Previous Article in Journal
Health Risk Assessment of Toluene and Formaldehyde Based on a Short-Term Exposure Scenario: A Comparison of the Reference Concentration, Reference Dose, and Minimal Risk Level
Previous Article in Special Issue
Chronic Low-Dose Phoxim Exposure Impairs Silk Production in Bombyx mori L. (Lepidoptera: Bombycidae) by Disrupting Juvenile Hormone Signaling-Mediated Fibroin Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mechanisms and Genetic Drivers of Resistance of Insect Pests to Insecticides and Approaches to Its Control

by
Yahya Al Naggar
1,*,
Nedal M. Fahmy
2,
Abeer M. Alkhaibari
2,
Rasha K. Al-Akeel
3,
Hend M. Alharbi
4,
Amr Mohamed
5,*,†,
Ioannis Eleftherianos
6,†,
Hesham R. El-Seedi
7,
John P. Giesy
8,9,10 and
Hattan A. Alharbi
11,12
1
Applied College, Center for Bee Research and Its Products, King Khalid University, Abha 61413, Saudi Arabia
2
Department of Biology, Faculty of Science, University of Tabuk, Tabuk 71491, Saudi Arabia
3
Department of Zoology, Faculty of Science, King Saud University, Riyadh 11451, Saudi Arabia
4
Department of Biology, College of Science, Princess Nourah bint Abdulrahman University, Riyadh 11671, Saudi Arabia
5
Department of Entomology, Faculty of Science, Cairo University, Giza 12613, Egypt
6
Department of Biological Sciences, The George Washington University, Washington, DC 20052, USA
7
Department of Chemistry, Faculty of Science, Islamic University of Madinah, Madinah 42351, Saudi Arabia
8
Department of Environmental Science, Baylor University, Waco, TX 76798-7266, USA
9
Department of Veterinary Biomedical Sciences and Toxicology Centre, Western College of Veterinary Medicine, University of Saskatchewan, Saskatoon, SK S7N 5B3, Canada
10
Department of Integrative Biology and Center for Integrative Toxicology, Michigan State University, East Lansing, MI 48824, USA
11
Department of Plant Protection, College of Food and Agriculture Sciences, King Saud University, Riyadh 11451, Saudi Arabia
12
Chair of Date Palm Research, College of Food and Agriculture Sciences, King Saud University, Riyadh 11451, Saudi Arabia
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Toxics 2025, 13(8), 681; https://doi.org/10.3390/toxics13080681 (registering DOI)
Submission received: 28 June 2025 / Revised: 7 August 2025 / Accepted: 14 August 2025 / Published: 16 August 2025
(This article belongs to the Special Issue Impacts of Agrochemicals on Insects and Soil Organisms)

Abstract

The escalating challenge of resistance to insecticides among agricultural and public health pests poses a significant threat to global food security and vector-borne disease control. This review synthesizes current understanding of the molecular mechanisms underpinning resistance, including well-characterized pathways such as target-site mutations affecting nicotinic acetylcholine receptors (nAChRs), acetylcholinesterase (AChE), voltage-gated sodium channels (VGSCs), and γ-aminobutyric acid (GABA) receptors, and metabolic detoxification mediated by cytochrome P450 monooxygenases (CYPs), esterases, and glutathione S-transferases (GSTs). Emerging resistance mechanisms are also explored, including protein sequestration by odorant-binding proteins and post-transcriptional regulation via non-coding RNAs, such as microRNAs (miRNAs) and long non-coding RNAs (lncRNAs). Focused case studies on Aedes aegypti and Spodoptera frugiperda illustrate the complex interplay of genetic and biochemical adaptations driving resistance. In Ae. aegypti, voltage-gated sodium channel (VGSCs) mutations (V410L, V1016I, F1534C) combined with metabolic enzyme amplification confer resistance to pyrethroids, accompanied by notable fitness costs and ecological impacts on vector populations. In S. frugiperda, multiple resistance mechanisms, including overexpression of cytochrome P450 genes (e.g., CYP6AE43, CYP321A8), target-site mutations in ryanodine receptors (e.g., I4790K), and behavioral avoidance, have rapidly evolved across global populations, undermining the efficacy of diamide, organophosphate, and pyrethroid insecticides. The review further evaluates integrated pest management (IPM) strategies, emphasizing the role of biopesticides, biological control agents, including entomopathogenic fungi and parasitoids, and molecular diagnostics for resistance management. Taken together, this analysis underscores the urgent need for continuous molecular surveillance, the development of resistance-breaking technologies, and the implementation of sustainable, multifaceted interventions to safeguard the long-term efficacy of insecticides in both agricultural and public health contexts.

1. Introduction

The ongoing arms race between humans and insect pests remains heavily dependent on synthetic insecticides, despite mounting resistance and growing concerns about their impacts on human health and biodiversity [1,2]. The widespread and often indiscriminate use of these chemicals has accelerated resistance, posing serious threats to ecosystems, public health, and agriculture. Insect pests not only inflict extensive damage on crops, forests, stored products, and animal feed, but also serve as vectors for diseases caused by viruses, bacteria, and parasites [1,2,3,4,5]. Their impact extends beyond direct yield loss, posing serious threats to food security, disrupting ecological balance, and imposing substantial economic burdens—especially in low- and middle-income countries. Global assessments estimate yield losses of approximately 21.5% for wheat, 30% for rice, 22.5% for maize, 17% for potatoes, and 21% for soybeans, with the highest losses concentrated in food-insecure regions [3]. Moreover, the economic cost of invasive insect pests alone exceeds US$70 billion annually worldwide, with associated health impacts surpassing US$6.9 billion [6]. Since its invasion of Africa in 2016, the fall armyworm (Spodoptera frugiperda) has spread to over 70 countries, causing annual maize yield losses of 8.3–20.6 million tons and economic damages of US$2.5–6.2 billion [7,8].
Globally, about 3.7 million tons of pesticides were applied in agriculture in 2022, with insecticides accounting for roughly 29.5% [9]. China alone used an estimated 1.4 million tons of insecticides, according to FAOSTAT data [9,10]. These pesticides primarily target insect physiological systems, including the nervous system and energy metabolism. While chemical control has driven significant productivity gains since the Green Revolution, it has also accelerated the evolution of resistance in many pest species.
While organophosphates, pyrethroids, and neonicotinoids have historically exhibited strong insecticidal efficacy, over 600 arthropod species have developed resistance to these compounds, with thousands of rigorously documented cases worldwide. Resistance to insecticides occurs when pests inherit reduced sensitivity, causing pesticides to fail repeatedly even when applied as directed [5]. The Insecticide Resistance Action Committee (IRAC) describes resistance as “a heritable change in the sensitivity of a pest population that is reflected in the repeated failure of a product to achieve the expected level of control when used according to the label recommendation for that pest species” [11]. Resistance not only leads to failure in controlling insect pests but also necessitates higher dosages and the use of newer—often more toxic—chemical agents, thereby compounding environmental and health concerns [12]. Figure 1 illustrates the process of resistance development.
Resistance is driven by various factors, including genetic variability, high reproductive rates, and operational practices such as repeated use of similar chemical classes [13]. These factors facilitate rapid selection of resistant phenotypes and increase the likelihood of resistance allele fixation within populations. Furthermore, globalization and trade have enhanced the spread of resistant pests across borders, turning localized problems into global challenges. This relentless selection pressure underscores resistance as an “evolutionary arms race” demanding disruptive solutions. Traditional chemical-centric approaches are outpaced by pest adaptation, necessitating paradigm-shifting strategies grounded in evolutionary theory—such as evolutionary rescue interventions or gene drives targeting resistance alleles—to break the cycle of resistance escalation [14].
These evolutionary pressures give rise to two major categories of resistance mechanisms: (1) target-site resistance, involving structural mutations at insecticide binding sites (e.g., [15]); and (2) metabolic resistance, characterized by enzymatic degradation of insecticides before they reach their site [5]. Other less common but increasingly recognized mechanisms include behavioral resistance, where pests modify their actions to avoid exposure, and cuticular resistance, involving changes in the insect cuticle that reduce insecticide penetration. Understanding these diverse mechanisms is critical for designing countermeasures and delaying the onset of resistance.
Considering the growing challenge of resistance to insecticides, there is an urgent need to review and evaluate existing pest management paradigms. Reliance on chemical controls alone is no longer tenable. A shift toward integrated pest management (IPM), which combines biological control agents, cultural practices, resistant crop varieties, and judicious use of chemicals, is imperative. Such an approach not only ensures long-term sustainability but also mitigates the adverse effects of chemical overuse on ecosystems and human health [16].
This review provides an examination of the molecular and biochemical foundations of resistance to insecticides, drawing attention to key genes and pathways involved. It also evaluates the limitations of conventional chemical control and explores integrated alternatives aimed at sustainable pest suppression. By consolidating current knowledge and identifying emerging trends, this review aims to support the development of informed, adaptable, and ecologically sound pest management strategies.

2. Key Mechanisms Underlying Insecticide Resistance

Resistance to insecticides arises through four main mechanisms. Behavioral resistance involves altered behaviors like outdoor resting or oviposition avoidance, offering low to moderate resistance but delaying physiological resistance. Penetration resistance results from thickened or modified cuticles, reducing insecticide uptake and providing moderate, cross-class protection. Target-site insensitivity includes mutations at binding sites (e.g., kdr, Ace-1, Rdl) and leads to high, class-specific resistance and rapid control failure. Metabolic resistance, the most potent, is driven by overexpression or amplification of detoxifying enzymes (P450s, esterases, GSTs), resulting in high, broad-spectrum resistance and cross-resistance to multiple insecticide classes [17,18] (Table 1). These mechanisms often act in combination.

2.1. Molecular Mechanisms of Target-Site Resistance

Target-site resistance arises when specific mutations alter the molecular conformation of an insecticide’s binding site, diminishing or abolishing its affinity. This prevents the compound from exerting its intended neurotoxic or biochemical effect. Resistance mechanisms can be classified into four broad categories based on the nature of the target site: (1) neurological targets (e.g., AChE, nAChR, sodium channel, chloride channel, etc.); (2) physiological and metabolic targets (e.g., chitin synthase, ryanodine receptor, mitochondrial respiratory chain, etc.); (3) growth and developmental targets (e.g., juvenile hormone receptors, ecdysteroid receptors, etc.); and (4) other novel or rare targets (e.g., octopamine receptor, midgut cell membranes, oxidative phosphorylation and decoupling agents, etc.).

2.1.1. Neurological Targets

Nicotinic Acetylcholine Receptor (nAChR) Mutations
Neonicotinoids, acting as high-affinity agonists of insect nAChRs, have become one of the most widely used classes of insecticides globally due to their systemic properties, potency, and relatively favorable mammalian toxicity profile [31]. In insects, nAChRs are pentameric ligand-gated ion channels composed of various combinations of α- and β-subunits. The neonicotinoid-binding site is formed at the subunit interface, involving loops A–C of the α-subunit and loops D–E of the β-subunit. Among these, residues in loop D of the β1 subunit play a critical role in neonicotinoid recognition and sensitivity [32].
Target-site resistance has emerged in several pest species through point mutations that impair neonicotinoid binding without disrupting acetylcholine-mediated neurotransmission. In Myzus persicae, two such substitutions in the β1 subunit—R81T and V101I—markedly reduce neonicotinoid binding affinity while maintaining receptor function [15,33]. These mutations confer high-level resistance to multiple compounds, including imidacloprid, thiamethoxam, and dinotefuran. All tested homozygous R81T clones (81TT) exhibited markedly higher resistance to imidacloprid and thiacloprid compared to other genotypes, while heterozygous clones (81RT) displayed slightly elevated but overlapping resistance levels with wild-type clones. This pattern is consistent with semi-recessive inheritance, with mean dominance coefficients of 0.316 for imidacloprid and 0.351 for thiacloprid [34]. While the R81T mutation alone confers substantial resistance, bioassays demonstrate that co-occurrence with elevated CYP6CY3 expression further amplifies resistance via synergistic interaction between target-site and metabolic mechanisms [35].
Radioligand binding assays revealed a complete loss of the high-affinity [3H]-imidacloprid binding site and altered low-affinity interactions in the R81T variant, leading to a marked reduction in overall binding affinity. This effect is likely attributable to the disruption of key electrostatic interactions at the R81 residue [15]. Computational modelling supports this mechanism, demonstrating a substantial decrease in binding affinity associated with the R81T substitution [36]. In thrips, mutations V65I and V104I in the β1 subunit of nAChRs contribute to neonicotinoid resistance by reducing receptor binding to neonicotinoids and sulfoxaflor (V65I) and by increasing neonicotinoid efficacy while decreasing agonist affinity (V104I) [37]. Similarly, in Bemisia tabaci, dual mutations (A58T&R79E) in the BTβ1 nAChR subunit confer target-site resistance to neonicotinoids, with molecular modeling suggesting a mechanism driven by electrostatic repulsion at the orthosteric site [38]. In Musca domestica, reduced expression of the nAChR α2 subunit has been implicated in neonicotinoid resistance, particularly when coupled with enhanced detoxification via CYPs [39]. In Thrips palmi, the nAChR α1 subunit serves as the primary target of neonicotinoids, exhibiting higher agonist affinity and efficacy compared to the α2 subunit [40]. Collectively, these studies underscore the importance of nAChR subunit variation in neonicotinoid resistance, albeit through distinct molecular mechanisms.
Spinosad, a macrocyclic lactone insecticide, targets a distinct nAChR subunit—α6 (Dα6)—and exhibits a different mode of action from neonicotinoids. In Drosophila melanogaster, a CRISPR/Cas9-engineered G275E substitution in Dα6 produces a ~66-fold decrease in spinosad sensitivity, confirming the causality of this mutation even in the absence of changes in other subunits [41]. Resistance to spinosad tends to be recessively inherited and polygenic, contrasting with the more dominant neonicotinoid resistance conferred by β1 mutations.
Sulfoxaflor, a sulfoximine insecticide, acts as an nAChR agonist but targets a different receptor subtype from neonicotinoids. Its mode of action involves sustained activation of nAChRs, leading to hyperexcitation, paralysis, and eventual death. Despite its chemical distinctiveness, sulfoxaflor remains toxic to non-target pollinators [42]. Resistance has been documented in multiple pest species—including Aphis gossypii, Nilaparvata lugens, and B. tabaci—and is primarily associated with overexpression of CYPs, although target-site mutations have also been implicated [43,44,45]. Partial cross-resistance with neonicotinoids has been observed, emphasizing the need for integrated resistance management.
Field studies of 35 M. persicae colonies across Europe revealed a strong correlation between resistance ratios to sulfoxaflor and imidacloprid (Pearson’s r = 0.939, p < 0.0001). However, this correlation was absent in Spanish colonies carrying the R81T mutation (r = 0.2901, p = 0.3604), indicating that R81T alone does not confer altered sulfoxaflor susceptibility [46]. Notably, synergist bioassays using PBO significantly reduced—but did not eliminate—resistance in a single field-collected R81T-harboring clone (FRC), with residual resistance factors of 234-fold for imidacloprid and 26-fold for thiamethoxam following PBO treatment. These results demonstrate that the R81T mutation alone confers substantial resistance, independent of metabolic detoxification pathways [15].
The FRC clone, collected from southern France in 2009, was homozygous for R81T and exhibited concurrent CYP6CY3 gene amplification [15]. Experimental studies showed that sustained absence of neonicotinoid selection did not alter nAChR protein levels in the R81T-harboring M. persicae FRC strain [47]. Inheritance assays confirm the semi-recessive nature of R81T, with dominance coefficients of 0.316 (imidacloprid) and 0.351 (thiacloprid). This implies slower allele fixation rates in the field, even under strong selection pressure [34].
Collectively, these findings demonstrate that specific mutations in the β1 and Dα6 subunits of nAChRs act as dominant or semi-dominant resistance mechanisms by structurally disrupting the insecticide binding interface. Table 2 provides an overview of recent nAChR mutations linked to resistance or susceptibility in major insect pests. Although such mutations often co-occur with metabolic resistance (e.g., CYP6CY3 overexpression), other factors—such as α2 subunit downregulation in M. domestica and the recessive inheritance of spinosad resistance—introduce substantial heterogeneity in resistance dynamics and fitness trade-offs across species.
Future research should prioritize resolving the precise pentameric stoichiometry of α-β subunit assemblies that determine high- versus low-affinity neonicotinoid binding, as subunit composition may significantly influence receptor sensitivity. The development of high-resolution homology models or cryo-electron microscopy (cryo-EM) structures of mutated nAChRs is essential for quantifying conformational changes in ligand binding and channel gating. Additionally, ecological studies are needed to evaluate the fitness costs associated with R81T and Dα6 mutations under realistic conditions, including interspecific competition, environmental stress, and cyclic insecticide exposure. Finally, elucidating the epistatic interactions and genetic background effects—such as those involving cytochrome P450 haplotypes or nAChR gene duplications—will be critical for predicting the spread and stability of resistance in field populations.
Voltage-Gated Sodium Channel Alterations (Knockdown Resistance, kdr)
Pyrethroids and DDT target VGSCs, essential for action potential propagation in insect neurons. Resistance is conferred through mutations in the VGSC gene, commonly known as vssc1, particularly at position 1014 in domain II, where leucine is substituted with phenylalanine (L1014F) [62]. This mutation disrupts insecticide binding and prolongs channel opening, reducing neuronal excitability. Kdr-type mutations have been reported across multiple species, including M. domestica, Cx. quinquefasciatus, and L. decemlineata [63,64]. Variants kdr and super-kdr are differentiated, the latter involving additional amino acid substitutions and higher resistance levels [65]. Additionally, oxadiazine insecticides, such as indoxacarb, also target sodium channels. Resistance to indoxacarb has been reported in several insect species, with mutations in sodium channel genes identified as responsible for reduced binding affinity [66]. For instance, the presence of the F1845Y and V1848I mutations in the VGSCs of T. absoluta was strongly associated with indoxacarb resistance, closely correlating with the resistant phenotype [67].

2.1.2. Physiological and Metabolic Targets

Aminobenzoic Acid Amide Insecticides (Ryanodine Receptor)
Aminobenzoic acid amide insecticides, commonly known as diamides—such as flubendiamide, chlorantraniliprole, tetrachlorantraniliprole, and cyantraniliprole—act by targeting the ryanodine receptor (RyR) in insects, disrupting calcium ion release in muscle cells and leading to paralysis and death. These compounds, including broflanilide and tetraniliprole, have shown consistently high field efficacy against S. frugiperda larvae and effectively suppressed S. exigua populations shortly after application [68,69]. Nonetheless, the development of resistance is a growing challenge. Alterations in the RyR gene have been associated with decreased binding affinity of these insecticides, resulting in reduced effectiveness, as reported in P. xylostella [70]. Moreover, research indicates that various chemical categories—such as ryanodine, phthalamides, and anthranilamides—interact with different sites on the insect RyR [71]. Computational docking studies have also identified species-specific differences in the chlorantraniliprole binding site across pests, including Sogatella furcifera, Leptinotarsa decemlineata, and Helicoverpa assulta [72]. These insights emphasize the importance of thoroughly understanding RyR binding mechanisms, particularly in pests like S. frugiperda, to elucidate resistance pathways and support the development of sustainable insecticide strategies [72].
Benzoylurea Insecticides (Chitin Synthase)
Benzoylurea insecticides (BUs), which target chitin synthase—a key enzyme in insect exoskeleton formation—remain widely used in IPM. Resistance to BUs in arthropods is primarily driven by nonsynonymous mutations in the chitin synthase 1 gene (CHS1; referred to as CHSA in Lepidoptera), which encodes the catalytic subunit essential for chitin biosynthesis during molting. These mutations localize to transmembrane helices (TMH) 5–7 of CHS1, where conserved isoleucine residues are substituted to induce steric hindrance, blocking insecticide binding while preserving enzymatic function. Examples include I1040M in S. frugiperda CHSA, which confers resistance to lufenuron [73]; I1042M in P. xylostella CHS1, which causes cross-resistance to diflubenzuron and flufenoxuron [74]; homologous mutations such as I1017F in Tetranychus urticae and F. occidentalis [74,75]; and I1043F/L/M in Culex pipiens, which reduces diflubenzuron efficacy [76]. Emerging mutations, such as I1040T/V in S. frugiperda, further reduce BU binding affinity [56], demonstrating that convergent substitutions (I → M/F/L/T/V) represent a dominant evolutionary mechanism. Secondary metabolic resistance occurs via glutathione S-transferase (GST) overexpression, such as PxGSTs1 in P. xylostella, which enhances BU detoxification by glutathione conjugation mediated by catalytic residues (Ser65, Tyr97), enabling π-π stacking interactions [77].

2.1.3. Growth and Developmental Targets

Juvenile Hormone Receptors and Ecdysteroid Receptors
Insects rely on hormonal signals like juvenile hormone (JH) and ecdysteroids for normal growth and development [78]. Juvenile hormone analogs (e.g., methoprene) and ecdysteroid inhibitors (e.g., tebufenozide) are commonly used insecticides. Resistance to insecticides by juvenile hormone receptors (JHRs) and ecdysteroid receptors (EcRs) involves complex genetic adaptations, including mutations, gene overexpression, and epigenetic modifications. These adaptations enable insects to evade the effects of insect growth regulators (IGRs) and other endocrine-disrupting insecticides. For JHRs, such as Methoprene-tolerant (Met), resistance mechanisms include target-site mutations like Y129F, which impair juvenile hormone binding [79,80], gene overexpression [81], and functional redundancy between Met and its paralog germ cell-expressed (Gce) [82,83]. Additionally, post-transcriptional regulation by non-coding RNAs and RNA methylation further modulates resistance [84,85,86]. In EcRs, mutations in the ligand-binding domain [87,88,89] and alterations in receptor complex interactions, such as JH-mediated competition for Ultraspiracle (USP), disrupt 20-hydroxyecdysone (20E) signaling. This disruption affects molting and detoxification pathways [90,91,92]. Additionally, cross-talk between JH and EcR pathways modulates resistance, with JH-Met complexes inhibiting EcR dimerization and recruiting co-repressors to suppress metamorphosis [93,94,95].

2.1.4. Other Novel or Rare Targets

Octopamine Receptors
In insects, octopamine exerts pleiotropic effects as a neurotransmitter, neuromodulator, and hormone, regulating behavior, physiology, and immunity [96,97]. Its receptors have become key targets for insecticides due to their central physiological role [98]. Resistance to octopamine receptor antagonists can arise through mutations in the receptor gene, presenting challenges for control strategies. Octopaminergic signaling is disrupted not only by synthetic antagonists but also by certain essential oils. These oils—such as eugenol, γ-terpineol, and cinnamic alcohol—block octopamine receptors, causing acute and sublethal behavioral effects in insects. Normally, octopamine binding elevates cyclic AMP (cAMP) levels, but this signaling is inhibited by these essential oil compounds. Notably, eugenol alone significantly reduces receptor binding even at low doses [99,100]. Conversely, octopamine receptor (OR) agonists—including formamidine pesticides like amitraz and chlordimeform—exert their effects by inhibiting monoamine oxidase and disrupting cAMP production, which leads to behavioral changes. These pesticides also directly bind to octopamine receptors as OR agonists, underscoring their complex biochemical and molecular mechanisms. Furthermore, OR agonists show synergistic effects when combined with other insecticide classes such as organophosphates, pyrethroids, and neonicotinoids, enhancing pest control efficacy. Consequently, they are important components of IPM programs [101,102,103,104].
Midgut Cell Membranes and Oxidative Phosphorylation/Decoupling Agents
Insecticides targeting midgut cell membranes or oxidative phosphorylation pathways are still being explored, with resistance often linked to mutations in ATP synthase subunits or increased expression of decoupling proteins. For example, periplocoside P (PSP) from Periploca sepium selectively depolarizes the apical membrane potential (Va) in Mythimna separata larvae by inhibiting V-type H+-ATPase, disrupting midgut ion transport similarly to Cry1Ab toxin but with a faster onset [105]. Likewise, the acetogenin squamocin from Annona mucosa uncouples oxidative phosphorylation in Ae. aegypti by inhibiting mitochondrial complex I (NADH-ubiquinone oxidoreductase) while also suppressing V-ATPase and autophagy-related genes (Atg1, Atg8), demonstrating multi-modal toxicity [106]. Resistance to these compounds may involve multiple mechanisms, including potential mutations in ATP synthase subunits (e.g., V-ATPase) or the upregulation of mitochondrial uncoupling proteins that reduce the proton gradient. While direct evidence in insects—particularly in midgut tissues—is still limited, similar uncoupling effects have been reported for certain insecticides and other bioactive agents, suggesting that mitochondrial dysfunction could contribute to resistance development [107,108,109,110,111].

2.1.5. GABA Receptor Mutations and Resistance to Dieldrin (Rdl)

GABA-gated chloride channels are essential for inhibitory neurotransmission in insects, and mutations affecting their subunits represent critical mechanisms of resistance to insecticides. The Rdl gene encodes one such subunit, and substitutions at position A302—most notably A302S and A302G—have been strongly linked to resistance against cyclodienes and phenylpyrazoles [112,113]. Ala301 mutations in Rdl confer moderate resistance to phenylpyrazoles in D. melanogaster. Gly301 provides better survival than Ser301. Gene duplication and mutations in D. simulans also contribute to resistance [112]. These mutations are believed to disrupt GABA binding or reduce chloride ion conductance, thereby impairing the action of GABA-targeting insecticides [114].
In D. melanogaster, the A2′N mutation in the RDL GABA receptor confers substantial in vivo resistance to fipronil, likely through steric hindrance from bulkier side chains at the mutation site, which interfere with insecticide binding—even in heterozygous individuals [115]. Additionally, the A301S variant not only mediates resistance to cyclodienes and phenylpyrazoles but also induces neurophysiological compensation through the cholinergic system. This includes upregulation of AChE expression and activity, leading to cross-resistance against carbamates (e.g., propoxur) and organophosphates (e.g., dichlorvos), as well as altered expression of muscarinic acetylcholine receptors that confer resistance to non-selective muscarinic compounds such as pilocarpine and atropine [116]. Homology modeling and ligand-docking study predict that the A301S substitution narrows the RDL channel pore and significantly reduces the binding free energy of dieldrin and fipronil, providing atomistic evidence for a steric exclusion mechanism [116]. In An. gambiae, a T345S/M substitution co-occurs with the codon 296 allele and partially restores normal channel gating kinetics—suggesting that secondary mutations in transmembrane regions may compensate for Rdl-mediated dysfunction, a possibility yet to be examined in D. melanogaster [117].
In A301S-mutant Drosophila, the GABAergic toxicity of fipronil and dieldrin is less effectively synergized by pilocarpine, unlike in susceptible strains, suggesting altered muscarinic modulation of GABAergic pathways [118]. Furthermore, analogies to ace-1 gene duplications in Anopheles, which buffer the fitness costs of resistance alleles, raise the possibility that structural variants—such as tandem duplications or heterogeneous Rdl alleles—may similarly stabilize the A301S genotype in Drosophila [119].
Despite extensive phenotypic evidence of persistence, comprehensive life-history studies (e.g., on larval development, adult longevity, fecundity, and competitive fitness) remain unpublished for the MD-RR strain (homozygous for the Rdl A301S resistance mutation). Such studies are crucial for understanding why the A301S mutation persists without reversion, contrasting with rapid genotype reversion observed in other species following relaxation of insecticide pressure [120].
Critically, resistance mechanisms rarely act in isolation. Epistatic interactions and compensatory adaptations—such as CYPs upregulation, mitigating the fitness costs of kdr alleles—contribute to polygenic resistance and adaptive resilience [121,122,123]. Future research should prioritize disentangling mechanistic hierarchies, distinguishing primary resistance drivers from secondary modifiers, to effectively disrupt resistance networks and delay further evolution of multi-class resistance.

