characterization of β-Cyclodextrin Inclusion Complexes of Thymol and Thyme oil for Antimicrobial Delivery Applications. LWT—Food Sci. Technol. 2014, 59, 247–255. [Google Scholar] [CrossRef]
  232. Anaya-Castro, M.A.; Ayala-Zavala, J.F.; Muñoz-Castellanos, L.; Hernández-Ochoa, L.; Peydecastaing, J.; Durrieu, V. β-Cyclodextrin Inclusion Complexes Containing Clove (Eugenia caryophyllata) and Mexican oregano (Lippia berlandieri) Essential Oils: Preparation, Physicochemical and Antimicrobial Characterization. Food Packag. Shelf Life 2017, 14, 96–101. [Google Scholar] [CrossRef]
  233. Rakmai, J.; Cheirsilp, B.; Mejuto, J.C.; Torrado-Agrasar, A.; Simal-Gándara, J. Physico-Chemical Characterization and Evaluation of Bio-Efficacies of Black Pepper Essential Oil Encapsulated in Hydroxypropyl-Beta-Cyclodextrin. Food Hydrocoll. 2017, 65, 157–164. [Google Scholar] [CrossRef]
  234. Barbieri, N.; Sanchez-Contreras, A.; Canto, A.; Cauich-Rodriguez, J.V.; Vargas-Coronado, R.; Calvo-Irabien, L.M. Effect of Cyclodextrins and Mexican Oregano (Lippia graveolens Kunth) Chemotypes on the Microencapsulation of Essential Oil. Ind. Crops Prod. 2018, 121, 114–123. [Google Scholar] [CrossRef]
  235. Maia, J.D.; La Corte, R.; Martinez, J.; Ubbink, J.; Prata, A.S. Improved Activity of Thyme Essential Oil (Thymus vulgaris) against Aedes aegypti Larvae Using a Biodegradable Controlled Release System. Ind. Crops Prod. 2019, 136, 110–120. [Google Scholar] [CrossRef]
  236. Saraiva, N.R.; Roncato, J.F.F.; Pascoli, M.; e Sousa, J.M.F.M.; Windberg, L.F.; Rossatto, F.C.P.; de Jesus Soares, J.; Denardin, E.L.G.; Puntel, R.L.; Zimmer, K.R.; et al. Clove Oil-Loaded Zein Nanoparticles as Potential Bioinsecticide Agent with Low Toxicity. Sustain. Chem. Pharm. 2021, 24, 100554. [Google Scholar] [CrossRef]
  237. López, A.; Castro, S.; Andina, M.J.; Ures, X.; Munguía, B.; Llabot, J.M.; Elder, H.; Dellacassa, E.; Palma, S.; Domínguez, L. Insecticidal Activity of Microencapsulated Schinus molle Essential Oil. Ind. Crops Prod. 2014, 53, 209–216. [Google Scholar] [CrossRef]
  238. Radünz, M.; dos Santos Hackbart, H.C.; Camargo, T.M.; Nunes, C.F.P.; de Barros, F.A.P.; Dal Magro, J.; Filho, P.J.S.; Gandra, E.A.; Radünz, A.L.; da Rosa Zavareze, E. Antimicrobial Potential of Spray Drying Encapsulated Thyme (Thymus vulgaris) Essential Oil on the Conservation of Hamburger-like Meat Products. Int. J. Food Microbiol. 2020, 330, 108696. [Google Scholar] [CrossRef]
  239. Rodea-González, D.A.; Cruz-Olivares, J.; Román-Guerrero, A.; Rodríguez-Huezo, M.E.; Vernon-Carter, E.J.; Pérez-Alonso, C. Spray-Dried Encapsulation of Chia Essential Oil (Salvia hispanica L.) in Whey Protein Concentrate-Polysaccharide Matrices. J. Food Eng. 2012, 111, 102–109. [Google Scholar] [CrossRef]
  240. Sugumar, S.; Ghosh, V.; Nirmala, M.J.; Mukherjee, A.; Chandrasekaran, N. Ultrasonic Emulsification of Eucalyptus Oil Nanoemulsion: Antibacterial Activity against Staphylococcus aureus and Wound Healing Activity in Wistar Rats. Ultrason. Sonochem. 2014, 21, 1044–1049. [Google Scholar] [CrossRef] [PubMed]
  241. Suresh, U.; Murugan, K.; Panneerselvam, C.; Aziz, A.T.; Cianfaglione, K.; Wang, L.; Maggi, F. Encapsulation of Sea Fennel (Crithmum maritimum) Essential Oil in Nanoemulsion and SiO2 Nanoparticles for Treatment of the Crop Pest Spodoptera litura and the Dengue Vector Aedes aegypti. Ind. Crops Prod. 2020, 158, 113033. [Google Scholar] [CrossRef]
  242. Ahsaei, S.M.; Rodríguez-Rojo, S.; Salgado, M.; Cocero, M.J.; Talebi-Jahromi, K.; Amoabediny, G. Insecticidal Activity of Spray Dried Microencapsulated Essential Oils of Rosmarinus officinalis and Zataria multiflora against Tribolium Confusum. Crop Prot. 2020, 128, 104996. [Google Scholar] [CrossRef]
  243. 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 Spodopteral littoralis. Ind. Crops Prod. 2022, 184, 115089. [Google Scholar] [CrossRef]
  244. Barros, J.M.H.F.; Santos, A.A.; Stadnik, M.J.; da Costa, C. Encapsulation of Eucalyptus and Litsea cubeba Essential Oils Using Zein Nanopolymer: Preparation, Characterization, Storage Stability, and Antifungal Evaluation. Int. J. Biol. Macromol. 2024, 278, 134690. [Google Scholar] [CrossRef] [PubMed]
  245. United States Department of Agriculture, USDA. The National List of Allowed and Prohibited Substances. Available online: https://www.ams.usda.gov/rules-regulations/national-list-allowed-and-prohibited-substances (accessed on 9 November 2020).
  246. Khayrova, A.; Lopatin, S.; Varlamov, V. Obtaining Chitin, Chitosan and Their Melanin Complexes from Insects. Int. J. Biol. Macromol. 2021, 167, 1319–1328. [Google Scholar] [CrossRef] [PubMed]
  247. Pal, P.; Pal, A.; Nakashima, K.; Yadav, B.K. Applications of Chitosan in Environmental Remediation: A review. Chemosphere 2021, 266, 128934. [Google Scholar] [CrossRef]
  248. Rashki, S.; Asgarpour, K.; Tarrahimofrad, H.; Hashemipour, M.; Ebrahimi, M.S.; Fathizadeh, H.; Khorshidi, A.; Khan, H.; Marzhoseyni, Z.; Salavati-Niasari, M.; et al. Chitosan-Based Nanoparticles against Bacterial Infections. Carbohydr. Polym. 2021, 251, 117108. [Google Scholar] [CrossRef] [PubMed]
  249. Dong, Z.-J.; Xia, S.-Q.; Hua, S.; Hayat, K.; Zhang, X.-M.; Xu, S.-Y. Optimization of Cross-Linking Parameters during Production of Transglutaminase-Hardened Spherical Multinuclear Microcapsules by Complex Coacervation. Colloids Surf. B Biointerfaces 2008, 63, 41–47. [Google Scholar] [CrossRef]
  250. Lv, Y.; Yang, F.; Li, X.; Zhang, X.; Abbas, S. Formation of Heat-Resistant Nanocapsules of Jasmine Essential Oil via Gelatin/Gum Arabic Based Complex Coacervation. Food Hydrocoll. 2014, 35, 305–314. [Google Scholar] [CrossRef]
  251. Prata, A.S.; Zanin, M.H.A.; Ré, M.I.; Grosso, C.R.F. Release Properties of Chemical and Enzymatic Crosslinked Gelatin-Gum Arabic Microparticles Containing a Fluorescent Probe Plus Vetiver Essential Oil. Colloids Surf. B Biointerfaces 2008, 67, 171–178. [Google Scholar] [CrossRef] [PubMed]
  252. Dong, Z.; Ma, Y.; Hayat, K.; Jia, C.; Xia, S.; Zhang, X. Morphology and Release Profile of Microcapsules Encapsulating Peppermint Oil by Complex Coacervation. J. Food Eng. 2011, 104, 455–460. [Google Scholar] [CrossRef]
  253. Costa, P.; Medronho, B.; Gonçalves, S.; Romano, A. Cyclodextrins Enhance the Antioxidant Activity of Essential Oils from Three Lamiaceae species. Ind. Crops Prod. 2015, 70, 341–346. [Google Scholar] [CrossRef]
  254. da Rocha Neto, A.C.; da Rocha, A.B.D.O.; Maraschin, M.; Di Piero, R.M.; Almenar, E. Factors Affecting the Entrapment Efficiency of β-Cyclodextrins and Their Effects on the Formation of Inclusion Complexes Containing Essential Oils. Food Hydrocoll. 2018, 77, 509–523. [Google Scholar] [CrossRef]
  255. Shishir, M.R.I.; Xie, L.; Sun, C.; Zheng, X.; Chen, W. Advances in Micro and Nano-Encapsulation of Bioactive Compounds Using Biopolymer and Lipid-Based Transporters. Trends Food Sci. Technol. 2018, 78, 34–60. [Google Scholar] [CrossRef]
  256. Tarhini, M.; Greige-Gerges, H.; Elaissari, A. Protein-based Nanoparticles: From Preparation to Encapsulation of Active Molecules. Int. J. Pharm. 2017, 522, 172–197. [Google Scholar] [CrossRef] [PubMed]
  257. Pateiro, M.; Gómez, B.; Munekata, P.E.S.; Barba, F.J.; Putnik, P.; Kovačević, D.B.; Lorenzo, J.M. Nanoencapsulation of Promising Bioactive Compounds to Improve Their Absorption, Stability, Functionality and the Appearance of the Final Food Products. Molecules 2021, 26, 1547. [Google Scholar] [CrossRef] [PubMed]
