Next Article in Journal / Special Issue
The State of the Art of Lethal Oviposition Trap-Based Mass Interventions for Arboviral Control
Previous Article in Journal
A Nutritional Profile of the Trap-Nesting Wasp Trypoxylon lactitarse (Hymenoptera: Crabronidae): Comparison of Sexes and Overwintering and Non-Overwintering Generations
Previous Article in Special Issue
Toxicity and Physiological Actions of Carbonic Anhydrase Inhibitors to Aedes aegypti and Drosophila melanogaster
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

RNA Interference for Mosquito and Mosquito-Borne Disease Control

Department of Pathobiological Sciences, University of Wisconsin-Madison, Madison, WI 53706, USA
*
Author to whom correspondence should be addressed.
Submission received: 2 November 2016 / Revised: 15 December 2016 / Accepted: 21 December 2016 / Published: 5 January 2017

Abstract

:
RNA interference (RNAi) is a powerful tool to silence endogenous mosquito and mosquito-borne pathogen genes in vivo. As the number of studies utilizing RNAi in basic research grows, so too does the arsenal of physiological targets that can be developed into products that interrupt mosquito life cycles and behaviors and, thereby, relieve the burden of mosquitoes on human health and well-being. As this technology becomes more viable for use in beneficial and pest insect management in agricultural settings, it is exciting to consider its role in public health entomology. Existing and burgeoning strategies for insecticide delivery could be adapted to function as RNAi trigger delivery systems and thereby expedite transformation of RNAi from the lab to the field for mosquito control. Taken together, development of RNAi-based vector and pathogen management techniques & strategies are within reach. That said, tools for successful RNAi design, studies exploring RNAi in the context of vector control, and studies demonstrating field efficacy of RNAi trigger delivery have yet to be honed and/or developed for mosquito control.

1. Introduction

Insecticide resistance to DDT was originally documented in Aedes, Culex, and Anopheles target species less than a decade after the introduction of DDT for public health campaigns in Europe and the USA [1,2,3,4]. This trend has continued the world over for every class of chemical insecticides (pyrethroids, organochlorines, organophosphates, carbamates), with documented resistance in hundreds of mosquito species across sixty countries [5,6,7]. To combat the proliferation of pesticide resistant mosquito vectors and continue to effect control of mosquito vector-borne diseases, Integrated Vector Management and Integrated Vector Borne Disease Management programs will require alternatives to chemical pesticides [7,8].
RNA interference (RNAi) may lend itself to the cause through the suppression of gene products involved in key physiologies that impact mosquito survival, fecundity, behavior, or vector status. The RNAi pathway functions as a powerful subcellular anti-viral mechanism that post-transcriptionally suppresses mRNA transcripts based on sequence complementarity to double-stranded RNAs (dsRNAs) [9]. The pathway can be manipulated to suppress a given target gene by virtue of introducing a complementary dsRNA to mosquito cells. With hundreds of documented effective RNAi triggers targeting mosquito and pathogen genes, there is an expansive arsenal of anti-vector and anti-pathogen targets that could be harnessed for mosquito and mosquito-borne disease control strategies. RNAi experiments in a number of Aedes, Anopheles, Culex, and Armigeres disease vector species have resulted in disruption of processes including: morphogenesis, olfaction for host seeking and oviposition, blood feeding, fertility, fecundity, and survival [10,11,12,13,14,15,16,17,18,19].
In order to achieve these same detrimental phenotypes in wild populations, RNAi triggers must be delivered to the target species and life stage with consideration for environmental and abiotic factors including: UV, ribonucleases, microbes, dissipation and dilution in aqueous environs and on solid substrates [20,21,22,23]. RNAi triggers must also be delivered to target mosquito species using field feasible applications. Infrastructure and techniques for a variety of interventions already exist to deliver chemical and biological pesticides to vector mosquitoes. Existing intervention frameworks include topical and contact applications for adults (e.g., aerial and residual spraying and long-lasting insecticidal nets (LLINs)) and per os or contact applications for aquatic stages (see Figure 1) [24,25,26]. RNAi knockdown in larvae by per os exposure is efficacious using scalable bacterial and yeast expression systems, demonstrating potential for RNAi in larval control applications [13,27]. Novel interventions have also been explored to provide oral applications to adults in the form of Attractive Toxic Sugar Baits (ATSB) [28,29,30,31,32,33,34,35,36,37]. Formulations for ATSBs include simple sucrose solutions and complex mixtures of fruit sugars with minimal effects on non-target organisms [30,36,38]. Formulations can be delivered either via spraying on plant sources or in bait stations. Surprisingly high ATSB efficacy has been found in spray formulations on flowering and non-flowering plants in arid and wet climates [30,31,35,36,39]. Additionally, strategically placed ATSB stations near breeding sites (dubbed Attractive Baited Oviposition Trap, ABOT) or indoors can attract and kill vector species in proximity to people [32,33,40,41]. Although ATSB have not been studied in conjunction with RNAi, successful gene silencing by oral exposure routes has been documented using sucrose meals and artificial blood meals demonstrating the potential in combining these control approaches [11,42]. By the same logic, mosquitocidal RNAi triggers could be applied to target essential genes for embryogenesis in Attractive Baited Oviposition Traps (ABOTs) [43,44,45]. In both baited strategies and more traditional insecticidal delivery approaches (ultra-low volume or residual sprays, or LLINs), RNAi triggers may be more efficacious in combination with biotic (e.g., a virus, yeast or bacterial expression system) or abiotic (e.g., nanoparticle) systems that mediate both protection and uptake of RNAi triggers [13,27,46,47,48,49,50,51].
Despite these knowledge and application gaps in field-relevant delivery systems, RNAi triggers offer vastly improved species-specificity with diminished environmental toxicity compared with chemical pesticides. The majority of chemical insecticides have neurotoxic activity with potential safety implications for a diverse array of organisms including beneficial arthropods, fish, and mammals [60,61,62,63,64]. By comparison, RNAi has the capacity to be as specific as the gene or sequence targeted. As such, RNAi triggers can even target specific splice variants within the target organism as demonstrated in silencing the doublesex sex dependency gene [13]. While there are numerous conserved sequences shared between distantly-related species, even ancient genes contain enough variation to diminish cross-species knockdown if an RNAi trigger can be designed to avoid regions with >21 contiguous matching nucleotides. For instance, in Drosophila species the γ-tubulin 23C gene shares up to 96% sequence identity but no cross-species mortality occurs following exposure to RNAi triggers designed for a specific species of Drosophila [65]. Likewise no cross-species mortality was found for RNAi triggers for vacuolar ATPase in Drosophila melanogaster, Manduca sexta, Tribolium castaneum, and Acyrthosiphon pisum [65]. Although further investigation and testing of cross-species effects should be performed, conceptually, control RNAi triggers can be designed around sequences unique to the target, with specificity far greater than chemical insecticides.

2. RNAi Triggers with Potential Mosquito Control Applications

2.1. RNAi to Induce a Lethal Phenotype

There is great interest in developing RNAi to act as direct alternative to chemical insecticides by virtue of suppressing essential genes leading to RNAi-induced mortality. The most direct pathway to inducing mortality is via manipulation of programmed cell death. Apoptosis in insects is regulated by the Inhibitor of Apoptosis Protein 1 (IAP1) [66]. IAP1 primarily functions in inhibiting caspases and thereby preventing a cascade of subcellular events that result in degradation of hundreds of cellular components and subsequent cell death [66,67,68]. In Ae. aegypti, knockdown of IAP1 by RNAi results in activation of apoptosis and rapid mortality both in vitro and in vivo, making the gene a key target for insecticidal RNAi [10,52,53,54,55]. Mortality induced by IAP1 silencing can be reversed if initiator caspases such as Ae. aegypti dronc, which IAP1 inhibits, also are suppressed [10]. In a screen of 109 genes in An. gambiae, IAP1 was noted as one of two genes for which suppression significantly reduced cell growth and viability in vitro [69]. The other gene identified in this study was a ubiquitin-like/ribosomal fusion gene (AGAP008001). Interestingly, a similar ubiquitin gene was found to be strongly linked to apoptosis and cell survival in a genome wide screen of D. melanogaster [70]. In total, the D. melanogaster screen revealed 438 dsRNAs that induced apoptosis in vitro. This screen both validated known knockout lethal phenotype genes and identified novel essential genes including the ubiquitin-like Ubi-p63E and CG11700 genes as well as DNA binding proteins bss and CG15455 [70]. Beyond IAP1, apoptosis-associated lethal phenotype genes are relatively unexplored in mosquitoes. Considering that at least half of all D. melanogaster genes have orthologs in An. gambiae, there likely is a cache of highly potent, mortality-inducing target genes in mosquito species [71].
Beyond apoptosis, other essential ubiquitously and constitutively expressed genes are targeted for insecticidal activity such as tubulins and Vacuolar ATPase proton pumps (V-ATPases). Tubulin suppression by RNAi induces high levels of mortality in a diverse array of insects including Ae. aegypti, Drosophila species, Blattella germanica, Diabrotica virgifera virgifera, and Rhodnius prolixus [12,65,72,73,74]. In Ae. aegypti, mortality is induced by soaking larvae with a dsRNA targeting β-tubulin without transfection [12]. RNAi-based suppression of another cytoskeletal element, actin, in Ae. aegypti larvae has also led to increased Cry11Aa toxin sensitivity [75].
V-ATPase mutations have long been associated with lethal phenotypes in Drosophila and as such are considered vital to maintaining membrane proton translocation [76,77]. RNAi knockdown of various V-ATPases in Drosophila, M. sexta, T. castaneum, A. pisum, Bactericerca cockerelli, Helicoverpa armigera, and D. virgifera virgifera all result in death of the insect [65,73,78,79]. Thus far, there are several reports of V-ATPase silencing in mosquitoes with mixed results. The impact of V-ATPase subunit silencing appears to differ based on subunit, exposure route, phenotypic assay, and observed knockdown. Kang et al. (2014) explored the impact of V-ATPases on Dengue virus in Ae. aegypti midguts, and noted a dramatic reduction in virus titer following suppression of V-ATPase subunits by RNAi [80]. In addition to reduced Dengue virus titer, suppression of V-ATPase subunit V0B by injection resulted in 80.5% knockdown (72 h post exposure) leading to reduced longevity (measured over 50 days), fecundity (98% reduction), and fertility (19% reduction). Therefore the reduction in virus may be due to an overall impact on mosquito health following V-ATPase silencing. Conversely, Coy et al. (2012) showed that dsRNA delivered orally suppresses V-ATPase subunit C by 60% (168 h post exposure) without any noticeable death [11]. Death may have not been observed by Coy et al. (2012) because animals were monitored only for 48 h following exposure (as compared to 50 days observed by Kang et al. (2014)) [11,80]. RNAi suppression of V-ATPases in mosquito larvae also leads to adverse health outcomes. For instance, RNAi-based V-ATPase subunit β suppression increased Cry11Aa toxin hypersensitivity in Ae. aegypti larvae [75]. Additionally, using a short hairpin RNA (shRNA) Densovirus expression system to silence ATPase subunit A significantly reduced the lifespan of Ae. albopictus larvae [81]. Variation in the target subunit could explain the variation of these results because not all V-ATPase genes are considered essential. A genome- wide screen in Drosophila identified 33 distinct V-ATPase genes, with many subunits having multiple copies or splice variants [77]. Of these genes, only those associated with plasma membranes are known to have lethal knockout phenotypes. As such, V-ATPase as an RNAi target can produce rapid mortality, but may require RNAi triggers designed to multiple, non-redundant, or essential subunit sequences.
Finally, death can also be a by-product of suppression of genes with roles in processes beyond cell death and cytoskeletal structure. For instance, suppression of An. gambiae Serpin2 (SRPN2), which functions in processing prophenoloxidase, proves lethal to older mosquitoes and could be useful as a late life insecticide [78]. An RNAi trigger for prophenoloxidase III in Ar. subalbatus induced significant mortality and gross deformities in eclosing adults. COatamer Proteins (COPs) are required for vesicle formation and nutrient trafficking in the midgut. RNAi suppression of COPs induces death in H. armigera and Ae. aegypti [79,82,83,84]. And RNAi suppression of chitin synthase I in Ae. aegypti larvae is lethal during development [19,49,85]. Ultimately, it is very likely that suppression of genes with any number of functions will produce a lethal outcome, but the immediacy and potency of insecticidal RNAi triggers will differ dramatically depending on RNAi design, delivery strategy and timing, and the particular target physiology [86].

2.2. Olfaction

Olfaction is essential for hematophagy and associated fundamental mosquito behaviors including host-seeking and oviposition. Disruption of olfactory processes is a particularly useful approach to repelling a host-seeking mosquito—witness DEET, the most widely used active ingredient in repellent formulations on the market [87,88]. There are hundreds of chemosensory odorant (OR) and gustatory receptors (GR), ionotropic glutamate (IR) receptors, and odorant binding proteins (OBP) characterized in vector species including An. gambiae, Ae. aegypti, and Cx. pipiens [89,90,91,92,93]. RNAi knockdown of key sensory receptors impacts numerous essential behaviors of interest for mosquito control. For instance, silencing GR1 & 3 limits CO2 detection required for host-seeking [15]. Beyond host-seeking, rapid probing and blood meal engorgement significantly decrease as a result of RNAi-based suppression of OR8 and OR49 in the stylet neurons of Ae. aegypti [94]. In Ae. albopictus, RNAi knockdown of OR7 results in reduced blood feeding success and loss of human host preference in a human or mouse choice experiment [95]. Olfaction also is essential for responding to ovipositional cues. In Cx. pipiens, RNAi suppression of OR37 & OR99 results in reduced egg raft production and ability to sense the stimulant 4-ethylphenol [96].
Suppressing olfaction genes may offer non-chemical approaches to altering blood-feeding behavior and oviposition with clear public health benefits. However achieving this goal requires better understanding of the sequence diversity and evolution rates of olfaction genes in wild mosquito populations. As mentioned above, mutations in the orco gene can alter the repellency effect of DEET in Ae. aegypti demonstrating the potential loss of efficacy that can occur through odorant receptor mutations [87]. Another potential hurdle for an RNAi-based intervention targeting olfaction may be the increased dosage required to suppress genes in the head and antennae. In a study by Das et al. (2008) “7–8 times more dsRNA” was required to achieve gene suppression in the antennae compared to the carcass [97].

2.3. Blood Feeding & Digestion

Blood feeding stimulates differential expression in 50% (2388 upregulated transcripts) of all genes in An. gambiae [98,99]. Disruption of any of these genes immediately post-feeding could disrupt co-ordination of digestion, diuresis, can hamper oogenesis, or cause death. Upstream of digestion, suppression of genes required for probing and feeding success can block successful blood feeding. In Ae. aegypti, suppression of the aegyptin gene results in drastically reduced blood meal uptake and egg output [17]. Additionally RNAi silencing of circadian rhythm (timeless, chryptochrome 1, takeout 1–3) odorant receptor genes have been linked to reduced blood feeding success in An. gambiae and Ae. aegypti respectively [94,97]. After blood feeding, suppression of genes that encode digestive proteins including gbf1, gap1/gap2, late trypsin, AaSPVI, and AaSPVII in Ae. aegypti result in impaired oogenesis but not mortality [83,100]. Other digestive genes including COPI and ARF1/ARF4 induce rapid mortality following a blood meal indicating their importance in digestion [82,83].
In addition to digestion, impairing diuresis via suppression of arginase and urate oxidase delays vitellogenesis and thereby stalls oogenesis following a blood meal in Ae. aegypti [101]. Overend et al. (2015) state the importance of diuresis with “genes that show conservation…in the Malpighian tubules of Anopheles and Drosophila are likely to be essential for survival” [102]. Indeed, RNAi induced suppression of diuresis via aquaporin suppression increases fluid retention resulting in mosquito tolerance to desiccation improving survival under stress [103,104].

2.4. Reproduction

Beyond inducing rapid mortality, the next logical course to reduce vector populations is to interfere with sexual development and reproduction. Preventing oogenesis, diminishing the number of female adults through sexual dependency genes, and producing sterile males are viable approaches to population control. Numerous genes limit egg batch size and nutrient uptake through disruption of nutrient regulation and developmental processes in the oocyte [18,27,56,83,100,105,106,107,108,109]. For instance, targeting genes such as kir in An. gambiae, Catalase 2 in Ae. aegypti, and Cx. pipiens (Insulin like peptides) adult females can have no impact on survival but significantly reduce egg output or egg length [107,110,111]. In these studies, knockdown of target genes persisted for 3–11 days post exposure, meaning oogenesis can be interrupted even if a mosquito takes a blood-meal many days after exposure.
Fecundity and sex bias also have been heavily investigated as a means of vector control. Sterile males can be generated via suppression of testis specific genes such as zero population growth or transglutaminase genes, which function in sperm development and storage respectively [14,112]. In a study by Whyard et al. (2015) [13], feeding larvae E. coli expressing dsRNAs targeting multiple male testes genes resulted in sterility in 92% of emerging male adults. This study also suppressed the female splice variant of the mosquito sex dependency gene doublesex in larvae, with the result that 97% of emerging adults were male. Suppression of other sex dependency genes such as transformer 2 using an inducible RNAi-plasmid transformation was also found to result in 70% male progeny following mating of transformed adult males and females [113].

2.5. Embryogenesis

Embryogenesis is a transcriptionally intensive process. In Ae. aegypti, 8400 genes have altered transcript levels, and in An. gambiae the embryo contains more differently transcribed sequences (624–1009) than any other life-stage [114,115]. The embryo is an attractive life stage for delivery of RNAi triggers because cell uptake in progenitor cells conceivably facilitates widespread bio-distribution of the trigger following development. This is evident in studies which utilize siRNAs without need for transfection reagents [57,116,117]. There are some examples of RNAi-based studies of gene suppression phenotypes in embryos including regulation of embryonic nerve cord development by semaphorin A, frazzled, and commissureless2 [57,116,117]. However, studies designed to explore genes essential to embryonic development in mosquitoes are limited and typically utilize invasive injection methods. Further, when embryos are exposed to an RNAi trigger, the impact of knockdown is measured in later life stages. For instance, suppression of a Na+ methionine symporter (AeNat5) in Ae. aegypti embryos resulted in larval mortality but no impact on embryos was measured [118]. In the context of delivery, the chorion may limit direct delivery of RNAi triggers; that said, there is one example of non-invasive RNAi trigger delivery to eggs via soaking [118]. In this case, eggs were incubated in dsRNA until the time of hatching resulting in 56%–90% reduction in target gene expression in a dose-dependent manner [118]. Knockdown in the embryo also has been demonstrated indirectly through plasmid expression vectors [119,120]. Here plasmid injected into adults passed to the progeny and induced up to 99% suppression of the target gene following heat shock in progeny. There are also examples of the longevity of RNAi knockdown in adults providing up to 95% knockdown for at least 11 days post-injection in An. gambiae [107]. Therefore, suppression of embryo targets is potentially feasible following exposure of the parent (deemed parental RNAi or pRNAi in other insects) to embryo-specific RNAi triggers [121]. However this avenue has yet to be explored.

2.6. Larval & Pupal Development

Management of larval stage insects is a key component of any pest or vector control strategy. In larvae, RNAi is functional throughout the course of development with knockdown being inducible either by direct exposure or via inheritance from exposed parents [119,120,122,123]. Direct exposure of larvae has proven effective when administering RNAi triggers through injection, per os, soaking, or rehydration in aqueous solutions [12,13,75,123,124,125]. Ae. aegypti larvae soaked in an RNAi trigger for β-tubulin showed enhanced mortality [12]. It is not clear if soaking serves to introduce the RNAi trigger per os, through the cuticle, via the anal papillae or by another route. Many different delivery approaches have been used for per os RNAi trigger delivery in larvae, including naked dsRNA in buffer or water [126,127]. Although naked dsRNA can be internalized at high concentration in sterile conditions in the lab, dsRNA in field settings could be subject to rapid environmental degradation [21,22]. To prevent RNAi trigger degradation, abiotic and biotic delivery systems including the use of Effectene® (Qiagen) liposomes, chitosan nanoparticles, E. coli expression systems, and Pichia pastoris expression systems have been explored [27,47,49,75,122,123,128,129]. Currently, at least five published studies have measured significant knockdown of 10 target genes using dsRNA encapsulated in liposomes [75,122,123,130,131]. In each case, neonate larvae were exposed en masse with knockdown measured at fourth instar. Chitosan is currently the most economical and environmentally safe nanoparticle delivery system for larvae. Chitosan has the added benefit of having anti-microbial activity when complexed in particles which could limit dsRNA degradation by microbes [132]. Numerous studies have utilized chitosan to form nanocomplexes with dsRNA and subsequently fed to larvae of both Ae. aegypti and An. gambiae [47,48,49,128,129]. Alternatively, both E. coli and P. pastoris expression systems facilitate even cheaper, scalable dsRNA delivery which protect dsRNA until consumption [13,27] (see Section 3.5 and Section 3.6).
For pupae, RNAi trigger exposure requires injection because pupae are a non-feeding stage. Nevertheless, injection of RNAi triggers into Aedes, Armigeres, and Culex species pupae produces knockdown of targets up to 90% [13,19,96]. Pupal development can be disrupted following direct injection of RNAi triggers for prophenoloxidase III [19,85]. Although pupae cannot be exposed per os, RNAi trigger exposure in larvae, or even adult female F0s could provide sufficient long term persistence to suppress pupal genes [119,120]. Indeed Mysore et al. (2014) suppressed pupal brain development following larval exposure to dsRNA chitosan nanoparticles targeting the single-minded gene (AAEL011013) [128].

2.7. Detoxification and Insecticide Metabolism/Resistance

Because RNAi can suppress detoxification and pesticide resistance genes, RNAi triggers could act as synergists to enhance insecticidal activity. Using RNAi to increase insect sensitivity to toxins was an early concept in agricultural pest control. First demonstrated in cotton bollworm, RNAi-silencing of cytochrome p450 resulted in larval susceptibility to toxic plant metabolites [133]. The use of synergists has long extended the use of many chemical pesticides and have potential for extending the use of LLINs for malaria control [134]. Although there is an abundance of literature surrounding genes identified to play roles in insecticide detoxification and desensitization, there is no immediate solution to counteract such mechanisms [7]. RNAi triggers could be applied to suppress genes which are upregulated in resistant species and strains of mosquito. Genes with p450 reductase activity as well as ABC transporters are related to detoxification and efflux of numerous insecticides [7]. In Ae. aegypti larvae, exposure to the oganophosphate temephos increases expression of a ABC type transporter P-glycoprotein eightfold [127]. RNAi silencing of the P-glycoprotein resulted in no mortality but increased larval sensitivity to temephos by ~25%.
Synergistic action of RNAi trigger knockdown is not limited to insecticide resistance genes such as noted in suppression of chitin synthase 1 in Ae. aegypti larval midguts. Here disruption of the peritrophic matrix facilitates uptake of diflubenzuron, calcofluor white, and dithiothreitol leading to mortality [49].

2.8. RNAi-Induced Pathogen Resistance

RNAi is highly attractive as a technology to parallel chemical insecticides, counter basic mosquito physiology and impact populations. Unlike chemical approaches, RNAi could also be harnessed to impact vector competence, or the inherent capacity of a mosquito to support the development and transmission of a pathogen. This concept originated in a body of mosquito-arbovirus studies that constituted the first evidence that RNAi is an antiviral innate immune response. In mosquito cells, expression of an antisense segment of the La Crosse Virus was noted as a form of “intracellular immunization” because these cells had far lower LACV titers than control cells [135]. The same was true for Dengue (DENV) or Yellow fever virus (YFV) infection [136,137,138]. This led eventually to the production of transgenic mosquitoes with RNAi-mediated resistance to DENV based on production of an RNAi trigger from a midgut-specific and blood meal inducible promoter [139,140]. Conversely, an RNAi-based approach that specifically targets virus-infected mosquitoes could be considered to selectively kill those mosquitoes that contract an infectious blood meal [141]. Disabling the RNAi machinery is lethal for virus-infected mosquitoes. Suppressing RNAi results in uncontrolled virus replication and dissemination and death of the mosquito host in Ae. aegypti-Sindbis virus interactions [142,143].
Although RNAi cannot be used to directly target malaria parasites developing in the mosquito host, leads for RNAi approaches to Plasmodium parasite control in the vector abound because the innate immune response to these parasites has been studied extensively. For example, RNAi suppression of a number of proteins involved in complement, signal modulation (e.g., serine proteases) and melanization pathways decreases numbers of midgut-stage parasites as reviewed by Blandin et al. (2008) [144]. Furthermore, Caspar, a negative regulator of an immune-responsive transcription factor, suppresses parasite numbers in the midgut, as does suppression of a midgut metalloprotease [145,146].
The mosquito-borne filarial worm parasites present yet another potential scenario for RNAi-based interventions for mosquito-borne disease control. Animal parasitic nematodes are often considered recalcitrant to RNAi-based gene suppression [147]. However, the mosquito-borne filarial worm, Brugia malayi, proved amenable to RNAi by injection of an RNAi trigger for a B. malayi cathepsin gene, with a suspected role in molting, into the mosquito host body cavity. Parasites exposed to this trigger at key life stages proved to be significantly less motile than those from control groups, and thereby are highly unlikely to be successfully transmitted [148].

3. Delivery Systems for RNAi Triggers In Vivo

3.1. Naked RNA and Modified Nucleic Acids

A variety of effective oral delivery systems using unmodified dsRNAs have been developed for larvae. In some cases larvae are soaked in large quantities of naked RNAi triggers resulting in suppression of the target after consumption [13,124,126,127]. However the aquatic environment of larvae in the field will inevitably lead to degradation of unprotected RNAi triggers [22]. Additionally the presence of midgut ribonucleases prevents successful RNAi trigger delivery in the desert locust Schistocerca gregaria and the German cockroach B. germanica [72,150]. Therefore, while oral delivery of dsRNA can successfully silence genes in mosquito larvae, knockdown may be reduced by degradation. This was noted in Ae. aegypti where vATPase C dsRNA was found partially degraded in tissues after 24 h [11]. The transcript was reduced by 60% but failed to kill the mosquito. Addition of chemical modifications to siRNA and dsRNA prevent degradation by nucleases without impeding RNAi. Modifications including incorporation of 2′-methoxyl-nucleotides and 5′ polyethylene glycol addition have sufficiently suppressed a Rieske iron–sulfur gene and an acetylcholine esterase gene in the diamondback moth Plutella xylostella [151,152]. These modified siRNAs were sufficient to induce mortality in P. xylostella when sprayed onto cabbage leaves demonstrating potential for modified siRNAs to be used in ATSB or other per os exposure applications targeting mosquito species. The ability for modified nucleotides to be delivered by per os exposure has been demonstrated in Anopheles stephensi. In this case a single stranded antisense morpholino reduced transcript of the target anti-mitogen-activated protein kinase by 60% following administration via a synthetic blood-meal [42]. These oligos are chemically altered preventing enzymatic degradation and so will not be naturally degraded in the environment but are still UV labile. While morpholino induced gene suppression is not considered ‘RNAi’ it is akin to the original interference hypothesis, and mimics early studies using anti-sense RNA to block viral transcripts [135].
In adults, oral exposure of naked dsRNA has also proven effective when exposed via a nuclease free medium. Feeding a 3%–5% sucrose solution via capillary tubes containing 3–16 mg/mL dsRNA targeting jmtA resulted in suppression and 47% reduction in egg production in Ae. aegypti [27]. In terms of environmental hardiness there are also good examples of potent naked dsRNA when attached to surfaces. In a study targeting the Colorado potato beetle larvae (Leptinotarsa decemlineata) potato plant leaves were coated with an aqueous solution of dsRNA targeting actin and left to dry [23]. dsRNA dried onto the leaf surface was found not only to be resistant to washing but also withstood greenhouse conditions and UV as larval mortality reached 100% at least four weeks post initial applications. These studies show that naked dsRNA may be effective in liquid or sprayed Attractive Toxic Sugar Bait (ATSB) applications.

3.2. Insect Plasmid Expression Systems

Plasmid expression systems offer major advantages over in vitro production of dsRNA. Firstly, expression of dsRNAs using bacterial systems have long been utilized to greatly reduce the cost of synthesis for nucleic acids and proteins. These plasmids contain a promoter followed by the target sequence of interest, a loop region, and the reverse complement of the sequence of interest, which when transcribed form a long hairpin dsRNA. Bacterial production of dsRNA still requires the dsRNA to be delivered to the mosquito following production, which has proven effective per os [13]. Alternatively, plasmids with insect specific promoters such as β-tubulin or actin 5C can facilitate production of the dsRNA within target cells of the mosquito [17,119,120,153,154,155,156]. Insect expression plasmids can also persist through multiple generations, with dsRNA production being inducible when using heat shock promoters [119,120]. The main drawback is delivery, since typically direct injection of embryos or adults is required to enable uptake into different tissues and bacterial expression systems release dsRNA into the gut. However, delivery of plasmids into insect cells by virtue of nanoparticle complexes using cell penetrating peptides has been demonstrated [157]. The next step will be to deliver stability expressed insect expression plasmids via a non-invasive route.

3.3. Nanoparticle Delivery Systems

Nanoparticle complexes can provide protection of RNAi triggers from degradation in the environment and the midgut. The composition of nanoparticles varies greatly from biomacromolecular material like chitosan or peptides to chemically produced liposomes and polyethylene glycol (PEG) [49,50,72,157]. Biomacromolecular nanoparticles are advantageous as they will be degraded rather than accumulating in the environment. Of these, chitosan has been most widely reported as an RNAi trigger delivery system for mosquitoes, with efficacy when delivered orally to larvae and or injected in pupae and adults tatgetting an array of tissues [46,47,48,49,128,129,158]. Addition of quantum dots has even led to death following oral delivery of dsRNAs suppressing several essential genes in Ae. aegypti larvae [46]. Cell Penetrating Peptides (CPPs) offer an alternative to chitosan worthy of exploration in mosquito systems. Cell penetrating peptides (CPPs) and inorganic complexes. CPPs as described by Meade et al. are “small cationic peptides of approximately 10–30 amino acids in length that…rapidly induce their own cellular internalization through various forms of endocytosis” [159]. RNAi trigger delivery using CPPs has been explored in a variety of forms in mammalian systems with great success [159,160]. In Spodoptera frugiperda Sf9 cells, CPPs can facilitate delivery of plasmids [157]. In Ae. aegypti Aag2 cells, CPPs increase the potency of IAP1 dsRNA without increasing cytotoxicity [161]). These reports expose the potential for CPPs to deliver RNAi triggers into mosquitoes though this avenue has yet to be explored in vivo.
Chemically produced liposome nanoparticles have also been explored extensively as RNAi delivery systems. In the cockroach, B. germanica, naked dsRNA is degraded in the midgut following ingestion but protected when encapsulated in liposomes (GenJet, SignaGen), which subsequently facilitated uptake and knockdown of β-tubulin with 100% mortality [72]. The same is found in Drosophila species where neonate larvae soaked in Lipofectamine 2000 (Invitrogen) reduced RNAi target transcriptinon by 50% while soaking without lipofectamine failed to induce silencing [65]. In Ae. aegypti and M. sexta naked dsRNA and E. coli expressed dsRNA failed to induce RNAi silencing, but Effectene® (Qiagen) liposomes achieved 90% knockdown of target genes [122]. Of all Effectene® liposome feeding assays 77% knockdown was achieved on average, with up to 97% knockdown across 11 genes in Ae. aegypti [75,122,123,130,131]. There is some general debate about the necessity of liposomes however, since soaking of 1st instart Ae. aegypti with or without liposomes induced mortality when using β-tubulin, Chitin synthase I, or Heat shock protein 83 RNAi triggers [12]. Regardless of necessity, liposomes are limited in their field efficacy due to extreme cost and lack of large scale production systems.
An alternative to both chitosan and liposome systems is the utilization of PRINT particles. Here numerous different reagents, such as PEG, can be molded to form complexes with defined shapes, sizes, and charges in a highly scalable production system. Uptake and bio-distribution without toxicity have been demonstrated in larval and adult An. gambiae [50,51].

3.4. Viral Expression Systems

Considering the variety of entomopathogenic viruses already safely used as biocontrol agents mosquito viral systems offer a sophisticated means of delivering and expressing RNAi triggers directly in the mosquito cell cytosol [162]. The specificity of arboviruses to their respective arthropod hosts offer a main advantage over other expression and delivery systems [162]. The utilization of viruses as molecular tools to study or modify insects is no new concept with both Sindbis and Densovirus expression systems developed over 20 years ago [163]. The first mosquito virus expression system came in 1987 when Levis et al. [164] engineered Sindbis to express a bacterial chloramphenicol acetyltransferase in avian cells replacing 1689 nucleotides with no impact on viral propagation. When using Sindbis virus, shRNA or long hairpin dsRNA sequences can be inserted to the viral genome, enabling expression of the RNAi trigger directly in the mosquito cells. One study inserted a long dsRNA hairpin targeting vitellogenin and successfully suppressed the gene in the ovum when inoculated into Ae. aegypti adults [165].
Densoviruses have also been utilized as expression and delivery vectors of RNAs in larvae. Here the Ae. aegypti densovirus (AaDNV) genome was expressed in a plasmid containing a Pol III promoter-driven expression cassettes containing short hairpin RNA (shRNA) targeting Ae. aegypti V-ATPase subunit A [81]. Expression of the plasmid in Ae. albopictus C6/36 cells produced viral stocks that when exposed to larvae resulted in silencing of V-ATPase A and significant reduction in larval longevity.

3.5. Bacterial Expression Systems

A major alternative to nanoparticle complexes are direct expression systems. The main advantage of such a system is scalability and cost compared to in silico production methods. For instance, the RNase III deficient HT115 E. coli strain can be transformed with dsRNA plasmid expression vectors and grown in bulk worry of RNA degradation. In one study, a Pet17B plasmid containing an ampicillin resistance cassette, origin of replication, and cloned inverted repeats targeting 3 genes of interest, was mass produced in HT115 E. coli [36]. Purified dsRNA inverted repeats were then exposed to 2 day old Ae. aegypti larvae in water (soaking method) resulting in 81%–97% gene suppression of 3 target genes in larval midguts. Others take this one step further and expose mosquitoes directly to the organism producing the dsRNA. There are several lines of thought for taking on such an approach. For one, biologically produced dsRNA does not need to be expressed in the mosquito and as such does not require transfection or a mosquito specific promoter. Also dsRNA is rapidly degraded in aqueous environments [22]. This concept was originally explored by feeding dsRNA expressing E. coli to C. elegans [38]. The method was translated to mosquito larvae with remarkable efficiency. In one study, live E. coli expressing Ae. aegypti sexual dependency genes (embedded in agar pellets) were fed to larvae, resulting in sterility in up to 90% of adults [33]. Direct injection of the same dsRNAs to pupae resulted in 70%–95% gene suppression and similar levels of sterility, deeming this approach highly efficacious.

3.6. The Pichia pastoris Expression System

Pichia pastoris yeast have also been successfully transformed with pPicZB plasmid containing a long hairpin RNA sequence encoding an Ae. aegypti juvenile hormone acid methyl transferase gene [39]. Ae. aegypti fed on fermented yeasts resulted in >90% knockdown and death of larvae up to 144 h post exposure.

4. Conclusions

To combat the proliferation of pesticide resistant mosquito vectors and control associated mosquito-borne diseases, Integrated Vector Management and Integrated Vector Borne Disease Management programs will require alternatives to chemical pesticides [7,8]. Herein, we reviewed some key examples from hundreds of documented effective RNAi triggers that impact mosquito physiology and pathogen fitness, and thereby constitute an expansive arsenal targets for mosquito and mosquito-borne disease control strategies. We contend that RNAi could be adapted and implemented using the framework for existing vector control tools, including larvicides, contact and residual sprays, toxic baits, and LLINs. To translate RNAi to field applicability, RNAi triggers likely will need to be combined with biotic (e.g., a virus, yeast or bacterial expression system) or abiotic (e.g., nanoparticle) systems that mediate both protection and uptake of the RNAi trigger [13,27,46,47,48,49,50,51]. An RNAi approach to mosquito control offers a number of advantages over traditional chemical pesticides, including vastly improved species-specificity with diminished environmental toxicity compared with chemical pesticides.

Acknowledgments

This work was supported by a grant from the Foundation for the National Institutes of Health through the Vector-Based Transmission of Control: Discovery Research (VCTR) program of the Grand Challenges in Global Health initiative. The authors gratefully acknowledge collaborators Barry Beaty and Jon Carlson for fruitful discussions that inspired this review.

Author Contributions

Contributions were made as follows: Conceptualization and design of sections and content: Lyric C. Bartholomay and Paul M. Airs Design or figures and tables: Paul M. Airs Writing of manuscript: Paul M. Airs and Lyric C. Bartholomay.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Livadas, G.A.; Georgopoulos, G. Development of resistance to DDT by Anopheles sacharovi in Greece. Bull. World Health Organ. 1953, 8, 497–511. [Google Scholar] [PubMed]
  2. Hammerstrom, R.J. Insect resistance to insecticides. Public Health Rep. 1958, 73, 1126–1131. [Google Scholar] [CrossRef] [PubMed]
  3. Gjullin, C.M.; Peters, R.F. Recent studies of mosquito resistance to insecticides in California. Mosquito News 1952, 12, 1–7. [Google Scholar]
  4. Brown, A.W. Insecticide resistance in mosquitoes: A pragmatic review. J. Am. Mosq. Control Assoc. 1986, 2, 123–140. [Google Scholar] [PubMed]
  5. World Health Organization. World Malaria Report; World Health Organization: Geneva, Switzerland, 2015. [Google Scholar]
  6. World Health Organization. Vector Resistance to Pesticides. Fifteenth Report of the WHO Expert Committee on Vector Biology and Control; World Health Organization Technical Report Series; World Health Organization: Geneva, Switzerland, 1992; Volume 818, pp. 1–68. [Google Scholar]
  7. Liu, N. Insecticide resistance in mosquitoes: Impact, mechanisms, and research directions. Annu. Rev. Entomol. 2015, 60, 537–559. [Google Scholar] [CrossRef] [PubMed]
  8. Zaim, M.; Guillet, P. Alternative insecticides: An urgent need. Trends Parasitol. 2002, 18, 161–163. [Google Scholar] [CrossRef]
  9. Sen, G.L.; Blau, H.M. A brief history of RNAi: The silence of the genes. FASEB J. 2006, 20, 1293–1299. [Google Scholar] [CrossRef] [PubMed]
  10. Wang, H.; Gort, T.; Boyle, D.L.; Clem, R.J. Effects of manipulating apoptosis on Sindbis virus infection of Aedes aegypti mosquitoes. J. Virol. 2012, 86, 6546–6554. [Google Scholar] [CrossRef] [PubMed]
  11. Coy, M.R.; Sanscrainte, N.D.; Chalaire, K.C.; Inberg, A.; Maayan, I.; Glick, E.; Paldi, N.; Becnel, J.J. Gene silencing in adult Aedes aegypti mosquitoes through oral delivery of double-stranded RNA. J. Appl. Entomol. 2012, 136, 741–748. [Google Scholar] [CrossRef]
  12. Singh, A.D.; Wong, S.; Ryan, C.P.; Whyard, S. Oral delivery of double-stranded RNA in larvae of the yellow fever mosquito, Aedes aegypti: Implications for pest mosquito control. J. Insect Sci. 2013, 13. [Google Scholar] [CrossRef] [PubMed]
  13. Whyard, S.; Erdelyan, C.N.; Partridge, A.L.; Singh, A.D.; Beebe, N.W.; Capina, R. Silencing the buzz: A new approach to population suppression of mosquitoes by feeding larvae double-stranded RNAs. Parasites Vectors 2015. [Google Scholar] [CrossRef] [PubMed]
  14. Thailayil, J.; Magnusson, K.; Godfray, H.C.; Crisanti, A.; Catteruccia, F. Spermless males elicit large-scale female responses to mating in the malaria mosquito Anopheles gambiae. Proc. Natl. Acad. Sci. USA 2011, 108, 13677–13681. [Google Scholar] [CrossRef] [PubMed]
  15. Erdelyan, C.N.; Mahood, T.H.; Bader, T.S.; Whyard, S. Functional validation of the carbon dioxide receptor genes in Aedes aegypti mosquitoes using RNA interference. Insect Mol. Biol. 2012, 21, 119–127. [Google Scholar] [CrossRef] [PubMed]
  16. Pelletier, J.; Guidolin, A.; Syed, Z.; Cornel, A.J.; Leal, W.S. Knockdown of a mosquito odorant-binding protein involved in the sensitive detection of oviposition attractants. J. Chem. Ecol. 2010, 36, 245–248. [Google Scholar] [CrossRef] [PubMed]
  17. Chagas, A.C.; Ramirez, J.L.; Jasinskiene, N.; James, A.A.; Ribeiro, J.M.; Marinotti, O.; Calvo, E. Collagen-binding protein, Aegyptin, regulates probing time and blood feeding success in the dengue vector mosquito, Aedes aegypti. Proc. Natl. Acad. Sci. USA 2014, 111, 6946–6951. [Google Scholar] [CrossRef] [PubMed]
  18. Attardo, G.M.; Hansen, I.A.; Shiao, S.H.; Raikhel, A.S. Identification of two cationic amino acid transporters required for nutritional signaling during mosquito reproduction. J. Exp. Biol. 2006, 209 Pt 16, 3071–3078. [Google Scholar] [CrossRef] [PubMed]
  19. Tsao, I.Y.; Lin, U.S.; Christensen, B.M.; Chen, C.C. Armigeres subalbatus prophenoloxidase III: Cloning, characterization and potential role in morphogenesis. Insect Biochem. Mol. Biol. 2009, 39, 96–104. [Google Scholar] [CrossRef] [PubMed]
  20. Blumberg, D.D. Creating a ribonuclease-free environment. Methods Enzymol. 1987, 152, 20–24. [Google Scholar] [PubMed]
  21. Dubelman, S.; Fischer, J.; Zapata, F.; Huizinga, K.; Jiang, C.; Uffman, J.; Levine, S.; Carson, D. Environmental fate of double-stranded RNA in agricultural soils. PLoS ONE 2014, 9, e93155. [Google Scholar] [CrossRef] [PubMed]
  22. Fischer, J.R.; Zapata, F.; Dubelman, S.; Mueller, G.M.; Uffman, J.P.; Jiang, C.; Jensen, P.D.; Levine, S.L. Aquatic fate of a double-stranded RNA in a sediment-water system following an over-water application. Environ. Toxicol. Chem. 2016. [Google Scholar] [CrossRef] [PubMed]
  23. San Miguel, K.; Scott, J.G. The next generation of insecticides: DsRNA is stable as a foliar-applied insecticide. Pest Manag. Sci. 2016, 72, 801–809. [Google Scholar] [CrossRef] [PubMed]
  24. World Health Organization. Dengue: Guidelines for Diagnosis, Treatment, Prevention, and Control; New Edition Special Programme for Research and Training in Tropical Diseases; World Health Organization: Geneva, Switzerland, 2009; pp. 59–86. [Google Scholar]
  25. World Health Organization. Indoor Residual Spraying: An Operational Manual for IRS for Malaria Transmission, Control and Elimination, 2nd ed.; World Health Organization: Geneva, Switzerland, 2015. [Google Scholar]
  26. World Health Organization. Malaria Vector Control and Personal Protection: Report of a WHO Study Group; WHO Technical Report Series; World Health Organization: Geneva, Switzerland, 2006. [Google Scholar]
  27. Van Ekert, E.; Powell, C.A.; Shatters, R.G., Jr.; Borovsky, D. Control of larval and egg development in Aedes aegypti with RNA interference against juvenile hormone acid methyl transferase. J. Insect Physiol. 2014, 70, 143–150. [Google Scholar] [CrossRef] [PubMed]
  28. Muller, G.C.; Beier, J.C.; Traore, S.F.; Toure, M.B.; Traore, M.M.; Bah, S.; Doumbia, S.; Schlein, Y. Field experiments of Anopheles gambiae attraction to local fruits/seedpods and flowering plants in Mali to optimize strategies for malaria vector control in Africa using attractive toxic sugar bait methods. Malar. J. 2010. [Google Scholar] [CrossRef] [PubMed]
  29. Allan, S.A. Susceptibility of adult mosquitoes to insecticides in aqueous sucrose baits. J. Vector Ecol. 2011, 36, 59–67. [Google Scholar] [CrossRef] [PubMed]
  30. Qualls, W.A.; Muller, G.C.; Revay, E.E.; Allan, S.A.; Arheart, K.L.; Beier, J.C.; Smith, M.L.; Scott, J.M.; Kravchenko, V.D.; Hausmann, A.; et al. Evaluation of attractive toxic sugar bait (ATSB)-Barrier for control of vector and nuisance mosquitoes and its effect on non-target organisms in sub-tropical environments in Florida. Acta Trop. 2014, 131, 104–110. [Google Scholar] [CrossRef] [PubMed]
  31. Zhu, L.; Marshall, J.M.; Qualls, W.A.; Schlein, Y.; McManus, J.W.; Arheart, K.L.; Hlaing, W.M.; Traore, S.F.; Doumbia, S.; Muller, G.C.; et al. Modelling optimum use of attractive toxic sugar bait stations for effective malaria vector control in Africa. Malar. J. 2015. [Google Scholar] [CrossRef] [PubMed]
  32. Qualls, W.A.; Muller, G.C.; Traore, S.F.; Traore, M.M.; Arheart, K.L.; Doumbia, S.; Schlein, Y.; Kravchenko, V.D.; Xue, R.D.; Beier, J.C. Indoor use of attractive toxic sugar bait (ATSB) to effectively control malaria vectors in Mali, West Africa. Malar. J. 2015. [Google Scholar] [CrossRef] [PubMed]
  33. Stewart, Z.P.; Oxborough, R.M.; Tungu, P.K.; Kirby, M.J.; Rowland, M.W.; Irish, S.R. Indoor application of attractive toxic sugar bait (ATSB) in combination with mosquito nets for control of pyrethroid-resistant mosquitoes. PLoS ONE 2013, 8, e84168. [Google Scholar] [CrossRef] [PubMed]
  34. Muller, G.C.; Beier, J.C.; Traore, S.F.; Toure, M.B.; Traore, M.M.; Bah, S.; Doumbia, S.; Schlein, Y. Successful field trial of attractive toxic sugar bait (ATSB) plant-spraying methods against malaria vectors in the Anopheles gambiae complex in Mali, West Africa. Malar. J. 2010. [Google Scholar] [CrossRef] [PubMed]
  35. Beier, J.C.; Muller, G.C.; Gu, W.; Arheart, K.L.; Schlein, Y. Attractive toxic sugar bait (ATSB) methods decimate populations of Anopheles malaria vectors in arid environments regardless of the local availability of favoured sugar-source blossoms. Malar. J. 2012. [Google Scholar] [CrossRef] [PubMed]
  36. Revay, E.E.; Muller, G.C.; Qualls, W.A.; Kline, D.L.; Naranjo, D.P.; Arheart, K.L.; Kravchenko, V.D.; Yefremova, Z.; Hausmann, A.; Beier, J.C.; et al. Control of Aedes albopictus with attractive toxic sugar baits (ATSB) and potential impact on non-target organisms in St. Augustine, Florida. Parasitol. Res. 2014, 113, 73–79. [Google Scholar] [CrossRef] [PubMed]
  37. Schlein, Y.; Muller, G.C. Experimental control of Phlebotomus papatasi by spraying attractive toxic sugar bait (ATSB) on vegetation. Trans. R. Soc. Trop. Med. Hyg. 2010, 104, 766–771. [Google Scholar] [CrossRef] [PubMed]
  38. Ding, Y.M.; Hu, Y.; Yu, B.T.; Mo, X.C.; Mo, J.C. Laboratory evaluation of differential attraction of Culex pipiens pallens to fruit-based sugar baits. Acta Trop. 2016, 163, 20–25. [Google Scholar] [CrossRef] [PubMed]
  39. Junnila, A.; Revay, E.E.; Muller, G.C.; Kravchenko, V.; Qualls, W.A.; Xue, R.D.; Allen, S.A.; Beier, J.C.; Schlein, Y. Efficacy of attractive toxic sugar baits (ATSB) against Aedes albopictus with garlic oil encapsulated in beta-cyclodextrin as the active ingredient. Acta Trop. 2015, 152, 195–200. [Google Scholar] [CrossRef] [PubMed]
  40. Muller, G.C.; Junnila, A.; Qualls, W.; Revay, E.E.; Kline, D.L.; Allan, S.; Schlein, Y.; Xue, R.D. Control of Culex quinquefasciatus in a storm drain system in Florida using attractive toxic sugar baits. Med. Vet. Entomol. 2010, 24, 346–351. [Google Scholar] [CrossRef] [PubMed]
  41. Muller, G.C.; Junnila, A.; Schlein, Y. Effective control of adult Culex pipiens by spraying an attractive toxic sugar bait solution in the vegetation near larval habitats. J. Med. Entomol. 2010, 47, 63–66. [Google Scholar] [CrossRef] [PubMed]
  42. Pietri, J.E.; Cheung, K.W.; Luckhart, S. Knockdown of mitogen-activated protein kinase (MAPK) signalling in the midgut of Anopheles stephensi mosquitoes using antisense morpholinos. Insect Mol. Biol. 2014, 23, 558–565. [Google Scholar] [CrossRef] [PubMed]
  43. Baak-Baak, C.M.; Rodriguez-Ramirez, A.D.; Garcia-Rejon, J.E.; Rios-Delgado, S.; Torres-Estrada, J.L. Development and laboratory evaluation of chemically-based baited ovitrap for the monitoring of Aedes aegypti. J. Vector Ecol. 2013, 38, 175–181. [Google Scholar] [CrossRef] [PubMed]
  44. Lampman, R.L.; Novak, R.J. Oviposition preferences of Culex pipiens and Culex restuans for infusion-baited traps. J. Am. Mosq. Control Assoc. 1996, 12, 23–32. [Google Scholar] [PubMed]
  45. Harris, C.; Kihonda, J.; Lwetoijera, D.; Dongus, S.; Devine, G.; Majambere, S. A simple and efficient tool for trapping gravid Anopheles at breeding sites. Parasites Vectors 2011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Das, S.; Debnath, N.; Cui, Y.; Unrine, J.; Palli, S.R. Chitosan, Carbon Quantum Dot, and Silica Nanoparticle Mediated dsRNA Delivery for Gene Silencing in Aedes aegypti: A Comparative Analysis. ACS Appl. Mater. Interfaces 2015, 7, 19530–19535. [Google Scholar] [CrossRef] [PubMed]
  47. Mysore, K.; Flannery, E.M.; Tomchaney, M.; Severson, D.W.; Duman-Scheel, M. Disruption of Aedes aegypti olfactory system development through chitosan/siRNA nanoparticle targeting of semaphorin-1A. PLoS Negl. Trop. Dis. 2013, 7, e2215. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, Q.; Hua, G.; Adang, M.J. Chitosan/DsiRNA nanoparticle targeting identifies AgCad1 cadherin in Anopheles gambiae larvae as an in vivo receptor of Cry11Ba toxin of Bacillus thuringiensis subsjegathesan. Insect Biochem. Mol. Biol. 2015, 60, 33–38. [Google Scholar] [CrossRef] [PubMed]
  49. Zhang, X.; Zhang, J.; Zhu, K.Y. Chitosan/double-stranded RNA nanoparticle-mediated RNA interference to silence chitin synthase genes through larval feeding in the African malaria mosquito (Anopheles gambiae). Insect Mol. Biol. 2010, 19, 683–693. [Google Scholar] [CrossRef] [PubMed]
  50. Paquette, C.C.; Phanse, Y.; Perry, J.L.; Sanchez-Vargas, I.; Airs, P.M.; Dunphy, B.M.; Xu, J.; Carlson, J.O.; Luft, J.C.; DeSimone, J.M.; et al. Biodistribution and trafficking of hydrogel nanoparticles in adult mosquitoes. PLoS Negl. Trop. Dis. 2015, 9, e0003745. [Google Scholar] [CrossRef] [PubMed]
  51. Phanse, Y.; Dunphy, B.M.; Perry, J.L.; Airs, P.M.; Paquette, C.C.; Carlson, J.O.; Xu, J.; Luft, J.C.; DeSimone, J.M.; Beaty, B.J.; et al. Biodistribution and Toxicity Studies of PRINT Hydrogel Nanoparticles in Mosquito Larvae and Cells. PLoS Negl. Trop. Dis. 2015, 9, e0003735. [Google Scholar] [CrossRef] [PubMed]
  52. Liu, Q.; Clem, R.J. Defining the core apoptosis pathway in the mosquito disease vector Aedes aegypti: The roles of iap1, ark, dronc, and effector caspases. Apoptosis 2011, 16, 105–113. [Google Scholar] [CrossRef] [PubMed]
  53. Pridgeon, J.W.; Zhao, L.; Becnel, J.J.; Strickman, D.A.; Clark, G.G.; Linthicum, K.J. Topically applied AaeIAP1 double-stranded RNA kills female adults of Aedes aegypti. J. Med. Entomol. 2008, 45, 414–420. [Google Scholar] [CrossRef] [PubMed]
  54. Puglise, J.M.; Estep, A.S.; Becnel, J.J. Expression Profiles and RNAi Silencing of Inhibitor of Apoptosis Transcripts in Aedes, Anopheles, and Culex Mosquitoes (Diptera: Culicidae). J. Med. Entomol. 2016, 53, 304–314. [Google Scholar] [CrossRef] [PubMed]
  55. Ocampo, C.B.; Caicedo, P.A.; Jaramillo, G.; Bedoya, R.U.; Baron, O.; Serrato, I.M.; Cooper, D.M.; Lowenberger, C. Differential expression of apoptosis related genes in selected strains of Aedes aegypti with different susceptibilities to dengue virus. PLoS ONE 2013, 8, e61187. [Google Scholar] [CrossRef] [PubMed]
  56. Vogel, K.J.; Brown, M.R.; Strand, M.R. Ovary ecdysteroidogenic hormone requires a receptor tyrosine kinase to activate egg formation in the mosquito Aedes aegypti. Proc. Natl. Acad. Sci. USA 2015, 112, 5057–5062. [Google Scholar] [CrossRef] [PubMed]
  57. Clemons, A.; Haugen, M.; Le, C.; Mori, A.; Tomchaney, M.; Severson, D.W.; Duman-Scheel, M. siRNA-mediated gene targeting in Aedes aegypti embryos reveals that frazzled regulates vector mosquito CNS development. PLoS ONE 2011, 6, e16730. [Google Scholar] [CrossRef] [PubMed]
  58. Zou, F.F.; Guo, Q.; Sun, Y.; Zhou, D.; Hu, M.X.; Hu, H.X.; Liu, B.Q.; Tian, M.M.; Liu, X.M.; Li, X.X.; et al. Identification of protease m1 zinc metalloprotease conferring resistance to deltamethrin by characterization of an AFLP marker in Culex pipiens pallens. Parasites Vectors 2016. [Google Scholar] [CrossRef] [PubMed]
  59. Clayton, A.M.; Cirimotich, C.M.; Dong, Y.; Dimopoulos, G. Caudal is a negative regulator of the Anopheles IMD pathway that controls resistance to Plasmodium falciparum infection. Dev. Comp. Immunol. 2013, 39, 323–332. [Google Scholar] [CrossRef] [PubMed]
  60. Papaefthimiou, C.; Theophilidis, G. The cardiotoxic action of the pyrethroid insecticide deltamethrin, the azole fungicide prochloraz, and their synergy on the semi-isolated heart of the bee Apis mellifera macedonica. Pest. Biochem. Physiol. 2001, 69, 77–91. [Google Scholar] [CrossRef]
  61. Desneux, N.; Decourtye, A.; Delpuech, J.M. The sublethal effects of pesticides on beneficial arthropods. Annu. Rev. Entomol. 2007, 52, 81–106. [Google Scholar] [CrossRef] [PubMed]
  62. Antwi, F.B.; Reddy, G.V. Toxicological effects of pyrethroids on non-target aquatic insects. Environ. Toxicol. Pharmacol. 2015, 40, 915–923. [Google Scholar] [CrossRef] [PubMed]
  63. Costa, L.G. Current issues in organophosphate toxicology. Clin. Chim. Acta 2006, 366, 1–13. [Google Scholar] [CrossRef] [PubMed]
  64. Bagchi, M.; Zafra, S.; Bagchi, D. DNA Damage, Gene Expression, and Carcinogenesis by Organophosphates and Carbamates. In Toxicology of Organophosphate & Carbamate Compounds; Gupta, R.C., Ed.; Academic Press: Burlington, VT, USA, 2006; pp. 533–548. [Google Scholar]
  65. Whyard, S.; Singh, A.D.; Wong, S. Ingested double-stranded RNAs can act as species-specific insecticides. Insect Biochem. Mol. Biol. 2009, 39, 824–832. [Google Scholar] [CrossRef] [PubMed]
  66. Kaiser, W.J.; Vucic, D.; Miller, L.K. The Drosophila inhibitor of apoptosis D-IAP1 suppresses cell death induced by the caspase drICE. FEBS Lett. 1998, 440, 243–248. [Google Scholar] [CrossRef]
  67. Taylor, R.C.; Cullen, S.P.; Martin, S.J. Apoptosis: Controlled demolition at the cellular level. Nat. Rev. Mol. Cell Biol. 2008, 9, 231–241. [Google Scholar] [CrossRef] [PubMed]
  68. Luthi, A.U.; Martin, S.J. The CASBAH: A searchable database of caspase substrates. Cell Death Differ. 2007, 14, 641–650. [Google Scholar] [CrossRef] [PubMed]
  69. Lombardo, F.; Ghani, Y.; Kafatos, F.C.; Christophides, G.K. Comprehensive genetic dissection of the hemocyte immune response in the malaria mosquito Anopheles gambiae. PLoS Pathog. 2013, 9, e1003145. [Google Scholar] [CrossRef] [PubMed]
  70. Boutros, M.; Kiger, A.A.; Armknecht, S.; Kerr, K.; Hild, M.; Koch, B.; Haas, S.A.; Paro, R.; Perrimon, N.; Array, C.H.F. Genome-wide RNAi analysis of growth and viability in Drosophila cells. Science 2004, 303, 832–835. [Google Scholar] [CrossRef] [PubMed]
  71. Zdobnov, E.M.; von Mering, C.; Letunic, I.; Torrents, D.; Suyama, M.; Copley, R.R.; Christophides, G.K.; Thomasova, D.; Holt, R.A.; Subramanian, G.M.; et al. Comparative genome and proteome analysis of Anopheles gambiae and Drosophila melanogaster. Science 2002, 298, 149–159. [Google Scholar] [CrossRef] [PubMed]
  72. Lin, Y.H.; Huang, J.H.; Liu, Y.; Belles, X.; Lee, H.J. Oral delivery of dsRNA lipoplexes to German cockroach protects dsRNA from degradation and induces RNAi response. Pest Manag. Sci. 2016. [Google Scholar] [CrossRef] [PubMed]
  73. Baum, J.A.; Bogaert, T.; Clinton, W.; Heck, G.R.; Feldmann, P.; Ilagan, O.; Johnson, S.; Plaetinck, G.; Munyikwa, T.; Pleau, M.; et al. Control of coleopteran insect pests through RNA interference. Nat. Biotechnol. 2007, 25, 1322–1326. [Google Scholar] [CrossRef] [PubMed]
  74. Whitten, M.M.; PFacey, D.; del Sol, R.; Fernandez-Martinez, L.T.; Evans, M.C.; Mitchell, J.J.; Bodger, O.G.; Dyson, P.J. Symbiont-mediated RNA interference in insects. Proc. Biol. Sci. 2016. [Google Scholar] [CrossRef] [PubMed]
  75. Cancino-Rodezno, A.; Lozano, L.; Oppert, C.; Castro, J.I.; Lanz-Mendoza, H.; Encarnacion, S.; Evans, A.E.; Gill, S.S.; Soberon, M.; Jurat-Fuentes, J.L.; et al. Comparative proteomic analysis of Aedes aegypti larval midgut after intoxication with Cry11Aa toxin from Bacillus thuringiensis. PLoS ONE 2012, 7, e37034. [Google Scholar] [CrossRef] [PubMed]
  76. Nelson, N.; Perzov, N.; Cohen, A.; Hagai, K.; Padler, V.; Nelson, H. The cellular biology of proton-motive force generation by V-ATPases. J. Exp. Biol. 2000, 203, 89–95. [Google Scholar] [PubMed]
  77. Allan, A.K.; Du, J.; Davies, S.A.; Dow, J.A. Genome-wide survey of V-ATPase genes in Drosophila reveals a conserved renal phenotype for lethal alleles. Physiol. Genom. 2005, 22, 128–138. [Google Scholar] [CrossRef] [PubMed]
  78. Wuriyanghan, H.; Rosa, C.; Falk, B.W. Oral delivery of double-stranded RNAs and siRNAs induces RNAi effects in the potato/tomato psyllid, Bactericerca cockerelli. PLoS ONE 2011, 6, e27736. [Google Scholar] [CrossRef] [PubMed]
  79. Mao, J.; Zhang, P.; Liu, C.; Zeng, F. Co-silence of the coatomer beta and V-ATPase A genes by siRNA feeding reduces larval survival rate and weight gain of cotton bollworm, Helicoverpa armigera. Pestic. Biochem. Physiol. 2015, 118, 71–76. [Google Scholar] [CrossRef] [PubMed]
  80. Kang, S.; Shields, A.R.; Jupatanakul, N.; Dimopoulos, G. Suppressing dengue-2 infection by chemical inhibition of Aedes aegypti host factors. PLoS Negl. Trop. Dis. 2014, 8, e3084. [Google Scholar] [CrossRef] [PubMed]
  81. Gu, J.; Liu, M.; Deng, Y.; Peng, H.; Chen, X. Development of an efficient recombinant mosquito densovirus-mediated RNA interference system and its preliminary application in mosquito control. PLoS ONE 2011, 6, e21329. [Google Scholar] [CrossRef] [PubMed]
  82. Isoe, J.; Collins, J.; Badgandi, H.; Day, W.A.; Miesfeld, R.L. Defects in coatomer protein I (COPI) transport cause blood feeding-induced mortality in Yellow Fever mosquitoes. Proc. Natl. Acad. Sci. USA 2011, 108, E211–E217. [Google Scholar] [CrossRef] [PubMed]
  83. Isoe, J.; Stover, W.; Miesfeld, R.B.; Miesfeld, R.L. COPI-mediated blood meal digestion in vector mosquitoes is independent of midgut ARF-GEF and ARF-GAP regulatory activities. Insect Biochem. Mol. Biol. 2013, 43, 732–739. [Google Scholar] [CrossRef] [PubMed]
  84. Zhou, G.; Isoe, J.; Day, W.A.; Miesfeld, R.L. Alpha-COPI coatomer protein is required for rough endoplasmic reticulum whorl formation in mosquito midgut epithelial cells. PLoS ONE 2011, 6, e18150. [Google Scholar] [CrossRef] [PubMed]
  85. Tsao, I.Y.; Christensen, B.M.; Chen, C.C. Armigeres subalbatus (Diptera: Culicidae) prophenoloxidase III is required for mosquito cuticle formation: Ultrastructural study on dsRNA-knockdown mosquitoes. J. Med. Entomol. 2010, 47, 495–503. [Google Scholar] [CrossRef] [PubMed]
  86. Scott, J.G.; Michel, K.; Bartholomay, L.C.; Siegfried, B.D.; Hunter, W.B.; Smagghe, G.; Zhu, K.Y.; Douglas, A.E. Towards the elements of successful insect RNAi. J. Insect Physiol. 2013, 59, 1212–1221. [Google Scholar] [CrossRef] [PubMed]
  87. DeGennaro, M.; McBride, C.S.; Seeholzer, L.; Nakagawa, T.; Dennis, E.J.; Goldman, C.; Jasinskiene, N.; James, A.A.; Vosshall, L.B. Orco mutant mosquitoes lose strong preference for humans and are not repelled by volatile DEET. Nature 2013, 498, 487–491. [Google Scholar] [CrossRef] [PubMed]
  88. Xu, P.; Choo, Y.M.; de la Rosa, A.; Leal, W.S. Mosquito odorant receptor for DEET and methyl jasmonate. Proc. Natl. Acad. Sci. USA 2014, 111, 16592–16597. [Google Scholar] [CrossRef] [PubMed]
  89. Pelletier, J.; Leal, W.S. Characterization of olfactory genes in the antennae of the Southern house mosquito, Culex quinquefasciatus. J. Insect Physiol. 2011, 57, 915–929. [Google Scholar] [CrossRef] [PubMed]
  90. Xu, P.X.; Zwiebel, L.J.; Smith, D. Identification of a distinct family of genes encoding atypical odorant-binding proteins in the malaria vector mosquito, Anopheles gambiae. Insect Mol. Biol. 2003, 12, 549–560. [Google Scholar] [CrossRef] [PubMed]
  91. Zhou, J.J.; He, X.L.; Pickett, J.A.; Field, L.M. Identification of odorant-binding proteins of the yellow fever mosquito Aedes aegypti: Genome annotation and comparative analyses. Insect Mol. Biol. 2008, 17, 147–163. [Google Scholar] [CrossRef] [PubMed]
  92. Rinker, D.C.; Pitts, R.J.; Zhou, X.; Suh, E.; Rokas, A.; Zwiebel, L.J. Blood meal-induced changes to antennal transcriptome profiles reveal shifts in odor sensitivities in Anopheles gambiae. Proc. Natl. Acad. Sci. USA 2013, 110, 8260–8265. [Google Scholar] [CrossRef] [PubMed]
  93. Rinker, D.C.; Zhou, X.; Pitts, R.J.; Consortium, A.G.C.; Rokas, A.; Zwiebel, L.J. Antennal transcriptome profiles of anopheline mosquitoes reveal human host olfactory specialization in Anopheles gambiae. BMC Genom. 2013. [Google Scholar] [CrossRef] [PubMed]
  94. Won Jung, J.; Baeck, S.-J.; Perumalsamy, H.; Hansson, B.S.; Ahn, Y.-J.; Kwon, H.W. A novel olfactory pathway is essential for fast and efficient blood-feeding in mosquitoes. Sci. Rep. 2015. [Google Scholar] [CrossRef] [PubMed]
  95. Liu, H.; Liu, T.; Xie, L.; Wang, X.; Deng, Y.; Chen, C.H.; James, A.A.; Chen, X.G. Functional analysis of Orco and odorant receptors in odor recognition in Aedes albopictus. Parasites Vectors 2016. [Google Scholar] [CrossRef] [PubMed]
  96. Zhu, F.; Xu, P.; Barbosa, R.M.; Choo, Y.M.; Leal, W.S. RNAi-based demonstration of direct link between specific odorant receptors and mosquito oviposition behavior. Insect Biochem. Mol. Biol. 2013, 43, 916–923. [Google Scholar] [CrossRef] [PubMed]
  97. Das, S.; Dimopoulos, G. Molecular analysis of photic inhibition of blood-feeding in Anopheles gambiae. BMC Physiol. 2008. [Google Scholar] [CrossRef] [PubMed]
  98. Marinotti, O.; Nguyen, Q.K.; Calvo, E.; James, A.A.; Ribeiro, J.M. Microarray analysis of genes showing variable expression following a blood meal in Anopheles gambiae. Insect Mol. Biol. 2005, 14, 365–373. [Google Scholar] [CrossRef] [PubMed]
  99. Marinotti, O.; Calvo, E.; Nguyen, Q.K.; Dissanayake, S.; Ribeiro, J.M.; James, A.A. Genome-wide analysis of gene expression in adult Anopheles gambiae. Insect Mol. Biol. 2006, 15, 1–12. [Google Scholar] [CrossRef] [PubMed]
  100. Isoe, J.; Rascon, A.A., Jr.; Kunz, S.; Miesfeld, R.L. Molecular genetic analysis of midgut serine proteases in Aedes aegypti mosquitoes. Insect Biochem. Mol. Biol. 2009, 39, 903–912. [Google Scholar] [CrossRef] [PubMed]
  101. Isoe, J.; Scaraffia, P.Y. Urea synthesis and excretion in Aedes aegypti mosquitoes are regulated by a unique cross-talk mechanism. PLoS ONE 2013, 8, e65393. [Google Scholar] [CrossRef] [PubMed]
  102. Overend, G.; Cabrero, P.; Halberg, K.A.; Ranford-Cartwright, L.C.; Woods, D.J.; Davies, S.A.; Dow, J.A. A comprehensive transcriptomic view of renal function in the malaria vector, Anopheles gambiae. Insect Biochem. Mol. Biol. 2015, 67, 47–58. [Google Scholar] [CrossRef] [PubMed]
  103. Drake, L.L.; Boudko, D.Y.; Marinotti, O.; Carpenter, V.K.; Dawe, A.L.; Hansen, I.A. The Aquaporin gene family of the yellow fever mosquito, Aedes aegypti. PLoS ONE 2010, 5, e15578. [Google Scholar] [CrossRef] [PubMed]
  104. Liu, K.; Tsujimoto, H.; Cha, S.J.; Agre, P.; Rasgon, J.L. Aquaporin water channel AgAQP1 in the malaria vector mosquito Anopheles gambiae during blood feeding and humidity adaptation. Proc. Natl. Acad. Sci. USA 2011, 108, 6062–6066. [Google Scholar] [CrossRef] [PubMed]
  105. Hansen, I.A.; Attardo, G.M.; Roy, S.G.; Raikhel, A.S. Target of rapamycin-dependent activation of S6 kinase is a central step in the transduction of nutritional signals during egg development in a mosquito. J. Biol. Chem. 2005, 280, 20565–20572. [Google Scholar] [CrossRef] [PubMed]
  106. Hansen, I.A.; Attardo, G.M.; Park, J.H.; Peng, Q.; Raikhel, A.S. Target of rapamycin-mediated amino acid signaling in mosquito anautogeny. Proc. Natl. Acad. Sci. USA 2004, 101, 10626–10631. [Google Scholar] [CrossRef] [PubMed]
  107. Raphemot, R.; Estevez-Lao, T.Y.; Rouhier, M.F.; Piermarini, P.M.; Denton, J.S.; Hillyer, J.F. Molecular and functional characterization of Anopheles gambiae inward rectifier potassium (Kir1) channels: A novel role in egg production. Insect Biochem. Mol. Biol. 2014, 51, 10–19. [Google Scholar] [CrossRef] [PubMed]
  108. Telang, A.; Rechel, J.A.; Brandt, J.R.; Donnell, D.M. Analysis of ovary-specific genes in relation to egg maturation and female nutritional condition in the mosquitoes Georgecraigius atropalpus and Aedes aegypti (Diptera: Culicidae). J. Insect Physiol. 2013, 59, 283–294. [Google Scholar] [CrossRef] [PubMed]
  109. Moretti, D.M.; Ahuja, L.G.; Nunes, R.D.; Cudischevitch, C.O.; Daumas-Filho, C.R.; Medeiros-Castro, P.; Ventura-Martins, G.; Jablonka, W.; Gazos-Lopes, F.; Senna, R.; et al. Molecular analysis of Aedes aegypti classical protein tyrosine phosphatases uncovers an ortholog of mammalian PTP-1B implicated in the control of egg production in mosquitoes. PLoS ONE 2014, 9, e104878. [Google Scholar] [CrossRef] [PubMed]
  110. Hansen, I.A.; Boudko, D.Y.; Shiao, S.H.; Voronov, D.A.; Meleshkevitch, E.A.; Drake, L.L.; Aguirre, S.E.; Fox, J.M.; Attardo, G.M.; Raikhel, A.S. AaCAT1 of the yellow fever mosquito, Aedes aegypti: A novel histidine-specific amino acid transporter from the SLC7 family. J. Biol. Chem. 2011, 286, 10803–10813. [Google Scholar] [CrossRef] [PubMed]
  111. Sim, C.; Denlinger, D.L. A shut-down in expression of an insulin-like peptide, ILP-1, halts ovarian maturation during the overwintering diapause of the mosquito Culex pipiens. Insect Mol. Biol. 2009, 18, 325–332. [Google Scholar] [CrossRef] [PubMed]
  112. Rogers, D.W.; Baldini, F.; Battaglia, F.; Panico, M.; Dell, A.; Morris, H.R.; Catteruccia, F. Transglutaminase-mediated semen coagulation controls sperm storage in the malaria mosquito. PLoS Biol. 2009, 7, e1000272. [Google Scholar] [CrossRef] [PubMed]
  113. Hoang, K.P.; Teo, T.M.; Ho, T.X.; Le, V.S. Mechanisms of sex determination and transmission ratio distortion in Aedes aegypti. Parasites Vectors 2016. [Google Scholar] [CrossRef] [PubMed]
  114. Akbari, O.S.; Antoshechkin, I.; Amrhein, H.; Williams, B.; Diloreto, R.; Sandler, J.; Hay, B.A. The developmental transcriptome of the mosquito Aedes aegypti, an invasive species and major arbovirus vector. G3 2013, 3, 1493–1509. [Google Scholar] [CrossRef] [PubMed]
  115. Koutsos, A.C.; Blass, C.; Meister, S.; Schmidt, S.; MacCallum, R.M.; Soares, M.B.; Collins, F.H.; Benes, V.; Zdobnov, E.; Kafatos, F.C.; et al. Life cycle transcriptome of the malaria mosquito Anopheles gambiae and comparison with the fruitfly Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 2007, 104, 11304–11309. [Google Scholar] [CrossRef] [PubMed]
  116. Sarro, J.; Andrews, E.; Sun, L.; Behura, S.K.; Tan, J.C.; Zeng, E.; Severson, D.W.; Duman-Scheel, M. Requirement for commissureless2 function during dipteran insect nerve cord development. Dev. Dyn. 2013, 242, 1466–1477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Haugen, M.; Flannery, E.; Tomchaney, M.; Mori, A.; Behura, S.K.; Severson, D.W.; Duman-Scheel, M. Semaphorin-1A is required for Aedes aegypti embryonic nerve cord development. PLoS ONE 2011, 6, e21694. [Google Scholar] [CrossRef] [PubMed]
  118. Meleshkevitch, E.A.; Voronov, D.A.; Miller, M.M.; Penneda, M.; Fox, J.M.; Metzler, R.; Boudko, D.Y. A novel eukaryotic Na+ methionine selective symporter is essential for mosquito development. Insect Biochem. Mol. Biol. 2013, 43, 755–767. [Google Scholar] [CrossRef] [PubMed]
  119. Peng, R.; Fu, Q.; Hong, H.; Schwaegler, T.; Lan, Q. THAP and ATF-2 regulated sterol carrier protein-2 promoter activities in the larval midgut of the yellow fever mosquito, Aedes aegypti. PLoS ONE 2012, 7, e46948. [Google Scholar] [CrossRef] [PubMed]
  120. Peng, R.; Maklokova, V.I.; Chandrashekhar, J.H.; Lan, Q. In vivo functional genomic studies of sterol carrier protein-2 gene in the yellow fever mosquito. PLoS ONE 2011, 6, e18030. [Google Scholar] [CrossRef] [PubMed]
  121. Bucher, G.; Scholten, J.; Klingler, M. Parental RNAi in Tribolium (Coleoptera). Curr. Biol. 2002, 12, R85–R86. [Google Scholar] [CrossRef]
  122. Cancino-Rodezno, A.; Alexander, C.; Villasenor, R.; Pacheco, S.; Porta, H.; Pauchet, Y.; Soberon, M.; Gill, S.S.; Bravo, A. The mitogen-activated protein kinase p38 is involved in insect defense against Cry toxins from Bacillus thuringiensis. Insect Biochem. Mol. Biol. 2010, 40, 58–63. [Google Scholar] [CrossRef] [PubMed]
  123. Bedoya-Perez, L.P.; Cancino-Rodezno, A.; Flores-Escobar, B.; Soberon, M.; Bravo, A. Role of UPR pathway in defense response of Aedes aegypti against Cry11Aa toxin from Bacillus thuringiensis. Int. J. Mol. Sci. 2013, 14, 8467–8478. [Google Scholar] [CrossRef] [PubMed]
  124. Lopez-Martinez, G.; Meuti, M.; Denlinger, D.L. Rehydration driven RNAi: A novel approach for effectively delivering dsRNA to mosquito larvae. J. Med. Entomol. 2012, 49, 215–218. [Google Scholar] [CrossRef] [PubMed]
  125. Liu, C.; Pitts, R.J.; Bohbot, J.D.; Jones, P.L.; Wang, G.; Zwiebel, L.J. Distinct olfactory signaling mechanisms in the malaria vector mosquito Anopheles gambiae. PLoS Biol. 2010, 8, e1000467. [Google Scholar] [CrossRef] [PubMed]
  126. Saengwiman, S.; Aroonkesorn, A.; Dedvisitsakul, P.; Sakdee, S.; Leetachewa, S.; Angsuthanasombat, C.; Pootanakit, K. In vivo identification of Bacillus thuringiensis Cry4Ba toxin receptors by RNA interference knockdown of glycosylphosphatidylinositol-linked aminopeptidase N transcripts in Aedes aegypti larvae. Biochem. Biophys. Res. Commun. 2011, 407, 708–713. [Google Scholar] [CrossRef] [PubMed]
  127. Figueira-Mansur, J.; Ferreira-Pereira, A.; Mansur, J.F.; Franco, T.A.; Alvarenga, E.S.; Sorgine, M.H.; Neves, B.C.; Melo, A.C.; Leal, W.S.; Masuda, H.; et al. Silencing of P-glycoprotein increases mortality in temephos-treated Aedes aegypti larvae. Insect Mol. Biol. 2013, 22, 648–658. [Google Scholar] [CrossRef] [PubMed]
  128. Mysore, K.; Andrews, E.; Li, P.; Duman-Scheel, M. Chitosan/siRNA nanoparticle targeting demonstrates a requirement for single-minded during larval and pupal olfactory system development of the vector mosquito Aedes aegypti. BMC Dev. Biol. 2014. [Google Scholar] [CrossRef] [PubMed]
  129. Zhang, X.; Mysore, K.; Flannery, E.; Michel, K.; Severson, D.W.; Zhu, K.Y.; Duman-Scheel, M. Chitosan/interfering RNA nanoparticle mediated gene silencing in disease vector mosquito larvae. J. Vis. Exp. 2015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Rodriguez-Almazan, C.; Reyes, E.Z.; Zuniga-Navarrete, F.; Munoz-Garay, C.; Gomez, I.; Evans, A.M.; Likitvivatanavong, S.; Bravo, A.; Gill, S.S.; Soberon, M. Cadherin binding is not a limiting step for Bacillus thuringiensis subsisraelensis Cry4Ba toxicity to Aedes aegypti larvae. Biochem. J. 2012, 443, 711–717. [Google Scholar] [CrossRef] [PubMed]
  131. Jimenez, A.I.; Reyes, E.Z.; Cancino-Rodezno, A.; Bedoya-Perez, L.P.; Caballero-Flores, G.G.; Muriel-Millan, L.F.; Likitvivatanavong, S.; Gill, S.S.; Bravo, A.; Soberon, M. Aedes aegypti alkaline phosphatase ALP1 is a functional receptor of Bacillus thuringiensis Cry4Ba and Cry11Aa toxins. Insect Biochem. Mol. Biol. 2012, 42, 683–689. [Google Scholar] [CrossRef] [PubMed]
  132. Jeon, S.J.; Oh, M.; Yeo, W.S.; Galvao, K.N.; Jeong, K.C. Underlying mechanism of antimicrobial activity of chitosan microparticles and implications for the treatment of infectious diseases. PLoS ONE 2014, 9, e92723. [Google Scholar] [CrossRef] [PubMed]
  133. Mao, Y.B.; Cai, W.J.; Wang, J.W.; Hong, G.J.; Tao, X.Y.; Wang, L.J.; Huang, Y.P.; Chen, X.Y. Silencing a cotton bollworm P450 monooxygenase gene by plant-mediated RNAi impairs larval tolerance of gossypol. Nat. Biotechnol. 2007, 25, 1307–1313. [Google Scholar] [CrossRef] [PubMed]
  134. Adams, P. Preserving pyrethroids. Bull. World Health Organ. 2014, 92, 158–159. [Google Scholar] [PubMed]
  135. Powers, A.M.; Olson, K.E.; Higgs, S.; Carlson, J.O.; Beaty, B.J. Intracellular immunization of mosquito cells to LaCrosse virus using a recombinant Sindbis virus vector. Virus Res. 1994, 32, 57–67. [Google Scholar] [CrossRef]
  136. Gaines, P.J.; Olson, K.E.; Higgs, S.; Powers, A.M.; Beaty, B.J.; Blair, C.D. Pathogen-derived resistance to dengue type 2 virus in mosquito cells by expression of the premembrane coding region of the viral genome. J. Virol. 1996, 70, 2132–2137. [Google Scholar] [PubMed]
  137. Olson, K.E.; Higgs, S.; Gaines, P.J.; Powers, A.M.; Davis, B.S.; Kamrud, K.I.; Carlson, J.O.; Blair, C.D.; Beaty, B.J. Genetically engineered resistance to dengue-2 virus transmission in mosquitoes. Science 1996, 272, 884–886. [Google Scholar] [CrossRef] [PubMed]
  138. Higgs, S.; Rayner, J.O.; Olson, K.E.; Davis, B.S.; Beaty, B.J.; Blair, C.D. Engineered resistance in Aedes aegypti to a West African and a South American strain of yellow fever virus. Am. J. Trop. Med. Hyg. 1998, 58, 663–670. [Google Scholar] [PubMed]
  139. Franz, A.W.; Sanchez-Vargas, I.; Raban, R.R.; Black, W.C.t.; James, A.A.; Olson, K.E. Fitness impact and stability of a transgene conferring resistance to dengue-2 virus following introgression into a genetically diverse Aedes aegypti strain. PLoS Negl. Trop. Dis. 2014, 8, e2833. [Google Scholar] [CrossRef] [PubMed]
  140. Franz, A.W.; Sanchez-Vargas, I.; Adelman, Z.N.; Blair, C.D.; Beaty, B.J.; James, A.A.; Olson, K.E. Engineering RNA interference-based resistance to dengue virus type 2 in genetically modified Aedes aegypti. Proc. Natl. Acad. Sci. USA 2006, 103, 4198–4203. [Google Scholar] [CrossRef] [PubMed]
  141. Goic, B.; Stapleford, K.A.; Frangeul, L.; Doucet, A.J.; Gausson, V.; Blanc, H.; Schemmel-Jofre, N.; Cristofari, G.; Lambrechts, L.; Vignuzzi, M.; et al. Virus-derived DNA drives mosquito vector tolerance to arboviral infection. Nat. Commun. 2016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Myles, K.M.; Wiley, M.R.; Morazzani, E.M.; Adelman, Z.N. Alphavirus-derived small RNAs modulate pathogenesis in disease vector mosquitoes. Proc. Natl. Acad. Sci. USA 2008, 105, 19938–19943. [Google Scholar] [CrossRef] [PubMed]
  143. Cirimotich, C.M.; Scott, J.C.; Phillips, A.T.; Geiss, B.J.; Olson, K.E. Suppression of RNA interference increases alphavirus replication and virus-associated mortality in Aedes aegypti mosquitoes. BMC Microbiol. 2009. [Google Scholar] [CrossRef] [PubMed]
  144. Blandin, S.A.; Marois, E.; Levashina, E.A. Antimalarial responses in Anopheles gambiae: From a complement-like protein to a complement-like pathway. Cell Host Microbe 2008, 3, 364–374. [Google Scholar] [CrossRef] [PubMed]
  145. Garver, L.S.; Bahia, A.C.; Das, S.; Souza-Neto, J.A.; Shiao, J.; Dong, Y.; Dimopoulos, G. Anopheles Imd pathway factors and effectors in infection intensity-dependent anti-Plasmodium action. PLoS Pathog. 2012, 8, e1002737. [Google Scholar] [CrossRef] [PubMed]
  146. Garver, L.S.; Dong, Y.; Dimopoulos, G. Caspar controls resistance to Plasmodium falciparum in diverse anopheline species. PLoS Pathog. 2009, 5, e1000335. [Google Scholar] [CrossRef] [PubMed]
  147. Geldhof, P.; Visser, A.; Clark, D.; Saunders, G.; Britton, C.; Gilleard, J.; Berriman, M.; Knox, D. RNA interference in parasitic helminths: Current situation, potential pitfalls and future prospects. Parasitology 2007, 134, 609–619. [Google Scholar] [CrossRef] [PubMed]
  148. Song, C.; Gallup, J.M.; Day, T.A.; Bartholomay, L.C.; Kimber, M.J. Development of an in vivo RNAi protocol to investigate gene function in the filarial nematode, Brugia malayi. PLoS Pathog. 2010, 6, e1001239. [Google Scholar] [CrossRef] [PubMed]
  149. Attardo, G.M.; Higgs, S.; Klingler, K.A.; Vanlandingham, D.L.; Raikhel, A.S. RNA interference-mediated knockdown of a GATA factor reveals a link to anautogeny in the mosquito Aedes aegypti. Proc. Natl. Acad. Sci. USA 2003, 100, 13374–13379. [Google Scholar] [CrossRef] [PubMed]
  150. Wynant, N.; Santos, D.; Verdonck, R.; Spit, J.; van Wielendaele, P.; Broeck, J.V. Identification, functional characterization and phylogenetic analysis of double stranded RNA degrading enzymes present in the gut of the desert locust, Schistocerca gregaria. Insect Biochem. Mol. Biol. 2014, 46, 1–8. [Google Scholar] [CrossRef] [PubMed]
  151. Gong, L.; Chen, Y.; Hu, Z.; Hu, M. Testing insecticidal activity of novel chemically synthesized siRNA against Plutella xylostella under laboratory and field conditions. PLoS ONE 2013, 8, e62990. [Google Scholar] [CrossRef] [PubMed]
  152. Gong, L.; Yang, X.; Zhang, B.; Zhong, G.; Hu, M. Silencing of Rieske iron-sulfur protein using chemically synthesised siRNA as a potential biopesticide against Plutella xylostella. Pest Manag. Sci. 2011, 67, 514–520. [Google Scholar] [CrossRef] [PubMed]
  153. Papathanos, P.A.; Windbichler, N.; Menichelli, M.; Burt, A.; Crisanti, A. The vasa regulatory region mediates germline expression and maternal transmission of proteins in the malaria mosquito Anopheles gambiae: A versatile tool for genetic control strategies. BMC Mol. Biol. 2009. [Google Scholar] [CrossRef] [PubMed]
  154. Brown, A.E.; Crisanti, A.; Catteruccia, F. Comparative analysis of DNA vectors at mediating RNAi in Anopheles mosquito cells and larvae. J. Exp. Biol. 2003, 206, 1817–1823. [Google Scholar] [CrossRef] [PubMed]
  155. Brown, A.E.; Bugeon, L.; Crisanti, A.; Catteruccia, F. Stable and heritable gene silencing in the malaria vector Anopheles stephensi. Nucleic Acids Res. 2003, 31, e85. [Google Scholar] [CrossRef] [PubMed]
  156. Adelman, Z.N.; Anderson, M.A.; Morazzani, E.M.; Myles, K.M. A transgenic sensor strain for monitoring the RNAi pathway in the yellow fever mosquito, Aedes aegypti. Insect Biochem. Mol. Biol. 2008, 38, 705–713. [Google Scholar] [CrossRef] [PubMed]
  157. Chen, Y.J.; Liu, B.R.; Dai, Y.H.; Lee, C.Y.; Chan, M.H.; Chen, H.H.; Chiang, H.J.; Lee, H.J. A gene delivery system for insect cells mediated by arginine-rich cell-penetrating peptides. Gene 2012, 493, 201–210. [Google Scholar] [CrossRef] [PubMed]
  158. Ramesh Kumar, D.; Kumar, P.S.; Gandhi, M.R.; Al-Dhabi, N.A.; Paulraj, M.G.; Ignacimuthu, S. Delivery of chitosan/dsRNA nanoparticles for silencing of wing development vestigial (vg) gene in Aedes aegypti mosquitoes. Int. J. Biol. Macromol. 2016, 86, 89–95. [Google Scholar] [CrossRef] [PubMed]
  159. Meade, B.R.; Dowdy, S.F. Enhancing the cellular uptake of siRNA duplexes following noncovalent packaging with protein transduction domain peptides. Adv. Drug Deliv. Rev. 2008, 60, 530–536. [Google Scholar] [CrossRef] [PubMed]
  160. Meade, B.R.; Dowdy, S.F. Exogenous siRNA delivery using peptide transduction domains/cell penetrating peptides. Adv. Drug Deliv. Rev. 2007, 59, 134–140. [Google Scholar] [CrossRef] [PubMed]
  161. Airs, P.M.; Bartholomay, L.C.; Department of Pathobiological Sciences, University of Wisconsin-Madison, Madison, WI, USA. Unpublished work. 2016.
  162. Lacey, L.A.; Grzywacz, D.; Shapiro-Ilan, D.I.; Frutos, R.; Brownbridge, M.; Goettel, M.S. Insect pathogens as biological control agents: Back to the future. J. Invertebr. Pathol. 2015, 132, 1–41. [Google Scholar] [CrossRef] [PubMed]
  163. Carlson, J.; Olson, K.; Higgs, S.; Beaty, B. Molecular genetic manipulation of mosquito vectors. Annu. Rev. Entomol. 1995, 40, 359–388. [Google Scholar] [CrossRef] [PubMed]
  164. Levis, R.; Huang, H.; Schlesinger, S. Engineered defective interfering RNAs of Sindbis virus express bacterial chloramphenicol acetyltransferase in avian cells. Proc. Natl. Acad. Sci. USA 1987, 84, 4811–4815. [Google Scholar] [CrossRef] [PubMed]
  165. Attardo, G.M.; Hansen, I.A.; Raikhel, A.S. Nutritional regulation of vitellogenesis in mosquitoes: Implications for anautogeny. Insect Biochem. Mol. Biol. 2005, 35, 661–675. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Mosquito life cycle events in the context of RNAi interventions (see also Table 1). Key life stage events illustrated in the context of targeted RNAi in mosquito disease vectors (1–8). Existing frameworks for mosquito control are shown in the context of mosquito life events to highlight points where RNAi triggers can be delivered. Control measures for particular life stages include: (A) Attractive Toxic Sugar Bait (ATSB); (B) residual spray & Long Lasting Insecticidal Bed-nets (LLIN); (C) Attractive Baited Oviposition Trap (ABOT); and (D) larvicides and pupacides. All of these control measures are at risk of resistance development (red arrows). Numerals in the figure correspond to Section 2 of the text and Table 1.
Figure 1. Mosquito life cycle events in the context of RNAi interventions (see also Table 1). Key life stage events illustrated in the context of targeted RNAi in mosquito disease vectors (1–8). Existing frameworks for mosquito control are shown in the context of mosquito life events to highlight points where RNAi triggers can be delivered. Control measures for particular life stages include: (A) Attractive Toxic Sugar Bait (ATSB); (B) residual spray & Long Lasting Insecticidal Bed-nets (LLIN); (C) Attractive Baited Oviposition Trap (ABOT); and (D) larvicides and pupacides. All of these control measures are at risk of resistance development (red arrows). Numerals in the figure correspond to Section 2 of the text and Table 1.
Insects 08 00004 g001
Figure 2. Diverse examples of mosquito RNAi trigger delivery systems (see also Table 2). The variety of (A) delivery vehicles and (B) RNAi trigger producing nucleic acids employed to suppress genes in mosquito species. Examples shown include: (1) naked RNAi triggers such as dsRNA, siRNA, or chemically modified siRNA (star shape); (2) transfection agents with dsRNA or shRNA expressing plasmids; (3) nanoparticles of abiotic or biotic origin in combination with dsRNA or plasmids; (4) viral expression systems carrying dsRNA or ssRNA that is converted to dsRNA in the cell; (5) bacterial expression systems containing dsRNA or shRNA plasmids; and (6) yeast expression systems containing dsRNA or shRNA plasmids. Numerals in the figure correspond to Section 3 of the text and Table 2.
Figure 2. Diverse examples of mosquito RNAi trigger delivery systems (see also Table 2). The variety of (A) delivery vehicles and (B) RNAi trigger producing nucleic acids employed to suppress genes in mosquito species. Examples shown include: (1) naked RNAi triggers such as dsRNA, siRNA, or chemically modified siRNA (star shape); (2) transfection agents with dsRNA or shRNA expressing plasmids; (3) nanoparticles of abiotic or biotic origin in combination with dsRNA or plasmids; (4) viral expression systems carrying dsRNA or ssRNA that is converted to dsRNA in the cell; (5) bacterial expression systems containing dsRNA or shRNA plasmids; and (6) yeast expression systems containing dsRNA or shRNA plasmids. Numerals in the figure correspond to Section 3 of the text and Table 2.
Insects 08 00004 g002
Table 1. Example target genes for RNAi and phenotypic outcomes (see also Figure 1).
Table 1. Example target genes for RNAi and phenotypic outcomes (see also Figure 1).
Legend *FunctionGene of InterestAccessionOutcomeSpeciesReference
1DeathInhibitor of Apoptosis 1AAEL009074DeathAe. aegypti[10,52,53,54,55]
2Olfaction & SensationGustatory receptors 1 & 3AAEL002380, AAEL010058Inability to detect CO2Ae. aegypti[15]
Odorant binding protein 1CPIJ007604Reduced oviposition attractant sensingCx. quinquefasciatus[16]
3Blood FeedingAegyptinAGAP009974Diminished blood feeding successAe. aegypti[17]
4ReproductionOvary ecdysteroidogenic hormone receptorAAEL001915Diminished egg developmentAe. aegypti[56]
zero population growthAGAP006241Spermless malesAn. gambiae[14]
5EmbryogenesisFrazzledAAEL014592Malformed ventral nerve cordAe. aegypti[57]
6Larval & Pupal DevelopmentChitin synthase 1AAEL002718Disrupted peritrophic matrixAn. gambiae[49]
MorphogenesisProphenoloxidase IIIAY487171.1Malformed pharate adult cuticleAm. subalbatus[19]
7Pesticide ResistanceProtease m1 zinc metalloproteaseCPIJ012471Death by deltamethrin susceptibilityCx. pipiens[58]
8Pathogen Uptake, Development & TransmissionCasparAGAP006473Suppresses malaria parasite numbers in the midgutAn. gambiae[59]
* Legend refers to both Section 2 and Figure 1.
Table 2. Examples of RNAi trigger delivery systems (see also Figure 2).
Table 2. Examples of RNAi trigger delivery systems (see also Figure 2).
Legend *(A) Delivery Vehicle(B) Nucleic AcidRNAi TargetDelivery RouteReference
1NonedsRNAvATPase AAdult per os[11]
NonedsRNAP-glycoproteinLarval per os[127]
NonesiRNASemaphorin AEmbryo Injection[117]
2NonepMOS-dsRED plasmidAegyptinEmbryo Injection[17]
3Liposome (Effectene®)dsRNACaspase 1Larval per os[123]
ChitosandsRNAChitin synthase 1Larval per os[49]
4Sindbis virusLong hairpin RNAGATA factorAdult Injection[149]
DensovirusShort hairpin RNAvATPase ATransfection (C6/36 cells)[81]
5E. coliLong hairpin RNAAAEL001684Larval per os of E. coli[13]
6P. pastorisLong hairpin RNAJH acid methyl transferaseLarval per os of P. pastoris[27]
* Legend refers to both Section 3 and Figure 2.

Share and Cite

MDPI and ACS Style

Airs, P.M.; Bartholomay, L.C. RNA Interference for Mosquito and Mosquito-Borne Disease Control. Insects 2017, 8, 4. https://doi.org/10.3390/insects8010004

AMA Style

Airs PM, Bartholomay LC. RNA Interference for Mosquito and Mosquito-Borne Disease Control. Insects. 2017; 8(1):4. https://doi.org/10.3390/insects8010004

Chicago/Turabian Style

Airs, Paul M., and Lyric C. Bartholomay. 2017. "RNA Interference for Mosquito and Mosquito-Borne Disease Control" Insects 8, no. 1: 4. https://doi.org/10.3390/insects8010004

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

Article Metrics

Back to TopTop