2.2. Mechanisms of Metabolic Resistance: Enzymatic Detoxification Pathways

Insects frequently overcome insecticidal exposure not only through structural mutations at target sites but also via enhanced detoxification systems. Metabolic resistance is primarily conferred through the upregulation, gene amplification, or structural modification of key enzymes that neutralize or degrade insecticides before they reach their intended molecular targets. The detoxification process is now widely understood to involve three integrated phases: Phase I (oxidation, hydrolysis, or reduction), Phase II (conjugation of metabolites for excretion), and Phase III (efflux transport and/or sequestration). The major enzyme families implicated in metabolic resistance include Phase I enzymes such as esterases and CYPs, Phase II enzymes including glutathione-S-transferases (GSTs) and UDP-glucosyltransferases (UGTs), and Phase III transporters such as ATP-binding cassette (ABC) proteins [124,125,126].

2.2.1. Esterases: Gene Amplification and Detoxification Efficiency

Esterases (particularly CESs) are integral to Phase I detoxification. They cleave ester bonds in insecticides such as pyrethroids, carbamates, organophosphates, and neonicotinoids, diminishing their toxicity before neuronal interference [124]. Esterase-based resistance in M. persicae is linked to two amplified carboxylesterase genes, E4 and FE4. The E4 form, associated with an A1–3 chromosomal translocation, is globally distributed, while FE4, typically found in the Mediterranean, lacks such rearrangements [127]. Resistant aphid clones can exhibit 5 to 11 times higher esterase gene copy numbers compared to susceptible populations, leading to significant degradation of insecticides before they reach their target. However, increased gene copy number or transcript abundance alone does not necessarily confirm a detoxification role.
Current approaches to verify enzymatic activity include in vitro metabolism assays using recombinant esterases, susceptibility shifts following RNA interference (RNAi) knockdown, and identification of metabolic products. For example, recombinant carboxylesterase E4 from M. persicae expressed in Escherichia coli hydrolyzed the carbamate insecticide carbaryl by 64% within 2.5 h and the organophosphate malathion by 80% within 1.25 h, directly confirming its catalytic detoxification capability [128]. Similarly, RNAi silencing of LbEST1, LbEST2, and LbEST3 in the psocid Liposcelis bostrychophila increased mortality by 1.83–2.69-fold following malathion exposure, demonstrating in vivo detoxification roles through phenotypic reversion assays [129]. Furthermore, RNAi-mediated silencing of the carboxylesterase gene Deg-CarE in Dermanyssus gallinae significantly reduced esterase activity and restored susceptibility to β-cypermethrin in resistant mites, confirming Deg-CarE’s critical role in detoxification and resistance [130].

2.2.2. Cytochrome P450 Monooxygenases: The Versatile Detoxifiers

Cytochrome P450 enzymes (CYPs) constitute a large, evolutionarily conserved family capable of metabolizing endogenous hormones, xenobiotics, and plant secondary compounds [131,132]. These monooxygenases catalyze oxidation reactions, including N-dealkylation, hydroxylation, desulfuration, and epoxidation, forming the biochemical cornerstone of insect metabolic resistance. In M. persicae, P450 enzymes (CYP6CY3 and CYP6CY4) play a key role in metabolic resistance. Elevated P450 expression correlates with resistance up to a threshold. The homozygous R81T mutation, combined with P450 overexpression, confers high neonicotinoid resistance. A synergistic interaction between P450s and R81T enhances resistance, though resistance variation across different R81T genotypes is only partially explained by P450 levels, suggesting other factors are involved [35].
In M. domestica, cytochrome P450 subunits CYP6A1, CYP6D1, and CYP6D3 exhibit sex-specific overexpression, conferring resistance to multiple insecticide classes [133]. Additionally, D. melanogaster P450 variants like CYP6w1 and CG10737 mediate DDT resistance via similar overexpression patterns [134]. Synergists such as piperonyl butoxide (PBO) inhibit CYP enzymes, restoring insecticide sensitivity and emphasizing their crucial role in the evolution of resistance [135]. Beyond transcriptomic associations, the metabolism of insecticides by specific P450s has been confirmed using in vitro expression systems. For example, recombinant expression of CYP enzymes has demonstrated direct metabolism of lambda-cyhalothrin into less toxic hydroxylated metabolites [136], further affirming their functional role in resistance.

2.2.3. Glutathione-S-Transferases (GSTs): Conjugation- and Sequestration-Based Resistance

GSTs are multifunctional enzymes catalyzing the conjugation of reduced glutathione (GSH) to electrophilic xenobiotics, facilitating their solubilization and excretion. These enzymes also contribute to oxidative stress mitigation, intracellular transport, and hormone biosynthesis [137]. In insects, GSTs are classified into three major families: Cytosolic GSTs (e.g., Delta, Epsilon, Omega, Sigma), Microsomal GSTs (also called MAPEGs), and Mitochondrial GSTs (Kappa, typically absent in insects) [138]. In S. litura, GSTe16 is crucial for pyrethroid detoxification and antioxidant defense. Knockdown and CRISPR/Cas9 validation confirm its role in resistance. Molecular analysis identifies key residues (Arg111, Asn122) for substrate specificity, indicating an evolutionary link between xenobiotic metabolism and oxidative stress response [139]. Allelic variation in epsilon GST genes contributes to DDT and pyrethroid resistance in An. funestus. Transcriptomic and genomic analyses reveal region-specific variants—such as A17D26T158-GSTe3, L135H191A189-GSTe4, and T169S201E210-GSTe6—that exhibit increased binding affinity for DDT and permethrin. Functional assays showed recombinant GSTs metabolize DDT (41–63%) and permethrin (13–25%), with the BN-GSTe4 variant demonstrating significantly enhanced deltamethrin metabolism [140]. Transgenic expression in Drosophila confirms reduced DDT mortality in flies with selected alleles, indicating a key role in resistance [140]. These enzymes operate primarily in Phase II detoxification, neutralizing oxidized metabolites produced by CYPs and esterases. Their upregulation is often driven by gene duplication, promoter mutations, or stress-induced expression networks [137,138]. Recent findings indicate that certain GST isoforms possess high-affinity hydrophobic binding pockets that enable sequestration of insecticides such as lambda-cyhalothrin, thereby reducing their bioavailability and interaction with neural targets. This detoxification-by-sequestration mechanism—distinct from enzymatic conjugation—has been experimentally confirmed in Cydia pomonella as a key factor contributing to resistance, where exposure to sublethal doses of lambda-cyhalothrin upregulated 17 GST genes, and recombinant CpGSTd3 exhibited the highest binding activity despite no detectable metabolites [141].

2.2.4. UDP-Glucosyltransferases (UGTs): An Overlooked Phase II Component

UDP-glucosyltransferases (UGTs) are increasingly recognized as critical Phase II detoxification enzymes in insects. They conjugate UDP-sugars to hydrophobic xenobiotics, enhancing their water solubility for excretion. Recent studies have identified overexpressed UGTs in resistant populations of B. tabaci, H. armigera, S. litura, Anopheles funestus, and parasitoid wasps, with direct involvement in pyrethroid and neonicotinoid resistance [142,143,144,145,146]. Functional assays using recombinant UGTs and RNAi-based knockdown approaches further confirm their metabolic roles.

2.2.5. ABC Transporters: Phase III Toxin Efflux Systems

Phase III detoxification involves the active export of both conjugated and unmodified toxins from insect cells, primarily mediated by ATP-binding cassette (ABC) transporters [147]. These transporters play a critical role in regulating toxin bioavailability by facilitating efflux from midgut epithelial cells, Malpighian tubules, and other tissues. Members of the ABC transporter family, such as ABCB1, ABCG4, and ABCC2, have been implicated in resistance to various insecticides, including spinosad, pyrethroids, and Bacillus thuringiensis (Bt) toxins [126]. Their overexpression is often induced by sublethal insecticide exposure and is associated with enhanced excretory capacity, forming an important defense mechanism against toxin accumulation [147]. Experimental evidence supports the functional role of these transporters in resistance to insecticides. In A. gossypii, RNAi targeting ABCC2 or ABCG15 significantly reduced thiamethoxam efflux and increased insecticide susceptibility by 2.3–3.5-fold [73]. Similarly, in T. urticae, abamectin resistance was reversed using beauvericin (BEA), a known ABC transporter inhibitor. Co-application of BEA with abamectin led to a 4.4–7.7-fold reduction in LC50 values compared to abamectin alone. Molecular docking studies further confirmed BEA’s high affinity for ABCC and ABCG substrate-binding sites, indicating that it competitively inhibits xenobiotic transport [148].

3. Emerging and Underexplored Mechanisms of Resistance of Insects to Insecticides

Recent advances have revealed novel classes of resistance to insecticides that extend beyond conventional metabolic and target-site models. Two such mechanisms, protein-based sequestration and RNA-mediated posttranscriptional regulation, are reshaping understanding of insect adaptation under chemical pressure.

3.1. Sequestration Resistance: Repurposing Olfactory Proteins as Insecticide Buffers

The concept of sequestration resistance is that insects co-opt chemosensory proteins to sequester insecticides, reducing their bioavailability and toxicity. Traditionally recognized for their role in the detection of odors, odorant-binding proteins (OBPs) and chemosensory proteins (CSPs) have been co-opted to function as molecular “sponges” for xenobiotics. Overexpression of these proteins in cuticular tissues, particularly the legs, creates localized sequestration zones that limit insecticide penetration and interaction with target sites [149,150]. For example, An. gambiae expresses SAP2, a sensory appendage protein that can bind with deltamethrin. Resistance collapses upon SAP2 silencing via RNAi [151]. In the brown planthopper, NIOBP3 binds the neonicotinoid nitenpyram, with RNAi-mediated silencing restoring susceptibility [152]. Results of studies with transgenic Drosophila further confirm functional sequestration activity across multiple species, including aphids, moths, and honeybees [153,154]. These findings support the hypothesis that olfactory proteins have evolved multifunctional roles, acting as frontline defenders against insecticide exposure [155].

3.2. RNA-Mediated Resistance: Noncoding RNAs as Posttranscriptional Regulators

The concept of RNA-mediated resistance is based on long noncoding RNAs (lncRNAs) and microRNAs (miRNAs) that orchestrate changes in the expression of genes that enhance resistance phenotypes. Noncoding RNAs modulate the expression of genes through various mechanisms, including mRNA stabilization, miRNA sponging, and epi-transcriptomic modification. In P. xylostella, the lncRNA lnc-GSTu1-AS protects GSTu1 transcripts from miRNA (miR-8525-5p)-mediated degradation, sustaining GST-based detoxification and chlorantraniliprole resistance [156]. In Drosophila, lnc19419 acts as a competing endogenous RNA, binding miR-944 and derepressing CPCFC—a cuticular protein gene implicated in malathion resistance [157]. Epi-transcriptomic modifications such as m6A methylation in the 5’UTR of CYP4C64 enhance gene expression and enzymatic detoxification efficiency in whiteflies [84]. These findings underscore the sophistication of RNA-based regulatory networks in modulating resistance beyond static genetic changes.

3.3. Role of the Microbiome in Detoxification: A Dynamic Evolutionary Defense

Insect gut microbiomes play a vital role beyond nutrition by acting as adaptive biochemical filters that metabolize insecticides into less toxic compounds, enhancing host survival and contributing to chemical resilience across diverse taxa [158,159,160]. These microbial communities dynamically respond to pesticide exposure through upregulation of degradative enzymes and shifts in community composition, often working synergistically with host detoxification systems [161,162]. Notably, symbionts play a key role in shaping the functional diversity and ecological dynamics of these interactions. This symbiotic detoxification represents an evolving defense strategy with implications for sustainable pest and pollinator management [163].

Role of the Microbiome in Insect Detoxification and Resistance: Mechanistic Insights

Members of the insect microbiome, including both gut-associated bacteria and intracellular endosymbionts, can influence resistance to insecticides through multiple mechanisms, not only by directly metabolizing xenobiotics but also by modulating host detoxification gene expression and immune-regulated defense pathways [164]. Endosymbionts such as Wolbachia and Rickettsia have been shown to alter insecticide susceptibility in a context-dependent manner, with effects that vary according to host species, symbiont strain, and the specific chemical exposure. These modulatory effects are primarily mediated through alterations in host metabolic processes, immune responses, and gene regulatory networks [165].
For example, Wolbachia enhances pesticide resistance in T. urticae by upregulating detoxification-related genes such as CYPs and GSTs, thereby increasing enzyme activity [166]. RNAi experiments confirmed that the genes TuCYP392D2 and TuGSTd05 are essential for this resistance phenotype. Notably, exposure to abamectin results in elevated Wolbachia levels, while the overall microbiome composition remains unchanged, highlighting the pivotal role of Wolbachia in augmenting detoxification and resistance to insecticides [166].
In N. lugens, Wolbachia enhances resistance to imidacloprid by promoting the expression of detoxification genes, including CYPs (e.g., CYP6ER1, CYP6AY1, CYP4CE1) and GSTs, potentially through activation of the CncC oxidative stress response pathway. This regulatory effect is temperature-sensitive: elevated temperatures reduce Wolbachia abundance, which in turn is associated with a decrease in detoxification gene expression and a corresponding decline in resistance to insecticides [167]. In contrast, an S-type strain of Arsenophonus has been shown to suppress host xenobiotic metabolism and reduce imidacloprid resistance in N. lugens, as evidenced by the downregulation of detoxification-related genes and altered metabolic profiles [168]. These findings demonstrate that the identity of the symbiont strain plays a crucial role in differentially modulating host detoxification capacity and insecticide susceptibility.
Further, the gut microbiota of honeybees, which is predominantly composed of Lactobacillus, exhibits temporal fluctuations during overwintering, with strain-specific variations that may influence immune function, metabolism, and overall health [169]. Microbial symbionts also contribute to toxin resistance by regulating cytochrome P450 gene expression in the midgut, thereby facilitating detoxification [170]. Collectively, these studies underscore the environmentally responsive nature of symbiont-mediated detoxification mechanisms and emphasize the vulnerability of these microbe-dependent resistance pathways to thermal stress.
In addition to influencing host gene expression, some microbiota confer resistance through direct metabolic detoxification. For example, N. lugens can acquire environmental Serratia marcescens capable of degrading buprofezin; gain or loss of this symbiont alters resistance levels, confirming its functional role in xenobiotic metabolism [171]. In Bactrocera dorsalis, Citrobacter sp. (CF-BD) carries phosphatase hydrolase genes that enhance survival under trichlorfon exposure [158]. Symbionts can also contribute to detoxification via distinct enzymatic pathways, as demonstrated in D. melanogaster, where gut microbiota mediate nitro-reductive imidacloprid metabolism in parallel with the host’s oxidative CYP6G1 pathway [172]. Similar dual-function mechanisms have been reported in other species. In A. gossypii, the gut symbiont Sphingomonas mediates imidacloprid resistance through both chemical degradation and upregulation of host cytochrome P450 genes, indicating a combined direct and indirect mode of action [173]. Likewise, in the bean bug Riptortus pedestris, Burkholderia symbionts degrade organophosphates such as fenitrothion while simultaneously inducing host gene expression changes that enhance resistance traits [161]. These findings highlight the complex and multifaceted contributions of insect-associated microbes to detoxification and the evolution of resistance.
Beyond direct enzymatic activity and gene regulation, symbiotic bacteria also influence resistance to insecticides by modulating host immune pathways. In Lymantria dispar larvae, shifts in the Gram-negative/Gram-positive composition of gut microbiota alter expression of Toll and IMD signaling components [174]. These immune pathways intersect with oxidative stress regulators such as Nrf2, which play a key role in the induction of detoxification enzymes. As a result, microbial regulation of host immunity represents an additional, indirect mechanism through which the microbiota can shape xenobiotic tolerance [175]. Table 3 summarizes key cases of microbiome-mediated insecticide resistance in insects, highlighting mechanisms like direct metabolism, gene regulation, and immune modulation.

3.4. Horizontal Gene Transfer (HGT) as a Driver of Novel Resistance Mechanisms

Emerging evidence indicates that horizontal gene transfer (HGT)—particularly from microbial partners such as bacteria or bacteriophages into insect genomes—can confer novel detoxification or immune effector capabilities. In M. persicae, fungal-origin horizontally transferred genes (HTGs) associated with cyanate detoxification have been shown to reduce reproduction when silenced, underscoring their functional role in managing toxic stress [176]. In gall midges (Contarinia nasturtii) and other Cecidomyiidae, microbial toxin genes including cdtB, aip56, rhs, and sltxB have been horizontally acquired and retain conserved catalytic residues, suggesting a contribution to insect adaptation through modulation of host interactions or immune responses [177]. Similarly, in the whitefly B. tabaci, a gene encoding a plant-derived detoxification enzyme appears to have been acquired via virus-mediated inter-kingdom transfer from host plants, potentially influencing both host-plant interactions and insecticide metabolism [178]. In Lepidoptera, the majority of identified HTGs are implicated in physiological roles such as nutritional metabolism and detoxification [179].
In R. pedestris, gut symbionts degrade the organophosphate insecticide fenitrothion into a non-toxic compound via a horizontally acquired enzyme. This bactericidal byproduct is subsequently excreted by the host, constituting a coordinated “host-symbiont detoxification relay” that maintains symbiosis while conferring effective resistance to insecticides. This mechanism parallels host-encoded detoxification pathways observed in other insects, highlighting the adaptive potential of microbial symbiosis and horizontal gene transfer in overcoming chemical stress [161].
Although direct evidence for HGT-derived detoxification genes in major agricultural pests remains limited, existing precedents from other insect species suggest these mechanisms may be more widespread and warrant further exploration. Recent reviews emphasize that bacterial symbionts and their associated bacteriophage communities represent a rich reservoir of genetic material capable of horizontal transfer, likely contributing to genome innovation in insects, particularly within ecological niches subjected to insecticide pressure [180,181]. These findings highlight the need to explore ecological, behavioral, and environmental factors in future resistance research.

3.5. Future Directions

Emerging evidence points toward even more complex and integrative resistance mechanisms that include: (1) Microbial symbioses may confer metabolic capabilities or modulate host immunity in ways that enhance insecticide tolerance [182]. (2) Behavioral adaptations, including altered movement and feeding patterns, can reduce exposure to treated surfaces [183]. (3) Climate change may accelerate resistance evolution not only through altered selection pressures and life histories [184], but also via temperature-dependent modulation of detoxification pathways—where elevated temperatures enhance cytochrome P450 activity and expression in pests like Ae. aegypti and B. tabaci, thereby increasing metabolic resistance to pyrethroids and neonicotinoids [184]. Together, these discoveries expand the resistance paradigm and highlight the adaptive plasticity of pest species subject to chemical control.

4. Cross-Resistance and Multiple Resistance in Insect Pests—Mechanisms, Environmental Factors, and Management Strategies

4.1. Resistance Types and Their Management Implications

4.1.1. Cross-Resistance: One Mechanism, Many Failures

Cross-resistance poses a significant challenge to insecticide rotation strategies. This form of resistance often emerges silently in field populations under diverse chemical exposure, highlighting the value of early diagnostic screening for broad-spectrum enzymatic activity [185,186]. Notably, cross-resistance is often concentrated in polyphagous pests exposed to diverse insecticide chemistries [187,188], such as Spodoptera spp. and P. xylostella, due to their broad host range and frequent insecticide encounters. It occurs when a single genetic or biochemical mechanism confers resistance to multiple insecticides, sometimes spanning chemically unrelated classes. The most common drivers are detoxification enzymes with broad substrate specificity, particularly P450s. For instance, in S. exigua, the enzyme CYP321A8 metabolizes pyrethroids (cypermethrin and deltamethrin) and organophosphates (chlorpyrifos). This multifunctionality is mediated by two mechanisms: transcription factor overexpression and promoter mutations creating a new cis-regulatory element for orphan nuclear receptor binding [189]. The P450 gene CYP321A8 also contributes to insecticide cross-resistance in field populations of S. frugiperda [190]. Further, Zuo et al. [191] studied the roles of different resistance mechanisms in S. exigua using nearly isogenic lines. They found that the L1014F mutation in the VGSC led to a resistance factor of 6.2, while the overproduction of the cytochrome P450 enzyme CYP9A9 resulted in a much higher resistance factor of 79. When both mechanisms were present, the resistance factor increased to 631 [191]. Similarly, P450s in S. litura drive indoxacarb resistance. Silencing CYP339A1, CYP340G2, and CYP321A19 increased sensitivity, while transgenic overexpression enhanced resistance in Drosophila. These P450s may also mediate cross-resistance to chlorantraniliprole, λ-cyhalothrin, and imidacloprid [192].
Not all mechanisms of resistance rely on metabolism. For instance, chlorfenapyr resistance in P. xylostella is primarily governed by one major gene or a few tightly linked loci, with autosomal and incompletely dominant inheritance [193]. This resistance exhibited minimal involvement of detoxification enzymes and showed no significant cross-resistance to other insecticides, which suggests a distinct mechanism, potentially involving mitochondrial disruption or non-specific energy metabolism effects rather than typical target-site or metabolic alterations.
Environmental stressors complicate resistance dynamics. Exposure to heavy metals like cadmium (Cd) can trigger overexpression of detoxification genes through activation of the ROS/CnCC signaling pathway, enhancing P450-mediated detoxification and contributing to cross-resistance in Cd-exposed insects [194]. This phenomenon, known as xenobiotic preconditioning, can predispose pest populations to insecticide cross-resistance even without direct exposure to insecticides [195], which illustrates how anthropogenic pollution can inadvertently amplify resistance risks. In some pest species, such as B. tabaci, both cross-resistance and multiple resistance coexist, with overlapping metabolic and target-site mechanisms evolving under sustained and diverse chemical pressures [196].

4.1.2. Multiple Resistance: Accumulation of Independent Mechanisms

Multiple resistance arises when pest populations independently acquire mechanisms that impart resistance to several insecticide classes, resulting in a polygenic and often synergistic defense system [197,198]. Such resistance profiles are especially concerning when different mechanisms converge in a single population, reducing the utility of rotation or combination treatments. Unlike cross-resistance, which is typically mediated by a single mechanism (e.g., detoxification enzyme or target-site mutation) conferring resistance to structurally related compounds, multiple resistance involves distinct mechanisms selected by sequential or overlapping exposure to different chemistries. Multiple resistance is frequently observed in pests subjected to intensive insecticide regimes in monoculture cropping systems, such as Chilo suppressalis and S. frugiperda, where diverse insecticide classes are applied over multiple generations, driving accumulation of resistance traits [197,199]. For example, C. suppressalis has evolved resistance to diamides through multiple mutations in the ryanodine receptor gene (e.g., I4758M, Y4667C, Y4667D), with combinations of these mutations contributing additively to high levels of tetraniliprole resistance [200]. Tetraniliprole is a pyrazole carboxamide insecticide, which has the structure of cyantraniliprole in which the bromine atom has been replaced by a [5-(trifluoromethyl)-2H-tetrazol-2-yl]methyl group. It has a role as a ryanodine receptor agonist [201]. In S. frugiperda, resistance-associated mutations have been identified across multiple target-site genes—ace-1, CHSA, GluCl, and RyR—indicating the potential for multiple resistance to organophosphates, chitin synthesis inhibitors, ivermectins, and diamides [56,186].
Regulatory mechanisms also play a role. In N. lugens, cis-regulatory variation upstream of CYP6CS1 enhances P450 expression, conferring resistance to pymetrozine and moderate cross-resistance to other insecticides [202]. Additionally, resistance evolution often entails fitness trade-offs. In A. gossypii, sulfoxaflor resistance declined after selection pressure was removed, but reduced phloem-feeding efficiency and lower relative fitness persisted, illustrating the lingering costs of resistance, even after phenotypic reversion [203].

4.1.3. Strategic Implications: Toward Mechanism-Informed IRM

Addressing these resistance patterns demands more than chemical rotation—it requires tactical integration of molecular diagnostics and ecological insights. The intertwined challenges of cross-resistance and multiple resistance call for a fundamental shift in the management of resistance to insecticides (IRM). Traditional rotation strategies, once the cornerstone of resistance delay, often fall short when resistance alleles are already widespread or when cross-resistance networks link multiple insecticides. T. absoluta populations resistant to tetraniliprole (a diamide) also show cross-resistance to chlorantraniliprole and flubendiamide, but remain susceptible to other insecticides like emamectin benzoate, broflanilide, spinosad, metaflumizone, and indoxacarb [204]. This pattern restricts the effectiveness of simple rotation schemes. These examples demonstrate how detailed resistance profiling—identifying specific cross-resistance patterns and independent resistance mechanisms—can guide more precise insecticide selections that minimize efficacy loss and delay resistance evolution. Rotations effectively delay resistance only when insecticides have non-overlapping modes of action and are applied to mostly susceptible pest populations—conditions that are rarely found in the field [205].
To overcome these challenges, IRM must incorporate integrated, mechanism-informed strategies such as (1) high-throughput diagnostics tools: including qPCR assays for detoxification gene expression and allele-specific PCR for key resistance mutations (e.g., kdr, RyR), enabling real-time resistance monitoring and targeted interventions [206]. (2) Synergist-enhanced formulations: Combining insecticides with metabolic enzyme inhibitors, such as PBO, can restore efficacy by blocking key resistance pathways. Novel approaches like RNAi offer promising routes to bypass known resistance mechanisms altogether [207]. (3) Ecological management: Mitigating environmental stressors, such as heavy metal contamination that can pre-activate detoxification pathways, reduces unintended induction of resistance. Concurrently, integrating biological control agents (e.g., Trichogramma spp.) diversifies pest mortality sources and disrupts selection for resistance [208,209]. (4) Exploiting fitness costs: Where resistance imposes biological trade-offs, temporarily withdrawing or rotating insecticides can enable susceptible genotypes to rebound, slowing resistance spread [203].
Ultimately, sustainable IRM demands abandoning one-size-fits-all prescriptions in favor of dynamic, data-driven strategies that leverage detailed biochemical, ecological, and genomic insights. Such integrated approaches will better guide insecticide deployment in both space and time, preserving control efficacy and agricultural productivity [205,210].