  258. Ghasemi, S.; Jafari, S.M.; Assadpour, E.; Khomeiri, M. Production of Pectin-Whey Protein Nano-Complexes as Carriers of Orange Peel Oil. Carbohydr. Polym. 2017, 177, 369–377. [Google Scholar] [CrossRef] [PubMed]
  259. Weisany, W.; Yousefi, S.; Tahir, N.A.; Golestanehzadeh, N.; McClements, D.J.; Adhikari, B.; Ghasemlou, M. Targeted Delivery and Controlled Released of Essential Oils Using Nanoencapsulation: A Review. Adv. Colloid Interface Sci. 2022, 303, 102655. [Google Scholar] [CrossRef] [PubMed]
  260. Marques, C.S.; Carvalho, S.G.; Bertoli, L.D.; Villanova, J.C.O.; Pinheiro, P.F.; dos Santos, D.C.M.; Yoshida, M.I.; de Freitas, J.C.C.; Cipriano, D.F.; Bernardes, P.C. β-Cyclodextrin Inclusion Complexes with Essential Oils: Obtention, Characterization, Antimicrobial Activity and Potential Application for Food Preservative Sachets. Food Res. Int. 2019, 119, 499–509. [Google Scholar] [CrossRef]
  261. Hill, L.E.; Gomes, C.; Taylor, T.M. Characterization of Beta-Cyclodextrin Inclusion Complexes Containing Essential Oils (Trans-Cinnamaldehyde, Eugenol, Cinnamon Bark, and Clove Bud Extracts) for Antimicrobial Delivery Applications. LWT—Food Sci. Technol. 2013, 51, 86–93. [Google Scholar] [CrossRef]
Figure 2. Nanoemulsion by the phase inversion method. Nanoemulsions are nanosized isotropic systems of two immiscible liquids, oily and aqueous, combined in a single and stable formulation.
Figure 2. Nanoemulsion by the phase inversion method. Nanoemulsions are nanosized isotropic systems of two immiscible liquids, oily and aqueous, combined in a single and stable formulation.
Agriculture 14 01766 g002
Figure 3. Generation of nanoparticles by ionic gelation. Ionic gelation is a technique based on ionic interactions between oppositely charged groups: a cationic polymer and an anionic molecule.
Figure 3. Generation of nanoparticles by ionic gelation. Ionic gelation is a technique based on ionic interactions between oppositely charged groups: a cationic polymer and an anionic molecule.
Agriculture 14 01766 g003
Figure 4. Formation of nanoparticles by complex coacervation. Complex coacervation is a technique based on the electrostatic interaction between oppositely charged polymers present in two separate immiscible phases: one that is polymer-rich used to coat the particle, and another that is polymer-poor.
Figure 4. Formation of nanoparticles by complex coacervation. Complex coacervation is a technique based on the electrostatic interaction between oppositely charged polymers present in two separate immiscible phases: one that is polymer-rich used to coat the particle, and another that is polymer-poor.
Agriculture 14 01766 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ayllón-Gutiérrez, R.; Díaz-Rubio, L.; Montaño-Soto, M.; Haro-Vázquez, M.d.P.; Córdova-Guerrero, I. Applications of Plant Essential Oils in Pest Control and Their Encapsulation for Controlled Release: A Review. Agriculture 2024, 14, 1766. https://doi.org/10.3390/agriculture14101766

AMA Style

Ayllón-Gutiérrez R, Díaz-Rubio L, Montaño-Soto M, Haro-Vázquez MdP, Córdova-Guerrero I. Applications of Plant Essential Oils in Pest Control and Their Encapsulation for Controlled Release: A Review. Agriculture. 2024; 14(10):1766. https://doi.org/10.3390/agriculture14101766

Chicago/Turabian Style

Ayllón-Gutiérrez, Rocío, Laura Díaz-Rubio, Myriam Montaño-Soto, María del Pilar Haro-Vázquez, and Iván Córdova-Guerrero. 2024. "Applications of Plant Essential Oils in Pest Control and Their Encapsulation for Controlled Release: A Review" Agriculture 14, no. 10: 1766. https://doi.org/10.3390/agriculture14101766

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