Next Article in Journal
Protium spruceanum Extract Enhances Mupirocin Activity When Combined with Nanoemulsion-Based Hydrogel: A Multi-Target Strategy for Treating Skin and Soft Tissue Infections
Previous Article in Journal
Targeted Drug Therapy for Senescent Cells Alleviates Unilateral Ureteral Obstruction-Induced Renal Injury in Rats
Previous Article in Special Issue
Epidrugs as Promising Tools to Eliminate Plasmodium falciparum Artemisinin-Resistant and Quiescent Parasites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Tiny Green Army: Fighting Malaria with Plants and Nanotechnology

by
Isabelle Moraes-de-Souza
1,2,
Bianca P. T. de Moraes
1,2,
Adriana R. Silva
2,
Stela R. Ferrarini
3 and
Cassiano F. Gonçalves-de-Albuquerque
1,2,*
1
Immunopharmacology Laboratory, Department of Physiological Sciences, Federal University of the State of Rio de Janeiro—UNIRIO, Rio de Janeiro 20211-010, Brazil
2
Immunopharmacology Laboratory, Oswaldo Cruz Foundation, FIOCRUZ, Rio de Janeiro 21040-361, Brazil
3
Pharmaceutical Nanotechnology Laboratory, Federal University of Mato Grosso of Sinop Campus—UFMT, Cuiabá 78550-728, Brazil
*
Author to whom correspondence should be addressed.
Pharmaceutics 2024, 16(6), 699; https://doi.org/10.3390/pharmaceutics16060699
Submission received: 30 August 2023 / Revised: 27 October 2023 / Accepted: 1 November 2023 / Published: 23 May 2024
(This article belongs to the Special Issue Recent Advances in Drug Therapy for Malaria)

Abstract

:
Malaria poses a global threat to human health, with millions of cases and thousands of deaths each year, mainly affecting developing countries in tropical and subtropical regions. Malaria’s causative agent is Plasmodium species, generally transmitted in the hematophagous act of female Anopheles sp. mosquitoes. The main approaches to fighting malaria are eliminating the parasite through drug treatments and preventing transmission with vector control. However, vector and parasite resistance to current strategies set a challenge. In response to the loss of drug efficacy and the environmental impact of pesticides, the focus shifted to the search for biocompatible products that could be antimalarial. Plant derivatives have a millennial application in traditional medicine, including the treatment of malaria, and show toxic effects towards the parasite and the mosquito, aside from being accessible and affordable. Its disadvantage lies in the type of administration because green chemical compounds rapidly degrade. The nanoformulation of these compounds can improve bioavailability, solubility, and efficacy. Thus, the nanotechnology-based development of plant products represents a relevant tool in the fight against malaria. We aim to review the effects of nanoparticles synthesized with plant extracts on Anopheles and Plasmodium while outlining the nanotechnology green synthesis and current malaria prevention strategies.

1. Introduction

In 2021, there were a total of 247 million cases of malaria, causing 619 thousand deaths, with a mortality rate of 14.8 per 100,000 population at risk. Compared to the previous year, there was an increase of 2 million malaria cases globally [1]. Changes in malaria epidemiology have been attributed to increased risk of transmission, parasite genetic variability, growing drug resistance, and the COVID-19 pandemic [2]. In the first year of the pandemic, malaria combat was compromised because of the reduced distribution of Insecticide-Treated Nets (ITNs) and the disruption of treatment and diagnosis of malaria cases, especially in the World Health Organization (WHO) African Region [1]. Due to the pandemic, the lockdown limited the action of malaria prevention programs, causing a reduction in the distribution of ITNs. Also, health systems were overwhelmed with a shortage of health workers due to a lack of protection equipment, the closure of clinics due to quarantines, and all efforts being directed to control the pandemic, directly impacting the management and prevention of new malaria cases [3]. The WHO reported that in 2021, there was an increase in malaria cases, mainly in the African and Asian continents, where 268 million people required assistance due to humanitarian crises, such as floods, hunger, and political conflicts [1].
Malaria is an infectious disease caused by the Plasmodium sp. parasite and transmitted by female Anopheles sp. mosquitoes during blood feeding. Around forty Anopheles species are known to be human malaria vectors, e.g., A. stephensi, A. gambiae, A. coluzzii, A. funestus [1,4,5]. Anophelinae is from predominantly tropical locations because warmer temperatures are optimal for its survival and development [6]. Gravid females lay eggs in still bodies of water, which hatch into larvae. Larvae turn to pupae after four stages (instars) of growth. Pupae is the resting stage where feeding stops, and the adult mosquito emerges above water after a few days. In the adult stage, female mosquitoes feed from blood to develop eggs [7] and can acquire the malaria parasite if it feeds from an infected host.
Plasmodium has an intricate life cycle, needing two different hosts to complete its development. The parasite enters the bloodstream at human infection and reaches the liver, where it matures. Then, the parasite is released back into the blood, invading erythrocytes and initiating sexual reproduction [8], generating gametes that can be transmitted to mosquitoes.
According to the WHO guidelines for malaria [9], case management involves early diagnosis and prompt treatment, assuring correct dosing, rational medicine use, and drug choice. The WHO established artemisinin-based combination therapy (ACT) as the primary treatment for most malaria cases. The treatment uses artemisinin and its derivatives, such as artemether, dihydroartemisinin, and artesunate, in combination with other drugs [9]. The ACT therapeutic scheme is necessary due to the emergence of artemisinin-resistant parasites [10,11]. Genomic studies enabled the identification and tracking of genetic mutations in the parasite (reference Chaves et al., 2022, for a better review on this topic) [12]. The proposed mechanisms of Plasmodium resistance are the effluxion or blocked entry of the drug through mutations in membrane transporters and the enhancement of metabolic processes to protect from drug-mediated protein damage [13,14]. Resistance to first-line and easily accessible antimalarial drugs, such as chloroquine, quinine, and mefloquine, directly impacts the increase in disease cases in some tropical countries, threatening the local health systems.
The immunity that patients develop after contact with the parasite is not permanent. So, individuals become susceptible again over time. The vast majority of antimalarials used in the treatment of malaria present toxicity, with commonly reported gastrointestinal adverse effects, like nausea and vomiting. Which makes it difficult for the patient to adhere to the treatment, compromising its efficacy [15]. Antimalarial treatment aims to eliminate all forms of plasmodia that produce clinical signs. Different drugs are used for this, depending on the parasite species, the patient’s age, and the disease’s severity.
While vector control strategies focus on using insecticides, resistance to these chemicals has also been observed, posing another threat to malaria transmission control. As an adaptation to ITNs, mosquitoes can change their feeding behavior to ensure survival [16]. Also, the frequent use of insecticides causes genetic modifications that improve the mosquito detoxification system, which allows for lower susceptibility to insecticides-based interventions [17,18,19].
Hence, despite being treatable and preventable, malaria management is complex because of the parasite’s resistance to current drug treatments, mosquitoes’ resistance to insecticides, and the environmental impacts of the latter. In light of these difficulties, there is a relevant need to improve research on novel treatments, drug targets, vector control, and the consequent development of better technology to deliver these strategies.

2. Green-Nanotechnology

Nanotechnology involves manipulating matter in atomic or molecular scales to generate new products and materials, with sizes ranging from 1 to 100 nanometers [20]. Nano systems can be classified as organic, produced from lipids or polymers widely used in the pharmaceutical industry, or inorganic, which are metal-based. Nano systems differ in terms of composition, the method of preparation, and structural organization [21]. The most common methods for synthesizing nanoparticles (NPs) utilize expensive and toxic substances as reducing agents, such as sodium borohydride, hydrazine, formaldehyde, and hydrogen [22]. So, to overcome toxicity, several nanocarriers are being evaluated for developing these particles, such as plant extracts, for a biogenic synthesis approach. Nanotechnology has numerous biomedical applications in wound healing [23], cancer treatment [24,25,26], dentistry [27], and the cosmetic industry [28]. Also, a clinical trial evaluated the effects of thyme and carvacrol NPs on Aspergillus fumigatus isolated from patients in intensive care (found in clicaltrial.gov with the ID NCT04431804).
Because of economic benefits, easy accessibility, and variety, plant extracts are a relevant alternative [29]. NPs synthesized with plant extracts have appropriate size and shape, are consistent with average NPs, and are efficient because of high yield in less synthesis time [30]. Also, using plant extracts requires the bottom-up method of NPs synthesis, characterized by the self-assembly of atoms and molecules from chemical interactions, which is simpler and more efficient [20]. A very commonly used metal in the synthesis of NPs, silver (Ag), has antimicrobial and anti-inflammatory effects [29]. Thus, using plant extracts to synthesize Ag NPs can enhance these activities because polyphenols also show beneficial properties [31]. Moreover, Ag NPs are harmful to aquatic organisms and are considered a pollutant. However, the green synthesis of these NPs can reduce Ag toxicity by increasing molecule stability and reducing the agglomeration of particles [30]. Although these advantages are significant, some drawbacks must be considered, such as possible toxicity, the lack of biocompatibility of the materials used, and the high cost of obtaining nano systems compared to conventional pharmaceutical formulations to treat a neglected disease [12].
The green synthesis of NPs refers to using biological materials as the capping agents to improve the form, functionality, and stability of NPs [32]. The process of green-synthesizing NPs is illustrated in Figure 1.
Capping agents refer to molecules with an amphiphilic character, featuring a polar head and a non-polar tail region. The polar head group forms coordination bonds with the metal atoms within nanocrystals formed during agglomeration, while the tail region interacts with the surrounding medium. The right reducing and capping agents used during the synthesis of NPs are necessary to ensure optimal activity. The bioactive compounds of plant-based products are currently underestimated due to degradation and volatilization in field conditions. One option to avoid these drawbacks is to formulate bioactive plant products using polymers, plasticizers, stabilizers, and biodegradable antioxidants [33]. A plant extract is composed of a variety of metabolites and reductive biomolecules that play a crucial role in reducing metal ions. These components include terpenoids, flavones, ketones, aldehydes, amides, carboxylic acids, carbohydrates, proteins, and vitamins [34]. During the process of green synthesis, bio-compounds form a coating on the surface of NPs [32]. This supplementary layer consequently amplifies the biological attributes of green NPs in contrast to those generated through other chemical reduction methods. The conventional green synthesis reaction involves a straightforward combination of a natural extract and a metal salt solution, serving as the precursor for NPs [35]. This strategy has garnered attention over the past decade, especially for Ag and Au NPs, which are considered safer than other types of metallic NPs [36]. Other metals, like nickel (Ni), manganese (Mn), Titanium (Ti), titanium trichloride (TiO2), palladium (Pd), cerium (Ce), platinum (Pt), and ZnO, have been utilized in the creation of plant-based NPs and green NPs synthesis [35,36].
Biomolecules exhibit the capacity to transfer electrons to metallic ions, functioning as potent reducing agents that initiate the ions’ reduction. Consequently, the reduced ions develop structured patterns that resemble a crystalline formation called a nucleus. The nucleation process begins with the monomer aggregation in supersaturated systems. The formation of an atomic nuclei is followed by the expansion of these nuclei into metal nanoparticles [37]. These nuclei offer a foundation for the reduced ions, facilitating their gradual deposition onto the surface, thereby contributing to the enlargement of the particles. The expansion process is halted by biomolecules that can also function as capping agents, binding to the particle surface and effectively stabilizing their size [37,38].
Besides metallic NPs, the discussion extends to nano emulsions and nanogels. Nanogels are hydrogel particles composed of crosslinked polymers that can swell in an aqueous environment in nanoscale size. Notably, nanogels can accommodate a substantial drug load and achieve efficient encapsulation. Also, they excel at encapsulating larger quantities of substances and sustaining the release of entrapped molecules [39]. On the other hand, nano emulsions are nanostructures formed by the dispersion of two immiscible liquids, such as oil in water or water in oil, that are stabilized by a surfactant. It has a variety of dosage forms, including liquids, creams, sprays, gels, aerosols, and foams. The pesticide industry uses it as an aqueous base for organic deliverables [40]. The characteristics of NPs directly influence their interaction with the environment. Hence, they should be thoroughly investigated to prevent any adverse effects. The small size of NPs and the negative charge on their surface were associated with the potential of Ag NPs to cause hemolysis [41].
The application of nanotechnology for both vector control and patient therapy represents the most promising solution to this challenge. The main objective of malaria treatment is to provide the ideal drug concentration for the intracellular parasite [42]. Nanotechnology protects drugs from chemical degradation, increases the drug’s circulation time in the blood, improves cellular interaction with the parasite, and promotes the maintenance of plasma levels at a constant concentration. Also, it increases therapeutic efficacy, the progressive and controlled release of the drug, decreases toxicity by reducing plasma peaks at maximum concentration, and adds to the possibility of directly affecting specific targets [43]. Nanoparticles can be functionalized with ligands that specifically bind to receptors on the surfaces of malaria parasites. This targeted drug delivery ensures that anti-malarial drugs are delivered directly to the parasites, increasing treatment efficacy while minimizing side effects on healthy cells. Also, nanoparticles can serve as carriers for vaccine antigens, enhancing their stability and immunogenicity [44].

3. Plants and Nano Formulations

In their natural habitat, plants already show defense mechanisms against plagues, predators, and other competing species [33,45]. Thus, to use those against a naturally occurring mosquito is to take full advantage of the biological power displayed by these plants in favor of human health whilst causing minimal environmental impact.
Especially in the malaria scenario, plants that grow in tropical and subtropical regions can facilitate green insecticide production because of easy access, rare toxicity against humans, and low-dose effectiveness [46]. For example, extracts of various seaweed species, especially Caulerpa acemose, are toxic to mosquito larvae, including A. stephensi, at the half-lethal concentration (LC50) of 0.06 µg/mL [47]. Zingiber cernuum essential oil is an effective larvicide and oviposition deterrent against malaria vectors A. stephensi and A. subpictus and does not affect non-target mosquito predators [48]. Heracleum sprengelianum essential oil is also effective against A. subpictus larvae with LC50 33.4 µg/mL [49].
Notably, diverse plant species are commonly used in traditional medicine practices to treat malaria and alleviate Campo’s symptoms [50] because these plants show antiparasitic properties, as seen in [51,52,53,54,55]. Historically, plants have originated the active substances used in malaria drug treatments, such as quinolines derived from Cinchona trees, which originate chloroquine [56]. Also, the current drug of choice to treat malaria, artemisinin, is derived from the plant Artemisia annua L. Its discovery can be traced back to the sixties when it was used as an herbal medicine in China [57].
The challenge in using crude plant extracts is the degradation and evaporation of the chemical compounds upon field conditions [33]. In addition, most isolated plant-derived substances have a short half-life [58], emphasizing the need to develop a better form of administration. Phytochemical nanoformulation (i.e., non-nutrients derived from plants) can improve bioavailability, water solubility, and protection from enzymatic degradation [59]. Botanical-based NPs can be an eco-friendly alternative to current chemical pesticides, which are highly toxic to the environment [60].
Apart from plants, many other biological sources have been used to synthesize NPs. Nano-pesticide synthesized using Metarhizium robertsii showed larvicidal activity against A. stephensi fourth instar larvae in low doses and was not toxic to non-target organisms, such as earthworms and crustacea [61]. Sardine-fish-scale Ag NPs showed potent larvicidal, pupicidal, and ovicidal activity against A. stephensi. Interestingly, the study showed increased predatory activity against larvae from the mosquitofish Gambusia affinis [62]. Ag NPs synthesized by Bacillus marisflavi showed great ovicidal, larvicidal, and pupicidal activity against A. stephensi, the mortality rates of eggs and pupae being the highest compared to other mosquito species [63]. Chitosan Ag NPs showed toxicity against A. stephensi larvae (fourth instar LC50 5.51 ppm) and pupae (LC50 6.54 ppm) [64]. Mycosynthesized Ag NPs using Cochliobolus lunatus were toxic to second–fourth instar larvae of A. stephensi at concentrations of 5 and 10 ppm while presenting no danger to non-target Poecilia reticulata [65]. Earthworms’ coelomic fluids have been extensively used due to their therapeutic properties, such as cytotoxic, proteolytic, and antibacterial. Researchers used the Eudrilus eugeniae earthworms as reducing and stabilizing agents in synthesizing Ag NPs tested against A. stephensi. E. eugeniae Ag NPs were effective against larvae and pupae, and the predation efficiency of the mosquitofish G. affinis was boosted from 68.50% to 89.25% in second instar larvae [66]. G. affinis and P. reticulata, among other species, are larvivorous fish commonly used in the biological control of Anopheles mosquito by malaria prevention programs [67] and demonstrated to be highly effective in a malaria-free country [68].
Moreover, nanoformulated natural products have proven effective against other disease-carrying mosquitoes, such as Aedes sp. [69,70,71,72,73,74,75,76,77], and Culex sp. [78,79,80]. Also, it shows antimicrobial [64,81,82,83], and acaricidal [84] effects.
Additionally, nanotechnology can be applied to other strategies to fight malaria, such as better drug delivery in malaria cases, as reviewed by our group [12], and the nanoformulation of repellents to control disease transmission [85].

