Next Article in Journal
Simple and Accurate Exchange Energy for Density Functional Theory
Next Article in Special Issue
Novel Methodologies for Chemical Activation in Organic Synthesis under Solvent-Free Reaction Conditions
Previous Article in Journal
Synthetic Life with Alternative Nucleic Acids as Genetic Materials
Previous Article in Special Issue
2-Methyloxolane (2-MeOx) as Sustainable Lipophilic Solvent to Substitute Hexane for Green Extraction of Natural Products. Properties, Applications, and Perspectives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plant Natural Products for the Control of Aedes aegypti: The Main Vector of Important Arboviruses

by
Maíra Rosato Silveiral Silvério
1,
Laila Salmen Espindola
2,
Norberto Peporine Lopes
1,* and
Paulo Cézar Vieira
1,*
1
NPPNS, Departamento de Ciências BioMoleculares, Faculdade de Ciências Farmacêuticas de Ribeirão Preto, Universidade de São Paulo, Ribeirão Preto 14040-903, São Paulo, Brazil
2
Laboratório de Farmacognosia, Universidade de Brasília, Brasília 70910-900, Brazil
*
Authors to whom correspondence should be addressed.
Submission received: 29 June 2020 / Revised: 24 July 2020 / Accepted: 28 July 2020 / Published: 31 July 2020

Abstract

:
The mosquito species Aedes aegypti is one of the main vectors of arboviruses, including dengue, Zika and chikungunya. Considering the deficiency or absence of vaccines to prevent these diseases, vector control remains an important strategy. The use of plant natural product-based insecticides constitutes an alternative to chemical insecticides as they are degraded more easily and are less harmful to the environment, not to mention their lower toxicity to non-target insects. This review details plant species and their secondary metabolites that have demonstrated insecticidal properties (ovicidal, larvicidal, pupicidal, adulticidal, repellent and ovipositional effects) against the mosquito, together with their mechanisms of action. In particular, essential oils and some of their chemical constituents such as terpenoids and phenylpropanoids offer distinct advantages. Thiophenes, amides and alkaloids also possess high larvicidal and adulticidal activities, adding to the wealth of plant natural products with potential in vector control applications.

Graphical Abstract

1. Introduction

The mosquito Aedes aegypti (Diptera: Culicidae) originated in Egypt and it is widely distributed in tropical and subtropical regions, including North America and Europe [1,2]. Ae. aegypti presents complete metamorphosis from immature egg, larva and pupa stages to the adult mosquito itself (Figure 1). The life cycle varies according to environmental temperature, food availability and quantity of larvae in the same breeding site. Under favorable conditions, after egg hatching, the mosquito transforms into the adult stage within 10 days, even though the eggs can be viable up to 450 days in the absence of water [3].
The female mosquito requires hematophagy for egg maturation. Viral transmission to humans occurs during this process if the mosquito is infected. The lifetime of a female mosquito is approximately 45 days [1]. Ae. aegypti population control is considered the principal measure to combat arboviral diseases as this species is the primary vector of dengue, Zika, chikungunya and urban yellow fever [4,5].
In 2012, dengue was considered the mosquito-borne disease of major importance in the world [5]. According to the World Health Organization, 390 million people are infected annually with the dengue virus, 96 million of which have clinical manifestations [6]. There are various symptoms, the first is usually high fever (39–40 °C) with headache, prostration, arthralgia, anorexia, asthenia, nausea, among others. Some clinical aspects often depend on patient age. There is no specific treatment for dengue and the more complicated cases of the disease can cause hemorrhage, shock and even death [7].
Ae. aegypti is considered the main Zika virus vector (Figure 1), but infection can also occur by sexual transmission or blood transfusion [8]. Symptoms are non-specific and self-limited, being easily confused with other arboviral diseases. Some important complications exist, such as microcephaly in fetuses and Guillan Barré syndrome. The virus has been reported in countries in Americas, Europe, Asia and the Pacific region [8]. Like dengue and Zika, chikungunya has no specific treatment. The disease emerged in the Americas in 2013, with about 1.7 million cases identified and 252 deaths reported by August 2015 [9].
In 2013, dengue generated a global cost of US $8.9 billion, with around 58.4 million symptomatic cases (13.5 million fatalities) in the 141 countries and territories. The per capita costs of dengue were $70.1 for hospital treatment, $51.1 for outpatient treatment and $12.9 for cases that did not reach the health system. According to this study, Brazil had an incidence of 751 to 1,000 cases per 100,000 people. The expenses were proportional to the incidence and ranged from $2.5 to $5 for each treated case [10].
A more recent study showed that in 2016 the Brazilian government spent around R $805 million (ca. 160 million US$) to treat diseases caused by the Ae. aegypti mosquito, including direct medical expenses and indirect costs. In addition, about R $1.5 billion were destined to combat the vector, totaling R $2.3 billion, that was 2% of the health budget for that year. More than 2 million cases of Ae. aegypti related diseases were verified. These numbers were underestimated as they did not include complications such as microcephaly and Guillain-Barré syndrome [11].
A study estimated that about 60% of the world population will be at risk of dengue in 2080, which represents over 6.1 (4.7–6.9) billion people [12]. Considering that vector control is the main tool for controlling these expensive arboviruses, investment in techniques to combat the Ae. aegypti mosquito is growing [3]. Investments are particularly focused on techniques with minimal negative impacts on non-target animals and the environment [4,13].

2. Mosquito Control

There are several techniques already used to combat mosquitoes, which act both in the immature phases (egg, larva and pupa) and in the adult [13]. Highly toxic synthetic insecticides such as organophosphates, pyrethroids and carbamates have been historically used to combat the mosquito, acting mainly on insect larvae [4]. More recently, insecticides with less toxicity which are less persistent in the environment have been developed, including neonicotinoids and oxadiazines [14]. However, these products are still widely used and harmful to living organisms and the environment, and the use of foogers and aerial applications of sintetic insecticides against adults, such as pyrethroids products, contributes to insect resistance problems [4]. Therefore, efforts must be made to ensure newly developed alternative insecticides are more eco-friendly.
Biological tools that control the adult stage are based on behavior, such as the Sterile Insect Technique (SIT), Incompatible Insect Technique (IIT) and Release of Insects carrying a Dominant Lethal gene (RIDL), which involve insect sterilization by chemical irradiation, natural bacteria which are pathogenic for mosquito (highly specific strains of Wolbachia) and genetic modifications to make sterile male mosquitoes, respectively [15,16]. The other technique applied against the adult stage is the use of entomopathogenic fungi, specifically in the orders Entomophthorales, Hypocreales and Pezizales due to their specificity, ability to manipulate and infectiveness to the host [17].
Ae. aegypti control using specific strains of Wolbachia bacteria is currently practiced in different locations around the world through the World Mosquito Program. This program involves the application of the bacteria to laboratory mosquitoes that are released into the local Ae. aegypti population during reproduction. The presence of bacteria in mosquitoes decreases the possibility of arbovirus transmission to people [18].
Biological control tools that act against the immature stages include the application of Bacillus thuringiensis in larvae habitats; products that prevent oviposition and/or inhibit growth and reproduction, including pheromones. There are also natural predators such as fish (especially of the genus Gambusia and Poecilia, family Poeciliidae) [19,20], copepods (including several species of the genus Mesocyclops) [21,22] and the “elephant mosquito” (genus Toxorhynchites) [23,24].
Finally, plant-based insecticides (ovicides, larvicides and pupicides) [25,26] deserve a special mention due to the vast biodiversity of species found in the world, estimated to be approximately 400,000 terrestrial species [27]. Botanical insecticides can be plant extracts, essential oils and/or secondary metabolites [4,14].

3. Plant Natural Products to Control Mosquitoes

The search for plant natural products to control Ae. aegypti dates back a number of years, with research published since the 1980s [28,29]. However, chemical insecticides are most commonly used, despite their enormous toxicity to non-target organisms, such as: (i) poisoning and death; (ii) cancer, by non-genotoxic mechanisms (immunosuppressants, cytotoxic) or by triggering the carcinogenic process in different ways; (iii) harmful effects on the nervous, renal, respiratory and reproductive systems and (iv) induction of oxidative stress [30].
In addition to toxicity, another concern is the increasing resistance of the mosquito vector to chemical insecticides. One example is the knockdown resistance (kdr) mutation, in which resistance to pyrethroid insecticides occurs, whereby the target site is the sodium channel of the Ae. aegypti nervous system [31,32]. In Brazil, of the five insecticides approved by the Public Health Ministry and recommended by the WHO for adult mosquito control, four belong to the pyrethroid class together with one organophosphate (malathion). However, in 2011 a technical note was issued suspending the use of pyrethroids in Brazil to control Ae. aegypti [33].
The level of resistance is dependent on the insecticide concentration, frequency and duration of application [34]. The resistance mechanisms of mosquitoes may be associated with changes in the insect cuticle resulting in less insecticide absorption [35], changes in insect metabolism involving biotransformation enzymes [36,37] and modifications of the insecticide target site, usually by genetic mutations [38,39].
The main esterases involved in the resistance process are carboxylesterases and cholinesterases. Carboxylesterases are usually resistant to organophosphates, with this resistance relating to both a quantitative mechanism (overproduction of enzyme) and qualitative mechanism (mutations that cause alterations in enzymatic properties) [40]. In the case of cholinesterases, the resistance is mainly caused by gene mutation. The main insecticides resistant to the acetylcholinesterase target site are organophosphates and carbamates [41].
Another problem associated with chemical insecticides is the damage caused to the environment and living organisms by their degradation products, which may prove more toxic than the original product itself. Examples include the degradation products of temephos, whose effects have already been documented in aquatic environments [42], and malathion, together with its metabolites, in non-target organisms such as Daphnia magna (Cladocera: Daphniidae) [43].
Insecticides derived from plant natural products therefore offer a promising source of safer new products for mosquito control due to minimal residues from its natural degradation in both the field and in water, minimizing ecosystem disruption [44,45]. There is considerable research on insecticides of natural origin, especially those of microbial and plant origin, due to their innumerable secondary metabolites produced especially as a defense mechanism against natural predators [46]. It is estimated that there are more than 100,000 plant metabolites, with hundreds or more exhibiting some activity against insects [47].
Botanical insecticides are advantageous as they are generally environmentally safe, non-toxic to non-target organisms including homeothermic animals and their residues biodegradable [25,26,30]. The synergic mixture of the active compounds in extracts induce several mechanisms of action and result in less pest resistance [30,48].
The present review focuses on the more recent studies of botanical extracts and active compounds in applications against Ae. aegypti, from immature to adult stages, in addition to their main proposed mechanisms of action. The crude extracts are obtained using different extraction methods with organic solvents or water. Essential oils are obtained by steam distillation or hydrodistillation. The classes of active compounds include terpenes, alkaloids and amides, steroids, flavonoids, furanochromones, phenylpropanoids and phenol derivatives, lignans and neolignans, naphthoquinones, fatty acids and their derivatives. The type of insecticide activity (ovicide, larvicide, pupicide, adulticide) is reported as mortality and lethal concentration values (LC50, LC90 and/or LC99), together with egg hatchability. The other activities tested are mosquito repellency, oviposition deterrence, growth regulation and the antifeedant effect.