4.2. Resistance Management: A Multitactic Framework for Insecticide Sustainability

The multifactorial nature of factors influencing IRM efficacy is summarized in Figure 2, which highlights the interplay between operational, biological, tactical, and monitoring components essential for successful IRM. Resistance to insecticides poses an escalating threat across agricultural, public health, and urban sectors, demanding comprehensive, science-based management. Mode-of-Action (MoA) rotation, a cornerstone strategy, reduces selection pressure by alternating chemistries with distinct biochemical targets. The IRAC classification system guides evidence-based rotations, particularly in agriculture, where chemical diversity and pyramided Bt crops (e.g., cotton and maize) delay resistance development [211,212,213]. For instance, Australia’s cotton industry implemented mandatory IRAC MoA rotation since the 1990s, restricting consecutive applications of any insecticide class to a single generation of Helicoverpa pests. This strategy—combined with refuge crops and threshold-based spraying—reduced pyrethroid resistance frequencies from >50% to <10% within 5 years while extending the efficacy of newer chemistries like diamides [211]. However, cross-resistance complicates this approach. Metabolic mechanisms such as CYPs upregulation and target-site mutations like kdr can confer resistance across multiple insecticide classes [131,185]. Therefore, effective IRM requires molecular diagnostics and surveillance data to inform timely and tactical shifts in insecticide use.
Complementary strategies enhance IRM resilience. Insecticide mixtures and synergists, which combine full doses of active ingredients with independent MoAs, significantly reduce the risk of pests developing dual resistance [214]. Innovations such as next-generation long-lasting insecticidal nets (LLINs) that combine pyrethroids with synergists like PBO or novel insecticides, such as chlorfenapyr, exemplify this strategy in vector control [215,216]. Refugia strategies, especially structured non-Bt plantings in transgenic crop systems, preserve susceptible genotypes and dilute resistance alleles through interbreeding [217,218]. Although refugia are less feasible in public health and urban contexts, the conservation of natural enemies offers partial mitigation by maintaining ecological balance [219,220,221].
The successful implementation of the management of resistance to insecticides depends on several pillars. Surveillance-driven decision thresholds, using bioassays [222] and molecular tools, such as allele frequency tracking, trigger interventions before resistance becomes widespread [223,224,225,226,227]. Diversified tactics beyond chemicals, including biological control agents like entomopathogens, RNAi, gene drives, and ecological methods such as attract-and-kill, reduce reliance on insecticides and delay resistance evolution [228,229,230,231]. Finally, robust policy frameworks like the WHO Global Plan for Integrated Resistance Management (GPIRM) and IRAC guidelines provide essential structures for coordinated action, while stakeholder education and capacity building ensure adaptive and effective execution [232,233,234]. In summary, the proactive integration of these elements across both institutional and ecological contexts is vital to sustaining insecticide efficacy and protecting human health and agricultural productivity into the future.

4.3. Divergent Strategies for RNAi Resistance to Insecticides Management, a Comparative Framework for Transgenic and Sprayable Applications

4.3.1. Distinct Exposure Dynamics Drive Divergent IRM Needs

The contrasting exposure profiles of RNAi-based insecticides, transgenic versus sprayable, fundamentally shape their IRM strategies. Transgenic crops such as SmartStax PRO® maize, which express DvSnf7 dsRNA, expose target pests like western corn rootworm to season-long, continuous selection pressure. This persistent exposure creates a high-risk environment for the evolution of resistance across multiple overlapping generations in the field [207,235]. In contrast, sprayable dsRNA products like Calantha® (Ledprona®), which target the Colorado potato beetle, were developed specifically for pest management and are applied intermittently [236]. The rapid environmental degradation of dsRNA [237] limits exposure to short, defined treatment windows, ideally targeting a single pest generation—often the early larval stage. These fundamental differences in exposure duration and intensity demand distinct IRM designs: long-term containment for transgenic traits versus generation-specific rotation for foliar sprays.

4.3.2. Tailored IRM Strategies Reflect Product Format and Biology

For transgenic RNAi insecticides regulated as Plant-Incorporated Protectants (PIPs), IRM strategies focus on pyramiding and refuge requirements. Products like SmartStax PRO® combine dsRNA with Bt proteins (e.g., Cry3Bb1 and Gpp34Aa/Tpp35Aa) to deliver multiple, non-cross-resistant modes of action within a single crop. This approach ensures that even if resistance arises to one component, the other maintains efficacy [235,238]. To further delay resistance, a structured refuge of 5% non-transgenic seed is included in seed mixtures. This practice promotes mating between susceptible and potentially resistant individuals, leveraging the recessive nature of dsRNA resistance inheritance to reduce the frequency of homozygous-resistant offspring [239].
Management of resistance of insects to management for sprayable dsRNA, categorized under IRAC Mode of Action Group 35, relies on temporal separation of exposures via mode-of-action (MoA) rotation. Products are applied within defined “treatment windows,” typically limited to one application per pest generation, followed by rotation to a distinct MoA group (e.g., IRAC Groups 4 or 28) for the next generation or application [233]. This strategy takes advantage of the lack of cross-resistance between dsRNA and small-molecule insecticides, reducing selection pressure continuity and preserving dsRNA efficacy [240].

4.3.3. Mechanisms of Resistance Inform Risk and Regulatory Design

Both strategies are informed by a shared mechanistic understanding of dsRNA resistance in coleopteran pests. The predominant mechanism, impaired cellular uptake of dsRNA, exhibits autosomal recessive inheritance and confers cross-resistance among RNAi targets, but not to Bt toxins or conventional chemical insecticides [239,240,241]. Laboratory selection studies have demonstrated that resistance can evolve rapidly, within as few as 7–11 generations, and resistant populations can suffer little due to a lack of fitness cost in the absence of selection [242]. These findings reinforce the necessity of early and robust implementation of IRM to prevent widespread resistance. Each approach aligns with regulatory frameworks designed for each system of delivery. Transgenic RNAi traits must adhere to EPA guidelines that emphasize high-dose expression, pyramiding of non-cross-resistant traits, structured refuges, and resistance monitoring protocols [238,243,244]. Alternatively, Sprayable RNAi insecticides are seamlessly incorporated into global MoA-based IRM systems developed by the Resistance to Insecticides Action Committee (IRAC), including principles of rotation and generational treatment limits [245]. Going forward, harmonizing these IRM strategies across international jurisdictions will be critical for supporting the long-term sustainability of RNAi technologies in IPM programs [207].

5. Case Studies

The mosquito Ae. aegypti and the fall armyworm S. frugiperda were selected as representative case studies due to their global economic and public health significance, their rapid evolution of resistance, and the extensive research characterizing their underlying mechanisms across diverse ecological and regulatory contexts [246,247].

5.1. Mechanisms of Resistance to Insecticides for the Mosquito, Aedes aegypti

The extensive and prolonged application of insecticides has exerted intense evolutionary pressure on mosquito populations, particularly Ae. aegypti, fostering the emergence of resistance (Table 4). Resistance to insecticides is widely considered a pre-adaptive process, whereby alleles conferring survival advantages may be present at low frequencies within populations prior to exposure to selective agents [248]. However, resistance can also arise from de novo mutations following insecticide application. Individuals harboring these alleles exhibit increased survival and reproductive success, leading to a rise in resistance allele frequencies. Two principal mechanisms underlie this resistance: alterations at the insecticide target site and enhanced metabolic detoxification [124].

5.1.1. Target Site Resistance

Target site resistance arises from nucleotide substitutions that lead to changes in amino acids in critical proteins, which reduce the binding affinities of insecticides. Notably, mutations in the AChE genes, Ace1 and Ace2, have been implicated in resistance to organophosphates in species such as An. gambiae, Cx. pipiens, and Culex tritaeniorhynchus [249,250,251]. However, conclusive evidence of this mechanism in Ae. aegypti remains to be done.
Resistance to pyrethroids, often termed “knockdown resistance” (kdr), involves point mutations in the VGSC (Figure 3), a primary target of these insecticides. Such mutations have been identified across a broad spectrum of arthropods, including both pest and vector species [64]. These mutations may influence channel gating dynamics, affecting insecticide binding efficacy. Eleven VGSC mutations associated with pyrethroid resistance have been identified in Ae. aegypti.
V1016I Mutation
The V1016I mutation, identified by [252], entails a G → A substitution at codon 1016 in domain II, segment 6 (IIS6) of the VGSC. This single-nucleotide polymorphism results in a valine-to-isoleucine amino acid change. While this mutation alone may not significantly alter insecticide susceptibility, it often co-occurs with F1534C, and their combination appears to synergistically impair insecticide binding [64,253].
F1534C Mutation
F1534C, first observed in populations from the Grand Cayman Islands and Thailand [254,255], results from a T → G transversion at the second position of codon 1534 in domain III, segment 6 (IIIS6). This leads to a substitution of phenylalanine with cysteine. Functional studies have demonstrated that F1534C confers resistance primarily to type I pyrethroids, with limited efficacy against type II compounds (Figure 3) [256].
V410L Mutation
The V410L mutation (Figure 3), first identified in 2017 in a Brazilian laboratory strain of Ae. aegypti [257], results from a G → T transversion at the first nucleotide of codon 410 in domain I, segment 6 (IS6), leading to a valine-to-leucine amino acid substitution. Although initially absent from wild populations, its frequency has increased substantially over time. In Tapachula, Mexico, by 2016, heterozygosity at the 410 locus had increased to 64% [258]. This trend paralleled increases in the frequencies of V1016I and F1534C, and a strong genotype–phenotype correlation was established for resistance to both type I (e.g., permethrin) and type II (e.g., deltamethrin) pyrethroids [258].
Figure 3. Schematic representation of three key amino acid substitutions—V410L, V1016I, and F1534C—in the Aedes aegypti voltage-gated sodium channel (VGSC). These mutations are located in domain I segment 6 (IS6), domain II segment 6 (IIS6), and domain III segment 6 (IIIS6), respectively. The domain segments and connecting helices contributing to pyrethroid receptor sites 1 (green) and 2 (purple) are also highlighted. These structural changes are associated with knockdown resistance (kdr) to pyrethroids. Reproduced from [258]. Copyright 2018, Nature Portfolio.
Figure 3. Schematic representation of three key amino acid substitutions—V410L, V1016I, and F1534C—in the Aedes aegypti voltage-gated sodium channel (VGSC). These mutations are located in domain I segment 6 (IS6), domain II segment 6 (IIS6), and domain III segment 6 (IIIS6), respectively. The domain segments and connecting helices contributing to pyrethroid receptor sites 1 (green) and 2 (purple) are also highlighted. These structural changes are associated with knockdown resistance (kdr) to pyrethroids. Reproduced from [258]. Copyright 2018, Nature Portfolio.
Toxics 13 00681 g003

5.1.2. Metabolic Resistance

Enzymatic detoxification represents another major mechanism whereby Ae. aegypti mitigates insecticidal toxicity. Elevated levels of CES activity have been associated with phenotypes resistant to insecticides [259]. Overexpression of several detoxifying enzyme families, including CYPs, carboxyl/cholinesterases (CCEs), and GSTs, has been documented in resistant populations of mosquitoes [246,260,261]. These enzymatic systems often work in concert, with CCEs initiating pyrethroid hydrolysis, followed by further metabolism via CYPs [262]. Such cross-functional enzymatic activity can result in broad-spectrum resistance, independent of the insecticide’s target site [131].

5.1.3. Resistance to Insecticides and Associated Fitness Costs

The evolutionary dynamics of resistance to insecticides are intertwined with vector fitness and epidemiological outcomes. While resistance mechanisms can impose physiological trade-offs that affect vectorial capacity and may lead to reversion toward susceptibility in insecticide-free environments, this is not universally the case [263]. Some resistance alleles, such as the Ae. aegypti V1016I kdr mutation, exhibit low fitness costs and can persist in populations even in the absence of selection pressure [264]. Thus, the persistence or decline of resistance alleles depends on the balance between fitness costs and benefits in the specific ecological context [265,266].
Several longitudinal studies have explored the consequences of resistance to insecticides on fitness [267,268]. Evaluating fitness requires careful measurement of key life-history traits, including adult longevity, host-seeking behavior, reproductive success (e.g., fecundity and egg viability), and larval development [265,269,270,271]. However, these comparisons must be cautiously interpreted. These physiological costs can reduce a mosquito’s vectorial capacity under pesticide stress. They do so by shortening its lifespan or impairing its pathogen defense mechanisms [272,273,274]. Resistance phenotypes may result from multiple interacting mechanisms, and variations in environmental adaptation among field and laboratory strains can confound conclusions [269,275].
Further research is required to elucidate the extent and ecological significance of these trade-offs, particularly in multi-generational contexts and under varying environmental conditions. A deeper understanding of resistance-associated fitness costs holds twofold importance: it can help predict the reversibility of resistance post-intervention and inform vector control strategies by evaluating how these costs affect mosquito population dynamics and pathogen transmission. Importantly, regional differences in insecticide regulations and use patterns contribute to the selection of distinct resistance alleles and mechanisms in Ae. aegypti populations worldwide. For example, pyrethroid resistance is more prevalent in Latin America and Southeast Asia due to historical overuse, while organophosphate resistance remains significant in Africa [246,276,277,278]. These differences underscore the need for locally adapted vector control strategies.

5.2. Mechanisms of Resistance to Insecticides in Spodoptera frugiperda

The fall armyworm (S. frugiperda, (FAW)) exhibits multiple, often concurrent, resistance mechanisms that complicate efforts to control with insecticides (Table 4). These include metabolic detoxification, target-site mutations, reduced cuticular penetration, and behavioral avoidance. Metabolic resistance is primarily mediated by elevated activity of detoxification enzymes, such as CYPs, GSTs, and CESs. For instance, overexpression of CYP genes, including CYP6AE43, CYP321A8, and CYP321A9, has been linked to enhanced degradation of diamide insecticides such as chlorantraniliprole [70].
Target-site resistance arises from mutations that alter the binding affinity of insecticides. The I4790K mutation in the RyR reduces the efficacy of diamides [186], while point mutations such as F290V and A201S in AChE confer resistance to organophosphates and carbamates [56,224,279,280]. Additional mechanisms, including decreased cuticular penetration and avoidance behaviors of larvae, reduce insecticide exposure [161]. These mechanisms frequently co-occur, resulting in widespread multi-resistance to insecticides in FAW populations globally [281,282].

5.2.1. Global Distribution and Severity of Resistance

Resistance levels vary markedly across regions, reflecting local insecticide use patterns. In China, field populations of S. frugiperda have developed moderate to high resistance to indoxacarb (resistance ratio [RR] = 9.23–45.53×) and spinetoram (RR = 4.32–18.05×), with incipient signs of reduced susceptibility to chlorantraniliprole [159]. Egyptian S. frugiperda populations exhibit moderate resistance to cypermethrin (RR = 3.65–5.75), with metabolic detoxification, particularly GST activity, contributing to resistance. Emamectin benzoate remains highly efficacious, while synergists enhance spinosad toxicity, supporting an enzyme-mediated resistance mechanism [283].
In Brazil, S. frugiperda populations have evolved high resistance levels to diamide insecticides—including flubendiamide, chlorantraniliprole, cyantraniliprole, and tetraniliprole—primarily mediated by target-site mutations in the RyR, notably I4790M and I4790K substitutions [284]. Genetic analyses indicate resistance is inherited as an autosomal, incompletely recessive, monogenic trait. Robust cross-resistance among diamides has been documented. Although the overall frequency of these resistance-associated alleles remains low in many regions, locally elevated frequencies underscore the urgent need for integrated resistance management strategies to sustain diamide efficacy in fall armyworm control [186].
Genetic analysis of invasive S. frugiperda populations in India revealed low nucleotide diversity and close genetic affinity to African populations, indicating recent common origins. Upregulation of detoxification genes (CYPs, GSTs) underlies observed resistance to insecticides, with resistance ratios significantly correlated to specific loci (e.g., GST1950, CYP9131, CYP9360), highlighting the contribution of heritable metabolic mechanisms to resistance development [285,286].
In Europe, invasive S. frugiperda populations—such as those in Greece—show alarmingly high frequencies (up to 80.9%) of ABCC2 gene mutations, molecular markers associated with resistance to Bt Cry toxins, emphasizing an urgent need for rigorous resistance monitoring [287].
The diversity of regulatory frameworks and insecticide usage across continents has resulted in geographically distinct resistance profiles in S. frugiperda. For example, the rapid emergence of diamide resistance in Brazil contrasts with the slower development of resistance in parts of Africa, where Bt crops and conventional chemistries are deployed differently [281,284].

5.2.2. Integrated Resistance Management Strategies

The effective management of resistance in S. frugiperda necessitates an integrated and multifaceted approach. Strategic rotation of insecticides with distinct modes of action—such as alternating diamides (Group 28) with spinosyns (Group 5)—is imperative to delay the onset of resistance [288,289]. Furthermore, the application of synergists, notably diethyl maleate (DEM), has been shown to significantly enhance insecticidal efficacy, augmenting spinosad toxicity by up to eightfold in resistant populations [283].
Biological control constitutes a sustainable and ecologically sound strategy for managing S. frugiperda; however, its global implementation remains constrained and methodologically fragmented. Although over 500 natural enemies have been documented—including approximately 304 parasitoids, 215 predators, and 46 entomopathogens—only approximately 40% have been subjected to laboratory or field evaluations, with fewer undergoing rigorous population-level efficacy assessments [290]. Notably, isolates of Beauveria bassiana have exhibited considerable potential, eliciting up to 87% egg mortality and 71–93% cumulative larval mortality within 14 days, in addition to suppressing feeding by second-instar larvae by 69–78% within 48 h [291]. Despite these promising results, geographic and taxonomic biases, insufficient bio-inventories, and the absence of standardized efficacy metrics impede the extrapolation of experimental findings to broad-scale field applications.
Unlike chemical insecticides, biological control agents seldom achieve complete eradication; instead, they promote sustained population suppression within ecological thresholds, preserving natural enemy communities and reducing the risk of pest resurgence [292]. Complementary cultural practices, such as push-pull strategies and optimized planting schedules, have been demonstrated to substantially enhance early-season pest suppression. For example, a climate-adapted push-pull system implemented in East Africa, which involves intercropping maize with drought-tolerant Desmodium species and deploying Brachiaria cv. Mulato II, as a border crop, has yielded significant reductions in pest infestation and crop damage, concurrently improving maize yields [293].
Ultimately, the successful reduction of insecticide overuse and the effective adoption of IPM frameworks are contingent upon comprehensive farmer education and active stakeholder engagement [294].

5.2.3. Future Directions for Resistance Mitigation

Long-term control of FAW will depend on genomics-guided surveillance as well as other emerging technologies. Molecular diagnostics allow for real-time detection of resistance-associated alleles, such as RyR I4790K, in field populations [287]. New insecticides, including fluxametamide—a novel isoxazoline insecticide (Group 30)—exhibit low resistance potential, with a realized heritability (h2) of 0.084 and only a 2.63-fold increase in LC50 after 10 generations of selection. The selected strain showed no significant cross-resistance except a modest increase to emamectin benzoate and displayed fitness costs, including reduced reproduction and development (relative fitness = 0.353) [163].
Biotechnological advances such as RNAi targeting essential genes like chitin synthase and transgenic crops expressing multiple Bt toxins offer promising control strategies. However, their success hinges on proper refuge management to mitigate resistance evolution [186]. The establishment and maintenance of global resistance databases, such as the FAO’s FAW portal, are essential for harmonized data reporting and informed decision-making [8]. Finally, future investment in next-generation control tools—including nanoparticle-based formulations and mycoinsecticides—holds potential to enhance efficacy while minimizing environmental impact [295].
Closing remarks—The diverse resistance mechanisms and control challenges observed in S. frugiperda contrast with those in Ae. aegypti, reflecting their distinct life histories, ecologies, and insecticide selection pressures. Together, these two species illustrate divergent evolutionary trajectories and multifaceted resistance adaptations shaped by their biology and exposure histories. This comparison underscores the critical importance of tailoring IRM strategies to species-specific vulnerabilities and local chemical selection landscapes.

6. Strategies for Managing Resistance to Insecticides: Chemical, Biological, and Integrated Approaches

The escalating threat of resistance to insecticides in pest populations necessitates a paradigm shift in pest control strategies. Traditional reliance on chemical insecticides has proven insufficient and environmentally detrimental in the long term. Therefore, modern pest management emphasizes integrated approaches that combine chemical, biological, and ecological tactics to delay resistance evolution and minimize ecological impacts (Table 5).

6.1. Conventional Insecticides: Effectiveness and Limitations

Insecticides such as organophosphates, pyrethroids, and neonicotinoids remain essential in current pest management frameworks due to their broad-spectrum activity and rapid knockdown effects [125]. However, their efficacy is increasingly undermined by widespread resistance—driven by metabolic, target-site, penetration, behavioral, and emerging sequestration mechanisms—and by cross-resistance across chemical classes [296]. Operational factors such as poor spray coverage, mistimed application, and pest cryptic behavior further reduce control success [297]. Moreover, their environmental persistence and broad-spectrum toxicity lead to non-target impacts, ecosystem imbalance, pest resurgence and secondary outbreaks, soil and water contamination, and human health hazards from acute and chronic exposure [298]. For instance, neonicotinoids, despite their systemic properties and efficacy against sap-feeding pests, have been implicated in the disruption of pollinator navigation and colony collapse disorder [299].
Sulfoxaflor, a next-generation sulfoximine insecticide, targets sap-feeding pests resistant to neonicotinoids by acting as a nAChR agonist at a distinct receptor subtype. Initially deemed safer for pollinators, recent studies reveal sublethal effects on bees, including impaired foraging, disrupted gut microbiota, and reduced colony performance [300,301,302,303]. Thus, while chemical insecticides remain vital tools, they must be deployed judiciously within IRM frameworks to sustain efficacy and reduce ecological footprints.
These limitations underscore the growing role of IPM—including strategic rotation with biorational agents—to mitigate resistance development, reduce environmental burden, and sustain long-term effectiveness [304].

6.2. Next-Generation Resistance-Breaking Technologies

While IPM remains essential, truly durable solutions require disrupting resistance at its genetic roots. Such as (1) Nanocarrier-based delivery of dsRNA targeting long non-coding RNAs (e.g., lnc-GSTu1-AS) or resistance-related genes (e.g., GSTu1, SAP2, NIOBP3) enables effective RNAi-mediated suppression of detoxification pathways associated with resistance to insecticides. In P. xylostella, silencing lnc-GSTu1-AS destabilized GSTu1 transcripts and increased susceptibility to chlorantraniliprole [156]. Similarly, silencing CYP6CY3 via nanocarrier-delivered dsRNA increased imidacloprid susceptibility in A. gossypii [305]. Broadly, nanoparticles—including liposomes, chitosan polymers, dendrimers, and lipid-protamine complexes—enhance delivery efficacy by protecting dsRNA from degradation, improving cellular uptake, and promoting endosomal escape, leading to stronger gene knockdown and pest control outcomes [306]. (2) CRISPR-Cas9 allelic-drive systems can replace resistant kdr mutations (e.g., V1016I) or amplified P450 alleles with wild-type sequences, driving susceptibility through pest populations—a CRISPR-based allelic-drive in D. melanogaster reversed a resistant kdr mutation to the wild-type form in ~80–87% of the population within 8–10 generations [307]. (3) Machine learning models that integrate real-time genomic data (e.g., allele frequencies of resistance loci), insecticide usage patterns, and climate projections have demonstrated high predictive accuracy in identifying resistance hotspots in malaria vectors across Africa. Ensemble models, such as XGBoost, Random Forest, and Bayesian Generalized Additive Models, have been employed to predict insecticide resistance phenotypes, utilizing bioassay mortality data and a suite of covariates. These models have been validated using out-of-sample data, showing robust performance in forecasting resistance trends and informing targeted vector control strategies [308].

6.3. Biopesticides: Eco-Friendly Alternatives

Biopesticides derived from plants, microorganisms, and natural enemies offer eco-compatible alternatives to synthetic chemicals. These agents are often species-specific, biodegradable, and pose minimal risks to humans and non-target organisms [309,310]. Plant-based biopesticides—such as azadirachtin, pyrethrins, rotenone, ryania, nicotine, and sabadilla—exert neurotoxic, antifeedant, or growth-disrupting effects on pests. They are species-specific, biodegradable, and safer for non-target organisms, making them effective, eco-friendly alternatives in IPM [311]. Moreover, the widespread use of botanical products has reinvigorated traditional pest control knowledge and offers viable solutions for organic agriculture.

6.4. Biological Control: Natural Regulation of Pest Populations

Biological control employs natural enemies—predators, parasitoids, and entomopathogens—to suppress pest populations without chemical inputs. This approach is integral to IPM due to its high specificity, environmental safety, and sustainable efficacy [312,313]. Key entomopathogenic viruses include baculoviruses such as H. armigera nucleopolyhedrovirus (HaNPV) and Spodoptera litura nucleopolyhedrovirus (SliNPV), widely utilized across Asia and South America [314,315]. Entomopathogenic fungi, notably B. bassiana and Metarhizium anisopliae, infect over 700 pest species under optimal conditions [316]. Inoculative releases of parasitoids and predators in greenhouse and field environments provide long-term pest suppression [317]. Integrating biological control with biopesticides and selective chemical applications enhances pest management resilience while promoting biodiversity conservation [318].

7. Conclusions

The relentless application of insecticides has shaped a powerful evolutionary force, enabling insect pests to rapidly adapt through genetic and biochemical innovations. Resistance to insecticides evolves through two principal pathways: (1) target-site modifications, which hinder the binding of insecticides to their molecular receptors (e.g., nAChRs, AChE, VGSCs, GABA receptors), and (2) metabolic detoxification mechanisms, where the upregulation or amplification of key enzyme systems (esterases, CYPs, GSTs) neutralizes toxic compounds. These resistance strategies are often polygenic, heritable, and strongly influenced by selection pressure from the overuse or misuse of insecticides. This evolutionary pressure has led to a rich array of resistance mechanisms—behavioral shifts, thicker cuticles, mutated target sites, and enhanced detoxification pathways. While chemical pesticides remain indispensable tools for pest control, their diminishing efficacy, ecotoxicological consequences, and non-target impacts underscore the urgent need for integrated, sustainable alternatives. Recent studies confirm that resistance continues to escalate, with new alleles and cross-resistances appearing each year. For example, pyrethroid resistance in mosquitoes has intensified to the point that only new-generation interventions remain effective. Resistance management strategies offer tools to sustain insecticide utility. Integrated approaches, such as combining chemical and non-chemical tactics, rotating and mixing insecticides by mode of action, and conserving refugia, have been repeatedly shown to delay resistance emergence. Ongoing monitoring and adaptive management are key: resistance patterns vary over time and space, so tactics must be adjusted accordingly. Cross-sector learning is also valuable; for instance, the success of IPM in crops can inform vector control, and lessons from mosquito net rotations can inspire better practices in agricultural irrigation. As the post-insecticide era demands a fundamental reimagining of pest control, from reactive chemistry to proactive genetic disruption, leveraging evolutionary principles such as trade-offs, epistasis, and gene drive dynamics can help us outpace adaptation. Priority actions include establishing global consortia for real-time resistance allele tracking, fast-tracking regulatory processes for RNAi and nanotech tools, and developing open-source AI platforms to model resistance evolution. These transformative, ecology-guided strategies will be essential in securing food systems and public health, ensuring that insecticides remain effective for the foreseeable future.

Author Contributions

Conceptualization, Y.A.N., I.E. and A.M.; methodology, A.M. and I.E.; writing—original draft preparation, Y.A.N., N.M.F. and A.M.; writing—review and editing, J.P.G., I.E., A.M., R.K.A.-A., H.R.E.-S., A.M.A., H.M.A. and H.A.A.; visualization, A.M. and R.K.A.-A.; supervision, J.P.G. and I.E.; funding acquisition, Y.A.N. and H.A.A. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge institutional support as detailed in the Acknowledgments.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors express their sincere appreciation to the Deanship of Research and Graduate Studies at King Khalid University for supporting this work through a Small Research Project (Grant No. RGP1/300/46). The authors also thank the Deanship of Scientific Research at King Saud University for funding provided through the Vice Deanship of Scientific Research Chairs, Chair of Date Palm Research. Curie for Microsoft Word and Scribbr Proofreader were used to improve the grammar, clarity, and readability of the manuscript. Subsequent critical reading was conducted by John P. Giesy and Ioannis Eleftherianos, co-authors of the manuscript. All authors reviewed and approved the final version to ensure accuracy and consistency. The authors also acknowledge the constructive feedback and valuable insights provided by various colleagues during the development of this review. All illustrations featured in this review were designed and prepared by co-author Amr Mohamed. Finally, the authors sincerely thank the three anonymous reviewers for their insightful and constructive comments, which significantly enriched the depth and clarity of the manuscript.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AChEAcetylcholinesterase
AIArtificial Intelligence
BtBacillus thuringiensis
CESsCarboxylesterases (commonly abbreviated as CESs or CarEs)
CRISPR-Cas9Clustered Regularly Interspaced Short Palindromic Repeats-CRISPR associated protein 9
CSPsChemosensory Proteins
CYPsCytochrome P450 Monooxygenases
DDTDichlorodiphenyltrichloroethane
DEMDiethyl Maleate
dsRNADouble-Stranded RNA
FAWFall Armyworm (Spodoptera frugiperda)
GABAGamma-Aminobutyric Acid
GPIRMWHO Global Plan for Insecticide Resistance Management
GSTsGlutathione-S-Transferases
HaNPVHelicoverpa armigera Nucleopolyhedrovirus
IPMIntegrated Pest Management
IRMInsecticide Resistance Management
kdrKnockdown Resistance
LLINsLong-Lasting Insecticidal Nets
lncRNAsLong Non-Coding RNAs
MACEModified Acetylcholinesterase
miRNAsMicroRNAs
MoAMode of Action
nAChRsNicotinic Acetylcholine Receptors
NPVNucleopolyhedrovirus
OBPsOdorant-Binding Proteins
OPOrganophosphates
PBOPiperonyl Butoxide
PIPsPlant-Incorporated Protectants
qPCRQuantitative Polymerase Chain Reaction
RDLResistance to Dieldrin
RNAiRNA Interference
RyRRyanodine Receptor
SliNPVSpodoptera litura Nucleopolyhedrovirus
VGSCVoltage-Gated Sodium Channel
WHOWorld Health Organization

References

  1. Ádám, B.; Cocco, P.; Godderis, L. Hazardous Effects of Pesticides on Human Health. Toxics 2024, 12, 186. [Google Scholar] [CrossRef]
  2. Wan, N.-F.; Fu, L.; Dainese, M.; Kiær, L.P.; Hu, Y.-Q.; Xin, F.; Goulson, D.; Woodcock, B.A.; Vanbergen, A.J.; Spurgeon, D.J.; et al. Pesticides have negative effects on non-target organisms. Nat. Commun. 2025, 16, 1360. [Google Scholar] [CrossRef] [PubMed]
  3. Savary, S.; Willocquet, L.; Pethybridge, S.J.; Esker, P.; McRoberts, N.; Nelson, A. The global burden of pathogens and pests on major food crops. Nat. Ecol. Evol. 2019, 3, 430–439. [Google Scholar] [CrossRef] [PubMed]
  4. Tonnang, H.E.; Sokame, B.M.; Abdel-Rahman, E.M.; Dubois, T. Measuring and modelling crop yield losses due to invasive insect pests under climate change. Curr. Opin. Insect Sci. 2022, 50, 100873. [Google Scholar] [CrossRef]
  5. Sparks, T.C.; Crossthwaite, A.J.; Nauen, R.; Banba, S.; Cordova, D.; Earley, F.; Ebbinghaus-Kintscher, U.; Fujioka, S.; Hirao, A.; Karmon, D.; et al. Insecticides, biologics and nematicides: Updates to IRAC’s mode of action classification—A tool for resistance management. Pestic. Biochem. Physiol. 2020, 167, 104587. [Google Scholar] [CrossRef]
  6. Bradshaw, C.J.; Leroy, B.; Bellard, C.; Roiz, D.; Albert, C.; Fournier, A.; Barbet-Massin, M.; Salles, J.M.; Simard, F.; Courchamp, F. Massive yet grossly underestimated global costs of invasive insects. Nat. Commun. 2016, 7, 12986. [Google Scholar] [CrossRef] [PubMed]
  7. Food and Agriculture Organization of the United Nations. 2023 Annual Meeting on the Implementation of FAO Global Action (GA) for Fall Armyworm (FAW) Control in Africa—Report; FAO: Rome, Italy, 2023; Available online: https://openknowledge.fao.org/handle/20.500.14283/cd0185en (accessed on 2 August 2025).
  8. Togola, A.; Beyene, Y.; Bocco, R.; Tepa-Yotto, G.; Gowda, M.; Too, A.; Boddupalli, P. Fall armyworm (Spodoptera frugiperda) in Africa: Insights into biology, ecology and impact on staple crops, food systems and management approaches. Front. Agron. 2025, 7, 1538198. [Google Scholar] [CrossRef]
  9. Quandahor, P.; Kim, L.; Kim, M.; Lee, K.; Kusi, F.; Jeong, I. Effects of Agricultural Pesticides on Decline in Insect Species and Individual Numbers. Environments 2024, 11, 182. [Google Scholar] [CrossRef]
  10. Food and Agriculture Organization of the United Nations (FAO). FAOSTAT Database: Insecticides Use Data. 2025. Available online: http://www.fao.org/faostat/en/#data (accessed on 2 August 2025).
  11. IRAC (Insecticide Resistance Action Committee): Introduction to Resistance. 2025. Available online: https://irac-online.org/training-centre/resistance/ (accessed on 4 August 2025).
  12. Rezende-Teixeira, P.; Dusi, R.G.; Jimenez, P.C.; Espindola, L.S.; Costa-Lotufo, L.V. What Can We Learn from Commercial Insecticides? Efficacy, Toxicity, Environmental Impacts, and Future Developments. Environ. Pollut. 2022, 300, 118983. [Google Scholar] [CrossRef]
  13. Stratonovitch, P.; Elias, J.; Denholm, I.; Slater, R.; Semenov, M.A. An Individual-Based Model of the Evolution of Pesticide Resistance in Heterogeneous Environments: Control of Meligethes aeneus Population in Oilseed Rape Crops. PLoS ONE 2014, 9, e115631. [Google Scholar] [CrossRef]
  14. Gould, F.; Brown, Z.S.; Kuzma, J. Wicked Evolution: Can We Address the Sociobiological Dilemma of Pesticide Resistance? Science 2018, 360, 728–732. [Google Scholar] [CrossRef]
  15. Bass, C.; Puinean, A.M.; Andrews, M.; Cutler, P.; Daniels, M.; Elias, J.; Paul, V.L.; Crossthwaite, A.J.; Denholm, I.; Field, L.M.; et al. Mutation of a Nicotinic Acetylcholine Receptor β Subunit Is Associated with Resistance to Neonicotinoid Insecticides in the Aphid Myzus persicae. BMC Neurosci. 2011, 12, 51. [Google Scholar] [CrossRef]
  16. Zhou, W.; Arcot, Y.; Medina, R.F.; Bernal, J.; Cisneros-Zevallos, L.; Akbulut, M.E.S. Integrated Pest Management: An Update on the Sustainability Approach to Crop Protection. ACS Omega 2024, 9, 41130–41147. [Google Scholar] [CrossRef] [PubMed]
  17. Nauen, R.; Slater, R.; Sparks, T.C.; Elbert, A.; Mccaffery, A. IRAC: Insecticide resistance and mode-of-action classification of insecticides. In Modern Crop Protection Compounds, Volume 3: Insecticides, 3rd ed.; Jeschke, P., Witschel, M., Krämer, W., Schirmer, U., Eds.; Wiley-VCH: Weinheim, Germany, 2019; Chapter 28. [Google Scholar] [CrossRef]
  18. Sparks, T.C.; Storer, N.; Porter, A.; Slater, R.; Nauen, R. Insecticide Resistance Management and Industry: The Origins and Evolution of the Insecticide Resistance Action Committee (IRAC) and the Mode of Action Classification Scheme. Pest Manag. Sci. 2021, 77, 2609–2619. [Google Scholar] [CrossRef] [PubMed]
  19. Russell, T.L.; Govella, N.J.; Azizi, S.; Drakeley, C.J.; Kachur, S.P.; Killeen, G.F. Increased Proportions of Outdoor Feeding among Residual Malaria Vector Populations Following Increased Use of Insecticide-Treated Nets in Rural Tanzania. Malar. J. 2011, 10, 80. [Google Scholar] [CrossRef] [PubMed]
  20. Clements, J.; Olson, J.M.; Sanchez-Sedillo, B.; Bradford, B.; Groves, R.L. Changes in emergence phenology, fatty acid composition, and xenobiotic-metabolizing enzyme expression is associated with increased insecticide resistance in the Colorado potato beetle. Arch. Insect Biochem. Physiol. 2020, 103, e21630. [Google Scholar] [CrossRef]
  21. Sarfraz, M.; Dosdall, L.M.; Keddie, B.A. Evidence for Behavioural Resistance by the Diamondback Moth, Plutella xylostella (L.). J. Appl. Entomol. 2005, 129, 340–341. [Google Scholar] [CrossRef]
  22. Bass, C.; Jones, C.M. Mosquitoes boost body armor to resist insecticide attack. Proc. Natl. Acad. Sci. USA 2016, 113, 9145–9147. [Google Scholar] [CrossRef]
  23. Ahmad, M.; Denholm, I.; Bromilow, R.H. Delayed cuticular penetration and enhanced metabolism of deltamethrin in pyrethroid-resistant strains of Helicoverpa armigera from China and Pakistan. Pest Manag. Sci. 2006, 62, 805–810. [Google Scholar] [CrossRef]
  24. Reddy, M.R.; Godoy, A.; Dion, K.; Matias, A.; Callender, K.; Kiszewski, A.E.; Kleinschmidt, I.; Ridl, F.C.; Powell, J.R.; Caccone, A.; et al. Insecticide resistance allele frequencies in Anopheles gambiae before and after anti-vector interventions in continental Equatorial Guinea. Am. J. Trop. Med. Hyg. 2013, 88, 897–907. [Google Scholar] [CrossRef]
  25. Djogbénou, L.; Chandre, F.; Berthomieu, A.; Dabiré, R.; Koffi, A.; Alout, H.; Weill, M. Evidence of introgression of the ace-1(R) mutation and of the ace-1 duplication in West African Anopheles gambiae s. s. PLoS ONE 2008, 3, e2172. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Meng, X.; Yang, Y.; Li, H.; Wang, X.; Yang, B.; Zhang, J.; Li, C.; Millar, N.S.; Liu, Z. Synergistic and compensatory effects of two point mutations conferring target-site resistance to fipronil in the insect GABA receptor RDL. Sci. Rep. 2016, 6, 32335. [Google Scholar] [CrossRef] [PubMed]
  27. Riveron, J.M.; Irving, H.; Ndula, M.; Barnes, K.G.; Ibrahim, S.S.; Paine, M.J.; Wondji, C.S. Directionally selected cytochrome P450 alleles are driving the spread of pyrethroid resistance in the major malaria vector Anopheles funestus. Proc. Natl. Acad. Sci. USA 2013, 110, 252–257. [Google Scholar] [CrossRef]
  28. Poulton, B.C.; Colman, F.; Anthousi, A.; Sattelle, D.B.; Lycett, G.J. Aedes aegypti CCEae3A carboxylase expression confers carbamate, organophosphate and limited pyrethroid resistance in a model transgenic mosquito. PLoS Negl. Trop. Dis. 2024, 18, e0011595. [Google Scholar] [CrossRef]
  29. Daborn, P.J.; Yen, J.L.; Bogwitz, M.R.; Le Goff, G.; Feil, E.; Jeffers, S.; Tijet, N.; Perry, T.; Heckel, D.; Batterham, P.; et al. A single p450 allele associated with insecticide resistance in Drosophila. Science 2002, 297, 2253–2256. [Google Scholar] [CrossRef]
  30. Lu, X.; Simma, E.A.; Spanoghe, P.; Van Leeuwen, T.; Dermauw, W. Recombinant expression and characterization of GSTd3 from a resistant population of Anopheles arabiensis and comparison of DDTase activity with GSTe2. Pestic. Biochem. Physiol. 2023, 192, 105397. [Google Scholar] [CrossRef]
  31. Taillebois, E.; Cartereau, A.; Jones, A.K.; Thany, S.H. Neonicotinoid insecticides mode of action on insect nicotinic acetylcholine receptors using binding studies. Pestic. Biochem. Physiol. 2018, 151, 59–66. [Google Scholar] [CrossRef] [PubMed]
  32. Matsuda, K.; Ihara, M.; Sattelle, D.B. Neonicotinoid Insecticides: Molecular Targets, Resistance, and Toxicity. Annu. Rev. Pharmacol. Toxicol. 2020, 60, 241–255. [Google Scholar] [CrossRef] [PubMed]
  33. Xu, X.; Ding, Q.; Wang, X.; Wang, R.; Ullah, F.; Gao, X.; Song, D. V101I and R81T mutations in the nicotinic acetylcholine receptor β1 subunit are associated with neonicotinoid resistance in Myzus persicae. Pest Manag. Sci. 2022, 78, 1500–1507. [Google Scholar] [CrossRef]
  34. Mottet, C.; Fontaine, S.; Caddoux, L.; Brazier, C.; Mahéo, F.; Simon, J.C.; Micoud, A.; Roy, L. Assessment of the Dominance Level of the R81T Target Resistance to Two Neonicotinoid Insecticides in Myzus persicae (Hemiptera: Aphididae). J. Econ. Entomol. 2016, 109, 2182–2189. [Google Scholar] [CrossRef]
  35. Mottet, C.; Caddoux, L.; Fontaine, S.; Plantamp, C.; Bass, C.; Barrès, B. Myzus persicae resistance to neonicotinoids—Unravelling the contribution of different mechanisms to phenotype. Pest Manag. Sci. 2024, 80, 5852–5863. [Google Scholar] [CrossRef] [PubMed]
  36. Beck, M.E.; Riplinger, C.; Neese, F.; Bistoni, G. Unraveling individual host–guest interactions in molecular recognition from first principles quantum mechanics: Insights into the nature of nicotinic acetylcholine receptor agonist binding. J. Comput. Chem. 2021, 42, 293–302. [Google Scholar] [CrossRef]
  37. Zhang, K.; Chen, L.; Chen, J.; Huang, H.; Liu, K.; Zhang, Y.; Yang, J.; Wu, S. Mutation V65I in the β1 Subunit of the Nicotinic Acetylcholine Receptor Confers Neonicotinoid and Sulfoxaflor Resistance in Insects. J. Agric. Food Chem. 2024, 72, 5671–5681. [Google Scholar] [CrossRef]
  38. Yin, C.; O’Reilly, A.O.; Liu, S.N.; Du, T.H.; Gong, P.P.; Zhang, C.J.; Wei, X.G.; Yang, J.; Huang, M.J.; Fu, B.L.; et al. Dual mutations in the whitefly nicotinic acetylcholine receptor β1 subunit confer target-site resistance to multiple neonicotinoid insecticides. PLoS Genet. 2024, 20, e1011163. [Google Scholar] [CrossRef]
  39. Markussen, M.D.; Kristensen, M. Low Expression of Nicotinic Acetylcholine Receptor Subunit Mdα2 in Neonicotinoid-Resistant Strains of Musca domestica L. Pest Manag. Sci. 2010, 66, 1257–1262. [Google Scholar] [CrossRef]
  40. Zhang, K.; Chen, J.; Huang, H.; Wen, H.; Yang, J.; Geng, J.; Wu, S. The Amino Acid Ser223 Acts as a Key Site for the Binding of Thrips palmi α1 Nicotinic Acetylcholine Receptor to Neonicotinoid Insecticides. Pestic. Biochem. Physiol. 2025, 213, 106484. [Google Scholar] [CrossRef] [PubMed]
  41. Zimmer, C.T.; Garrood, W.T.; Puinean, A.M.; Eckel-Zimmer, M.; Williamson, M.S.; Davies, T.G.; Bass, C. A CRISPR/Cas9 mediated point mutation in the alpha 6 subunit of the nicotinic acetylcholine receptor confers resistance to spinosad in Drosophila melanogaster. Insect Biochem. Mol. Biol. 2016, 73, 62–69. [Google Scholar] [CrossRef] [PubMed]
  42. Watson, G.B.; Siebert, M.W.; Wang, N.X.; Loso, M.R.; Sparks, T.C. Sulfoxaflor – A Sulfoximine Insecticide: Review and Analysis of Mode of Action, Resistance and Cross-Resistance. Pestic. Biochem. Physiol. 2021, 178, 104924. [Google Scholar] [CrossRef]
  43. Ma, K.; Tang, Q.; Zhang, B.; Liang, P.; Wang, B.; Gao, X. Overexpression of Multiple Cytochrome P450 Genes Associated with Sulfoxaflor Resistance in Aphis gossypii Glover. Pestic. Biochem. Physiol. 2019, 157, 204–210. [Google Scholar] [CrossRef]
  44. Zhao, X.; Chen, X.; He, M.; Wu, S.; Shi, Y.; Luo, D.; Zhang, H.; Wang, Z.; Wan, H.; Li, R.; et al. miRNAs modulate altered expression of cytochrome P450s and nicotinic acetylcholine receptor subunits conferring both metabolic and target resistance to sulfoxaflor in Nilaparvata lugens (Stål). Int. J. Biol. Macromol. 2025, 290, 138992. [Google Scholar] [CrossRef]
  45. Longhurst, C.; Babcock, J.M.; Denholm, I.; Gorman, K.; Thomas, J.D.; Sparks, T.C. Cross-Resistance Relationships of the Sulfoximine Insecticide Sulfoxaflor with Neonicotinoids and Other Insecticides in the Whiteflies Bemisia tabaci and Trialeurodes vaporariorum. Pest Manag. Sci. 2013, 69, 809–813. [Google Scholar] [CrossRef] [PubMed]
  46. Mezei, I.; Valverde-Garcia, P.; Siebert, M.W.; Gomez, L.E.; Torne, M.; Watson, G.B.; Raquel, A.M.; Fereres, A.; Sparks, T.C. Impact of the nicotinic acetylcholine receptor mutation R81T on the response of European Myzus persicae populations to imidacloprid and sulfoxaflor in laboratory and in the field. Pestic. Biochem. Physiol. 2022, 187, 105187. [Google Scholar] [CrossRef]
  47. Beckingham, C.; Phillips, J.; Gill, M.; Crossthwaite, A.J. Investigating nicotinic acetylcholine receptor expression in neonicotinoid resistant Myzus persicae FRC. Pestic. Biochem. Physiol. 2013, 107, 293–298. [Google Scholar] [CrossRef]
  48. Samal, R.R.; Panmei, K.; Lanbiliu, P.; Kumar, S. Metabolic detoxification and ace-1 target site mutations associated with acetamiprid resistance in Aedes aegypti L. Front. Physiol. 2022, 13, 988907. [Google Scholar] [CrossRef]
  49. Lan, J.; Wang, Z.; Chen, Z.; Zhang, L.; Zhao, J.; Guan, Q.; Liao, C.; Liu, N.; Han, Q. Identification of the Aedes aegypti nAChR gene family and molecular target of spinosad. Pest Manag. Sci. 2021, 77, 1633–1641. [Google Scholar] [CrossRef]
  50. Fouet, C.; Pinch, M.J.; Ashu, F.A.; Ambadiang, M.M.; Bouaka, C.; Batronie, A.J.; Hernandez, C.A.; Rios, D.E.; Penlap-Beng, V.; Kamdem, C. Field-Evolved Resistance to Neonicotinoids in the Mosquito, Anopheles gambiae, Is Associated with Mutations of Nicotinic Acetylcholine Receptor Subunits Combined with Cytochrome P450-Mediated Detoxification. Pestic. Biochem. Physiol. 2024, 206, 106205. [Google Scholar] [CrossRef] [PubMed]
  51. Homem, R.A.; Buttery, B.; Richardson, E.; Tan, Y.; Field, L.M.; Williamson, M.S.; Davies, T.G.E. Evolutionary Trade-Offs of Insecticide Resistance—The Fitness Costs Associated with Target-Site Mutations in the nAChR of Drosophila melanogaster. Mol. Ecol. 2020, 29, 2661–2675. [Google Scholar] [CrossRef] [PubMed]
  52. Ureña, E.; Guillem-Amat, A.; Couso-Ferrer, F.; Beroiz, B.; Perera, N.; López-Errasquín, E.; Castañera, P.; Ortego, F.; Hernández-Crespo, P. Multiple mutations in the nicotinic acetylcholine receptor Ccα6 gene associated with resistance to spinosad in medfly. Sci. Rep. 2019, 9, 2961. [Google Scholar] [CrossRef] [PubMed]
  53. Ilias, A.; Lagnel, J.; Kapantaidaki, D.E.; Roditakis, E.; Tsigenopoulos, C.S.; Vontas, J.; Tsagkarakou, A. Transcription analysis of neonicotinoid resistance in Mediterranean (MED) populations of B. tabaci reveal novel cytochrome P450s, but no nAChR mutations associated with the phenotype. BMC Genom. 2015, 16, 939. [Google Scholar] [CrossRef] [PubMed]
  54. Silva, W.M.; Berger, M.; Bass, C.; Williamson, M.; Moura, D.M.; Ribeiro, L.M.; Siqueira, H.A. Mutation (G275E) of the nicotinic acetylcholine receptor α6 subunit is associated with high levels of resistance to spinosyns in Tuta absoluta (Meyrick) (Lepidoptera: Gelechiidae). Pestic. Biochem. Physiol. 2016, 131, 125432. [Google Scholar] [CrossRef]
  55. Zuo, Y.; Xue, Y.; Lu, W.; Ma, H.; Chen, M.; Wu, Y.; Yang, Y.; Hu, Z. Functional validation of nicotinic acetylcholine receptor (nAChR) α6 as a target of spinosyns in Spodoptera exigua utilizing the CRISPR/Cas9 system. Pest Manag. Sci. 2020, 76, 2415–2422. [Google Scholar] [CrossRef]
  56. Cai, Y.; Chen, H.; Hu, M.; Wang, X.; Zhang, L. Discovery of Novel Potential Insecticide-Resistance Mutations in Spodoptera frugiperda. Insects 2024, 15, 186. [Google Scholar] [CrossRef]
  57. Wang, J.; Wang, X.; Lansdell, S.J.; Zhang, J.; Millar, N.S.; Wu, Y. A three amino acid deletion in the transmembrane domain of the nicotinic acetylcholine receptor α6 subunit confers high-level resistance to spinosad in Plutella xylostella. Insect Biochem. Mol. Biol. 2016, 71, 29–36. [Google Scholar] [CrossRef]
  58. Mocchetti, A.; Nikoloudi, A.A.; Vontas, J.; De Rouck, S.; Van Leeuwen, T. CRISPR/Cas9 Knock-Out of nAChR α6 Confers Resistance to Spinosyns in Frankliniella occidentalis and Is Associated with a Higher Fitness Cost than Target-Site Mutation G275E. Pestic. Biochem. Physiol. 2025, 212, 106455. [Google Scholar] [CrossRef] [PubMed]
  59. Bao, W.X.; Narai, Y.; Nakano, A.; Kaneda, T.; Murai, T.; Sonoda, S. Spinosad resistance of melon thrips, Thrips palmi, is conferred by G275E mutation in α6 subunit of nicotinic acetylcholine receptor and cytochrome P450 detoxification. Pestic. Biochem. Physiol. 2014, 112, 51–55. [Google Scholar] [CrossRef] [PubMed]
  60. Hirata, K.; Jouraku, A.; Kuwazaki, S.; Kanazawa, J.; Iwasa, T. The R81T mutation in the nicotinic acetylcholine receptor of Aphis gossypii is associated with neonicotinoid insecticide resistance with differential effects for cyano- and nitro-substituted neonicotinoids. Pestic. Biochem. Physiol. 2017, 143, 57–65. [Google Scholar] [CrossRef]
  61. Zhang, Y.-C.; Zhuang, Z.-X.; Zhang, F.; Song, X.-Y.; Ye, W.-N.; Wu, S.-F.; Bass, C.; O’Reilly, A.O.; Gao, C. Contribution of Nilaparvata lugens Nicotinic Acetylcholine Receptor Subunits Toward Triflumezopyrim and Neonicotinoid Susceptibility. Environ. Sci. Technol. 2025, 59, 7054–7065. [Google Scholar] [CrossRef] [PubMed]
  62. 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]
  63. Liu, N. Insecticide Resistance in Mosquitoes: Impact, Mechanisms, and Research Directions. Annu. Rev. Entomol. 2015, 60, 537–559. [Google Scholar] [CrossRef]
  64. Rinkevich, F.D.; Su, C.; Lazo, T.A.; Hawthorne, D.J.; Tingey, W.M.; Naimov, S.; Scott, J.G. Multiple Evolutionary Origins of Knockdown Resistance (kdr) in Pyrethroid-Resistant Colorado Potato Beetle, Leptinotarsa decemlineata. Pestic. Biochem. Physiol. 2012, 104, 192–200. [Google Scholar] [CrossRef]
  65. Sun, H.; Tong, K.P.; Kasai, S.; Scott, J.G. Overcoming Super-Knock Down Resistance (Super-Kdr) Mediated Resistance: Multi-Halogenated Benzyl Pyrethroids Are More Toxic to Super-Kdr Than Kdr House Flies. Insect Mol. Biol. 2016, 25, 126–137. [Google Scholar] [CrossRef] [PubMed]
  66. von Stein, R.T.; Silver, K.S.; Soderlund, D.M. Indoxacarb, Metaflumizone, and Other Sodium Channel Inhibitor Insecticides: Mechanism and Site of Action on Mammalian Voltage-Gated Sodium Channels. Pestic. Biochem. Physiol. 2013, 106, 101–112. [Google Scholar] [CrossRef]
  67. Roditakis, E.; Mavridis, K.; Riga, M.; Vasakis, E.; Morou, E.; Rison, J.L.; Vontas, J. Identification and Detection of Indoxacarb Resistance Mutations in the Para Sodium Channel of the Tomato Leafminer, Tuta absoluta. Pest Manag. Sci. 2017, 73, 1679–1688. [Google Scholar] [CrossRef] [PubMed]
  68. Xiao, Y.; Wang, X.; Li, Z.; Lei, C.; Wang, S. Insecticidal potential and risk assessment of diamide pesticides against Spodoptera frugiperda in maize crops. Ecotoxicol. Environ. Saf. 2024, 282, 116682. [Google Scholar] [CrossRef]
  69. Du, J.; Fu, Y. Diamide insecticides targeting insect ryanodine receptors: Mechanism and application prospect. Biochem. Biophys. Res. Commun. 2023, 670, 19–26. [Google Scholar] [CrossRef] [PubMed]
  70. Chen, H.-L.; Hasnain, A.; Cheng, Q.-H.; Xia, L.-J.; Cai, Y.-H.; Hu, R.; Gong, C.-W.; Liu, X.-M.; Pu, J.; Zhang, L.; et al. Resistance Monitoring and Mechanism in the Fall Armyworm Spodoptera frugiperda (Lepidoptera: Noctuidae) for Chlorantraniliprole from Sichuan Province, China. Front. Physiol. 2023, 14, 1180655. [Google Scholar] [CrossRef] [PubMed]
  71. Li, K.; Wang, K.B.; Tao, L.H.; Li, L.; Li, J.J.; Wu, W.W.; Xiao, C.; Ye, M. Analysis of binding modes and resistance risk between ryanodine receptor of Spodoptera frugiperda and diamide. J. Pestic. Sci. 2021, 23, 856–868. [Google Scholar]
  72. Ma, R.; Haji-Ghassemi, O.; Ma, D.; Jiang, H.; Lin, L.; Yao, L.; Samurkas, A.; Li, Y.; Wang, Y.; Cao, P. Structural basis for diamide modulation of ryanodine receptor. Nat. Chem. Biol. 2020, 16, 1246–1254. [Google Scholar] [CrossRef]
  73. Lv, Y.; Li, J.; Yan, K.; Ding, Y.; Gao, X.; Bi, R.; Zhang, H.; Pan, Y.; Shang, Q. Functional Characterization of ABC Transporters Mediates Multiple Neonicotinoid Resistance in a Field Population of Aphis gossypii Glover. Pestic. Biochem. Physiol. 2022, 188, 105264. [Google Scholar] [CrossRef]
  74. Douris, V.; Steinbach, D.; Panteleri, R.; Livadaras, I.; Pickett, J.A.; Van Leeuwen, T.; Nauen, R.; Vontas, J. Resistance Mutation Conserved Between Insects and Mites Unravels the Benzoylurea Insecticide Mode of Action on Chitin Biosynthesis. Proc. Natl. Acad. Sci. USA 2016, 113, 14692–14697. [Google Scholar] [CrossRef]
  75. Suzuki, Y.; Shiotsuki, T.; Jouraku, A.; Miura, K.; Minakuchi, C. Benzoylurea Resistance in Western Flower Thrips Frankliniella occidentalis (Thysanoptera: Thripidae): The Presence of a Point Mutation in Chitin Synthase 1. J. Pestic. Sci. 2017, 42, 93–96. [Google Scholar] [CrossRef] [PubMed]
  76. Fotakis, E.A.; Mastrantonio, V.; Grigoraki, L.; Porretta, D.; Puggioli, A.; Chaskopoulou, A.; Osório, H.; Weill, M.; Bellini, R.; Urbanelli, S.; et al. Identification and Detection of a Novel Point Mutation in the Chitin Synthase Gene of Culex pipiens Associated with Diflubenzuron Resistance. PLoS Negl. Trop. Dis. 2020, 14, e0008284. [Google Scholar] [CrossRef] [PubMed]
  77. Li, Y.; Li, R.; Shao, H.; Liu, Z.; Gao, X.; Tian, Z.; Zhang, Y.; Liu, J. Unraveling Key Amino Acid Residues Crucial for PxGSTs1 Conferring Benzoylurea Insecticide Resistance in Plutella xylostella. J. Agric. Food Chem. 2024, 72, 25549–25559. [Google Scholar] [CrossRef]
  78. Koyama, T.; Saeed, U.; Rewitz, K.; Halberg, K.V. The integrative physiology of hormone signaling: Insights from insect models. Physiology 2025, 40, 343–362. [Google Scholar] [CrossRef]
  79. Charles, J.P.; Iwema, T.; Epa, V.C.; Takaki, K.; Rynes, J.; Jindra, M. Ligand-binding properties of a juvenile hormone receptor, Methoprene-tolerant. Proc. Natl. Acad. Sci. USA 2011, 108, 21128–21133. [Google Scholar] [CrossRef]
  80. Ononamadu, C.J.; Abdalla, M.; Ihegboro, G.O.; Li, J.; Owolarafe, T.A.; John, T.D.; Tian, Q. In silico identification and study of potential anti-mosquito juvenile hormone binding protein (MJHBP) compounds as candidates for dengue virus-vector insecticides. Biochem. Biophys. Rep. 2021, 28, 101178. [Google Scholar] [CrossRef]
  81. Barry, J.; Wang, S.; Wilson, T.G. Overexpression of Methoprene-tolerant, a Drosophila melanogaster gene that is critical for juvenile hormone action and insecticide resistance. Insect Biochem. Mol. Biol. 2008, 38, 346–353. [Google Scholar] [CrossRef]
  82. Liu, Y.; Sheng, Z.; Liu, H.; Wen, D.; He, Q.; Wang, S.; Shao, W.; Jiang, R.J.; An, S.; Sun, Y.; et al. Juvenile hormone counteracts the bHLH-PAS transcription factors MET and GCE to prevent caspase-dependent programmed cell death in Drosophila. J. Biol. Chem. 2009, 136, 2015–2025. [Google Scholar] [CrossRef]
  83. Jindra, M.; Uhlirova, M.; Charles, J.-P.; Smykal, V.; Hill, R.J. Genetic evidence for the function of the bHLH-PAS protein Gce/Met as a juvenile hormone receptor. PLoS Genet. 2015, 11, e1005394. [Google Scholar] [CrossRef]
  84. Yang, X.; Wei, X.; Yang, J.; Du, T.; Yin, C.; Fu, B.; Huang, M.; Liang, J.; Gong, P.; Liu, S.; et al. Epitranscriptomic regulation of insecticide resistance. Sci. Adv. 2021, 7, eabe5903. [Google Scholar] [CrossRef] [PubMed]
  85. Bavithra, C.M.L.; Murugan, M.; Pavithran, S.; Naveena, K. Enthralling genetic regulatory mechanisms meddling insecticide resistance development in insects: Role of transcriptional and post-transcriptional events. Front. Mol. Biosci. 2023, 10, 1257859. [Google Scholar] [CrossRef]
  86. Mahalle, R.M.; Sun, W.; Posos-Parra, O.A.; Jung, S.; Mota-Sanchez, D.; Pittendrigh, B.R.; Seong, K.M. Identification of differentially expressed miRNAs associated with diamide detoxification pathways in Spodoptera frugiperda. Sci. Rep. 2024, 14, 4308. [Google Scholar] [CrossRef] [PubMed]
  87. Kumar, M.B.; Fujimoto, T.; Potter, D.W.; Deng, Q.; Palli, S.R. A single point mutation in ecdysone receptor leads to increased ligand specificity: Implications for gene switch applications. Proc. Natl. Acad. Sci. USA 2002, 99, 14710–14715. [Google Scholar] [CrossRef]
  88. Ullah, F.; Gul, H.; Tariq, K.; Hafeez, M.; Desneux, N. RNA interference-mediated silencing of ecdysone receptor (EcR) gene causes lethal and sublethal effects on melon aphid, Aphis gossypii. Entomol. Gen. 2022, 42, 791. [Google Scholar] [CrossRef]
  89. Saito, K.; Kanno, M.; Tanimoto, H.; Ichinose, T. Ecdysteroid-DopEcR signaling in neuronal and midgut cells mediates toxin avoidance and detoxification in Drosophila. Curr. Biol. 2025, 35, 3418–3428.e4. [Google Scholar] [CrossRef]
  90. Uchibori-Asano, M.; Uchiyama, T.; Jouraku, A.; Ozawa, A.; Akiduki, G.; Yamamura, K.; Shinoda, T. Tebufenozide resistance in the smaller tea tortrix, Adoxophyes honmai (Lepidoptera: Tortricidae): Establishment of a molecular diagnostic method based on EcR mutation and its application for field-monitoring. Appl. Entomol. Zool. 2019, 54, 223–230. [Google Scholar] [CrossRef]
  91. Gao, Y.; Liu, S.; Jia, Q.; Wu, L.; Yuan, D.; Li, E.Y.; Feng, Q.; Wang, G.; Palli, S.R.; Wang, J.; et al. Juvenile hormone membrane signaling phosphorylates USP and thus potentiates 20-hydroxyecdysone action in Drosophila. Sci. Bull. 2022, 67, 186–197. [Google Scholar] [CrossRef]
  92. Luo, J.; Ouyang, H.; Liu, S.; Wu, H.; Xiong, F.; Yang, T.; Li, H.; Li, X. Juvenile hormone signaling regulates apoptosis-mediated chlorantraniliprole sensitivity in Spodoptera frugiperda (Lepidoptera: Noctuidae). Pestic. Biochem. Physiol. 2025, 214, 106603. [Google Scholar] [CrossRef]
  93. Ling, L.; Raikhel, A.S. Cross-talk of insulin-like peptides, juvenile hormone, and 20-hydroxyecdysone in regulation of metabolism in the mosquito Aedes aegypti. Proc. Natl. Acad. Sci. USA 2021, 118, e2023470118. [Google Scholar] [CrossRef] [PubMed]
  94. Geng, D.Q.; Wang, X.L.; Lyu, X.Y.; Raikhel, A.S.; Zou, Z. Ecdysone-controlled nuclear receptor ERR regulates metabolic homeostasis in the disease vector mosquito Aedes aegypti. PLoS Genet. 2024, 20, e1011196. [Google Scholar] [CrossRef] [PubMed]
  95. Niu, Z.-L.; Wei, J.; Wang, Q.; Yang, Y.-B.; Gao, Y.-N.; Li, L.; Jia, J.; Guo, W.; Zhao, D. RNAi-mediated knockdown of nuclear receptors impairs cuticle formation by disrupting chitin metabolic pathway in Holotrichia oblita. J. Insect Physiol. 2025, 165, 104861. [Google Scholar] [CrossRef] [PubMed]
  96. Verlinden, H.; Vleugels, R.; Marchal, E.; Badisco, L.; Pflüger, H.J.; Blenau, W.; Vanden Broeck, J. The role of octopamine in locusts and other arthropods. J. Insect Physiol. 2010, 56, 854–867. [Google Scholar] [CrossRef] [PubMed]
  97. Roeder, T. The control of metabolic traits by octopamine and tyramine in invertebrates. J. Exp. Biol. 2020, 223, jeb194282. [Google Scholar] [CrossRef]
  98. Ohta, H.; Ozoe, Y. Molecular signalling, pharmacology, and physiology of octopamine and tyramine receptors as potential insect pest control targets. Adv. Insect Physiol. 2014, 46, 73–166. [Google Scholar] [CrossRef]
  99. Enan, E. Insecticidal activity of essential oils: Octopaminergic sites of action. Comp. Biochem. Physiol. C Toxicol. 2001, 130, 325–337. [Google Scholar] [CrossRef] [PubMed]
  100. Kostyukovsky, M.; Rafaeli, A.; Gileadi, C.; Demchenko, N.; Shaaya, E. Activation of octopaminergic receptors by essential oil constituents isolated from aromatic plants: Possible mode of action against insect pests. Pest Manag. Sci. 2002, 58, 1101–1106. [Google Scholar] [CrossRef]
  101. Ahmed, M.A.I.; Vogel, C.F.; Matsumura, F. Unique biochemical and molecular biological mechanism of synergistic actions of formamidine compounds on selected pyrethroid and neonicotinoid insecticides on the fourth instar larvae of Aedes aegypti (Diptera: Culicidae). Pest Biochem. Physiol. 2015, 120, 57–63. [Google Scholar] [CrossRef]
  102. Kita, T.; Hayashi, T.; Ohtani, T.; Takao, H.; Takasu, H.; Liu, G.; Ohta, H.; Ozoe, F.; Ozoe, Y. Amitraz and its metabolite differentially activate α- and β-adrenergic-like octopamine receptors. Pest Manag. Sci. 2017, 73, 984–990. [Google Scholar] [CrossRef]
  103. Monteiro, H.R.; Lemos, M.F.L.; Novais, S.C.; Soares, A.M.V.M.; Pestana, J.L.T. Amitraz toxicity to the midge Chironomus riparius: Life-history and biochemical responses. Chemosphere 2019, 221, 324–332. [Google Scholar] [CrossRef]
  104. Ahmed, M.I.I.; Vogel, C.F.A. The synergistic effect of octopamine receptor agonists on selected insect growth regulators on Culex quinquefasciatus Say (Diptera: Culicidae) mosquitoes. One Health 2020, 10, 100138. [Google Scholar] [CrossRef]
  105. Wang, Y.; Qi, Z.; Qi, M.; Hu, Z.; Wu, W. Effects of Periplocoside P from Periploca sepium on the Midgut Transmembrane Potential of Mythimna separata Larvae. Sci. Rep. 2016, 6, 36982. [Google Scholar] [CrossRef]
  106. da Silva Costa, M.; de Paula, S.O.; Martins, G.F.; Zanuncio, J.C.; Santana, A.E.G.; Serrão, J.E. Multiple Modes of Action of the Squamocin in the Midgut Cells of Aedes aegypti Larvae. PLoS ONE 2016, 11, e0160928. [Google Scholar] [CrossRef]
  107. Vidau, C.; González-Polo, R.A.; Niso-Santano, M.; Gómez-Sánchez, R.; Bravo-San Pedro, J.M.; Pizarro-Estrella, E.; Blasco, R.; Brunet, J.L.; Belzunces, L.P.; Fuentes, J.M. Fipronil is a powerful uncoupler of oxidative phosphorylation that triggers apoptosis in human neuronal cell line SHSY5Y. Neurotoxicology 2011, 32, 935–943. [Google Scholar] [CrossRef] [PubMed]
  108. Xu, T.; Pagadala, V.; Mueller, D.M. Understanding structure, function, and mutations in the mitochondrial ATP synthase. Microb. Cell 2015, 2, 105–125. [Google Scholar] [CrossRef]
  109. Demine, S.; Renard, P.; Arnould, T. Mitochondrial uncoupling: A key controller of biological processes in physiology and diseases. Cells 2019, 8, 795. [Google Scholar] [CrossRef] [PubMed]
  110. Che-Mendoza, A.; González-Olvera, G.; Medina-Barreiro, A.; Arisqueta-Chablé, C.; Bibiano-Marin, W.; Correa-Morales, F.; Kirstein, O.D.; Manrique-Saide, P.; Vazquez-Prokopec, G.M. Efficacy of targeted indoor residual spraying with the pyrrole insecticide chlorfenapyr against pyrethroid-resistant Aedes aegypti. PLoS Negl. Trop. Dis. 2021, 15, e0009822. [Google Scholar] [CrossRef]
  111. Zhou, X.; Lv, H.; Yang, C.; Jiang, D. ROS-Mediated COXI Overexpression by AMPKα Confers Resistance to Dimefluthrin and Cross-Resistance to Indoxacarb in Aedes albopictus. J. Agric. Food Chem. 2025, 73, 16736–16745. [Google Scholar] [CrossRef]
  112. Remnant, E.J.; Morton, C.J.; Daborn, P.J.; Lumb, C.; Yang, Y.T.; Ng, H.L.; Parker, M.W.; Batterham, P. The Role of Rdl in Resistance to Phenylpyrazoles in Drosophila melanogaster. Insect Biochem. Mol. Biol. 2014, 54, 11–21. [Google Scholar] [CrossRef]
  113. Gassel, M.; Wolf, C.; Noack, S.; Williams, H.; Ilg, T. The Novel Isoxazoline Ectoparasiticide Fluralaner: Selective Inhibition of Arthropod γ-Aminobutyric Acid- and l-Glutamate-Gated Chloride Channels and Insecticidal/Acaricidal Activity. Insect Biochem. Mol. Biol. 2014, 45, 111–124. [Google Scholar] [CrossRef]
  114. Nakao, T. Mechanisms of Resistance to Insecticides Targeting RDL GABA Receptors in Planthoppers. Neurotoxicology 2017, 60, 293–298. [Google Scholar] [CrossRef] [PubMed]
  115. Qiao, X.; Zhou, T.; Zhang, J.; Zhang, L.; Lu, Y.; Huang, J. Functional Validation of A2′N Mutation of the RDL GABA Receptor against Fipronil via Molecular Modeling and Genome Engineering in Drosophila. Pest Manag. Sci. 2024, 80, 1924–1929. [Google Scholar] [CrossRef]
  116. Xie, N.; Bickley, B.A.; Gross, A.D. GABA-Gated Chloride Channel Mutation (Rdl) Induces Cholinergic Physiological Compensation Resulting in Cross Resistance in Drosophila melanogaster. Pestic. Biochem. Physiol. 2024, 203, 105972. [Google Scholar] [CrossRef]
  117. Grau-Bové, X.; Tomlinson, S.; O’Reilly, A.O.; Harding, N.J.; Miles, A.; Kwiatkowski, D.; Donnelly, M.J.; Weetman, D.; Anopheles gambiae 1000 Genomes Consortium. Evolution of the Insecticide Target Rdl in African Anopheles Is Driven by Interspecific and Interkaryotypic Introgression. Mol. Biol. Evol. 2020, 37, 2900–2917. [Google Scholar] [CrossRef]
  118. Xie, N.; Gross, A.D. Muscarinic Acetylcholine Receptor Activation Synergizes the Knockdown and Toxicity of GABA-Gated Chloride Channel Insecticides. Pest Manag. Sci. 2022, 78, 4599–4607. [Google Scholar] [CrossRef]
  119. Ffrench-Constant, R.H.; Bass, C. Does Resistance Really Carry a Fitness Cost? Curr. Opin. Insect Sci. 2017, 21, 39–46. [Google Scholar] [CrossRef] [PubMed]
  120. Gomard, Y.; Alout, H.; Lebon, C.; Latreille, A.; Benlali, A.; Mavingui, P.; Tortosa, P.; Atyame, C. Fitness Costs Associated with a GABA Receptor Mutation Conferring Dieldrin Resistance in Aedes albopictus. Heredity 2022, 129, 273–280. [Google Scholar] [CrossRef]
  121. Mastrantonio, V.; Porretta, D.; Lucchesi, V.; Güz, N.; Çağatay, N.S.; Bellini, R.; Vontas, J.; Urbanelli, S. Evolution of Adaptive Variation in the Mosquito Culex pipiens: Multiple Independent Origins of Insecticide Resistance Mutations. Insects 2021, 12, 676. [Google Scholar] [CrossRef]
  122. Hobbs, N.P.; Weetman, D.; Hastings, I.M. Insecticide Resistance Management Strategies for Public Health Control of Mosquitoes Exhibiting Polygenic Resistance: A Comparison of Sequences, Rotations, and Mixtures. Evol. Appl. 2023, 16, 936–959. [Google Scholar] [CrossRef] [PubMed]
  123. Chamnanya, S.; Kiddela, B.; Saingamsook, J.; Nachaiwieng, W.; Lumjuan, N.; Somboon, P.; Yanola, J. Overexpression of Multiple Cytochrome P450 Genes with and without Knockdown Resistance Mutations Confers High Resistance to Deltamethrin in Culex quinquefasciatus. Infect. Dis. Poverty 2025, 14, 2. [Google Scholar] [CrossRef] [PubMed]
  124. Casida, J.E.; Bryant, R.J. The ABCs of Pesticide Toxicology: Amounts, Biology, and Chemistry. Toxicol. Res. 2017, 6, 755–763. [Google Scholar] [CrossRef]
  125. Casida, J.E. Pesticide Interactions: Mechanisms, Benefits, and Risks. J. Agric. Food Chem. 2017, 65, 4553–4561. [Google Scholar] [CrossRef] [PubMed]
  126. Amezian, D.; Nauen, R.; Le Goff, G. Transcriptional Regulation of Xenobiotic Detoxification Genes in Insects—An Overview. Pestic. Biochem. Physiol. 2021, 174, 104822. [Google Scholar] [CrossRef] [PubMed]
  127. Rivi, M.; Monti, V.; Mazzoni, E.; Cassanelli, S.; Panini, M.; Anaclerio, M.; Cigolini, M.; Corradetti, B.; Bizzaro, D.; Mandrioli, M.; et al. A1-3 Chromosomal Translocations in Italian Populations of the Peach Potato Aphid Myzus persicae (Sulzer) Not Linked to Esterase-Based Insecticide Resistance. Bull. Entomol. Res. 2013, 103, 278–285. [Google Scholar] [CrossRef]
  128. Lan, W.S.; Cong, J.; Jiang, H.; Jiang, S.R.; Qiao, C.L. Expression and Characterization of Carboxylesterase E4 Gene from Peach–Potato Aphid (Myzus persicae) for Degradation of Carbaryl and Malathion. Biotechnol. Lett. 2005, 27, 1141–1146. [Google Scholar] [CrossRef] [PubMed]
  129. Wei, D.D.; He, W.; Miao, Z.Q.; Tu, Y.Q.; Wang, L.; Dou, W.; Wang, J.J. Characterization of Esterase Genes Involving Malathion Detoxification and Establishment of an RNA Interference Method in Liposcelis bostrychophila. Front. Physiol. 2020, 11, 274. [Google Scholar] [CrossRef]
  130. Zhang, X.; Zhang, Y.; Xu, K.; Qin, J.; Wang, D.; Xu, L.; Wang, C. Identification and Biochemical Characterization of a Carboxylesterase Gene Associated with β-Cypermethrin Resistance in Dermanyssus gallinae. Poult. Sci. 2024, 103, 103612. [Google Scholar] [CrossRef]
  131. Nauen, R.; Bass, C.; Feyereisen, R.; Vontas, J. The Role of Cytochrome P450s in Insect Toxicology and Resistance. Annu. Rev. Entomol. 2022, 67, 105–124. [Google Scholar] [CrossRef] [PubMed]
  132. Ye, M.; Nayak, B.; Xiong, L.; Xie, C.; Dong, Y.; You, M.; Yuchi, Z.; You, S. The Role of Insect Cytochrome P450s in Mediating Insecticide Resistance. Agriculture 2022, 12, 53. [Google Scholar] [CrossRef]
  133. Markussen, M.D.K.; Kristensen, M. Cytochrome P450 Monooxygenase-Mediated Neonicotinoid Resistance in the House Fly Musca domestica L. Pestic. Biochem. Physiol. 2010, 98, 50–58. [Google Scholar] [CrossRef]
  134. Schmidt, J.M.; Battlay, P.; Gledhill-Smith, R.S.; Good, R.T.; Lumb, C.; Fournier-Level, A.; Robin, C. Insights into DDT Resistance from the Drosophila melanogaster Genetic Reference Panel. Genetics 2017, 207, 1181–1193. [Google Scholar] [CrossRef]
  135. Zimmer, C.T.; Panini, M.; Singh, K.S.; Randall, E.L.; Field, L.M.; Roditakis, E.; Mazzoni, E.; Bass, C. Use of the Synergist Piperonyl Butoxide Can Slow the Development of Alpha-Cypermethrin Resistance in the Whitefly Bemisia tabaci. Insect Mol. Biol. 2017, 26, 152–163. [Google Scholar] [CrossRef]
  136. Li, P.-R.; Shi, Y.; Ju, D.; Liu, Y.-X.; Wang, W.; He, Y.-S.; Zhang, Y.-Y.; Yang, X.-Q. Metabolic Functional Redundancy of the CYP9A Subfamily Members Leads to P450-Mediated Lambda-Cyhalothrin Resistance in Cydia pomonella. Pest Manag. Sci. 2023, 79, 1452–1466. [Google Scholar] [CrossRef] [PubMed]
  137. Pavlidi, N.; Vontas, J.; Van Leeuwen, T. The Role of Glutathione S-Transferases (GSTs) in Insecticide Resistance in Crop Pests and Disease Vectors. Curr. Opin. Insect Sci. 2018, 27, 97–102. [Google Scholar] [CrossRef] [PubMed]
  138. Koirala, B.K.S.; Moural, T.; Zhu, F. Functional and Structural Diversity of Insect Glutathione S-Transferases in Xenobiotic Adaptation. Int. J. Biol. Sci. 2022, 18, 5713–5723. [Google Scholar] [CrossRef]
  139. Xiao, T.; Huang, X.; Deng, M.; Jiang, Y.; Wang, W.; Xu, X.; Li, J.; Zhao, X.; Pan, B.; He, Z.; et al. Unveiling the Functional Contribution of GSTe16 to Pyrethroid Detoxification in Spodoptera litura. Insect Biochem. Mol. Biol. 2025, 182, 104355. [Google Scholar] [CrossRef] [PubMed]
  140. Kouamo, M.F.M.; Ibrahim, S.S.; Muhammad, A.; Gadji, M.; Hearn, J.; Wondji, C.S. Allelic Variation in a Cluster of Epsilon Glutathione S-Transferase Genes Contributes to DDT and Pyrethroid Resistance in the Major African Malaria Vector Anopheles funestus. BMC Genom. 2025, 26, 452. [Google Scholar] [CrossRef]
  141. Hu, C.; Liu, J.-Y.; Wang, W.; Mota-Sanchez, D.; He, S.; Shi, Y.; Yang, X.-Q. Glutathione S-Transferase Genes Are Involved in Lambda-Cyhalothrin Resistance in via Sequestration. J. Agric. Food Chem. 2022, 70, 2265–2279. [Google Scholar] [CrossRef]
  142. Yan, M.-W.; Xing, X.-R.; Wu, F.-A.; Wang, J.; Sheng, S. UDP-Glycosyltransferases Contribute to the Tolerance of Parasitoid Wasps Towards Insecticides. Pestic. Biochem. Physiol. 2021, 174, 104967. [Google Scholar] [CrossRef]
  143. Du, T.-H.; Yin, C.; Gui, L.-Y.; Liang, J.-J.; Liu, S.-N.; Fu, B.-L.; He, C.; Yang, J.; Wei, X.-G.; Gong, P.-P.; et al. Over-Expression of UDP-Glycosyltransferase UGT353G2 Confers Resistance to Neonicotinoids in Whitefly (Bemisia tabaci). Pestic. Biochem. Physiol. 2023, 196, 105635. [Google Scholar] [CrossRef]
  144. Yang, X.-Y.; Yang, W.; Zhao, H.; Wang, B.-J.; Shi, Y.; Wang, M.-Y.; Liu, S.-Q.; Liao, X.-L.; Shi, L. Functional Analysis of UDP-Glycosyltransferase Genes Conferring Indoxacarb Resistance in Spodoptera litura. Pestic. Biochem. Physiol. 2023, 196, 105589. [Google Scholar] [CrossRef]
  145. Al-Yazeedi, T.; Muhammad, A.; Irving, H.; Ahn, S.J.; Hearn, J.; Wondji, C.S. Overexpression and Nonsynonymous Mutations of UDP-Glycosyltransferases Are Potentially Associated with Pyrethroid Resistance in Anopheles funestus. Genomics 2024, 116, 110798. [Google Scholar] [CrossRef]
  146. Zheng, J.; Chen, X.; Xie, Y.; Zhang, Y.; Huang, Y.; Wu, P.; Lv, J.; Qiu, L. Knocking Out of UDP-Glycosyltransferase Gene UGT2B10 via CRISPR/Cas9 in Helicoverpa armigera Reveals Its Function in Detoxification of Insecticides. J. Agric. Food Chem. 2024, 72, 20862–20871. [Google Scholar] [CrossRef]
  147. Wu, C.; Chakrabarty, S.; Jin, M.; Liu, K.; Xiao, Y. Insect ATP-Binding Cassette (ABC) Transporters: Roles in Xenobiotic Detoxification and Bt Insecticidal Activity. Int. J. Mol. Sci. 2019, 20, 2829. [Google Scholar] [CrossRef]
  148. Al Khoury, C.; Nemer, N.; Nemer, G. Beauvericin Potentiates the Activity of Pesticides by Neutralizing the ATP-Binding Cassette Transporters in Arthropods. Sci. Rep. 2021, 11, 10865. [Google Scholar] [CrossRef]
  149. Qiu, Y.-L.; Wu, F.; Zhang, L.; Jiang, H.-Q.; Chen, J.-T.; Pan, Y.-J.; Li, H.-L. A Sublethal Dose of Neonicotinoid Imidacloprid Precisely Sensed and Detoxified by a C-Minus Odorant-Binding Protein 17 Highly Expressed in the Legs of Apis cerana. Sci. Total Environ. 2023, 885, 163762. [Google Scholar] [CrossRef]
  150. Pelosi, P.; Iovinella, I.; Zhu, J.; Wang, G.; Dani, F.R. Beyond Chemoreception: Diverse Tasks of Soluble Olfactory Proteins in Insects. Biol. Rev. 2018, 93, 184–200. [Google Scholar] [CrossRef] [PubMed]
  151. Ingham, V.A.; Anthousi, A.; Douris, V.; Harding, N.J.; Lycett, G.; Morris, M.; Vontas, J.; Ranson, H. A Sensory Appendage Protein Protects Malaria Vectors from Pyrethroids. Nature 2020, 577, 376–380. [Google Scholar] [CrossRef] [PubMed]
  152. Zhang, J.; Mao, K.; Ren, Z.; Jin, R.; Zhang, Y.; Cai, T.; He, S.; Li, J.; Wan, H. Odorant Binding Protein 3 Is Associated with Nitenpyram and Sulfoxaflor Resistance in Nilaparvata lugens. Int. J. Biol. Macromol. 2022, 209, 1352–1358. [Google Scholar] [CrossRef] [PubMed]
  153. Yao, Q.; Liang, Z.; Chen, B. Evidence for the Participation of Chemosensory Proteins in Response to Insecticide Challenge in Conopomorpha sinensis. J. Agric. Food Chem. 2023, 71, 1360–1368. [Google Scholar] [CrossRef] [PubMed]
  154. Li, F.; Venthur, H.; Wang, S.; Homem, R.A.; Zhou, J.-J. Evidence for the Involvement of the Chemosensory Protein AgosCSP5 in Resistance to Insecticides in the Cotton Aphid, Aphis gossypii. Insects 2021, 12, 335. [Google Scholar] [CrossRef]
  155. Pu, J.; Chung, H. New and Emerging Mechanisms of Insecticide Resistance. Curr. Opin. Insect Sci. 2024, 63, 101184. [Google Scholar] [CrossRef]
  156. Zhu, B.; Li, L.; Wei, R.; Liang, P.; Gao, X. Regulation of GSTu1-Mediated Insecticide Resistance in Plutella xylostella by miRNA and lncRNA. PLoS Genet. 2021, 17, e1009888. [Google Scholar] [CrossRef]
  157. Kang, J.; Ramirez-Calero, S.; Paula, J.R.; Chen, Y.; Schunter, C. Gene Losses, Parallel Evolution and Heightened Expression Confer Adaptations to Dedicated Cleaning Behaviour. BMC Biol. 2023, 21, 180. [Google Scholar] [CrossRef] [PubMed]
  158. Cheng, D.; Guo, Z.; Riegler, M.; Xi, Z.; Liang, G.; Xu, Y. Gut Symbiont Enhances Insecticide Resistance in a Significant Pest, the Oriental Fruit Fly Bactrocera dorsalis (Hendel). Microbiome 2017, 5, 13. [Google Scholar] [CrossRef] [PubMed]
  159. Soh, L.-S.; Veera Singham, G. Bacterial Symbionts Influence Host Susceptibility to Fenitrothion and Imidacloprid in the Obligate Hematophagous Bed Bug, Cimex hemipterus. Sci. Rep. 2022, 12, 4919. [Google Scholar] [CrossRef]
  160. Zhao, M.; Lin, X.; Guo, X. The Role of Insect Symbiotic Bacteria in Metabolizing Phytochemicals and Agrochemicals. Insects 2022, 13, 583. [Google Scholar] [CrossRef]
  161. Sato, Y.; Jang, S.; Takeshita, K.; Itoh, H.; Koike, H.; Tago, K.; Hayatsu, M.; Hori, T.; Kikuchi, Y. Insecticide Resistance by a Host–Symbiont Reciprocal Detoxification. Nat. Commun. 2021, 12, 6432. [Google Scholar] [CrossRef]
  162. Wei, X.; Peng, H.; Li, Y.; Meng, B.; Wang, S.; Bi, S.; Zhao, X. Pyrethroids Exposure Alters the Community and Function of the Internal Microbiota in Aedes albopictus. Ecotoxicol. Environ. Saf. 2023, 252, 114579. [Google Scholar] [CrossRef] [PubMed]
  163. Lv, C.; Huang, Y.Z.; Luan, J.B. Insect–Microbe Symbiosis-Based Strategies Offer a New Avenue for the Management of Insect Pests and Their Transmitted Pathogens. Crop Health 2024, 2, 18. [Google Scholar] [CrossRef]
  164. Kline, O.; Joshi, N.K. Microbial Symbiont-Based Detoxification of Different Phytotoxins and Synthetic Toxic Chemicals in Insect Pests and Pollinators. J. Xenobiot. 2024, 14, 753–771. [Google Scholar] [CrossRef]
  165. Liu, X.D.; Guo, H.F. Importance of Endosymbionts Wolbachia and Rickettsia in Insect Resistance Development. Curr. Opin. Insect Sci. 2019, 33, 84–90. [Google Scholar] [CrossRef]
  166. Ye, Q.-T.; Gong, X.; Liu, H.-H.; Wu, B.-X.; Peng, C.-W.; Hong, X.-Y.; Bing, X.-L. The Symbiont Wolbachia Alleviates Pesticide Susceptibility in the Two-Spotted Spider Mite Tetranychus urticae through Enhanced Host Detoxification Pathways. Insect Sci. 2024, 31, 1822–1837. [Google Scholar] [CrossRef]
  167. Zhang, Y.; Cai, T.; Ren, Z.; Liu, Y.; Yuan, M.; Cai, Y.; Yu, C.; Shu, R.; He, S.; Wong, A.C.N.; et al. Decline in symbiont-dependent host detoxification metabolism contributes to increased insecticide susceptibility of insects under high temperature. ISME J. 2021, 15, 3693–3703. [Google Scholar] [CrossRef] [PubMed]
  168. Pang, R.; Chen, M.; Yue, L.; Xing, K.; Li, T.; Kang, K.; Liang, Z.; Yuan, L.; Zhang, W. A distinct strain of Arsenophonus symbiont decreases insecticide resistance in its insect host. PLoS Genet. 2018, 14, e1007725. [Google Scholar] [CrossRef] [PubMed]
  169. Brar, G.; Ngor, L.; McFrederick, Q.S.; Torson, A.S.; Rajamohan, A.; Rinehart, J.; Singh, P.; Bowsher, J.H. High Abundance of Lactobacilli in the Gut Microbiome of Honey Bees During Winter. Sci. Rep. 2025, 15, 7409. [Google Scholar] [CrossRef]
  170. Wu, Y.; Zheng, Y.; Chen, Y.; Wang, S.; Chen, Y.; Hu, F.; Zheng, H. Honey Bee (Apis mellifera) Gut Microbiota Promotes Host Endogenous Detoxification Capability via Regulation of P450 Gene Expression in the Digestive Tract. Microb. Biotechnol. 2020, 13, 1201–1212. [Google Scholar] [CrossRef]
  171. Zeng, B.; Zhang, F.; Liu, Y.-T.; Wu, S.-F.; Bass, C.; Gao, C.-F. Symbiotic Bacteria Confer Insecticide Resistance by Metabolizing Buprofezin in the Brown Planthopper, Nilaparvata lugens (Stål). PLoS Pathog. 2023, 19, e1011828. [Google Scholar] [CrossRef] [PubMed]
  172. Fusetto, R.; Denecke, S.; Perry, T.; O’Hair, R.A.J.; Batterham, P. Partitioning the Roles of CYP6G1 and Gut Microbes in the Metabolism of the Insecticide Imidacloprid in Drosophila melanogaster. Sci. Rep. 2017, 7, 11339. [Google Scholar] [CrossRef] [PubMed]
  173. Lv, N.; Li, R.; Cheng, S.; Zhang, L.; Liang, P.; Gao, X. The Gut Symbiont Sphingomonas Mediates Imidacloprid Resistance in the Important Agricultural Insect Pest Aphis gossypii Glover. BMC Biol. 2023, 21, 86. [Google Scholar] [CrossRef]
  174. Zhao, P.; Rensing, C.; Wang, D. Symbiotic Bacteria Modulate Lymantria dispar Immunity by Altering Community Proportions after Infection with LdMNPV. Int. J. Mol. Sci. 2023, 24, 9694. [Google Scholar] [CrossRef]
  175. Bottino-Rojas, V.; Talyuli, O.A.C.; Carrara, L.; Martins, A.J.; James, A.A.; Oliveira, P.L.; Paiva-Silva, G.O. The redox-sensing gene Nrf2 affects intestinal homeostasis, insecticide resistance, and Zika virus susceptibility in the mosquito Aedes aegypti. J. Biol. Chem. 2018, 293, 9053–9063. [Google Scholar] [CrossRef]
  176. Feng, H.; Chen, W.; Hussain, S.; Shakir, S.; Tzin, V.; Adegbayi, F.; Ugine, T.; Fei, Z.; Jander, G. Horizontally transferred genes as RNA interference targets for aphid and whitefly control. Plant Biotechnol. J. 2023, 21, 754–768. [Google Scholar] [CrossRef]
  177. Verster, K.I.; Tarnopol, R.L.; Akalu, S.M.; Whiteman, N.K. Horizontal Transfer of Microbial Toxin Genes to Gall Midge Genomes. Genome Biol. Evol. 2021, 13, evab202. [Google Scholar] [CrossRef]
  178. Xia, J.; Guo, Z.; Yang, Z.; Han, H.; Wang, S.; Xu, H.; Yang, X.; Yang, F.; Wu, Q.; Xie, W.; et al. Whitefly hijacks a plant detoxification gene that neutralizes plant toxins. Cell 2021, 184, 1693–1705.e17. [Google Scholar] [CrossRef]
  179. Li, Y.; Zhou, Y.; Jing, W.; Xu, S.; Jin, Y.; Xu, Y.; Wang, H. Horizontally acquired cysteine synthase genes undergo functional divergence in lepidopteran herbivores. Heredity 2021, 127, 21–34. [Google Scholar] [CrossRef]
  180. Liu, C.; Li, Y.; Chen, Y.; Chen, X.; Huang, J.; Rokas, A.; Shen, X. How Has Horizontal Gene Transfer Shaped the Evolution of Insect Genomes? Environ. Microbiol. 2023, 25, 642–645. [Google Scholar] [CrossRef] [PubMed]
  181. Yadav, A.; Suhag, A.; Jaiwal, R.; Chaudhary, D.; Jaiwal, P.K. Current progress and challenges of horizontal gene transfers in whiteflies (Bemisia tabaci) for their sustainable management. J. Asia-Pac. Entomol. 2024, 27, 102216. [Google Scholar] [CrossRef]
  182. Zhang, Y.; Cai, T.; Yuan, M.; Li, Z.; Jin, R.; Ren, Z.; Qin, Y.; Yu, C.; Cai, Y.; Shu, R.; et al. Microbiome variation correlates with the insecticide susceptibility in different geographic strains of a significant agricultural pest, Nilaparvata lugens. npj Biofilms Microbiomes 2023, 9, 2. [Google Scholar] [CrossRef] [PubMed]
  183. Zalucki, M.; Furlong, M. Behavior as a Mechanism of Insecticide Resistance: Evaluation of the Evidence. Curr. Opin. Insect Sci. 2017, 21, 19–25. [Google Scholar] [CrossRef] [PubMed]
  184. Pu, J.; Wang, Z.; Chung, H. Climate Change and the Genetics of Insecticide Resistance. Pest Manag. Sci. 2020, 76, 846–852. [Google Scholar] [CrossRef]
  185. Quintela, V.; Pereira, D.L.; Langa, T.P.; de Oliveira, M.; Siqueira, H.A.A. Field-evolved resistance in Phthorimaea absoluta to abamectin: Genetic foundations, female-linked traits, and cross-resistance pattern. Pest Manag. Sci. 2025, 81, 3867–3877. [Google Scholar] [CrossRef] [PubMed]
  186. Thiesen, L.V.; Gonçalves, G.C.; Guidolin, A.S.; Nascimento, A.R.B.; Coutinho, E.F.; Borba, J.P.; Picelli, E.C.M.; Omoto, C. Characterization of cyantraniliprole resistance in Spodoptera frugiperda: Selection, inheritance pattern, and cross-resistance to other diamide insecticides. Pest Manag. Sci. 2025, 81, 4677–4689. [Google Scholar] [CrossRef] [PubMed]
  187. Venkatesan, T.; Chethan, B.R.; Mani, M. Insecticide Resistance and Its Management in the Insect Pests of Horticultural Crops. In Trends in Horticultural Entomology; Mani, M., Ed.; Springer: Singapore, 2022. [Google Scholar] [CrossRef]
  188. Gul, H.; Gadratagi, B.G.; Güncan, A.; Tyagi, S.; Ullah, F.; Desneux, N.; Liu, X. Fitness costs of resistance to insecticides in insects. Front. Physiol. 2023, 14, 1238111. [Google Scholar] [CrossRef] [PubMed]
  189. Hu, B.; Huang, H.; Hu, S.; Ren, M.; Wei, Q.; Tian, X.; Elzaki, M.E.A.; Bass, C.; Su, J.; Palli, S.R. Changes in Both Trans- and Cis-Regulatory Elements Mediate Insecticide Resistance in a Lepidopteron Pest, Spodoptera exigua. PLoS Genet. 2021, 17, e1009403. [Google Scholar] [CrossRef] [PubMed]
  190. Chen, Y.; Cen, Y.; Liu, Y.; Peng, Y.; Lin, Y.; Feng, Q.; Xiao, Y.; Zheng, S. P450 gene CYP321A8 is responsible for cross-resistance of insecticides in field populations of Spodoptera frugiperda. Insect Sci. 2025, 32, 227–242. [Google Scholar] [CrossRef]
  191. Zuo, Y.; Pei, Y.; Li, Y.; Wen, S.; Ren, X.; Li, L.; Wu, Y.; Hu, Z. The Synergism Between Metabolic and Target-Site Resistance Enhances the Intensity of Resistance to Pyrethroids in Spodoptera exigua. Insect Biochem. Mol. Biol. 2025, 180, 104313. [Google Scholar] [CrossRef]
  192. Li, W.; Yang, W.; Shi, Y.; Yang, X.; Liu, S.; Liao, X.; Shi, L. Comprehensive analysis of the overexpressed cytochrome P450-based insecticide resistance mechanism in Spodoptera litura. J. Hazard. Mater. 2024, 461, 132605. [Google Scholar] [CrossRef]
  193. Qian, C.; Li, J.; Wu, S.; Yang, Y.; Wu, Y.; Wang, X. Cross-Resistance and Genetics of Field-Evolved Resistance to Chlorfenapyr in Plutella xylostella. Insect Sci. 2024, 31, 533–541. [Google Scholar] [CrossRef]
  194. Tan, M.; Jiang, H.; Chai, R.; Fan, M.; Niu, Z.; Sun, G.; Yan, S.; Jiang, D. Cd exposure confers β-cypermethrin tolerance in Lymantria dispar by activating the ROS/CnCC signaling pathway-mediated P450 detoxification. J. Hazard. Mater. 2024, 478, 135566. [Google Scholar] [CrossRef]
  195. Yan, S.; Tan, M.; Zhang, A.; Jiang, D. The Exposure Risk of Heavy Metals to Insect Pests and Their Impact on Pests Occurrence and Cross-Tolerance to Insecticides: A Review. Sci. Total Environ. 2024, 916, 170274. [Google Scholar] [CrossRef]
  196. Horowitz, A.R.; Ghanim, M.; Roditakis, E.; Nauen, R.; Ishaaya, I. Insecticide resistance and its management in Bemisia tabaci species. J. Pest Sci. 2020, 93, 893–910. [Google Scholar] [CrossRef]
  197. Guo, Z.; Tang, J.; Ling, S.; Zheng, C.; Bass, C.; He, S.; Wan, H.; Li, J.; Ma, K. Upregulation of a cytochrome P450 gene, CYP6B50, confers multi-insecticide resistance in Spodoptera frugiperda. J. Adv. Res. 2025, in press. [Google Scholar] [CrossRef]
  198. Kiuru, C.; Constantino, L.; Cole, G.; Karisa, J.; Wanjiku, C.; Okoko, M.; Candrinho, B.; Saute, F.; Rabinovich, N.R.; Chaccour, C.; et al. Multiple insecticide resistance in Anopheles funestus from Mopeia, Central Mozambique. Malar. J. 2025, 24, 81. [Google Scholar] [CrossRef] [PubMed]
  199. Madgwick, P.G.; Slater, R.; Kanitz, R. The evolution of pesticide resistance: A data-driven case study of chlorantraniliprole resistance in Chilo suppressalis and other Lepidopteran pests in China. Evol. Appl. 2025, 18, e70131. [Google Scholar] [CrossRef] [PubMed]
  200. Sun, H.; Wang, S.; Liu, C.; Hu, W.K.; Liu, J.W.; Zheng, L.J.; Gao, M.Y.; Guo, F.R.; Qiao, S.T.; Liu, J.L.; et al. Risk assessment, fitness cost, cross-resistance, and mechanism of tetraniliprole resistance in the rice stem borer, Chilo suppressalis. Insect Sci. 2024, 31, 835–846. [Google Scholar] [CrossRef]
  201. Cordova, D.; Benner, E.A.; Clark, D.A.; Bolgunas, S.P.; Lahm, G.P.; Gutteridge, S.; Rhoades, D.F.; Wu, L.; Sopa, J.S.; Rauh, J.J.; et al. Pyrrole-2 Carboxamides—A Novel Class of Insect Ryanodine Receptor Activators. Pestic. Biochem. Physiol. 2021, 174, 104798. [Google Scholar] [CrossRef]
  202. Lin, P.-X.; Peng, Y.-X.; Xing, J.-Y.; Liu, Z.-Y.; Guo, F.-R.; Thia, J.A.; Gao, C.-F.; Wu, S.-F. Cis-Regulation of the CYP6CS1 Gene and Its Role in Mediating Cross-Resistance in a Pymetrozine-Resistant Strain of Nilaparvata lugens. Insect Biochem. Mol. Biol. 2025, 177, 104261. [Google Scholar] [CrossRef]
  203. Lv, M.; Wang, W.; Fang, F.; Fu, X.; Liang, G. The Changes in Cross-Resistance, Fitness, and Feeding Behavior in Aphis gossypii as Their Resistance to Sulfoxaflor Declines. Insects 2024, 15, 920. [Google Scholar] [CrossRef]
  204. Qu, C.; Yao, J.; Huang, J.; Che, W.; Fang, Y.; Luo, C.; Wang, R. Tetraniliprole Resistance in Field-Collected Populations of Tuta absoluta (Lepidoptera: Gelechiidae) from China: Baseline Susceptibility, Cross-Resistance, Inheritance, and Biochemical Mechanism. Pestic. Biochem. Physiol. 2024, 203, 106019. [Google Scholar] [CrossRef]
  205. Madgwick, P.G.; Kanitz, R. What Is the Value of Rotations to Insecticide Resistance Management? Pest Manag. Sci. 2024, 80, 1671–1680. [Google Scholar] [CrossRef]
  206. Hu, B.; Xing, Z.; Dong, H.; Chen, X.; Ren, M.; Liu, K.; Rao, C.; Tan, A.; Su, J. Cytochrome P450 CYP6AE70 Confers Resistance to Multiple Insecticides in a Lepidopteran Pest, Spodoptera exigua. J. Agric. Food Chem. 2024, 72, 23141–23150. [Google Scholar] [CrossRef] [PubMed]
  207. Narva, K.; Toprak, U.; Alyokhin, A.; Groves, R.; Jurat-Fuentes, J.L.; Moar, W.; Nauen, R.; Whipple, S.; Head, G. Insecticide Resistance Management Scenarios Differ for RNA-Based Sprays and Traits. Insect Mol. Biol. 2025, in press. [Google Scholar] [CrossRef]
  208. Ullah, F.; Guru-Pirasanna-Pandi, G.; Murtaza, G.; Sarangi, S.; Gul, H.; Li, X.; Chavarín-Gómez, L.E.; Ramírez-Romero, R.; Guedes, R.N.C.; Desneux, N.; et al. Evolving strategies in agroecosystem pest control: Transitioning from chemical to green management. J. Pest Sci. 2025, 1–18. [Google Scholar] [CrossRef]
  209. Kamran, M.; Ahmad, M.; Shad, S.A. Genetic Analysis of Chlorantraniliprole Resistance in the Non-Target Bio-Control Agent Trichogramma chilonis. Chemosphere 2025, 370, 143952. [Google Scholar] [CrossRef] [PubMed]
  210. Afzal, M.B.S.; Ijaz, M.; Abbas, N.; Shad, S.A.; Serrão, J.E. Resistance of Lepidopteran Pests to Bacillus thuringiensis Toxins: Evidence of Field and Laboratory Evolved Resistance and Cross-Resistance, Mode of Resistance Inheritance, Fitness Costs, Mechanisms Involved and Management Options. Toxins 2024, 16, 315. [Google Scholar] [CrossRef]
  211. Anderson, J.A.; Ellsworth, P.C.; Faria, J.C.; Head, G.P.; Owen, M.D.K.; Pilcher, C.D.; Shelton, A.M.; Meissle, M. Genetically Engineered Crops: Importance of Diversified Integrated Pest Management for Agricultural Sustainability. Front. Bioeng. Biotechnol. 2019, 7, 24. [Google Scholar] [CrossRef]
  212. Gassmann, A.J.; Reisig, D.D. Management of Insect Pests with Bt Crops in the United States. Annu. Rev. Entomol. 2023, 68, 31–49. [Google Scholar] [CrossRef]
  213. Downes, S.; Kriticos, D.; Parry, H.; Paull, C.; Schellhorn, N.; Zalucki, M.P. A perspective on management of Helicoverpa armigera: Transgenic Bt cotton, IPM, and landscapes. Pest Manag. Sci. 2017, 73, 485–492. [Google Scholar] [CrossRef]
  214. Suh, P.F.; Elanga-Ndille, E.; Tchouakui, M.; Sandeu, M.M.; Tagne, D.; Wondji, C.; Ndo, C. Impact of Insecticide Resistance on Malaria Vector Competence: A Literature Review. Malar. J. 2023, 22, 19. [Google Scholar] [CrossRef]
  215. Demissie, D.B.; Fetensa, G.; Desta, T.; Tiyare, F.T. Effectiveness and Efficacy of Long-Lasting Insecticidal Nets for Malaria Control in Africa: Systematic Review and Meta-Analysis of Randomized Controlled Trials. Int. J. Environ. Res. Public Health 2025, 22, 1045. [Google Scholar] [CrossRef]
  216. Kouassi, B.L.; Edi, C.; Tia, E.; Konan, L.Y.; Akré, M.A.; Koffi, A.A.; Ouattara, A.F.; Tanoh, A.M.; Zinzindohoue, P.; Kouadio, B.; et al. Susceptibility of Anopheles gambiae from Côte d’Ivoire to insecticides used on insecticide-treated nets: Evaluating the additional entomological impact of piperonyl butoxide and chlorfenapyr. Malar. J. 2020, 19, 454. [Google Scholar] [CrossRef] [PubMed]
  217. Li, Y.; Gao, Y.; Wu, K. Function and Effectiveness of Natural Refuge in IRM Strategies for Bt Crops. Curr. Opin. Insect Sci. 2017, 21, 1–6. [Google Scholar] [CrossRef] [PubMed]
  218. Dively, G.P.; Huang, F.; Oyediran, I.; Burd, T.; Morsello, S. Evaluation of Gene Flow in Structured and Seed Blend Refuge Systems of Non-Bt and Bt Corn. J. Pest Sci. 2020, 93, 439–447. [Google Scholar] [CrossRef]
  219. Zhu, F.; Lavine, L.; O’Neal, S.; Lavine, M.; Foss, C.; Walsh, D. Insecticide Resistance and Management Strategies in Urban Ecosystems. Insects 2016, 7, 2. [Google Scholar] [CrossRef] [PubMed]
  220. Gontijo, L.M. Matching Landscapes and Managing Habitats to Enhance Classical Biological Control of Arthropods. Biol. Control 2022, 169, 104896. [Google Scholar] [CrossRef]
  221. Hatt, S.; Döring, T.F. Designing Pest Suppressive Agroecosystems: Principles for an Integrative Diversification Science. J. Clean. Prod. 2023, 432, 139701. [Google Scholar] [CrossRef]
  222. World Health Organization. Test Procedures for Insecticide Resistance Monitoring in Malaria Vector Mosquitoes, 2nd ed.; World Health Organization: Geneva, Switzerland, 2016. Available online: https://apps.who.int/iris/handle/10665/250677 (accessed on 4 August 2025).
  223. Chen, L.; Lang, K.; Mei, Y.; Shi, Z.; He, K.; Li, F.; Xiao, H.; Ye, G.; Han, Z. FastD: Fast Detection of Insecticide Target-Site Mutations and Overexpressed Detoxification Genes in Insect Populations from RNA-Seq Data. Ecol. Evol. 2020, 10, 14346–14358. [Google Scholar] [CrossRef] [PubMed]
  224. Guan, F.; Zhang, J.; Shen, H.; Wang, X.; Padovan, A.; Walsh, T.K.; Tay, W.T.; Gordon, K.H.J.; James, W.; Czepak, C.; et al. Whole-Genome Sequencing to Detect Mutations Associated with Resistance to Insecticides and Bt Proteins in Spodoptera frugiperda. Insect Sci. 2021, 28, 627–638. [Google Scholar] [CrossRef]
  225. Mertz, R.W.; Dressel, A.E.; Fisher, C.R.; Moon, R.D.; Donahue, W.A.; Kasai, S.; Scott, J.G. Frequencies and Distribution of Kdr and Ace Alleles that Cause Insecticide Resistance in House Flies in the United States. Pestic. Biochem. Physiol. 2023, 194, 105497. [Google Scholar] [CrossRef]
  226. Chang, C.-C.; Dai, S.-M.; Chen, C.-Y.; Huang, L.-H.; Chen, Y.-H.; Hsu, J.-C. Insecticide Resistance and Characteristics of Mutations Related to Target Site Insensitivity of Diamondback Moths in Taiwan. Pestic. Biochem. Physiol. 2024, 203, 106001. [Google Scholar] [CrossRef]
  227. Jeon, J.; Kim, H.C.; Klein, T.A.; Jun, H.; Choi, K.S. Insecticide Resistance Mutations of Anopheles Species in the Republic of Korea. PLoS Negl. Trop. Dis. 2025, 19, e0012748. [Google Scholar] [CrossRef]
  228. Helps, J.C.; Paveley, N.D.; White, S.; van den Bosch, F. Determinants of Optimal Insecticide Resistance Management Strategies. J. Theor. Biol. 2020, 503, 110383. [Google Scholar] [CrossRef]
  229. Willow, J.; Cook, S.M.; Veromann, E.; Smagghe, G. Uniting RNAi Technology and Conservation Biocontrol to Promote Global Food Security and Agrobiodiversity. Front. Bioeng. Biotechnol. 2022, 10, 871651. [Google Scholar] [CrossRef]
  230. Liu, W.; Wang, X.; Zhou, A.; Zhang, J.; Ge, X.; Moussian, B.; Yan, C.; Gao, S.; Wang, Y. Trends and Emerging Hotspots in RNAi-Based Arthropod Pest Control: A Comprehensive Bibliometric Analysis. J. Insect Physiol. 2025, 161, 104754. [Google Scholar] [CrossRef]
  231. Zhong, X.; Yu, X.; Zhang, J.; Xu, J.; Qin, M.; Cao, M.; Francis, F.; Xia, L. RNAi Technologies for Insect Control in Crop Protection. Crop J. 2025, in press. [Google Scholar] [CrossRef]
  232. Mnzava, A.P.; Knox, T.B.; Temu, E.A.; Trett, A.; Fornadel, C.; Hemingway, J.; Renshaw, M. Implementation of the Global Plan for Insecticide Resistance Management in Malaria Vectors: Progress, Challenges and the Way Forward. Malar. J. 2015, 14, 173. [Google Scholar] [CrossRef]
  233. IRAC Group 30 Task Team. Insecticide Resistance Management Guidelines: IRAC Group 30 Insecticides. Version 3.1; Updated 15 December 2021 & 16 February 2022. Available online: https://irac-online.org/documents/irac-moa-group-30-irm-guidelines/ (accessed on 5 August 2025).
  234. Food and Agriculture Organization of the United Nations. FAO Guidelines on Prevention and Management of Pesticide Resistance; Food and Agriculture Organization of the United Nations: Rome, Italy, 2012; Available online: https://openknowledge.fao.org/items/64bb88d3-2a15-4f23-b94e-d25a8777ae5b (accessed on 5 August 2025).
  235. Head, G.P.; Carroll, M.W.; Evans, S.P.; Rule, D.M.; Willse, A.R.; Clark, T.L.; Storer, N.P.; Flannagan, R.D.; Samuel, L.W.; Meinke, L.J. Evaluation of SmartStax and SmartStax PRO Maize against Western Corn Rootworm and Northern Corn Rootworm: Efficacy and Resistance Management. Pest Manag. Sci. 2017, 73, 1883–1899. [Google Scholar] [CrossRef]
  236. Narva, K.; Otto, E.; Sridharan, K.; Flannagan, R.; Barnes, E.; Mézin, L.; Manley, B. Calantha™: The First Commercialized Sprayable dsRNA Product for Insect Control. In RNA Interference in Agriculture: Basic Science to Applications; Smagghe, G., Palli, S.R., Swevers, L., Eds.; Springer: Cham, Switzerland, 2025; pp. 679–715. [Google Scholar] [CrossRef]
  237. Bachman, P.; Fischer, J.; Song, Z.; Urbanczyk-Wochniak, E.; Watson, G. Environmental Fate and Dissipation of Applied dsRNA in Soil, Aquatic Systems, and Plants. Front. Plant Sci. 2020, 11, 21. [Google Scholar] [CrossRef] [PubMed]
  238. Vatanparast, M.; Merkel, L.; Amari, K. Exogenous Application of dsRNA in Plant Protection: Efficiency, Safety Concerns and Risk Assessment. Int. J. Mol. Sci. 2024, 25, 6530. [Google Scholar] [CrossRef] [PubMed]
  239. Khajuria, C.; Ivashuta, S.; Wiggins, E.; Flagel, L.; Moar, W.; Pleau, M.; Miller, K.; Zhang, Y.; Ramaseshadri, P.; Jiang, C.; et al. Development and characterization of the first dsRNA-resistant insect population from western corn rootworm, Diabrotica virgifera virgifera LeConte. PLoS ONE 2018, 13, e0197059. [Google Scholar] [CrossRef]
  240. Mishra, S.; Moar, W.; Jurat-Fuentes, J.L. Larvae of Colorado Potato Beetle (Leptinotarsa decemlineata Say) Resistant to Double-Stranded RNA (dsRNA) Remain Susceptible to Small-Molecule Pesticides. Pest Manag. Sci. 2024, 80, 905–909. [Google Scholar] [CrossRef]
  241. Liao, C.; Zhang, M.; Zhang, J. Characterization and Potential Mechanism of Resistance to Double-Stranded RNA in Willow Leaf Beetle, Plagiodera versicolora. J. Pest Sci. 2024, 97, 2217–2226. [Google Scholar] [CrossRef]
  242. Pinto, M.M.D.; Ferreira dos Santos, R.; De Bortoli, S.A.; Moar, W.; Jurat-Fuentes, J.L. Lack of Fitness Costs in dsRNA-Resistant Leptinotarsa decemlineata (Coleoptera: Chrysomelidae). J. Econ. Entomol. 2023, 116, 1352–1359. [Google Scholar] [CrossRef] [PubMed]
  243. EPA FIFRA Scientific Advisory Panel. Subpanel on Bacillus thuringiensis (Bt) Plant-Pesticides and Resistance Management; U.S. Environmental Protection Agency: Washington, DC, USA, 1998. Available online: https://archive.epa.gov/scipoly/sap/meetings/web/pdf/finalfeb.pdf (accessed on 4 August 2025).
  244. EPA FIFRA Scientific Advisory Panel. Sets of Scientific Issues Being Considered by the Environmental Protection Agency Regarding Bt Plant-Pesticides Risk and Benefit Assessments; U.S. Environmental Protection Agency: Washington, DC, USA, 2001. Available online: https://archive.epa.gov/scipoly/sap/meetings/web/pdf/octoberfinal.pdf (accessed on 4 August 2025).
  245. Swevers, L.; Palli, S.R.; Smagghe, G. The Road to RNAi-Based Pest Control: A Historical Perspective. In RNA Interference in Agriculture: Basic Science to Applications; Smagghe, G., Palli, S.R., Swevers, L., Eds.; Springer: Cham, Switzerland, 2025; pp. 1–23. [Google Scholar] [CrossRef]
  246. Zulfa, R.; Lo, W.-C.; Cheng, P.-C.; Martini, M.; Chuang, T.-W. Updating the Insecticide Resistance Status of Aedes aegypti and Aedes albopictus in Asia: A Systematic Review and Meta-Analysis. Trop. Med. Infect. Dis. 2022, 7, 306. [Google Scholar] [CrossRef]
  247. Kenis, M.; Benelli, G.; Biondi, A.; Calatayud, P.-A.; Day, R.; Desneux, N.; Harrison, R.D.; Kriticos, D.; Rwomushana, I.; van den Berg, J.; et al. Invasiveness, biology, ecology, and management of the fall armyworm, Spodoptera frugiperda. Entomol. Gen. 2023, 43, 187–241. [Google Scholar] [CrossRef]
  248. Hawkins, N.J.; Bass, C.; Dixon, A.; Neve, P. The Evolutionary Origins of Pesticide Resistance. Biol. Rev. 2019, 94, 135–155. [Google Scholar] [CrossRef]
  249. Weill, M.; Malcolm, C.; Chandre, F.; Mogensen, K.; Berthomieu, A.; Marquine, M.; Raymond, M. The Unique Mutation in ace-1 Giving High Insecticide Resistance is Easily Detectable in Mosquito Vectors. Insect Mol. Biol. 2004, 13, 1–7. [Google Scholar] [CrossRef] [PubMed]
  250. Nabeshima, T.; Mori, A.; Kozaki, T.; Iwata, Y.; Hidoh, O.; Harada, S.; Kasai, S.; Severson, D.W.; Kono, Y.; Tomita, T. An Amino Acid Substitution Attributable to Insecticide-Insensitivity of Acetylcholinesterase in a Japanese Encephalitis Vector Mosquito, Culex tritaeniorhynchus. Biochem. Biophys. Res. Commun. 2004, 313, 794–801. [Google Scholar] [CrossRef]
  251. Alout, H.; Berthomieu, A.; Cui, F.; Tan, Y.; Berticat, C.; Qiao, C.; Weill, M. Different Amino-Acid Substitutions Confer Insecticide Resistance through Acetylcholinesterase 1 Insensitivity in Culex vishnui and Culex tritaeniorhynchus (Diptera: Culicidae) from China. J. Med. Entomol. 2007, 44, 463–469. [Google Scholar] [CrossRef] [PubMed]
  252. Saavedra-Rodriguez, K.; Urdaneta-Marquez, L.; Rajatileka, S.; Moulton, M.; Flores, A.E.; Fernandez-Salas, I.; Bisset, J.; Rodriguez, M.; McCall, P.J.; Donnelly, M.J.; et al. A mutation in the voltage-gated sodium channel gene associated with pyrethroid resistance in Latin American Aedes aegypti. Insect Mol. Biol. 2007, 16, 785–798. [Google Scholar] [CrossRef] [PubMed]
  253. Du, Y.; Nomura, Y.; Satar, G.; Hu, Z.; Nauen, R.; He, S.Y.; Zhorov, B.S.; Dong, K. Molecular Evidence for Dual Pyrethroid-Receptor Sites on a Mosquito Sodium Channel. Proc. Natl. Acad. Sci. USA 2013, 110, 11785–11790. [Google Scholar] [CrossRef]
  254. Yanola, J.; Somboon, P.; Walton, C.; Nachaiwieng, W.; Somwang, P.; Prapanthadara, L. High-Throughput Assays for Detection of the F1534C Mutation in the Voltage-Gated Sodium Channel Gene in Permethrin-Resistant Aedes aegypti and the Distribution of This Mutation throughout Thailand. Trop. Med. Int. Health 2011, 16, 501–509. [Google Scholar] [CrossRef]
  255. Harris, A.F.; Rajatileka, S.; Ranson, H. Pyrethroid Resistance in Aedes aegypti from Grand Cayman. Am. J. Trop. Med. Hyg. 2010, 83, 277–284. [Google Scholar] [CrossRef]
  256. Hu, Z.; Du, Y.; Nomura, Y.; Dong, K. A Sodium Channel Mutation Identified in Aedes aegypti Selectively Re-duces Cockroach Sodium Channel Sensitivity to Type I, but Not Type II Pyrethroids. Insect Biochem. Mol. Biol. 2011, 41, 9–13. [Google Scholar] [CrossRef]
  257. Haddi, K.; Tomé, H.V.V.; Du, Y.; Valbon, W.R.; Nomura, Y.; Martins, G.F.; Dong, K.; Oliveira, E.E. Detection of a New Pyrethroid Resistance Mutation (V410L) in the Sodium Channel of Aedes aegypti: A Potential Challenge for Mosquito Control. Sci. Rep. 2017, 7, 46549. [Google Scholar] [CrossRef]
  258. Saavedra-Rodriguez, K.; Maloof, F.V.; Campbell, C.L.; Garcia-Rejon, J.; Lenhart, A.; Penilla, P.; Rodriguez, A.; Sandoval, A.A.; Flores, A.E.; Ponce, G.; et al. Parallel evolution of vgsc mutations at domains IS6, IIS6 and IIIS6 in pyrethroid resistant Aedes aegypti from Mexico. Sci. Rep. 2018, 8, 6747. [Google Scholar] [CrossRef] [PubMed]
  259. Poupardin, R.; Srisukontarat, W.; Yunta, C.; Ranson, H. Identification of Carboxylesterase Genes Implicated in Temephos Resistance in the Dengue Vector Aedes aegypti. PLoS Negl. Trop. Dis. 2014, 8, e2743. [Google Scholar] [CrossRef]
  260. Asgarian, T.S.; Vatandoost, H.; Hanafi-Bojd, A.A.; Nikpoor, F. Worldwide Status of Insecticide Resistance of Aedes aegypti and Ae. albopictus, Vectors of Arboviruses of Chikungunya, Dengue, Zika and Yellow Fever. J. Arthropod Borne Dis. 2023, 17, 1–27. [Google Scholar] [CrossRef] [PubMed]
  261. Nurul-Nastasea, S.; Yu, K.-X.; Ahmad, R.; Mohamed Nor, Z.; Tengku Idzzan Nadzirah, T.-I.; Roza, D.; Masse Rezki, S.; Wan Mohamad Ali, W.N. Insecticide Resistance Status of Aedes aegypti and Aedes albopictus in Malaysia (2010 to 2022): A Review. Asian Pac. J. Trop. Med. 2023, 16, 434–445. [Google Scholar] [CrossRef]
  262. Chandor-Proust, A.; Bibby, J.; Régent-Kloeckner, M.; Roux, J.; Guittard-Crilat, E.; Poupardin, R.; Riaz, M.A.; Paine, M.; Dauphin-Villemant, C.; Reynaud, S.; et al. The Central Role of Mosquito Cytochrome P450 CYP6Zs in Insecticide Detoxification Revealed by Functional Expression and Structural Modelling. Biochem. J. 2013, 455, 75–85. [Google Scholar] [CrossRef]
  263. Rocha dos Santos, C.; Rodovalho, C.M.; Jablonka, W.; Martins, A.J.; Pereira Lima, J.B.; dos Santos Dias, L.; da Silva Neto, M.A.C.; Atella, G.C. Insecticide Resistance, Fitness and Susceptibility to Zika Infection of an Interbred Aedes aegypti Population from Rio de Janeiro, Brazil. Parasites Vectors 2020, 13, 496. [Google Scholar] [CrossRef]
  264. Vera-Maloof, F.Z.; Saavedra-Rodriguez, K.; Penilla-Navarro, R.P.; Rodriguez-Ramirez, A.D.; Dzul, F.; Manrique-Saide, P.; Black, W.C., IV. Loss of pyrethroid resistance in newly established laboratory colonies of Aedes aegypti. PLoS Negl. Trop. Dis. 2020, 14, e0007753. [Google Scholar] [CrossRef]
  265. Smith, L.B.; Silva, J.J.; Chen, C.; Harrington, L.C.; Scott, J.G. Fitness Costs of Individual and Combined Pyrethroid Resistance Mechanisms, kdr and CYP-Mediated Detoxification, in Aedes aegypti. PLoS Negl. Trop. Dis. 2021, 15, e0009271. [Google Scholar] [CrossRef]
  266. Fisher, C.R.; Dressel, A.E.; Silva, J.J.; Scott, J.G. A Globally Distributed Insecticide Resistance Allele Confers a Fitness Cost in the Absence of Insecticide in Aedes aegypti (Diptera: Culicidae), the Yellow Fever Mosquito. J. Med. Entomol. 2023, 60, 494–499. [Google Scholar] [CrossRef]
  267. Diniz, D.F.A.; de Melo-Santos, M.A.V.; Santos, E.M.M.; Beserra, E.B.; Helvecio, E.; de Carvalho-Leandro, D.; dos Santos, B.S.; de Menezes Lima, V.L.; Ayres, C.F.J. Fitness Cost in Field and Laboratory Aedes aegypti Populations Associated with Resistance to the Insecticide Temephos. Parasites Vectors 2015, 8, 662. [Google Scholar] [CrossRef]
  268. Penilla-Navarro, P.; Solis-Santoyo, F.; Lopez-Solis, A.; Rodriguez, A.D.; Vera-Maloof, F.; Lozano, S.; Contreras-Mejía, E.; Vázquez-Samayoa, G.; Torreblanca-Lopez, R.; Perera, R.; et al. Pyrethroid Susceptibility Reversal in Aedes aegypti: A Longitudinal Study in Tapachula, Mexico. PLoS Negl. Trop. Dis. 2024, 18, e0011369. [Google Scholar] [CrossRef]
  269. Rigby, L.M.; Rašić, G.; Peatey, C.L.; Hugo, L.E.; Beebe, N.W.; Devine, G.J. Identifying the Fitness Costs of a Pyrethroid-Resistant Genotype in the Major Arboviral Vector Aedes aegypti. Parasites Vectors 2020, 13, 358. [Google Scholar] [CrossRef]
  270. Vinauger, C.; Chandrasegaran, K. Context-Specific Variation in Life History Traits and Behavior of Aedes aegypti Mosquitoes. Front. Insect Sci. 2024, 4, 1426715. [Google Scholar] [CrossRef] [PubMed]
  271. Silalahi, C.N.; Yasin, A.; Chen, M.E.; Ahmad, I.; Neoh, K.B. Behavioral Responses and Life History Traits of Taiwanese and Indonesian Populations of Aedes aegypti Surviving Deltamethrin-Clothianidin Treatment. Parasites Vectors 2024, 17, 117. [Google Scholar] [CrossRef] [PubMed]
  272. Souza-Neto, J.A.; Powell, J.R.; Bonizzoni, M. Aedes aegypti Vector Competence Studies: A Review. Infect. Genet. Evol. 2019, 67, 191–209. [Google Scholar] [CrossRef]
  273. Chen, T.-Y.; Smartt, C.T.; Shin, D. Permethrin Resistance in Aedes aegypti Affects Aspects of Vectorial Capacity. Insects 2021, 12, 71. [Google Scholar] [CrossRef]
  274. Deng, J.; Guo, Y.; Su, X.; Liu, S.; Yang, W.; Wu, Y.; Wu, K.; Yan, G.; Chen, X.G. Impact of deltamethrin-resistance in Aedes albopictus on its fitness cost and vector competence. PLoS Negl. Trop. Dis. 2021, 15, e0009391. [Google Scholar] [CrossRef] [PubMed]
  275. Tantowijoyo, W.; Tanamas, S.K.; Nurhayati, I.; Setyawan, S.; Budiwati, N.; Fitriana, I.; Ernesia, I.; Wardana, D.S.; Supriyati, E.; Arguni, E.; et al. Aedes aegypti Abundance and Insecticide Resistance Profiles in the Applying Wolbachia to Eliminate Dengue Trial. PLoS Negl. Trop. Dis. 2022, 16, e0010284. [Google Scholar] [CrossRef]
  276. Moyes, C.L.; Vontas, J.; Martins, A.J.; Ng, L.C.; Koou, S.Y.; Dusfour, I.; Raghavendra, K.; Pinto, J.; Corbel, V.; David, J.P.; et al. Contemporary Status of Insecticide Resistance in the Major Aedes Vectors of Arboviruses Infecting Humans. PLoS Negl. Trop. Dis. 2017, 11, e0005625. [Google Scholar] [CrossRef]
  277. Guedes, R.N.C.; Beins, K.; Costa, D.N.; Coelho, G.E.; Bezerra, H.S.S. Patterns of Insecticide Resistance in Aedes aegypti: Meta-Analyses of Surveys in Latin America and the Caribbean. Pest Manag. Sci. 2020, 76, 2144–2157. [Google Scholar] [CrossRef]
  278. van den Berg, H.; da Silva Bezerra, H.S.; Al-Eryani, S.; Chanda, E.; Nagpal, B.N.; Knox, T.B.; Velayudhan, R.; Yadav, R.S. Recent Trends in Global Insecticide Use for Disease Vector Control and Potential Implications for Resistance Management. Sci. Rep. 2021, 11, 23867. [Google Scholar] [CrossRef]
  279. Zheng, C.; Li, S.; Wu, M.; Li, J.; Ma, K.; You, H. Insecticide Binding Mode Analysis and Biological Effects of Acetylcholinesterase Target-Site Resistance Mutations in Spodoptera frugiperda. Pestic. Biochem. Physiol. 2024, 205, 106164. [Google Scholar] [CrossRef]
  280. Saladini di Rovetino, M.; Lueke, B.; Masawang, K.; Piyasaengthong, N.; Kaewwongse, M.; Nobsathian, S.; Fricaux, T.; Nam, K.; d’Alençon, E.; Bullangpoti, V.; et al. Monitoring the Molecular Mechanisms of Insecticide Resistance in Spodoptera frugiperda Populations from Thailand. Pestic. Biochem. Physiol. 2025, 214, 106599. [Google Scholar] [CrossRef]
  281. Ngegba, P.M.; Khalid, M.Z.; Jiang, W.; Zhong, G. An Overview of Insecticide Resistance Mechanisms, Challenges, and Management Strategies in Spodoptera frugiperda. Crop Prot. 2025, 197, 107322. [Google Scholar] [CrossRef]
  282. Chen, Y.; Nguyen, D.T.; Spafford, H.; Herron, G.A. A High-Throughput Multilocus-Amplicon Sequencing Panel to Monitor Insecticide Resistance in Fall Armyworm (Spodoptera frugiperda) (Lepidoptera: Noctuidae). Pest Manag. Sci. 2024, 80, 1510–1522. [Google Scholar] [CrossRef]
  283. Moustafa, M.A.M.; El-Said, N.A.; Alfuhaid, N.A.; Abo-Elinin, F.M.A.; Mohamed, R.M.B.; Aioub, A.A.A. Monitoring and Detection of Insecticide Resistance in Spodoptera frugiperda (Lepidoptera: Noctuidae): Evidence for Field-Evolved Resistance in Egypt. Insects 2024, 15, 705. [Google Scholar] [CrossRef]
  284. Samanta, S.; Barman, M.; Thakur, H.; Chakraborty, S.; Upadhyaya, G.; Roy, D.; Banerjee, A.; Samanta, A.; Tarafdar, J. Evidence of Population Expansion and Insecticide Resistance Mechanism in Invasive Fall Armyworm (Spodoptera frugiperda). BMC Biotechnol. 2023, 23, 17. [Google Scholar] [CrossRef]
  285. Bolzan, A.; Padovez, F.E.O.; Nascimento, A.R.B.; Kaiser, I.S.; Lira, E.C.; Amaral, F.S.A.; Kanno, R.H.; Malaquias, J.B.; Omoto, C. Selection and inheritance of resistance in Spodoptera frugiperda (Lepidoptera: Noctuidae) to chlorantraniliprole and cross-resistance to other diamides. Pest Manag. Sci. 2019, 75, 2682–2689. [Google Scholar] [CrossRef]
  286. Okuma, D.M.; Cuenca, A.; Nauen, R.; Omoto, C. Large-Scale Monitoring of the Frequency of Ryanodine Receptor Target-Site Mutations Conferring Diamide Resistance in Brazilian Field Populations of Fall Armyworm, Spodoptera frugiperda (Lepidoptera: Noctuidae). Insects 2022, 13, 626. [Google Scholar] [CrossRef] [PubMed]
  287. Mavridis, K.; Evangelou, V.; Grigoriadou, A.M.; Papachristos, D.P.; Vontas, J. Molecular Surveillance of Resistance Mutations in Invasive Populations of Spodoptera frugiperda in Europe, for Evidence-Based Pest Control. Pest Manag. Sci. 2025, 81, 4821–4830. [Google Scholar] [CrossRef]
  288. Barbosa, M.G.; André, T.P.P.; Pontes, A.D.S.; Souza, S.A.; Oliveira, N.R.X.; Pastori, P.L. Insecticide Rotation and Adaptive Fitness Cost Underlying Insecticide Resistance Management for Spodoptera frugiperda (Lepidoptera: Noctuidae). Neotrop. Entomol. 2020, 49, 882–892. [Google Scholar] [CrossRef]
  289. Kafle, L.; Joshi, R. Fall Armyworm Threatens Asian Rice Security: A Review of Sustainable Management Strategies. CABI Rev. 2025, 20, 0017. [Google Scholar] [CrossRef]
  290. Wyckhuys, K.A.G.; Akutse, K.S.; Amalin, D.M.; Araj, S.-E.; Barrera, G.; Beltran, M.J.B.; Ben Fekih, I.; Calatayud, P.-A.; Cicero, L.; Cokola, M.C.; et al. Global Scientific Progress and Shortfalls in Biological Control of the Fall Armyworm Spodoptera frugiperda. Biol. Control 2024, 191, 105460. [Google Scholar] [CrossRef]
  291. Idrees, A.; Afzal, A.; Qadir, Z.A.; Li, J. Bioassays of Beauveria bassiana Isolates against the Fall Armyworm, Spodoptera frugiperda. J. Fungi 2022, 8, 717. [Google Scholar] [CrossRef]
  292. Ratto, F.; Bruce, T.; Chipabika, G.; Mwamakamba, S.; Mkandawire, R.; Khan, Z.; Mkindi, A.; Pittchar, J.; Sallu, S.M.; Whitfield, S.; et al. Biological control interventions reduce pest abundance and crop damage while maintaining natural enemies in sub-Saharan Africa: A meta-analysis. Proc. Biol. Sci. 2022, 289, 20221695. [Google Scholar] [CrossRef]
  293. Midega, C.A.O.; Pittchar, J.O.; Pickett, J.A.; Hailu, G.W.; Khan, Z.R. A climate-adapted push-pull system effectively controls fall armyworm, Spodoptera frugiperda (J E Smith), in maize in East Africa. Crop Protect. 2018, 105, 10–15. [Google Scholar] [CrossRef]
  294. Kinyanjui, G.; Mawcha, K.T.; Malinga, L.N.; Soobramoney, K.; Ṋethononda, P.; Assefa, Y.; Okonkwo, C.O.; Ndolo, D. Managing African Armyworm Outbreaks in Sub-Saharan Africa: Current Strategies and Future Directions. Insects 2025, 16, 645. [Google Scholar] [CrossRef]
  295. Yin, J.; Su, X.; Yan, S.; Shen, J. Multifunctional Nanoparticles and Nanopesticides in Agricultural Application. Nanomaterials 2023, 13, 1255. [Google Scholar] [CrossRef] [PubMed]
  296. Araújo, M.F.; Castanheira, E.M.S.; Sousa, S.F. The Buzz on Insecticides: A Review of Uses, Molecular Structures, Targets, Adverse Effects, and Alternatives. Molecules 2023, 28, 3641. [Google Scholar] [CrossRef]
  297. Ataide, L.M.S.; Vargas, G.; Velazquez-Hernandez, Y.; Reyes-Arauz, I.; Villamarin, P.; Canon, M.A.; Yang, X.; Riley, S.S.; Revynthi, A.M. Efficacy of Conventional and Biorational Insecticides against the Invasive Pest Thrips parvispinus (Thysanoptera: Thripidae) under Containment Conditions. Insects 2024, 15, 48. [Google Scholar] [CrossRef]
  298. Zhou, W.; Li, M.; Achal, V. A Comprehensive Review on Environmental and Human Health Impacts of Chemical Pesticide Usage. Emerg. Contam. 2025, 11, 100410. [Google Scholar] [CrossRef]
  299. Li, F.; Xiong, W.; Zhang, C.; Wang, D.; Zhou, C.; Li, W.; Zeng, G.; Song, B.; Zeng, Z. Neonicotinoid insecticides in non-target organisms: Occurrence, exposure, toxicity, and human health risks. J. Environ. Manag. 2025, 383, 125432. [Google Scholar] [CrossRef]
  300. Al Naggar, Y.; Paxton, R.J. The Novel Insecticides Flupyradifurone and Sulfoxaflor Do Not Act Synergistically with Viral Pathogens in Reducing Honey Bee (Apis mellifera) Survival but Sulfoxaflor Modulates Host Immunocompetence. Microb. Biotechnol. 2021, 14, 227–240. [Google Scholar] [CrossRef] [PubMed]
  301. Al Naggar, Y.; Singavarapu, B.; Paxton, R.J.; Wubet, T. Bees under Interactive Stressors: The Novel Insecticides Flupyradifurone and Sulfoxaflor along with the Fungicide Azoxystrobin Disrupt the Gut Microbiota of Honey Bees and Increase Opportunistic Bacterial Pathogens. Sci. Total Environ. 2022, 849, 157941. [Google Scholar] [CrossRef] [PubMed]
  302. El-Din, H.S.; Helmy, W.S.; Al Naggar, Y.; Ahmed, F.S. Chronic Exposure to a Field-Realistic Concentration of Closer® SC (24% Sulfoxaflor) Insecticide Impacted the Growth and Foraging Activity of Honey Bee Colonies. Apidologie 2022, 53, 22. [Google Scholar] [CrossRef]
  303. Al Naggar, Y.; Wubet, T. Chronic Exposure to Pesticides Disrupts the Bacterial and Fungal Co-Existence and the Cross-Kingdom Network Characteristics of Honey Bee Gut Microbiome. Sci. Total Environ. 2024, 906, 167530. [Google Scholar] [CrossRef] [PubMed]
  304. Barathi, S.; Sabapathi, N.; Kandasamy, S.; Lee, J. Present Status of Insecticide Impacts and Eco-Friendly Approaches for Remediation—A Review. Environ. Res. 2024, 240, 117432. [Google Scholar] [CrossRef]
  305. Wang, L.; Zhang, L.; Liu, N.; Gao, X.; Liu, X. Effect of RNAi targeting CYP6CY3 on the growth, development and insecticide susceptibility of Aphis gossypii by using nanocarrier-based transdermal dsRNA delivery system. Pestic. Biochem. Physiol. 2021, 177, 104878. [Google Scholar] [CrossRef]
  306. Arjunan, N.; Thiruvengadam, V.; Sushil, S. Nanoparticle-Mediated dsRNA Delivery for Precision Insect Pest Control: A Comprehensive Review. Mol. Biol. Rep. 2024, 51, 355. [Google Scholar] [CrossRef]
  307. Kaduskar, B.; Kushwah, R.B.S.; Auradkar, A.; Guichard, A.; Li, M.; Bennett, J.B.; Julio, A.H.F.; Marshall, J.M.; Montell, C.; Bier, E. Reversing Insecticide Resistance with Allelic-Drive in Drosophila melanogaster. Nat. Commun. 2022, 13, 291. [Google Scholar] [CrossRef] [PubMed]
  308. Hancock, P.A.; Hendriks, C.J.M.; Tangena, J.A.; Gibson, H.; Hemingway, J.; Coleman, M.; Gething, P.W.; Cameron, E.; Bhatt, S.; Moyes, C.L. Mapping Trends in Insecticide Resistance Phenotypes in African Malaria Vectors. PLoS Biol. 2020, 18, e3000633. [Google Scholar] [CrossRef]
  309. Ayilara, M.S.; Adeleke, B.S.; Akinola, S.A.; Fayose, C.A.; Adeyemi, U.T.; Gbadegesin, L.A.; Omole, R.K.; Johnson, R.M.; Uthman, Q.O.; Babalola, O.O. Biopesticides as a Promising Alternative to Synthetic Pesticides: A Case for Microbial Pesticides, Phytopesticides, and Nanobiopesticides. Front. Microbiol. 2023, 14, 1040901, Erratum in Front. Microbiol. 2024, 14, 1258968. [Google Scholar] [CrossRef] [PubMed]
  310. Priya, A.K.; Alagumalai, A.; Balaji, D.; Song, H. Bio-based Agricultural Products: A Sustainable Alternative to Agrochemicals for Promoting a Circular Economy. RSC Sustain. 2023, 1, 746–762. [Google Scholar] [CrossRef]
  311. Ngegba, P.M.; Cui, G.; Khalid, M.Z.; Zhong, G. Use of Botanical Pesticides in Agriculture as an Alternative to Synthetic Pesticides. Agriculture 2022, 12, 600. [Google Scholar] [CrossRef]
  312. Kansman, J.T.; Jaramillo, J.L.; Ali, J.G.; Hermann, S.L. Chemical ecology in conservation biocontrol: New perspectives for plant protection. Trends Plant Sci. 2023, 28, 1166–1177. [Google Scholar] [CrossRef]
  313. Villavicencio-Vásquez, M.; Espinoza-Lozano, F.; Espinoza-Lozano, L.; Coronel-León, J. Biological Control Agents: Mechanisms of Action, Selection, Formulation and Challenges in Agriculture. Front. Agron. 2025, 7, 1578915. [Google Scholar] [CrossRef]
  314. Liu, T.-X.; Chen, X.-X. Biological control of aphids in China: Successes and prospects. Annu. Rev. Entomol. 2025, 70, 401–419. [Google Scholar] [CrossRef]
  315. Haase, S.; Sciocco-Cap, A.; Romanowski, V. Baculovirus Insecticides in Latin America: Historical Overview, Current Status and Future Perspectives. Viruses 2015, 7, 2230–2267. [Google Scholar] [CrossRef]
  316. Karthi, S.; Vasantha-Srinivasan, P.; Senthil-Nathan, S.; Han, Y.S.; Shivakumar, M.S.; Murali-Baskaran, R.K.; Kalaivani, K.; Radhakrishnan, N.; Park, K.B.; Malafaia, G. Entomopathogenic Fungi Promising Biocontrol Agents for Managing Lepidopteran Pests: Review of Current Knowledge. Biocatal. Agric. Biotechnol. 2024, 58, 103146. [Google Scholar] [CrossRef]
  317. Pijnakker, J.; Vangansbeke, D.; Duarte, M.; Moerkens, R.; Wäckers, F.L. Predators and parasitoids-in-first: From inundative releases to preventative biological control in greenhouse crops. Front. Sustain. Food Syst. 2020, 4, 595630. [Google Scholar] [CrossRef]
  318. Angon, P.B.; Mondal, S.; Jahan, I.; Datto, M.; Antu, U.B.; Ayshi, F.J.; Islam, M.S. Integrated pest management (IPM) in agriculture and its role in maintaining ecological balance and biodiversity. Int. J. Agron. 2023, 2023, 5546373. [Google Scholar] [CrossRef]
Figure 1. Conceptual illustration of insecticide resistance in insects. (A) The diagram contrasts insecticide-susceptible insects (left) with insecticide-resistant insects (right). Susceptible individuals are killed upon exposure to insecticides, whereas resistant individuals possess one or more adaptive mechanisms that enable survival. These mechanisms include reduced cuticular penetration (limiting insecticide entry), enhanced metabolic detoxification, and target-site insensitivity (e.g., mutations at the insecticide’s site of action). (B) Insecticide resistance emerges when these adaptive traits allow certain individuals to survive exposure to otherwise lethal doses, resulting in the selection, persistence, and eventual spread of resistant populations.
Figure 1. Conceptual illustration of insecticide resistance in insects. (A) The diagram contrasts insecticide-susceptible insects (left) with insecticide-resistant insects (right). Susceptible individuals are killed upon exposure to insecticides, whereas resistant individuals possess one or more adaptive mechanisms that enable survival. These mechanisms include reduced cuticular penetration (limiting insecticide entry), enhanced metabolic detoxification, and target-site insensitivity (e.g., mutations at the insecticide’s site of action). (B) Insecticide resistance emerges when these adaptive traits allow certain individuals to survive exposure to otherwise lethal doses, resulting in the selection, persistence, and eventual spread of resistant populations.
Toxics 13 00681 g001
Figure 2. Conceptual framework illustrating the multifactorial influences on the efficacy of Insect Resistance Management (IRM) strategies. The diagram organizes contributing factors into four major domains: Operational Factors, which include grower compliance, adoption rates, and regulatory oversight; Biological Factors, encompassing insect biology, fitness costs linked to resistance alleles, and interactions with natural enemies; Tactical Factors, covering specific IRM implementation strategies such as refuge size and placement, integration with cultural control methods, and broader IPM programs; and Monitoring & Management, emphasizing the critical role of continuous resistance surveillance. The arrows indicate the hierarchical and interconnected relationships among these elements, demonstrating how biological constraints, management practices, and stakeholder behavior collectively influence IRM outcomes.
Figure 2. Conceptual framework illustrating the multifactorial influences on the efficacy of Insect Resistance Management (IRM) strategies. The diagram organizes contributing factors into four major domains: Operational Factors, which include grower compliance, adoption rates, and regulatory oversight; Biological Factors, encompassing insect biology, fitness costs linked to resistance alleles, and interactions with natural enemies; Tactical Factors, covering specific IRM implementation strategies such as refuge size and placement, integration with cultural control methods, and broader IPM programs; and Monitoring & Management, emphasizing the critical role of continuous resistance surveillance. The arrows indicate the hierarchical and interconnected relationships among these elements, demonstrating how biological constraints, management practices, and stakeholder behavior collectively influence IRM outcomes.
Toxics 13 00681 g002
Table 1. Major mechanisms of resistance of insects to insecticides.
Table 1. Major mechanisms of resistance of insects to insecticides.
MechanismDescriptionMolecular/Behavioral Basis & Key ExamplesLevel & Pattern of ResistanceReferences
Behavioral resistanceModified behaviors that reduce contact with treated surfaces.
-
Anopheles arabiensis: Shift from indoor to outdoor resting
-
Leptinotarsa decemlineata (Colorado potato beetle): Avoidance of treated foliage
-
Plutella xylostella (Diamondback moth): Reduced oviposition on sprayed crops
Low–moderate (2–10-fold); delays physiological resistance; compromises contact-dependent interventions[19,20,21]
Penetration resistanceThickened or altered cuticle reduces insecticide uptake.
-
Aedes aegypti, Culex quinquefasciatus: Epicuticular thickening via hydrocarbon remodeling
-
Helicoverpa armigera: Reduced deltamethrin penetration through enhanced cuticular esterases
Moderate (<5-fold); offers cross-class protection; synergizes with other mechanisms[22,23]
Target-site insensitivityMutations at insecticide binding sites reduce compound efficacy.
-
kdr (L1014F/Vssc): Pyrethroid/DDT resistance in Anopheles gambiae
-
Ace-1: OP/carbamate resistance via acetylcholinesterase (AChE) insensitivity
-
Rdl (A302S): Cyclodiene/fipronil resistance via GABA receptor changes
High (100–1000-fold); class-specific; causes rapid control failure; detectable via molecular diagnostics[24,25,26]
Metabolic resistanceIncreased detoxification via overexpressed or amplified enzymes.
-
P450s: CYP6P9a/b overexpression in An. funestus (pyrethroid detoxification)
-
Carboxylesterases (CESs): Coeae1f amplification in Ae. aegypti (OP/pyrethroid hydrolysis)
-
GSTs: GSTe2 upregulation in An. gambiae (DDT dehydrochlorination)
High (10–500-fold); often polygenic; causes broad cross-resistance; undermines new insecticide chemistries[27,28,29,30]
Table 2. Summary of nicotinic acetylcholine receptor (nAChR) mutations associated with resistance or susceptibility to neonicotinoid and spinosyn insecticides in major insect pests in the last decade.
Table 2. Summary of nicotinic acetylcholine receptor (nAChR) mutations associated with resistance or susceptibility to neonicotinoid and spinosyn insecticides in major insect pests in the last decade.
Insect SpeciesnAChR Mutation(s)Insecticide(s)Resistance?References
Aedes aegypti (yellow fever mosquito)No known nAChR target-site mutationNeonicotinoidsNo nAChR target-site mutations found (metabolic resistance via CYPs is highlighted elsewhere (Section 5.1.2))[48]
Frameshift mutation in α6 subunit (32-bp deletion)SpinosadYes (320-fold resistance)[49]
Anopheles gambiae (malaria mosquito)Multiple subunits. Reduced expression (β1, α3, α7)Neonicotinoids (clothianidin, acetamiprid)Partial resistance; 15–23-fold downregulation in field populations[50]
Drosophila melanogaster (fruit fly)β1 subunit: R81T (engineered)NeonicotinoidsYes (Increased tolerance)—CRISPR/Cas9 genome editing; Fitness costs observed[51]
α6 subunit: KnockoutSpinosadYes (High resistance)—CRISPR/Cas9 deletion; No fitness deficits[51]
Ceratitis capitata (medfly)α6 subunit: 3aQ68 * and K352 * (* premature stops)Spinosyns (spinosad)Yes—truncated α6 isoforms (stop codons at Q68 and K352) confer resistance[52]
Bemisia tabaci (silverleaf whitefly)β1 subunit: A58T and R79E (target-site) [38]Neonicotinoids (imidacloprid, thiamethoxam, etc.)Yes—confers resistance [38].
(earlier studies suggested an absence of nAChR mutations [53])
[38,53]
None identifiedSpinosyns (spinosad)No—no target-site mutations reported (resistance mainly metabolic)[53]
Tuta absoluta (tomato borer)α6 subunit: G275E (target-site)Spinosyns (spinosad, spinetoram)Yes—high resistance via G275E[54]
None identifiedNeonicotinoidsNo—no nAChR target-site mutations reported in recent studies
Spodoptera exigua (beet armyworm)α6 subunit: G275E (target-site)Spinosyns (spinosad, spinetoram)Yes—CRISPR knock-in G275E confers resistance[55]
None identifiedNeonicotinoidsNo—no target-site mutations reported
Spodoptera frugiperda (fall armyworm)α6 subunit: G275E (low frequency)SpinosynsPotential transitional resistance—Amplicon sequencing (0.1–1% allele frequency)[56]
Plutella xylostella (diamondback moth)α6 subunit: 3-amino-acid deletion (TM4)Spinosyns (spinosad, spinetoram)Yes—3-aa deletion (IIA) in nAChR α6 underlies ~940-fold resistance[57]
None identifiedNeonicotinoidsNo—no nAChR target-site mutation documented in recent literature
Frankliniella occidentalis (western flower thrips)α6 subunit: G275E (target-site)Spinosyns (spinosad)Yes—G275E associated with spinosad resistance, but α6 knockout confers complete resistance to spinosad[58]
Thrips palmi (melon thrips)α6 subunit: G275E (target-site)Spinosyns (spinosad)Yes—G275E confers spinosad resistance[59]
Myzus persicae (green peach aphid)β1 subunit: R81T (major), V101INeonicotinoids (imidacloprid)Yes—R81T and V101I linked to imidacloprid resistance[33]
Aphis gossypii (cotton aphid)β1 subunit: R81TNeonicotinoids (imidacloprid)Yes—R81T confers neonicotinoid resistance[60]
Nilaparvata lugens (brown planthopper)Nlα2 subunitNeonicotinoids (imidacloprid, dinotefuran)Yes—CRISPR/Cas9 knockout. Nlα2 knockout confers cross-resistance to neonicotinoids[61]
Key: “Yes” indicates nAChR target-site resistance documented; “No” indicates no target-site mutation reported.
Table 3. Microbiome-mediated insecticide resistance mechanisms across insects.
Table 3. Microbiome-mediated insecticide resistance mechanisms across insects.
MechanismInsect HostSymbiont(s)Pesticide(s)Evidence TypeNotesReferences
Direct metabolismBactrocera dorsalisCitrobacter sp. (CF-BD)TrichlorfonGenomics + MetabolomicsDegrades trichlorfon into chloral hydrate and dimethyl phosphite; enhances host survival[158]
Direct metabolismRiptortus pedestrisBurkholderiaFenitrothionBioassays + Functional Enzyme TestsSoil-acquired strains detoxify fenitrothion; induces resistance and modulates host gene expression[161]
Direct metabolismNilaparvata lugensSerratia marcescensBuprofezinGain/loss of symbiontAcquisition alters resistance; symbiont breaks down pesticide[171]
Direct metabolismDrosophila melanogasterMixed gut microbiotaImidaclopridMetabolic comparisonsSymbiont-mediated nitro-reduction complements host oxidative CYP6G1 pathway[172]
Gene regulationNilaparvata lugensWolbachia, ArsenophonusImidaclopridGene expression profilingWolbachia upregulates CYPs and GSTs; Arsenophonus suppresses detox genes; response is strain-specific[167,168]
Gene regulationApis melliferaGut bacteria (e.g., Pantoea, Enterobacter)ClothianidinTranscriptomics + Probiotic RescueDisrupted microbiota leads to P450 gene suppression; reintroduction restores detox response[170]
Gene regulation + Enzyme activityTetranychus urticaeWolbachiaAbamectin, Pyridaben, CyflumetofenRNAi, qPCR, Enzyme AssaysUpregulates TuCYP392D2 and TuGSTd05; increases GST activity; abamectin increases Wolbachia abundance[166]
Direct + Indirect (dual)Aphis gossypiiSphingomonasImidaclopridDual-mode functional profilingChemical degradation and host P450 upregulation demonstrated[173]
Immune modulationLymantria disparMixed gut microbiotaPyrethroidsImmune gene profilingMicrobial shifts affect Toll/IMD signaling and Nrf2-mediated detox enzyme expression[175]
Table 4. Comparative analysis of resistance to insecticides traits in Aedes aegypti and Spodoptera frugiperda, highlighting mechanisms, evolutionary dynamics, and control implications.
Table 4. Comparative analysis of resistance to insecticides traits in Aedes aegypti and Spodoptera frugiperda, highlighting mechanisms, evolutionary dynamics, and control implications.
TraitAedes aegyptiSpodoptera frugiperdaEvolutionary Insight
Primary mechanismTarget-site (kdr) + P450sMetabolic (P450s) + RyR mutationVector-herbivore divergence in adaptation
Resistance spread~10–20 years regionally<5 years globallyTrade-facilitated gene flow accelerates resistance
Fitness costReduced fecundity, prolonged developmentMinimal (behavioral compensation)Metabolic flexibility buffers trade-offs
Key vulnerabilitySequestration (OBPs)RNAi susceptibilityTaxon-specific weaknesses enable precision control
Table 5. Resistance to insecticide management (IRM) strategies.
Table 5. Resistance to insecticide management (IRM) strategies.
ApproachCore PrincipleImplementation Examples and Scientific Rationale
MoA rotationAlternate MoA classes to disrupt resistance selection.
-
Agriculture: Rotate neonicotinoids (MoA 4A) → spinosyns (MoA 5) → diamides (MoA 28)
-
Public health: IRS switch pyrethroids (MoA 3A) → carbamates (MoA 1A) annually
Mixtures & synergistsCombine compounds with distinct targets or inhibit detoxification.
-
PBO-nets: Pyrethroid + piperonyl butoxide (P450 inhibitor) for An. gambiae control
-
AChE/GABA co-targeting: Neonicotinoid (MoA 4A) + fipronil (MoA 2B) tank mixes
Refugia & IPMMaintain susceptible alleles; reduce pest density non-chemically.
-
Structured refugia: ≥20% non-Bt maize adjacent to Cry1Ab fields
-
Cultural: Trap cropping (Brassica spp. for diamondback moth) + sanitation (container removal for Ae. aegypti)
Diagnostic monitoringDeploy resistance data for threshold-based interventions.
-
Molecular: kdr allele frequency qPCR (threshold: >60% → switch MoA)
-
Bioassay: WHO tube tests (mortality < 90% → avoid pyrethroids)
Biological/genetic toolsExploit natural enemies or genetic mechanisms.
-
Biocontrol: Beauveria bassiana GHA strain (EPF) against resistant Bemisia tabaci
-
Genetic: Wolbachia (wMel)-mediated population suppression in Ae. aegypti
Policy & stewardshipEnforce regulations to delay resistance.
-
IRAC MoA classification: Mandatory rotation schedules per crop/pest
-
WHO GPIRM: National bans on agriculture-grade pyrethroids for malaria control
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

Al Naggar, Y.; Fahmy, N.M.; Alkhaibari, A.M.; Al-Akeel, R.K.; Alharbi, H.M.; Mohamed, A.; Eleftherianos, I.; El-Seedi, H.R.; Giesy, J.P.; Alharbi, H.A. Mechanisms and Genetic Drivers of Resistance of Insect Pests to Insecticides and Approaches to Its Control. Toxics 2025, 13, 681. https://doi.org/10.3390/toxics13080681

AMA Style

Al Naggar Y, Fahmy NM, Alkhaibari AM, Al-Akeel RK, Alharbi HM, Mohamed A, Eleftherianos I, El-Seedi HR, Giesy JP, Alharbi HA. Mechanisms and Genetic Drivers of Resistance of Insect Pests to Insecticides and Approaches to Its Control. Toxics. 2025; 13(8):681. https://doi.org/10.3390/toxics13080681

Chicago/Turabian Style

Al Naggar, Yahya, Nedal M. Fahmy, Abeer M. Alkhaibari, Rasha K. Al-Akeel, Hend M. Alharbi, Amr Mohamed, Ioannis Eleftherianos, Hesham R. El-Seedi, John P. Giesy, and Hattan A. Alharbi. 2025. "Mechanisms and Genetic Drivers of Resistance of Insect Pests to Insecticides and Approaches to Its Control" Toxics 13, no. 8: 681. https://doi.org/10.3390/toxics13080681

APA Style

Al Naggar, Y., Fahmy, N. M., Alkhaibari, A. M., Al-Akeel, R. K., Alharbi, H. M., Mohamed, A., Eleftherianos, I., El-Seedi, H. R., Giesy, J. P., & Alharbi, H. A. (2025). Mechanisms and Genetic Drivers of Resistance of Insect Pests to Insecticides and Approaches to Its Control. Toxics, 13(8), 681. https://doi.org/10.3390/toxics13080681

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

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop