Next Article in Journal
Long-Term Incubation PrPCWD with Soils Affects Prion Recovery but Not Infectivity
Next Article in Special Issue
Opportunistic Mapping of Strongyloides stercoralis and Hookworm in Dogs in Remote Australian Communities
Previous Article in Journal
Effects of Mycoplasmas on the Host Cell Signaling Pathways
Previous Article in Special Issue
A Host-Specific Blocking Primer Combined with Optimal DNA Extraction Improves the Detection Capability of a Metabarcoding Protocol for Canine Vector-Borne Bacteria
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mosquito-Borne Diseases Emergence/Resurgence and How to Effectively Control It Biologically

by
Handi Dahmana
1,2 and
Oleg Mediannikov
1,2,*
1
Aix Marseille Univ, IRD, AP-HM, MEPHI, 13005 Marseille, France
2
IHU-Méditerranée Infection, 13005 Marseille, France
*
Author to whom correspondence should be addressed.
Pathogens 2020, 9(4), 310; https://doi.org/10.3390/pathogens9040310
Submission received: 13 March 2020 / Revised: 20 April 2020 / Accepted: 21 April 2020 / Published: 23 April 2020

Abstract

:
Deadly pathogens and parasites are transmitted by vectors and the mosquito is considered the most threatening vector in public health, transmitting these pathogens to humans and animals. We are currently witnessing the emergence/resurgence in new regions/populations of the most important mosquito-borne diseases, such as arboviruses and malaria. This resurgence may be the consequence of numerous complex parameters, but the major cause remains the mismanagement of insecticide use and the emergence of resistance. Biological control programmes have rendered promising results but several highly effective techniques, such as genetic manipulation, remain insufficiently considered as a control mechanism. Currently, new strategies based on attractive toxic sugar baits and new agents, such as Wolbachia and Asaia, are being intensively studied for potential use as alternatives to chemicals. Research into new insecticides, Insect Growth Regulators, and repellent compounds is pressing, and the improvement of biological strategies may provide key solutions to prevent outbreaks, decrease the danger to at-risk populations, and mitigate resistance.

1. Introduction

The significant connection between fauna and flora in the world today is due to many factors, including the highest increase ever experienced in population growth accompanied by the evolution of transport systems. These factors disrupt biogeographic barriers and are followed by the first appearance of species in novel habitats [1,2]. In the Americas, incursions of these species are estimated to cause more than $120 billion in damage every year [3].
Deadly pathogens and parasites may be transmitted by arthropods [4], and the increasing global human and animal populations are threatened by such epidemics and pandemics [5]. Mosquitoes (Diptera: Culicidae) represent the most threatening vector due to their role in the transmission of dangerous pathogens [1]. Through trade and travel, key mosquito species are being introduced into novel habitats [2,6,7].
A number of chemical products formulated to provide a high safety profile are commercially available, but their toxicity to human skin and the nervous system can lead to several serious problems, such as rashes, swelling, and eye irritation [8]. The most important drawback of these products is the incidence of insecticide resistance, which has increased rapidly in recent years [9], and the extremely challenging or downright impossible task of finding and treating all mosquito breeding sites. New approaches and vector-control tools targeting aquatic stages and adults are urgently needed [10].
In this review, we discuss the current state of knowledge about mosquito-borne diseases and the latest figures from these resurgences, highlighting current techniques for their control and their limitations. We then focus on new innovative alternatives currently known but rarely used, others that are not used at all, and those that are still in the test or design phase but are very promising, which we suggest to be considered in the biological control of mosquito-borne diseases.

2. Resurgence of Diseases Transmitted by Mosquitoes

The three main mosquito genera, Anopheles, Aedes, and Culex, transmit the causative agents of numerous important diseases to humans as well as animals [11,12,13,14]. In this chapter, we briefly describe the resurgence of essential disease agents transmitted by mosquitoes and their impact on humans and animals.
Malaria is considered the most important parasitic disease of human beings and is currently endemic and transmitted by anopheline mosquitoes in more than 80 countries inhabited by approximately three billion people (Table S1; Figure 1), especially in sub-Saharan Africa, where more than 85% of cases and 90% of deaths occur, mainly in children younger than 5 years old. Malaria continues to cause phenomenal damage to public health (228 million cases worldwide, with 213 million (93%) reported in Africa alone, and severe outbreaks have recently ravaged many areas [15,16,17,18,19].
Wuchereria bancrofti and Brugia spp. can be transmitted by numerous mosquito species [13,20,21,22,23,24,25] (Table S1), and cause various clinical manifestations (25 million men with hydrocele and over 15 million people with lymphoedema) and at least 36 million people continue to have these chronic disease manifestations [26]. However, it is clear that eliminating lymphatic filariasis is not possible without controlling their vectors.
Dengue virus (DENV): Flaviviridae is responsible for dengue disease, caused by four distinct serotypes. Currently, it is the predominant arthropod-borne viral disease affecting humans [27], with 3.6 billion people living in areas at risk of transmission and hundreds of millions of dengue fever cases reported each year [28,29], causing ongoing epidemics in several countries [29,30] (https://www.outbreakobservatory.org/outbreak-thursday) (Table S1; Figure 1)
Zika virus (ZIKV): Flaviviridae also causes ongoing epidemics in several countries in Latin America and the Pacific [30,31,32,33,34] (https://www.who.int/emergencies/diseases/zika/en/) (Figure 1). Aedes aegypti is considered to be the primary vector associated with ZIKV outbreaks [35], while Ae. albopictus is considered a secondary vector [36]. However, several other species are also involved in the occurrence and transmission of this rapidly spreading virus [34,37,38] (Table S1). Currently, it is considered one of the most serious diseases threatening public health.
Chikungunya virus (CHIKV): Togaviridae is the causal agent of chikungunya fever (CHIKF) (Figure 1), known for producing an antalgic stance gait with severe articular pain [39]. Infected patients evolving to the chronic stage may range from 1.4% to 90% (52% in the American continent) [39]. Numerous outbreaks have recently been reported in several countries [30,40,41,42].
Yellow fever virus: Flaviviridae [43] is a haemorrhagic and potentially lethal RNA virus that causes outbreaks in several countries, especially in unvaccinated populations [44,45,46,47,48] (Table S1). Its emergence is cyclical, and outbreaks occur approximately 7–10 years apart [49]. In the summer of 2016, 47 countries declared YFV endemic, and 42 countries identified a risk of transmission, with 29 of them in Africa in 2017 [45]. With the highest fatality rate of up to 33.6%, numerous outbreaks continue to be registered [44,50]. Vaccination is safe, affordable, and the most effective way to prevent YF: “70 to 90 million doses are annually produced worldwide” [45].
Annually, the WHO reports approximately 67,000 cases of Japanese encephalitis, 20% to 30% of which are fatal, while 30% to 50% of survivors have significant neurological sequelae [51]. New strains genetically close to strains involved in previous outbreaks continue to be identified [52]. The St. Louis encephalitis virus was the major cause of epidemic encephalitis by an arbovirus in the USA [53]. It is re-emerging, causing numerous cases [54] (Table S1).
Similar to humans, horses are the domesticated animal that is most commonly affected by West Nile virus; 80% of cases are asymptomatic, while neurological signs are the most commonly reported symptom, with 90% of the 20% developing clinical signs, and the mortality rate may reach 30% [55]. Nevertheless, recent outbreaks in humans have been highlighted [56,57].
Different pathogenic blood-borne bacteria are regularly detected in mosquitoes [58,59]. It is not yet clear whether the presence of these bacteria in mosquitoes may be explained by occasional ingestion with blood meals or acquisition from the environment, or whether these bacteria may multiply and eventually be transmitted during blood meals. Different pathogenic alpha-proteobacteria, including Anaplasma spp., Ehrlichia spp., Candidatus Neoehrlichia, Bartonella spp., and Rickettsia spp., have been identified (xeno-monitoring studies) in adult mosquitoes [59,60]. More interestingly, the agent of febrile rickettsiosis, Rickettsia felis, has not only been identified in mosquitoes [58,61] but also shown to be potentially transmitted by Anopheles mosquitoes in laboratory experiments [62]. Francisella tularensis [63] is also carried by mosquitoes (Aedes), which act as a main vector in Sweden and Finland, making it the first reported mosquito-borne bacterium [63].
Several complex factors may explain the expansion of these diseases, such as population growth, globalisation of the economy, international travel (recreational, business, and military), inadequate vector-control efforts, limited access to good healthcare, rapid and unplanned urbanisation of tropical regions coupled with poor sanitary conditions, and a deterioration of public health infrastructures, all of which are related to climate change [64]; but, the major factors remain the mismanagement of insecticide use and the emergence of resistance.

3. Actual Insecticide-Based Vector-Control Strategies

The debate regarding dichlorodiphenyltrichloroethane (DDT) use for prevention, especially for malaria control, is polarised because it saved millions of lives worldwide but is unsafe. This has led to the invocation of precautions to enable choices to be made for healthier lives [65]. Some studies have focused on predicting mosquito abundance and assessing aquatic and adult mosquito control strategies [66], but despite the added efforts to develop new insecticides, other new alternative classes are slowly emerging [67,68].

3.1. Indoor Residual Spraying (IRS)

This is a well-developed and effective but potentially underused approach in vector control. It consists of treating the surfaces upon which common mosquitoes rest inside houses with a long-lasting insecticide. The most affected species among the endophilic species will be Ae. aegypti, which rests mainly indoors, feeds on humans, and is thus more likely to be reached by IRS than by space sprays [69]. IRS has some limitations and imperfections, such as the need for specialised training, which is time consuming in terms of obtaining public acceptance within a region. It does not prevent people from being bitten but above all, it must be adapted to several factors specific to a region, such as insecticide resistance, which is expensive and takes several years [70]. IRS has had a considerable impact on the mortality of Ae. aegypti, and used alone [71] or in combination with larval control [72] contributed to the elimination of Ae. aegypti in Guyana and the Cayman Islands, respectively [70]. In 2006, the WHO reaffirmed the importance of IRS for malaria transmission control, which was supported by the President’s Malaria Initiative (PMI) in 2012 [73]. Malaria eradication campaigns using IRS in the Mediterranean region seem to have led to the elimination of malaria [74]. New formulations could last between five and eight months [70]. The potential evolution of insecticide resistance in the vector to pyrethroids can be controlled using alternative formulations, such as bendiocarb [75] and other new IRS formulations [67,73,76,77]. Good insecticide management is based on an alternation of formulations to combat the evolution of resistance, which may maintain efficacy over time, especially for location-specific interventions [70].

3.2. Peridomestic Space Spraying

This strategy is attractive because it is highly visible and conveys the message that health authorities use vector-control activities [78]. The risks to humans due to the management of adult mosquitoes are probably negligible [79]. This has no direct impact on immature stages (egg, larvae, or pupae) [80], targeting adult mosquitoes only, and is performed by spraying small droplets of insecticide into the air. It is used mainly in emergency situations to limit the massive production of adult mosquitoes, thus decreasing the risk of existing outbreaks expanding [78].
To perform this intervention, two forms of space sprays are commonly used for control: thermal fog and cold fog, also known as ultralow volume (ULV) sprays. Both can be distributed using a vehicle-mounted or hand-held machine [78]. The insecticide concentration ranges from 2% (pyrethroids) to 95% (organophosphates), depending on the amount of active ingredient in the formulation. The applied volume is dependent on the compound concentration and toxicity to the target species [80]. Aerial spraying of pyrethrin significantly impacts small organisms found in the sprayed zones, which is not the case on large bodies [81].
For dengue control, mosquitoes emerging after treatment can still be vectors because the viruses can be transmitted transovarially. Therefore, their exposure to successive treatments seems necessary and should be done at intervals shorter than the extrinsic incubation period of the virus [82].
A high resurgence of mosquitoes was reported after six days of ULV treatment as a single method in Thailand [81], while good results were observed with a decrease in the incidence of dengue fever after a large emergency vector-control campaign included several space sprays [83].

3.3. Long-Lasting Insecticide-Treated Nets (LLINs)

Designed as a solution to the problems of conventional insecticide-treated nets (ITNs), and based on novel fabric technologies [84], LLINs were developed to resist multiple washes and remain effective for a prolonged time (at least three years). LLINs are considered one of the most successful mosquito control tools, especially for malaria prevention [85]. ITNs with synthetic pyrethroid insecticides either incorporated into or coated around their fibres have resulted in a considerable decline in malaria morbidity and mortality in several countries, especially in sub-Saharan Africa, where over 427 million nets were delivered between 2012 and 2014 [85]. Th annual cost of an LLIN can be as high as US$2.6, while IRS costs about US$4, and standard ITN costs range from US$1.5 to US$6 [85]. The level of use of LLINs varies according to several factors, such as temperature, humidity, season, and, especially, the density of mosquitoes, and access to them plays a major determinant of their use [86,87]. LLINs have contributed to the reduction in malaria over the past 15 years, combined with other new control measures, such as IRS and artemisinin-based combination therapies [88] in children and pregnant women [87,89]. Other important advantages of LLINS include reduced consumption of insecticides and insecticide released into the environment because they do not need retreatment [85]. The efficacy of LLINs is closely related to the molecules used (the choice depends on the presence or absence of its resistance) [90], and their correct use may enhance their efficiency [91]. In a study, the use of LLINs led to a dramatic reduction (97%) in the prevalence of malaria compared to a group of LLIN non-users [92].

3.4. Mosquito Repellents

Mosquitos are mostly attracted to humans by the lactic acid and CO2 present in our sweat that are detected by chemoreceptors present in their antennae, and repellents mask the human scent [8]. DEET (N,N-diethyl-meta-toluamide) is the most widely used and effective repellent against mosquitoes [93].
Biobased mosquito repellents are pest management tools that are based on safe, biologically based active ingredients derived from plants [94,95], fungi [96], or bacteria [93].
In terms of the effective control of mosquitoes and to ensure human and environmental safety where endemic mosquito resistance and environmental concerns limit the use of products, the use of biobased natural mosquito repellents is preferable to that of chemical repellents [8].
The most effective synthetic repellents are DEET (N,N-diethyl-m-toluamide) and IR3535 (3-(NButyl-N-acetyl)-aminopropionic acid [97]. Several nanoparticles synthesised and successfully impregnated into cotton fabrics in insect-repellent clothing show high efficacy against mosquito larvae and adult populations, which gives them the potential to be used as eco-friendly approaches to control mosquitoes if applied in long-lasting insect-repellent clothing [98,99]. The fact that the use of synthetic repellents causes insecticide resistance in mosquitoes, has a harmful effect on non-target organisms, and threatens the environment has led to widespread discussions around this method of control [97].
The increased involvement of governments and authorities on scientific projects coupled with correct individual action may help to combat the spread of mosquito-borne diseases and limit their devastating transmission.

4. Biological Control

Every year, promising new “eco-friendly” compounds are developed to progressively replace the oldest compounds, which are the most toxic and harmful. The use of biological control programmes, such as genetic modification or biological agents such as predatory fish, bacteria, protozoa, nematodes, and fungi, have rendered some promising results.

4.1. Genetic Modification

The sterile insect technique (SIT) is a species-specific and environmentally benign method for insect population control based on mass rearing, radiation-mediated sterilisation, and the release of a large number of male insects into a given target area, which compete for mates with wild males. A wild female mating with a released sterile male has no or fewer progeny, so the population tends to decline [100,101,102,103], which was an improvement on RIDL (release of insects carrying a dominant lethal gene). The lethal dominant gene could be controlled by a female-specific promoter and its expression could be inactivated by antibiotic treatment (tetracycline), allowing the mosquito-colony to be maintained. When male and female separation is required, the antibiotic is removed from the system, causing the death of all females [10]. Some projects cost approximately US$1.1 million [104], and some reports of failure have been published [105]. Mosquito egg production and mass rearing problems were also highlighted [106,107].
Paratransgenic strategies based on genetically modified symbiotic bacteria reintroduced in mosquitoes reveal a very high potential of casually controlling all-important mosquito species, including Culex, which is difficult to transform [14]. New studies on RNAi-based bioinsecticides (RNA interference) show promising results [108].

4.2. Fungi

Particular attention has been paid to fungal species belonging to the genera Lagenidium, Coelomomyces, Entomophthora, Culicinomyces, Beauveria, and Metarhizium for their power to reduce mosquito populations, but unfortunately, none of them have been specifically adapted as larvicidal agents against important vector species [109,110,111], even transgenic ones [112]. Application to surfaces on which mosquitoes land or need to pass through, such as fungus-impregnated cloths around bed nets, attractive bait stations, and adult mosquito traps and PET traps, show promising results, with a 39–50% reduction in survival rates of malaria-carrying mosquitoes and elimination of 95% of Anopheles arabiensis mosquitoes in a bait station [113]. One of the most effective fungi studied recently against simultaneously Ae. albopictus and Cx. pipiens mosquito adults is Beauveria bassiana. The production and persistence of its conidia was remarkably high [109].

4.3. Control of Aquatic Stages Using Elephant Mosquito and Fish Predators

The use of fish to control the aquatic stages of mosquitoes was an important tool in the pre-DDT era. These fish were introduced into all potential mosquito-breeding habitats and their use decreased after the introduction of DDT and then was rekindled after the development of resistance and harmful effects [114]. The use of indigenous larvivorous fishes is suggested [115], and a limited number of species are used, primarily Gambusia affinnis and Poecilia reticulata, although several failures have been reported in the literature [114]. Other aquatic predators may play a role in reducing mosquito populations, especially in rainy periods [116,117], and the combination of multiple predators can reduce mosquito populations [118].
The naturally occurring non-biting Toxorhynchites species, which exhibit predatory behaviour during their larval stages, have been explored for their potential use as biological control alternatives to chemical insecticides (the 4th instar larva is the most predaceous) [119,120]. Important progress was made concerning their production for use as biological agents and they demonstrated remarkable effectiveness against numerous mosquito species, such as Ae. aegypti, Ae. albopictus, and Cx. quinquefasciatus [119,121]. In certain situations, they have demonstrated practical potential, but their use continues to be limited by several problems, such as cannibalism during the early instars, temperature (limited by low temperatures), and also the inadequate overlap in the larval habitats between the prey and the predator mosquito [120].

4.4. Protozoan Control

Chilodonella uncinata is a protozoan parasite with many beneficial properties associated with a good microbial pathogen [122]. It causes low to very high (25–100%) mortality in mosquito larvae. It exhibits high virulence and resistance to desiccation and also demonstrates a high reproductive potential when cultured in vitro. Through its mosquito host, C. uncinata has the ability to spread in nature by the way of transovarian transmission [122].

5. Bacterial Agents Tested or Used in Control Strategies

Most of the attention of pest control scientists focuses on bacterial agents targeting both aquatic and adult stages. Several studies have shown their efficiency, and their use is recommended by the WHO. Here, we list some bacterial agents currently in use or undergoing tests with promising results.

5.1. Bacillus spp.

Before the discovery of Bacillus thuringiensis israelensis (Bti) (Figure 2) and Bacillus sphaericus (Bs) (Figure 3), little attention was paid to bacteria as sources of agents for microbial control of mosquitoes. Around 1500 microorganisms were recently identified as good potential insecticidal agents, and looking for insecticidal activity, metabolites from approximately a thousand microbial isolates were examined [123]. Bti formulations are the predominant nonchemical means employed for controlling mosquito larvae [124]. In addition, several studies indicate the highly effective and safe use of individuals or the mixture of Bti and Bs for mosquito control, and they are considered to be safe to non-target organisms cohabiting with mosquito larvae [125].
B. thuringiensis produces three classes of larvicidal proteins: Cry (exert intoxication through toxin activation, receptor binding, and pore formation in a suitable larval gut environment), Cyt (cytolytic toxicity) when sporulating (parasporal crystals), and Vip proteins throughout the vegetative phase (ionic, non-ionic detergents and pore-forming mechanisms of action were suggested), some of which are toxic against a wide range of insect orders, nematodes, and human-cancer cells. This has been widely employed as an effective biopesticide to control pests that are harmful to crops, forests, and humans. Cyt toxins possess less toxicity against mosquito larvae than Cry toxins [126,127].
Several species of B. thuringiensis exhibit a high mortality rate toward all mosquito larval instars, such as Bti [128], B. thuringiensis var. krustaki [129], B. thuringiensis var. jegathesan (Btjeg) [130], B. thuringiensis var. kenyae, and B. thuringiensis var. entomocidus [131]. Other species with homology to Bacillus show the highest toxicity against dipterans, such as Clostridium bifermentans (serovar malaysia) [132], B. circulans [133], and B. laterosporus [134,135]. Bacillus spp. remains a massive source of active compounds against pests, which are currently being explored to fill public health needs.

5.2. Insect Growth Regulators (IGRs)

Due to several advantages, such as low toxicity to the environment and selectivity, IGRs present an effective tool to control mosquito populations. They are substances that are analogues or antagonists of hormones and interfere with insect development [136]. There is growing interest in the use of IGRs, such as methoprene and pyriproxyfen, two juvenile hormone agonists belonging to IGR insecticides. They are effective against mosquito larvae and may inhibit the emergence of adults [137]; others include novaluron and diflubenzuron [138] for mosquito control [139]. Numerous recent studies have highlighted that mosquitoes and other pests have developed resistance to commonly used IGRs, such as methoprene and pyriproxyfen [140,141,142], which reinforces the need to develop new compounds and identify new targets in mosquitoes [143].

5.3. Wolbachia spp.

Mosquito symbiont-associated bacteria may exert a pathogenic effect on their host, interfering with its reproduction and also reducing vector competence [144]. Wolbachia are endosymbiotic bacteria that naturally infect approximately 40% of insect species [145,146], and Wolbachia pipientis is a unique valid species of the genus [147]. They are present in some major mosquito disease vectors, such as Cx. quinquefasciatus, Ae. albopictus, and anopheline species, including malaria vectors such as An. gambiae and An. coluzzii but never Ae. aegypti [144,145,148,149,150]. This maternally transmitted bacterium allowing the invasion of host populations can induce feminisation of males (turning genetic males into females), parthenogenesis (reproduction without males) [144], and cytoplasmic incompatibility, leading to the generation of inviable offspring when a Wolbachia-infected male mates with an uninfected female, but not in the contrary case [145]. Successfully used in Myanmar in the 1960s to eradicate Cx. quinquefasciatus [151], it is currently also being used to target Ae. albopictus, using a triple Wolbachia-infected strain [152], and to target Ae. polynesiensis (2012) [153]. To date, it has been used in several countries, such as Australia, Brazil, Indonesia, Vietnam, and Colombia. The fear of resistance to the inhibitory effect of Wolbachia has been highlighted, but no studies have demonstrated that this scenario is likely to happen, and the creation of Wolbachia-superinfected lines, such as Ae. aegypti with stable infection, could help to mitigate potential resistance [145,154] and add to their role in reducing vector competence. Studies have reported that Wolbachia inhibits the transmission of CHIKV [155], YFV [156], malaria parasites in An. stephensi [157] and An. gambiae [158], DENV [159], and ZIKV [160]. Recent reviews clearly explain Wolbachia as a form of biological control [144,161,162].
Wolbachia-based control constitutes a potentially promising strategy for the control of mosquitoes and their transmitted diseases that urgently needs to be considered and associated with biological control programmes in countries suffering from malaria and arbovirus outbreaks.

5.4. Asaia

To make malaria vectors inefficient, interruption of the cycle within the vector to stop parasite development before the Anopheles host becomes infective is a good solution [163]. The simplest approach to this is paratransgenesis, consisting of producing bacterial strains that are able to both live in the midgut of various mosquito species and spread rapidly among wild mosquito populations [164]. Several studies have been performed on the identification and use of competent microorganisms to combat vector-borne diseases [165]. The genus Asaia, first discovered in plant nectar, is an excellent candidate [166]; it is localized in many organs of mosquitoes, and can disperse inside the mosquito body through the haemolymph [165,167]. Its distribution in the mosquito population is made possible through several mechanisms (co-feeding, sexual mating, paternal, maternal, and horizontal transmission) [168,169,170]. Asaia bacteria may be genetically modified in order to be recolonised in a new host, resulting in spread within wild populations [166]. Recently, it was isolated and characterised from several Anopheles species, which would be beneficial if applied toward achieving paratransgenesis against malaria [165]. Advanced studies recently showed that Asaia may activate the mosquito’s immune system, leading to a reduction in the development of malaria parasites [171]. In the future, additional assets to which the bacterium may be used in mosquito control may be identified because it seems that Asaia plays a key role in the health of the mosquito host, even during its larval stage, allowing the larvae to develop rapidly [172]. Engineering of Asaia to produce an antiplasmodial effector causing the mosquito to become refractory to Plasmodium berghei is a perfect demonstration of the power of a transgenic microbiota [173], which makes it beneficial to microbial ecology and a potential candidate not only for paratransgenesis but also for general control of mosquitoes and mosquito-borne diseases.

5.5. Spinosyns

Spinosad is a biopesticide derived via fermentation from an actinomycete, Saccharopolyspora spinosa, a naturally occurring soil-dwelling bacterium. It contains two insecticidal factors, A (C41H65NO10) and D (C42H67NO10) [174,175]. It is categorised as a Group 5 insecticide by the Insecticide Resistance Action Committee (IRAC), forming a new class of polyketide-macrolide insecticides that act as nicotinic acetylcholine receptor (nAChR) allosteric modulators. Discovered in the 1980s in an early-stage insecticide screen that included Ae. aegypti, it was shown to be highly active against numerous pests in the Lepidoptera, Diptera, Thysanoptera, Coleoptera, Orthoptera, and Hymenoptera orders, and others. Its application to mosquito control is relatively new due to its pesticidal activity after ingestion and cuticle absorption and its highly favourable toxicology profiles in mammals and the environment [176]. It was also recently approved for use as a mosquito larvicide in human drinking water sources and containers [177]. Its applications in natural habitats are too few, but in laboratories it has been demonstrated to be very efficient at preventing and reducing larval development in important medical and veterinary vector species, such as Ae. aegypti, Ae. albopictus (Figure 4), Anopheles gambiae (Figure 5), An. pseudopunctipennis, An. albimanus, Cx. pipiens (Figure 6), Cx. quinquefasicatus [175,178], and some anopheline species [179,180].
As the best solution, biological control requires several components for the design of effective plans for mosquito control, and spinosad, which has no resistance because it was introduced recently into control programmes, will play a very important role.

5.6. Bacterial-Based Feeding Deterrents and Repellents

The fear of the occurrence of possible side effects of DEET, such as toxic encephalopathy, seizures, acute manic psychosis, cardiovascular toxicity, and dermatitis [181], as well as potential resistance that has become a reality with Ae. aegypti mosquitoes [182] and An. gambiae [183], has led to the use of innovative technologies to create other products free of DEET that are marketed in the form of sprays or creams and include other active ingredients [184], such as picaridin [185] and IR 3535 [186], and a wide range of essential oils, which synergistically use various components and have been reported to provide a higher repellent activity than single isolated components [187]. Recently, a mixture of compounds isolated from Xenorhabdus budapestensis (entomopathogenic-associated bacteria) exhibited potent feeding-deterrent activity against three mosquito species considered to be the most important vectors of diseases affecting public health. They belong to the fabclavine class and exhibit a high activity comparable to or better than that of DEET or picaridin in side-by-side assays [93], which supports the attempt to replace toxic molecules by considering bacteria as a very promising source of new alternative molecules for exploitation as mosquito repellents.

6. Biological Insecticide Resistance

In view of their efficacy and safety, the importance of bacterio-insecticides seems to be increasing in insect control activities, which has led researchers to investigate and characterise new bacterial strains with insecticidal properties and identify their active compounds.

6.1. Resistance to Bti

Numerous factors, such as wild proliferation or environmental accumulation, as well as the persistence of human-spread Bti in treated larvae breeding sites, may lead to a long exposure time of insects, which may increase the risk of acquiring resistance in target insects and also have a negative impact on non-target insects [188]. A study showed that a high resistance to each individual Bti toxin can be obtained under some conditions in the laboratory after only a few generations of selection, and this resistance seems to be lowest for commercial and environmental Bti, which might act as a first step in resistance to a complete Bti toxin mixture (Table 1). Studies reporting that individuals show resistance to one toxin but not to another suggest that different resistance mechanisms exist [189]. The mechanisms of resistance to Bti Cry toxins are widely studied in Culex and Aedes species [176,190,191,192,193,194], and marginal cross-resistances have been identified [193,195]. To date, resistance of malaria-carrying species to Bti has not been found [196].

6.2. Resistance to Bs

B. sphaericus (Bs) is found in numerous habitats, especially in soils and aquatic habitats. It is known as a producer of a characteristic spherical spore inside the swollen sporangium. Over the past 25 years, scientists had much interest and focused on isolating numerous strains because of their potential use as mosquito larvicides [200]. Several formulations used in biocontrol are highly effective against mosquitoes [125]. Recently renamed Lysinibacillus sphaericus (2007) [201], numerous studies have reported various levels of resistance to Bs in laboratory and field populations from different countries [176] (Table 1), mostly on Culex populations. Different mechanisms [202] have been observed in numerous locations, such as France, China, India, and Brazil [126,203]. If mosquitoes develop resistance to one strain of Bs, it appears that they will develop resistance to other Bs strains due to the similarity of the binary toxins in most strains; but, they remain susceptible to Bti [176]. Resistant strain fitness was found to be heavily impacted, especially fecundity and fertility, which became very low in a study [204,205], although opposite results were achieved in another study [206].
Although it has been tested widely for controlling malaria vectors [207,208,209,210], no laboratory or field resistance has been highlighted for Anopheles species to date.

6.3. Resistance Management

When it comes to managing the rapid increase in insecticide resistance [9], Bti can be used as a powerful tool to mitigate resistance to Bs in mosquitoes, although it has been reported that using them in rotation or in a mixture leads to a steady decline in resistance over 30 generations. They can also delay or prevent the emergence of resistance due to the synergistic action between their toxins, and recent formulations have shown greater larvicidal activity and efficacy [176]. Other combinations with botanical pesticides are considered alternatives to mitigating the development of resistance to Bs in mosquitoes [211].
The best solution for the management of insecticide resistance is to systematically replace most of the molecules used in chemical control with eco-friendly biological control, but that strategy will depend on the plans conceived, the number of molecules chosen and the associations between the various formulations to prevent and reduce current resistance and avoid the appearance of new resistance.

7. Current Challenges for a Prosperous Future

In view of the current situation and the failures that have been experienced with mosquitos invading new territories associated with devastating outbreaks, new tools, molecules, plans, synergistic associations, and methods of mosquito control are being developed to facilitate strategic objectives, such as protecting at-risk populations, especially in endemic areas; preventing the international spread of mosquitoes and the diseases they carry; and rapidly containing epidemics. Some strategies are in the stage of preliminary testing or in the validation phase and others have recently been introduced into use.

7.1. New Insecticide, IGR, and Repellent Compounds

The most urgent need is to develop new insecticides to fight mosquito-borne diseases due to their crucial efficiency and their economic importance [212]. The Innovative Vector Control Consortium (IVCC) has released new product classes, especially for malaria eradication, and manages international efforts to establish new methods, including producing a new ATSB (attractive targeted sugar bait) product class and programming next-generation IRS projects [213]. Due to their eco-friendly properties and efficiency, entomopathogenic Ascomycete fungi have been suggested for the control of both larval and adult stages of dengue vectors [12,110]. Several other bacteria showing promising results on numerous pests have been suggested to have the same effect on mosquitoes, such as the entomopathogenic nematode-associated bacteria Xenorhabdus sp. [93,214]; Serratia marcescens, which is often associated with insect infection and shows high insecticidal effects alone [215] or when associated with other insecticides [216]; and entomopathogens [217]. Other bacteria exhibiting toxic effects on mosquitoes, such as Clostridium bifermentans [132], may also be considered in control strategies.
Recently, a new compound class, chalcones with JHAN activity, showed impressive insecticide and IGR activity when tested against Ae. albopictus larvae and could be useful for the development of environmentally benign IGR insecticides to control mosquitoes [143]. Moreover, the beneficial effects of diterpene and their derivatives as well as their potential use as biological alternatives in dengue fever control has been highlighted [218].
Auto-dissemination is a phenomenon where the dispersal and transfer of active compounds is carried out by contaminated adult mosquitoes to treat undeveloped habitats that are difficult to locate and treat [219]. It can occur through treated materials or dissemination stations, such as modified ovitraps, and can also be combined with other methods, such as SIT [12], which may increase their effectiveness.

7.2. Attractive Toxic Sugar Baits (ATSB)

Bait aims to attract mosquitoes in order to feed them on toxic sugar meals broadly sprayed on plants or placed in bait stations [220,221]. They show the highest efficacy in laboratory and field studies [222] against Aedes species, culicines, and sand flies [12].
Whether for indoor or outdoor control, ATBS can reduce mosquito populations through direct mortality caused by feeding them on insecticide-treated bait but also through the spread of mosquito pathogens or non-chemical toxins [223]. Developing mosquito-specific attractants to avoid their effects on non-target species make baits one of the best solutions, and their combination with other strategies, such as genetic ones, will maximise their effectiveness.

7.3. Parasitic Nematodes

Lot of nematodes belonging to numerous orders and families are known to be parasites of insects [224]. Some insect parasitic nematodes that are specific to mosquitoes [225] may be considered alternatives to chemical insecticides [226]. When tested, they were effective against malaria vectors and several other important mosquito species, such as Ae. aegypti, Ae. albopictus, Cx. quinquefasciatus, and An. gambiae [226,227,228,229]. As they are naturally adapted to their host, such nematodes are highly specific to theirs hosts, which they can kill by producing high levels of parasitism. They are free swimming and disseminate easily in the infective stage [225], and species such as Romanomermis iyengari are widely suggested to be a component of integrated mosquito control programmes in lymphatic filariasis endemic countries [229].

7.4. Acoustic Larvicides and Traps

These are emerging technologies designed to combat the aquatic stages of mosquitoes by killing them with sound waves resulting in instantaneous mortality or inhibited emergence. They have proven to be effective as a beneficial non-chemical alternative for the treatment of drinking water supplies [230]. This approach has been shown to be highly effective in a range of typical volumes found in peri-domestic water containers [230] without causing resistance within mosquito populations or harming non-target organisms when used properly [231]. Furthermore, even simple and cheap mobile phones can sensitively acquire acoustic data on the species-specific level of adult wingbeat sounds. This makes it possible to simultaneously record the time and location of the encounter between humans and mosquitos, which forms a powerful tool for acoustically mapping mosquito species distribution worldwide [232]. Other innovative acoustic-based tools have been developed to control mosquitoes during rear-and-release operations, such as the low-cost and battery-powered sound-baited gravid Aedes trap, which may be an effective replacement for the costly Biogents Sentinel (BGS) trap [233].

7.5. Advanced Genetic Studies

Recently, a new RNAi-based bioinsecticide was developed from D-RNA molecules, which was subsequently tested on Aedes larval breeding water [108]. A significant reduction in the viability of the larvae treated with dsRNA was reported while in the surviving larvae and adults, altered morphology and chitin content was observed. In combination with diflubenzuron, this innovative bioinsecticide had insecticidal adjuvant properties [108].
In another study, a considerable reduction in the fertility of Ae. aegypti adult males was observed when feeding their larvae double-stranded RNAs (dsRNAs) targeting testis genes. Moreover, several dsRNAs were reported to be inducing males and were remarkably effective in competing for mates. RNAi-mediated knockdown of the female-specific isoform of double-sex was also effective in producing a highly male-biased population of mosquitoes, making it possible to overcome the need to sex-sort insects before release [234].

8. Conclusions

Despite currently deployed methods, epidemics and the spread of mosquito-borne diseases continue as a result of a range of complex reasons, including insecticide resistance, inappropriate design of control programmes, ineffective coverage, missing and poorly trained manpower, as well as a lack of financial resources and infrastructure [12].
Many strategies have been designed for the control of mosquito-borne diseases, each with their strengths and weaknesses. However, approaches such as integrated vector management that adopts receding horizon control strategies, which may consider multiple objectives, seem to provide optimal control solutions that are fast and sustainable but that also offer the most cost-effective control choices [235].
Improving current strategies, such as the sterile insect technique, the release of insects with dominant lethality, or transgenesis, may provide key solutions to preventing outbreaks, decreasing the danger to at-risk populations and mitigating resistance. Meanwhile, promising techniques, such as those discussed in this manuscript, have already proven their effectiveness but remain under-used and require more attention and consideration in vector-control plans.

Supplementary Materials

The following are available online at https://www.mdpi.com/2076-0817/9/4/310/s1, Table S1: Updates concerning important mosquito borne diseases. We listed most of the mosquito-borne diseases, including their actual distribution, transmission, natural occurrence or animal infection, and virulence as well as the existence or absence of treatments or vaccines to date.

Author Contributions

H.D.: Investigation; software; writing—original draft preparation; O.M.: Conceptualisation; methodology; validation; writing—review and editing; supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

We thank Jean Michel Beranger for providing us with high-quality pictures of insects from our insectary.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Benelli, G. Research in mosquito control: Current challenges for a brighter future. Parasitol. Res. 2015, 114, 2801–2805. [Google Scholar] [CrossRef] [PubMed]
  2. Kilpatrick, A.M. Globalization, land use and the invasion of West Nile virus NIH Public Access. Science 2011, 334, 323–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Pimentel, D.; Zuniga, R.; Morrison, D. Update on the environmental and economic costs associated with alien-invasive species in the United States. Ecol. Econ. 2005, 52, 273–288. [Google Scholar] [CrossRef]
  4. Mehlhorn, H. Encyclopedia of Parasitology; Institut für Zoomorphologie, Zellbiologie und Parasitologie: Düsseldorf, Germany, 2016; Volume 1, ISBN 978-3-662-43977-7. [Google Scholar]
  5. Mehlhorn, H.; Al-Rasheid, K.A.S.; Al-Quraishy, S.; Abdel-Ghaffar, F. Research and increase of expertise in arachno-entomology are urgently needed. Parasitol. Res. 2012, 110, 259–265. [Google Scholar] [CrossRef]
  6. Hubálek, Z.; Halouzka, J. West Nile fever—A reemerging mosquito-borne viral disease in Europe. Emerg. Infect. Dis. 1999, 5, 643–650. [Google Scholar] [CrossRef]
  7. Vila, M.; Hulme, P.E. Impact of Biological Invasions on Ecosystem Services; Springer International Publishing: Berlin/Heidelberg, Germany, 2017; ISBN 978-3-319-45121-3. [Google Scholar]
  8. Shukla, D.; Wijayapala, S.; Vankar, P.S. Effective mosquito repellent from plant based formulation. Int. J. Mosq. Res. 2018, 5, 19–24. [Google Scholar]
  9. 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]
  10. Toledo Marrelli, M.; Barretto Bruno WILKE, A.; Toledo Marrelli, M. Genetic Control of Mosquitoes: Population suppression strategies. Rev. Inst. Med. Trop. São Paulo 2012, 54, 287–292. [Google Scholar]
  11. Benelli, G.; Beier, J.C. Current vector control challenges in the fight against malaria. Acta Trop. 2017, 174, 91–96. [Google Scholar] [CrossRef]
  12. Achee, N.L.; Grieco, J.P.; Vatandoost, H.; Seixas, G.; Pinto, J.; Ching-Ng, L.; Martins, A.J.; Juntarajumnong, W.; Corbel, V.; Gouagna, C.; et al. Alternative strategies for mosquito-borne arbovirus control. PLoS Negl. Trop. Dis. 2019, 13, e0006822. [Google Scholar]
  13. Omori, N. A review of the role of mosquitoes in the transmission of Malayan and bancroftian filariasis in Japan. Bull. World Health Organ. 1962, 27, 585–594. [Google Scholar] [PubMed]
  14. Barretto, A.; Wilke, B.; Toledo Marrelli, M. Paratransgenesis: A promising new strategy for mosquito vector control. Parasites Vectors 2015, 8, 342. [Google Scholar]
  15. World Malaria Report. 2019. Available online: https://www.who.int/news-room/feature-stories/detail/world-malaria-report-2019 (accessed on 26 March 2020).
  16. Rahmah, Z.; Sasmito, S.D.; Siswanto, B.; Sardjono, T.W.; Fitri, L.E. Malaria. Malays. J. Med. Sci. 2015, 22, 25–32. [Google Scholar] [PubMed]
  17. Feged-Rivadeneira, A.; Ángel, A.; González-Casabianca, F.; Rivera, C. Malaria intensity in Colombia by regions and populations. PLoS ONE 2018, 13, e0203673. [Google Scholar] [CrossRef]
  18. Amato, R.; Pearson, R.D.; Almagro-Garcia, J.; Amaratunga, C.; Lim, P.; Suon, S.; Sreng, S.; Drury, E.; Stalker, J.; Miotto, O.; et al. Origins of the current outbreak of multidrug-resistant malaria in southeast Asia: A retrospective genetic study. Lancet Infect. Dis. 2018, 18, 337–345. [Google Scholar] [CrossRef] [Green Version]
  19. Lok, P.; Dijk, S. Malaria outbreak in Burundi reaches epidemic levels with 5.7 million infected this year. BMJ 2019, 366. [Google Scholar] [CrossRef]
  20. Erickson, S.M.; Thomsen, E.K.; Keven, J.B.; Vincent, N.; Koimbu, G.; Siba, P.M.; Christensen, B.M.; Reimer, L.J. Mosquito-parasite interactions can shape filariasis transmission dynamics and impact elimination programs. PLoS Negl. Trop. Dis. 2013, 7, e2433. [Google Scholar] [CrossRef]
  21. Gleave, K.; Cook, D.; Taylor, M.J.; Reimer, L.J. Filarial infection influences mosquito behaviour and fecundity. Sci. Rep. 2016, 6, 36319. [Google Scholar] [CrossRef]
  22. Ughasi, J.; Bekard, H.E.; Coulibaly, M.; Adabie-Gomez, D.; Gyapong, J.; Appawu, M.; Wilson, M.D.; Boakye, D.A. Mansonia africana and Mansonia uniformis are vectors in the transmission of Wuchereria bancrofti lymphatic filariasis in Ghana. Parasit. Vectors 2012, 5, 89. [Google Scholar] [CrossRef] [Green Version]
  23. Joseph, H.; Moloney, J.; Maiava, F.; McClintock, S.; Lammie, P.; Melrose, W. First evidence of spatial clustering of lymphatic filariasis in an Aedes polynesiensis endemic area. Acta Trop. 2011, 120, S39–S47. [Google Scholar] [CrossRef] [PubMed]
  24. Southgate, B.A.; Bryan, J.H. Factors affecting transmission of Wuchereria bancrofti by anopheline mosquitoes. 4. Facilitation, limitation, proportionality and their epidemiological significance. Trans. R Soc. Trop. Med. Hyg. 1992, 86, 523–530. [Google Scholar] [CrossRef]
  25. Wada, Y. Vector mosquitoes of filariasis in Japan. Trop. Med. Health 2011, 39, 39–45. [Google Scholar] [PubMed]
  26. Lymphatic Filariasis. Available online: https://www.who.int/news-room/fact-sheets/detail/lymphatic-filariasis (accessed on 26 March 2020).
  27. Ramos-Castañeda, J.; Barreto Dos Santos, F.; Martínez-Vega, R.; Lio, J.; Galvão De Araujo, M.; Joint, G.; Sarti, E. Dengue in Latin America: Systematic Review of Molecular Epidemiological Trends. PLoS Negl. Trop. Dis. 2017, 11, e0005224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Dengue Worldwide Overview. Available online: https://www.ecdc.europa.eu/en/dengue-monthly (accessed on 26 March 2020).
  29. Katzelnick, L.C.; Coloma, J.; Harris, E. Dengue: Knowledge gaps, unmet needs, and research priorities. Lancet Infect. Dis. 2017, 17, e88–e100. [Google Scholar] [CrossRef] [Green Version]
  30. Villamil-Gómez, W.E.; Rodríguez-Morales, A.J.; Uribe-García, A.M.; González-Arismendy, E.; Castellanos, J.E.; Calvo, E.P.; Álvarez-Mon, M.; Musso, D. Zika, dengue, and chikungunya co-infection in a pregnant woman from Colombia. Int. J. Infect. Dis. 2016, 51, 135–138. [Google Scholar] [CrossRef] [Green Version]
  31. Giron, S.; Franke, F.; Decoppet, A.; Cadiou, B.; Travaglini, T.; Thirion, L.; Durand, G.; Jeannin, C.; L’Ambert, G.; Grard, G.; et al. Vector-borne transmission of Zika virus in Europe, southern France, August 2019. Euro Surveill. 2019, 24. Available online: https://doi.org/10.2807/1560-7917.ES.2019.24.45.1900655 (accessed on 20 April 2020). [CrossRef] [Green Version]
  32. Ruchusatsawat, K.; Wongjaroen, P.; Posanacharoen, A.; Rodriguez-Barraquer, I.; Sangkitporn, S.; Cummings, D.A.; Salje, H. Long-term circulation of Zika virus in Thailand: An observational study. Lancet Infect. Dis. 2019, 19, 439–446. [Google Scholar] [CrossRef] [Green Version]
  33. Brady, O.J.; Hay, S.I. The first local cases of Zika virus in Europe. Lancet 2019, 394, 1991–1992. [Google Scholar] [CrossRef] [Green Version]
  34. Ledermann, J.P.; Guillaumot, L.; Yug, L.; Saweyog, S.C.; Tided, M.; Machieng, P.; Pretrick, M.; Marfel, M.; Griggs, A.; Bel, M.; et al. Aedes hensilli as a potential vector of Chikungunya and Zika viruses. PLoS Negl. Trop. Dis. 2014, 8, e3188. [Google Scholar] [CrossRef]
  35. Diallo, D.; Sall, A.A.; Diagne, C.T.; Faye, O.; Faye, O.; Ba, Y.; Hanley, K.A.; Buenemann, M.; Weaver, S.C.; Diallo, M. Zika virus emergence in mosquitoes in southeastern Senegal, 2011. PLoS ONE 2014, 9, e109442. [Google Scholar] [CrossRef] [Green Version]
  36. Grard, G.; Caron, M.; Mombo, I.M.; Nkoghe, D.; Mboui Ondo, S.; Jiolle, D.; Fontenille, D.; Paupy, C.; Leroy, E.M. Zika virus in Gabon (Central Africa)—2007: A new threat from Aedes albopictus? PLoS Negl. Trop. Dis. 2014, 8, e2681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Song, B.H.; Yun, S.I.; Woolley, M.; Lee, Y.M. Zika virus: History, epidemiology, transmission, and clinical presentation. J. Neuroimmunol. 2017, 308, 50–64. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Musso, D.; Nilles, E.J.; Cao-Lormeau, V.-M. Rapid spread of emerging Zika virus in the Pacific area. Clin. Microbiol. Infect. 2014, 20, O595–O596. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Edington, F.; Varjão, D.; Melo, P. Incidence of articular pain and arthritis after chikungunya fever in the Americas: A systematic review of the literature and meta-analysis. Jt. Bone Spine 2018, 85, 669–678. [Google Scholar] [CrossRef]
  40. Spoto, S.; Riva, E.; Fogolari, M.; Cella, E.; Costantino, S.; Angeletti, S.; Ciccozzi, M. Diffuse maculopapular rash: A family cluster during the last Chikungunya virus epidemic in Italy. Clin. Case Rep. 2018, 6, 2322–2325. [Google Scholar] [CrossRef] [Green Version]
  41. Rahman, M.M.; Jakaria, S.K.; Sayed, B.; Moniruzzaman, M.; Humayon Kabir, A.K.M.; Mallik, M.U.; Hasan, M.R.; Siddique, A.B.; Hossain, M.A.; Uddin, N.; et al. Clinical and Laboratory Characteristics of an Acute Chikungunya Outbreak in Bangladesh in 2017. Am. J. Trop. Med. Hyg. 2018, 100, 405–410. [Google Scholar] [CrossRef]
  42. Da Silva Junior, G.B.; Pinto, J.R.; Mota, R.M.S.; da Pires Neto, R.J.; Daher, E.D.F. Risk factors for death among patients with Chikungunya virus infection during the outbreak in northeast Brazil, 2016–2017. Trans. R. Soc. Trop. Med. Hyg. 2018, 113, 221–226. [Google Scholar] [CrossRef]
  43. De Azevedo Fernandes, N.C.C.; Cunha, M.S.; Guerra, J.M.; Réssio, R.A.; Cirqueira, C.D.S.; Iglezias, S.D.; de Carvalho, J.; Araujo, E.L.L.; Catão-Dias, J.L.; Díaz-Delgado, J. Outbreak of Yellow Fever among Nonhuman Primates, Espirito Santo, Brazil, 2017. Emerg. Infect. Dis. 2017, 23, 2038–2041. [Google Scholar] [CrossRef] [Green Version]
  44. Possas, C.; Lourenço-de-Oliveira, R.; Tauil, P.L.; de Pinheiro, F.P.; Pissinatti, A.; da Cunha, R.V.; Freire, M.; Martins, R.M.; Homma, A. Yellow fever outbreak in Brazil: The puzzle of rapid viral spread and challenges for immunisation. Mem. Inst. Oswaldo Cruz 2018, 113, e180278. [Google Scholar] [CrossRef] [Green Version]
  45. Simon, L.V.; Hashmi, M.F.; Torp, K.D. Yellow Fever; StatPearls: Tampa/St. Petersburg, FL, USA, 2018. [Google Scholar]
  46. Nwachukwu, W.E.; Yusuff, H.; Nwangwu, U.; Okon, A.; Ogunniyi, A.; Imuetinyan-Clement, J.; Besong, M.; Ayo-Ajayi, P.; Nikau, J.; Baba, A.; et al. The response to re-emergence of yellow fever in Nigeria, 2017. Int. J. Infect. Dis. 2020, 92, 189–196. [Google Scholar] [CrossRef] [Green Version]
  47. WHO. Yellow Fever—Nigeria; WHO: Geneva, Switzerland, 2019. [Google Scholar]
  48. Silva, N.I.O.; Sacchetto, L.; De Rezende, I.M.; Trindade, G.D.S.; Labeaud, A.D.; De Thoisy, B.; Drumond, B.P. Recent sylvatic yellow fever virus transmission in Brazil: The news from an old disease. Virol. J. 2020, 17, 9. [Google Scholar] [CrossRef] [Green Version]
  49. Auguste, A.J.; Lemey, P.; Bergren, N.A.; Giambalvo, D.; Moncada, M.; Morón, D.; Hernandez, R.; Navarro, J.-C.; Weaver, S.C. Enzootic transmission of yellow fever virus, Venezuela. Emerg. Infect. Dis. 2015, 21, 99–102. [Google Scholar] [CrossRef] [Green Version]
  50. Selemane, I. Epidemiological monitoring of the last outbreak of yellow fever in Brazil—An outlook from Portugal. Travel Med. Infect. Dis. 2019, 28, 46–51. [Google Scholar] [CrossRef]
  51. Solomon, T.; Hombach, J.; Jacobson, J.; Hoke, C.; Marfin, A.; Campbell, G.; Ginsburg, A.; Fischer, M.; Tsai, T.; Hills, S.; et al. Estimated Global Incidence of Japanese Encephalitis: A Systematic Review. Bull. World Health Organ. 2011, 89, 766–774. [Google Scholar]
  52. Fang, Y.; Zhang, Y.; Zhou, Z.B.; Xia, S.; Shi, W.Q.; Xue, J.B.; Li, Y.Y.; Wu, J.T. New strains of Japanese encephalitis virus circulating in Shanghai, China after a ten-year hiatus in local mosquito surveillance. Parasites Vectors 2019, 12, 22. [Google Scholar] [CrossRef] [PubMed]
  53. Griesemer, S.B.; Kramer, L.D.; Van Slyke, G.A.; Pata, J.D.; Gohara, D.W.; Cameron, C.E.; Ciota, A.T. Mutagen resistance and mutation restriction of St. Louis encephalitis virus. J. Gen. Virol. 2017, 98, 201–211. [Google Scholar] [CrossRef] [PubMed]
  54. Diaz, A.; Coffey, L.L.; Burkett-Cadena, N.; Day, J.F. Reemergence of St. Louis encephalitis virus in the Americas. Emerg. Infect. Dis. 2018, 24, 2150–2157. [Google Scholar] [CrossRef] [Green Version]
  55. Castillo-Olivares, J.; Wood, J. West Nile virus infection of horses. Vet. Res. 2004, 35, 467–483. [Google Scholar] [CrossRef] [Green Version]
  56. Barrett, A.D.T. West Nile in Europe: An increasing public health problem. J. Travel Med. 2018, 25, tay096. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. López-Ruiz, N.; del Montaño-Remacha, M.C.; Durán-Pla, E.; Pérez-Ruiz, M.; Navarro-Marí, J.M.; Salamanca-Rivera, C.; Miranda, B.; Oyonarte-Gómez, S.; Ruiz-Fernández, J. West nile virus outbreak in humans and epidemiological surveillance, West Andalusia, Spain, 2016. Eurosurveillance 2018, 23, 17-00261. [Google Scholar] [CrossRef] [Green Version]
  58. Zhang, J.; Lu, G.; Li, J.; Kelly, P.; Li, M.; Wang, J.; Huang, K.; Qiu, H.; You, J.; Zhang, R.; et al. Molecular Detection of Rickettsia felis and Rickettsia bellii in Mosquitoes. Vector Borne Zoonotic Dis. 2019, 19, 802–809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Guo, W.P.; Tian, J.H.; Lin, X.D.; Ni, X.B.; Chen, X.P.; Liao, Y.; Yang, S.Y.; Dumler, J.S.; Holmes, E.C.; Zhang, Y.Z. Extensive genetic diversity of Rickettsiales bacteria in multiple mosquito species. Sci. Rep. 2016, 6, 38770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Krajacich, B.J.; Huestis, D.L.; Dao, A.; Yaro, A.S.; Diallo, M.; Krishna, A.; Xu, J.; Lehmann, T. Investigation of the seasonal microbiome of Anopheles coluzzii mosquitoes in Mali. PLoS ONE 2018, 13, e0194899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Socolovschi, C.; Pagés, F.; Raoult, D. Rickettsia felis in aedes albopictus mosquitoes, libreville, gabon. Emerg. Infect. Dis. 2012, 18, 1688–1689. [Google Scholar] [CrossRef] [PubMed]
  62. Dieme, C.; Bechah, Y.; Socolovschi, C.; Audoly, G.; Berenger, J.M.; Faye, O.; Raoult, D.; Parola, P. Transmission potential of rickettsia felis infection by Anopheles gambiae mosquitoes. Proc. Natl. Acad. Sci. USA 2015, 112, 8088–8093. [Google Scholar] [CrossRef] [Green Version]
  63. Eliasson, H.; Broman, T.; Forsman, M.; Bäck, E. Tularemia: Current Epidemiology and Disease Management. Infect. Dis. Clin. North Am. 2006, 20, 289–311. [Google Scholar] [CrossRef]
  64. San Martín, J.L.; Brathwaite, O.; Zambrano, B.; Solórzano, J.O.; Bouckenooghe, A.; Dayan, G.H.; Guzmán, M.G. The epidemiology of dengue in the americas over the last three decades: A worrisome reality. Am. J. Trop. Med. Hyg. 2010, 82, 128–135. [Google Scholar] [CrossRef]
  65. Bouwman, H.; van den Berg, H.; Kylin, H. DDT and Malaria Prevention: Addressing the Paradox. Environ. Health Perspect. 2011, 119, 744–747. [Google Scholar] [CrossRef] [Green Version]
  66. Cailly, P.; Tran, A.; Balenghien, T.; L’Ambert, G.; Toty, C.; Ezanno, P. A climate-driven abundance model to assess mosquito control strategies. Ecol. Model. 2012, 227, 7–17. [Google Scholar] [CrossRef]
  67. Hemingway, J.; Beaty, B.J.; Rowland, M.; Scott, T.W.; Sharp, B.L. The Innovative Vector Control Consortium: Improved control of mosquito-borne diseases. Trends Parasitol. 2006, 22, 308–312. [Google Scholar] [CrossRef]
  68. Oxborough, R.M.; Kitau, J.; Jones, R.; Mosha, F.W.; Rowland, M.W. Experimental hut and bioassay evaluation of the residual activity of a polymer-enhanced suspension concentrate (SC-PE) formulation of deltamethrin for IRS use in the control of Anopheles arabiensis. Parasites Vectors 2014, 7, 454. [Google Scholar] [CrossRef] [PubMed]
  69. Dzul-Manzanilla, F.; Ibarra-López, J.; Marín, W.B.; Martini-Jaimes, A.; Leyva, J.T.; Correa-Morales, F.; Huerta, H.; Manrique-Saide, P.; Vazquez-Prokopec, G.M.; Day, J. Indoor resting behavior of Aedes aegypti (Diptera: Culicidae) in Acapulco, Mexico. J. Med. Entomol. 2018, 54, 501–504. [Google Scholar]
  70. Hladish, T.J.; Pearson, C.A.B.; Rojas, D.P.; Gomez-Dantes, H.; Halloran, M.E.; Vazquez-Prokopec, G.M.; Longini, I.M. Forecasting the effectiveness of indoor residual spraying for reducing dengue burden. PLoS Negl. Trop. Dis. 2018, 12, e0006570. [Google Scholar] [CrossRef] [PubMed]
  71. Giglioli, G. An Investigation of the House-Frequenting Habits of Mosquitoes of the British Guiana Coastland in Relation to the Use of DDT 1. Am. J. Trop. Med. Hyg. 1948, 1, 43–70. [Google Scholar] [CrossRef]
  72. Nathan, M.B.; Giglioli, M.E. Eradication of Aedes aegypti on Cayman Brac and Little Cayman, West Indies, with Abate (Temephos) in 1970–1971. Bull. Pan Am. Health Organ. 1982, 16, 28–39. [Google Scholar]
  73. Oxborough, R.M.; Kitau, J.; Jones, R.; Feston, E.; Matowo, J.; Mosha, F.W.; Rowland, M.W. Long-lasting control of Anopheles arabiensis by a single spray application of micro-encapsulated pirimiphos-methyl (Actellic® 300 CS). Malar. J. 2014, 13, 37. [Google Scholar] [CrossRef] [Green Version]
  74. WHO. Pesticides and Their Application: For the Control of Vectors and Pests of Public Health Importance, 6th ed.; WHO: Geneva, Switzerland, 2006; Available online:https://apps.who.int/iris/handle/10665/69223 (accessed on 20 April 2020).
  75. Vazquez-Prokopec, G.M.; Medina-Barreiro, A.; Che-Mendoza, A.; Dzul-Manzanilla, F.; Correa-Morales, F.; Guillermo-May, G.; Bibiano-Marín, W.; Uc-Puc, V.; Geded-Moreno, E.; Vadillo-Sánchez, J.; et al. Deltamethrin resistance in Aedes aegypti results in treatment failure in Merida, Mexico. PLoS Negl. Trop. Dis. 2017, 11, e0005656. [Google Scholar] [CrossRef] [Green Version]
  76. Uragayala, S.; Kamaraju, R.; Tiwari, S.N.; Sreedharan, S.; Ghosh, S.K.; Valecha, N. Village-scale (Phase III) evaluation of the efficacy and residual activity of SumiShield ® 50 WG (Clothianidin 50%, w/w) for indoor spraying for the control of pyrethroid-resistant Anopheles culicifacies Giles in Karnataka state, India. Trop. Med. Int. Heal. 2018, 23, 605–615. [Google Scholar] [CrossRef]
  77. Zaim, M.; Guillet, P. Alternative insecticides: An urgent need. Trends Parasitol. 2002, 18, 161–163. [Google Scholar] [CrossRef]
  78. Esu, E.; Lenhart, A.; Smith, L.; Horstick, O. Effectiveness of peridomestic space spraying with insecticide on dengue transmission; Systematic review. Trop. Med. Int. Heal. 2010, 15, 619–631. [Google Scholar] [CrossRef]
  79. Peterson, R.K.D.; Macedo, P.A.; Davis, R.S. A human-health risk assessment for West Nile virus and insecticides used in mosquito management. Environ. Health Perspect. 2006, 114, 366–372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Bonds, J.A.S. Ultra-low-volume space sprays in mosquito control: A critical review. Med. Vet. Entomol. 2012, 26, 121–130. [Google Scholar] [CrossRef] [PubMed]
  81. Boyce, W.M.; Lawler, S.P.; Schultz, J.M.; McCauley, S.J.; Kimsey, L.S.; Niemela, M.K.; Nielsen, C.F.; Reisen, W.K. Nontarget effects of the mosquito adulticide pyrethrin applied aerially during a West Nile virus outbreak in an urban California environment. J. Am. Mosq. Control Assoc. 2007, 23, 335–339. [Google Scholar] [CrossRef] [Green Version]
  82. Nathan, M.; Reiter, P.; WHO. Guidelines for Assessing the Efficacy of Insecticidal Space Sprays for Control of the Dengue Vector: Aedes aegypti; WHO: Geneva, Switzerland, 2001; Available online: https://apps.who.int/iris/handle/10665/67047 (accessed on 20 April 2020).
  83. Teng, H.-J.; Chen, T.-J.; Tsai, S.-F.; Lin, C.-P.; Chiou, H.-Y.; Lin, M.-C.; Yang, S.-Y.; Lee, Y.-W.; Kang, C.-C.; Hsu, H.-C.; et al. Emergency vector control in a DENV-2 outbreak in 2002 in Pingtung City, Pingtung County, Taiwan. Jpn. J. Infect. Dis. 2007, 60, 271–279. [Google Scholar] [PubMed]
  84. Guillet, P.; Alnwick, D.; Cham, M.K.; Neira, M.; Zaim, M.; Heymann, D.; Mukelabai, K. Long-lasting treated mosquito nets: A breakthrough in malaria prevention. Bull. World Health Organ. 2001, 79, 998. [Google Scholar] [PubMed]
  85. Yang, G.G.; Kim, D.; Pham, A.; Paul, C.J. A meta-regression analysis of the effectiveness of mosquito nets for malaria control: The value of long-lasting insecticide nets. Int. J. Environ. Res. Public Health 2018, 15, 1–12. [Google Scholar] [CrossRef] [Green Version]
  86. Rowland, M.; Webster, J.; Saleh, P.; Chandramohan, D.; Freeman, T.; Pearcy, B.; Durrani, N.; Rab, A.; Mohammed, N. Prevention of malaria in Afghanistan through social marketing of insecticide-treated nets: Evaluation of coverage and effectiveness by cross-sectional surveys and passive surveillance. Trop. Med. Int. Heal. 2002, 7, 813–822. [Google Scholar] [CrossRef] [Green Version]
  87. Arroz, J.A.H.; Candrinho, B.; Mendis, C.; Varela, P.; Pinto, J.; Do, M.; Martins, R.O. Effectiveness of a new long-lasting insecticidal nets delivery model in two rural districts of Mozambique: A before-after study. Malar J. 2018, 17, 66. [Google Scholar] [CrossRef] [Green Version]
  88. Girond, F.; Madec, Y.; Kesteman, T.; Randrianarivelojosia, M.; Randremanana, R.; Randriamampionona, L.; Randrianasolo, L.; Ratsitorahina, M.; Herbreteau, V.; Hedje, J.; et al. Evaluating effectiveness of mass and continuous long-lasting insecticidal net distributions over time in Madagascar: A sentinel surveillance based epidemiological study. EClinicalMedicine 2018, 1, 62–69. [Google Scholar] [CrossRef] [Green Version]
  89. Hounkonnou, C.; Djènontin, A.; Egbinola, S.; Houngbegnon, P.; Bouraima, A.; Soares, C.; Fievet, N.; Accrombessi, M.; Yovo, E.; Briand, V.; et al. Impact of the use and efficacy of long lasting insecticidal net on malaria infection during the first trimester of pregnancy—A pre-conceptional cohort study in southern Benin. BMC Public Health 2018, 18, 683. [Google Scholar] [CrossRef] [Green Version]
  90. Trape, J.F.; Tall, A.; Diagne, N.; Ndiath, O.; Ly, A.B.; Faye, J.; Dieye-Ba, F.; Roucher, C.; Bouganali, C.; Badiane, A.; et al. Malaria morbidity and pyrethroid resistance after the introduction of insecticide-treated bednets and artemisinin-based combination therapies: A longitudinal study. Lancet Infect. Dis. 2011, 11, 925–932. [Google Scholar] [CrossRef]
  91. Sougoufara, S.; Thiaw, O.; Cailleau, A.; Diagne, N.; Harry, M.; Bouganali, C.; Sembène, P.M.; Doucoure, S.; Sokhna, C. The Impact of Periodic Distribution Campaigns of Long-Lasting Insecticidal-Treated Bed Nets on Malaria Vector Dynamics and Human Exposure in Dielmo, Senegal. Am. J. Trop. Med. Hyg. 2018, 98, 1343–1352. [Google Scholar] [CrossRef] [Green Version]
  92. Soleimani-Ahmadi, M.; Vatandoost, H.; Shaeghi, M.; Raeisi, A.; Abedi, F.; Eshraghian, M.R.; Madani, A.; Safari, R.; Oshaghi, M.A.; Abtahi, M.; et al. Field evaluation of permethrin long-lasting insecticide treated nets (Olyset®) for malaria control in an endemic area, southeast of Iran. Acta Trop. 2012, 123, 146–153. [Google Scholar] [CrossRef]
  93. Kajla, M.K.; Barrett-Wilt, G.A.; Paskewitz, S.M. Bacteria: A novel source for potent mosquito feeding-deterrents. Sci. Adv. 2019, 5, eaau6141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Nerio, L.S.; Olivero-Verbel, J.; Stashenko, E. Repellent activity of essential oils: A review. Bioresour. Technol. 2010, 101, 372–378. [Google Scholar] [CrossRef]
  95. Trongtokit, Y.; Rongsriyam, Y.; Komalamisra, N.; Apiwathnasorn, C. Comparative repellency of 38 essential oils against mosquito bites. Phyther. Res. 2005, 19, 303–309. [Google Scholar] [CrossRef] [PubMed]
  96. Daisy, B.H.; Strobel, G.A.; Castillo, U.; Ezra, D.; Sears, J.; Weaver, D.K.; Runyon, J.B. Naphthalene, an insect repellent, is produced by Muscodor vitigenus, a novel endophytic fungus. Microbiology 2002, 148, 3737–3741. [Google Scholar] [CrossRef] [Green Version]
  97. Naseem, S.; Munir, T.; Faheem Malik, M. Mosquito management: A review. J. Entomol. Zool. Stud. 2016, 4, 73–79. [Google Scholar]
  98. Balaji, A.P.B.; Ashu, A.; Manigandan, S.; Sastry, T.P.; Mukherjee, A.; Chandrasekaran, N. Polymeric nanoencapsulation of insect repellent: Evaluation of its bioefficacy on Culex quinquefasciatus mosquito population and effective impregnation onto cotton fabrics for insect repellent clothing. J. King Saud Univ. Sci. 2017, 29, 517–527. [Google Scholar] [CrossRef]
  99. Soni, N.; Prakash, S. Green nanoparticles for mosquito control. Sci. World J. 2014, 2014, 496362. [Google Scholar] [CrossRef] [Green Version]
  100. Knipling, E.F. Possibilities of Insect Control or Eradication Through the Use of Sexually Sterile Males1. J. Econ. Entomol. 1955, 48, 459–462. [Google Scholar] [CrossRef]
  101. Phuc, H.K.; Andreasen, M.H.; Burton, R.S.; Vass, C.; Epton, M.J.; Pape, G.; Fu, G.; Condon, K.C.; Scaife, S.; Donnelly, C.A.; et al. Late-acting dominant lethal genetic systems and mosquito control. BMC Biol. 2007, 5, 11. [Google Scholar] [CrossRef] [Green Version]
  102. Reiter, P. Oviposition, Dispersal, and Survival in Aedes aegypti: Implications for the Efficacy of Control Strategies. Vector Borne Zoonotic Dis. 2007, 7, 261–273. [Google Scholar] [CrossRef]
  103. Onyekwere, J.; Nnamonu, E.; Bede, E.; Okoye, C.; Okoro, J.; Nnamonu, E.; Bede, E.; Okoye, I.C.; Onyekwere, J.; Nnamonu, E.; et al. Application of genetically modified mosquitoes (Anopheles species) in the control of malaria transmission. Asian J. Biotechnol. Genet. Eng. 2018, 1, 1–16. [Google Scholar]
  104. Meghani, Z.; Boëte, C. Genetically engineered mosquitoes, Zika and other arboviruses, community engagement, costs, and patents: Ethical issues. PLoS Negl. Trop. Dis. 2018, 12, e0006501. [Google Scholar] [CrossRef]
  105. GMWATCH. New Documents Show Oxitec’s GM Mosquitoes Ineffective and Risky. Available online: https://www.gmwatch.org/en/news/latest-news/17828-new-documents-show-oxitec-s-gm-mosquitoes-ineffective-and-risky (accessed on 6 April 2019).
  106. Yamada, H.; Kraupa, C.; Lienhard, C.; Parker, A.G.; Maiga, H.; de Oliveira Carvalho, D.; Zheng, M.; Wallner, T.; Bouyer, J. Mosquito mass rearing: Who’s eating the eggs? Parasite 2019, 26, 75. [Google Scholar] [CrossRef] [PubMed]
  107. Mukherjee, S.; Blaustein, L. Effects of predator type and alternative prey on mosquito egg raft predation and destruction. Hydrobiologia 2019, 846, 215–221. [Google Scholar] [CrossRef]
  108. Lopez, S.B.G.; Guimarães-Ribeiro, V.; Rodriguez, J.V.G.; Dorand, F.A.P.S.; Salles, T.S.; Sá-Guimarães, T.E.; Alvarenga, E.S.L.; Melo, A.C.A.; Almeida, R.V.; Moreira, M.F. RNAi-based bioinsecticide for Aedes mosquito control. Sci. Rep. 2019, 9, 1–13. [Google Scholar] [CrossRef]
  109. Lee, J.Y.; Woo, R.M.; Choi, C.J.; Shin, T.Y.; Gwak, W.S.; Woo, S.D. Beauveria bassiana for the simultaneous control of Aedes albopictus and Culex pipiens mosquito adults shows high conidia persistence and productivity. AMB Express 2019, 9, 206. [Google Scholar] [CrossRef] [Green Version]
  110. Scholte, E.-J.; Knols, B.G.J.; Samson, R.A.; Takken, W. Entomopathogenic fungi for mosquito control: A review. J. Insect Sci. 2004, 4, 19. [Google Scholar] [CrossRef]
  111. Noskov, Y.A.; Polenogova, O.V.; Yaroslavtseva, O.N.; Belevich, O.E.; Yurchenko, Y.A.; Chertkova, E.A.; Kryukova, N.A.; Kryukov, V.Y.; Glupov, V.V. Combined effect of the entomopathogenic fungus Metarhizium robertsii and avermectins on the survival and immune response of Aedes aegypti larvae. PeerJ 2019, 7, e7931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Lovett, B.; Bilgo, E.; Diabate, A.; St Leger, R. A review of progress toward field application of transgenic mosquitocidal entomopathogenic fungi. Pest Manag. Sci. 2019, 75, 2316–2324. [Google Scholar] [CrossRef] [PubMed]
  113. Silva, L.E.I.; Paula, A.R.; Ribeiro, A.; Butt, T.M.; Silva, C.P.; Samuels, R.I. A new method of deploying entomopathogenic fungi to control adult Aedes aegypti mosquitoes. J. Appl. Entomol. 2018, 142, 59–66. [Google Scholar] [CrossRef] [Green Version]
  114. Louca, V.; Lucas, M.C.; Green, C.; Majambere, S.; Fillinger, U.; Lindsay, S.W. Role of fish as predators of mosquito larvae on the floodplain of the Gambia River. J. Med. Entomol. 2009, 46, 546–556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Aditya, G.; Pal, S.; Saha, N.; Saha, G. Efficacy of indigenous larvivorous fishes against Culex quinquefasciatus in the presence of alternative prey: Implications for biological control. J. Vector Borne Dis. 2012, 49, 217–225. [Google Scholar]
  116. Sareein, N.; Phalaraksh, C.; Rahong, P.; Techakijvej, C.; Seok, S.; Bae, Y.J. Relationships between predatory aquatic insects and mosquito larvae in residential areas in northern Thailand. J. Vector Ecol. 2019, 44, 223–232. [Google Scholar] [CrossRef] [Green Version]
  117. Früh, L.; Kampen, H.; Schaub, G.A.; Werner, D. Predation on the invasive mosquito Aedes japonicus (Diptera: Culicidae) by native copepod species in Germany. J. Vector Ecol. 2019, 44, 241–247. [Google Scholar] [CrossRef] [Green Version]
  118. Cuthbert, R.N.; Callaghan, A.; Sentis, A.; Dalal, A.; Dick, J.T.A. Additive multiple predator effects can reduce mosquito populations. Ecol. Entomol. 2019, 45, 243–250. [Google Scholar] [CrossRef]
  119. Digma, J.R.; Sumalde, A.C.; Salibay, C.C. Laboratory evaluation of predation of Toxorhynchites amboinensis (Diptera:Culicidae) on three mosquito vectors of arboviruses in the Philippines. Biol. Control 2019, 137, 104009. [Google Scholar] [CrossRef]
  120. Focks, D.A. Toxorhynchites as biocontrol agents. J. Am. Mosq. Control Assoc. 2007, 23, 118–127. [Google Scholar] [CrossRef]
  121. Schiller, A.; Allen, M.; Coffey, J.; Fike, A.; Carballo, F. Updated Methods for the Production of Toxorhynchites rutilus septentrionalis (Diptera, Culicidae) for Use as Biocontrol Agent Against Container Breeding Pest Mosquitoes in Harris County, Texas. J. Insect Sci. 2019, 19, 8. [Google Scholar] [CrossRef] [PubMed]
  122. Das, B.P. Chilodonella uncinate—A protozoa pathogenic to mosquito larvae. Curr. Sci. 2003, 85, 483–489. [Google Scholar]
  123. Dhanasekaran, D.; Thangaraj, R. Microbial secondary metabolites are an alternative approaches against insect vector to prevent zoonotic diseases. Asian Pac. J. Trop. Dis. 2014, 4, 253–261. [Google Scholar] [CrossRef]
  124. Lacey, L.A. Bacillus thuringiensis serovariety israelensis and Bacillus sphaericus for mosquito control. J. Am. Mosq. Control Assoc. 2007, 23, 93–109. [Google Scholar] [CrossRef]
  125. Derua, Y.A.; Kahindi, S.C.; Mosha, F.W.; Kweka, E.J.; Atieli, H.E.; Wang, X.; Zhou, G.; Lee, M.; Githeko, A.K.; Yan, G. Microbial larvicides for mosquito control: Impact of long lasting formulations of Bacillus thuringiensis var. israelensis and Bacillus sphaericus on non-target organisms in western Kenya highlands. Ecol. Evol. 2018, 8, 7563–7573. [Google Scholar] [CrossRef] [Green Version]
  126. Zhang, Q.; Hua, G.; Adang, M.J. Effects and mechanisms of Bacillus thuringiensis crystal toxins for mosquito larvae. Insect Sci. 2017, 24, 714–729. [Google Scholar] [CrossRef]
  127. Palma, L.; Muñoz, D.; Berry, C.; Murillo, J.; Caballero, P. Bacillus thuringiensis toxins: An overview of their biocidal activity. Toxins 2014, 6, 3296–3325. [Google Scholar] [CrossRef] [Green Version]
  128. Chee Dhang, C.; Han Lim, L.; Wasi Ahmad, N.; Benjamin, S.; Koon Weng, L.; Abdul Rahim, D.; Syafinaz Safian, E.; Sofian-Azirun, M. Field effectiveness of Bacillus thuringiensis israelensis (Bti) against Aedes (Stegomyia) aegypti (Linnaeus) in ornamental ceramic containers with common aquatic plants. Trop. Biomed. 2009, 26, 100–105. [Google Scholar]
  129. Saliha, B.; Wafa, H.; Laid, O.M. Effect of Bacillus thuringiensis var krustaki on the mortality and development of Culex pipiens (Diptera; Cullicidae). Int. J. Mosq. Res. 2017, 4, 20–23. [Google Scholar]
  130. Delécluse, A.; Rosso, M.L.; Ragni, A. Cloning and expression of a novel toxin gene from Bacillus thuringiensis subsp. jegathesan encoding a highly mosquitocidal protein. Appl. Environ. Microbiol. 1995, 61, 4230–4235. [Google Scholar] [CrossRef] [Green Version]
  131. López-Meza, J.; Federici, B.A.; Poehner, W.J.; Martinez-Castillo, A.; Ibarra, J.E. Highly mosquitocidal isolates of Bacillus thuringiensis subspecies kenyae and entomocidus from Mexico. Biochem. Syst. Ecol. 1995, 23, 461–468. [Google Scholar] [CrossRef]
  132. De Barjac, H.; Sebald, M.; Charles, J.F.; Cheong, W.H.; Lee, H.L. Clostridium bifermentans serovar malaysia, a new anaerobic bacterium pathogen to mosquito and blackfly larvae. C. R. Acad. Sci. Iii. 1990, 310, 383–387. [Google Scholar] [PubMed]
  133. Darriet, F.; Hougard, J.-M. An isolate of Bacillus circulans toxic to mosquito larvae. J. Am. Mosq. Control Assoc. 2002, 18, 65–67. [Google Scholar] [PubMed]
  134. Favret, M.E.; Yousten, A.A. Insecticidal activity of Bacillus laterosporus. J. Invertebr. Pathol. 1985, 45, 195–203. [Google Scholar] [CrossRef]
  135. Orlova, M.V.; Smirnova, T.A.; Ganushkina, L.A.; Yacubovich, V.Y.; Azizbekyan, R.R. Insecticidal activity of Bacillus laterosporus. Appl. Environ. Microbiol. 1998, 64, 2723–2725. [Google Scholar] [CrossRef] [Green Version]
  136. Pener, M.P. An Overview of Insect Growth Disruptors; Applied Aspects. Adv. Insect Phys. 2012, 43, 1–162. [Google Scholar]
  137. Yapabandara, A.M.G.M.; Curtis, C.F. Laboratory and field comparisons of pyriproxyfen, polystyrene beads and other larvicidal methods against malaria vectors in Sri Lanka. Acta Trop. 2002, 81, 211–223. [Google Scholar] [CrossRef]
  138. Ansari, M.A.; Razdan, R.K.; Sreehari, U. Laboratory and field evaluation of Hilmilin against mosquitoes. J. Am. Mosq. Control Assoc. 2005, 21, 432–436. [Google Scholar] [CrossRef]
  139. Raghavendra, K.; Barik, T.K.; Reddy, B.P.N.; Sharma, P.; Dash, A.P. Malaria vector control: From past to future. Parasitol. Res. 2011, 108, 757–779. [Google Scholar] [CrossRef]
  140. Paul, A.; Harrington, L.C.; Zhang, L.; Scott, J.G. Insecticide resistance in Culex pipiens from New York. J. Am. Mosq. Control Assoc. 2005, 21, 305–309. [Google Scholar] [CrossRef]
  141. De Silva, J.J.; Mendes, J. Susceptibility of Aedes aegypti (L) to the insect growth regulators diflubenzuron and methoprene in Uberlândia, State of Minas Gerais. Rev. Soc. Bras. Med. Trop. 2007, 40, 612–616. [Google Scholar] [CrossRef] [Green Version]
  142. Dennehy, T.J.; Degain, B.A.; Harpold, V.S.; Zaborac, M.; Morin, S.; Fabrick, J.A.; Nichols, R.L.; Brown, J.K.; Byrne, F.J.; Li, X. Extraordinary resistance to insecticides reveals exotic Q biotype of Bemisia tabaci in the New World. J. Econ. Entomol. 2010, 103, 2174–2186. [Google Scholar] [CrossRef] [PubMed]
  143. Lee, S.-H.; Young CHOI, J.; Ram LEE, B.; Fang, Y.; Hoon KIM, J.; Hwan Park, D.; Gu Park, M.; Mi Woo, R.; Jin Kim, W.; Ho, Y.J.; et al. Insect growth regulatory and larvicidal activity of chalcones against Aedes albopictus. Entomol. Rep. 2018, 48, 55–59. [Google Scholar] [CrossRef] [Green Version]
  144. Niang, E.H.A.; Bassene, H.; Fenollar, F.; Mediannikov, O. Biological Control of Mosquito-Borne Diseases: The Potential of Wolbachia -Based Interventions in an IVM Framework. J. Trop. Med. 2018, 2018, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Benelli, G.; Jeffries, C.L.; Walker, T. Biological Control of Mosquito Vectors: Past, Present, and Future. Insects 2016, 7, 52. [Google Scholar] [CrossRef]
  146. Zug, R.; Hammerstein, P. Still a Host of Hosts for Wolbachia: Analysis of Recent Data Suggests That 40% of Terrestrial Arthropod Species Are Infected. PLoS ONE 2012, 7, 38544. [Google Scholar] [CrossRef] [Green Version]
  147. Skerman, V.B.D.; McGowan, V.F.; Sneath, P.H.A.; Peter, H.A. Approved Lists of Bacterial Names; Skerman, V.B.D., McGowan, V.F., Sneath, P.H.A., Eds.; American Society for Microbiology: Washington, DC, USA, 1989; ISBN 9781555810146. [Google Scholar]
  148. Ogunbiyi, T.S.; Eromon, P.; Oluniyi, P.; Ayoade, F.; Oloche, O.; Oguzie, J.U.; Folarin, O.; Happi, C.; Komolafe, I. First Report of Wolbachia from Field Populations of Culex Mosquitoes in South-Western Nigeria. Afr. Zool. 2019, 54, 181–185. [Google Scholar] [CrossRef]
  149. Balaji, S.; Jayachandran, S.; Prabagaran, S.R. Evidence for the natural occurrence of Wolbachia in Aedes aegypti mosquitoes. Fems Microbiol. Lett. 2019, 366. Available online: https://doi.org/10.1093/femsle/fnz055 (accessed on 20 April 2020). [CrossRef]
  150. Mohanty, I.; Rath, A.; Swain, S.P.; Pradhan, N.; Hazra, R.K. Wolbachia Population in Vectors and Non-vectors: A Sustainable Approach Towards Dengue Control. Curr. Microbiol. 2019, 76, 133–143. [Google Scholar] [CrossRef]
  151. Laven, H. Eradication of Culex pipiens fatigans through Cytoplasmic Incompatibility. Nature 1967, 216, 383–384. [Google Scholar] [CrossRef]
  152. Zhang, D.; Zheng, X.; Xi, Z.; Bourtzis, K.; Gilles, J.R.L. Combining the Sterile Insect Technique with the Incompatible Insect Technique: I-Impact of Wolbachia Infection on the Fitness of Triple- and Double-Infected Strains of Aedes albopictus. PLoS ONE 2015, 10, e0121126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. O’Connor, L.; Plichart, C.; Sang, A.C.; Brelsfoard, C.L.; Bossin, H.C.; Dobson, S.L. Open Release of Male Mosquitoes Infected with a Wolbachia Biopesticide: Field Performance and Infection Containment. PLoS Negl. Trop. Dis. 2012, 6, e1797. [Google Scholar] [CrossRef]
  154. Joubert, D.A.; Walker, T.; Carrington, L.B.; De Bruyne, J.T.; Kien, D.H.T.; Hoang, N.L.T.; Chau, N.V.V.; Iturbe-Ormaetxe, I.; Simmons, C.P.; O’Neill, S.L. Establishment of a Wolbachia Superinfection in Aedes aegypti Mosquitoes as a Potential Approach for Future Resistance Management. PLoS Pathog. 2016, 12, e1005434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Aliota, M.T.; Walker, E.C.; Yepes, A.U.; Velez, I.D.; Christensen, B.M.; Osorio, J.E. The wMel Strain of Wolbachia Reduces Transmission of Chikungunya Virus in Aedes aegypti. PLoS Negl. Trop. Dis. 2016, 10, e0004677. [Google Scholar] [CrossRef] [Green Version]
  156. Van den Hurk, A.F.; Hall-Mendelin, S.; Pyke, A.T.; Frentiu, F.D.; McElroy, K.; Day, A.; Higgs, S.; O’Neill, S.L. Impact of Wolbachia on infection with chikungunya and yellow fever viruses in the mosquito vector Aedes aegypti. PLoS Negl. Trop. Dis. 2012, 6, e1892. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Bian, G.; Joshi, D.; Dong, Y.; Lu, P.; Zhou, G.; Pan, X.; Xu, Y.; Dimopoulos, G.; Xi, Z. Wolbachia Invades Anopheles stephensi Populations and Induces Refractoriness to Plasmodium Infection. Science 2013, 340, 748–751. [Google Scholar] [CrossRef]
  158. Hughes, G.L.; Koga, R.; Xue, P.; Fukatsu, T.; Rasgon, J.L. Wolbachia Infections Are Virulent and Inhibit the Human Malaria Parasite Plasmodium Falciparum in Anopheles Gambiae. PLoS Pathog. 2011, 7, e1002043. [Google Scholar] [CrossRef] [Green Version]
  159. Ford, S.A.; Allen, S.L.; Ohm, J.R.; Sigle, L.T.; Sebastian, A.; Albert, I.; Chenoweth, S.F.; McGraw, E.A. Selection on Aedes aegypti alters Wolbachia-mediated dengue virus blocking and fitness. Nat. Microbiol. 2019, 4, 1832–1839. [Google Scholar] [CrossRef]
  160. Dutra, H.L.C.; Rocha, M.N.; Dias, F.B.S.; Mansur, S.B.; Caragata, E.P.; Moreira, L.A. Wolbachia Blocks Currently Circulating Zika Virus Isolates in Brazilian Aedes aegypti Mosquitoes. Cell Host Microbe 2016, 19, 744–771. [Google Scholar] [CrossRef] [Green Version]
  161. Mariño, Y.; Verle Rodrigues, J.; Bayman, P. Wolbachia Affects Reproduction and Population Dynamics of the Coffee Berry Borer (Hypothenemus hampei): Implications for Biological Control. Insects 2017, 8, 8. [Google Scholar] [CrossRef] [Green Version]
  162. Chegeni, T.N.; Fakhar, M. Promising Role of Wolbachia as Anti-parasitic Drug Target and Eco-Friendly Biocontrol Agent. Recent Pat. Antiinfect. Drug Discov. 2019, 14, 69–79. [Google Scholar] [CrossRef] [PubMed]
  163. Bongio, N.J.; Lampe, D.J. Inhibition of Plasmodium berghei Development in Mosquitoes by Effector Proteins Secreted from Asaia sp. Bacteria Using a Novel Native Secretion Signal. PLoS ONE 2015, 10. Available online: https://doi.org/10.1371/journal.pone.0143541 (accessed on 20 April 2020). [CrossRef] [PubMed] [Green Version]
  164. Wang, S.; Jacobs-Lorena, M. Genetic approaches to interfere with malaria transmission by vector mosquitoes. Trends Biotechnol. 2013, 31, 185–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Rami, A.; Raz, A.; Zakeri, S.; Dinparast Djadid, N. Isolation and identification of Asaia sp. in Anopheles spp. mosquitoes collected from Iranian malaria settings: Steps toward applying paratransgenic tools against malaria. Parasit Vectors 2018, 11, 367. [Google Scholar] [CrossRef] [PubMed]
  166. Favia, G.; Ricci, I.; Marzorati, M.; Negri, I.; Alma, A.; Sacchi, L.; Bandi, C.; Daffonchio, D. Bacteria of the Genus Asaia: A Potential Paratransgenic Weapon Against Malaria. In Transgenesis and the Management of Vector-Borne Disease; Springer: New York, NY, USA, 2008; pp. 49–59. [Google Scholar]
  167. Favia, G.; Ricci, I.; Damiani, C.; Raddadi, N.; Crotti, E.; Marzorati, M.; Rizzi, A.; Urso, R.; Brusetti, L.; Borin, S.; et al. Bacteria of the genus Asaia stably associate with Anopheles stephensi, an Asian malarial mosquito vector. Proc. Natl. Acad. Sci. USA 2007, 104, 9047–9051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Damiani, C.; Ricci, I.; Crotti, E.; Rossi, P.; Rizzi, A.; Scuppa, P.; Esposito, F.; Bandi, C.; Daffonchio, D.; Favia, G. Paternal transmission of symbiotic bacteria in malaria vectors. Curr. Biol. 2008, 18, R1087–R1088. [Google Scholar] [CrossRef] [Green Version]
  169. Nugapola, N.W.N.P.; De Silva, W.A.P.P.; Karunaratne, S.H.P.P. Distribution and phylogeny of Wolbachia strains in wild mosquito populations in Sri Lanka. Parasites Vectors 2017, 10, 230. [Google Scholar] [CrossRef] [Green Version]
  170. Mamlouk, D.; Gullo, M. Acetic Acid Bacteria: Physiology and Carbon Sources Oxidation. Indian J. Microbiol. 2013, 53, 377. [Google Scholar] [CrossRef] [Green Version]
  171. Cappelli, A.; Damiani, C.; Mancini, M.V.; Valzano, M.; Rossi, P.; Serrao, A.; Ricci, I.; Favia, G. Asaia Activates Immune Genes in Mosquito Eliciting an Anti-Plasmodium Response: Implications in Malaria Control. Front. Genet. 2019, 10, 836. [Google Scholar] [CrossRef] [Green Version]
  172. Mitraka, E.; Stathopoulos, S.; Siden-Kiamos, I.; Christophides, G.K.; Louis, C. Asaia accelerates larval development of Anopheles gambiae. Pathog. Glob. Health 2013, 107, 305. [Google Scholar] [CrossRef] [Green Version]
  173. Shane, J.L.; Grogan, C.L.; Cwalina, C.; Lampe, D.J. Blood meal-induced inhibition of vector-borne disease by transgenic microbiota. Nat. Commun. 2018, 9, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Mertz, F.P.; YAO, R.C. Saccharopolyspora spinosa sp. nov. Isolated from Soil Collected in a Sugar Mill Rum Still. Int. J. Syst. Bacteriol. 1990, 40, 34–39. [Google Scholar] [CrossRef]
  175. Hertlein, M.B.; Mavrotas, C.; Jousseaume, C.; Lysandrou, M.; Thompson, G.D.; Jany, W.; Ritchie, S.A. A Review of Spinosad as a Natural Product for Larval Mosquito Control. J. Am. Mosq. Control Assoc. 2010, 26, 67–87. [Google Scholar] [CrossRef]
  176. Su, T. Resistance and Its Management to Microbial and Insect Growth Regulator Larvicides in Mosquitoes; InTech Europe: Rijeka, Croatia, 2016; pp. 135–154. [Google Scholar] [CrossRef] [Green Version]
  177. WHO. Spinosad DT in Drinking-Water: Use for Vector Control in Drinking-water Sources and Containers; WHO: Geneva, Switzerland, 2010; Available online: https://www.who.int/water_sanitation_health/dwq/chemicals/spinosadbg.pdf (accessed on 20 April 2020).
  178. Marina, C.F.; Bond, J.G.; Muñoz, J.; Valle, J.; Chirino, N.; Williams, T. Spinosad: A biorational mosquito larvicide for use in car tires in southern Mexico. Parasit. Vectors 2012, 5, 95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Marina, C.F.; Bond, J.; Muñoz, J.; Valle, J.; Novelo-Gutiérrez, R.; Williams, T. Efficacy and non-target impact of spinosad, Bti and temephos larvicides for control of Anopheles spp. in an endemic malaria region of southern Mexico. Parasit. Vectors 2014, 7, 55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. Prabhu, K.; Murugan, K.; Nareshkumar, A.; Bragadeeswaran, S. Larvicidal and pupicidal activity of spinosad against the malarial vector Anopheles stephensi. Asian Pac. J. Trop. Med. 2011, 4, 610–613. [Google Scholar] [CrossRef] [Green Version]
  181. Qiu, H.; Jun, H.W.; Mccall, J.W. Pharmacokinetics, formulation, and safety of insect repellent N,N-Diethyl-3-methylbenzamide (deet): A review. J. Am. Mosq. Control Assoc. 1998, 14, 12–27. [Google Scholar]
  182. Stanczyk, N.M.; Brookfield, J.F.Y.; Field, L.M.; Logan, J.G. Aedes aegypti Mosquitoes Exhibit Decreased Repellency by DEET following Previous Exposure. PLoS ONE 2013, 8, e54438. [Google Scholar] [CrossRef]
  183. Deletre, E.; Martin, T.; Duménil, C.; Chandre, F. Insecticide resistance modifies mosquito response to DEET and natural repellents. Parasit. Vectors 2019, 12, 89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Debboun, M.; Frances, S.P.; Strickman, D. Insect Repellents Handbook; CRC Press: Boca Raton, FL, USA, 2014; ISBN 9781466553552. [Google Scholar]
  185. Van Roey, K.; Sokny, M.; Denis, L.; Van den Broeck, N.; Heng, S.; Siv, S.; Sluydts, V.; Sochantha, T.; Coosemans, M.; Durnez, L. Field evaluation of picaridin repellents reveals differences in repellent sensitivity between Southeast Asian vectors of malaria and arboviruses. PLoS Negl. Trop. Dis. 2014, 8, e3326. [Google Scholar] [CrossRef]
  186. Carroll, S.P. Prolonged efficacy of IR3535 repellents against mosquitoes and blacklegged ticks in North America. J. Med. Entomol. 2008, 45, 706–714. [Google Scholar] [CrossRef] [PubMed]
  187. Lee, M.Y. Essential Oils as Repellents against Arthropods. Biomed Res. Int. 2018, 2018, 6860271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Tilquin, M.; Paris, M.; Reynaud, S.; Despres, L.; Ravanel, P.; Geremia, R.A.; Gury, J. Long lasting persistence of Bacillus thuringiensis Subsp. israelensis (Bti) in mosquito natural habitats. PLoS ONE 2008, 3, e3432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Paris, M.; Tetreau, G.; Laurent, F.; Lelu, M.; Despres, L.; David, J.-P. Persistence of Bacillus thuringiensis israelensis (Bti) in the environment induces resistance to multiple Bti toxins in mosquitoes. Pest Manag. Sci. 2011, 67, 122–128. [Google Scholar] [CrossRef] [PubMed]
  190. Bravo, A.; Soberón, M. How to cope with insect resistance to Bt toxins? Trends Biotechnol. 2008, 26, 573–579. [Google Scholar] [CrossRef]
  191. Georghiou, G.P.; Wirth, M.C. Influence of Exposure to Single versus Multiple Toxins of Bacillus thuringiensis subsp. israelensis on Development of Resistance in the Mosquito Culex quinquefasciatus (Diptera: Culicidae). Appl. Environ. Microbiol. 1997, 63, 1095–1101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Cadavid-Restrepo, G.; Sahaza, J.; Orduz, S. Treatment of an Aedes aegypti colony with the Cry11Aa toxin for 54 generations results in the development of resistance. Mem. Inst. Oswaldo Cruz 2012, 107, 74–79. [Google Scholar] [CrossRef] [Green Version]
  193. Stalinski, R.; Tetreau, G.; Gaude, T.; Després, L. Pre-selecting resistance against individual Bti Cry toxins facilitates the development of resistance to the Bti toxins cocktail. J. Invertebr. Pathol. 2014, 119, 50–53. [Google Scholar] [CrossRef]
  194. Paris, M.; Melodelima, C.; Coissac, E.; Tetreau, G.; Reynaud, S.; David, J.-P.; Despres, L. Transcription profiling of resistance to Bti toxins in the mosquito Aedes aegypti using next-generation sequencing. J. Invertebr. Pathol. 2012, 109, 201–208. [Google Scholar] [CrossRef]
  195. Cheong, H.; Dhesi, R.K.; Gill, S.S. Marginal cross-resistance to mosquitocidal Bacillus thuringiensis strains in Cry11A-resistant larvae: Presence of Cry11A-like toxins in these strains. Fems Microbiol. Lett. 2006, 153, 419–424. [Google Scholar] [CrossRef]
  196. Demissew, A.; Balkew, M.; Girma, M. Larvicidal activities of chinaberry, neem and Bacillus thuringiensis israelensis (Bti) to an insecticide resistant population of Anopheles arabiensis from Tolay, Southwest Ethiopia. Asian Pac. J. Trop. Biomed. 2016, 6, 554–561. [Google Scholar] [CrossRef] [Green Version]
  197. Boyer, S.; Tilquin, M.; Ravanel, P. differential sensitivity to Bacillus thuringiensis var. israelensis and temephos in field mosquito populations of Ochlerotatus cataphylla (Diptera: Culicidae): Toward resistance? Environ. Toxicol. Chem. 2007, 26, 157. [Google Scholar] [CrossRef] [PubMed]
  198. Paris, M.; Boyer, S.; Bonin, A.; Collado, A.; David, J.-P.; Despres, L. Genome scan in the mosquito Aedes rusticus: Population structure and detection of positive selection after insecticide treatment. Mol. Ecol. 2010, 19, 325–337. [Google Scholar] [CrossRef] [PubMed]
  199. Su, T.; Thieme, J.; White, G.S.; Lura, T.; Mayerle, N.; Faraji, A.; Cheng, M.L.; Brown, M.Q. High Resistance to Bacillus sphaericus and Susceptibility to Other Common Pesticides in Culex pipiens (Diptera: Culicidae) from Salt Lake City, UT. J. Med. Entomol. 2019, 56, 506–513. [Google Scholar] [CrossRef]
  200. Park, H.-W.; Bideshi, D.K.; Federici, B.A. Properties and applied use of the mosquitocidal bacterium, Bacillus sphaericus. J. Asia. Pac. Entomol. 2010, 13, 159–168. [Google Scholar] [CrossRef] [Green Version]
  201. Ahmed, I.; Yokota, A.; Yamazoe, A.; Fujiwara, T. Proposal of Lysinibacillus boronitolerans gen. nov. sp. nov., and transfer of Bacillus fusiformis to Lysinibacillus fusiformis comb. nov. and Bacillus sphaericus to Lysinibacillus sphaericus comb. nov. Int. J. Syst. Evol. Microbiol. 2007, 57, 1117–1125. [Google Scholar] [CrossRef] [Green Version]
  202. Nielsen-Leroux, C.; Pasquier, F.; Charles, J.F.; Sinègre, G.; Gaven, B.; Pasteur, N. Resistance to Bacillus sphaericus involves different mechanisms in Culex pipiens (Diptera:Culicidae) larvae. J. Med. Entomol. 1997, 34, 321–327. [Google Scholar] [CrossRef]
  203. Rao, D.R.; Mani, T.R.; Rajendran, R.; Joseph, A.S.; Gajanana, A.; Reuben, R. Development of a high level of resistance to Bacillus sphaericus in a field population of Culex quinquefasciatus from Kochi, India. J. Am. Mosq. Control Assoc. 1995, 11, 1–5. [Google Scholar]
  204. Rodcharoen, J.; Mulla, M.S. Biological Fitness of Culex quinquefasciatus (Diptera: Culicidae) Susceptible and Resistant to Bacillus sphaericus. J. Med. Entomol. 1997, 34, 5–10. [Google Scholar] [CrossRef]
  205. De Oliveira, C.M.F.; Filho, F.C.; Beltràn, J.E.N.; Silva-Filha, M.H.; Regis, L. Biological fitness of a Culex quinquefasciatus population and its resistance to Bacillus sphaericus. J. Am. Mosq. Control Assoc. 2003, 19, 125–129. [Google Scholar]
  206. Amorim, L.B.; de Barros, R.A.; de Melo Chalegre, K.D.; de Oliveira, C.M.F.; Narcisa Regis, L.; Silva-Filha, M.H.N.L. Stability of Culex quinquefasciatus resistance to Bacillus sphaericus evaluated by molecular tools. Insect Biochem. Mol. Biol. 2010, 40, 311–316. [Google Scholar] [CrossRef] [PubMed]
  207. Rojas, J.E.; Mazzarri, M.; Sojo, M.; García, A.G.Y. Effectiveness of Bacillus sphaericus strain 2362 on larvae of Anopheles nuñeztovari. Investig. Clin. 2001, 42, 131–146. [Google Scholar] [PubMed]
  208. Nicolas, L.; Darriet, F.; Hougard, J.M. Efficacy of Bacillus sphaericus 2362 against larvae of Anopheles gambiae under laboratory and field conditions in West Africa. Med. Vet. Entomol. 1987, 1, 157–162. [Google Scholar] [CrossRef] [PubMed]
  209. Derua, Y.A.; Kahindi, S.C.; Mosha, F.W.; Kweka, E.J.; Atieli, H.E.; Zhou, G.; Lee, M.-C.; Githeko, A.K.; Yan, G. Susceptibility of Anopheles gambiae complex mosquitoes to microbial larvicides in diverse ecological settings in western Kenya. Med. Vet. Entomol. 2019, 33, 220–227. [Google Scholar] [CrossRef]
  210. Skovmand, O.; Bauduin, S. Efficacy of a granular formulation of Bacillus sphaericus against Culex quinquefasciatus and Anopheles gambiae in West African countries. J. Vector Ecol. 1997, 22, 43–51. [Google Scholar]
  211. Poopathi, S.; Mani, T.R.; Rao, D.R.; Kabilan, L. Evaluation of Synergistic Interaction between Bacillus sphaericus and a Neem-based Biopesticide on Bsph-Susceptible Culex quinquefasciatus Say Larvae. Int. J. Trop. Insect Sci. 2011, 22, 303–306. [Google Scholar] [CrossRef]
  212. Raoult, D.; Abat, C. Developing new insecticides to prevent chaos: The real future threat. Lancet Infect. Dis. 2017, 17, 804–805. [Google Scholar] [CrossRef] [Green Version]
  213. IVCC Annual Report 2017–2018. 2018. Available online: http://www.ivcc.com/about/governance/annual-reports (accessed on 20 April 2020).
  214. Sergeant, M.; Baxter, L.; Jarrett, P.; Shaw, E.; Ousley, M.; Winstanley, C.; Morgan, J.A.W. Identification, typing, and insecticidal activity of Xenorhabdus isolates from entomopathogenic nematodes in United Kingdom soil and characterization of the xpt toxin loci. Appl. Environ. Microbiol. 2006, 72, 5895–5907. [Google Scholar] [CrossRef] [Green Version]
  215. Pineda-Castellanos, M.L.; Rodríguez-Segura, Z.; Villalobos, F.J.; Hernández, L.; Lina, L.; Nuñez-Valdez, M.E. Pathogenicity of Isolates of Serratia Marcescens towards Larvae of the Scarab Phyllophaga Blanchardi (Coleoptera). Pathogens 2015, 4, 210–228. [Google Scholar] [CrossRef] [Green Version]
  216. Niu, H.; Wang, N.; Liu, B.; Xiao, L.; Wang, L.; Guo, H. Synergistic and additive interactions of Serratia marcescens S-JS1 to the chemical insecticides for controlling Nilaparvata lugens (Hemiptera: Delphacidae). J. Econ. Entomol. 2018, 111, 823–828. [Google Scholar] [CrossRef]
  217. Wei, G.; Lai, Y.; Wang, G.; Chen, H.; Li, F.; Wang, S. Insect pathogenic fungus interacts with the gut microbiota to accelerate mosquito mortality. Proc. Natl. Acad. Sci. USA 2017, 114, 5994–5999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Islam, M.T.; Mubarak, M.S. Diterpenes and their derivatives as promising agents against dengue virus and dengue vectors: A literature-based review. Phyther. Res. 2019. Available online: https://doi.org/10.1002/ptr.6562 (accessed on 20 April 2020).
  219. Itoh, T.; Kawada, H.; Abe, A.; Eshita, Y.; Rongsriyam, Y.; Igarashi, A. Utilization of bloodfed females of Aedes aegypti as a vehicle for the transfer of the insect growth regulator pyriproxyfen to larval habitats. J. Am. Mosq. Control Assoc. 1994, 10, 344–347. [Google Scholar] [PubMed]
  220. Revay, E.E.; Müller, 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] [Green Version]
  221. Naranjo, D.P.; Qualls, W.A.; Müller, G.C.; Samson, D.M.; Roque, D.; Alimi, T.; Arheart, K.; Beier, J.C.; Xue, R.-D. Evaluation of boric acid sugar baits against Aedes albopictus (Diptera: Culicidae) in tropical environments. Parasitol. Res. 2013, 112, 1583–1587. [Google Scholar] [CrossRef]
  222. Müller, 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, 9, 210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Allan, S.A. Susceptibility of adult mosquitoes to insecticides in aqueous sucrose baits. J. Vector Ecol. 2011, 36, 59–67. [Google Scholar] [CrossRef]
  224. George, O. Poinar Nematodes for Biological Control of Insects; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar] [CrossRef]
  225. Petersen, J.J. Role of mermithid nematodes in biological control of mosquitoes. Exp. Parasitol. 1973, 33, 239–247. [Google Scholar] [CrossRef]
  226. Abagli, A.Z.; Alavo, T.B.; Platzer, E.G. Efficacy of the insect parasitic nematode, Romanomermis iyengari, for malaria vector control in Benin West Africa. Malar. J. 2012, 11, P5. [Google Scholar] [CrossRef] [Green Version]
  227. Paily, K.P.; Balaraman, K. Susceptibility of ten species of mosquito larvae to the parasitic nematode Romanomermis iyengari and its development. Med. Vet. Entomol. 2000, 14, 426–429. [Google Scholar] [CrossRef]
  228. Abagli, A.Z.; Alavo, T.B.C.; Perez-Pacheco, R.; Platzer, E.G. Efficacy of the mermithid nematode, Romanomermis iyengari, for the biocontrol of Anopheles gambiae, the major malaria vector in sub-Saharan Africa. Parasit. Vectors 2019, 12, 253. [Google Scholar] [CrossRef]
  229. Abagli, A.Z.; Alavo. Biocontrol of Culex quinquefasciatus using the insect parasitic nematode, Romanomermis iyengari (Nematoda: Mermithidae). Trop. Biomed. 2019, 36, 1003–1013. Available online: http://msptm.org/files/Vol36No4/1003-1013-Alavo-TBC.pdf (accessed on 20 April 2020).
  230. Britch, S.C.; Nyberg, H.; Aldridge, R.L.; Swan, T.; Linthicum, K.J. Acoustic Control of Mosquito Larvae In Artificial Drinking Water Containers. J. Am. Mosq. Control Assoc. 2016, 32, 341–344. [Google Scholar] [CrossRef]
  231. Fredregill, C.L.; Motl, G.C.; Dennett, J.A.; Bueno, R.; Debboun, M. Efficacy of Two LarvasonicTM Units Against Culex Larvae and Effects on Common Aquatic Nontarget Organisms in Harris County, Texas 1. J. Am. Mosq. Control Assoc. 2015, 31, 366–370. [Google Scholar] [CrossRef]
  232. Mukundarajan, H.; Hol, F.J.H.; Castillo, E.A.; Newby, C.; Prakash, M. Using mobile phones as acoustic sensors for high-throughput mosquito surveillance. Elife 2017, 6, e27854. [Google Scholar] [CrossRef] [PubMed]
  233. Johnson, B.J.; Rohde, B.B.; Zeak, N.; Staunton, K.M.; Prachar, T.; Ritchie, S.A. A low-cost, battery-powered acoustic trap for surveilling male Aedes aegypti during rear-and-release operations. PLoS ONE 2018, 13, e0201709. [Google Scholar] [CrossRef] [PubMed]
  234. Whyard, S.; Erdelyan, C.N.G.; 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. Parasit. Vectors 2015, 8, 96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Jesus, T.; Wanner, E.; Cardoso, R. A receding horizon control approach for integrated vector management of Aedes aegypti using chemical and biological control: A mono and a multiobjective approach. Math. Methods Appl. Sci. 2019. Available online: https://doi.org/10.1002/mma.6115 (accessed on 20 April 2020).
Figure 1. Cartography of significant resurgences of mosquito-borne diseases worldwide (until September 2019). We listed all reported outbreaks and imported and autochthon cases of malaria, dengue fever, yellow fever, chikungunya fever, and Zika fever between 2017 and 2019. This figure clearly shows their resurgence in almost all tropical countries. In many cases they were imported to several northern countries where the competent vector has become established, which may lead to potential local transmission.
Figure 1. Cartography of significant resurgences of mosquito-borne diseases worldwide (until September 2019). We listed all reported outbreaks and imported and autochthon cases of malaria, dengue fever, yellow fever, chikungunya fever, and Zika fever between 2017 and 2019. This figure clearly shows their resurgence in almost all tropical countries. In many cases they were imported to several northern countries where the competent vector has become established, which may lead to potential local transmission.
Pathogens 09 00310 g001
Figure 2. Bacillus thuringiensis israelensis (Bti), 3 days of culture, 4.9 µm in length (Hitachi TM4000) (personal image).
Figure 2. Bacillus thuringiensis israelensis (Bti), 3 days of culture, 4.9 µm in length (Hitachi TM4000) (personal image).
Pathogens 09 00310 g002
Figure 3. Lysinibacillus sphaericus (Bs) strain CSURP827, 3 days of culture, 5.02 µm in length (Hitachi TM4000) (personal image).
Figure 3. Lysinibacillus sphaericus (Bs) strain CSURP827, 3 days of culture, 5.02 µm in length (Hitachi TM4000) (personal image).
Pathogens 09 00310 g003
Figure 4. Aedes albopictus strain: (A) larvae (personal images), (B) pupa (personal images), and (C) adult (personal images).
Figure 4. Aedes albopictus strain: (A) larvae (personal images), (B) pupa (personal images), and (C) adult (personal images).
Pathogens 09 00310 g004
Figure 5. Anopheles gambiae: (A) larvae (personal images), (B) pupa (personal images), and (C) adult (personal images).
Figure 5. Anopheles gambiae: (A) larvae (personal images), (B) pupa (personal images), and (C) adult (personal images).
Pathogens 09 00310 g005
Figure 6. Culex pipiens: (A) larvae (personal images), (B) pupa (CC BY-SA 4.0_Andrei Savitsky), and (C) adult (personal images).
Figure 6. Culex pipiens: (A) larvae (personal images), (B) pupa (CC BY-SA 4.0_Andrei Savitsky), and (C) adult (personal images).
Pathogens 09 00310 g006
Table 1. Highlights of field and laboratory insecticide resistances to Bti and Bs.
Table 1. Highlights of field and laboratory insecticide resistances to Bti and Bs.
BacteriaMosquitoSiteType of StudyNumber of Studied RegionsDateReference
Bti + BsCulex pipiens–complexOnondaga County, USAField2June 2003[140]
BtiOchlerotatuscataphyllaRhône-Alpes, FranceField4April 2003[197]
BtiAedes rusticusRhône-Alpes, FranceField13Winters 2005 and 2006[198]
BtiCulex quinquefasciatusUSALaboratory1Summer 1990[191]
BtiAedes aegyptiUSALaboratory12011[192]
BsCulex pipiens–complexUtah, USAField3September 2016[199]
BtiAedes aegyptiFranceLaboratory12010[189]

Share and Cite

MDPI and ACS Style

Dahmana, H.; Mediannikov, O. Mosquito-Borne Diseases Emergence/Resurgence and How to Effectively Control It Biologically. Pathogens 2020, 9, 310. https://doi.org/10.3390/pathogens9040310

AMA Style

Dahmana H, Mediannikov O. Mosquito-Borne Diseases Emergence/Resurgence and How to Effectively Control It Biologically. Pathogens. 2020; 9(4):310. https://doi.org/10.3390/pathogens9040310

Chicago/Turabian Style

Dahmana, Handi, and Oleg Mediannikov. 2020. "Mosquito-Borne Diseases Emergence/Resurgence and How to Effectively Control It Biologically" Pathogens 9, no. 4: 310. https://doi.org/10.3390/pathogens9040310

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