4. Essential Oils

Essential oils deserve special attention as they have yields of 0.5 to 2.0% in the extraction process, contain a high concentratration of secondary metabolites and generally present potent activity due to the synergic effect of the constituents. An important advantage is, with few exceptions, their relatively low, or no, toxicity to mammals (Figure 2). Some pure compounds constituents of essential oils are moderately toxic to mammals (LD50 800–3000 mg/kg in rodents) while formulated products usually are low or non-toxic to mammals, birds and fish (LD50 above 5000 mg/kg for rodents) [47,49].
These essential oils are mainly obtained from aromatic plants, of which there are more than 3000 species. Approximately 10% of these are already produced in large quantities for other uses, such as flavorings and fragrances, and are therefore readily available at reasonable prices [30]. Essential oils are composed of volatile compounds, which give an important advantage of non-persistence in the environment [49,50].
It is important to note that the same volatility may be a disadvantage in terms of instability. However, this property can be overcome using pharmaceutical technology such as micro and nanoencapsulation [51,52]. Formulation development is therefore critical for essential oils to be used effectively and safely as pesticides. A number of studies have demonstrated that a suitable vehicle prolongs the insecticidal effect [51,53,54].
Of the plant families affording essentials oils, those most tested against Ae. aegypti larvae were Myrtaceae, in particular Eucalyptus species, followed by Fabaceae, Asteraceae, Apiaceae and Lamiaceae. Asteraceae was the most important for the adulticide, repellent and oviposition effects.
Regarding larvicidal activity there is currently no value specified by the WHO to discriminate whether a compound or extract is active against insects. However, researchers usually consider that an LC50 < 50 µg/mL is very active; an LC50 50–100 µg/mL is active, and an LC50 > 100 µg/mL is weak/inactive [55,56,57].
Considering this classification, this review highlights 11 species with 12 very active essential oils, 11 species with 14 active essential oils and 6 species with weak/no activity for 7 essential oils. Eight species do not have reported LC50 values and are not considered in this classification. However, these values can change significantly after formulation, as discussed in Section 8 “Limitations and/or Expectations of Plant Natural Product Insecticide Applications”.
Some studies made the identification of secondary metabolites in essential oils evaluated for insectidal activities described abouve and its chemical structures are illustrated in Figure 3.
Table 1 and Table 2 summarise the publications selected for this review and discussion of the essential oils active against the Ae. aeygpti mosquito. Table 1 describes larvicidal activities, while Table 2 details the adulticidal, repellent and oviposition activities.
The essential oil from Acacia nilotica (L.) Delile (Fabaceae) seeds had the highest insecticidal activity (LC50 3.17 µg/mL). The major compounds were hexadecane (1) and heptacosane (2) [58]. This species has been investigated for several medicinal uses including spasmogenic and antiplasmodial activities of seed extracts [59].
Another species is Myristica fragans Houtt. (Myristicaceae), which is popularly known as nutmeg and is used as a flavoring. Essential oil from its seeds demonstrated high toxicity against Ae. aegypti, in both the L3 larval phase (LC50 28.2 µg/mL) and the adult phase (LC50 18.5 µg/mg female). The major compounds identified were sabinene (3, 52%), α-pinene (no stereochemistry defined, 4) (13%) and terpinen-4-ol (5) (11%). Regarding neurotoxic effects, this essential oil is non-toxic to humans as its IC50 values for human acetylcholinesterase and human butyrylcholinesterase are higher than 4000 µg/mL [54]. Nutmeg flower essential oil presented higher larvicidal activity (LC50 47.42 μg/mL) than the ethanolic extract (LC50 75.45 μg/mL). This result suggests that the constituents of the essential oil either exhibit higher larvicidal activity, or that the synergy between them favors the toxicity to the mosquito [60].
Eucalyptus species (Myrtaceae) leaf essential oils showed LC50 values in the range of 31.0–95.5 µg/mL for the larvae stage and 100% repellency for 1.5 h. Eucalyptus camaldulensis Dehnh. had the highest larvicidal activity while Eucalyptus saligna Sm. displayed higher repellency than Eucalyptus nitens (H. Deane & Maiden) Maiden [61,62,63].
Cinnamomum osmophloeum Kaneh. (Lauraceae) is commonly named pseudocinnamomum and the essential oil from leaves of different places demonstrating larvicidal activities of LC50 36 to 177 µg/mL [64]. Similar variation was observed in the larvicidal activity of different guava crops. The LC50 values of Psidium guajava L. (Myrtaceae) leaf essential oil ranged from 39.48 to 64.25 µg/mL [65].
Other edible plant essential oils that showed strong larvicidal activity were parsley [Petroselinum crispum (Mill.) A.W. Hill (Apiaceae)], fennel [Foeniculum vulgare Mill. (Apiaceae)], star anise [Illicium verum Hook. f. (Illiciaceae)], Piper sarmentosum Roxb. ex Hunt. (Piperaceae) and sucupira [Pterodon emarginatus Vogel (Fabaceae)] [54,60,66].
For parsley, the toxicity was evaluated for pyrethroid-susceptible and pyrethroid-resistant mosquitoes (LC50 40 µg/mL) and the major metabolite was thymol (6). Similar toxicity was observed for fennel (LC50 44.84 µg/mL), for star anise (LC50 39.8 µg/mL) and for Piper sarmentosum (LC50 49.19 µg/mL) [54,60]. The major constituent of star anise essential oil was trans-anethole (7, 90%) [54]. Furthermore, parsley and star anise essential oils demonstrated adulticidal activity with LC50 6.01 µg/mg female for pyrethroid-susceptible, LC50 6.15 µg/mg female for pyrethroid-resistant (Petroselinum crispum) and LC50 10.3 µg/mg female for Illicium verum [54,60].
A sucupira (Pterodon emarginatus) fruit essential oil incorporated into a nanoemulsion to improve water solubility and increase product stability presented an LC50 34.75 µg/mL. The metabolites were identified as β-caryophyllene (8), geranylgeraniol (9) and 6α,7β-dihydroxyvouacapan-17-β-oic acid (10). The toxicity for non-target organisms was tested in adult female Swiss albino mice [Mus musculus (Rodentia: Muridae)] with no behavioral effects, macroscopical changes or deaths reported [66].
Table 1. Larvicidal activity of essential oils against the Ae. aegypti mosquito.
Table 1. Larvicidal activity of essential oils against the Ae. aegypti mosquito.
Plant SpeciesFamilyCountryPart UsedLarval StageMortalityTime of Analysis (h)Reference
% DeathConcentration (ppm)LC50 (ppm)LC90 (ppm)
Acacia nilótica (L.) DelileFabaceaeIndiaSeedsL4NDND3.1711.7324[58]
Alpinia purpurata (Viell.) K. Schum.ZingiberaceaeBrazilRed flowersL4NDND80.70ND24[67]
Pink flowersL4NDND71.50
Baccharis reticularia DC.AsteraceaeBrazilLeavesL4NDND221.27457.4724[68]
Bauhinia pulchella Benth.FabaceaeBrazilLeavesL3NDND105.90ND24[69]
Bauhinia ungulata L.FabaceaeBrazilLeavesL3NDND75.10ND24[69]
Cinnamomum osmophloeum Kaneh.LauraceaeTaiwanLeavesL4NDND36.0 to 177.079.0 to 296.024[64]
Croton rhamnifolioides Pax & K. Hoffm.EuphorbiaceaeBrazilLeavesL4NDND89.0 and 122.30ND24[70]
Cunninghamia konishii HayataTaxodiaceaeTaiwanWoodL4NDND85.70171.4024[71]
LeavesL4NDND91.70176.50
Curcuma longa L.ZingiberaceaeThailandRhizomeL4 (p-s)NDND65.51110.9324[60]
Eucalyptus camaldulensis Dehnh.MyrtaceaeTaiwanLeavesL4NDND31.071.8024[61]
Eucalyptus nitens (H. Deane & Maiden) MaidenMyrtaceaeArgentinaLeavesL3/L4NDND52.83ND24[62]
Eucalyptus urophylla S.T. BlakeMyrtaceaeTaiwanLeavesL4NDND95.50166.3024[61]
Ferula galbaniflua Boiss. & BuhseApiaceaeCoreaResinL390100.0NDND48[45]
Foeniculum vulgare Mill.ApiaceaeThailandFruitL4 (p-s)NDND44.8457.0524[60]
Hyssopus officinalis L.LamiaceaeCoreaFlowersL395100.0NDND48[45]
Illicium verum Hook. f.IlliciaceaeBrazilFruitL3NDND39.8053.024[54]
Larix europea Lam. & A. DC.PinaceaeCoreaResinL387100.0NDND48[45]
Limnophila aromatica (Lamk.) Merr.ScrophulariaceaeThailandWhole plantL4 (p-s)NDND47.9465.1424[60]
Mentha spicata L.LamiaceaeIndiaLeavesL3NDND56.08110.2824[72]
Myristica fragrans Houtt.MyristicaceaeThailandFlowersL4 (p-s)NDND47.4269.2824[60]
BrazilSeedsL3NDND28.2041.7024[54]
Myroxylon pereirae (Royle) KlotzschFabaceaeCoreaResinL397.5100.0NDND48[45]
9550.0NDND
Pelargonium graveolens L’Hér. ex AitonGeraniaceaeCoreaLeavesL382100.0NDND48[45]
Petroselinum crispum (Mill.) A.W. HillApiaceaeThailandFruitL4 (p-s)NDND43.2266.6024[60]
L4 (p-s 1)NDND44.5068.2924
L4 (p-s 2)NDND44.0367.7124
Pimenta dioica (L.) Merr.MyrtaceaeBrazilFruitL3NDND104.40137.3024[54]
Pinus sylvestris L.PinaceaeNigeriaNeedlesL4NDND100.39ND24[73]
Piper aduncum L.PiperaceaeBrazilLeavesL3/L4NDND289.90654.9024[74]
NDND134.10527.1048
Piper sarmentosum Roxb. ex Hunt.PiperaceaeThailandStem and LeavesL4 (p-s)NDND49.1975.1024[60]
Pogostemon patchouli Pellet.LamiaceaeCoreaWhole plantL397100.0NDND48[45]
Porophyllum ruderale (Jacq.) Cass.AsteraceaeBrazilFlowers and leavesL3NDND60.90132.4824[75]
L4NDND72.28173.6524
Psidium guajava L.MyrtaceaeBrazilLeavesL4NDND39.48 to 64.2557.34 to 86.024[65]
Pterodon emarginatus VogelFabaceaeBrazilFruitL4NDND34.75ND48[66]
Rosmarinus officinalis L.LamiaceaeBrazilLeavesL480250.0NDND24[76]
90250.0NDND48
Sphaeranthus indicus L.AsteraceaeIndiaLeavesL4NDND140.0350.024[77]
Syzygium aromaticum (L.) Merr. & L.M. PerryMyrtaceaeNigeriaBudL4NDND92.56ND24[73]
Tanacetum argenteum (Lam.) Willd. subsp. argenteum (Lam.)AsteraceaeTurkeyAerial partsL1NDND93.30241.7024[78]
Trachyspermum ammi (L.) SpragueApiaceaeCoreaSeedsL3100100.0NDND48[45]
8050.0NDND
LC50 lethal concentration required to kill 50% of the larval population, LC90 lethal concentration required to kill 90% of the larval population, ND not described, p-s pyrethroid-susceptible.
Table 2. Adulticidal, repellent and oviposition activities of essential oils against the Ae. aegypti mosquito.
Table 2. Adulticidal, repellent and oviposition activities of essential oils against the Ae. aegypti mosquito.
Plant SpeciesFamilyCountryPart UsedActivityResultsTime of Analysis (h)Reference
Acantholippia seriphioides (A. Gray) MoldenkeVerbenaceaeArgentinaNDRepellent100% of repellency at 50%1.2[63]
Aloysia citriodora PalauVerbenaceaeArgentinaNDRepellent100% of repellency at 12.5%1.5[63]
Alpinia purpurata (Viell.) K. Schum.ZingiberaceaeBrazilRed flowersOvipositionOviposition disruptive effectND[67]
Pink flowersOvipositionOviposition disruptive effectND[67]
Baccharis spartioides (Hook. & Arn.) RemyAsteraceaeArgentinaNDRepellent100% of repellency at 12.5%1.5[63]
Croton rhamnifolioides Pax & K. Hoffm.EuphorbiaceaeBrazilLeavesOvipositionOnly 30% of oviposition at 100.0 µg/mL16[70]
Eucalyptus nitens (H. Deane & Maiden) MaidenMyrtaceaeArgentinaLeavesRepellent100% pure repellency1.5[62]
Eucalyptus saligna Sm.MyrtaceaeArgentinaNDRepellent100% of repellency at 50%1.5[63]
Illicium verum Hook. f.IlliciaceaeBrazilFruitAdulticideLC50 10.30 µg/mg female
LC90 17.50 µg/mg female
24[54]
Minthostachys mollis GrisebLamiaceaeArgentinaNDRepellent100% of repellency at 50%1.0[63]
Myristica fragrans Houtt.MyristicaceaeBrazilSeedsAdulticideLC50 18.50 µg/mg female
LC90 31.90 µg/mg female
24[54]
Petroselinum crispum (Mill.) A.W. HillApiaceaeThailandFruitAdulticide (p-s)LC50 6.01 µg/mg female
LC90 9.39 µg/mg female
24[60]
Adulticide (p-r)LC50 6.15 µg/mg female
LC90 9.82 µg/mg female
24[60]
Pimenta dioica (L.) Merr.MyrtaceaeBrazilFruitAdulticideLC50 16.6 µg/mg female
LC90 31.4 µg/mg female
24[54]
Pluchea carolinensis (Jack.) G.AsteraceaeMartinicaLeaves and flowersRepellent36.6% of repellency at 1.0%ND[79]
Irritant66.2% of irritation at 0.1%ND[79]
Rosmarinus officinalis L.LamiaceaeArgentinaNDRepellent100% of repellency at 50%1.5[63]
Sphaeranthus indicus L.AsteraceaeIndiaLeavesRepellent100% of repellency at 200 ppm3.5[77]
Adulticide100% of mortality at 800 ppm24[77]
Tagetes minuta L.AsteraceaeArgentinaNDRepellent100% of repellency at 25%1.5[63]
LC50 lethal concentration required to kill 50% of the mosquito population, LC90 lethal concentration required to kill 90% of the mosquito population, ND not described, p-s pyrethroid-susceptible, p-r pyrethroid-resistant.
The essential oil of spearmint [Mentha spicata L. (Lamiaceae)] leaves also showed larvicidal activity (LC50 56.08 µg/mL). The main constituents were carvone (11) (48.6%), cis-carveol (12, 21.3%) and limonene (13, 11.3%) [72]. Another spice with larvicidal acitivity was turmeric [Curcuma longa L. (Zingiberaceae)] whose rhizome essential oil demonstrated LC50 65.51 µg/mL [62]. Similarly, Porophyllum ruderale (Jacq.) Cass. (Asteraceae) leaves essential oil, a herb used for seasoning food, showed LC50 60.9 µg/mL for L3 larvae and LC50 72.3 µg/mL for L4. The main metabolite identified was β-ocimene (14, 94%) [75].
Two other species exhibited very similar LC50 values (ca. 93 µg/mL): clove [Syzygium aromaticum (L.) Merr. & L.M. Perry (Myrtaceae)] against L4 larvae and Tanacetum argenteum (Lam.) Willd. subsp. argenteum (Lam.) (Asteraceae) against L1 [73,78]. The major constituent of clove bud essential oil was eugenol 15 (80%) [73]. The clove bud alcoholic extract is popularly used as a repellent.
Cheng et al. (2013) demonstrated that essential oils of different parts of Cunninghamia konishii Hayata (Taxodiaceae) were toxic to L4 larvae (wood LC50 85.7 µg/mL) and (leaves LC50 91.7 µg/mL). The ethanolic extracts were inactive: wood (LC50 240 µg/mL) and leaves (LC50 > 400 µg/mL). The essential oil major compounds were: cedrol (16, 53.0%) and α-pinene (4, 25.6%, wood), and α-pinene (4, 35.9%) and p-cymene (17, 16.7%, leaves) [71]. The results suggest that compound synergy is important for essential oil larvicidal activity as isolated compounds showed lower toxicity, with the exception of p-cymene [71].
Alpinia purpurata (Viell.) K. Schum. (Zingiberaceae) and Croton rhamnifolioides Pax & K. Hoffm. (Euphorbiaceae) demonstrated toxicity against L4 larvae and oviposition deterrent effect. The LC50 of A. purpurata essential oil was 71.5 µg/mL (pink flowers) and 80.7 µg/mL (red flowers) [67], while the LC50 for C. rhamnifolioides was 89.0 µg/mL (fresh leaves) and 122.3 µg/mL (stored leaves) [70].
Other species, such as Baccharis reticularia DC. (Asteraceae); Piper aduncum L. (Piperaceae) and Pinus sylvestris L. (Pinaceae) showed weak larvicidal activities (LC50 100.4–290.0 µg/mL) [68,73,74]. Bauhinia pulchella Benth. (Fabaceae) displayed weak activity (105.9 µg/mL) whereas Bauhinia ungulata L. (Fabaceae) was active (75.1 µg/mL) [69] and Pimenta dioica (L.) Merr. (Myrtaceae) was investigated for both larvicidal (weak activity, LC50 104.4 µg/mL) [54] and adulticidal action (very active, LC50 16.6 µg/mL) [24].
The essential oil of Sphaeranthus indicus L. (Asteraceae) leaves showed poor activity: larvae (LC50 140 µg/mL), adult (800 µg/mL for 100% mortality) and repellent (200 µg/mL). This essential oil showed low toxicity (1500 µg/mL) to a non-target aquatic predator [Toxorhynchites splendens (Diptera: Culicidae)] [77]. Similarly, the essential oil of rosemary [Rosmarinus officinalis L. (Lamiaceae)] was poor active when incorporated into a nanoemulsion and tested for larvicidal activity. The mortality of L3 larvae at 250 µg/mL of product was 80% at 24 h and 90% at 48 h [76].
Seo et al. (2012) investigated the activity of 7 plant species essential oils against L3 larvae at 50 µg/mL and 100 µg/mL. Mortality ranged from 80 to 100% after 48 h. The species tested were Ferula galbaniflua Boiss. & Buhse and Trachyspermum ammi (L.) Sprague (Apiaceae); Hyssopus officinalis L. and Pogostemon patchouli Pellet. (Lamiaceae); Larix europea Lam. & A. DC. (Pinaceae); Myroxylon pereirae (Royle) Klotzsch (Fabaceae), and Pelargonium graveolens L’Hér. ex Aiton (Geraniaceae) [45]. Additional LC50 data is required to determine the degree of activity against Ae. aegypti larvae.
Repellent activity was reported for essentials oils of Acantholippia seriphioides (A. Gray) Moldenke (Verbenaceae) and Aloysia citriodora Palau (Verbenaceae); Baccharis spartioides (Hook. & Arn.) Remy (Asteraceae) and Tagetes minuta L. (Asteraceae), and Minthostachys mollis Griseb and Rosmarinus officinalis. All of the aforementioned demonstrated 100% repellency (12.5–50%) [63]. In another study, Pluchea carolinensis (Jack.) G. (Asteraceae) demonstrated repellency (36.6%) at 1.0% concentration and irritation (62.2%) at 0.1% concentration [79].
The results listed and discussed in this section clearly suggest that essential oils present a promising alternative to develop an effective natural and potentially more eco-friendly insecticide for the control of Ae. aegypti, especially during the larval phase. The challenges for these materials are to improve solubility in water and prolong the insecticidal effect. It is also important to understand the synergism and/or antagonism of their constituents, together with the optimum ratio.

5. Organic/Aqueous Extracts

Concerning organic/aqueous extracts, the plant families with the highest number of species tested against Ae. aegypti larvae were Fabaceae, Asteraceae, Piperaceae and Euphorbiaceae. Similarly, as described by Isman (2015), India was the country with the most publications in this field, followed by Brazil [14].
Of the 20 plant species, at least one organic/aqueous extract showed high larvicidal activity (LC50 < 50 µg/mL); 12 were active (LC50 50–100 µg/mL) and 26 had weak activity (LC50 > 100 µg/mL). Nevertheless, these values can change significantly after formulation in a similar way to essential oils, as described in Section 8 “Limitations and/or Expectations of Plant Natural Product Insecticidal Applications”.
Figure 4 details the chemical structures of the secondary metabolites identified in the organic extracts.
Table 3 and Table 4 summarise the scientific literature selected for the discussion of insecticidal activities of organic/aqueous extracts against Ae. aeygpti mosquito. Table 3 describes larvicidal activities, while Table 4 describes adulticidal, pupicidal, ovicidal, repellent and oviposition activities.
Piper species (Piperaceae) demonstrated LC50 ranging from 2.23 to 567 μg/mL for L3 and L4 larval stages [106,109,110]. The most active species extracts were Piper longum L. (fruit ethanolic), followed by P. sarmentosum (entire plant ethanolic LC50 4.06 μg/mL) and Piper ribesoides Wall. (wood LC50 8.13 μg/mL) [109]. Piper nigrum L. peppercorn ethanolic extract was active and purified fractions were highly active, with possible toxicity due to oleic acid (18) [110]. P. aduncum and Piper hispidum Sw. displayed weak activity against L3 larvae (LC50 > 150 μg/mL) [106].
An Echinops transiliensis Golosk. (Asteraceae) root dichloromethane extract showed strong toxicity against larvae (LC50 3.21 μg/mL). The nine thiophene derivatives isolated showed strong activity (LC50 0.16—19.97 μg/mL) [90]. Similarly, a Euphorbia tirucalli L. (Euphorbiaceae) stem bark petroleum ether extract presented LC50 4.25 μg/mL against L4 larvae, while a Euphorbia hirta L. leaves extract exhibited weak activity (LC50 272.36 μg/mL) [92].
The Hypericum japonicum Thunb. (Hypericaceae) species was also very toxic to mosquito larvae. The organic extracts of the whole plant presented LC50 values between 7.37 and 13.15 μg/mL. The methanolic extract proved the most active, its purified fraction displaying LC50 0.95 μg/mL. The major constituents of this active fraction were isopropyl palmitate (19), 2,4-Di-tert-butylphenol (20) and hydrocinnamic acid (21) [97]. Similarly active, the Nerine sarniensis (L.) Herb. (Amaryllidaceae) bulb ethyl acetate extract demonstrated toxicity against L1 larvae (LC50 8.0 μg/mL) and adult females (LC50 4.6 μg/mosquito) [104].
Purified fractions of a n-hexane extract of Limonia acidissima L. (Rutaceae) leaves showed interesting ovicidal, larvicidal and pupicidal activities. The LC50 for L3 larvae ranged from 4.11 to 23.53 µg/mL; for pupae 4.19 to 39.48 µg/mL, and the maximum inhibition of egg hatching was 78.4% at 10 ppm. Bioguided fractionation resulted in the isolation of nilocetin (88) [99].
The biological activity of each plant species extract is specific to the plant part(s) and the polarity of the extraction solvent used. Furthermore, activity can differ significantly for the 4 different larval stages [58,86,89,92,93,94,96,98,100,108,113,115]. This variation is discussed below.
A Jatropha curcas L. (Euphorbiaceae) leaves petroleum ether extract showed LC50 8.79 μg/mL whereas the root ethanolic extract demonstrated less activity (LC50 44.75 μg/mL) [92,98]. A methanolic extract of Gardenia ternifolia Schumach. & Thonn. (Rubiaceae) leaves was very active (LC50 32.01 μg/mL) whereas an acetone extract exhibited lower activity (LC50 83.31 μg/mL) [94]. This study exemplifies the effect of different solvents on the toxicity of the same plant part.
n-Hexane, chloroform and acetone extracts of A. nilotica seed pods showed weak activity (LC50 103—169 μg/mL), while ethyl acetate was more active (LC50 59.12 μg/mL). Interestingly, a benzene extract demonstrated the most activity (LC50 45.32 μg/mL) [58]. A Ficus benghalensis L. (Moraceae) leaves methanolic extract showed activity against L2, L3 and L4 larvae, but the activity was low for benzene and acetone extracts [93]. Another example of solvent variation was an aqueous extract of Solanum nigrum L. (Solanaceae) fruits which displayed very low activity (LC50 359.0 μg/mL) against L3/L4 larvae, while higher activity was observed for a hexanic extract (LC50 17.63 μg/mL) [113].
With not so discrepant, but nonetheless different results, a Pemphis acidula J.R. Forst. & G. Forst. (Lythraceae) leaves methanolic extract (LC50 22.10 μg/mL) was more active than a benzenic leaves extract (LC50 43.99 μg/mL) [108], while a Tagetes patula L. (Asteraceae) seeds acetone extract (LC50 15.74 μg/mL) was more active than the 50% ethanol extract (LC50 25.46 μg/mL) [93]. P. acidula extracts also demonstrated complete inhibition of egg hatchability at 500 ppm acetone and 450 ppm methanolic [108].
In addition, different parts of Lonchocarpus urucu Killip & A.C. Sm. (Fabaceae) extracted with the same solvent (methanol) showed different toxicity. The root bark extract was more active (LC50 17.6 μg/mL) than the root medulla extract (LC50 33.32 μg/mL) against L4 larvae [100]. The toxicity of Heracleum rigens Wall. (Apiaceae) seed extracts was evaluated against different larval stages (LC50 40.64 to 308.65 μg/mL), with the petroleum ether extract the most toxic to all larval stages and acetone the least toxic [96].
Different organic solvent extracts of Cassia fistula L. (Fabaceae) leaves were evaluated against the mosquito (larvicide, ovicide and repellent). The methanolic extract was the most active for all activities, notably as a larvicidal (LC50 10.69 μg/mL). Other extracts also demonstrated high activity against larvae: benzene (LC50 18.27 μg/mL) and acetone (LC50 23.95 μg/mL). The non-hatching concentration for eggs ranged from 120 to 160 mg/L and the repellent action (100% at 5 mg/cm2) ranged from 6.0 to 4.3 h [86].
Dalbergia brasiliensis Vogel (Fabaceae), commonly known as Jacarandá-da-Bahia in Brazil, is a tree native to the states of Bahia, Minas Gerais, Espírito Santo, Rio de Janeiro and São Paulo. Larvicidal activity of its leaves and trunk bark ethanolic extracts, together with fractions purified by partitioning with n-hexane, ethyl acetate and chloroform, were similar (LC50 between 24.0 and 44.0 μg/mL) [89].
Studies using the incorporation of inorganic nanoparticles, such as zinc oxide and silver in plant extracts, have shown an increase in their biological activity. They are generally easy to obtain, inexpensive, not to mention non-toxic to humans and animals [53]. All of the plant extracts described below showed higher larvicidal activity when incorporated into nanoparticles [53,82,116].
An aqueous extract of Artemisia herba-alba Asso (Asteraceae) leaves was tested against L4 larvae strains from India and Saudi Arabia. The LC50 values were 117.18 μg/mL and 614.54 μg/mL for India and Saudi Arabia larvae, respectively. When the extract was incorporated into silver nanoparticles the activity increased significantly to 10.70 μg/mL and 33.58 μg/mL, respectively. Similar results were observed against adult mosquitoes [82].
Aqueous extracts of Ventilago madraspatana Gaertn. (Rhamnaceae) and Zeuxine gracilis (Berda) Bl. (Orchidaceae) leaves were also more toxic to egg, larvae and adult mosquitoes when incorporated into silver nano particle. No egg hatching was observed at 120 μg/mL and 12 μg/mL, respectively. Corresponding larvicidal (LC50 26.92 μg/mL and 10.39 μg/mL) and adulticidal effective dose activities (44.85 μg/mL and 27.90 μg/mL) were observed [112,116].
The activity of a zinc oxide nanoparticle incorporating a Myristica fragans leaf methanolic extract was compared with the crude extract. The activity of the crude extract against the 4 larvae stages (LC50 162.03 to 273.9 μg/mL) was less than the nanoparticles (LC50 3.44 to 10.28 μg/mL). Similar activity was reported against the pupa (crude extract LC50 359.08 μg/mL and nanoparticles LC50 14.63 μg/mL), and female adult forms (crude extract LC50 180.26 μg/mosquito and nanoparticles LC50 15 μg/mosquito) [53].
Following are the results of plant species that demonstrated at least one active extract for larvicidal action. A 90% ethanol extract of the leaves of Scoparia dulcis L. (Plantaginaceae), a plant used in Brazilian indigenous medicine, demonstrated activity against L4 larvae (LC50 83.43 μg/mL). The compounds isolated were betulinic acid (22); scopadulcic acid A (23); scopadulcic acid B (24); scopadiol (25); scopadulciol (26) and scopadulin (27) [95].
A methanolic extract of Cinnamosma fragrans Baill. (Canellaceae) root bark demonstrated potential as an insecticide acting in different ways: L1 larvae (LC50 52.5 μg/mL), adult (0.17 μg/mg) and 80% repellency at 20.8 μg/cm2 [87]. Similarly, extracts of Ervatamia coronaria (Jacq.) Stapf. (Apocynaceae), commonly known as Capre jasmine, were evaluated for different insecticidal applications. L3 larvicides: benzene (LC50 89.59 µg/mL) and ethyl acetate (LC50 97.53 µg/mL) [85]. Ovicide/no hatchability: methanol (200 ppm), benzene (250 ppm) and ethyl acetate (300 ppm) [85,117]. The repellent activity was the same for all extracts (100% repellency at 5 mg/cm2) [85,117].
Different extracts were obtained from Mirabilis jalapa L. (Nyctaginaceae) leaves and investigated for their L3 larvicidal activity: methanol (LC50 64.58 µg/mL), ethyl acetate (LC50 72.77 µg/mL), chloroform (LC50 88.20 µg/mL) and benzene (LC50 97.03 µg/mL) [103]. Extracts of three other species obtained from the same plant part with different organic solvents were also evaluated. The methanolic extracts were the most active: Acalypha alnifolia Klein ex Willd. (Euphorbiaceae) (L4 instar LC50 128.55–202.15 µg/mL), Orthosiphon thymiflorus (Roth) Sleesen (Labiatae) (L3 instar LC50 149.96–228.13 µg/mL) and Ocimum sanctum L. (Labiatae) (L4 instar LC50 175.67–425.94 µg/mL) [80,105,107].
Petroleum ether leaf extracts of Pedilanthus tithymaloides (L.) Poit. (Euphorbiaceae), Citrullus colocynthis (L.) Schrad. (Cucurbitaceae) and Phyllanthus amarus Schumach. & Thonn. (Euphorbiaceae) were active against L4 instar, with LC50 55.26 µg/mL, 74.57 µg/mL and 90.92 µg/mL, respectively [88,92]. Similar extracts of Catharanthus roseus (L.) G. Don (Apocynaceae), Eupatorium odoratum L. (Asteraceae), Nyctanthes arbor-tristis L. (Oleaceae), Boenninghausenia albiflora (Hook.) Rchb. ex Meisn. (Rutaceae) and Valeriana hardwickii Wall. (Valerianaceae) presented weak activity (LC50 > 100 µg/mL) [83].
Larvicidal activity was also reported for Maytenus oblongata Reissek (Celasteraceae) bark ethyl acetate extract (LC50 74.4 µg/mL) and Millettia pachycarpa Benth. (Fabaceae) root ethanol extract (LC50 98.47 µg/mL) [101,102]. The latter also demonstrated ovicidal activity at 200 ppm (100% non-hatchability) [102].
Ethanol, n-hexane and n-butanol extracts of flowers and aerial parts of Buddleja polystachya Fresen. (Buddlejaceae) were evaluated for adulticidal activity at 5 µg/mosquito: the most active was n-hexane aerial parts (100% mortality), followed by n-hexane flowers (96.7% mortality), ethanol aerial parts (90% mortality) and ethanol flowers (83.3% mortality). Only the n-butanol flower extract was investigated for L1 larvicidal activity, demonstrating 100% mortality at 1 µg/µL [84].
Finally, the species that presented poor larvicidal action for all tested extracts. Aristolochia bracteata Retz. (Aristolochiaceae) leaves methanolic extract was evaluated against L3 larvae (LC50 114.89 µg/mL), egg (100% non-hatchability at 240 ppm) and adult mosquito (100% repellency at 6 mg/cm2) [81]. Similarly, Caesalpinia pulcherrima (L.) Sw. (Fabaceae) extracts were evaluated against larvae: benzene (LC50 136.37 µg/mL) and ethyl acetate (LC50 144.67 µg/mL), and 100% non-hatchability for benzene (375 ppm) and ethyl acetate (450 ppm). However, both extracts demonstrated 100% repellency at 5 mg/cm2 [85].
In another study involving a methanol extract of C. pulcherrima, complete inhibition of egg hatching was reported at 300 ppm. The repellency was the same as the aforementioned study (5 mg/cm2) [117]. Coccinia indica Wight & Arn. (Cucurbitaceae) presented similar insecticidal properties for different extracts, with a methanolic extract the most active in terms of ovicidal activity (zero hatchability at 200 ppm) and a hexanic extract having the more effective repellency (100% of repellency at 1 mg/cm2). For 100% non-hatchability, the concentrations were between 200 ppm and 300 ppm and for 100% of repellency were between 1 and 5 mg/cm2 [119].
The methanolic extract of Eclipta alba (L.) Hassk (Asteraceae) leaves was also the most active among the solvents of different polarities used to evaluate larvicidal and ovicidal activities of this plant. The LC50 values against L3 larvae were between 127 and 165 µg/mL. Complete inhibition of egg hatching occurred at 300 ppm for the methanolic extract and 350 ppm for the other solvents [91].
The methanolic extract of Mentha piperita L. (Lamiaceae) and different extracts of Cardiospermum halicacabum L. (Sapindaceae) showed repellent activity [118,120]. Essential oils and aqueous extracts of the red and pink flowers of A. purpurata were investigated for both larvicidal activity and oviposition effect. Similar to the essential oils, the extract of the pink flower was more active than the red, and both disrupted oviposition [67].
In general, organic extracts from different parts of Parthenium hysterophorus (Asteraceae), Pithecellobium dulce (Roxb.) Benth. (Fabaceae) and Solanum xanthocarpum Schrad. & J.C. Wendl. (Solanaceae) showed weak insecticidal action, requiring high concentrations to demonstrate some biological activity [111,114,121]. Other species that were inactive were Helicteres velutina K. Schum. (Malvaceae), Momordica charantia L. (Cucurbitaceae), Ormosia arborea Vell (Fabaceae), Solanum variabile Mart. (Solanaceae), Spermacoce latifolia Aubl. (Rubiaceae) and Turnera ulmifolia L. (Turneraceae) [88,95,106].
Several organic extracts have shown important insecticidal activities against the Aedes mosquito, including E. transiliensis, E. tirucalli, H. japonicum, N. sarnisiensis, P. longum, P. ribesoides and P. sarmentosum. Studies in this section show the potential of plant natural products as insecticides targeting different stages of the mosquito life cycle and how different formulation approaches, such as the incorporation of botanical extracts into silver and zinc nanoparticles, can increase the insecticidal effects. Prominent examples include nanoformulations of Z. gracilis, M. fragans and A. herba-herba extracts. These data reinforce the large diversity of plants with toxic effect in different life stages of the Aedes aegypti mosquito. However, an important consideration for these materials is the type of extraction solvent employed, such as n-hexane, chloroform, benzene, given their toxicity to humans associated with harmful residues [122].

6. Secondary Metabolites

6.1. Terpenes

Terpenoids are a very promising target for the development of products of natural origin to be used in the control of the Ae. aegypti mosquito. These compounds were the most identified in the essential oils, extracts and purified fractions, generally having better results against the mosquito, especially in terms of larvicidal activity. Of the terpenes, monoterpenes are the most active and present great possibilities in bioinsecticide applications due to their low toxicity against mammals and non-target organisms [50].
This significant activity against the mosquito can be explained by the hydrophobicity of this class. Terpene toxicity against Ae. aegypti larvae may be associated with their nonpolar property as reported for other insects [123,124]. This property increases the ability of the compound to penetrate the hydrophobic larvae cuticle and renders them more toxic to the insect in comparison to polar compounds [123]. The chemical structures of the terpenes tested are shown in Figure 5, Figure 6 and Figure 7.
Diterpene 7-oxo-8,11,13-cleistanthatrien-3-ol (28), isolated from the dichloromethane extract of Vellozia gigantea N.L. Menezes & Mello-Silva (Velloziaceae) adventitious roots, caused 100% larvae mortality at 416.06 µM [125]. The adulticidal activity of the diterpene phytol (29) was LC50 4.23 µM/mosquito [84].
The triterpenoids ursolic acid (30) and betulinic acid (22) showed larvicidal activity with LC50 245.24 µM and 310.83 µM, respectively. Their corresponding structures are illustrated in Figure 5. Bioassays of their chemical derivatives, with esterification of the hydroxyl group at the C-3 position, demonstrated less activity, suggesting that the hydroxyl group plays an important role in larvicidal activity [126].
The sesquiterpenes α-costic acid (31) and inuloxin A (32), both isolated from Inula viscosa (L.) Aiton (Asteraceae), demonstrated strong activity against L1 larvae. The concentration of each terpene required for 100% mortality was 4.27 µM and 4.03 µM, respectively [127]. Other sesquiterpene alkaloids with strong larvicidal activity (L3 and L4 instar) were 1-O-benzoyl-1-deacetyl-4-deoxyalatamine (33) (LC50 9.4 µM) and 1,2-O-dibenzoyl-1,2-deacetyl-4-deoxyalatamine (34) (LC50 2.3 µM). These sesquiterpenes with a β-dihydroagrofuran skeleton were isolated from M. oblongata stems [101].
β-caryophyllene (8) and caryophyllene oxide (35) demonstrated lower larvicidal activity, with LC50 values of 127.23 and 135.24 µM, respectively [78], together with the sesquiterpene cinnamodial (36), isolated from C. fragrans (LC50 70 µM) [87]. Regarding adulticidal activity, sesquiterpenes isolated from C. fragrans—cinnamodial (36), cinnafragrin A (37) and cinnamosmolide (38)—showed strong activity with ED50 0.29, 2.85 and 12.79 nmol/mg mosquito, respectively [87].
Monoterpenes were the most evaluated for larvicidal activity with LC50 values ranging from 88 to 540 µM. The terpene hydrocarbons: limonene (13) (LC50 88.16 µM); α-terpinene (39) (LC50 107.90 µM); α-phellandrene (40) (LC50 121.85 µM), and ρ-cymene (17) (LC50 143.05 µM) were the most active [61,72]. Terpinolene (41), γ-terpinene (42), β-myrcene (43) and sabinene (3) showed LC50 208.46 µM, 225.35 µM, 262.78 µM and 543.92 µM, respectively [61,71]. The oxygenated terpene hydrocarbons carvone (11) and cis-carveol (12) demonstrated activity with LC50 of 155.62 µM and 218.88 µM, respectively [72].
A study of the larvicidal activity of α-pinene and β-pinene enantiomers reported the following LC50 values: (-)-β-pinene (44) (263.52 µM) and (+)-β-pinene (45) (414.73 µM); (-)-α-pinene (46) (363.35 µM) and (+)-α-pinene (47) (484.26 µM). The (-) enantiomers displayed higher activity than (+), so these results showed that the type of enantiomer and even the racemic mixture could directly interfere with the activity [78].
The monoterpene limonene (13) was incorporated into a nanoemulsion to improve its water solubility and therefore increase its activity [68]. This compound deserves to be highlighted as it presented high oral LC50 (>4000 mg/kg) and dermatological LC50 (>5000 mg/kg) values for rodents. It is therefore considered safe and non-toxic to mammals [30].

6.2. Phenylpropanoids and Phenolic Derivatives

The chemical structures of the phenylpropanoids and esters discussed in this section are illustrated in Figure 8.
Among the phenylpropanoids and phenolic derivatives classes cinnamaldehyde (48) and cinnamyl acetate (49) can be highlighted as they presented interesting larvicidal activity (LC50 219.43 and 187.27 µM, respectively) [64]. It is important to note that cinnamaldehyde is present in commercial insect-fighting formulations such as Cinamite® and Valero®. This information suggests that these secondary metabolites have a high potential for use against Ae. aegypti due to their possible toxicological safety, given that they have been authorized as insecticides since 2001.
Important larvicidal properties were also reported for eugenol (15), a phenolic compound that presents some advantages such as its non-persistence in water and soil, together with its natural degradation in organic acids through the action of Pseudomonas, a soil-dwelling bacterium. Furthermore, it is 1500 times less toxic than pyrethrins and 15,000 times less toxic than azinphos-methyl, an organophosphate [50]. The LC50 value for larvicidal activity was 200.97 μM [64]. The phenylpropanoid trans-anethole (7) also showed important action against Ae. aegypti larvae (LC50 283.40 μM) [64].
Metabolites isolated from Zingiber officinale Roscoe (Zingiberaceae) demonstrated strong larvicidal activity, presenting LC50 values of 15.96 µM for 4-gingerol (50), 37.36 µM for 6-dehydrogingerdione (51) and 61.86 µM for 6-gingerol (52) [128]. The phenolic derivatives esters benzyl benzoate (53) and benzyl cinnamate (54)—demonstrated interesting activity as their 100% larvae mortality concentrations were 117.79 μM and 104.92 μM, respectively [45].

6.3. Alkaloids and Amides

The Piperaceae family has numerous compounds with promising activity against Ae. aegypti. The chemical structures of the aforementioned alkaloids and amides are illustrated in Figure 9.
N-Isobutylamide alkaloids from Piper species presented potent larvicidal and adulticidal activities. Secondary metabolites isolated from P. nigrum: pellitorine (55), guineensine (56), pipercide (57) and retrofractamide A (58) presented respective larvicidal LC50 values of 4.12 μM, 2.32 μM, 0.28 μM and 0.12 μM [129]. Regarding the adulticidal activity, LC50 values (μM/female mosquito) were 0.76, 4.43, 6.11 and 4.22, respectively [130]. Pipernonaline (59) isolated from the methanolic extract of P. longum, also showed potent larvicidal activity (LC50 0.73 μM) [127], while piperine (60) and pipwaqarine (61) isolated from P. nigrum demonstrated LC50 of 17.87 and 75.46 μM, respectively [131,132].
The analysis of the structure-activity relationship for the N-isobutylamide alkaloids 5561, it is reasonable to hypothesise that the N-isobutylamine moiety is of crucial importance in terms of larvicidal activity, while the methylenedioxyphenyl moiety does not appear to be essential.
The mesembrine-type alkaloid sarniensinol (62) isolated from N. sarniensis exhibited strong larvicidal (LC50 24.24 μM) and adulticidal (LC50 13.88 μM/female mosquito) activities [104]. The crinine-type alkaloid crinsarnine (63) only demonstrated strong adulticidal activity (LC50 5.78 μM/female mosquito) [133]. N-hydroxyaristolactam I (64), an aristololactam derivative, also showed strong larvicidal activity (LC50 11.45 μM) [130], whereas the pyrrolidine alkaloid (Z)-3-(4-hydroxybenzylidene)-4-(4-hydroxyphenyl)-1-methylpyrrolidin-2-one (65) demonstrated weak activity (LC50 785.86 μM) [134,135].

6.4. Thiophenes and Acids

Thiophene and fatty acid chemical structures are illustrated in Figure 10.
Nine thiophenes, with different numbers of thiophene rings, isolated from E. transiliensis exhibited strong larvicidal activity and a positive correlation was reported between the number of thiophene rings and larvicidal activity, with thiophene derivatives composed of more rings demonstrating more activity [90].
The terthiophene 2,2′: 5′,2′′-terthiophene (66) was the most active (LC50 0.65 μM). The activity (LC50) of bithiophenes was: 4-(2,2′-bithiophen-5-yl)but-3-yne-1,2-diyl diacetate (67) (12.54 μM); 4-(2,2′-bithiophen-5-yl)-2-hydroxybut-3-yn-1-yl acetate (68) (25.31 μM) and 4-(2,2′-bithiophen-5-yl)but-3-yne-1,2-diol (69) (39.19 μM). Lower larvicidal activity was observed for monothiophenes: 2-chloro-4-[5-(penta-1,3-diyn-1-yl)thiophen-2-yl]but-3-yn-1-yl acetate (70) (49.56 μM); 4-[5-(penta-1,3-diyn-1-yl)thiophen-2-yl]but-3-yne-1,2-diyl diacetate (71) (56.02 μM); 4-[5-(penta-1,3-diyn-1-yl)thiophen-2-yl]but-3-yne-1,2-diol (72) (56.68 μM); 2-hydroxy-4-[5-(penta-1,3-diyn-1-yl)thiophen-2-yl]but-3-yn-1-yl acetate (73) (66.64 μM) and 1-hydroxy-4-[5-(penta-1,3-diyn-1-yl)thiophen-2-yl]but-3-yn-2-yl acetate (74) (71.74 μM) [90].
Tetradecanoic acid (75) showed both larvicidal action (LC50 131.37 μM) and oviposition attraction (78.2% at 43.79 μM) [136,137]. Other fatty acids, hexadecanoic acid (76) and dodecanoic acid (77) displayed lower attraction: 57.4% for (76) at 3.9 μM and 68.8% for (77) at 249.6 μM [137].

6.5. Flavonoids

The corresponding chemical structures are shown in Figure 11. Quercetin-4′,7-O-dimethyl ether (78), naringenin-7-O-methyl ether (79) and kaempferol-7-O-methyl ether (80) flavonoids isolated from G. ternifolia were active against L2 larvae. The LC50 values were 108.09 μM, 85.37 μM and 102.08 μM, respectively [94].

6.6. Neolignans

Figure 12 details the neolignan chemical structures.
Eupomatenoid-6 (81), a neolignan isolated from Piper solmsianum C. DC. (Piperaceae), demonstrated strong larvicidal acitivity (LC50 19.33 μM) with probably low toxicity to mammals (IC50 39.30 μM for human fibroblast cells, MRC5, with an estimated LD50 of 42.26 mmol/kg) [138]. Grandisin (82) presented larvicidal activity (LC50 346.82 μM). Histological analysis revealed that this neolignan damages the anterior-middle midgut of the larvae [139].

6.7. Furanochromones and Furanocoumarin

The chemical structures of the coumarins are illustrated in Figure 13.
Khellin (83), a natural furanochromone isolated from Ammi visnaga (L.) Lam. (Umbelliferae), caused L3 mortality (LC50 192.1 μM) [140]. In another study, 83 demonstrated 100% larvae mortality at 3.84 μM and 75% adult mortality at 19.21 μM/mosquito for permethrin-susceptible strains [141]. Another furanochromone isolated from A. visnaga, visnagin (84) presented 93% mortality at 4.34 μM, together with moderate adulticidal activity (65% mortality at 21.72 μM/mosquito). In addition, this study investigated 2 furanocoumarins isolated from Ruta graveolens L. Royle (Rutaceae): 5-methoxypsoralen (5-MOP) (85) and 8-methoxypsoralen (8-MOP) (86), which showed moderate adulticidal activity (55% mortality at 23.13 μM/mosquito and 67.5% mortality at 23.13 μM/mosquito, respectively). Compound 86 presented weak activity against L3 (53.3% mortality at 4.63 μM) while 85 was inactive [141].

6.8. Other Secondary Metabolites

The chemical structures of the compounds referenced in this section are illustrated in Figure 14.
Several other classes of plant natural products have also been investigated regarding their insecticidal activities. Naphthoquinone 2-methoxy-1,4-naphthoquinone (87) isolated from Impatiens glandulifera (Balsaminaceae) showed extremely potent larvicidal acitivity (LC50 0.45 μM) and moderate adulticidal activity (40% mortality at 26.6 μM) [127].
Nilocetin (88), a protolimonoid isolated from L. acidissima, also demonstrated very strong larvicidal (LC50 0.96 μM) and pupicidal (LC50 1.36 μM) activity which was higher than temephos, a well-documented chemical insecticide. The triterpenoid also caused 83% egg mortality at 4.38 μM [99]. Already a study evaluating volatile plant metabolites capable of eliciting an Ae. aegypti behavioral response reported that acetophenone (89) attracted adult mosquitoes whereas 1-octanol (90) acted as a repellent (flight aversive response) [142].

7. Mechanisms of Action

The mechanisms of action for the Ae. aegypti control relate more to the use of conventional chemical insecticides. Table 5 summarizes the mechanisms of action data of the botanical samples discussed in this section. Ae. aegypti control relies primarily on the use of conventional chemical insecticides which target different critical sites in the mosquito life cycle. Organophosphates and carbamates, for example, target acetylcholinesterase enzyme inhibition. Pyrethroids and some organochlorines target sodium channels. Cyclodienes and polychloroterpenes target gamma-aminobutyric acid (GABA) receptors [34].
Other mechanisms of alternative insecticides authorized by regulatory agencies vary in terms of their action. For example, a biological approach employs the use of entomopathogenic bacteria (Bacillus thuringiensis israelensis and Bacillus sphaericus), which act via the toxic action of their spores damaging the intestinal epithelium of larvae. Insect growth regulators (IGR) differ in that they inhibit insect chitin synthesis, and therefore disrupt the moulting process, while juvenile hormone analogs (JHA) act by interfering with the insect’s endocrine system [138].
Further studies are required in order to completely understand the various toxic action mechanisms of botanical insecticides. However, a number of mechanisms have been proposed and proven. The majority of mechanism of action studies have focused on the larval stage, particularly feeding and/or contact. In the case of ingestion, the action is usually through digestive toxicity whereas contact may involve enzymatic inhibition, endocrine disruption (acting especially during the moulting process), toxicity to the nervous system and other mechanisms depending on the target site [143].
The rapid toxic action of essential oils against the insect indicates a possible neurotoxic mode of action [144]. Phytochemicals may act in cholinergic, GABA, mitochondrial and octopaminergic systems [145]. A study of five volatile compounds commonly found in plant essential oils—eugenol, geraniol, coumarin, eucalyptol and carvacrol—investigated docking against octopamine and acetylcholinesterase receptors in Ae. aegypti and Homo sapiens protein models. All compounds were found to dock in both protein models, with some more selectivity for insect proteins [146].
Effects on the larval nervous system were observed after treatment with Piper species extracts. Tremor, convulsion, excitement, followed by paralysis and death were verified after exposure of larvae to P. longum, P. ribesoides and P. sarmentosum extracts. In addition, the larvae showed morphological changes in the anal papillae [109].
Essential oils of I. verum, P. dioica and M. fragrans inhibited acetylcholinesterase causing acetylcholine accumulation in the synapses, with the membrane in a constant state of excitement, culminating in ataxia, lack of neuromuscular coordination and eventual death [54]. The neurotoxic effect was also observed for a nanoemulsion with P. emarginatus essential oil. It probably causes reversible inhibition of acetylcholinesterase and consequently larval death [66].
D. brasiliensis extracts caused external morphological alterations in the larvae, resulting in interference in the moulting process. The authors also reported digestive toxicity and morphological changes in the anal papillae and respiratory siphon of the larvae which interfered with swimming and oxygen flow [89]. Similarly, nilocetin (88) (Figure 14) induced morphological deformations together with moulting symptoms and growth disruption in all mosquito life cycle stages. These compounds also totally ruptured the peritrophic membrane [99].
A Lonchocarpus urucu extract caused disruption in the peritrophic matrix, a medium intestine lining composed of chitin and proteins, whose functions are to protect against abrasion caused by food and micro-organisms, among others such as decreasing the excretion of digestive enzymes through their recycling. In addition, this extract caused extensive damage to the midgut epithelium (Table 5) [100].
Table 5. Mechanisms of action of botanicals against the Ae. aegypti mosquito.
Table 5. Mechanisms of action of botanicals against the Ae. aegypti mosquito.
Target SiteMechanism of ActionCompoundPlant SpeciesReference
Nervous systemInhibition of acetylcholinesterase (AChE)Essential oilIllicium verum, Pimenta dioica and Myristica fragrans[54]
Nanoemulsion with essential oilPterodon emarginatus[66]
Not specifiedEthanolic extractPiper longum, Piper ribesoides and Piper sarmentosum[109]
Gut trypsinInhibition of trypsin and consequent decreased absorption of nutrientes and essential aminoacidsAqueous extractMoringa oleifera[147]
Essential oilCroton rhamnifolioides[70]
Peritrophic matrixChange in internal morphology and consequent insect protection dysfuntionMethanol extractsDerris (Lonchocarpus) urucu[100]
NilocetinLimonia acidissima[99]
Midgut epitheliumTissue destruction and cell disorganizationMethanol extractsDerris (Lonchocarpus) urucu[100]
PellitorineAsarum heterotropoides[148]
GrandisinPiper solmsianum[139]
Anal papillaeMorphological changes, interference with the larva swinCrude extractsDalbergia brasiliensis[89]
Ethanolic extractPiper longum, Piper ribesoides and Piper sarmentosum[109]
Respiratory siphonMorphological changes, interference with the oxygen flowCrude extractsDalbergia brasiliensis[89]
Nanoemulsion with limoneneBaccharis reticularia[68]
Anal gillsComprehensive damage; debris in hemolymphPellitorineAsarum heterotropoides[148]
Thorax and exoskeletonChanges in external morphology, interfering with the molting processCrude extractsDalbergia brasiliensis[89]
PellitorineAsarum heterotropoides[148]
NilocetinLimonia acidissima[99]
Nanoemulsion with limoneneBaccharis reticularia[68]
Digestive systemDigestive toxicityCrude extractsDalbergia brasiliensis[89]
Pellitorine (55), an isobutylamide alkaloid, whose structure is illustrated in Figure 9, promoted histological changes in the thorax, midgut and anal gills. These toxic effects probably occur as a result of compound action on the larval osmoregulation system [148]. Already a nanoemulsion with limonene (13) (Figure 7) promoted morphological alterations to the head, siphon, abdomen cuticles and thorax, promoting larvae fragility and low mobility [68].
C. rhamnifolioides essential oil induced toxicity in the larvae by trypsin-like activity. Trypsin is a serine protease that widely occurs in the gut of insects. A decrease in its activity may result in poor nutrient absorption and non-availability of essential aminoacids, causing insect death [72]. A Moringa oleifera (Moringaceae) extract also caused larval toxicity by inhibiting trypsin in the gut [147]. After treatment with grandisin (82) (Figure 12), larvae presented intense tissue destruction and cell disorganization in the anterior midgut [139].

8. Limitations and/or Expectations of Plant Natural Product Insecticide Applications

As demonstrated in this review, natural products of botanical origin are promising for control of the Ae. aegypti mosquito, although there remain several limitations and challenges to overcome for their application as insecticidal products. From 1998 until early 2011, the number of patents of essential oil-containing mosquito repellent inventions has almost doubled every 4 years [149], but the number of new products does not reflect this. There are several possible reasons for this disparity, such as: (i) the onerous regulatory processes involved in the registration of a pesticide product; (ii) the quantity of raw material biomass required to obtain sufficient extract and/or its isolated active compound, and (iii) most of the research is conducted at the laboratory scale often without field evaluation to confirm the product application [13,47,150].
The complex process of registering an insecticide discourages companies from investing in new products, especially in some places such as Brazil and the European Union. There are numerous criteria, including provision of non-target toxicology and environmental destination data, extensive data to guarantee plant stability and extract standardization, together with physico-chemical and microbiological procedures establishing quality control of the raw material and final product [47,48].
The low availability of raw materials due to limited yields and cultivation usually makes botanical insecticides more expensive than chemicals. The study of bioactive compound synthesis through biotechnology, such as tissue culture in bioreactors, constitutes an alternative to this limitation [150,151]. In moving from laboratory to industrial scale, a number of different factors must be considered: botanical material, analysis technique, formulations, toxicological tests, mechanisms of action, among others.
Considering the botanical material, it is essential to correctly identify the botanical species and determine the chemical composition of the extract (standardization) [13,50]. Chemical composition may vary depending on numerous factors, such as crop period, seasonality, phenological stage, temperature, humidity, luminosity, altitude, pluviometry, ultraviolet radiation, soil and nutrient conditions, geographical locations, collection method, drying and the part of the plant used, among others, and consequently impact insecticidal activity [152,153,154,155].
Regarding larvicidal analysis techniques, important considerations are: (i) larval phase, (ii) analysis time, and (iii) the use of positive and negative controls. In different studies, the younger the larval stage, the more susceptible it is to toxic effects, as reported for Ficus benghalensis, Heracleum rigens, Myristica fragans and Solanum xanthocarpum (Table 3) [53,93,96,114]. This characteristic may relate to the reduced feeding of larvae in the late L4 instar. If the toxic effect of the insecticide is ingestion-dependent, the effect may be less pronounced the closer the larva is to the pupa stage of metamorphosis [156]. Although most studies use 24 h as the contact time to express the mortality result, it is important to note that some materials may have delayed activity, actually causing larvae death after 48, 72 or even 96 h. Thus, during product development, it is necessary to assess the toxic effect at different time intervals. Finally, the use of negative and positive controls is essential to ensure results reliability, although a number of studies did not report this data [157]. Therefore, test non-uniformity makes it difficult to compare the results of different studies. This constitutes another obstacle to overcome for the development of plant natural product insecticides [158].
Understanding the mechanism of action is fundamental in using a material as an inseticidal product. Knowledge of the mechanism of action makes it possible to understand which non-target organisms could be harmed by the use of such products [145,159]. In addition, this information facilitates prospecting other possibly more active materials using biotechnology tools and in silico models [160,161,162]. However, in general it is not easy to understand the mechanisms of action of plant natural products. Normally there are multiple modes of action pertaining to the complex composition of the materials [30,49,145], that usually occur in different target sites, as described for Piper spp, Derris (Lonchocarpus) urucu, Asarum heterotropoides and Dalbergia brasiliensis (Table 5) [89,100,109,148].
During the development stage, it is essential to evaluate toxicity in non-target organisms for promising insecticides using suitable models, such as the fish embryo acute toxicity (FET) test [163]. This model has been proposed to determine the acute or lethal toxicity of materials in the embryonic stages of zebrafish (Danio rerio) and for environmental assessments [163,164]. In addition, it is important to consider other aquatic and terrestrial organisms according to the intended application location, such as fish, amphibians, bees, birds and mammals [159]. Considering that natural product insecticides have natural degradation mechanisms, they possibly present advantages in comparison with insecticides of synthetic origin [26,30,50].
In general, raw materials (essential oils, extracts and isolated compounds) from plant natural products are poorly soluble in water and do not persist in the environment, which complicates the application and reduces the effectiveness of the desired action [26,51,165]. Therefore, the use of pharmaceutical technology is of fundamental importance in the development of formulations. Among the techniques used, nanotechnology, encapsulation and use of hydrophobic matrices with an extended and controlled release system should be highlighted as they can prolong the residual effect of formulations [51,149,165,166,167] due to controlled release. Formulation development of natural products also poses a challenge for the application of these materials but it is imperative to improve both efficiency and cost-effectiveness [150,165]. Investing in botanical natural product formulations is an important advance in increasing the availability of commercial eco-friendly insecticides for Ae. aegypti control.

9. Conclusions

Considering the several stages of the insect development, the larvicidal test is the most evaluated bioassay in the search for insecticides to control Ae. aegypti for a number of reasons: (i) the larval phase is the longest in the immature stage; (ii) larvae are generally more sensitive to the toxic effects of compounds, and (iii) larvae breeding sites are localized and usually accessible. The search for ovicidal action is complex, especially due to its composition that hinders the toxic action of compounds. For the adult phase, there are compounds that cause toxicity by contact as well as those with repellent action.
Some very common edible botanical species such as Petroselinum crispum, Foeniculum vulgare, Curcuma longa, Mentha spicata, Ocimum gratissimium and Rosmarinus officinalis are highlighted, especially in the larval phase of Ae. Aegypti, due to their possible low toxicity to non-target organisms. However, other non-edible species have shown strong larvicidal extract activity, among them Echinops transiliensis, Piper ssp, Hypericum japonicum and Nerine sarniensis.
Essential oils provide a promising source for insecticidal applications due to their important insecticidal activities and possible toxicological safety for mammals and the environment. Moreover, they generally possess high oral and dermal LC50 values for these animals and are more readily degraded by natural ecosystem mechanisms.
Among the secondary metabolites, terpenes, especially monoterpenes, and phenylpropanoids are highlighted for larvicidal activity. These compounds are present in large quantities in essential oils. In addition, thiophenes, amides and alkaloids demonstrate high larvicidal and adulticidal activity.
Regarding the mechanisms of action, botanical natural products extracts, and pure compounds have displayed acitivities that include altering insect morphogenesis and therefore impairing the moulting process, respiration, feeding, and self-defense, among others. In addition, they altered biochemical processes and the nervous system.
Despite the limitations and obstacles to overcome, plant natural products are a suitable alternative source of eco-friendly botanical insecticides to control the Ae. aegypti mosquito, popularly known as dengue mosquito. Ever increasing mosquito resistance to conventional chemical insecticides warrants alternative products, which are safer for the environment and pose less risk to human health.

Author Contributions

M.R.S.S. and P.C.V. contribuited to the design and writing of the manuscript; P.C.V., L.S.E. and N.P.L. supervised and critically revised the final version of the manuscript; All authors have read and agreed to the published version of the manuscript.

Funding

The authors also acknowledge the Ministry of Health and University of Brasilia (TED 74/2016 and TED 42/2017), FAPESP CNPq.

Acknowledgments

The authors acknowledge the Ministry of Health, Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq); Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES; finance code 001) and The Research Network Natural Products against Neglected Diseases (ResNet NPND).

Conflicts of Interest

The authors have no conflict of interest to declare.

References

  1. Consoli, R.A.G.B.; Oliveira, R.L. Principais Mosquitos de Importância Sanitária no Brasil, 1st ed.; Fiocruz: Rio de Janeiro, Brazil, 1994; pp. 115–117. [Google Scholar]
  2. Kraemer, M.U.F.; Sinka, M.E.; Duda, K.A.; Mylne, A.Q.; Shearer, F.M.; Barker, C.M.; Moore, C.G.; Carvalho, R.G.; Coelho, G.E.; Van Bortel, W.; et al. The global distribution of the arbovirus vectors Aedes aegypti and Ae. albopictus. eLife 2015, 4, 08347. [Google Scholar] [CrossRef]
  3. Geris, R.; Ribeiro, P.R.; Brandão, M.D.S.; Da Silva, H.H.G.; Da Silva, I.G. Bioactive Natural Products as Potential Candidates to Control Aedes aegypti, the Vector of Dengue. In Bioactive Natural Products; Elsevier BV: Amsterdam, The Netherlands, 2012; Volume 37, pp. 277–376. [Google Scholar]
  4. Benelli, G.; Jeffries, C.L.; Walker, T. Biological Control of Mosquito Vectors: Past, Present, and Future. Insects 2016, 7, 52. [Google Scholar] [CrossRef]
  5. World Health Organization. Global Strategy for Dengue Prevention and Control. Available online: https://apps.who.int/iris/bitstream/handle/10665/75303/9789241504034_eng.pdf;jsessionid=59CD5C25000FE7F9ABE29B06C33D6461?sequence=1 (accessed on 27 November 2019).
  6. Bhatt, S.; Gething, P.W.; Brady, O.J.; Messina, J.P.; Farlow, A.W.; Moyes, C.L.; Drake, J.M.; Brownstein, J.S.; Hoen, A.G.; Sankoh, O.; et al. The global distribution and burden of dengue. Nature 2013, 496, 504–507. [Google Scholar] [CrossRef] [PubMed]
  7. World Health Organization. Dengue Guidelines for Diagnosis, Treatment, Prevention and Control. Available online: https://apps.who.int/iris/bitstream/handle/10665/44188/9789241547871_eng.pdf?sequence=1&isAllowed=y (accessed on 2 December 2019).
  8. World Health Organization. Zika Virus: An Epidemiological Update. Available online: https://apps.who.int/iris/bitstream/handle/10665/255008/WER9215.pdf?sequence=1 (accessed on 5 December 2019).
  9. World Health Organization. Chikungunya Disease: Gaps and Opportunities in Public Health and Research in the Americas. Available online: https://www.who.int/wer/2015/wer9042.pdf?ua=1 (accessed on 9 December 2019).
  10. Shepard, D.S.; Undurraga, E.A.; Halasa, Y.A.; Stanaway, J.D. The global economic burden of dengue: A systematic analysis. Lancet Infect. Dis. 2016, 16, 935–941. [Google Scholar] [CrossRef]
  11. Teich, V.; Arinelli, R.; Fahham, L. Aedes aegypti e sociedade: O impacto econômico das arboviroses no Brasil. J. Bras. Econ. 2017, 9, 267–276. [Google Scholar] [CrossRef]
  12. Messina, J.; Brady, O.J.; Golding, N.; Kraemer, M.U.F.; Wint, G.R.W.; Ray, S.E.; Pigott, D.M.; Shearer, F.M.; Johnson, K.; Earl, L.; et al. The current and future global distribution and population at risk of dengue. Nat. Microbiol. 2019, 4, 1508–1515. [Google Scholar] [CrossRef] [PubMed]
  13. Benelli, G. Research in mosquito control: Current challenges for a brighter future. Parasitol. Res. 2015, 114, 2801–2805. [Google Scholar] [CrossRef]
  14. Isman, M.B. A renaissance for botanical insecticides? Pest Manag. Sci. 2015, 71, 1587–1590. [Google Scholar] [CrossRef]
  15. Lees, R.; Gilles, J.R.; Hendrichs, J.; Vreysen, M.J.B.; Bourtzis, K. Back to the future: The sterile insect technique against mosquito disease vectors. Curr. Opin. Insect Sci. 2015, 10, 156–162. [Google Scholar] [CrossRef] [Green Version]
  16. Alphey, L.; McKemey, A.; Nimmo, D.; Oviedo, M.N.; Lacroix, R.; Matzen, K.; Beech, C. Genetic control of Aedes mosquitoes. Pathog. Glob. Heal. 2013, 107, 170–179. [Google Scholar] [CrossRef] [Green Version]
  17. Evans, H.C.; Elliot, S.L.; Barreto, R.W. Entomopathogenic fungi and their potential for the management of Aedes aegypti (Diptera: Culicidae) in the Americas. Mem. Inst. Oswaldo Cruz 2018, 113, 206–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Durovni, B.; Saraceni, V.; Eppinghaus, A.; Riback, T.I.S.; Neill, S.L.O.; Moreira, L.A.; Jewell, N.P.; Dufault, S.M.; Simmons, C.P.; Tanamas, S.K.; et al. The impact of large-scale deployment of Wolbachia mosquitoes on arboviral disease incidence in Rio de Janeiro and Niterói, Brazil: Study protocol for a controlled interrupted time series analysis using routine disease surveillance data. F1000 Res. 2019, 8, 1328. [Google Scholar] [CrossRef] [Green Version]
  19. Sarwar, M. Control of dengue carrier Aedes mosquitoes (Diptera: Culicidae) larvae by larvivorous fishes and putting it into practice within water bodies. Int. J. Prev. Med. Res. 2015, 1, 232–237. [Google Scholar]
  20. Cavalcanti, L.P.D.G.; Pontes, R.J.S.; Regazzi, A.C.F.; de Paula Júnior, F.J.; Frutuoso, R.L.; Sousa, E.P.; Dantas Filho, F.D.; Lima, J.W.D.O. Efficacy of fish as predators of Aedes aegypti larvae, under laboratory conditions. Rev. Saude Pública 2007, 41, 638–644. [Google Scholar] [CrossRef] [PubMed]
  21. Nam, V.S.; Yen, N.T.; Duc, H.M.; Tu, T.C.; Thang, V.T.; Le, N.H.; San, L.H.; Loan, L.L.; Huong, V.T.Q.; Khanh, L.H.K.; et al. Community-based control of Aedes aegypti by using Mesocyclops in southern Vietnam. Am. J. Trop. Med. Hyg. 2012, 86, 850–859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Torres-Estrada, J.L.; Rodríguez, M.H.; Cruz-López, L.; Arredondo-Jimenez, J.I. Selective oviposition by Aedes aegypti (Diptera: Culicidae) in response to Mesocyclops longisetus (Copepoda: Cyclopoidea) under laboratory and field conditions. J. Med. Entomol. 2001, 38, 188–192. [Google Scholar] [CrossRef] [PubMed]
  23. Albeny, D.S.; Martins, G.F.; Andrade, M.R.; Krüger, R.F.; Vilela, E.F. Aedes aegypti survival in the presence of Toxorhynchites violaceus (Diptera: Culicidae) fourth instar larvae. Zoologia 2011, 28, 538–540. [Google Scholar] [CrossRef] [Green Version]
  24. Zuharah, W.F.; Fadzly, N.; Yusof, N.A.; Dieng, H.; Ni, X. Risky behaviors: Effects of Toxorhynchites splendens (Diptera: Culicidae) predator on the behavior of three mosquito species. J. Insect Sci. 2015, 15, 128. [Google Scholar] [CrossRef] [Green Version]
  25. Benelli, G. Plant-borne ovicides in the fight against mosquito vectors of medical and veterinary importance: A systematic review. Parasitol. Res. 2015, 114, 3201–3212. [Google Scholar] [CrossRef]
  26. Pavela, R. Essential oils for the development of eco-friendly mosquito larvicides: A review. Ind. Crops Prod. 2015, 76, 174–187. [Google Scholar] [CrossRef]
  27. Lughadha, E.N.; Govaerts, R.; Belyaeva, I.; Black, N.; Lindon, H.; Allkin, R.; Magill, R.E.; Nicolson, N. Counting counts: Revised estimates of numbers of accepted species of flowering plants, seed plants, vascular plants and land plants with a review of other recent estimates. Phytotaxa 2016, 272, 82–88. [Google Scholar] [CrossRef] [Green Version]
  28. Zebitz, C.P.W. Effects of three different neem seed kernel extracts and azadirachtin on larvae of different mosquito species. J. Appl. Entomol. 1986, 102, 455–463. [Google Scholar] [CrossRef]
  29. Sharma, R.N.; Bhosale, A.S.; Joshi, V.N.; Hebbalkar, D.S.; Tungikar, V.B.; Gupta, A.S.; Patwardhan, S.A. Lavandula gibsonii: A plant with insectistatic potential. Phytoparasitica 1981, 9, 101–109. [Google Scholar] [CrossRef]
  30. Pavela, R. History, presence and perspective of using plant extracts as commercial botanical insecticides and farm products for protection against Insects—A review. Plant Prot. Sci. 2016, 52, 229–241. [Google Scholar]
  31. Yanola, J.; Somboon, P.; Walton, C.; Nachaiwieng, W.; Prapanthadara, L.-a. A novel F1552/C1552 point mutation in the Aedes aegypti voltage-gated sodium channel gene associated with permethrin resistance. Pestic. Biochem. Physiol. 2010, 96, 127–131. [Google Scholar] [CrossRef]
  32. Saavedra-Rodriguez, K.; Urdaneta-Marquez, L.; Rajatileka, S.; Moulton, M.; Flores, A.E.; Fernandez-Salas, I.; Bisset, J.; Rodriguez, M.; Mccall, P.J.; Donnelly, M.J.; et al. A mutation in the voltage-gated sodium channel gene associated with pyrethroid resistance in Latin American Aedes aegypti. Insect Mol. Biol. 2007, 16, 785–798. [Google Scholar] [CrossRef] [Green Version]
  33. Fundação Osvaldo Cruz. Recomendação Técnica Sobre a Interrupção do Uso de Inseticidas Piretróides no Controle do Aedes aegypti no Brasil. Available online: http://www.fiocruz.br/ioc/media/NotaTecnica%202%20IOC%20v1%20dengue%206%20set%202011.pdf (accessed on 18 December 2019).
  34. Hemingway, J.; Ranson, H. Insecticide resistance in insect vectors of human diseases. Annu. Rev. Entomol. 2000, 45, 371–391. [Google Scholar] [CrossRef]
  35. Priester, T.M.; Georghiou, G.P. Penetration of permethrin and knockdown in larvae of pyrethroid-resistant and -susceptible strains of the southern house mosquito. J. Econ. Entomol. 1980, 73, 165–167. [Google Scholar] [CrossRef]
  36. Raymond, M.; Chevillon, C.; Guillemaud, T.; Lenormand, T.; Pasteur, N. An overview of the evolution of overproduced esterases in the mosquito Culex pipiens. Philos. Trans. R. Soc. Lond. B 1998, 353, 1707–1711. [Google Scholar] [CrossRef] [Green Version]
  37. Cariño, F.A.; Koener, J.F.; Plapp, F.W.; Feyereisen, R. Constitutive overexpression of the cytochrome P450 gene CYP6A1 in a house fly strain with metabolic resistance to insecticides. Insect Biochem. Mol. Biol. 1994, 24, 411–418. [Google Scholar] [CrossRef]
  38. Mutero, A.; Pralavorio, M.; Bride, J.M.; Fournier, D. Resistance-associated point mutations in insecticide-insensitive acetylcholinesterase. Proc. Natl. Acad. Sci. USA 1994, 91, 5922–5926. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Bisset, J.; Rodríguez, M.M.; Fernández, D. Selection of insensitive acetylcholinesterase as a resistance mechanism in Aedes aegypti (Diptera: Culicidae) from Santiago de Cuba. J. Med. Entomol. 2006, 43, 1185–1189. [Google Scholar] [CrossRef] [PubMed]
  40. Cui, F.; Li, M.X.; Chang, H.J.; Mao, Y.; Zhang, H.Y.; Lu, L.X.; Yan, S.G.; Lang, M.L.; Liu, L.; Qiao, C.L. Carboxylesterase-mediated insecticide resistance: Quantitative increase induces broader metabolic resistance than qualitative change. Pestic. Biochem. Physiol. 2015, 121, 88–96. [Google Scholar] [CrossRef] [PubMed]
  41. Anthony, N.; Rocheleau, T.; Mocelin, G.; Lee, H.J.; Ffrench-Constant, R. Cloning, sequencing and functional expression of an acetylcholinesterase gene from the yellow fever mosquito Aedes aegypti. FEBS Lett. 1995, 368, 461–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Lacorte, S.; Ehresmann, N.; Barceló, D. Persistence of temephos and its transformation products in rice crop field waters. Environ. Sci. Technol. 1996, 30, 917–923. [Google Scholar] [CrossRef]
  43. Zhang, A.; Xie, X.; Ye, J.; Lin, C.; Hu, X. Stereoselective toxicity of malathion and its metabolites, malaoxon and isomalathion. Environ. Chem. Lett. 2011, 9, 369–373. [Google Scholar] [CrossRef]
  44. Park, H.-M.; Kim, J.; Chang, K.-S.; Kim, B.-S.; Yang, Y.-J.; Kim, G.-H.; Shin, S.-C.; Park, I.-K. Larvicidal activity of Myrtaceae essential oils and their components against Aedes aegypti, acute toxicity on Daphnia magna, and aqueous residue. J. Med. Entomol. 2011, 48, 405–410. [Google Scholar] [CrossRef]
  45. Seo, S.M.; Park, H.M.; Park, I.K. Larvicidal activity of ajowan (Trachyspermum ammi) and Peru balsam (Myroxylon pereira) oils and blends of their constituents against mosquito, Aedes aegypti, acute toxicity on water flea, Daphnia magna, and aqueous residue. J. Agric. Food Chem. 2012, 60, 5909–5914. [Google Scholar] [CrossRef]
  46. Williams, D.H.; Stone, M.J.; Hauck, P.R.; Rahman, S.K. Why are secondary metabolites (natural products) biosynthesized. J. Nat. Prod. 1989, 52, 1189–1208. [Google Scholar] [CrossRef]
  47. Isman, M.B. Bridging the gap: Moving botanical insecticides from the laboratory to the farm. Ind. Crops Prod. 2017, 110, 10–14. [Google Scholar] [CrossRef]
  48. Miresmailli, S.; Isman, M.B. Botanical insecticides inspired by plant–herbivore chemical interactions. Trends Plant Sci. 2014, 19, 29–35. [Google Scholar] [CrossRef] [PubMed]
  49. Isman, M.B.; Miresmailli, S.; MacHial, C. Commercial opportunities for pesticides based on plant essential oils in agriculture, industry and consumer products. Phytochem. Rev. 2011, 10, 197–204. [Google Scholar] [CrossRef]
  50. Isman, M.B. Plant essential oils for pest and disease management. Crop Prot. 2000, 19, 603–608. [Google Scholar] [CrossRef]
  51. Pavela, R. Encapsulation—A convenient way to extend the persistence of the effect of eco-friendly mosquito larvicides. Curr. Org. Chem. 2016, 20, 2674–2680. [Google Scholar] [CrossRef] [Green Version]
  52. Benelli, G. Plant-mediated biosynthesis of nanoparticles as an emerging tool against mosquitoes of medical and veterinary importance: A review. Parasitol. Res. 2016, 115, 23–34. [Google Scholar] [CrossRef] [PubMed]
  53. Ashokan, A.P.; Paulpandi, M.; Dinesh, D.; Murugan, K.; Vadivalagan, C.; Benelli, G. Toxicity on Dengue mosquito vectors through Myristica fragrans—synthesized zinc oxide nanorods, and their cytotoxic effects on liver cancer cells (HepG2). J. Clust. Sci. 2017, 28, 205–226. [Google Scholar] [CrossRef]
  54. Gomes da Rocha Voris, D.; Dos Santos Dias, L.; Alencar Lima, J.; Dos Santos Cople Lima, K.; Pereira Lima, J.B.; Dos Santos Lima, A.L. Evaluation of larvicidal, adulticidal, and anticholinesterase activities of essential oils of Illicium verum Hook. f., Pimenta dioica (L.) Merr., and Myristica fragrans Houtt. against Zika virus vectors. Environ. Sci. Pollut. Res. 2018, 25, 22541–22551. [Google Scholar] [CrossRef] [PubMed]
  55. Cheng, S.; Chang, H.; Chang, S. Bioactivity of selected plant essential oils against the yellow fever mosquito Aedes aegypti larvae. Bioresour. Technol. 2003, 89, 99–102. [Google Scholar] [CrossRef]
  56. Kiran, S.R.; Bhavani, K.; Devi, P.S.; Rao, B.R.R.; Reddy, K.J. Composition and larvicidal activity of leaves and stem essential oils of Chloroxylon swietenia DC against Aedes aegypti and Anopheles stephensi. Bioresour. Technol. 2006, 97, 2481–2484. [Google Scholar] [CrossRef]
  57. Komalamisra, N.; Trongtokit, Y.; Rongsriyam, Y.; Apiwathnasorn, C. Screening for lavicidal activity in some Thai plants against four mosquito vector species. Southeast Asian J. Trop. Med. Public Health 2005, 36, 1412–1422. [Google Scholar]
  58. Vivekanandhan, P.; Venkatesan, R.; Ramkumar, G.; Karthi, S.; Senthil-Nathan, S.; Shivakumar, M.S. Comparative analysis of major mosquito vectors response to seed-derived essential oil and seed pod-derived extract from Acacia nilotica. Int. J. Environ. Res. Public Health 2018, 15, 388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Ali, A.; Akhtar, N.; Khan, B.A.; Khan, M.S.; Rasul, A. Acacia nilotica: A plant of multipurpose medicinal uses. J. Med. Plants Res. 2012, 6, 1492–1496. [Google Scholar] [CrossRef]
  60. Intirach, J.; Junkum, A.; Lumjuan, N.; Chaithong, U.; Jitpakdi, A.; Riyong, D.; Wannasan, A.; Champakaew, D.; Muangmoon, R.; Chansang, A.; et al. Antimosquito property of Petroselinum crispum (Umbellifereae) against the pyrethroid resistant and susceptible strains of Aedes aegypti (Diptera: Culicidae). Environ. Sci. Pollut. Res. 2016, 23, 23994–24008. [Google Scholar] [CrossRef] [PubMed]
  61. Cheng, S.S.; Huang, C.G.; Chen, Y.J.; Yu, J.J.; Chen, W.J.; Chang, S.T. Chemical compositions and larvicidal activities of leaf essential oils from two Eucalyptus species. Bioresour. Technol. 2009, 100, 452–456. [Google Scholar] [CrossRef]
  62. Costa, A.A.; Naspi, C.V.; Lucia, A.; Masuh, H.M.; Barrera, R. Repellent and larvicidal activity of the essential oil from Eucalyptus nitens against Aedes aegypti and Aedes albopictus (Diptera: Culicidae). J. Med. Entomol. 2017, 54, 670–676. [Google Scholar] [CrossRef]
  63. Gillij, Y.G.; Gleiser, R.M.; Zygadlo, J.A. Mosquito repellent activity of essential oils of aromatic plants growing in Argentina. Bioresour. Technol. 2008, 99, 2507–2515. [Google Scholar] [CrossRef]
  64. Cheng, S.-S.; Liu, J.-Y.; Tsai, K.-H.; Chen, W.-J.; Chang, S.-T. Chemical composition and mosquito larvicidal activity of essential oils from leaves of different Cinnamomum osmophloeum provenances. J. Agric. Food Chem. 2004, 52, 4395–4400. [Google Scholar] [CrossRef]
  65. Mendes, L.A.; Martins, G.F.; Valbon, W.R.; da Silva de Souza, T.; Menini, L.; Ferreira, A.; da Silva Ferreira, M.F. Larvicidal effect of essential oils from Brazilian cultivars of Guava on Aedes aegypti L. Ind. Crops Prod. 2017, 108, 684–689. [Google Scholar] [CrossRef]
  66. Oliveira, A.E.; Duarte, J.L.; Amado, J.R.R.; Cruz, R.A.S.; Rocha, C.F.; Souto, R.N.P.; Ferreira, R.M.A.; Santos, K.; da Conceição, E.C.; de Oliveira, L.A.R.; et al. Development of a larvicidal nanoemulsion with Pterodon emarginatus Vogel oil. PLoS ONE 2016, 11, e0145835. [Google Scholar] [CrossRef]
  67. 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]
  68. Botas, G.D.S.; Cruz, R.A.S.; de Almeida, F.B.; Duarte, J.L.; Araújo, R.S.; Souto, R.N.P.; Ferreira, R.; Carvalho, J.C.T.; Santos, M.G.; Rocha, L.; et al. Baccharis reticularia DC. and limonene nanoemulsions: Promising larvicidal agents for Aedes aegypti (Diptera: Culicidae) control. Molecules 2017, 22, 1990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Duarte, J.L.; Amado, J.R.R.; Oliveira, A.E.M.F.M.; Cruz, R.A.S.; Ferreira, A.M.; Souto, R.N.P.; Falcão, D.Q.; Carvalho, J.C.T.; Fernandes, C.P. Evaluation of larvicidal activity of a nanoemulsion of Rosmarinus officinalis essential oil. Braz. J. Pharmacog. 2015, 25, 189–192. [Google Scholar] [CrossRef] [Green Version]
  70. Santos, G.K.N.; Dutra, K.A.; Lira, C.S.; Lima, B.N.; Napoleão, T.H.; Paiva, P.M.G.; Maranhão, C.A.; Brandão, S.S.F.; Navarro, D.M.A.F. Effects of Croton rhamnifolioides essential oil on Aedes aegypti oviposition, larval toxicity and trypsin activity. Molecules 2014, 19, 16573–16587. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Cheng, S.S.; Lin, C.Y.; Chung, M.J.; Liu, Y.H.; Huang, C.G.; Chang, S.T. Larvicidal activities of wood and leaf essential oils and ethanolic extracts from Cunninghamia konishii Hayata against the Dengue mosquitoes. Ind. Crops Prod. 2013, 47, 310–315. [Google Scholar] [CrossRef]
  72. Govindarajan, M.; Sivakumar, R.; Rajeswari, M.; Yogalakshmi, K. Chemical composition and larvicidal activity of essential oil from Mentha spicata (Linn.) against three mosquito species. Parasitol. Res. 2012, 110, 2023–2032. [Google Scholar] [CrossRef]
  73. Fayemiwo, K.A.; Adeleke, M.A.; Okoro, O.P.; Awojide, S.H.; Awoniyi, I.O.; Ademolu, K.O. Larvicidal efficacies and chemical composition of essential oils of Pinus sylvestris and Syzygium aromaticum against mosquitoes. Asian Pac. J. Trop. Biomed. 2014, 4, 30–34. [Google Scholar] [CrossRef]
  74. Oliveira, G.; Cardoso, S.; Junior, C.L. Chemical study and larvicidal activity against Aedes aegypti of essential oil of Piper aduncum L. (Piperaceae). An. Acad. Bras. Cienc. 2013, 85, 1227–1234. [Google Scholar] [CrossRef] [Green Version]
  75. Fontes, U.R., Jr.; Ramos, C.S.; Serafini, M.R.; Cavalcanti, S.C.H.; Alves, P.B.; Lima, G.M.; Andrade, P.H.S.; Bonjardim, L.R.; Quintans, L.J., Jr.; Araujo, A.A.S. Evaluation of the lethality of Porophyllum ruderale essential oil against Biomphalaria glabrata, Aedes aegypti and Artemia salina. Afr. J. Biotechnol. 2012, 11, 3169–3172. [Google Scholar] [CrossRef]
  76. Ali, A.; Tabanca, N.; Kurkcuoglu, M.; Duran, A.; Blythe, E.K.; Khan, I.A.; Baser, K.H.C. Chemical composition, larvicidal, and biting deterrent activity of essential oils of two subspecies of Tanacetum argenteum (Asterales: Asteraceae) and individual constituents against Aedes aegypti (Diptera: Culicidae). J. Med. Entomol. 2014, 51, 824–830. [Google Scholar] [CrossRef] [PubMed]
  77. Chellappandian, M.; Thanigaivel, A.; Vasantha-Srinivasan, P.; Edwin, E.S.; Ponsankar, A.; Selin-Rani, S.; Kalaivani, K.; Senthil-Nathan, S.; Benelli, G. Toxicological effects of Sphaeranthus indicus Linn. (Asteraceae) leaf essential oil against human disease vectors, Culex quinquefasciatus Say and Aedes aegypti Linn., and impacts on a beneficial mosquito predator. Environ. Sci. Pollut. Res. 2018, 25, 10294–10306. [Google Scholar] [CrossRef]
  78. De Sousa, L.M.; de Carvalho, J.L.; Gois, R.W.S.; da Silva, H.C.; Santiago, G.M.P.; Lemos, T.L.G.; Arriaga, A.M.C.; Alves, P.B.; de Matos, I.L.; Militão, G.C.G.; et al. Chemical composition, larvicidal and cytotoxic activities of the essential oils from two Bauhinia species. Rec. Nat. Prod. 2016, 10, 341–348. [Google Scholar]
  79. Kerdudo, A.; Gonnot, V.; Ellong, E.N.; Boyer, L.; Chandre, F.; Adenet, S.; Rochefort, K.; Michel, T.; Fernandez, X. Composition and bioactivity of Pluchea carolinensis (Jack.) G. essential oil from Martinique. Ind. Crops Prod. 2016, 89, 295–302. [Google Scholar] [CrossRef]
  80. Kovendan, K.; Murugan, K.; Vincent, S. Evaluation of larvicidal activity of Acalypha alnifolia Klein ex Willd. (Euphorbiaceae) leaf extract against the malarial vector, Anopheles stephensi, dengue vector, Aedes aegypti and Bancroftian filariasis vector, Culex quinquefasciatus (Diptera: Culicidae). Parasitol. Res. 2012, 110, 571–581. [Google Scholar] [CrossRef] [PubMed]
  81. Krishnappa, K.; Elumalai, K. Toxicity of Aristolochia bracteata methanol leaf extract against selected medically important vector mosquitoes (Diptera: Culicidae). Asian Pacific J. Trop. Dis. 2012, 2, S553–S557. [Google Scholar] [CrossRef]
  82. Aziz, A.T.; Alshehri, M.A.; Panneerselvam, C.; Murugan, K.; Trivedi, S.; Mahyoub, J.A.; Hassan, M.M.; Maggi, F.; Sut, S.; Dall’Acqua, S.; et al. The desert wormwood (Artemisia herba-alba)—From arabian folk medicine to a source of green and effective nanoinsecticides against mosquito vectors. J. Photochem. Photobiol. B Biol. 2018, 180, 225–234. [Google Scholar] [CrossRef]
  83. Alam, M.F.; Safhi, M.M.; Chopra, A.K.; Dua, V.K. Toxicological properties of several medicinal plants from the Himalayas (India) against vectors of malaria, filariasis and dengue. Trop. Biomed. 2011, 28, 343–350. [Google Scholar]
  84. El-Gamal, A.; Al-Massarani, S.; Fawzy, G.; Ati, H.; Al-Rehaily, A.; Basudan, O.; Abdel-Kader, M.; Tabanca, N.; Becnel, J. Chemical composition of Buddleja polystachya aerial parts and its bioactivity against Aedes aegypti. Nat. Prod. Res. 2017, 32, 2775–2782. [Google Scholar] [CrossRef]
  85. Govindarajan, M.; Mathivanan, T.; Elumalai, K.; Krishnappa, K.; Anandan, A. Mosquito larvicidal, ovicidal, and repellent properties of botanical extracts against Anopheles stephensi, Aedes aegypti, and Culex quinquefasciatus (Diptera: Culicidae). Parasitol. Res. 2011, 109, 353–367. [Google Scholar] [CrossRef]
  86. Govindarajan, M. Bioefficacy of Cassia fistula Linn. (Leguminosae) leaf extract against chikungunya vector, Aedes aegypti (Diptera: Culicidae). Eur. Rev. Med. Pharmacol. Sci. 2009, 13, 99–103. [Google Scholar]
  87. Inocente, E.A.; Shaya, M.; Acosta, N.; Rakotondraibe, L.H.; Piermarini, P.M. A natural agonist of mosquito TRPA1 from the medicinal plant Cinnamosma fragrans that is toxic, antifeedant, and repellent to the yellow fever mosquito Aedes aegypti. PLoS Negl. Trop. Dis. 2018, 12, e0006265. [Google Scholar] [CrossRef] [Green Version]
  88. Rahuman, A.A.; Venkatesan, P. Larvicidal efficacy of five Cucurbitaceous plant leaf extracts against mosquito species. Parasitol. Res. 2008, 103, 133–139. [Google Scholar] [CrossRef]
  89. Dalarmi, L.; Silva, C.B.d.; Ocampos, F.M.; Burci, M.L.; Nascimento, K.F.d.; Jesus, C.d.; Dias Gaspari, J.d.F.; Miguel, M.D.; Miguel, O.G.; Zanin, S.M.W. Larvicidal activity of Dalbergia brasiliensis (Fabaceae—Papilionoideae). Afr. J. Pharm. Pharmacol. 2015, 9, 881–885. [Google Scholar]
  90. Nakano, H.; Ali, A.; Ur Rehman, J.; Mamonov, L.K.; Cantrell, C.L.; Khan, I.A. Toxicity of thiophenes from Echinops transiliensis (Asteraceae) against Aedes aegypti (Diptera: Culicidae) larvae. Chem. Biodivers. 2014, 11, 1001–1009. [Google Scholar] [CrossRef]
  91. Govindarajan, M.; Karuppannan, P. Mosquito larvicidal and ovicidal properties of Eclipta alba (L.) Hassk (Asteraceae) against chikungunya vector, Aedes aegypti (Linn.) (Diptera: Culicidae). Asian Pac. J. Trop. Med. 2011, 4, 24–28. [Google Scholar] [CrossRef] [Green Version]
  92. Rahuman, A.A.; Gopalakrishnan, G.; Venkatesan, P.; Geetha, K. Larvicidal activity of some Euphorbiaceae plant extracts against Aedes aegypti and Culex quinquefasciatus (Diptera: Culicidae). Parasitol. Res. 2008, 102, 867–873. [Google Scholar] [CrossRef]
  93. Govindarajan, M. Larvicidal efficacy of Ficus benghalensis L. plant leaf extracts against Culex quinquefasciatus Say, Aedes aegypti L. and Anopheles stephensi L. (Diptera: Culicidae). Eur. Rev. Med. Pharmacol. Sci. 2010, 14, 107–111. [Google Scholar]
  94. Ochieng, C.O.; Mid, J.O.; Owu, P.O. Anti-plasmodial and larvicidal effects of surface exudates of Gardenia ternifolia aerial parts. Res. J. Pharmacol. 2010, 4, 45–50. [Google Scholar] [CrossRef] [Green Version]
  95. Dos Santos, E.A.; de Carvalho, C.M.D.; Costa, A.L.S.; Conceição, A.S.; Moura, F.D.B.P.; Santana, A.E.G. Bioactivity evaluation of plant extracts used in indigenous medicine against the snail, Biomphalaria glabrata, and the larvae of Aedes aegypti. Evid. Based Complement. Altern. Med. 2012, 2012, 846583. [Google Scholar] [CrossRef] [Green Version]
  96. Kumuda, S.S.; Mohankumar, T.K.; Prathibha, K.P.; Vijayan, V.A. Efficacy of plant extracts against the larvae of filariasis vector, Culex quinquefasciatus Say and the Dengue vector Aedes aegypti Linn at Mysore. Int. J. Curr. Microbiol. Appl. Sci. 2015, 4, 242–249. [Google Scholar]
  97. Puthur, S.; Anoopkumar, A.N.; Rebello, S.; Aneesh, E.M. Hypericum japonicum: A double-headed sword to combat vector control and cancer. Appl. Biochem. Biotechnol. 2018, 186, 1–11. [Google Scholar] [CrossRef]
  98. Komansilan, A.; Suriani, N.W.; Lawalata, H. Test toxic tuba root extract as a natural insecticide on larvae of Aedes aegypti mosquito vector of dengue fever. Int. J. ChemTech Res. 2017, 10, 522–528. [Google Scholar]
  99. Reegan, A.D.; Gandhi, M.R.; Paulraj, M.G.; Balakrishna, K.; Ignacimuthu, S. Effect of niloticin, a protolimonoid isolated from Limonia acidissima L. (Rutaceae) on the immature stages of dengue vector Aedes aegypti L. (Diptera: Culicidae). Acta Trop. 2014, 139, 67–76. [Google Scholar] [CrossRef]
  100. Gusmão, D.S.; Páscoa, V.; Mathias, L.; Vieira, I.J.C.; Braz-Filho, R.; Lemos, F.J.A. Derris (Lonchocarpus) urucu (Leguminosae) extract modifies the peritrophic matrix structure of Aedes aegypti (Diptera: Culicidae). Mem. Inst. Oswaldo Cruz 2002, 97, 371–375. [Google Scholar] [CrossRef] [Green Version]
  101. Touré, S.; Nirma, C.; Falkowski, M.; Dusfour, I.; Boulogne, I.; Jahn-Oyac, A.; Coke, M.; Azam, D.; Girod, R.; Moriou, C.; et al. Aedes aegypti larvicidal sesquiterpene alkaloids from Maytenus oblongata. J. Nat. Prod. 2017, 80, 384–390. [Google Scholar] [CrossRef]
  102. Lalchhandama, K. Mosquitocidal activity of Millettia pachycarpa on the larvae and eggs of Aedes aegypti. Ann. Biol. Res. 2011, 2, 217–222. [Google Scholar]
  103. Govindarajan, M.; Ramya, A.; Sivakumar, R. Mosquito larvicidal properties of Mirabilis jalapa (Nyctaginaceae) against Anopheles stephensi, Aedes aegypti & Culex quinquefasciatus (Diptera: Culicidae). Indian J. Med. Res. 2014, 140, 438–440. [Google Scholar]
  104. Masi, M.; van der Westhuyzen, A.E.; Tabanca, N.; Evidente, M.; Cimmino, A.; Green, I.R.; Bernier, U.R.; Becnel, J.J.; Bloomquist, J.R.; van Otterlo, W.A.L.; et al. Sarniensine, a mesembrine-type alkaloid isolated from Nerine sarniensis, an indigenous South African Amaryllidaceae, with larvicidal and adulticidal activities against Aedes aegypti. Fitoterapia 2017, 116, 34–38. [Google Scholar] [CrossRef] [Green Version]
  105. Anees, A.M. Larvicidal activity of Ocimum sanctum Linn. (Labiatae) against Aedes aegypti (L.) and Culex quinquefasciatus (Say). Parasitol. Res. 2008, 103, 1451–1453. [Google Scholar] [CrossRef]
  106. Porto, K.R.d.A.; Motti, P.R.; Yano, M.; Roel, A.R.; Cardoso, C.A.L.; Matias, R. Screening of plant extracts and fractions on Aedes aegypti larvae found in the state of Mato Grosso do Sul (Linnaeus, 1762) (Culicidae). An. Acad. Bras. Cienc. 2017, 89, 895–906. [Google Scholar] [CrossRef] [Green Version]
  107. Kovendan, K.; Murugan, K.; Vincent, S.; Barnard, D.R. Mosquito larvicidal properties of Orthosiphon thymiflorus (Roth) Sleesen. (Family: Labiatae) against mosquito vectors, Anopheles stephensi, Culex quinquefasciatus and Aedes aegypti (Diptera: Culicidae). Asian Pac. J. Trop. Med. 2012, 5, 299–305. [Google Scholar] [CrossRef] [Green Version]
  108. Samidurai, K.; Jebanesan, A.; Saravanakumar, A.; Govindarajan, M.; Pushpanathan, T. Larvicidal, ovicidal and repellent activities of Pemphis acidula Forst. (Lythraceae) against filarial and dengue vector mosquitoes. Acad. J. Entomol. 2009, 2, 62–66. [Google Scholar]
  109. Chaitong, U.; Choochote, W.; Kamsuk, K.; Jitpakdi, A.; Tippawangkosol, P.; Chaiyasit, D.; Champakaew, D.; Tuetun, B.; Pitasawat, B. Larvicidal effect of pepper plants on Aedes aegypti (L.) (Diptera: Culicidae). J. Vector Ecol. 2006, 31, 138–144. [Google Scholar] [CrossRef]
  110. Santiago, V.S.; Alvero, R.G.; Villaseñor, I.M. Aedes aegypti larvicide from the ethanolic extract of Piper nigrum black peppercorns. Nat. Prod. Res. 2015, 29, 441–443. [Google Scholar] [CrossRef]
  111. Govindarajan, M.; Rajeswary, M.; Sivakumar, R. Larvicidal & ovicidal efficacy of Pithecellobium dulce (Roxb.) Benth. (Fabaceae) against Anopheles stephensi Liston & Aedes aegypti Linn. (Diptera: Culicidae). Indian J. Med. Res. 2013, 138, 129–134. [Google Scholar]
  112. Azarudeen, R.M.S.T.; Govindarajan, M.; AlShebly, M.M.; AlQahtani, F.S.; Amsath, A.; Benelli, G. One pot green synthesis of colloidal silver nanocrystals using the Ventilago maderaspatana leaf extract: Acute toxicity on malaria, zika virus and filariasis mosquito vectors. J. Clust. Sci. 2017, 28, 369–392. [Google Scholar] [CrossRef] [Green Version]
  113. Raghavendra, K.; Singh, S.P.; Subbarao, S.K.; Dash, A.P. Laboratory studies on mosquito larvicidal efficacy of aqueous & hexane extracts of dried fruit of Solanum nigrum Linn. Indian J. Med. Res. 2009, 130, 74–77. [Google Scholar]
  114. Kumar, P.M.; Murugan, K.; Kovendan, K.; Panneerselvam, C.; Kumar, K.P.; Amerasan, D.; Subramaniam, J.; Kalimuthu, K.; Nataraj, T. Mosquitocidal activity of Solanum xanthocarpum fruit extract and copepod Mesocyclops thermocyclopoides for the control of dengue vector Aedes aegypti. Parasitol. Res. 2012, 111, 609–618. [Google Scholar] [CrossRef]
  115. Munhoz, V.M.; Longhini, R.; Souza, J.R.P.; Zequi, J.A.C.; Mello, E.V.S.L.; Lopes, G.C.; Mello, J.C.P. Extraction of flavonoids from Tagetes patula: Process optimization and screening for biological activity. Braz. J. Pharmacog. 2014, 24, 576–583. [Google Scholar] [CrossRef] [Green Version]
  116. Kovendan, K.; Chandramohan, B.; Govindarajan, M.; Jebanesan, A.; Kamalakannan, S.; Vincent, S.; Benelli, G. Orchids as sources of novel nanoinsecticides? Efficacy of Bacillus sphaericus and Zeuxine gracilis—fabricated silver nanoparticles against dengue, malaria and filariasis mosquito vectors. J. Clust. Sci. 2018, 29, 345–357. [Google Scholar] [CrossRef]
  117. Govindarajan, M.; Mathivanan, T.; Elumalai, K.; Krishnappa, K.; Anandan, A. Ovicidal and repellent activities of botanical extracts against Culex quinquefasciatus, Aedes aegypti and Anopheles stephensi (Diptera: Culicidae). Asian Pac. J. Trop. Biomed. 2011, 1, 43–48. [Google Scholar] [CrossRef] [Green Version]
  118. Govindarajan, M.; Sivakumar, R. Repellent properties of Cardiospermum halicacabum Linn. (Family: Sapindaceae) plant leaf extracts against three important vector mosquitoes. Asian Pac. J. Trop. Med. 2012, 2, 602–607. [Google Scholar] [CrossRef] [Green Version]
  119. Govindarajan, M. Ovicidal and repellent properties of Coccinia indica Wight and Arn. (Family: Cucurbitaceae) against three important vector mosquitoes. Eur. Rev. Med. Pharmacol. Sci. 2011, 15, 1010–1019. [Google Scholar] [PubMed]
  120. Bhargava, S. Mosquito repellency potential of the essential oil from Mentha spp. Int. J. Pharm. Pharm. Res. 2016, 6, 509–522. [Google Scholar]
  121. Kumar, S.; Singh, A.P.; Nair, G.; Batra, S.; Seth, A.; Wahab, N.; Warikoo, R. Impact of Parthenium hysterophorus leaf extracts on the fecundity, fertility and behavioural response of Aedes aegypti L. Parasitol. Res. 2011, 108, 853–859. [Google Scholar] [CrossRef]
  122. Dick, F.D. Solvent Neurotoxicity. Occup. Environ. Med. 2006, 63, 221–226. [Google Scholar] [CrossRef]
  123. Vincent, J.F.V.; Wegst, U.G.K. Design and mechanical properties of insect cuticle. Arthropod Struct. Dev. 2004, 33, 187–199. [Google Scholar] [CrossRef]
  124. Fahmy, M.A.H.; Mallipudi, N.M.; Fukuto, T.R. Selective toxicity of N,N′-thiodicarbamates. J. Agric. Food Chem. 1978, 26, 550–557. [Google Scholar] [CrossRef]
  125. Ferreira, M.C.; Cantrell, C.L.; Duke, S.O.; Ali, A.; Rosa, L.H. New pesticidal diterpenoids from Vellozia gigantea (Velloziaceae), an endemic neotropical plant living in the endangered Brazilian biome rupestrian grasslands. Molecules 2017, 22, 175. [Google Scholar] [CrossRef] [Green Version]
  126. Da Silva, G.N.; Trindade, F.T.; Dos Santos, F.; Gosmann, G.; E Silva, A.A.; Gnoatto, S.C. Larvicidal activity of natural and modified triterpenoids against Aedes aegypti (Diptera: Culicidae). Pest Manag. Sci. 2016, 72, 1883–1887. [Google Scholar] [CrossRef]
  127. Masi, M.; Cimmino, A.; Tabanca, N.; Becnel, J.J.; Bloomquist, J.R.; Evidente, A. A survey of bacterial, fungal and plant metabolites against Aedes aegypti (Diptera: Culicidae), the vector of yellow and dengue fevers and zika virus. Open Chem. 2017, 15, 156–166. [Google Scholar] [CrossRef]
  128. Rahuman, A.A.; Gopalakrishnan, G.; Venkatesan, P.; Geetha, K.; Bagavan, A. Mosquito larvicidal activity of isolated compounds from the rhizome of Zingiber officinale. Phyther. Res. 2008, 22, 1035–1039. [Google Scholar] [CrossRef] [PubMed]
  129. Park, I.K.; Lee, S.G.; Shin, S.C.; Park, J.D.; Ahn, Y.J. Larvicidal activity of isobutylamides identified in Piper nigrum fruits against three mosquito species. J. Agric. Food Chem. 2002, 50, 1866–1870. [Google Scholar] [CrossRef] [PubMed]
  130. Park, I.K. Insecticidal activity of isobutylamides derived from Piper nigrum against adult of two mosquito species, Culex pipiens pallens and Aedes aegypti. Nat. Prod. Res. 2012, 26, 2129–2131. [Google Scholar] [CrossRef]
  131. Yang, Y.C.; Lee, S.G.; Lee, H.K.; Kim, M.K.; Lee, S.H.; Lee, H.S. A Piperidine amide extracted from Piper longum L. fruit shows activity against Aedes aegypti mosquito larvae. J. Agric. Food Chem. 2002, 50, 3765–3767. [Google Scholar] [CrossRef] [PubMed]
  132. Siddiqui, B.S.; Gulzar, T.; Begum, S.; Afshan, F.; Sattar, F.A. Insecticidal amides from fruits of Piper nigrum Linn. Nat. Prod. Res. 2005, 19, 143–150. [Google Scholar] [CrossRef] [PubMed]
  133. Masi, M.; Cala, A.; Tabanca, N.; Cimmino, A.; Green, I.R.; Bloomquist, J.R.; Van Otterlo, W.A.L.; Macias, F.A.; Evidente, A. Alkaloids with activity against the Zika virus vector Aedes aegypti (L.)-crinsarnine and sarniensinol, two new crinine and mesembrine type alkaloids isolated from the South African plant Nerine sarniensis. Molecules 2016, 21, 1432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Ma, H.-W.; Dong, C.-Z.; Zhou, X.-M.; Bu, M.-M.; Yu, S.-Q. Aristololactam derivatives from the fruits of Aristolochia contorta Bunge. Nat. Prod. Res. 2018, 32, 2505–2509. [Google Scholar] [CrossRef]
  135. Liu, X.C.; Lai, D.; Liu, Q.Z.; Zhou, L.; Liu, Q.; Liu, Z.L. Bioactivities of a new pyrrolidine alkaloid from the root barks of Orixa japonica. Molecules 2016, 21, 1665. [Google Scholar] [CrossRef] [Green Version]
  136. Sivakumar, R.; Jebanesan, A.; Govindarajan, M.; Rajasekar, P. Larvicidal and repellent activity of tetradecanoic acid against Aedes aegypti (Linn.) and Culex quinquefasciatus (Say.) (Diptera: Culicidae). Asian Pac. J. Trop. Med. 2011, 4, 706–710. [Google Scholar] [CrossRef] [Green Version]
  137. Sivakumar, R.; Jebanesan, A.; Govindarajan, M.; Rajasekar, P. Oviposition attractancy of dodecanoic, hexadecanoic and tetradecanoic acids against Aedes aegypti and Culex quinquefasciatus (Diptera: Culicidae). Eur. Rev. Med. Pharmacol. Sci. 2011, 15, 1172–1175. [Google Scholar]
  138. Macedo, A.L.; Duprat, R.C.; Moreira, D.d.L.; Kaplan, M.A.C.; Vasconcelos, T.R.A.; Pinto, L.C.; Montenegro, R.C.; Ratcliffe, N.A.; Mello, C.B.; Valverde, A.L. Isolation of a larvicidal compound from Piper solmsianum C.DC. (Piperaceae). Nat. Prod. Res. 2017, 32, 2701–2704. [Google Scholar] [CrossRef]
  139. Leite, A.C.C.F.; Kato, M.J.; Soares, R.O.A.; Guimarães, A.E.; Santos-Mallet, J.R.; Cabral, M.M.O. Grandisin caused morphological changes larval and toxicity on Aedes aegypti. Braz. J. Pharmacog. 2012, 22, 517–521. [Google Scholar] [CrossRef] [Green Version]
  140. Maleck, M.; Dos Santos, F.C.; Serdeiro, M.; Ferreira, B.; Gunaydin, K.; De Almeida, A. Khellin: A furanochromone with toxicity against Oncopeltus fasciatus (Hemiptera) and Aedes aegypti (Diptera). J. Nat. Pharm. 2013, 4, 32–36. [Google Scholar] [CrossRef] [Green Version]
  141. Meepagala, K.M.; Estep, A.S.; Becnel, J.J. Mosquitocidal activity of extracts from Ammi visnaga (Apiaceae) seeds. J. Agric. Chem. Environ. 2016, 5, 170–178. [Google Scholar] [CrossRef] [Green Version]
  142. von Oppen, S.; Masuh, H.; Licastro, S.; Zerba, E.; Gonzalez-Audino, P. A Floral-derived attractant for Aedes aegypti mosquitoes. Entomol. Exp. Appl. 2015, 155, 184–192. [Google Scholar]
  143. Braga, I.A.; Valle, D. Aedes aegypti: Inseticidas, mecanismos de ação e resistência. Epidemiol. Serv. Saúde 2007, 16, 279–293. [Google Scholar] [CrossRef] [Green Version]
  144. 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]
  145. Rattan, R.S. Mechanism of action of insecticidal secondary metabolites of plant origin. Crop Prot. 2010, 29, 913–920. [Google Scholar] [CrossRef]
  146. Khanikor, B.; Parida, P.; Yadav, R.N.S.; Bora, D. Comparative mode of action of some terpene compounds against octopamine receptor and acetylcholinesterase of mosquito and human system by the help of homology modeling and docking studies. J. Appl. Pharm. Sci. 2013, 3, 6–12. [Google Scholar]
  147. Pontual, E.V.; Napoleão, T.H.; Dias de Assis, C.R.; de Souza Bezerra, R.; Xavier, H.S.; Navarro, D.M.d.A.F.; Coelho, L.C.B.B.; Paiva, P.M.G. Effect of Moringa oleifera flower extract on larval trypsin and acethylcholinesterase activities in Aedes aegypti. Arch. Insect Biochem. Physiol. 2012, 79, 135–152. [Google Scholar] [CrossRef]
  148. Perumalsamy, H.; Kim, J.R.; Oh, S.M.; Jung, J.W.; Ahn, Y.J.; Kwon, H.W. Novel histopathological and molecular effects of natural compound pellitorine on larval midgut epithelium and anal gills of Aedes aegypti. PLoS ONE 2013, 8, e80226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Pohlit, A.M.; Lopes, N.P.; Gama, R.A.; Tadei, W.P.; Neto, V.F. Patent literature on mosquito repellent inventions which contain plant essential oils—A review. Planta Med. 2011, 77, 598–617. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Oliveira, J.L.; Campos, E.V.R.; Fraceto, L.F. Recent developments and challenges for nanoscale formulation of botanical pesticides for use in sustainable agriculture. J. Agric. Food Chem. 2018, 66, 8898–8913. [Google Scholar] [CrossRef] [PubMed]
  151. Valdiani, A.; Hansen, O.K.; Nielsen, U.B.; Johannsen, V.K.; Shariat, M.; Georgiev, M.I.; Omidvar, V.; Ebrahimi, M.; Dinanai, E.T.; Abiri, R. Bioreactor-based advances in plant tissue and cell culture: Challenges and prospects. Crit. Ver. Biotechnol. 2018, 39, 20–34. [Google Scholar] [CrossRef]
  152. Liu, Y.; Chen, P.; Zhou, M.; Wang, T.; Fang, S.; Shang, X.; Fu, X. Geographic variation in the chemical composition and antioxidant properties of phenolic compounds from Cyclocarya paliurus (Batal) Iljinskaja leaves. Molecules 2018, 23, 2440. [Google Scholar] [CrossRef] [Green Version]
  153. Zouari, N.; Ayadi, I.; Fakhfakh, N.; Rebai, A.; Zouari, S. Variation of chemical composition of essential oils in wild populations of Thymus algeriensis Boiss. et Reut., a North African endemic species. Lipids Health Dis. 2012, 11, 28. [Google Scholar] [CrossRef] [Green Version]
  154. Al-Hamwi, M.; Bakkour, Y.; Abou-Ela, M.; El-Lakany, A.; Tabcheh, M.; El-Omar, F. Chemical composition and seasonal variation of the essential oil of Micromeria fruticosa. J. Nat. Prod. 2011, 4, 147–150. [Google Scholar]
  155. Gobbo-Neto, L.; Lopes, N.P. Plantas medicinais: Fatores de influência no conteúdo de metabólitos secundários. Quim. Nova 2007, 30, 374–381. [Google Scholar] [CrossRef]
  156. Workman, M.J.; Gomes, B.; Weng, J.-L.; Ista, L.K.; Jesus, C.P.; David, M.R.; Ramalho-Ortigao, M.; Genta, F.A.; Matthews, S.K.; Durvasula, R.; et al. Yeast-encapsulated essential oils: A new perspective as an environmentally friendly larvicide. Parasite Vector 2020, 13, 1–19. [Google Scholar] [CrossRef]
  157. World Health Organization. Guidelines for Laboratory and Field Testing of Mosquito Larvicides. 2005. Available online: http://apps.who.int/iris/bitstream/handle/10665/69101/WHO_CDS_WHOPES_GCDPP_2005.13.pdf?sequence=1 (accessed on 21 February 2020).
  158. Isman, M.B.; Grieneisen, M.L. Botanical insecticide research: Many publications, limited useful data. Trends Plant Sci. 2014, 19, 140–145. [Google Scholar] [CrossRef]
  159. Sánchez-Bayo, F. Insecticides mode of action in relation to their toxicity to non-target organisms. J. Environ. Anal. Toxicol. 2011, S4, S4-002. [Google Scholar] [CrossRef] [Green Version]
  160. Borrero-Landazabal, M.A.; Duque, J.E.; Mendez-Sanchez, S.C. Model to design insecticides against Aedes aegypti using in silico and in vivo analysis of different pharmacological targets. Comp. Biochem. Phys. C 2020, 229, 108664. [Google Scholar] [CrossRef] [PubMed]
  161. Rueda, A.G.; Otero, A.L.C.; Duque, J.E.; Kouznetsov, V.V. Synthesis of new α-amino nitriles with insecticidal action on Aedes aegypti (Diptera: Culicidae). Rev. Bras. Entomol. 2018, 62, 112–118. [Google Scholar] [CrossRef]
  162. Devillers, J.; Lagneau, C.; Lattes, A.; Garrigues, J.C.; Clémenté, M.M.; Yébakima, A. In silico models for predicting vector control chemicals targeting Aedes aegypti. SAR QSAR Environ. Res. 2014, 25, 805–835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Dai, Y.-J.; Jia, Y.-F.; Chen, N.; Bian, W.-P.; Li, Q.-K.; Ma, Y.-B.; Chen, Y.-L.; Pei, D.-S. Zebrafish as a model system to study toxicology. Environ. Toxicol. Chem. 2014, 33, 11–17. [Google Scholar] [CrossRef] [PubMed]
  164. Organisation for Economic Co-operation and Development. OECD Guidelines for the Testing of Chemicals. 2013. Available online: https://0-www-oecd--ilibrary-org.brum.beds.ac.uk/docserver/9789264203709-en.pdf?expires=1591301187&id=id&accname=guest&checksum=30910D5C63FC44FE9C43DDAB70F56739 (accessed on 23 March 2020).
  165. Maia, J.D.; Corte, R.L..; 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]
  166. Benelli, G.; Mehlhorn, H. Declining malaria, rising of dengue and Zika virus: Insights for mosquito vector control. Parasitol. Res. 2016, 115, 1747–1754. [Google Scholar] [CrossRef]
  167. Ferreira, T.P.; Haddi, K.; Corrêa, R.F.T.; Zapata, V.L.B.; Piau, T.B.; Souza, L.F.N.; Santos, S.-M.G.; Oliveira, E.E.; Jumbo, L.O.V.; Ribeiro, B.M.; et al. Prolonged mosquitocidal activity of Siparuna guianensis essential oil encapsulated in chitosan nanoparticles. PLoS Negl. Trop. Dis. 2019, 13, e0007624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Aedes aegypti life cycle and the main arboviruses transmited by the female mosquito.
Figure 1. Aedes aegypti life cycle and the main arboviruses transmited by the female mosquito.
Molecules 25 03484 g001
Figure 2. Essential oils to combat the Ae. aegypti mosquito: an ecologically safe alternative.
Figure 2. Essential oils to combat the Ae. aegypti mosquito: an ecologically safe alternative.
Molecules 25 03484 g002
Figure 3. Secondary metabolites identified in essential oils with insecticidal activity against the Ae. aegypti. (A) Alkanes (B) Terpenes and (C) Phenylpropanoids.
Figure 3. Secondary metabolites identified in essential oils with insecticidal activity against the Ae. aegypti. (A) Alkanes (B) Terpenes and (C) Phenylpropanoids.
Molecules 25 03484 g003
Figure 4. Secondary metabolites identified in organic extracts with insecticidal activity against Ae. aegypti. (A) Fatty acid and derivatives, (B) Diterpenes and triterpenes and (C) Others.
Figure 4. Secondary metabolites identified in organic extracts with insecticidal activity against Ae. aegypti. (A) Fatty acid and derivatives, (B) Diterpenes and triterpenes and (C) Others.
Molecules 25 03484 g004
Figure 5. Diterpenes (A) and triterpenes (B) active against Ae. aegypti.
Figure 5. Diterpenes (A) and triterpenes (B) active against Ae. aegypti.
Molecules 25 03484 g005
Figure 6. Sesquiterpenes active against Ae. aegypti.
Figure 6. Sesquiterpenes active against Ae. aegypti.
Molecules 25 03484 g006
Figure 7. Monoterpenes active against Ae. aegypti.
Figure 7. Monoterpenes active against Ae. aegypti.
Molecules 25 03484 g007
Figure 8. Phenylpropanoids and phenolic derivatives active against Ae. aegypti.
Figure 8. Phenylpropanoids and phenolic derivatives active against Ae. aegypti.
Molecules 25 03484 g008
Figure 9. Alkaloids and amides active against Ae. aegypti.
Figure 9. Alkaloids and amides active against Ae. aegypti.
Molecules 25 03484 g009
Figure 10. Fatty acids and thiophene derivatives active against Ae. aegypti.
Figure 10. Fatty acids and thiophene derivatives active against Ae. aegypti.
Molecules 25 03484 g010
Figure 11. Flavonoids active against Ae. aegypti.
Figure 11. Flavonoids active against Ae. aegypti.
Molecules 25 03484 g011
Figure 12. Neolignans active against Ae. aegypti.
Figure 12. Neolignans active against Ae. aegypti.
Molecules 25 03484 g012
Figure 13. Furanochromones and furanocoumarins active against Ae. aegypti.
Figure 13. Furanochromones and furanocoumarins active against Ae. aegypti.
Molecules 25 03484 g013
Figure 14. Naphthoquinone, protolimonoid, alcohol and ketone with activity against Ae. aegypti.
Figure 14. Naphthoquinone, protolimonoid, alcohol and ketone with activity against Ae. aegypti.
Molecules 25 03484 g014
Table 3. Larvicidal activity of organic/aqueous extracts against the Ae. aegypti mosquito.
Table 3. Larvicidal activity of organic/aqueous extracts against the Ae. aegypti mosquito.
Plant SpeciesFamilyCountryPart UsedExtraction SolventLarval StageMortalityTime (h)Reference
LC50 (ppm)LC90 (ppm)
Acacia nilótica (L.) DelileFabaceaeIndiaSeed podn-HexaneL4169.25201.6224[58]
BenzeneL445.3299.3224[58]
ChloroformL4158.13198.2424[58]
Ethyl acetateL459.1275.8224[58]
AcetoneL4103.68162.0324[58]
Acalypha alnifolia Klein ex Willd.EuphorbiaceaeIndiaLeavesn-HexaneL4202.15476.5724[80]
ChloroformL4182.58460.8324[80]
Ethyl acetateL4160.35440.7824[80]
AcetoneL4146.07415.3824[80]
MethanolL4128.55381.6724[80]
Aristolochia bracteata Retz.AristolochiaceaeIndiaLeavesMethanolL3114.89216.2424[81]
Artemisia herba-alba AssoAsteraceaeSaudi ArabiaLeavesWaterL4—India117.18227.6324[82]
WaterL4—Saudi Arabia614.521273.3324[82]
Water + AgNPL4—India10.7021.2424[82]
Water + AgNPL4—Saudi Arabia33.5857.024[82]
Boenninghausenia albiflora (Hook.) Rchb. ex Meisn.RutaceaeIndiaLeavesPetroleum etherL4125.0190.0ND[83]
Buddleja polystachya Fresen.BuddlejaceaeSaudi ArabiaFlowersn-ButanolL1NDNDND[84]
Caesalpinia pulcherrima (L.) Sw.FabaceaeIndiaLeavesEthyl acetateL3144.67276.9924[85]
BenzeneL3136.36272.1524[85]
Cassia fistula L.FabaceaeIndiaLeavesMethanolL310.6920.4724[86]
BenzeneL318.2735.6724[86]
AcetoneL323.9547.1324[86]
Catharanthus roseus (L.) G. DonApocynaceaeIndiaLeavesPetroleum etherL4145.0255.024[83]
Cinnamosma fragrans Baill.CanellaceaeMadagascarRoot BarkMethanolL152.5ND24[87]
Citrullus colocynthis (L.) Schrad.CucurbitaceaeIndiaLeavesPetroleum etherL474.57538.3024[88]
Cunninghamia konishii HayataTaxodiaceaeTaiwanWoodEthanolL4240.0>400.024[73]
LeavesEthanolL4>400.0>400.024[73]
Dalbergia brasiliensis VogelFabaceaeBrazilLeavesEthanolL330.091.024[89]
n-Hexane FractionL344.081.024[89]
Chloroform FractionL333.075.024[89]
Ethyl acetate FractionL324.066.024[89]
BarkEthanolL332.071.024[89]
n-Hexane FractionL331.072.024[89]
Chloroform FractionL325.050.024[89]
Ethyl acetate FractionL328.093.024[89]
Echinops transiliensis Golosk.AsteraceaeKazakhstanRootDichlorometaneND3.216.8124[90]
Eclipta alba (L.) HasskAsteraceaeIndiaLeavesBenzeneL3151.38274.3424[91]
n-HexaneL3165.10297.7024[91]
Ethyl acetateL3154.88288.6124[91]
MethanolL3127.64245.7324[91]
ChloroformL3146.28274.4224[91]
Ervatamia coronaria (Jacq.) Stapf.ApocynaceaeIndiaLeavesEthyl acetateL397.53179.3724[85]
BenzeneL389.59166.0424[85]
Eupatorium odoratum L.AsteraceaeIndiaLeavesPetroleum etherL4155.0290.0ND[83]
Euphorbia hirta L.EuphorbiaceaeIndiaLeavesPetroleum etherL4272.36703.7624[92]
Euphorbia tirucalli L.EuphorbiaceaeIndiaStem barkPetroleum etherL44.2513.1424[92]
Ficus benghalensis L.MoraceaeIndiaLeavesMethanolL256.54109.2924[93]
MethanolL370.29137.2324[93]
MethanolL480.85169.5824[93]
BenzeneL2108.95227.1324[93]
BenzeneL3116.09235.5424[93]
BenzeneL4159.15430.9124[93]
AcetoneL2189.11444.4224[93]
AcetoneL3244.41573.5224[93]
AcetoneL4288.10668.7124[93]
Gardenia ternifolia Schumach. & Thonn.RubiaceaeKenyaLeavesAcetoneL283.31ND24[94]
MethanolL232.01ND24[94]
Helicteres velutina K. Schum.MalvaceaeBrazilRootEthanol 90%L4171.68403.6148[95]
StemEthanol 90%L4138.90319.3748[95]
Heracleum rigens Wall.ApiaceaeIndiaSeedPetroleum etherL240.6465.4924[96]
ChloroformL269.22132.9524[96]
Ethyl acetateL270.6593.1124[96]
MethanolL274.70135.0724[96]
AcetoneL297.07198.9924[96]
Petroleum etherL391.55162.0924[96]
ChloroformL3114.25179.9924[96]
Ethyl acetateL3143.48207.4524[96]
MethanolL3195.57348.5624[96]
AcetoneL3234.77ND24[96]
Petroleum etherL4113.69171.1224[96]
ChloroformL4144.64209.1624[96]
Ethyl acetateL4165.43230.2124[96]
MethanolL4231.26361.6824[96]
AcetoneL4308.65577.1424[96]
Hypericum japonicum Thunb.HypericaceaeIndiaWhole plantAcetoneL413.1518.5424[97]
n-HexaneL49.6319.5324[97]
Petroleum etherL48.2715.224[97]
MethanolL47.3711.5924[97]
Jatropha curcas L.EuphorbiaceaeIndiaLeavesPetroleum etherL48.7935.3924[92]
IndonesiaRootEthanolND44.75ND24[98]
Limonia acidíssima L.RutaceaeIndiaLeavesn-Hexane—Purified fractionsL34.11 to 23.53ND24[99]
Lonchocarpus urucu Killip & A.C. Sm.FabaceaeBrazilMedulla RootMethanolL433.3283.6924[100]
Bark RootMethanolL417.6055.4024[100]
Maytenus oblongata ReissekCelasteraceaeFrench GuianaBarkEthyl acetateL3/L474.40ND24[101]
Millettia pachycarpa Benth.FabaceaeIndiaRootEthanolL398.47ND24[102]
Mirabilis jalapa L.NyctaginaceaeIndiaLeavesBenzeneL397.03172.1524[103]
ChloroformL388.20162.1624[103]
Ethyl acetateL372.77127.9124[103]
MethanolL364.58120.2824[103]
Momordica charantia L.CucurbitaceaeIndiaLeavesMethanolL4199.14780.1024[88]
Myristica fragans Houtt.MyristicaceaeIndiaLeavesMethanolL1162.03502.0424[53]
MethanolL2194.11542.5624[53]
MethanolL3240.10604.7824[53]
MethanolL4273.90660.9624[53]
Methanol + ZnO NPL13.4418.3524[53]
Methanol + ZnO NPL25.2530.3724[53]
Methanol + ZnO NPL38.0239.1424[53]
Methanol + ZnO NPL410.2844.0724[53]
Myristica fragrans Houtt.MyristicaceaeThailandFlowersEthanolL4 (p-s)75.45123.6024[60]
Nerine sarniensis (L.) Herb.AmaryllidaceaeSouth AfricaBulbEthyl acetateL18.0ND24[104]
Nyctanthes arbor-tristis L.OleaceaeIndiaLeavesPetroleum etherL4180.0340.0 [83]
Ocimum sanctum L.LabiataeIndiaLeaves and FlowersAcetoneL4425.94ND24[105]
ChloroformL4150.40ND24[105]
Ethyl acetateL4350.78ND24[105]
n-HexaneL4575.26ND24[105]
MethanolL4175.67ND24[105]
Ormosia arborea VellFabaceaeBrazilLeavesEthanolL3238.0347.024[106]
SeedsEthanolL3111.0194.024[106]
Orthosiphon thymiflorus (Roth) SleesenLabiataeIndiaLeavesn-HexaneL3228.13526.1224[107]
ChloroformL3209.72502.8424[107]
Ethyl acetateL3183.35463.3524[107]
AcetoneL3163.55442.3224[107]
MethanolL3149.96426.1624[107]
Pedilanthus tithymaloides (L.) Poit.EuphorbiaceaeIndiaLeavesPetroleum etherL455.26256.7724[92]
Pemphis acidula J.R. Forst. & G. Forst.LythraceaeIndiaLeavesMethanolL322.1043.7124[108]
BenzeneL343.9984.8724[108]
AcetoneL357.66106.5124[108]
Phyllanthus amarus Schumach. & Thonn.EuphorbiaceaeIndiaLeavesPetroleum etherL490.92384.1924[92]
Piper aduncum L.PiperaceaeBrazilLeavesn-HexaneL3342.0473.024[106]
ChloroformL3192.0346.024[106]
Piper hispidum Sw.PiperaceaeBrazilLeavesEthanolL3169.0474.024[106]
ChloroformL3567.01003.024[106]
Piper longum L.PiperaceaeThailandFruitsEthanolL42.23ND24[109]
Piper nigrum L.PiperaceaePhilippinesPeppercornsEthanolL3/L471.259.3724[110]
Ethanol—Fraction 1AL3/L417.103.8424[110]
Ethanol—Fraction 1BL3/L418.103.8424[110]
Piper ribesoides Wall.PiperaceaeThailandWoodEthanolL48.13ND24[109]
Piper sarmentosum Roxb. ex Hunt.PiperaceaeThailandWhole plantEthanolL44.06ND24[109]
Pithecellobium Dulce (Roxb.) Benth.FabaceaeIndiaLeavesMethanolL3155.78279.7324[111]
Ethyl acetateL3162.36283.4324[111]
ChloroformL3169.08293.1724[111]
BenzeneL3176.02308.8824[111]
n-HexaneL3185.14316.4624[111]
SeedsMethanolL3193.66377.3924[111]
Ethyl acetateL3215.63416.5124[111]
ChloroformL3240.39461.2824[111]
BenzeneL3259.42489.4124[111]
n-HexaneL3281.18516.3324[111]
Scoparia dulcis L.PlantaginaceaeBrazilLeavesEthanol 90%L483.43158.8348[112]
Solanum nigrum L.SolanaceaeIndiaFruitWaterL3/L4359.0931.024[113]
n-HexaneL3/L417.6365.2224[113]
Solanum variabile Mart.SolanaceaeBrazilLeavesEthanolL3188.0284.024[106]
Solanum xanthocarpum Schrad. & J.C. Wendl.SolanaceaeIndiaFruitMethanolL1170.91320.6224[114]
MethanolL2195.07366.4824[114]
MethanolL3221.45410.2024[114]
MethanolL4253.18435.1624[114]
Spermacoce latifólia Aubl.RubiaceaeBrazilLeavesn-HexaneL3415.0901.024[106]
MethanolL3625.01122.024[106]
Tagetes patula L.AsteraceaeBrazilSeedsAcetoneL415.74ND48[115]
Ethanol 50%L425.46ND48[115]
Turnera ulmifolia L.TurneraceaBrazilLeavesEthanolL3242.0899.024[106]
Valeriana hardwickii Wall.ValerianaceaeIndiaLeavesPetroleum etherL4235.0415.0ND[83]
Ventilago madraspatana Gaertn.RhammnaceaeIndiaLeavesWater + AgNPL326.92ND24[112]
WaterL3267.27ND24[112]
Zeuxine gracilis (Berda) Bl.OrchidaceaeIndiaLeavesWater + AgNPL310.3923.5824[116]
LC50 lethal concentration required to kill 50% of the larval population, LC90 lethal concentration required to kill 90% of the larval population, ND not described, p-s pyrethroid-susceptible, AgNP silver nanoparticle, ZnONP zinc oxide nanoparticle.
Table 4. Adulticidal, pupicidal, ovicidal, repellent and oviposition activities of organic/aqueous extracts against the Ae. aegypti mosquito.
Table 4. Adulticidal, pupicidal, ovicidal, repellent and oviposition activities of organic/aqueous extracts against the Ae. aegypti mosquito.
Plant SpeciesFamilyCountryPart UsedExtraction SolventActivityResultsTime (h)Reference
Alpinia purpurata (Viell.) K. Schum.ZingiberaceaeBrazilRed FlowersWaterOvipositionOviposition disruptive effect24[67]
Pink FlowersWaterOvipositionOviposition disruptive effect24[67]
Aristolochia bracteata Retz.AristolochiaceaeIndiaLeavesMethanolOvicideZero hatchability at 240 ppm48[81]
MethanolRepellent100% of repellency at 6 mg/cm23[81]
Artemisia herba-alba AssoAsteraceaeSaudi ArabiaLeavesWaterAdulticide—strain from IndiaLC50 327.15 µg/mL
LC90 = 779.98 µg/mL
24[82]
WaterAdulticide—strain from Saudi ArabiaLC50 450.21 µg/mL
LC90 1153.18 µg/mL
24[82]
Water + AgNPAdulticide—strain from IndiaLC50 8.71 µg/mL
LC90 39.88 µg/mL
24[82]
Water + AgNPAdulticide—strain from Saudi ArabiaLC50 25.62 µg/mL
LC90 48.88 µg/mL
24[82]
Buddleja polystachya Fresen.BuddlejaceaeSaudi ArabiaFlowersn-HexaneAdulticide96.7% mortality at 5 µg/mg femaleND[84]
EthanolAdulticide83.3% mortality at 5 µg/mg femaleND[84]
Aerial partsn-HexaneAdulticide100% mortality at 5 µg/mg femaleND[84]
EthanolAdulticide90% mortality at 5 µg/mg femaleND[84]
Caesalpinia pulcherrima (L.) Sw.FabaceaeIndiaLeavesMethanolOvicideZero hatchability at 300 ppm48[85,117]
MethanolRepellent100% of repellency at 5 mg/cm23[85,114]
Ethyl acetateOvicideZero hatchability at 450 ppm48[85,117]
Ethyl acetateRepellent100% of repellency at 5 mg/cm21.5[85,117]
BenzeneOvicideZero hatchability at 375 ppm48[85,117]
BenzeneRepellent100% of repellency at 5 mg/cm22[85,117]
Cardiospermum halicacabum L.SapindaceaeIndiaLeavesMethanolRepellent100% of repellency at 5 mg/cm23[118]
n-HexaneRepellent100% of repellency at 5 mg/cm23[118]
Ethyl acetateRepellent100% of repellency at 5 mg/cm23[118]
ChloroformRepellent100% of repellency at 5 mg/cm23[118]
BenzeneRepellent100% of repellency at 5 mg/cm23[118]
Cassia fistula L.FabaceaeIndiaLeavesMethanolOvicideZero hatchability at 120 ppm48[86]
MethanolRepellent100% of repellency at 5 mg/cm26[86]
BenzeneOvicideZero hatchability at 140 ppm48[86]
BenzeneRepellent100% of repellency at 5 mg/cm25[86]
AcetoneOvicideZero hatchability at 160 ppm48[86]
AcetoneRepellent100% of repellency at 5 mg/cm24.3[86]
Cinnamosma fragrans Baill.CanellaceaeMadagascarRoot barksMethanolAdulticideLC50 0.17 µg/mg female24[87]
MethanolRepellent80% of repellency at 20.8 µg/cm23[87]
Coccinia indica Wight & Arn.CucurbitaceaeIndiaLeavesBenzeneOvicideZero hatchability at 250 ppm48[119]
BenzeneRepellent100% of repellency at 5 mg/cm23[119]
n-HexaneOvicideZero hatchability at 300 ppm48[119]
n-HexaneRepellent100% of repellency at 1 mg/cm22.5[119]
Ethyl acetateOvicideZero hatchability at 250 ppm48[119]
Ethyl acetateRepellent100% of repellency at 2.5 mg/cm22.5[119]
MethanolOvicideZero hatchability at 200 ppm48[119]
MethanolRepellent100% of repellency at 5 mg/cm23.5[119]
ChloroformOvicideZero hatchability at 250 ppm48[119]
ChloroformRepellent100% of repellency at 2.5 mg/cm22.5[119]
Eclipta alba (L.) HasskAsteraceaeIndiaLeavesBenzeneOvicideZero hatchability at 350 ppm48[119]
n-HexaneOvicide21% hatchability at 350 ppm48[119]
Ethyl acetateOvicideZero hatchability at 350 ppm48[119]
MethanolOvicideZero hatchability at 300 ppm48[91]
ChloroformOvicideZero hatchability at 350 ppm48[119]
Ervatamia coronaria (Jacq.) Stapf.ApocynaceaeIndiaLeavesMethanolOvicideZero hatchability at 200 ppm48[117]
MethanolRepellent100% of repellency at 5 mg/cm23[117]
Ethyl acetateOvicideZero hatchability at 300 ppm48[117]
Ethyl acetateRepellent100% of repellency at 5 mg/cm22[117]
BenzeneOvicideZero hatchability at 250 ppm48[117]
BenzeneRepellent100% of repellency at 5 mg/cm22.5[117]
Limonia acidíssima L.RutaceaeIndiaLeavesn-Hexane—Purified fractionsOvicide78.4 hatchability at 10 ppm120[99]
n-Hexane—Purified fractionsPupicideLC50 4.19—39.48 µg/mL24[99]
Mentha piperita L.LamiaceaeIndiaWhole plantMethanolRepellentRepellencyND[120]
Millettia pachycarpa Benth.FabaceaeIndiaRootEthanolOvicideZero hatchability at 200 ppm24[102]
Myristica fragans Houtt.MyristicaceaeIndiaLeavesMethanolPupicideLC50 359.08 µg/mL
LC90 803.52 µg/mL
24[53]
Methanol + ZnONPPupicideLC50 14.63 µg/mL
LC90 51.22 µg/mL
24[53]
MethanolAdulticideLC50 180.26 µg/mL
LC90 368.93 µg/mL
24[53]
Methanol + ZnONPAdulticideLC50 15.0 µg/mL
LC90 34.2 µg/mL
24[53]
Nerine sarniensis (L.) Herb.AmaryllidaceaeSouth AfricaBulbsEhtyl acetateAdulticideLC50 4.6 µg/mg female24[104]
Parthenium hysterophorusAsteraceaeIndiaLeavesEtherRepellent99.6% of repellency at 1000 ppm48[121]
EtherOvicideZero hatchability at 1000 ppm48[121]
BenzeneRepellent93.8% of repellency at 1000 ppm48[121]
BenzeneOvicideZero hatchability at 1000 ppm48[121]
Pemphis acidula J.R. Forst. & G. Forst.LythraceaeIndiaLeavesMethanolOvicideZero hatchability at 450 ppm48[108]
AcetoneOvicideZero hatchability at 500 ppm48[108]
Pithecellobium Dulce (Roxb.) Benth.FabaceaeIndiaLeavesMethanolOvicideZero hatchability at 400 ppm48[111]
Ehtyl acetateOvicideZero hatchability at 500 ppm48[111]
ChloroformOvicideZero hatchability at 500 ppm48[111]
BenzeneOvicideZero hatchability at 600 ppm48[111]
n-HexaneOvicideZero hatchability at 600 ppm48[111]
SeedsMethanolOvicideZero hatchability at 625 ppm48[111]
Ethyl acetateOvicideZero hatchability at 750 ppm48[111]
ChloroformOvicideZero hatchability at 750 ppm48[111]
BenzeneOvicideZero hatchability at 750 ppm48[111]
Solanum xanthocarpum Schrad. & J.C. Wendl.SolanaceaeIndiaFruitMethanolPupicideLC50 279.52 µg/mL
LC90 462.10 µg/mL
24[114]
Ventilago madraspatana Gaertn.RhammnaceaeIndiaLeavesWater + AgNPOvicideZero hatchability at 120 ppm48[112]
WaterOvicideZero hatchability at 400 ppm48[112]
Water + AgNPAdulticideLC50 44.85 µg/mL24[112]
WaterAdulticideLC50 334.46 µg/mL24[112]
Zeuxine gracilis (Berda) Bl.OrchidaceaeIndiaLeavesWater + AgNPOvicideZero hatchability at 12 ppm48[116]
AdulticideLC50 27.90 µg/mL
LC90 = 59.20 µg/mL
24[116]
LC50 lethal concentration required to to kill 50% of the mosquito population, LC90 lethal concentration required to kill 90% of the mosquito population, ND not described, AgNP silver nanoparticle, ZnONP zinc oxide nanoparticle.

Share and Cite

MDPI and ACS Style

Silvério, M.R.S.; Espindola, L.S.; Lopes, N.P.; Vieira, P.C. Plant Natural Products for the Control of Aedes aegypti: The Main Vector of Important Arboviruses. Molecules 2020, 25, 3484. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25153484

AMA Style

Silvério MRS, Espindola LS, Lopes NP, Vieira PC. Plant Natural Products for the Control of Aedes aegypti: The Main Vector of Important Arboviruses. Molecules. 2020; 25(15):3484. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25153484

Chicago/Turabian Style

Silvério, Maíra Rosato Silveiral, Laila Salmen Espindola, Norberto Peporine Lopes, and Paulo Cézar Vieira. 2020. "Plant Natural Products for the Control of Aedes aegypti: The Main Vector of Important Arboviruses" Molecules 25, no. 15: 3484. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25153484

Article Metrics

Back to TopTop