4. Malaria Prevention Strategies

Whereas targeting the mosquito vector, the main combat strategies are Indoor Residual Spraying (IRS) and ITNs [9]. In sub-Saharan Africa, ITNs are the primary prevention strategy, and even though there has been an increase in ownership and use of ITNs over the years, population-level access has decreased since 2017. In addition, the levels of people protected by IRS in endemic countries have decreased, reaching 2.4% in 2021 compared to 5.5% in 2010 [1]. Although IRS and ITNs have effectively reduced the malaria burden, these strategies have faced challenges because of household accessibility [86] and vector resistance to insecticides [87]. Pyrethroid-only long-lasting insecticidal nets (LLINs) are the current main WHO recommendation, but mosquito resistance to pyrethroids has compromised its efficacy. Pyrethroids are the synthetic version of pyrethrins derived from the flowers of the Pyrethrum genus, which have insecticidal effects. The development of pyrethroids was incited due to the molecular instability of pyrethrins [88]. Pyrethroids increase the opening of insects’ sodium channels, disrupting nervous impulse propagation and the overactivation of nerve cells [89], causing mosquito paralysis, the same effect observed in nanoformulated plant extracts and further discussed in this review. Other synthetic insecticides are proposed with conditional recommendation because cost-effectiveness is still uncertain and lacks evidence [9]. Pesticides are known to be hazardous to the environment [90]. They are lethal to aquatic organisms [91,92] and potentially affect human development due to prenatal exposure [93] and hormone signaling dysfunction [94].
Currently, the WHO does not recommend insecticide space-spraying due to a lack of evidence about its impact on malaria and the short life of the used chemicals [9]. This review summarizes the effects of numerous eco-friendly alternatives to synthetic pesticides, emphasizing nanotechnology’s use to improve efficacy (Table 1). Senna occidentalis and Ocimum basilicum coils showed significant toxicity against A. stephensi, causing mortality rates of 38% and 52%, respectively, similar to the pyrethrin-based control coil, which caused a 42% mortality rate [95]. Pteridium aquilinum Ag NPs coils synthesized with different plant parts caused mortality rates comparable to pyrethrin-based coils [96]. Also, in both studies, the plant-containing smoke could reduce parasite transmission because it caused a reduction in the percentage of fed mosquitoes. Coils from Ulva lactuca were effective against A. stephensi, causing a mortality rate of 66%, higher than that of the pyrethrin-based coil, which was 41% [97].
In addition, the green nanotechnology strategy can promote a biological form of vector control because most plant extracts can increase the predatory activity of non-target organisms against the early developmental stages of the mosquito. For instance, chitosan-synthesized Ag NPs increased the predation efficacy of Danio rerio against larvae of A. stephensi at a concentration of 1 ppm [64]. Citrus limon gold–palladium (Au-Pd) NPs had no negative impact on the predatory habit of flying insects, which increased over time [98]. Likewise, Lagenaria siceraria zinc oxide (ZnO) NPs increased P. reticulata predation efficacy from 45.8% in the aqueous extract to 61.13% [99]. Also, Cymbopogon citratus Au NPs boosted Mesocyclops aspericornis predation against A. stephensi from 26.8 to 45.6% [100]. The predation efficacy improvement could be due to the decreased mobility of larvae, which are typically fast-paced, after exposure to the green-synthesized NPs [98]. Additionally, Ag NPs synthesized with Ichnocarpus frutescensi [101], Rubus ellipticus [102], Naregamia alata [103], and Hugonia mystax [104] had a negligible toxic effect on non-target aquatic organisms, such as G. affinis, Diplonychus indicus, and Anisops bouvieri. Solanum xanthocarpum-synthesized Ag NPs [105] and Momordica charantia leaf extract TiO2 NPs [106] had no toxicity against P. reticulata at the same concentrations lethal to mosquitos. It is crucial to outline the choice of material used to synthesize NPs because CdS showed toxicity against non-target aquatic organisms [107]. These are predators of mosquito breeding sites, hence the importance of developing insecticides harmful only to the intended species to maintain the ecologic niche balance. Malva sylvestris Ag NPs were not toxic to D. indicus and G. affinis, as the longevity and swimming habits did not change after exposure to NPs, and the suitability index showed that treatment with NPs was more toxic to the mosquito larvae [108].
The studies above follow the recently defined WHO preferred product characteristics (PPCs) for vector-control products. PPCs were determined to incentivize the search and development of new strategies to control Anopheles mosquitoes while addressing multiple parameters, including indoor and outdoor transmission control, complementarity to existing methods, and safety for humans and the environment [87]. Green-synthesized nanotechnology fulfills most of the demanded PCCs, thereby being a suitable candidate to fight malaria.
Table 1. Characteristics of NPs synthesized with plant extracts that have beneficial use and its effects on malaria vector and parasite. The nanoparticle sizes are according to transmission electron microscopy (TEM) or scanning electron microscopy (SEM) analysis.
Table 1. Characteristics of NPs synthesized with plant extracts that have beneficial use and its effects on malaria vector and parasite. The nanoparticle sizes are according to transmission electron microscopy (TEM) or scanning electron microscopy (SEM) analysis.
Plant ExtractPlant UseType of NanotechnologySize of NanoparticlesEffectsReferences
Alchornea cordifolia leafInfections, inflammation, rheumatism, analgesia, and arthritisAg NPs5–25 nm
(TEM)
Larvicidal, impact on larvae behavior and morphology, antiplasmodial and hemolytic[109]
Aristolochia indica leafPost–partum infections, snakebites, fever, malaria, skin diseases, helminths, edema, intestinal disorders, and antibacterialAg NPs30–55 nm
(SEM)
Larvicidal and pupicidal in laboratory and field, predation booster[110]
Azadirachta indica leaf and barkAntimalarial, intestinal disorders, and diabetesAg NPs4–28 nm
(TEM)
Antiplasmodial and hemolytic[41]
Cedrus deodara oilInsecticidalPectin nanocapsules40–80 nm (TEM)Larvicidal and impact on larvae behavior and morphology[111]
Cocos nucifera mesocarpFoodAg NPsmean 23 nm (TEM)Larvicidal [112]
Couroupita guianensis flowerAntibiotic, antifungal, antidepressant, antiseptic, analgesia, stomach ache, and skin diseasesAu NPs29.2–43.8 nm
(TEM)
Larvicidal and pupicidal in laboratory and field, adulticidal, predation booster, and antiplasmodial[113]
Codium tomentosumAntioxidantAg NPs20–40 nm
(SEM)
Larvicidal, pupicidal, antiplasmodial, antioxidant, antibacterial, and predation booster[114]
Eclipta prostrata leafLipidemia, atherosclerosis, hepatic diseases, and snake-venom poisoningPd NPs18–64 nm
(TEM)
Antiplasmodial and cytotoxic[84]
Mimusops elengi leafCardiotonic, stomach ache, anti-helminthic, dysentery, antimicrobial, anti-ulcer, antianxiety, anti-oxidant, antihyperglycemic, anti-hyperlipidemic, anti-inflammatory, and feverAg NPs25–40 nm (TEM)Larvicidal, pupicidal, adulticidal, and predation booster in laboratory and field[115]
Arachis hypogaea peelCattle foodAg NPs20–50 nm (TEM)Larvicidal and impact on larvae morphology[116]
Vitex negundo leafBactericidal, diabetes, inflammation, and asthmaZnO NPs28–42 nm
(TEM)
Larvicidal, pupicidal, antioxidant, cytotoxic, and photocatalytic[117]
Pteridium aquilinum leafAnalgesia and foodAg NPs35–65 nm
(SEM)
Larvicidal and pupicidal in laboratory and field, adulticidal, ovicidal, antiplasmodial, and repellent smoke[96]
Ulva lactucaAntioxidant, antibacterial, and antiviralAg NPs20–35 nm
(SEM)
Larvicidal, pupicidal, antiplasmodial, and repellent smoke[97]
Citrus limon leafNatural pesticide, insect repellent, and antimicrobialAu-Pd NPs1.5–18.5 nm
(TEM)
Larvicidal and predation booster[98]
Lagenaria siceraria peelAnti-swelling, diuretic, antibacterial, and cytotoxicZnO NPs-Larvicidal, impact on larvae behavior and morphology, antiplasmodial, predation booster, and cytotoxic[99]
Cymbopogon citratus leaf-Au NPs20–50 nm
(TEM)
Larvicidal, pupicidal, and predation booster[100]
Ichnocarpus frutescens leafAnti-diabetes, anti-tumor, anti-inflammatory, and analgesiaAg NPs5–47 nm
(TEM)
Larvicidal and biocompatible with non-target organisms[101]
Rubus ellipticus leafDiabetes, diarrhea, gastralgia, wound healing, dysentery, antifertility, antimicrobial, analgesia, and epilepsyAg NPs1–25 nm
(TEM)
Larvicidal, ovicidal, and adulticidal, oviposition deterrent, and biocompatible with non-target organisms[102]
Naregamia alata leafWound healing, ulcers, halitosis, cough, asthma, bronchitis, splenomegaly, scabies, pruritus, dysentery, dyspepsia, catarrh, anemia, and malarial feversAg NPs0–5.5 nm
(TEM)
Larvicidal, ovicidal, adulticidal, and biocompatible with non-target organisms[103]
Hugonia mystax leafAnthelmintic and rheumatismAg NPs40–90 nm
(SEM)
Larvicidal and biocompatible with non-target organisms[104]
Solanum xanthocarpum leafAnti-cancer, antioxidant, anti-HIV, antibacterial, and insecticidalAg NPs10–20 nm
(TEM)
Larvicidal and biocompatible with non-target organisms[105]
Eucalyptus globulus oilNatural pesticideNanoemulsion22–40 nm
(TEM)
Larvicidal in laboratory and semi-field[118]
Mangifera indica leafAntioxidant and antibacterialTiO2 NPs30 nm
(TEM)
Larvicidal and acaricidal[84]
Barleria cristata leafAntioxidant, cytotoxic, and antimicrobialAg NPs38–41 nm
(TEM)
Larvicidal, biocompatible with non-target organisms[119]
Malva sylvestris leafAntioxidant, anti–inflammatory, and antimicrobialAg NPs18–25 nm
(TEM)
Larvicidal and biocompatible with non-target organisms[108]
Ammania baccifera aerialAnalgesia, antifertility, hypothermic, hypertensive, antiurolithiasis, antisteroidogenic,
antimicrobial, antiurolithic, anti–inflammatory, and antioxidant
Ag NPs10–30 nm
(TEM)
Larvicidal[120]
Artemisia nilagirica leafAntimicrobial, anthelmintic, antiseptic, and larvicidalAg NPs<30 nm
(SEM)
Larvicidal and pupicidal[121]
Valoniopsis pachynema-CdS NPs<100 nm
(SEM)
Larvicidal, pupicidal, antiplasmodial, and toxic to non–target organisms[107]
Vitex negundo leafAntimicrobial, anti–inflammatory, diabetes, asthma, cytotoxic, and larvicidalZnO NPs28.48–42.14 nm
(TEM)
Larvicidal and antioxidant[117]
Annona squamosa leafPesticidal, cytotoxic, and antioxidantAg NPs200–500 nm
(SEM)
Larvicidal, pupicidal, ovicidal, and adulticidal[122]
Momordica charantia leafAntidiabetic, antiviral, antitumor, antileukemic, antibacterial, anthelmintic, antimutagenic, antimycobacterial, antioxidant, antiulcer, anti-inflammatory, hypocholesterolemic, hypotriglyceridemic, hypotensive, immunostimulant, and insecticidalZnO NPs21.32 nm
(SEM)
Larvicidal, acaricidal, and pediculicidal[123]
TiO2 NPs70 nm
(TEM)
Larvicidal, pupicidal, antiplasmodial, and biocompatible with non–target organisms[106]
Euphorbia hirta leafNatural insecticideAg NPs30–60 nm
(SEM)
Larvicidal and pupicidal[124]
Nerium oleander leafAnticancer, antimicrobial, anxiolytic, and antipsychoticAg NPs20–35 nm
(SEM)
Larvicidal and pupicidal[125]
Heliotropium indicum leafFever, throat infection, ulcer, gonorrhea, localized
inflammation, rheumatism, ring worm, wounds, aphrodisiac, astringent, and expectorant
Ag NPs18–45 nm
(TEM)
Adulticidal[126]
Zornia diphylla leafDysentery, venereal diseases, and sleep inductionAg NPs30–60 nm
(SEM)
Larvicidal and biocompatible with non-target organisms[127]
Mussaenda glabra leaf-Ag NPs15–25 nm
(TEM)
Larvicidal and biocompatible with non-target organisms[128]
Anisomeles indica leafInflammatory skin diseases, liver protection, intestinal infections,
abdominal pain and immune system deficiencies, expectorant, diaphoretic, rheumatism, and psoriasis
Ag NPs18–35 nm
(TEM)
Larvicidal[129]
Holostemma adakodien leaf-Ag NPs-Larvicidal, antibacterial, and biocompatible with non-target organisms[81]
Quisqualis indica leafAntimicrobial and anticoccidialAg NPs<30 nm
(SEM)
Larvicidal and biocompatible with non-target organisms[130]
Nicandra physalode leafAntidiuretic,
mydriasis, analgesia, antibacterial, anti-inflammatory, and insecticidal
Ag NPs5–35 nm
(SEM)
Larvicidal and biocompatible with non-target organisms[131]
Gmelina asiatica leafHepatic diseasesAg NPs20–64 nm
(TEM)
Larvicidal[132]

5. Plant Extracts and Green NPs Action Mechanisms

The NPs’ action mechanisms are still unclear, and the current knowledge is limited to mosquitoes’ morphological, molecular, biochemical, and physiological levels [133]. Table 1 summarizes the effects of plant extracts in these numerous levels with the nanoparticle’s corresponding type and size. The high mortality rates exerted by NPs on mosquito larvae may be explained by the small size of the particles, which enables the passage through the respiratory tubes and/or insect cuticle and into cells where they interfere with molting and other physiological processes [66,122]. After penetrating the exoskeleton, the Ag NPs exert their effects in the intracellular space, interacting directly with proteins or DNA and causing genotoxic effects. The depletion of antioxidants induces oxidative stress, the oxidative dissolution of Ag NPs, and mitochondria perturbation [134]. It also can induce apoptosis through the activation of procaspase-3, the downregulation of the pro-survival protein Bcl-2, the enhanced expression of pro-apoptotic gene products, and the release of cytochrome c into the cytosol [135,136].
The toxicity of NPs is intricately linked to their propensity to accumulate within various organs. Upon entering the bloodstream, NPs exhibit the capacity to disperse throughout the body, accumulating within organs such as the liver, spleen, lungs, and kidneys. This distribution of NPs within the body is primarily influenced by their surface area-to-size ratio, a factor that affects their tendency to accumulate in distinct tissues and organs [137]. The pathways through which NPs exert toxicity within the organs go through several critical factors, including producing reactive oxygen species (ROS), DNA damage, the alteration of protein structures and functions, and compromising membrane integrity. Notably, NP characteristics that seem to enhance these mechanisms involve their expansive surface areas, which promote molecular interactions within the specific target sites [138]. Toxicity is not limited to the target species but also to the environment, especially in the management of the vector in-field conditions. NPs released in water affect aquatic organisms, consequently resulting in reports of blood acidosis, causing circulatory collapse, bioaccumulation, hepatotoxic effects, oxidative stress, and embryonic development. However, such effects were observed at elevated concentrations [29], in opposition to the low concentrations used in nanoparticles.
The neurotoxic effects of plant-derived compounds are linked to acetylcholinesterase (AChE) inhibition, the modulation of gamma-aminobutyric acid (GABA), and octopamine (OA) receptors (GABAr and OAr, respectively). Essential oils are a complex mixture of green chemicals, mainly terpenoids, that act through these mechanisms, resulting in the overstimulation or paralysis of the insect’s nervous system [139]. Studies using Ag NPs evaluate the expression of glutathione S-transferase (GST) in midges, pointing to the importance of GST genes protecting against oxidative stress induced after exposure to Ag NPs [140].
Exposure to Ag NPs synthesized with Alchornea cordifolia leaf extract triggered behavioral and morphological changes in larvae, with reduced swimming [109]. L. siceraria ZnO NPs caused the loss of body hair, the decomposition of the outer cuticle’s epithelial layer, and the disintegration of abdominal structures and midgut on A. stephensi third instar larvae [99]. In addition, the abnormal restless movement of adult A. stephensi upon exposure to high concentrations of nanoformulated Heliotropium indicum extract was observed [126]. The histopathological analysis of fourth instar larvae of A. culicifacies exposed to Cedrus deodara essential oil pectin nanocapsules revealed severe morphological deformities and damage to muscle, adipose tissue, and epidermal cells. C. deodara essential oil has been reported to possess insecticidal activities attributed to the presence of compounds like terpenes, terpenoids, and sesquiterpenes, such as β-himachalene and α-himachalene [111]. H. sprengelianum essential oil components had different efficacy levels, lavandulyl acetate (LC50: 4.17 μg/mL) being more lethal to A. subpictus larvae than bicyclogermacrene (LC50: 10.3 μg/mL) [49]. Also, α-humulene and β-elemene isolated from Syzygium zeylanicum essential oil were highly toxic to A. subpictus larvae with LC50 6.19 and 10.26 μg/mL, respectively, which is even lower than the complete essential oil, which had LC50 83 μg/mL [141]. Carvacrol and terpinen-4-ol from Origanum vulgare essential oil were more effective than whole oil against A. subpictus, with an LC50 of 24.06 and 47.73 μg/mL, and A. stephensi, with an LC50 of 21.15 and 43.27 μg/mL, respectively [142]. Plectranthus barbatus essential oil’s major components, eugenol, α-pinene, and β-caryophyllene, were more effective against A. subpictus than whole extract, presenting an LC50 of 25.45, 32.09, and 41.66 μg/mL, respectively [143]. The main components from Hedychium larsenii essential oil, ar-curcumene and epi-β-bisabolol, were toxic against A. stephensi larvae (LC50: 10.45 and 14.68 µg/mL) and had oviposition-deterrent activity [144].
In field or semi-field experiments, it is hypothesized that the lethal effect of plant extracts may be due to the formation of an oil layer in the aqueous medium that blocks proper water oxygenation and allows for the active penetration of phytochemicals in larvae [110]. The water quality and compounds also implicate mosquito habits, such as egg-laying, because female gravid mosquitoes are attracted to breeding sites according to olfactory cues, such as pheromones and environmental odors, with species-dependent preferences [145]. A. coluzzii were more attracted to sites with early stages larvae [146] and A. gambiae had a preference to water sites treated with nonane, a pheromone produced by larvae [147]. Plant extracts may prevent oviposition due to their larvicidal effects, thus eliminating a mosquito attraction factor. Conversely, volatile plant compounds from natural habitats can attract mosquitoes and favor egg-laying, such as limonene, β-pinene, β-elemene, and β-caryophyllene detected in preferred A. gambiae breeding sites [148]. In this case, we hypothesize that the oviposition-deterrent activity by nanoformulations is due to the stability of these compounds in the water and the higher capacity for penetrating the insect’s body, causing toxicity.
The disadvantage of inorganic nanoparticles is the metal ion’s toxicity to the environment, highlighting the demand for an eco-friendly and modern intervention. For that reason, NPs synthesized from biomolecules is a relevant tool in fighting malaria whilst causing minimal impact on the environment. To summarize, the effects and possible action mechanisms are depicted in Figure 2, and we hypothesize that the toxicity of green NPs may be caused by the known effects of phytochemicals, which play a part in the functionality of NPs (as seen in Figure 1).

6. Antimalarial Effects of Green NPs

Multiple studies have evaluated the plant extracts potential to fight malaria at its many fronts, parasite elimination, or vector control. Regarding the vector population control, the main parameters assessed were the induction of larvae, pupae, and adult mortality, the impediment of oviposition, and egg hatchability. The main effects of green-synthesized NPs against mosquitoes and parasites are illustrated in Figure 3.
Among the studied vector species are A. stephensi and A. subpictus. A. stephensi is an essential vector of both P. falciparum and P. vivax, and has been documented to spread to other geographic areas, especially urban environments, where it can thrive, posing a challenge to vector control [1]. Thus, the use of natural products in these settings would be beneficial. Using plant extracts to synthesize NPs increases the mortality rate of mosquitoes in all developmental stages, and this effect is obtained using lower concentrations of the extract. For example, Ag NPs synthesized with Euphorbia hirta [124], Nerium oleander [125], U. lactuca [97], Codium tomentosum [114], M. sylvestris [108], Holostemma adakodien [81], H. mystax [104], Quisqualis indica [130], Nicandra physalode [131], Gmelina asiatica [132], Musa paradisiaca [149], Drypetes roxburghii [150], Vinca rosea [151], Cocos nucifera [112], Pergularia daemia [152], L. siceraria ZnO NPs [99], and C. citratus Au NPs [100] had the highest larvicidal and pupicidal effect against A. stephensi, in comparison to the crude plant extract. S. xanthocarpum-synthesized Ag NPs showed significant larvicidal effects up to 72 h after A. stephensi larvae exposure [105]. Also, against larvae of A. subpictus, Ag NPs synthesized with extracts of diverse plant species, such as Zornia diphylla [127], Mussaenda glabra [128], Anisomeles indica [129], Barleria cristata [119], Ammannia baccifera [120], Nelumbo nucifera [153], Tinospora cordifolia [154], and Mimosa pudica [155], and Vitex negundo ZnO NPs [117] were more toxic than the crude leaf extract, presenting lower lethal concentrations. The LC50 of green NPs and the affected mosquito species are summarized in Table 2.
A eucalyptus nanoemulsion had larvicidal activity against A. stephensi in laboratory and semi-field conditions. The larvae population had low density for up to 6 days in breeding sites sprayed with the nanoemulsion. In contrast to the breeding sites treated with regular eucalyptus oil, which had increased larvae after only two days [118]. In laboratory conditions, Aristolochia indica leaf extract-synthesized Ag NPs were toxic against A. stephensi larvae and pupae, with an LC50 ranging from 3.94 to 15.65 ppm, respectively. In field experiments, the crude extract and NPs had similar larvicidal activity, reaching about 50% larvae mortality in 24 and 100% in 72 h [110]. Artemisia dracunculus essential oil nanoemulsion and nano gel were toxic to A. stephensi larvae. The nanogel preparation was more lethal than the nanoemulsion, with an LC50 of 6.6 and 13.5 μg/mL, respectively [82]. Nanoliposomes containing A. dracunculus essential oil showed more significant larvicidal effect against A. stephensi when compared to other Artemisia species, such as A. annua and A. sieberi, even though all three presented toxic effects to larvae. A. dracunculus essential oil nanoliposomes caused 100% mortality at 50, 100 and 200 μg/mL, while A. annua reached this rate only at 200 μg/mL, and A. sieberi caused 77% [156]. On the other hand, Acroptilon repens essential oil nanoemulsion showed low larvicidal activity against A. stephensi because LC50 and LC90 values were 7 and 35 ppm, respectively [157].
C. limon bimetallic AuPd NPs showed larvicidal activity against A. stephensi. In the same study, larval mortality of A. aegypti was also assessed and proven to be lower because a 100% mortality rate was seen for the first instar larvae of A. stephensi but not A. aegypti [98]. Likewise, A. cordifolia Ag NPs were more lethal to A. stephensi larvae than other mosquito species [109]. Conversely, Annona squamosa Ag NPs LC50 values were higher for A. stephensi than A. aegypti and C. quinquefasciatus [122].
Mangifera indica-synthesized TiO2 NPs effectively controlled A. subpictus larvae. Even at the lowest concentration of 5 mg/L, TiO2 NP-exposed larvae had a higher mortality rate compared to TiO(OH) solution (TiO2 NPs: 37%; TiO(OH): 11%). TiO2 NPs reached the maximum mortality rate at 25 mg/L, whereas TiO(OH) solution had 89% mortality at the same concentration [84]. I. frutescens-synthesized Ag NPs showed a dose-dependent toxic effect on larvae of A. subpictus. The LC50 for Ag NPs was 14.22 μg/mL, and for the aqueous extract, it was 185.83 μg/mL. The highest mortality rate, 100%, was only obtained by NPs at a concentration of 35 μg/mL, while the crude extract reached 99.2% mortality at 450 μg/mL [101]. A. nilagirica Ag NPs had significant toxic effects against A. stephensi first–fourth instar larvae and pupae, with LC50 ranging from 0.343 to 0.05% after 24 h of exposure at a 0.25% concentration level. The A. nilagirica Ag NPs reached 100% mortality of late instars and pupae at 0.20% concentration. The plant aqueous extract did not reach these mortality rates even at the highest concentration of 1% [121], in addition to M. charantia leaf extract-synthesized TiO2 NPs [106] and ZnO NPs [123]. TiO2 NPs were toxic to first–fourth instar larvae and pupae of A. stephensi, with an LC50 ranging from 2.5 to 5.04 mg/L. The ZnO NPs had a larvicidal effect against A. subpictus, with an LC50 of 5.42 mg/L. Both M. charantia NPs had higher toxicity than the plant aqueous extract. P. aquilinum Ag NPs showed high toxicity to larvae, pupae, and adult A. stephensi in laboratory and field conditions. The NP treatment reduced adult mosquitoes’ longevity and fecundity. Applying plant-based treatments in water reservoirs reduced the larval population to zero within 72 h of exposure [96]. R. ellipticus-synthesized Ag NPs also displayed ovicidal (LC50: 60 μg/mL) and adulticidal (LD50: 21.10 μg/mL) effects and prevented 89% oviposition at 60 ppm concentration [102].
Also, researchers synthesized essential oil from C. deodara into pectin nanocapsules and applied the solution to cotton-bag fibers in water containing A. culicifacies third instar larvae. Within five days, the larval population was reduced by 90%, reaching a 98% reduction after 28 days [111]. Mimusops elengi Ag NP has also displayed toxicity against mosquito larvae, pupae, and adult mosquitoes at ultra-low doses. Intriguingly, the ultra-low dosages of green-synthesized Ag NP reduced the motility of mosquito larvae, consequently boosting the efficiency of natural mosquito predators, like mosquitofish [115]. Datura metell leaf extract that contains alkaloids and atropine, used for medicinal purposes [158], and Aloe vera extract [159] Ag NPs showed acute toxicity against larvae and pupae of A. stephensi, in addition to enhancing predation to larval populations. In field conditions, A. vera Ag NP led to a larval reduction of 97.7% after 72 h [159]. Annona muricata extract possesses several properties, such as anti-viral, anti-cancer, anti-fungal, and especially larvicidal activity. Ag NPs from A. muricata showed toxicity against A. stephensi third instar larvae with LC50 and LC90 values of 15.28 and 31.91 μg mL (−1) [160].
In addition, Chrysanthemum indicum, a flower rich in pyrethrin, has gained widespread recognition for its insecticidal properties attributed to its impact on membrane integrity, causing damage to both lipid and aqueous components of the gill membrane. Notably, C. indicum-derived Ag NPs exhibited remarkable mortality rates when tested against larvae and pupae of A. stephensi [161]. The peanut (Arachis hypogaea) peel extracts were utilized to synthesize Ag NP and evaluated against fourth instar larvae of A. stephensi. When exposed to a concentration of 15 mg/L, a 100% mortality rate was observed among the larvae within 24 h, accompanied by significant morphological changes in the cuticular membrane [116]. A similar outcome was observed with the Ag NPs derived from the aqueous leaf extracts of Leucas aspera and Hyptis suaveolens, where a mortality rate of 100% was achieved at a concentration of 10 mg/L [162]. Biosynthesized Au NPs of Couroupita guianensis flower extract were used in water storage reservoirs on larval populations of A. stephensi. The Au NP flower extract had larvicidal and pupicidal effects, and the adulticidal activity showed 95% mortality. The predation efficiency of Aplocheilus lineatus after treatment with ultra-low dosages of C. guianensis Au NP was enhanced (96.04%) [113].
A few reports indicated that green-synthesized NPs may also have toxic effects against malaria parasites in cellular and animal models. The most studied Plasmodium strains are the P. falciparum chloroquine-sensitive (CQs) strain 3D7, and chloroquine-resistant (CQr) strains RKL9 and INDO. An early in vivo report assessed the potential of orally administered palladium (Pd) NPs synthesized from Eclipta prostrata in mice infected with the P. berghei CQs strain NK 65. Pd NPs were more effective in reducing parasitemia than the aqueous extract [84]. A. cordifolia Ag NPs were more toxic to 3D7 and RKL9 strains than crude extract, with IC50 values 8.05 and 20.27 μg/mL for 3D7, 10.31 and 32.55 μg/mL for RKL9, respectively [109]. Ag NPs synthesized with Azadirachta indica inhibited parasite growth. The IC50 (µg/mL) values for A. indica bark Ag NPs against 3D7 and RKL9 strains were 8 and 7.8, respectively. A. indica leaf Ag NPs had an IC50 of 9.2 for 3D7 and 11.1 for RKL9. Both NPs were effective at lower concentrations than the aqueous extract [41]. Antiplasmodial effects were also seen for L. siceraria ZnO NPs, with parasitemia suppression at IC50 4.32 µg/mL, comparable to chloroquine with IC50 2.5 µg/mL [99]. Compared to the crude leaf extract, M. charantia-synthesized TiO2 NPs were more effective in the antiplasmodial assay against 3D7 and INDO strains [106]. C. tomentosum Ag NPs inhibited the parasitemia of 3D7 and INDO strains, with an IC50 of 72.45 and 76.08 μg/mL, respectively. The treatment was similar to chloroquine, with an IC50 of 80 (CQs strain) and 85 μg/mL (CQr strain) [114]. Interestingly, Ag NPs and crude extract from U. lactuca [97] and P. aquilinum [96] showed higher toxicity to 3D7 and INDO parasite strains than chloroquine. The C. guianensis extract Au NPs showed a higher growth inhibition than chloroquine against 3D7 and INDO strains [113]. Also, the L. siceraria ZnO NPs exhibited the potent inhibition of hematin formation [99]. E. eugeniae Ag NPs showed higher toxic activity in antiplasmodial assays than chloroquine against P. falciparum [66]. After the infection of red blood cells, P. falciparum digests hemoglobin, generating a free heme group, which is toxic to the parasite. Within its digestive vacuole, Plasmodium converts the heme group into hemozoin, also called β-hematin, a non-toxic crystallized molecule [163]. So, the inhibition of hemozoin formation is crucial in the research for novel antimalarials. Thus, pilling evidence shows the potential use of green NPs to fight malaria.

7. Conclusions

While it is well-documented in the literature, there needs to be more research discussions regarding the relevance of green-synthesized NPs in malaria prevention and its impact on the vector. Many studies have indeed explored its potential in treatment, but there needs to be more exploration of its role in vector prevention. In this paper, our primary objective was to conduct a comprehensive review of previously synthesized nanoparticles that are potentially used in controlling malaria transmission. By addressing this gap in research, we aim to contribute valuable insights into malaria prevention. Nanotechnology is leading with the most recent research for better drug delivery and the development of novel tools. The nanoformulation of compounds enables low concentrations compared to the bulk chemicals, thereby reducing toxicity. Indeed, nanotechnology characteristics and materials are an essential topic in discussion since they can influence the biological impact and behavior of the NPs. There are disadvantages to inorganic NPs, so the biogenic synthesis of NPs has been extensively reported in the literature, especially regarding the use of plant extracts in the synthesis of inorganic NPs with potent activity against the malaria vector and parasite. This green synthesis of NPs allows the bio compounds plant polyphenols to enhance the effects of metallic NPs, causing an overall toxic impact on all developmental stages of the mosquito vector, reflected in larval mortality, reduced oviposition, impaired feeding habit, and morphological and behavioral changes. In addition, NPs synthesized with plant extracts show antiplasmodial effects by inhibiting parasite growth. Furthermore, choosing interventions against malaria should take into account equal access, sustainability, and human health impact.

8. Future Perspectives

The lack of evidence of the plant extracts’ mechanism of action becomes a significant issue in this review. The complex chemical composition of plants can hinder the emergence of resistance because the toxic compounds can vary between plant species. Although plant-derived nanoformulations are not currently used, we advocate for robust research regarding the long-term bioavailability of green nanoparticles, potential toxicity, production scaling, and administration feasibility.

Author Contributions

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

Funding

This research was funded by Universidade Federal Fluminense (PROPPI/UFF); Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) Grant [001]; Programa de Biotecnologia da Universidade Federal Fluminense (UFF); Programa de Pós Graduação em Biologia Molecular Celular (UNIRIO); Universidade Federal do Estado do Rio de Janeiro (UNIRIO); Fundação Carlos Chagas Filho de Amparo à Pesquisa do Estado do Rio de Janeiro (FAPERJ) Grants (E-26/010.000983/2019—Rede Rio de Inovação em Nanossistemas para a Saúde—NanoSAÚDE/FAPERJ; E-26/203.290/2017, and E-26/2010.592/2019, E-26/201.448/2021); an FINEP Grant (01.23.0139.00); Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq); and Instituto Oswaldo Cruz, FIOCRUZ.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the study’s design, in the collection, analyses, or interpretation of data, in the writing of the manuscript, or in the decision to publish the results.

References

  1. WHO. World Malaria Report; WHO: Geneva, Switzerland, 2022; p. 293.
  2. Gonzalez-Sanz, M.; Berzosa, P.; Norman, F.F. Updates on Malaria Epidemiology and Prevention Strategies. Curr. Infect. Dis. Rep. 2023, 25, 131–139. [Google Scholar] [CrossRef]
  3. Lu, H.Z.; Sui, Y.; Lobo, N.F.; Fouque, F.; Gao, C.; Lu, S.; Lv, S.; Deng, S.Q.; Wang, D.Q. Challenge and opportunity for vector control strategies on key mosquito-borne diseases during the COVID-19 pandemic. Front. Public Health 2023, 11, 1207293. [Google Scholar] [CrossRef] [PubMed]
  4. Loonen, J.; Dery, D.B.; Musaka, B.Z.; Bandibabone, J.B.; Bousema, T.; van Lenthe, M.; Pop-Stefanija, B.; Fesselet, J.F.; Koenraadt, C.J.M. Identification of main malaria vectors and their insecticide resistance profile in internally displaced and indigenous communities in Eastern Democratic Republic of the Congo (DRC). Malar. J. 2020, 19, 425. [Google Scholar] [CrossRef] [PubMed]
  5. Yokoly, F.N.; Zahouli, J.B.Z.; Small, G.; Ouattara, A.F.; Opoku, M.; de Souza, D.K.; Koudou, B.G. Assessing Anopheles vector species diversity and transmission of malaria in four health districts along the borders of Cote d’Ivoire. Malar. J. 2021, 20, 409. [Google Scholar] [CrossRef] [PubMed]
  6. Okuneye, K.; Eikenberry, S.E.; Gumel, A.B. Weather-driven malaria transmission model with gonotrophic and sporogonic cycles. J. Biol. Dyn. 2019, 13, 288–324. [Google Scholar] [CrossRef] [PubMed]
  7. Gallichotte, E.N.; Dobos, K.M.; Ebel, G.D.; Hagedorn, M.; Rasgon, J.L.; Richardson, J.H.; Stedman, T.T.; Barfield, J.P. Towards a method for cryopreservation of mosquito vectors of human pathogens. Cryobiology 2021, 99, 1–10. [Google Scholar] [CrossRef]
  8. Tripathi, H.; Bhalerao, P.; Singh, S.; Arya, H.; Alotaibi, B.S.; Rashid, S.; Hasan, M.R.; Bhatt, T.K. Malaria therapeutics: Are we close enough? Parasit. Vectors 2023, 16, 130. [Google Scholar] [CrossRef]
  9. WHO. WHO Guidelines for Malaria; WHO: Geneva, Switzerland, 2023; p. 447.
  10. Azmi, W.A.; Rizki, A.F.M.; Djuardi, Y.; Artika, I.M.; Siregar, J.E. Molecular insights into artemisinin resistance in Plasmodium falciparum: An updated review. Infect. Genet. Evol. 2023, 112, 105460. [Google Scholar] [CrossRef]
  11. Wang, S.; Huang, F.; Yan, H.; Yin, J.; Xia, Z. A review of malaria molecular markers for drug resistance in Plasmodium falciparum and Plasmodium vivax in China. Front. Cell. Infect. Microbiol. 2023, 13, 1167220. [Google Scholar] [CrossRef]
  12. Chaves, J.B.; Portugal Tavares de Moraes, B.; Regina Ferrarini, S.; Noe da Fonseca, F.; Silva, A.R.; Goncalves-de-Albuquerque, C.F. Potential of nanoformulations in malaria treatment. Front. Pharmacol. 2022, 13, 999300. [Google Scholar] [CrossRef]
  13. Haldar, K.; Bhattacharjee, S.; Safeukui, I. Drug resistance in Plasmodium. Nat. Rev. Microbiol. 2018, 16, 156–170. [Google Scholar] [CrossRef]
  14. Wicht, K.J.; Mok, S.; Fidock, D.A. Molecular Mechanisms of Drug Resistance in Plasmodium falciparum Malaria. Annu. Rev. Microbiol. 2020, 74, 431–454. [Google Scholar] [CrossRef] [PubMed]
  15. Mishra, M.; Mishra, V.K.; Kashaw, V.; Iyer, A.K.; Kashaw, S.K. Comprehensive review on various strategies for antimalarial drug discovery. Eur. J. Med. Chem. 2017, 125, 1300–1320. [Google Scholar] [CrossRef] [PubMed]
  16. Sougoufara, S.; Diedhiou, S.M.; Doucoure, S.; Diagne, N.; Sembene, P.M.; Harry, M.; Trape, J.F.; Sokhna, C.; Ndiath, M.O. Biting by Anopheles funestus in broad daylight after use of long-lasting insecticidal nets: A new challenge to malaria elimination. Malar. J. 2014, 13, 125. [Google Scholar] [CrossRef]
  17. Matowo, N.S.; Martin, J.; Kulkarni, M.A.; Mosha, J.F.; Lukole, E.; Isaya, G.; Shirima, B.; Kaaya, R.; Moyes, C.; Hancock, P.A.; et al. An increasing role of pyrethroid-resistant Anopheles funestus in malaria transmission in the Lake Zone, Tanzania. Sci. Rep. 2021, 11, 13457. [Google Scholar] [CrossRef] [PubMed]
  18. Soma, D.D.; Zogo, B.M.; Some, A.; Tchiekoi, B.N.; Hien, D.F.S.; Pooda, H.S.; Coulibaly, S.; Gnambani, J.E.; Ouari, A.; Mouline, K.; et al. Anopheles bionomics, insecticide resistance and malaria transmission in southwest Burkina Faso: A pre-intervention study. PLoS ONE 2020, 15, e0236920. [Google Scholar] [CrossRef] [PubMed]
  19. Riveron, J.M.; Chiumia, M.; Menze, B.D.; Barnes, K.G.; Irving, H.; Ibrahim, S.S.; Weedall, G.D.; Mzilahowa, T.; Wondji, C.S. Rise of multiple insecticide resistance in Anopheles funestus in Malawi: A major concern for malaria vector control. Malar. J. 2015, 14, 344. [Google Scholar] [CrossRef]
  20. Bayda, S.; Adeel, M.; Tuccinardi, T.; Cordani, M.; Rizzolio, F. The History of Nanoscience and Nanotechnology: From Chemical-Physical Applications to Nanomedicine. Molecules 2019, 25, 112. [Google Scholar] [CrossRef]
  21. Zahoor, M.; Nazir, N.; Iftikhar, M.; Naz, S.; Zekker, I.; Burlakovs, J.; Uddin, F.; Kamran, A.W.; Kallistova, A.; Pimenov, N.; et al. A Review on Silver Nanoparticles: Classification, Various Methods of Synthesis, and Their Potential Roles in Biomedical Applications and Water Treatment. Water 2021, 13, 2216. [Google Scholar] [CrossRef]
  22. Duan, H.; Wang, D.; Li, Y. Green chemistry for nanoparticle synthesis. Chem. Soc. Rev. 2015, 44, 5778–5792. [Google Scholar] [CrossRef]
  23. Aldakheel, F.M.; Sayed, M.M.E.; Mohsen, D.; Fagir, M.H.; El Dein, D.K. Green Synthesis of Silver Nanoparticles Loaded Hydrogel for Wound Healing; Systematic Review. Gels 2023, 9, 530. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, L.; Li, X.; Yue, G.; Guo, L.; Hu, Y.; Cui, Q.; Wang, J.; Tang, J.; Liu, H. Nanodrugs systems for therapy and diagnosis of esophageal cancer. Front. Bioeng. Biotechnol. 2023, 11, 1233476. [Google Scholar] [CrossRef] [PubMed]
  25. Hao, Y.; Ji, Z.; Zhou, H.; Wu, D.; Gu, Z.; Wang, D.; Ten Dijke, P. Lipid-based nanoparticles as drug delivery systems for cancer immunotherapy. MedComm 2023, 4, e339. [Google Scholar] [CrossRef] [PubMed]
  26. Shoeb, E.; Badar, U.; Venkataraman, S.; Hefferon, K. Frontiers in Bioengineering and Biotechnology: Plant Nanoparticles for Anti-Cancer Therapy. Vaccines 2021, 9, 830. [Google Scholar] [CrossRef]
  27. Yazdanian, M.; Rostamzadeh, P.; Rahbar, M.; Alam, M.; Abbasi, K.; Tahmasebi, E.; Tebyaniyan, H.; Ranjbar, R.; Seifalian, A.; Yazdanian, A. The Potential Application of Green-Synthesized Metal Nanoparticles in Dentistry: A Comprehensive Review. Bioinorg. Chem. Appl. 2022, 2022, 2311910. [Google Scholar] [CrossRef] [PubMed]
  28. Gupta, V.; Mohapatra, S.; Mishra, H.; Farooq, U.; Kumar, K.; Ansari, M.J.; Aldawsari, M.F.; Alalaiwe, A.S.; Mirza, M.A.; Iqbal, Z. Nanotechnology in Cosmetics and Cosmeceuticals-A Review of Latest Advancements. Gels 2022, 8, 173. [Google Scholar] [CrossRef]
  29. Ghobashy, M.M.; Elkodous, M.A.; Shabaka, S.H.; Younis, S.A.; Alshangiti, D.M.; Madani, M.; Al-Gahtany, S.A.; Elkhatib, W.F.; Noreddin, A.M.; Nady, N.; et al. An overview of methods for production and detection of silver nanoparticles, with emphasis on their fate and toxicological effects on human, soil, and aquatic environment. Nanotechnol. Rev. 2021, 10, 954–977. [Google Scholar] [CrossRef]
  30. Noga, M.; Milan, J.; Frydrych, A.; Jurowski, K. Toxicological Aspects, Safety Assessment, and Green Toxicology of Silver Nanoparticles (AgNPs)— Critical Review: State of the Art. Int. J. Mol. Sci. 2023, 24, 5133. [Google Scholar] [CrossRef]
  31. Mikhailova, E.O. Silver Nanoparticles: Mechanism of Action and Probable Bio-Application. J. Funct. Biomater. 2020, 11, 84. [Google Scholar] [CrossRef]
  32. Javed, R.; Zia, M.; Naz, S.; Aisida, S.O.; Ain, N.U.; Ao, Q. Role of capping agents in the application of nanoparticles in biomedicine and environmental remediation: Recent trends and future prospects. J. Nanobiotechnol. 2020, 18, 172. [Google Scholar] [CrossRef]
  33. Borges, D.F.; Lopes, E.A.; Fialho Moraes, A.R.; Soares, M.S.; Visôtto, L.E.; Oliveira, C.R.; Moreira Valente, V.M. Formulation of botanicals for the control of plant-pathogens: A review. Crop. Protect. 2018, 110, 135–140. [Google Scholar] [CrossRef]
  34. Jha, A.K.; Prasad, K.; Prasad, K.; Kulkarni, A.R. Plant system: Nature’s nanofactory. Colloids Surf. B Biointerfaces 2009, 73, 219–223. [Google Scholar] [CrossRef] [PubMed]
  35. Miu, B.A.; Dinischiotu, A. New Green Approaches in Nanoparticles Synthesis: An Overview. Molecules 2022, 27, 6472. [Google Scholar] [CrossRef] [PubMed]
  36. Gour, A.; Jain, N.K. Advances in green synthesis of nanoparticles. Artif. Cells Nanomed. Biotechnol. 2019, 47, 844–851. [Google Scholar] [CrossRef] [PubMed]
  37. Ahmadi Tehrani, A.; Omranpoor, M.M.; Vatanara, A.; Seyedabadi, M.; Ramezani, V. Formation of nanosuspensions in bottom-up approach: Theories and optimization. Daru 2019, 27, 451–473. [Google Scholar] [CrossRef]
  38. Chugh, D.; Viswamalya, V.S.; Das, B. Green synthesis of silver nanoparticles with algae and the importance of capping agents in the process. J. Genet. Eng. Biotechnol. 2021, 19, 126. [Google Scholar] [CrossRef]
  39. Ahmed, S.; Alhareth, K.; Mignet, N. Advancement in nanogel formulations provides controlled drug release. Int. J. Pharm. 2020, 584, 119435. [Google Scholar] [CrossRef]
  40. Singh, Y.; Meher, J.G.; Raval, K.; Khan, F.A.; Chaurasia, M.; Jain, N.K.; Chourasia, M.K. Nanoemulsion: Concepts, development and applications in drug delivery. J. Control. Release 2017, 252, 28–49. [Google Scholar] [CrossRef]
  41. Hawadak, J.; Kojom Foko, L.P.; Pande, V.; Singh, V. In vitro antiplasmodial activity, hemocompatibility and temporal stability of Azadirachta indica silver nanoparticles. Artif. Cells Nanomed. Biotechnol. 2022, 50, 286–300. [Google Scholar] [CrossRef] [PubMed]
  42. Santos-Magalhaes, N.S.; Mosqueira, V.C. Nanotechnology applied to the treatment of malaria. Adv. Drug Deliv. Rev. 2010, 62, 560–575. [Google Scholar] [CrossRef]
  43. Joshi, M.; Pathak, S.; Sharma, S.; Patravale, V. Solid microemulsion preconcentrate (NanOsorb) of artemether for effective treatment of malaria. Int. J. Pharm. 2008, 362, 172–178. [Google Scholar] [CrossRef]
  44. Kumar, R.; Ray, P.C.; Datta, D.; Bansal, G.P.; Angov, E.; Kumar, N. Nanovaccines for malaria using Plasmodium falciparum antigen Pfs25 attached gold nanoparticles. Vaccine 2015, 33, 5064–5071. [Google Scholar] [CrossRef] [PubMed]
  45. Zaker, M. Natural Plant Products as Eco-friendly Fungicides for Plant Diseases Control—A Review. Agriculturists 2016, 14, 134–141. [Google Scholar] [CrossRef]
  46. Benelli, G.; Maggi, F.; Pavela, R.; Murugan, K.; Govindarajan, M.; Vaseeharan, B.; Petrelli, R.; Cappellacci, L.; Kumar, S.; Hofer, A.; et al. Mosquito control with green nanopesticides: Towards the One Health approach? A review of non-target effects. Environ. Sci. Pollut. Res. Int. 2018, 25, 10184–10206. [Google Scholar] [CrossRef]
  47. Ali, M.Y.S.; Ravikumar, S.; Beula, J.M. Mosquito larvicidal activity of seaweeds extracts against Anopheles stephensi, Aedes aegypti and Culex quinquefasciatus. Asian Pac. J. Trop. Dis. 2013, 3, 196–201. [Google Scholar] [CrossRef]
  48. Rajeswary, M.; Govindarajan, M.; Alharbi, N.S.; Kadaikunnan, S.; Khaled, J.M.; Benelli, G. Zingiber cernuum (Zingiberaceae) essential oil as effective larvicide and oviposition deterrent on six mosquito vectors, with little non-target toxicity on four aquatic mosquito predators. Environ. Sci. Pollut. Res. Int. 2018, 25, 10307–10316. [Google Scholar] [CrossRef]
  49. Govindarajan, M.; Benelli, G. Eco-friendly larvicides from Indian plants: Effectiveness of lavandulyl acetate and bicyclogermacrene on malaria, dengue and Japanese encephalitis mosquito vectors. Ecotoxicol. Environ. Saf. 2016, 133, 395–402. [Google Scholar] [CrossRef] [PubMed]
  50. Ribeiro, G.D.J.G.; Rei Yan, S.L.; Palmisano, G.; Wrenger, C. Plant Extracts as a Source of Natural Products with Potential Antimalarial Effects: An Update from 2018 to 2022. Pharmaceutics 2023, 15, 1638. [Google Scholar] [CrossRef]
  51. Bagavan, A.; Rahuman, A.A.; Kaushik, N.K.; Sahal, D. In vitro antimalarial activity of medicinal plant extracts against Plasmodium falciparum. Parasitol. Res. 2011, 108, 15–22. [Google Scholar] [CrossRef]
  52. Elango, G.; Rahuman, A.A.; Kamaraj, C.; Bagavan, A.; Zahir, A.A. Efficacy of medicinal plant extracts against malarial vector, Anopheles subpictus Grassi. Parasitol. Res. 2011, 108, 1437–1445. [Google Scholar] [CrossRef]
  53. Kamaraj, C.; Kaushik, N.K.; Mohanakrishnan, D.; Elango, G.; Bagavan, A.; Zahir, A.A.; Rahuman, A.A.; Sahal, D. Antiplasmodial potential of medicinal plant extracts from Malaiyur and Javadhu hills of South India. Parasitol. Res. 2012, 111, 703–715. [Google Scholar] [CrossRef]
  54. Kamaraj, C.; Kaushik, N.K.; Rahuman, A.A.; Mohanakrishnan, D.; Bagavan, A.; Elango, G.; Zahir, A.A.; Santhoshkumar, T.; Marimuthu, S.; Jayaseelan, C.; et al. Antimalarial activities of medicinal plants traditionally used in the villages of Dharmapuri regions of South India. J. Ethnopharmacol. 2012, 141, 796–802. [Google Scholar] [CrossRef] [PubMed]
  55. Kaushik, N.K.; Bagavan, A.; Rahuman, A.A.; Zahir, A.A.; Kamaraj, C.; Elango, G.; Jayaseelan, C.; Kirthi, A.V.; Santhoshkumar, T.; Marimuthu, S.; et al. Evaluation of antiplasmodial activity of medicinal plants from North Indian Buchpora and South Indian Eastern Ghats. Malar. J. 2015, 14, 65. [Google Scholar] [CrossRef] [PubMed]
  56. Alaithan, H.; Kumar, N.; Islam, M.Z.; Liappis, A.P.; Nava, V.E. Novel Therapeutics for Malaria. Pharmaceutics 2023, 15, 1800. [Google Scholar] [CrossRef]
  57. Yuan, H.; Ma, Q.; Ye, L.; Piao, G. The Traditional Medicine and Modern Medicine from Natural Products. Molecules 2016, 21, 559. [Google Scholar] [CrossRef] [PubMed]
  58. Patel, N.; Zinzuvadia, A.; Prajapati, M.; Tyagi, R.K.; Dalai, S. Swertiamarin-mediated immune modulation/adaptation confers protection against Plasmodium berghei. Future Microbiol. 2022, 17, 931–941. [Google Scholar] [CrossRef] [PubMed]
  59. Zuccari, G.; Alfei, S. Development of Phytochemical Delivery Systems by Nano-Suspension and Nano-Emulsion Techniques. Int. J. Mol. Sci. 2023, 24, 9824. [Google Scholar] [CrossRef]
  60. Ahmed, T.; Hyder, M.Z.; Liaqat, I.; Scholz, M. Climatic Conditions: Conventional and Nanotechnology-Based Methods for the Control of Mosquito Vectors Causing Human Health Issues. Int. J. Environ. Res. Public Health 2019, 16, 3165. [Google Scholar] [CrossRef]
  61. Vivekanandhan, P.; Swathy, K.; Thomas, A.; Kweka, E.J.; Rahman, A.; Pittarate, S.; Krutmuang, P. Insecticidal Efficacy of Microbial-Mediated Synthesized Copper Nano-Pesticide against Insect Pests and Non-Target Organisms. Int. J. Environ. Res. Public Health 2021, 18, 10536. [Google Scholar] [CrossRef]
  62. Murugan, K.; Subramaniam, J.; Rajaganesh, R.; Panneerselvam, C.; Amuthavalli, P.; Vasanthakumaran, M.; Jayashanthini, S.; Dinesh, D.; Anitha, J.; Wang, L.; et al. Efficacy and side effects of bio-fabricated sardine fish scale silver nanoparticles against malarial vector Anopheles stephensi. Sci. Rep. 2021, 11, 19567. [Google Scholar] [CrossRef]
  63. Thelma, J.; Balasubramanian, C. Ovicidal, larvicidal and pupicidal efficacy of silver nanoparticles synthesized by Bacillus marisflavi against the chosen mosquito species. PLoS ONE 2021, 16, e0260253. [Google Scholar] [CrossRef]
  64. Murugan, K.; Anitha, J.; Dinesh, D.; Suresh, U.; Rajaganesh, R.; Chandramohan, B.; Subramaniam, J.; Paulpandi, M.; Vadivalagan, C.; Amuthavalli, P.; et al. Fabrication of nano-mosquitocides using chitosan from crab shells: Impact on non-target organisms in the aquatic environment. Ecotoxicol. Environ. Saf. 2016, 132, 318–328. [Google Scholar] [CrossRef] [PubMed]
  65. Salunkhe, R.B.; Patil, S.V.; Patil, C.D.; Salunke, B.K. Larvicidal potential of silver nanoparticles synthesized using fungus Cochliobolus lunatus against Aedes aegypti (Linnaeus, 1762) and Anopheles stephensi Liston (Diptera; Culicidae). Parasitol. Res. 2011, 109, 823–831. [Google Scholar] [CrossRef] [PubMed]
  66. Jaganathan, A.; Murugan, K.; Panneerselvam, C.; Madhiyazhagan, P.; Dinesh, D.; Vadivalagan, C.; Aziz, A.T.; Chandramohan, B.; Suresh, U.; Rajaganesh, R.; et al. Earthworm-mediated synthesis of silver nanoparticles: A potent tool against hepatocellular carcinoma, Plasmodium falciparum parasites and malaria mosquitoes. Parasitol. Int. 2016, 65, 276–284. [Google Scholar] [CrossRef]
  67. Roux, O.; Robert, V. Larval predation in malaria vectors and its potential implication in malaria transmission: An overlooked ecosystem service? Parasit. Vectors 2019, 12, 217. [Google Scholar] [CrossRef]
  68. Al Mukhaini, S.K.; Mohammed, O.A.; Gerbers, S.; Al Awaidy, S.T. The Progress Towards National Malaria Elimination: The Experience of Oman. Oman Med. J. 2023, 38, e500. [Google Scholar] [CrossRef]
  69. Bosly, H.A.E. Evaluation of larvicidal enhanced activity of sandalwood oil via nano-emulsion against Culex pipiens and Ades aegypti. Saudi J. Biol. Sci. 2022, 29, 103455. [Google Scholar] [CrossRef]
  70. Ferreira, R.; D’Haveloose, N.P.; Cruz, R.A.S.; Araujo, R.S.; Carvalho, J.C.T.; Rocha, L.; Fernandes, L.P.; Da Costa, T.S.; Fernandes, C.P.; Souto, R.N.P. Nano-emulsification Enhances the Larvicidal Potential of the Essential Oil of Siparuna guianensis (Laurales: Siparunaceae) Against Aedes (Stegomyia) aegypti (Diptera: Culicidae). J. Med. Entomol. 2020, 57, 788–796. [Google Scholar] [CrossRef]
  71. Gnanadesigan, M.; Anand, M.; Ravikumar, S.; Maruthupandy, M.; Vijayakumar, V.; Selvam, S.; Dhineshkumar, M.; Kumaraguru, A.K. Biosynthesis of silver nanoparticles by using mangrove plant extract and their potential mosquito larvicidal property. Asian Pac. J. Trop. Med. 2011, 4, 799–803. [Google Scholar] [CrossRef]
  72. Hajra, A.; Dutta, S.; Mondal, N.K. Mosquito larvicidal activity of cadmium nanoparticles synthesized from petal extracts of marigold (Tagetes sp.) and rose (Rosa sp.) flower. J. Parasit. Dis. 2016, 40, 1519–1527. [Google Scholar] [CrossRef]
  73. Lobato Rodrigues, A.B.; Martins, R.L.; Rabelo, E.M.; Tomazi, R.; Santos, L.L.; Brandao, L.B.; Faustino, C.G.; Ferreira Farias, A.L.; Dos Santos, C.B.R.; de Castro Cantuaria, P.; et al. Development of nano-emulsions based on Ayapana triplinervis essential oil for the control of Aedes aegypti larvae. PLoS ONE 2021, 16, e0254225. [Google Scholar] [CrossRef]
  74. Rajaganesh, R.; Murugan, K.; Panneerselvam, C.; Jayashanthini, S.; Aziz, A.T.; Roni, M.; Suresh, U.; Trivedi, S.; Rehman, H.; Higuchi, A.; et al. Fern-synthesized silver nanocrystals: Towards a new class of mosquito oviposition deterrents? Res. Vet. Sci. 2016, 109, 40–51. [Google Scholar] [CrossRef] [PubMed]
  75. Suresh, U.; Murugan, K.; Benelli, G.; Nicoletti, M.; Barnard, D.R.; Panneerselvam, C.; Kumar, P.M.; Subramaniam, J.; Dinesh, D.; Chandramohan, B. Tackling the growing threat of dengue: Phyllanthus niruri-mediated synthesis of silver nanoparticles and their mosquitocidal properties against the dengue vector Aedes aegypti (Diptera: Culicidae). Parasitol. Res. 2015, 114, 1551–1562. [Google Scholar] [CrossRef]
  76. Udappusamy, V.; Mohan, H.; Thinagaran, R. Lantana camara L. essential oil mediated nano-emulsion formulation for biocontrol application: Anti-mosquitocidal, anti-microbial and antioxidant assay. Arch. Microbiol. 2022, 204, 388. [Google Scholar] [CrossRef] [PubMed]
  77. Wu, W.; Yang, Y.; Feng, Y.; Ren, X.; Li, Y.; Li, W.; Huang, J.; Kong, L.; Chen, X.; Lin, Z.; et al. Study of the Repellent Activity of 60 Essential Oils and Their Main Constituents against Aedes albopictus, and Nano-Formulation Development. Insects 2022, 13, 1077. [Google Scholar] [CrossRef]
  78. Abd Elghani, E.M.; El Sayed, A.M.; Abdel-Aziz Emam, M.M.; Al-Mahallawi, A.M.; Tadros, S.H.; Soliman, F.M.; Youssef, F.S. Seasonal metabolic profiling of Valencia orange leaf essential oil using GC coupled with chemometrics, nano-formulation, and insecticidal evaluation: In vivo and in silico. RSC Adv. 2023, 13, 1659–1671. [Google Scholar] [CrossRef]
  79. Murugan, K.; Dinesh, D.; Nataraj, D.; Subramaniam, J.; Amuthavalli, P.; Madhavan, J.; Rajasekar, A.; Rajan, M.; Thiruppathi, K.P.; Kumar, S.; et al. Iron and iron oxide nanoparticles are highly toxic to Culex quinquefasciatus with little non-target effects on larvivorous fishes. Environ. Sci. Pollut. Res. Int. 2018, 25, 10504–10514. [Google Scholar] [CrossRef] [PubMed]
  80. Vincent, S.; Kovendan, K.; Chandramohan, B.; Kamalakannan, S.; Kumar, P.M.; Vasugi, C.; Praseeja, C.; Subramaniam, J.; Govindarajan, M.; Murugan, K.; et al. Swift Fabrication of Silver Nanoparticles Using Bougainvillea glabra: Potential Against the Japanese Encephalitis Vector, Culex tritaeniorhynchus Giles (Diptera: Culicidae). J. Clust. Sci. 2017, 28, 37–58. [Google Scholar] [CrossRef]
  81. Alyahya, S.A.; Govindarajan, M.; Alharbi, N.S.; Kadaikunnan, S.; Khaled, J.M.; Mothana, R.A.; Al-anbr, M.N.; Vaseeharan, B.; Ishwarya, R.; Yazhiniprabha, M.; et al. Swift fabrication of Ag nanostructures using a colloidal solution of Holostemma ada-kodien (Apocynaceae)—Antibiofilm potential, insecticidal activity against mosquitoes and non-target impact on water bugs. J. Photochem. Photobiol. B Biol. 2018, 181, 70–79. [Google Scholar] [CrossRef]
  82. Osanloo, M.; Firooziyan, S.; Abdollahi, A.; Hatami, S.; Nematollahi, A.; Elahi, N.; Zarenezhad, E. Nanoemulsion and nanogel containing Artemisia dracunculus essential oil; larvicidal effect and antibacterial activity. BMC Res. Notes 2022, 15, 276. [Google Scholar] [CrossRef]
  83. Rawani, A.; Ghosh, A.; Chandra, G. Mosquito larvicidal and antimicrobial activity of synthesized nano-crystalline silver particles using leaves and green berry extract of Solanum nigrum L. (Solanaceae: Solanales). Acta Trop. 2013, 128, 613–622. [Google Scholar] [CrossRef]
  84. Rajakumar, G.; Rahuman, A.A.; Roopan, S.M.; Chung, I.M.; Anbarasan, K.; Karthikeyan, V. Efficacy of larvicidal activity of green synthesized titanium dioxide nanoparticles using Mangifera indica extract against blood-feeding parasites. Parasitol. Res. 2015, 114, 571–581. [Google Scholar] [CrossRef]
  85. Abrantes, D.C.; Rogerio, C.B.; Campos, E.V.R.; Germano-Costa, T.; Vigato, A.A.; Machado, I.P.; Sepulveda, A.F.; Lima, R.; de Araujo, D.R.; Fraceto, L.F. Repellent active ingredients encapsulated in polymeric nanoparticles: Potential alternative formulations to control arboviruses. J. Nanobiotechnol. 2022, 20, 520. [Google Scholar] [CrossRef]
  86. Chang, M.A.; Impoinvil, D.; Hamre, K.E.S.; Dalexis, P.E.; Merilien, J.B.; Dismer, A.M.; Fouche, B.; Desir, L.; Holmes, K.; Lafortune, W.; et al. Acceptability, Feasibility, Drug Safety, and Effectiveness of a Pilot Mass Drug Administration with a Single Round of Sulfadoxine-Pyrimethamine Plus Primaquine and Indoor Residual Spraying in Communities with Malaria Transmission in Haiti, 2018. Am. J. Trop. Med. Hyg. 2023, 108, 1127–1139. [Google Scholar] [CrossRef] [PubMed]
  87. WHO. Vector Control Products Targeting Outdoor Malaria Transmission: Preferred Product Characteristics; WHO: Geneva, Switzerland, 2023; p. 8.
  88. Ujihara, K. The history of extensive structural modifications of pyrethroids. J. Pestic. Sci. 2019, 44, 215–224. [Google Scholar] [CrossRef]
  89. Dong, K.; Du, Y.; Rinkevich, F.; Nomura, Y.; Xu, P.; Wang, L.; Silver, K.; Zhorov, B.S. Molecular biology of insect sodium channels and pyrethroid resistance. Insect. Biochem. Mol. Biol. 2014, 50, 1–17. [Google Scholar] [CrossRef] [PubMed]
  90. Kansal, I.; Kapoor, A.; Solanki, S.; Singh, R. Cypermethrin toxicity in the environment: Analytical insight into detection methods and microbial degradation pathways. J. Appl. Microbiol. 2023, 134, lxad105. [Google Scholar] [CrossRef]
  91. Miao, W.; Jiang, Y.; Hong, Q.; Sheng, H.; Liu, P.; Huang, Y.; Cheng, J.; Pan, X.; Yu, Q.; Wu, Y.; et al. Systematic evaluation of the toxicological effects of deltamethrin exposure in zebrafish larvae. Environ. Toxicol. Pharmacol. 2023, 100, 104155. [Google Scholar] [CrossRef] [PubMed]
  92. Ramesh, M.; Bindu, C.F.; Mohanthi, S.; Hema, T.; Poopal, R.K.; Ren, Z.; Li, B. Efficiency of hematological, enzymological and oxidative stress biomarkers of Cyprinus carpio to an emerging organic compound (alphamethrin) toxicity. Environ. Toxicol. Pharmacol. 2023, 101, 104186. [Google Scholar] [CrossRef] [PubMed]
  93. Gimenez-Asensio, M.J.; Hernandez, A.F.; Romero-Molina, D.; Gonzalez-Alzaga, B.; Luzardo, O.P.; Henríquez-Hernández, L.A.; Boada, L.D.; García-Cortés, H.; Lopez-Flores, I.; Sanchez-Piedra, M.D.; et al. Effect of prenatal exposure to organophosphates and pyrethroid pesticides on neonatal anthropometric measures and gestational age. Environ. Res. 2023, 232, 116410. [Google Scholar] [CrossRef]
  94. Sheikh, I.A.; Beg, M.A.; Hamoda, T.A.; Mandourah, H.M.S.; Memili, E. Androgen receptor signaling and pyrethroids: Potential male infertility consequences. Front. Cell. Dev. Biol. 2023, 11, 1173575. [Google Scholar] [CrossRef]
  95. Murugan, K.; Aarthi, N.; Kovendan, K.; Panneerselvam, C.; Chandramohan, B.; Kumar, P.M.; Amerasan, D.; Paulpandi, M.; Chandirasekar, R.; Dinesh, D.; et al. Mosquitocidal and antiplasmodial activity of Senna occidentalis (Cassiae) and Ocimum basilicum (Lamiaceae) from Maruthamalai hills against Anopheles stephensi and Plasmodium falciparum. Parasitol. Res. 2015, 114, 3657–3664. [Google Scholar] [CrossRef]
  96. Panneerselvam, C.; Murugan, K.; Roni, M.; Aziz, A.T.; Suresh, U.; Rajaganesh, R.; Madhiyazhagan, P.; Subramaniam, J.; Dinesh, D.; Nicoletti, M.; et al. Fern-synthesized nanoparticles in the fight against malaria: LC/MS analysis of Pteridium aquilinum leaf extract and biosynthesis of silver nanoparticles with high mosquitocidal and antiplasmodial activity. Parasitol. Res. 2016, 115, 997–1013. [Google Scholar] [CrossRef]
  97. Murugan, K.; Samidoss, C.M.; Panneerselvam, C.; Higuchi, A.; Roni, M.; Suresh, U.; Chandramohan, B.; Subramaniam, J.; Madhiyazhagan, P.; Dinesh, D.; et al. Seaweed-synthesized silver nanoparticles: An eco-friendly tool in the fight against Plasmodium falciparum and its vector Anopheles stephensi? Parasitol. Res. 2015, 114, 4087–4097. [Google Scholar] [CrossRef]
  98. Minal, S.P.; Prakash, S. Laboratory analysis of Au–Pd bimetallic nanoparticles synthesized with Citrus limon leaf extract and its efficacy on mosquito larvae and non-target organisms. Sci. Rep. 2020, 10, 21610. [Google Scholar] [CrossRef]
  99. Kalpana, V.N.; Alarjani, K.M.; Rajeswari, V.D. Enhancing malaria control using Lagenaria siceraria and its mediated zinc oxide nanoparticles against the vector Anopheles stephensi and its parasite Plasmodium falciparum. Sci. Rep. 2020, 10, 21568. [Google Scholar] [CrossRef]
  100. Murugan, K.; Benelli, G.; Panneerselvam, C.; Subramaniam, J.; Jeyalalitha, T.; Dinesh, D.; Nicoletti, M.; Hwang, J.S.; Suresh, U.; Madhiyazhagan, P. Cymbopogon citratus-synthesized gold nanoparticles boost the predation efficiency of copepod Mesocyclops aspericornis against malaria and dengue mosquitoes. Exp. Parasitol. 2015, 153, 129–138. [Google Scholar] [CrossRef] [PubMed]
  101. Govindarajan, M.; Hoti, S.L.; Benelli, G. Facile fabrication of eco-friendly nano-mosquitocides: Biophysical characterization and effectiveness on neglected tropical mosquito vectors. Enzyme Microb. Technol. 2016, 95, 155–163. [Google Scholar] [CrossRef] [PubMed]
  102. AlQahtani, F.S.; AlShebly, M.M.; Govindarajan, M.; Senthilmurugan, S.; Vijayan, P.; Benelli, G. Green and facile biosynthesis of silver nanocomposites using the aqueous extract of Rubus ellipticus leaves: Toxicity and oviposition deterrent activity against Zika virus, malaria and filariasis mosquito vectors. J. Asia-Pac. Entomol. 2017, 20, 157–164. [Google Scholar] [CrossRef]
  103. Azarudeen, R.M.S.T.; Govindarajan, M.; Amsath, A.; Muthukumaran, U.; Benelli, G. Single-Step Biofabrication of Silver Nanocrystals Using Naregamia alata: A Cost Effective and Eco-Friendly Control Tool in the Fight Against Malaria, Zika Virus and St. Louis Encephalitis Mosquito Vectors. J. Clust. Sci. 2017, 28, 179–203. [Google Scholar] [CrossRef]
  104. Govindarajan, M.; Kadaikunnan, S.; Alharbi, N.S.; Benelli, G. Single-step biological fabrication of colloidal silver nanoparticles using Hugonia mystax: Larvicidal potential against Zika virus, dengue, and malaria vector mosquitoes. Artif. Cells Nanomed. Biotechnol. 2017, 45, 1317–1325. [Google Scholar] [CrossRef]
  105. Kumar, P.; Kumar, D.; Kumar, V.; Chauhan, R.; Singh, H. Mosquito larvicidal potential of Solanum xanthocarpum leaf extract derived silver nanoparticles and its bio-toxicity on non-target aquatic organism. J. Vector Borne Dis. 2022, 59, 216–227. [Google Scholar] [CrossRef]
  106. Gandhi, P.R.; Jayaseelan, C.; Kamaraj, C.; Rajasree, S.R.R.; Regina Mary, R. In vitro antimalarial activity of synthesized TiO2 nanoparticles using Momordica charantia leaf extract against Plasmodium falciparum. J. Appl. Biomed. 2018, 16, 378–386. [Google Scholar] [CrossRef]
  107. Sujitha, V.; Murugan, K.; Dinesh, D.; Pandiyan, A.; Aruliah, R.; Hwang, J.S.; Kalimuthu, K.; Panneerselvam, C.; Higuchi, A.; Aziz, A.T.; et al. Green-synthesized CdS nano-pesticides: Toxicity on young instars of malaria vectors and impact on enzymatic activities of the non-target mud crab Scylla serrata. Aquat. Toxicol. 2017, 188, 100–108. [Google Scholar] [CrossRef] [PubMed]
  108. Govindarajan, M.; Hoti, S.L.; Rajeswary, M.; Benelli, G. One-step synthesis of polydispersed silver nanocrystals using Malva sylvestris: An eco-friendly mosquito larvicide with negligible impact on non-target aquatic organisms. Parasitol. Res. 2016, 115, 2685–2695. [Google Scholar] [CrossRef]
  109. Kojom Foko, L.P.; Hawadak, J.; Verma, V.; Belle Ebanda Kedi, P.; Eboumbou Moukoko, C.E.; Kamaraju, R.; Pande, V.; Singh, V. Phytofabrication and characterization of Alchornea cordifolia silver nanoparticles and evaluation of antiplasmodial, hemocompatibility and larvicidal potential. Front. Bioeng. Biotechnol. 2023, 11, 1109841. [Google Scholar] [CrossRef]
  110. Murugan, K.; Labeeba, M.A.; Panneerselvam, C.; Dinesh, D.; Suresh, U.; Subramaniam, J.; Madhiyazhagan, P.; Hwang, J.S.; Wang, L.; Nicoletti, M.; et al. Aristolochia indica green-synthesized silver nanoparticles: A sustainable control tool against the malaria vector Anopheles stephensi? Res. Vet. Sci. 2015, 102, 127–135. [Google Scholar] [CrossRef] [PubMed]
  111. Kala, S.; Sogan, N.; Naik, S.N.; Agarwal, A.; Kumar, J. Impregnation of pectin-cedarwood essential oil nanocapsules onto mini cotton bag improves larvicidal performances. Sci. Rep. 2020, 10, 14107. [Google Scholar] [CrossRef]
  112. Roopan, S.M.; Rohit; Madhumitha, G.; Rahuman, A.A.; Kamaraj, C.; Bharathi, A.; Surendra, T.V. Low-cost and eco-friendly phyto-synthesis of silver nanoparticles using Cocos nucifera coir extract and its larvicidal activity. Ind. Crops Prod. 2013, 43, 631–635. [Google Scholar] [CrossRef]
  113. Subramaniam, J.; Murugan, K.; Panneerselvam, C.; Kovendan, K.; Madhiyazhagan, P.; Dinesh, D.; Kumar, P.M.; Chandramohan, B.; Suresh, U.; Rajaganesh, R.; et al. Multipurpose effectiveness of Couroupita guianensis-synthesized gold nanoparticles: High antiplasmodial potential, field efficacy against malaria vectors and synergy with Aplocheilus lineatus predators. Environ. Sci. Pollut. Res. Int. 2016, 23, 7543–7558. [Google Scholar] [CrossRef] [PubMed]
  114. Murugan, K.; Panneerselvam, C.; Subramaniam, J.; Madhiyazhagan, P.; Hwang, J.S.; Wang, L.; Dinesh, D.; Suresh, U.; Roni, M.; Higuchi, A.; et al. Eco-friendly drugs from the marine environment: Spongeweed-synthesized silver nanoparticles are highly effective on Plasmodium falciparum and its vector Anopheles stephensi, with little non-target effects on predatory copepods. Environ. Sci. Pollut. Res. Int. 2016, 23, 16671–16685. [Google Scholar] [CrossRef]
  115. Subramaniam, J.; Murugan, K.; Panneerselvam, C.; Kovendan, K.; Madhiyazhagan, P.; Kumar, P.M.; Dinesh, D.; Chandramohan, B.; Suresh, U.; Nicoletti, M.; et al. Eco-friendly control of malaria and arbovirus vectors using the mosquitofish Gambusia affinis and ultra-low dosages of Mimusops elengi-synthesized silver nanoparticles: Towards an integrative approach? Environ. Sci. Pollut. Res. 2015, 22, 20067–20083. [Google Scholar] [CrossRef]
  116. Velu, K.; Elumalai, D.; Hemalatha, P.; Janaki, A.; Babu, M.; Hemavathi, M.; Kaleena, P.K. Evaluation of silver nanoparticles toxicity of Arachis hypogaea peel extracts and its larvicidal activity against malaria and dengue vectors. Environ. Sci. Pollut. Res. 2015, 22, 17769–17779. [Google Scholar] [CrossRef]
  117. Kamaraj, C.; Gandhi, P.R.; Ragavendran, C.; Sugumar, V.; Kumar, R.C.S.; Ranjith, R.; Priyadharsan, A.; Cherian, T. Sustainable development through the bio-fabrication of ecofriendly ZnO nanoparticles and its approaches to toxicology and environmental protection. Biomass Convers. Biorefin. 2022, 1–17. [Google Scholar] [CrossRef]
  118. Sabzalizade, S.; Amani, A.; Vatandoost, H.; Hosseini, F.; Najafi-Taher, R.; Basseri, H.R. Evaluation of Nanoemulsion of Eucalyptus globulus Oil as Potent Botanical Larvicide against Malaria Vector, Anopheles stephensi and West Nile Vector, Culex pipiens Under Laboratory and Semi-Field Conditions. J. Arthropod. Borne Dis. 2021, 15, 380–388. [Google Scholar] [CrossRef]
  119. Govindarajan, M.; Benelli, G. Facile biosynthesis of silver nanoparticles using Barleria cristata: Mosquitocidal potential and biotoxicity on three non-target aquatic organisms. Parasitol. Res. 2016, 115, 925–935. [Google Scholar] [CrossRef] [PubMed]
  120. Suman, T.Y.; Elumalai, D.; Kaleena, P.K.; Rajasree, S.R.R. GC–MS analysis of bioactive components and synthesis of silver nanoparticle using Ammannia baccifera aerial extract and its larvicidal activity against malaria and filariasis vectors. Ind. Crops Prod. 2013, 47, 239–245. [Google Scholar] [CrossRef]
  121. Nalini, M.; Lena, M.; Sumathi, P.; Sundaravadivelan, C. Effect of phyto-synthesized silver nanoparticles on developmental stages of malaria vector, Anopheles stephensi and dengue vector, Aedes aegypti. Egypt. J. Basic Appl. Sci. 2019, 4, 212–218. [Google Scholar] [CrossRef]
  122. Arjunan, N.K.; Murugan, K.; Rejeeth, C.; Madhiyazhagan, P.; Barnard, D.R. Green synthesis of silver nanoparticles for the control of mosquito vectors of malaria, filariasis, and dengue. Vector Borne Zoonotic Dis. 2012, 12, 262–268. [Google Scholar] [CrossRef] [PubMed]
  123. Gandhi, P.R.; Jayaseelan, C.; Mary, R.R.; Mathivanan, D.; Suseem, S.R. Acaricidal, pediculicidal and larvicidal activity of synthesized ZnO nanoparticles using Momordica charantia leaf extract against blood feeding parasites. Exp. Parasitol. 2017, 181, 47–56. [Google Scholar] [CrossRef]
  124. Priyadarshini, K.A.; Murugan, K.; Panneerselvam, C.; Ponarulselvam, S.; Hwang, J.S.; Nicoletti, M. Biolarvicidal and pupicidal potential of silver nanoparticles synthesized using Euphorbia hirta against Anopheles stephensi Liston (Diptera: Culicidae). Parasitol. Res. 2012, 111, 997–1006. [Google Scholar] [CrossRef]
  125. Roni, M.; Murugan, K.; Panneerselvam, C.; Subramaniam, J.; Hwang, J.S. Evaluation of leaf aqueous extract and synthesized silver nanoparticles using Nerium oleander against Anopheles stephensi (Diptera: Culicidae). Parasitol. Res. 2013, 112, 981–990. [Google Scholar] [CrossRef]
  126. Veerakumar, K.; Govindarajan, M.; Hoti, S.L. Evaluation of plant-mediated synthesized silver nanoparticles against vector mosquitoes. Parasitol. Res. 2014, 113, 4567–4577. [Google Scholar] [CrossRef]
  127. Govindarajan, M.; Rajeswary, M.; Muthukumaran, U.; Hoti, S.L.; Khater, H.F.; Benelli, G. Single-step biosynthesis and characterization of silver nanoparticles using Zornia diphylla leaves: A potent eco-friendly tool against malaria and arbovirus vectors. J. Photochem. Photobiol. B 2016, 161, 482–489. [Google Scholar] [CrossRef] [PubMed]
  128. Govindarajan, M.; Rajeswary, M.; Hoti, S.L.; Nicoletti, M.; Benelli, G. Facile synthesis of mosquitocidal silver nanoparticles using Mussaenda glabra leaf extract: Characterisation and impact on non-target aquatic organisms. Nat. Prod. Res. 2016, 30, 2491–2494. [Google Scholar] [CrossRef] [PubMed]
  129. Govindarajan, M.; Rajeswary, M.; Veerakumar, K.; Muthukumaran, U.; Hoti, S.L.; Benelli, G. Green synthesis and characterization of silver nanoparticles fabricated using Anisomeles indica: Mosquitocidal potential against malaria, dengue and Japanese encephalitis vectors. Exp. Parasitol. 2016, 161, 40–47. [Google Scholar] [CrossRef] [PubMed]
  130. Govindarajan, M.; Vijayan, P.; Kadaikunnan, S.; Alharbi, N.S.; Benelli, G. One-pot biogenic fabrication of silver nanocrystals using Quisqualis indica: Effectiveness on malaria and Zika virus mosquito vectors, and impact on non-target aquatic organisms. J. Photochem. Photobiol. B 2016, 162, 646–655. [Google Scholar] [CrossRef] [PubMed]
  131. Govindarajan, M.; Khater, H.F.; Panneerselvam, C.; Benelli, G. One-pot fabrication of silver nanocrystals using Nicandra physalodes: A novel route for mosquito vector control with moderate toxicity on non-target water bugs. Res. Vet. Sci. 2016, 107, 95–101. [Google Scholar] [CrossRef]
  132. Muthukumaran, U.; Govindarajan, M.; Rajeswary, M.; Hoti, S.L. Synthesis and characterization of silver nanoparticles using Gmelina asiatica leaf extract against filariasis, dengue, and malaria vector mosquitoes. Parasitol. Res. 2015, 114, 1817–1827. [Google Scholar] [CrossRef]
  133. Kumar, D.; Kumar, P.; Singh, H.; Agrawal, V. Biocontrol of mosquito vectors through herbal-derived silver nanoparticles: Prospects and challenges. Environ. Sci. Pollut. Res. Int. 2020, 27, 25987–26024. [Google Scholar] [CrossRef]
  134. Foldbjerg, R.; Jiang, X.; Miclăuş, T.; Chen, C.; Autrup, H.; Beer, C. Silver nanoparticles—wolves in sheep’s clothing? Toxicol. Res. 2015, 4, 563–575. [Google Scholar] [CrossRef]
  135. Gopinath, P.; Gogoi, S.K.; Sanpui, P.; Paul, A.; Chattopadhyay, A.; Ghosh, S.S. Signaling gene cascade in silver nanoparticle induced apoptosis. Colloids Surf. B Biointerfaces 2010, 77, 240–245. [Google Scholar] [CrossRef] [PubMed]
  136. Piao, M.J.; Kang, K.A.; Lee, I.K.; Kim, H.S.; Kim, S.; Choi, J.Y.; Choi, J.; Hyun, J.W. Silver nanoparticles induce oxidative cell damage in human liver cells through inhibition of reduced glutathione and induction of mitochondria-involved apoptosis. Toxicol. Lett. 2011, 201, 92–100. [Google Scholar] [CrossRef]
  137. Almeida, J.P.; Chen, A.L.; Foster, A.; Drezek, R. In vivo biodistribution of nanoparticles. Nanomedicine 2011, 6, 815–835. [Google Scholar] [CrossRef]
  138. Najahi-Missaoui, W.; Arnold, R.D.; Cummings, B.S. Safe Nanoparticles: Are We There Yet? Int. J. Mol. Sci. 2020, 22, 385. [Google Scholar] [CrossRef]
  139. Jankowska, M.; Rogalska, J.; Wyszkowska, J.; Stankiewicz, M. Molecular Targets for Components of Essential Oils in the Insect Nervous System—A Review. Molecules 2017, 23, 34. [Google Scholar] [CrossRef] [PubMed]
  140. Nair, P.M.; Choi, J. Identification, characterization and expression profiles of Chironomus riparius glutathione S-transferase (GST) genes in response to cadmium and silver nanoparticles exposure. Aquat. Toxicol. 2011, 101, 550–560. [Google Scholar] [CrossRef]
  141. Govindarajan, M.; Benelli, G. alpha-Humulene and beta-elemene from Syzygium zeylanicum (Myrtaceae) essential oil: Highly effective and eco-friendly larvicides against Anopheles subpictus, Aedes albopictus, and Culex tritaeniorhynchus (Diptera: Culicidae). Parasitol. Res. 2016, 115, 2771–2778. [Google Scholar] [CrossRef] [PubMed]
  142. Govindarajan, M.; Rajeswary, M.; Hoti, S.L.; Benelli, G. Larvicidal potential of carvacrol and terpinen-4-ol from the essential oil of Origanum vulgare (Lamiaceae) against Anopheles stephensi, Anopheles subpictus, Culex quinquefasciatus and Culex tritaeniorhynchus (Diptera: Culicidae). Res. Vet. Sci. 2016, 104, 77–82. [Google Scholar] [CrossRef]
  143. Govindarajan, M.; Rajeswary, M.; Hoti, S.L.; Bhattacharyya, A.; Benelli, G. Eugenol, alpha-pinene and beta-caryophyllene from Plectranthus barbatus essential oil as eco-friendly larvicides against malaria, dengue and Japanese encephalitis mosquito vectors. Parasitol. Res. 2016, 115, 807–815. [Google Scholar] [CrossRef]
  144. AlShebly, M.M.; AlQahtani, F.S.; Govindarajan, M.; Gopinath, K.; Vijayan, P.; Benelli, G. Toxicity of ar-curcumene and epi-beta-bisabolol from Hedychium larsenii (Zingiberaceae) essential oil on malaria, chikungunya and St. Louis encephalitis mosquito vectors. Ecotoxicol. Environ. Saf. 2017, 137, 149–157. [Google Scholar] [CrossRef] [PubMed]
  145. Khan, Z.; Bohman, B.; Ignell, R.; Hill, S.R. Odour-mediated oviposition site selection in Aedes aegypti depends on aquatic stage and density. Parasit. Vectors 2023, 16, 264. [Google Scholar] [CrossRef]
  146. Mwingira, V.S.; Spitzen, J.; Mboera, L.E.G.; Torres-Estrada, J.L.; Takken, W. The Influence of Larval Stage and Density on Oviposition Site-Selection Behavior of the Afrotropical Malaria Mosquito Anopheles coluzzii (Diptera: Culicidae). J. Med. Entomol. 2020, 57, 657–666. [Google Scholar] [CrossRef]
  147. Mwingira, V.S.; Mboera, L.E.G.; Takken, W. Synergism between nonane and emanations from soil as cues in oviposition-site selection of natural populations of Anopheles gambiae and Culex quinquefasciatus. Malar. J. 2021, 20, 52. [Google Scholar] [CrossRef] [PubMed]
  148. Bokore, G.E.; Svenberg, L.; Tamre, R.; Onyango, P.; Bukhari, T.; Emmer, A.; Fillinger, U. Grass-like plants release general volatile cues attractive for gravid Anopheles gambiae sensu stricto mosquitoes. Parasit. Vectors 2021, 14, 552. [Google Scholar] [CrossRef] [PubMed]
  149. Jayaseelan, C.; Rahuman, A.A.; Rajakumar, G.; Santhoshkumar, T.; Kirthi, A.V.; Marimuthu, S.; Bagavan, A.; Kamaraj, C.; Zahir, A.A.; Elango, G.; et al. Efficacy of plant-mediated synthesized silver nanoparticles against hematophagous parasites. Parasitol. Res. 2012, 111, 921–933. [Google Scholar] [CrossRef]
  150. Haldar, K.M.; Haldar, B.; Chandra, G. Fabrication, characterization and mosquito larvicidal bioassay of silver nanoparticles synthesized from aqueous fruit extract of putranjiva, Drypetes roxburghii (Wall.). Parasitol. Res. 2013, 112, 1451–1459. [Google Scholar] [CrossRef]
  151. Subarani, S.; Sabhanayakam, S.; Kamaraj, C. Studies on the impact of biosynthesized silver nanoparticles (AgNPs) in relation to malaria and filariasis vector control against Anopheles stephensi Liston and Culex quinquefasciatus Say (Diptera: Culicidae). Parasitol. Res. 2013, 112, 487–499. [Google Scholar] [CrossRef]
  152. Patil, C.D.; Borase, H.P.; Patil, S.V.; Salunkhe, R.B.; Salunke, B.K. Larvicidal activity of silver nanoparticles synthesized using Pergularia daemia plant latex against Aedes aegypti and Anopheles stephensi and nontarget fish Poecillia reticulata. Parasitol. Res. 2012, 111, 555–562. [Google Scholar] [CrossRef] [PubMed]
  153. Santhoshkumar, T.; Rahuman, A.A.; Rajakumar, G.; Marimuthu, S.; Bagavan, A.; Jayaseelan, C.; Zahir, A.A.; Elango, G.; Kamaraj, C. Synthesis of silver nanoparticles using Nelumbo nucifera leaf extract and its larvicidal activity against malaria and filariasis vectors. Parasitol. Res. 2011, 108, 693–702. [Google Scholar] [CrossRef]
  154. Jayaseelan, C.; Rahuman, A.A.; Rajakumar, G.; Vishnu Kirthi, A.; Santhoshkumar, T.; Marimuthu, S.; Bagavan, A.; Kamaraj, C.; Zahir, A.A.; Elango, G. Synthesis of pediculocidal and larvicidal silver nanoparticles by leaf extract from heartleaf moonseed plant, Tinospora cordifolia Miers. Parasitol. Res. 2011, 109, 185–194. [Google Scholar] [CrossRef] [PubMed]
  155. Marimuthu, S.; Rahuman, A.A.; Rajakumar, G.; Santhoshkumar, T.; Kirthi, A.V.; Jayaseelan, C.; Bagavan, A.; Zahir, A.A.; Elango, G.; Kamaraj, C. Evaluation of green synthesized silver nanoparticles against parasites. Parasitol. Res. 2011, 108, 1541–1549. [Google Scholar] [CrossRef]
  156. Sanei-Dehkordi, A.; Ghasemian, A.; Zarenezhad, E.; Qasemi, H.; Nasiri, M.; Osanloo, M. Nanoliposomes containing three essential oils from the Artemisia genus as effective larvicides against Aedes aegypti and Anopheles stephensi. Sci. Rep. 2023, 13, 11002. [Google Scholar] [CrossRef] [PubMed]
  157. Firooziyan, S.; Osanloo, M.; Moosa-Kazemi, S.H.; Basseri, H.R.; Hajipirloo, H.M.; Sadaghianifar, A.; Amani, A.; Sedaghat, M.M. Preparation of a Nanoemulsion of Essential Oil of Acroptilon repens Plant and Evaluation of Its Larvicidal Activity agianst Malaria Vector, Anopheles stephensi. J. Arthropod. Borne Dis. 2021, 15, 333–346. [Google Scholar] [CrossRef] [PubMed]
  158. Murugan, K.; Dinesh, D.; Kumar, P.J.; Panneerselvam, C.; Subramaniam, J.; Madhiyazhagan, P.; Suresh, U.; Nicoletti, M.; Alarfaj, A.A.; Munusamy, M.A.; et al. Datura metel-synthesized silver nanoparticles magnify predation of dragonfly nymphs against the malaria vector Anopheles stephensi. Parasitol. Res. 2015, 114, 4645–4654. [Google Scholar] [CrossRef]
  159. Dinesh, D.; Murugan, K.; Madhiyazhagan, P.; Panneerselvam, C.; Mahesh Kumar, P.; Nicoletti, M.; Jiang, W.; Benelli, G.; Chandramohan, B.; Suresh, U. Mosquitocidal and antibacterial activity of green-synthesized silver nanoparticles from Aloe vera extracts: Towards an effective tool against the malaria vector Anopheles stephensi? Parasitol. Res. 2015, 114, 1519–1529. [Google Scholar] [CrossRef]
  160. Santhosh, S.B.; Yuvarajan, R.; Natarajan, D. Annona muricata leaf extract-mediated silver nanoparticles synthesis and its larvicidal potential against dengue, malaria and filariasis vector. Parasitol. Res. 2015, 114, 3087–3096. [Google Scholar] [CrossRef] [PubMed]
  161. Arokiyaraj, S.; Dinesh Kumar, V.; Elakya, V.; Kamala, T.; Park, S.K.; Ragam, M.; Saravanan, M.; Bououdina, M.; Arasu, M.V.; Kovendan, K.; et al. Biosynthesized silver nanoparticles using floral extract of Chrysanthemum indicum L.—Potential for malaria vector control. Environ. Sci. Pollut. Res. 2015, 22, 9759–9765. [Google Scholar] [CrossRef]
  162. Elumalai, D.; Hemavathi, M.; Deepaa, C.V.; Kaleena, P.K. Evaluation of phytosynthesised silver nanoparticles from leaf extracts of Leucas aspera and Hyptis suaveolens and their larvicidal activity against malaria, dengue and filariasis vectors. Parasite Epidemiol. Control 2017, 2, 15–26. [Google Scholar] [CrossRef] [PubMed]
  163. Acosta, M.E.; Gotopo, L.; Gamboa, N.; Rodrigues, J.R.; Henriques, G.C.; Cabrera, G.; Romero, A.H. Antimalarial Activity of Highly Coordinative Fused Heterocycles Targeting β-Hematin Crystallization. ACS Omega 2022, 7, 7499–7514. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram for the biosynthesis of nanoparticles (NPs) via a green route using plant extract. A plant extract is a rich mixture of various metabolites and reductive biomolecules, each playing a pivotal role in the intricate process of reducing metal ions. Biomolecules stand out for their ability to transfer electrons to metallic ions, acting as potent reducing agents that instigate the ions’ reduction. The consequence of this reduction process is the development of structured patterns in the form of a crystalline formation known as a nucleus. The initiation of nucleation involves the aggregation of monomers in supersaturated systems. This crucial step establishes the foundation for the subsequent growth of these nuclei, leading to the formation of metal nanoparticles. As these atomic nuclei form, they provide a foundational platform for the reduced ions, facilitating their gradual deposition onto the surface. This deposition process significantly contributes to the enlargement of the particles. Biomolecules, in addition to their role as reducing agents, also serve as capping agents. They bind to the particle surface, acting as molecular guardians that effectively stabilize the size of the particles.
Figure 1. Schematic diagram for the biosynthesis of nanoparticles (NPs) via a green route using plant extract. A plant extract is a rich mixture of various metabolites and reductive biomolecules, each playing a pivotal role in the intricate process of reducing metal ions. Biomolecules stand out for their ability to transfer electrons to metallic ions, acting as potent reducing agents that instigate the ions’ reduction. The consequence of this reduction process is the development of structured patterns in the form of a crystalline formation known as a nucleus. The initiation of nucleation involves the aggregation of monomers in supersaturated systems. This crucial step establishes the foundation for the subsequent growth of these nuclei, leading to the formation of metal nanoparticles. As these atomic nuclei form, they provide a foundational platform for the reduced ions, facilitating their gradual deposition onto the surface. This deposition process significantly contributes to the enlargement of the particles. Biomolecules, in addition to their role as reducing agents, also serve as capping agents. They bind to the particle surface, acting as molecular guardians that effectively stabilize the size of the particles.
Pharmaceutics 16 00699 g001
Figure 2. Action mechanisms of plant extracts and plant-derived nanoparticles. In the left panel, we summarize the phytochemicals’ effects. These molecules are recognized for their larvicidal and mosquitocidal properties, modulating the mosquito nervous system, by inducing either overstimulation or inhibition. This neuro-modulation results in observable behavioral changes, such as the reduced swimming and restless movement of larvae. In the right panel, effects of green nanoparticles (NPs) are illustrated. NPs with their optimal size effortlessly penetrate the membranes of mosquitos. Their mechanism of action involves interactions with DNA, contributing to enhanced apoptosis and oxidative stress within the target organisms. Notably, when deployed in field or semi-field conditions, these nanoparticles form an oily layer, a phenomenon correlated with a reduction in water oxygenation. Furthermore, both phytochemicals and green NPs influence factors such as the attraction of female mosquitoes to breeding sites, which are guided by olfactory cues susceptible to water quality. The effects of green-synthesized NPs may be due to the known effects of biomolecules present in plant extracts.
Figure 2. Action mechanisms of plant extracts and plant-derived nanoparticles. In the left panel, we summarize the phytochemicals’ effects. These molecules are recognized for their larvicidal and mosquitocidal properties, modulating the mosquito nervous system, by inducing either overstimulation or inhibition. This neuro-modulation results in observable behavioral changes, such as the reduced swimming and restless movement of larvae. In the right panel, effects of green nanoparticles (NPs) are illustrated. NPs with their optimal size effortlessly penetrate the membranes of mosquitos. Their mechanism of action involves interactions with DNA, contributing to enhanced apoptosis and oxidative stress within the target organisms. Notably, when deployed in field or semi-field conditions, these nanoparticles form an oily layer, a phenomenon correlated with a reduction in water oxygenation. Furthermore, both phytochemicals and green NPs influence factors such as the attraction of female mosquitoes to breeding sites, which are guided by olfactory cues susceptible to water quality. The effects of green-synthesized NPs may be due to the known effects of biomolecules present in plant extracts.
Pharmaceutics 16 00699 g002
Figure 3. Effects of green-synthesized NPs on malaria prevention. NPs synthesized with plant extracts are toxic to early developmental stages of the malaria mosquito vector, Anopheles, acting against all four larvae stages (instar) and pupae, preventing the emergence of adult mosquitos capable of disease transmission. These nanoformulations also show adulticidal effects, impacting mosquito behavior and causing morphological damage. Furthermore, some studies showed that plant-extract NPs reduced the feeding of mosquitoes by preventing the biting habit. Oviposition is hindered in water treated with the nanoformulated plant extracts, thus reducing breeding site suitability. Green-synthesized NPs also improve the feeding activity of mosquito predators, lowering the survival chances of mosquito larvae and pupae. These formulations are eco-friendly because predators are not harmed. In addition, nanotechnology emphasizes the well-known antiparasitic activity of plant extracts. The antiplasmodial effects of plant-derived nanoformulations are parasitemia suppression by increased parasite mortality and metabolic impairment. Therefore, plant-derived nanoformulations are a potential tool to fight malaria.
Figure 3. Effects of green-synthesized NPs on malaria prevention. NPs synthesized with plant extracts are toxic to early developmental stages of the malaria mosquito vector, Anopheles, acting against all four larvae stages (instar) and pupae, preventing the emergence of adult mosquitos capable of disease transmission. These nanoformulations also show adulticidal effects, impacting mosquito behavior and causing morphological damage. Furthermore, some studies showed that plant-extract NPs reduced the feeding of mosquitoes by preventing the biting habit. Oviposition is hindered in water treated with the nanoformulated plant extracts, thus reducing breeding site suitability. Green-synthesized NPs also improve the feeding activity of mosquito predators, lowering the survival chances of mosquito larvae and pupae. These formulations are eco-friendly because predators are not harmed. In addition, nanotechnology emphasizes the well-known antiparasitic activity of plant extracts. The antiplasmodial effects of plant-derived nanoformulations are parasitemia suppression by increased parasite mortality and metabolic impairment. Therefore, plant-derived nanoformulations are a potential tool to fight malaria.
Pharmaceutics 16 00699 g003
Table 2. Half-lethal concentration (LC50) of nanoformulated plant products and bulk formulations in late developmental stages (third–fourth instar and pupae) of Anopheles species after 24 h of exposure.
Table 2. Half-lethal concentration (LC50) of nanoformulated plant products and bulk formulations in late developmental stages (third–fourth instar and pupae) of Anopheles species after 24 h of exposure.
Plant SpeciesType of NanoparticleMosquito SpeciesMosquito StageLC50 of NanoformulationLC50 of Bulk ExtractReference
Eucalyptus globulusNanoemulsionA. stephensiLarvae80.8730 ppm122.8343 ppm[118]
Mangifera indicaTiO2A. subpictusLarvae7.72 mg/L49.45 mg/L[84]
Ichnocarpus frutescensAgA. subpictusLarvae14.22 μg/mL185.83 μg/mL[101]
Citrus limonAuPdA. stephensiLarvae5.12 mL/L-[98]
Rubus ellipticusAgA. stephensiLarvae12.50 μg/mL167.19 μg/mL[102]
Naregamia alataAgA. stephensiLarvae12.40 μg/mL165.15 μg/mL[103]
Alchornea cordifoliaAgA. stephensiLarvae10.67 μg/mL53.15 μg/mL[109]
Solanum xanthocarpumAgA. stephensiLarvae9.927 ppm1764.528 ppm[105]
Barleria cristataAgA. subpictusLarvae12.46 μg/mL124.27 μg/mL[119]
Malva sylvestrisAgA. stephensiLarvae10.33 μg/mL143.61 μg/mL[108]
Ammania bacciferaAgA. subpictusLarvae29.54 mg/L257.61 mg/L[120]
Artemisia annuaNanoliposomesA. stephensiLarvae90 μg/mL-[156]
Artemisia sieberi140 μg/mL
Artemisia dracunculus23 μg/mL
Artemisia nilagiricaAgA. stephensiLarvae0.141%0.224%[121]
Pupae0.050%0.066%
Valoniopsis pachynema (algae)CdSA. stephensiLarvae29.429 μg/mL334.084 μg/mL[107]
Pupae31.905 μg/mL396.868 μg/mL
A. sundaicusLarvae22.496 μg/mL296.922 μg/mL
Pupae25.009 μg/mL311.860 μg/mL
Vitex negundoZnOA. subpictusLarvae2.48 mg/L36.89 mg/L[117]
Pupae3.63 mg/L45.76 mg/L
Aristolochia indicaAgA. stephensiLarvae10.48 ppm490.31 ppm[110]
Pupae15.65 ppm565.02 ppm
Annona squamosaAgA. stephensiLarvae2.12 ppm-[122]
Pupae3.74 ppm
Ulva lactucaAgA. stephensiLarvae5.261 ppm32.692 ppm[97]
Pupae6.860 ppm37.603 ppm
Momordica charantiaZnOA. stephensiLarvae5.42 mg/L52.86 mg/L[123]
TiO2Larvae3.43 mg/L85.35 mg/L[106]
Pupae5.04 mg/L96.09 mg/L
Pteridium aquilinumAgA. stephensiLarvae18.45 ppm395.12 ppm[96]
Pupae31.51 ppm502.20 ppm
Euphorbia hirtaAgA. stephensiLarvae27.89 ppm197.40 ppm[124]
Pupae34.52 ppm219.15 ppm
Nerium oleanderAgA. stephensiLarvae33.99 ppm369.96 ppm[125]
Pupae39.55 ppm426.01 ppm
Codium tomentosum (algae)AgA. stephensiLarvae29.6 ppm410.7 ppm[114]
Pupae40.7 ppm487.1 ppm
Cymbopogon citratusAuA. stephensiLarvae31.466 ppm362.292 ppm[100]
Pupae38.327 ppm434.649 ppm
Hugonia mystaxAgA. stephensiLarvae14.45 μg/mL162.66 μg/mL[104]
Lagenaria sicerariaZnOA. stephensiLarvae56.46 ppm261.67 ppm[99]
Heliotropium indicumAgA. stephensiAdult26.712 μg/mL111.680 μg/mL[126]
Zornia diphyllaAgA. subpictusLarvae12.53 μg/mL61.23 μg/mL[127]
Mussaenda glabraAgA. subpictusLarvae17 μg/mL81 μg/mL[128]
Anisomeles indicaAgA. subpictusLarvae31.56 μg/mL108.98 μg/mL[129]
Holostemma adakodienAgA. stephensiLarvae12.18 μg/mL185.79 μg/mL[81]
Quisqualis indicaAgA. stephensiLarvae12.52 μg/mL185.98 μg/mL[130]
Nicandra physalodesAgA. stephensiLarvae12.39 μg/mL202.82 μg/mL[131]
Gmelina asiaticaAgA. stephensiLarvae22.44 μg/mL113.53 μg/mL[132]
Couroupita guianensisAuA. stephensiLarvae24.57 ppm307.72 ppm[113]
Pupae28.78 ppm363.25 ppm
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Moraes-de-Souza, I.; de Moraes, B.P.T.; Silva, A.R.; Ferrarini, S.R.; Gonçalves-de-Albuquerque, C.F. Tiny Green Army: Fighting Malaria with Plants and Nanotechnology. Pharmaceutics 2024, 16, 699. https://doi.org/10.3390/pharmaceutics16060699

AMA Style

Moraes-de-Souza I, de Moraes BPT, Silva AR, Ferrarini SR, Gonçalves-de-Albuquerque CF. Tiny Green Army: Fighting Malaria with Plants and Nanotechnology. Pharmaceutics. 2024; 16(6):699. https://doi.org/10.3390/pharmaceutics16060699

Chicago/Turabian Style

Moraes-de-Souza, Isabelle, Bianca P. T. de Moraes, Adriana R. Silva, Stela R. Ferrarini, and Cassiano F. Gonçalves-de-Albuquerque. 2024. "Tiny Green Army: Fighting Malaria with Plants and Nanotechnology" Pharmaceutics 16, no. 6: 699. https://doi.org/10.3390/pharmaceutics16060699

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop