Next Article in Journal
Five Silkworm 30K Proteins Are Involved in the Cellular Immunity against Fungi
Next Article in Special Issue
Salivary Digestion Extends the Range of Sugar-Aversions in the German Cockroach
Previous Article in Journal
Subfamily Anischiinae (Coleoptera: Eucnemidae) in Early Cretaceous of Northeast China
Previous Article in Special Issue
Alarm Pheromone Responses Depend on Genotype, but Not on the Presence of Facultative Endosymbionts in the Pea Aphid Acyrthosiphon pisum
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advancements in Studies on Chemosensory Mechanisms Underlying Detection of Semiochemicals in Dacini Fruit Flies of Economic Importance (Diptera: Tephritidae)

1
Graduate School of Agriculture, Kyoto University, Kyoto 606-8502, Japan
2
Department of Biology, Faculty of Science, Universiti Putra Malaysia, 43400 UPM Serdang, Selangor Darul Ehsan, Malaysia
3
Key Laboratory of Entomology and Pest Control Engineering, College of Plant Protection, Southwest University, Chongqing 400716, China
4
Academy of Agricultural Sciences, Southwest University, Chongqing 400716, China
*
Author to whom correspondence should be addressed.
Insects 2021, 12(2), 106; https://doi.org/10.3390/insects12020106
Submission received: 31 December 2020 / Revised: 22 January 2021 / Accepted: 23 January 2021 / Published: 26 January 2021
(This article belongs to the Special Issue Cross Talking between Insects and Environment)

Abstract

:

Simple Summary

Chemical information, among various environmental cues, is the most important factor for insect survival and reproduction. Insects usually rely on pheromone signals, including sex, aggregation, and alarm pheromones, in conspecific communications. For phytophagous insects, plant semiochemicals associated with insect–plant interactions profoundly affect insect behavior and physiology. Therefore, elucidation of the mechanisms underlying the perception of various chemical signals is essential to understand the adaptation of insects to their surrounding environment. Dacini fruit flies are interesting models to investigate chemosensory mechanisms because they utilize various chemical cues, such as floral fragrances in mutualistic relationships, volatiles released from host fruits, and sex pheromones in mating behaviors. Since many Dacini species are serious pests, novel insights into chemosensory mechanisms not only contribute to understanding basic principles of chemoreception, but also provide cues for the development of more effective agents for pest control. Herein, we review current knowledge on chemosensory mechanisms and related topics in Dacini fruit flies from various aspects of chemical ecology, physiology, neuroscience, and molecular biology. We also discuss future perspectives based on recent advancements in these studies.

Abstract

Dacini fruit flies mainly contain two genera, Bactrocera and Zeugodacus, and include many important pests of fruits and vegetables. Their life cycle is affected by various environmental cues. Among them, multiple characteristic semiochemicals have remarkable effects on their reproductive and host-finding behaviors. Notably, floral fragrances released from so-called fruit fly orchids strongly attract males of several Dacini fruit fly species. Focusing on the strong attraction of male flies to particular chemicals, natural and synthetic lures have been used for pest management. Thus, the perception of semiochemicals is important to understand environmental adaptation in Dacini fruit flies. Since next-generation sequencers are available, a large number of chemosensory-related genes have been identified in Dacini fruit flies, as well as other insects. Furthermore, recent studies have succeeded in the functional analyses of olfactory receptors in response to semiochemicals. Thus, characterization of molecular components required for chemoreception is under way. However, the mechanisms underlying chemoreception remain largely unknown. This paper reviews recent findings on peripheral mechanisms in the perception of odors in Dacini fruit flies, describing related studies in other dipteran species, mainly the model insect Drosophila melanogaster. Based on the review, important themes for future research have also been discussed.

1. Introduction

Tephritid fruit flies (Diptera: Tephritidae) contain many notorious pests of commercially important fruits and vegetables. Females lay their eggs directly into fruits, and subsequently hatched larvae feed and develop in the fruits. Therefore, the consequent damage causes heavy economic losses by rendering infested fruits inedible and prohibiting fruit exports due to strict quarantine restrictions. Among the tephritid fruit flies, the Tribe Dacini is one of the most species-rich clades with more than 700 species, including three crucial genera: Bactrocera, Zeugodacus, and Dacus [1]. Dacini fruit flies have been extensively studied in many aspects such as their ecology, behavior, and physiology. In general, phytophagous insects respond to environmental cues, including the olfactory and visual elements associated with their host plants [2]. Among various environmental information, plant semiochemicals play crucial roles in the life cycles of Dacini fruit flies. For example, male Oriental fruit flies, Bactrocera dorsalis, are strongly attracted to a specific phenylpropanoid, methyl eugenol (ME), which is an element of floral fragrance present in “fruit fly orchids” in the genus Bulbophyllum (Orchidaceae) [3]. ME is found in various plant organs of at least 450 species from 80 families [3]. Due to its strong attraction, ME has been widely used as a lure in pest control programs against B. dorsalis. It is noteworthy that the use of ME for male annihilation successfully led to the eradication of B. dorsalis from entire islands in the Marianas and Okinawa located in the North Pacific Ocean [4,5,6]. In addition to the aromatic floral component, volatiles derived from host fruits play a crucial role in the search for oviposition sites by gravid females of Dacini fruit flies. Electroantennogram (EAG) and gas chromatography-flame ionization detection coupled with electroantennographic detection (GC-EAD) studies have identified various volatiles released from host fruits that are detected by the antennae of B. dorsalis females [7,8,9,10,11]. The active compounds mainly contained terpenes and phenylpropanoids. Moreover, several attractive compounds have been identified through behavioral tests using olfactometers [7,8,9]. Although the perception of various semiochemicals is very important in the diverse life histories of Dacini fruit flies, the mechanisms by which chemoreception occurs have not been fully elucidated. In this review, we highlight recent advancements in the elucidation of chemosensory mechanisms of Dacini fruit flies.

2. Semiochemicals Involved in the Life Cycle of Dacini Fruit Flies

2.1. Male Attractants and Sex Pheromones

Both male attractants and sex pheromones have been extensively characterized in Dacini fruit flies (Figure 1). As there are excellent reviews [12,13,14], we briefly describe major discoveries in Dacini male attractants and sex pheromones. The existence of male lures such as ME was first discovered approximately a hundred years ago [15,16]. Since then, many male lures, e.g., isoeugenol analogs for Z. diversus, have been identified [16,17]. In particular, males of many Bactrocera and Zeugodacus species are strongly attracted to certain floral fragrances, either ME, or raspberry ketone (RK: a phenylbutanoid) [18]. Based on male attractiveness, Dacini flies can be classified into two groups: ME- and RK-sensitive species [19]. Following the attraction, males voraciously ingest the attractants [20,21]. Notably, the synthetic attractant, cue-lure, which is analogous to the acetyl derivative of RK, has been developed and used for pest management of RK-sensitive fruit fly species, particularly the melon fly Zeugodacus cucurbitae [22]. The robust attractiveness and feeding behavior has been thought to partially enhance mating success by the acquisition of male sex pheromones or precursors to attract conspecific females [13]. Furthermore, zingerone (a phenylbutanoid) has been characterized as an attractant for both ME- and RK-sensitive fruit fly species, likely owing to the hybrid chemical structure between ME and RK [23]. ME-sensitive fruit flies convert the ingested ME into other phenylpropanoids, such as (E)-coniferyl alcohol (E-CF) and 2-allyl-4,5-dimethoxyphenol (DMP), as male sex pheromones [24]. The converted compounds accumulate in the rectal gland, which is considered to be a secretory organ of male sex pheromones [25,26]. RK-sensitive fruit flies accumulate ingested RK without conversion into rectal glands [27,28]. Thus, the robust attractiveness of male attractants seems to be intimately associated with male sex pheromone production and accumulation.
Attraction of female to male sex pheromones has been well studied in the ME-sensitive fruit fly B. dorsalis. Females were found to be attracted not only to live conspecific males, but also to male rectal glands in laboratory assays [29]. Of the two sex pheromones, E-CF triggered more pheromone-mediated zigzag behavior in females than DMP, although mixing of DMP and E-CF enhanced this behavior in a wind tunnel bioassay [30,31]. Unlike the male attractant ME, the male sex pheromones released from the rectal glands as smoke could function at close range rather than long range for attraction [29,32]. Light intensity greatly affects the female response, as females are most sensitive to both sex pheromones under light intensity corresponding to that at dusk [33]. For the RK-sensitive fruit fly, Z. cucurbitae, females were attracted to live conspecific males, but did not respond to male rectal glands [29]. Females were attracted more to males fed with cue-lure or zingerone than to unfed males [33]. However, to the best of our knowledge, an obvious pheromonal activity of the RK-related phenylbutanoids has not been reported [13,14].
Unlike the production of sex pheromones by males of ME- and RK-sensitive fruit flies, females released sex pheromones to attract males in the olive fly B. oleae [34]. A major component of the sex pheromones has been identified as the spiroketal 1,7-dioxaspiro[5,5]undecane, more commonly known as olean [35,36]. There are the (R)- and (S)-enantiomers in olean, and female flies release it as a racemic mixture. Remarkably, males responded only to (R)-olean, which functions as a sex pheromone. In contrast, females responded only to (S)-olean in a short range, probably owing to its aphrodisiac action [37]. As various spiroketals are commonly present in the rectal glands of Dacini fruit flies [38], a process to acquire the sexually dimorphic response to the enantiomers in B. oleae is interesting in terms of its evolutionary and physiological aspects.
Since the success of effective pest management of RK-sensitive fruit flies using cue-lure [22], several lures have been developed as practical agents for monitoring pest fruit flies [13]. Perception of these attractants and male sex pheromones by the peripheral sensory organs, as well as the processing of this information in the central nervous system, are crucial themes, as described below.

2.2. Common Plant Volatiles Other Than Phenylpropanoids and Phenylbutanoids as Attractants

Common plant volatiles released from host or non-host plants are also important for the reproductive and host-finding behaviors of Dacini fruit flies [39]. Electrophysiological studies using EAG and GC-EAD have been used to identify multiple volatiles released from host fruits that are detected in the female antennae of several fruit flies. For example, multiple compounds, including aliphatic alcohols, aliphatic aldehydes, esters, terpenes, and phenolic compounds, have been identified as antennal responsive compounds of B. dorsalis from host fruits, such as tropical almond fruit (Terminalia catappa) and mango (Mangifera indica), by extensive electrophysiological studies [7,8,9,10,40]. Furthermore, some of the volatiles exhibited attractiveness of female flies in the olfactometer tests [8,9]. In addition, several volatiles from mango elicited oviposition behavior [10,41]. The Chinese citrus fruit fly B. minax is a univoltine and oligophagous species as its host range is restricted to citrus fruits. Female B. minax was attracted to four volatile components, namely linalool, nonanal, citral, and limonene, contained in citrus peels among the 13 compounds tested by Y-tube olfactometer tests. In contrast, furyl alcohol and guaiacol exhibited deterrent activities against females [42]. The attractiveness of analogous monoterpenes to females is an interesting aspect of chemoreception associated with host-finding and oviposition behaviors in citrus fruit flies.
Among the common plant volatiles, sesquiterpenoids, including β-caryophyllene and α-humulene, along with ME-derived phenylpropanoids, were sequestered in male rectal glands of the ME-sensitive guava fruit fly B. correcta [43]. A field trap test at a guava orchard in Thailand revealed that a mixture of β-caryophyllene and α-humulene attracted only a few males, in contrast to the significant attractiveness of ME [43]. In another field trap test conducted in Thailand, using β-caryophyllene resulted in the capture of seven times more B. correcta males than ME-baited traps [44]. While ME attracted only mature males, β-caryophyllene could also attract immature males, suggesting an effective lure candidate [44,45]. The attractiveness of female B. correcta and B. dorsalis to β-caryophyllene at high concentrations has also been reported using a Y-tube olfactometer [46].
Remarkably, α-ionone analogs (norsesquiterpenes) are attractive to B. latifrons males, although this species does not respond to typical aromatic attractants such as ME, RK, and cue-lure. First, α-ionol analogs present in essential oils have been identified as potential male attractants in laboratory behavioral bioassays [47]. Furthermore, 3-oxo-α-ionol analogs exhibited stronger attractiveness as well as phagostimulant activity to B. latifrons [48]. Interestingly, males of B. latifrons ingested 3-hydroxy-α-ionone, present in eggplant as a minor component, and subsequently accumulated it and the 3-oxidized derivatives in their rectal glands [49]. The accumulation reminds us the transformation of ingested ME into male sex pheromones in B. dorsalis. However, 3-oxo-α-ionol analogs have not been characterized as natural male attractants and sex pheromones. Based on this information, effective male attractants such as 3-oxo-7,8-dihydro-α-ionone modified from 3-oxo-α-ionol have been developed [50]. As B. latifrons and B. dorsalis are closely related species in phylogenetic analyses [1], the differential attraction of these species to ME or 3-oxo-α-ionone, respectively, is intriguing.

3. Detection of Male Attractants by the Peripheral Sensory Organs

The detection of male attractants in Dacini fruit flies is an important issue for understanding their chemosensory mechanisms. Flies are equipped with two peripheral olfactory organs: the antennae and maxillary palps [51]. In general, the antennae mainly detect pheromones and various semiochemicals as the primary olfactory organs. In contrast, the role of the maxillary palps has not been well characterized in many insects. Identifying the organ involved in the detection of male attractants is of much interest. Notably, recent studies have demonstrated that male attractants are detected not only by the antennae, but also by the maxillary palps, entailing different effects among fruit fly species [52,53,54]. For ME-sensitive B. dorsalis, ablation of male antennae remarkably reduced responsiveness to ME at a short range [52]. Furthermore, males lost the ability to detect ME at a long distance. In contrast, ablation of male maxillary palps did not cause a loss of ME-detection, although they took longer to detect ME. Therefore, the antennae and maxillary palps seem to play different roles in the detection of ME, depending on the distance from the lure source. Interestingly, the effects of ablation of the chemosensory organs in the Queensland fruit fly B. tryoni were different from those in B. dorsalis [53]. Males of B. tryoni are strongly attracted to RK and cue-lure. Behavioral experiments have shown that the preference for cue-lure in a short range was abolished by the ablation of maxillary palps, but not the antennae, which is different from the case in B. dorsalis. Furthermore, electrophysiological analyses in RK-sensitive fruit flies have provided valuable insights into the recognition of attractants by maxillary palps. The maxillary palps of B. tryoni responded strikingly to cue-lure, whereas only a weak response was observed in the antennae [53]. Likewise, the maxillary palps also exhibited remarkable electrophysiological responses to cue-lure as well as RK in another RK-sensitive striped fruit fly, Z. scutellata [54]. The maxillary palps of both males and females responded to cue-lure in B. tryoni and Z. scutellata, suggesting that the male-specific attraction is triggered by sexually dimorphic processing of chemosensory information in the central nervous system, rather than in the peripheral tissues, in these species. It should be noted that the association of the maxillary palps with gustatory stimulation has been reported in other dipteran species. Detection of pleasant odors via the maxillary palps enhanced feeding behaviors in D. melanogaster and the blow fly Phormia regina [55,56]. Furthermore, axonal projection of olfactory receptor neurons from the maxillary palps to the primary gustatory center, the subesophageal ganglion, was observed using an anterograde staining method [56]. Considering that the attraction to male lures triggers compulsive feeding behavior in Dacini fruit flies, this raises the possibility that the maxillary palps convey olfactory information to the primary gustatory center to induce the gustatory response.
It is possible that gustatory organs, such as the proboscis and tarsi, are involved in the detection of aromatic attractants. To confirm this, electrophysiological analyses, such as the tip-recording method, would be useful [57]. However, to the best of our knowledge, there is no report on the response of the gustatory sensilla to male attractants. Analysis of gustatory responses to male attractants at the level of sensilla will be necessary to understand the mechanisms underlying the feeding behavior triggered by olfactory perception in Dacini fruit flies.

4. Chemosensory-Related Behaviors Modulated by Physiological Status

The response thresholds of insects to external chemical and visual stimuli are modulated by their physiological status [58,59]. Therefore, an understanding of the daily rhythm, sexual maturation, mating, and feeding status of flies is important in undertaking Dacini chemosensory behavioral bioassays. For example, diurnal rhythm affects the attractiveness of Dacini males to lures. In Z. cucurbitae, males were more attracted to cue-lure in the morning as compared to that in the mid-afternoon, while the attractiveness of cue-lure was lowest in the afternoon [60]. Similar observations have been reported for two ME-sensitive species, B. opiliae and B. cacuminata [61,62]. In B. dorsalis males, laboratory bioassays revealed that the attraction to ME occurred throughout the day until the onset of mating behavior at dusk, whereas field studies using synthetic ME demonstrated that peak attraction of males to the lure occurred in the morning [63]. After the male attraction, ME is ingested, biotransformed into sex pheromonal components, and subsequently transported from the crop to the rectal gland through the haemolymph in ME-sensitive species. [24,64,65,66,67]. This process may require time to ensure sufficient accumulation and release of the pheromones [65]. Thus, acquisition of ME in the morning may be advantageous for ME-fed males to mate at dusk [31].
Attraction of certain Dacini male flies to lures is known to be age-dependent, which is correlated with sexual maturity in ME-sensitive species, such as B. dorsalis, B. carambolae, B. opiliae, and B. umbrosa, and RK-sensitive species, such as Z. cucurbitae [62,68,69,70,71,72,73]. However, attraction of male flies to the lure, zingerone, also occurred before sexual maturation in the Jarvis’s fruit fly B. jarvisi, under normal laboratory breeding conditions, suggesting this species has an olfactory system that could respond to zingerone even at immature stages [74].
Photoperiodicity involving light intensity is a crucial factor that affects mating in the tropics at consistent temperatures [75]. At dusk, concomitant with decreased light intensity below 100 lx, mating entailed male intrinsic behavioral patterns, such as short bursts of wing fanning (also termed pheromone-calling) to release male pheromone compounds from the rectal gland through the anus [76]. Furthermore, light intensity has been shown to modulate female mating and oviposition affected by contact and/or volatile fruit stimuli in B. oleae [77]. Following mating, female tephritid flies have been shown to behaviorally switch their preference for host fruit odors rather than pheromones of conspecific males [78]. The olfactory-mediated behavior in females is modulated by male accessory gland fluid [78]. Additionally, nutritional status can also modulate the odor preference of mature females. In B. dorsalis, protein-fed females were more attracted to host fruit odors than to protein odors, regardless of their mating status [79]. In contrast, protein-deprived mated females were equally attracted to the fruit and protein odors [79].
Nutritional status also affects the mating behavior of males. Adult Dacini flies are usually provided with a diet consisting of a mixture of sugar and yeast (enzymatic hydrolysate) as a rich source of carbohydrates and proteins in laboratories [80]. Dietary protein strongly affected mating success in B. dorsalis and B. tryoni [81,82]. Food deprivation, even in short-term treatments (30 h), decreased mating success in B. dorsalis [83]. In B. tryoni, feeding a yeast hydrolysate-supplemented diet increased male attraction to cue-lure [84]. Conversely, protein deprivation reduced the attractiveness of males to cue-lure in B. tryoni males [84], perhaps due to delayed sexual maturation. It is noteworthy that the ingestion of male attractants also affects mating success. In B. dorsalis, ME ingestion enhanced wing-fanning (pheromone calling) behavior, thereby increasing pheromone release and attracting more females [85]. Even when providing a low-quality diet, ME-supplementation could enhance mating success in males [83]. For RK-sensitive species, ingestion of RK along with yeast hydrolysate also promoted sexual maturation and mating success in B. tryoni [86,87]. In contrast, supplementation with RK or cue-lure reduced attraction to cue-lure-baited traps in B. tryoni and Z. cucurbitae [88,89,90]. When newly emerged adults were ingested with RK, no significant effects on sexual maturation, mating success, and attraction to cue-lure-baited traps were observed in Z. cucurbitae, suggesting that different metabolic processes, including transport and accumulation of male attractants, distinctly affect the physiology and behavior among Dacini species [91].
As various factors associated with physiological status affect chemosensory-related behaviors in Dacini fruit flies as mentioned above, an evaluation of any behavioral changes associated with chemosensory perception requires careful conditioning. In field behavioral bioassays, the conditions of the bioassays should be similar to the field conditions of the wild population, lest the bioassays are improperly designed and interpreted [92]. In laboratory behavioral bioassays, we should be aware of the possible differences between laboratory and field strains of insects. In mass-rearing facilities, genetic changes in Z. cucurbitae related to developmental and behavioral traits have been observed in more than 40 successive generations [93,94]. Furthermore, various abiotic and biotic environmental factors, such as light intensity, temperature, and dietary conditions, must be acknowledged [58]. Therefore, appropriate behavioral bioassays focusing on the problems underlying experimental design are required to identify specific chemicals that influence the behavior of flies.

5. Major Molecular Components in Insect Chemoreception

Insect chemosensory receptors consist of three types of insect-specific superfamilies: olfactory receptors (ORs), gustatory receptors (GRs), and ionotropic receptors (IRs) [51,95,96,97,98]. In addition, odorant binding proteins (OBPs) are thought to facilitate the transport of hydrophobic ligands through the aqueous sensillum lymph to appropriate ORs [95,99,100]. The roles of these chemosensory components have been mainly characterized by studies on D. melanogaster, taking advantage of a rich repertoire of genetic tools. The chemosensory receptors are thought to form ligand-gated ion channels [101,102,103,104,105,106]. Among the chemosensory receptors, insects use ORs, IRs, and carbon dioxide receptors belonging to the GR family to detect olfactory signals. GRs and IRs mediate the gustatory signals. Thus, IRs function not only in olfaction, but also in gustation. Since the development of next-generation sequencing (NGS), numerous chemosensory-related genes have been reported or registered in insects, including Dacini fruit flies [42,107,108,109,110,111]. Furthermore, draft genome sequences of five Dacini fruit flies (Bioproject ID)—B. dorsalis (PRJNA273958), B. latifrons (PRJNA351211), B. tryoni (PRJNA241080), B. oleae (PRJNA293367), and Z. cucurbitae (PRJNA273817)—have been deciphered, and their reference sequences are now available on the NCBI website. Genetic information enables identification of sequences coding candidate chemosensory components.

5.1. OBPs

OBPs are small, water-soluble, extracellular proteins that are involved in the first step of insect chemosensory processing [100]. A member of the OBP family was first reported as a pheromone binding protein (PBP) in Lepidoptera [99]. OBPs are responsible for transporting external odorants that pass through the pores of the sensilla into ORs on the odorant receptor neurons, thereby initiating olfactory signal transduction. Nevertheless, there is controversy regarding whether ORs are activated by an odorant-OBP complex or a single odorant molecule. Currently, there are two hypotheses on the mode of action of OBPs [112,113]: 1) OBPs bind with odorants, then form a complex in which the allosteric complex activates the receptor; 2) OBPs bind with odorants to delay odorant release in the sensillum lymph. In the second hypothesis, OBPs have been suggested to maintain the stability of the insect olfactory system by buffering the effects of sudden changes caused by odorant stimuli, rather than activating the receptor, while the odorants can directly activate the ORs [113]. Notably, a recent study has shown that CRISPR/Cas9-mediated knockout of PBP genes in the tobacco cutworm Spodoptera litura reduced antennal and behavioral responses to sex pheromone components, but did not functionally abolish them [114], which supports the idea that OBPs act as mediators for the transport of hydrophobic ligands in the sensillum lymph.
ME and RK are powerful lure components that attract Dacini fruit flies. OBPs that participate in the perception of ME in fruit flies have been identified, although none have been characterized for RK perception. For instance, there are 49 OBPs annotated in B. dorsalis [115]. In this review, we describe the names of OBPs in B. dorsalis according to the nomenclature of D. melanogaster [115], and also indicate names used in previous papers (within parentheses). Among the OBPs, BdorOBP56f-2 (BdorOBP2) was found to be involved in the perception of ME, with significant ME-induced transcriptional expression in B. dorsalis. Furthermore, BdorOBP56f-2 gene silencing by RNAi reduced the sensitivity of B. dorsalis to ME significantly [116]. In addition to BdorOBP56f-2, BdorOBP83b (BdorOBP83a-2) also affected the sensitivity of B. dorsalis to ME [117].
Other than Bactrocera species, OBPs have also been characterized in Ceratitidini and Carpomyini fruit flies. (E,E)-α-Farnesene has been regarded as one of the pheromone components of the Mediterranean fruit fly Ceratitis capitata (Ceratitidini), which is released by mature males [118,119]. At present, 46 OBP genes have been identified in Ce. capitata [120]. Among these OBPs, CcapOBP22 (previously named CcapOBP69a) and CcapOBP83a-2 showed high affinity for (E,E)-α-farnesene [119,121]. In addition, CvesOBP5 and CvesOBP6 also showed strong binding capacity to nonanal, which is a pheromone component of the tephritid species, Carpomya vesuviana (Carpomyini) [122].
Recent studies have indicated that γ-octalactone, benzothiazole, ethyl tiglate, 1-octen-3-ol, and β-caryophyllene can attract gravid female Dacini flies for oviposition [10,41,123]. OBPs participating in the perception of these odorants have been progressively identified. It has been documented that CvesOBP1, CvesOBP4, and BminOBP9 identified from B. minax showed high affinity for β-caryophyllene in an in vitro binding assay [122,124]. Furthermore, BdorOBP84a-1, identified from B. dorsalis, also showed moderate affinity to several volatiles, including 1-octen-3-ol [117].
Additionally, several OBPs have been reported to have a strong affinity for several common host volatiles. Research on BminOBP9 (BminGOBP99a) of B. minax showed that it bound not only to β-caryophyllene but also to other citrus volatiles, such as ocimene and myrcene [124]. It was also observed that BdorOBP99b (BdorOBP99a) of B. dorsalis had a high affinity for host volatiles, such as ocimene and limonene [124]. CvesOBP2 of Ca. vesuviana specifically bound well with the (Z)-3-hexenyl acetate contained in the host fruit [122]. Characterized OBPs and their main ligands are summarized in Table 1 and Figure 2.

5.2. ORs

Among the insect chemosensory receptors, ORs have been relatively well characterized as heteromeric ligand-gated ion channels that consist of a specific OR and a highly conserved co-receptor ORCO [104,105,125]. Recently, high-resolution structural analysis using cryo-electron microscopy demonstrated that ORCO and OR subunits assemble into heterotetramers [101]. Since in vitro functional analyses using Xenopus oocytes and culture cell lines have been established [95], insect ORs including Dacini fruit flies have been characterized. Since a specific odorous ligand is tuned to a specific OR in the insect olfactory system, the identification of the ligands for uncharacterized ORs could provide essential information to understand how various chemicals in the external environment are recognized by insects. However, only a limited number of ORs have been functionally characterized in Dacini fruit flies, although extensive transcriptome data derived from chemosensory organs have been accumulated. Most recently, several ORs of B. dorsalis and B. minax that respond to semiochemicals, including ME and host plant volatiles, have been characterized, as mentioned later.
The identification of chemosensory receptors responsible for the detection of male attractants is a significant challenge. Among the three chemosensory receptor families, ORs have been relatively well analyzed in Dacini fruit flies. Sex-specific behaviors, such as male attraction or female oviposition, stimulated by plant semiochemicals are observed in the life cycles of fruit flies. Although statistical differences in the transcriptional levels of several ORs between the sexes have been reported [111,126], sex-specific expression of any OR has not been reported. This is different from the cases of lepidopteran sex pheromone receptors responding to female pheromones, all of which are specifically expressed in male antennae [127,128,129,130,131,132,133]. The ME receptor has been considered as an OR, because a previous study has reported that ORCO is required for ME detection in B. dorsalis [134]. In this study, ME-exposure enhanced the transcriptional level of ORCO. Furthermore, RNAi-mediated knockdown of ORCO reduced the attractiveness of ME in the treated males. Most importantly, a recent study characterized OR88a as the ME receptor [135]. This study showed that the transcriptional level of OR88a was enhanced by exposure to ME, as seen for ORCO. Crucially, in vitro functional analysis using the Xenopus oocyte system revealed an obvious response of oocytes co-expressing OR88a and ORCO to ME. RNAi-mediated knockdown of OR88a decreased the attraction of male flies to ME. Further experiments, such as knockout of this gene using genome editing, will provide clear evidence for OR88a as a bona fide ME-receptor. It is also important to identify chemosensory receptors responding to male sex pheromones, because male attractants are converted into sex pheromones in ME-sensitive fruit flies. As mentioned above, ME is oxidized into DMP and E-CF as male sex pheromones in B. dorsalis. The structures of DMP and E-CF are very similar to ME; therefore, it is intriguing whether ME-receptors or other homologous chemosensory receptor(s) respond to both sex pheromones with a broad spectrum. Otherwise, distinct chemosensory receptors may specifically respond to individual pheromones. It is also interesting to refer to the different patterns of pheromone components among the related species of B. dorsalis. The carambola fruit fly B. carambolae, which is very close to B. dorsalis, produces only E-CF from ME [136]. In contrast, the peach fruit fly B. zonata and B. correcta possess binary mixtures of ME-oxidates in an approximate ratio of 1:1 in the male rectal glands, as shown in B. dorsalis, but their components were different [43,137]. B. zonata accumulated DMP and (Z)-coniferyl alcohol (Z-CF), whereas B. correcta accumulated (Z)-3,4-dimethoxycinnamyl alcohol (Z-DMC) and Z-CF (Figure 3). The patterns in which the related species partially share a phenylpropanoid component, as both B. dorsalis and B. zonata contain DMP, raising a question regarding similarities of male sex pheromone receptors. Since the different binary patterns in the pheromone components might play a critical role in sexual isolation of related species, comparative analysis of pheromone receptors provides insight into evolutionary speciation driven by differentiated chemosensory receptors. It is also interesting to note the difference in informative processing between sexes to male attractants and sex pheromones.
Other than phenolic metabolites, various compounds also accumulate in male rectal glands. Although most components have not been functionally characterized, 1-nonanol analogs that accumulate in the rectal glands as major aliphatic components have shown semiochemical activities in some fruit flies. For example, 6-oxo-1-nonanol was detected in the male rectal glands of B. carambolae, a species closely related to B. dorsalis, together with a minor component, 1,6-nonanediol [138,139]. The amount of 6-oxo-1-nonanol increased concomitantly with sexual maturity and triggered a chemotactic behavior known as zigzag flight in conspecific females [140]. In Z. cucurbitae, 1,3-nonanediol was detected as one of the major components in the male rectal glands at a level similar to that of accumulated RK [141]. Although the defensive role of 1,3-nonanediol against a natural enemy has been demonstrated, its pheromonal role is unknown [23]. A recent study characterized the ORs of three Dacini fruit flies, B. dorsalis, B. latifrons, and Z. cucurbitae, responding to 1-nonanol analogs by focusing on OR74a homologs, because OR74a of D. melanogaster responds to 1-nonanol [142]. The three OR74a homologs responded to 1-nonanol, but their sensitivities were different. The OR74a homologs identified from B. dorsalis and Z. cucurbitae responded significantly to 6-oxo-1-nonanol, but not to 1,3-nonanediol, indicating similar binding properties of the homologous ORs.
Most recently, several ORs that respond to common plant volatiles have been characterized in B. dorsalis and B. minax by co-expression with their cognate ORCO using Xenopus oocytes. For B. dorsalis, BdorOR13a and BdorOR82a have been functionally characterized [126]. BdorOR13a specifically responds to 1-octen-3-ol among the 24 tested compounds, as reported in its homologs of D. melanogaster and the mosquito species Anopheles gambiae and Aedes aegypti [143,144,145]. BdorOR82a robustly responds to geranyl acetate, as reported in its homologs in D. melanogaster and An. gambiae [146,147]. Furthermore, BdorOR82a responds weakly, but significantly to linalyl acetate and farnesene isomers [126]. There are no reports examining whether OR82a homologs also respond to these terpenes in other insects. As the three terpenes have been detected in T. catappa, one of the hosts of B. dorsalis [8], it is intriguing whether BdorOR82a is associated with host-finding behavior. Indeed, geranyl acetate and ME elicited the strongest EAD responses from the antennae of B. dorsalis among the volatiles collected from T. catappa [8]. Linalyl acetate and farnesene isomers also elicited medial EAD responses [8]. In accordance with the antennal responses, the landing behavior of female B. dorsalis on spheres treated with geranyl acetate or farnesene isomers was observed in the laboratory experiments [126].
For B. minax, four ORs—BminOR3, BminOR12, BminOR16, and BminOR24—have been functionally characterized [148,149]. Of them, BminOR3, a homolog of OR13a, also showed a specific response to 1-octen-3-ol, as shown by BdorOR13a. Other ORs have been newly characterized in insect species. BminOR12 responds to eight various compounds, namely a linear aliphatic alcohol, three aliphatic esters, a monoterpenoid, an aromatic alcohol, an aromatic aldehyde, and an aromatic ester. BminOR12 responded to 1-octanol, but not to 1-nonanol or undecanol, suggesting that this receptor does not respond to aliphatic alcohols with chains longer than C-8. BminOR12 did not respond to the unsaturated C-8 alcohol, cis-3-hexen-1-ol. Thus, it would be interesting to know if the olefin moiety reduces the responsiveness of BminOR12. For aliphatic esters, BminOR12 responded to butyl acrylate, butyl propionate, and (Z)-3-hexenyl acetate, but not to ethyl acetate, ethyl butyrate, butyl butyrate, or isoamyl acetate. The structure-activity relationship between BminOR12 and its ligands will be clarified by further examining its response to various esters. For aromatic compounds, BminOR12 responds to benzyl alcohol, benzaldehyde, (S)-(+)-carvone, and methyl salicylate, but not to p-cymeme, 2-methoxy phenol, or ME. How BminOR12 distinguishes different compounds belonging to the same category must be addressed in the future. BminOR16 specifically responds to undecanol. The phylogenetic tree constructed from the ORs of five dipteran species, including D. melanogaster, revealed a unique clade comprising three Bactrocera ORs, namely BminOR16, BdorOR67c.2, and BdorOR67c.3 [149]. This clade does not contain an OR of other dipteran species, suggesting that the three ORs are probably specific to Dacini fruit flies.
BminOR24, a homolog of OR7a, responds robustly to linalool, but weakly to ME [148]. Notably, female flies were attracted to linalool, a component of their host citrus fruits [150]. Considering the robust response of BminOR24 to linalool, gravid females may search for their host fruits by recognition of linalool via this receptor. Multiple homologous OR7a genes diverged in Dacini fruit flies, although only one corresponding OR7a gene has been identified in D. melanogaster [111,126,142]. Therefore, understanding the functional roles and divergent processes of OR members belonging to the OR7a subfamily in Dacini fruit flies is interesting. The characterized ORs and their main ligands are summarized in Table 2 and Figure 2.

5.3. Other Chemosensory Receptors

Candidate GRs and IRs have been identified by transcriptome sequencing in several Dacini fruit flies [42,107,108,110,111], although their functional analysis has not been addressed. In vitro functional analysis of GRs and IRs has not been successful even in the model insect D. melanogaster, probably due to the complexity of receptor structures, with a few exceptions [106,151]. As male attractants trigger male feeding behaviors, it would be interesting if gustatory receptors are involved in the recognition of attractants. The tip-recording method [57] is one of the methods available for investigating whether gustatory sensilla are responsible for perceiving attractants. If chemosensory receptors are necessary for gustation, orphan receptors belonging to the GR and/or IR family could be candidates, because their roles as gustatory receptors have been mainly characterized in D. melanogaster [95,96,97]. As male attractants have a hydrophobic nature, binding proteins such as OBPs and CSPs might be necessary to dissolve them in an aqueous lymphatic fluid.

6. Future Research

6.1. Characterization of Chemosensory Receptors Responding to Male Attractants and Sex Pheromones

ORs are considered to be chemosensory receptors for male attractants in Dacini fruit flies. Indeed, OR88a has been characterized as a candidate ME receptor in B. dorsalis by heterologous expression using Xenopus oocytes, and behavioral bioassays of male flies treated with RNAi-mediated OR88a knockdown [135]. The involvement of the OR system in the chemoreception of ME has also been shown by RNAi-mediated ORCO knockdown in B. dorsalis [134]. These studies have shown that reduced expression of OR88a or ORCO leads to a decrease in the attraction of male flies to ME. However, a complete impairment of attractiveness was not observed in the male flies. Targeted mutagenesis using the CRISPR/Cas9 system is now available for Bactrocera species [152]. Therefore, the loss of function of specific chemosensory receptors of interest will clarify their roles in vivo. Though it is plausible that a specific OR responds to ME, studying IRs is also intriguing, because they have been characterized as molecular components required for both olfaction and gustation [97]. The requirement of multiple IRs to form a functional receptor complex has made it difficult to analyze their functional properties [102,103]. IRs expressed in sexually dimorphic taste neurons have been characterized as candidate pheromone receptors in D. melanogaster [153]. As male attractants in Dacini fruit flies have both odorant and taste properties, certain IR complexes may act as bifunctional receptors for both olfactory and gustatory functions. Since IRs are thought to form heteromeric complexes consisting of the co-receptor IR8a or/and IR25a ion channels [102], the genetic loss of function of IR8a and IR25a using the CRISPR/Cas9 system will provide an answer to this question.
As the responsiveness of male Dacini flies to their respective attractants are different among species [13], a comparative study of chemosensory receptors responsible for male attractants will provide insights into the structure-activity relationships from the point of view of the structural properties of chemosensory receptors. It will also unravel the evolutionary process of the distinct ME- and RK-sensitive fruit fly groups to adapt to their surrounding environment, such as the source of attractants derived from plants, including fruit fly orchids. As of the structural resemblance of ME and RK, the difference between the corresponding chemosensory receptors is interesting. A few differences in the critical amino acids in receptors possibly impart the distinct ligand-binding properties. Furthermore, both ME- and RK-sensitive fruit fly species are attracted to zingerone owing to its structural resemblance to ME and RK [23]. A recent study on the chemical modification of zingerone has shown the structural properties required for the attraction of B. jarvisi [154]. Therefore, it will be interesting to study partial structures in ME- and RK-receptors that are necessary for binding to zingerone. Notably, males of B. latifrons are attracted to and feed on α-ionone analogs [47,48,49,50]. The different skeletons of specific male attractants, i.e., α-ionone analogs for B. latifrons and aromatic compounds for B. dorsalis and Z. cucurbitae, suggest that chemosensory receptors in both these species bear greater similarity than those in B. latifrons. Nevertheless, the phylogenetic analysis of ORs shows that B. dorsalis and B. latifrons share more similar homologous genes than Z. cucurbitae [111,142], reflecting their phylogenetic relationships. Structural properties of chemosensory receptors responsible for the different affinities for the α-ionone analogs and aromatic compounds are important from various aspects, such as structure-activity relationships, evolutionary changes in chemoreception, and development of new lures.
Unlike male attractants, male sex pheromones released from the rectal glands as smoke affect females at a close distance [29,32]. Therefore, the nature of chemosensory receptors responsible for the detection of sex pheromones is of interest. ME-sensitive fruit flies utilize the analogous male sex pheromones derived from ME. Moreover, the related species share the same components partially (Figure 3) [43,136,137]. Therefore, slight differences in amino acids that are critical for ligand binding in chemosensory receptors may change the ligand-binding properties among the related species. In this case, the mode of acquisition for different amino acid sequences during the evolution of pheromone production and reception is very important. Although males accumulate ingested RK in their rectal glands, the role of RK is unclear [14]. Identification of chemosensory receptors for RK, and the subsequent loss of function analysis could provide a clue to understanding the roles of RK in sexual communication of fruit flies.
Various components other than the aromatic compounds such as ME and RK have been identified in the rectal glands. The most interesting case is the spiroketal, olean, in B. oleae [35,36,37]. The different responses to the enantiomers between males and females raise the question of how chiral discrimination is achieved. It is plausible that different, but similar, chemosensory receptors are responsible for sensing the (R)- and (S)-enantiomers. If so, the structural properties for enantiomeric specificity must be key for the different recognition between sexes. Other than spiroketals, various acetamides, cyclolactones, and aliphatic compounds have been identified [38], but their biological roles remain unknown. We may be able to presume the roles of the rectal gland components by characterizing chemosensory receptors responding to the uncharacterized components.

6.2. Characterization of Chemosensory Receptors Responding to Common Plant Volatiles

Recently, ORs responding to common plant volatiles have been functionally characterized in B. dorsalis and B. minax [126,148,149]. The results provide knowledge of the properties of ORs which may narrowly or broadly tune to plant volatiles. Among the common plant volatiles, the sesquiterpenoid β-caryophyllene attracts males of B. correcta, but not those of B. dorsalis [44]. As β-caryophyllene affects phytophagous insects in many aspects of plant-insect interactions, identification of chemosensory receptor(s) will provide information about the common recognition of plants via chemoreception in phytophagous insects.

6.3. Information Processing of Chemosensory Inputs

It is important to understand how chemosensory information is processed in the brain, because olfactory and gustatory inputs seem to be associated with each other in Dacini fruit flies. Olfactory neurons within the chemosensilla project their axons into the glomeruli within the antennal lobe of the brain in insects [51,155]. The glomeruli are the first relay sites where olfactory neurons are synaptically connected with the projection neurons conveying processed olfactory information to higher brain centers. Individual glomeruli receive convergent input from multiple olfactory neurons expressing the same ORs, indicating a functional module reflecting the nature of odor stimuli [156,157]. Thus, a comprehensive understanding of the neural mechanisms of odor detection and processing requires an olfactory sensory map of the brain. Recently, glomerular organization in the antennal lobe has been described in B. dorsalis [158]. In total, 64–65 glomeruli were anatomically classified in both sexes using synaptic antibody staining. The number of glomeruli was in the range of 50–70, in other dipteran species. By comparing male and female antennal lobes, eight glomeruli showing enlargement in males or females were identified [158]. As only males are strongly attracted to ME, it is of interest to know whether these sex-dimorphic glomeruli are involved in the recognition of ME. A recent study used Ca2+ imaging in B. tryoni to identify a region in the antennal lobe, which is the primary olfactory center in the brain responsible for the recognition of odors, including cue-lure [53]. The standard reference stimulus, octanol, elicited a distinct enhancement of Ca2+ concentration in the glomeruli of both sexes. It is expected that maxillary palp-innervated glomeruli are activated by cue-lure stimulation, because the maxillary palps, but not the antennae, respond to cue-lure in electrophysiological analyses. However, no evident response in the antennal lobe, including the glomeruli, was observed by cue-lure stimulation, probably due to technical difficulties in tracing projection from the maxillary palps [53].
In addition to the attractants, information processing of male sex pheromones in females of ME-sensitive fruit flies is also a crucial issue. Due to the structural similarities between ME and sex pheromones, identification of receptors and characterization of their neural projection patterns will help us better understand the mechanisms underlying the different responses and information processing of similar compounds between males and females. Similarly, differences in information processing of the enantiomers of the spiroketal, olean, between sexes in B. oleae is intriguing. It must be explored whether chemosensory receptor neurons discriminating the enantiomers project into similar or different area in the brain between males and females.
It is possible that ORs modulate male feeding behavior. In the blowfly P. regina, axonal projections of olfactory receptor neurons from the maxillary palps to the primary gustatory center, the subesophageal ganglion (SOG), as well as to the antennal lobes have been shown by an anterograde staining method [56]. Furthermore, overlapping projection of maxillary olfactory receptor neurons and labellar gustatory neurons in the SOG was observed, suggesting an interaction between olfactory input from the maxillary palps and gustatory stimuli in the labellum. In behavioral bioassays, an attractive odor, 1-octen-3-ol, increased the response of the proboscis extension reflex to sucrose via the maxillary palps, but not via the antennae, which supports the proposed integration of olfactory and gustatory information in the SOG. This framework may be applicable to attractant-feeding behavior in Dacini fruit flies, as voracious feeding behavior is observed after attraction to lures. In the mosquito, An. gambiae, the innervation of ORCO-expressing neurons from the proboscis into the subesophageal zone was observed [159]. However, no innervation of ORCO-expressing neurons from the olfactory organs, antennae or maxillary palps, was detected in this species. Thus, neural connections seem to differ among dipteran species. Comprehensive studies on the relationships between neural networks and behavioral outputs would provide insights into the mechanisms underlying information processing in multiple chemosensory systems.

6.4. Mechanisms of Physiological Change Associated with Mating Behavior

One of the reasons why Dacini males voraciously feed on attractants is to elevate mating success. Ingestion of male attractants activates the characteristic pheromone calling behavior, wing-fanning, and simultaneously disperses pheromones from their rectal glands [31,83,85,86,87,90]. Activation of male behavior by ingestion of attractants is an intriguing question from the point of view of insect endocrinology. In the Caribbean fruit fly Anastrepha suspensa (Toxotrypanini), treatment with juvenile hormone promoted male sexual maturation; therefore, males produced sex pheromones at earlier ages than untreated males [160]. Similarly, treatment with a juvenile hormone mimic, methoprene, accelerated sexual maturation and enhanced mating behavior of males in RK-sensitive Z. cucurbitae and B. tryoni [161,162]. Taken together, the accelerated sexual maturation by male attractants is possibly associated with the action of juvenile hormone. In contrast, no significant effect of methoprene on male sexual activity was observed in B. dorsalis or Ce. capitata [163]. The different effects on sexual maturation may be caused by different sensitivities to or the mode of action of juvenile hormone among fruit fly species. Recently, the molecular action of juvenile hormone, including binding of ligand to receptors and subsequent signaling pathways, has been well elucidated from studies on various insects, such as dipteran, lepidopteran, coleopteran, and hemipteran species [164,165]. This information will provide cues to elucidate the mechanisms underlying the sexual maturation-promoting action of male attractants and the possible involvement of juvenile hormone in Dacini fruit flies.
Recent studies have shown that the gut microbiome can modulate various aspects of physiology and behavior, such as reproduction and feeding, in Dacini fruit flies. Symbiotic bacteria affect the foraging behavior of adult flies towards nutritional resources in B. dorsalis [166,167], and oviposition behavior in B. oleae [168]. Considering that chemical stimuli are a crucial cue for locating appropriate food sources and oviposition sites, it is intriguing whether chemosensory perception is affected by the gut microbiome in Dacini fruit flies.

6.5. New Approaches Such as Promoter Analysis and Ectopic Expression Using New Genetic Techniques

Insect chemosensory mechanisms have been elucidated in D. melanogaster taking advantage of the availability of genetic techniques. For example, visualization of chemosensory neurons using fluorescent proteins linked to the promoter region of a target gene has revealed neuronal projections into an information processing region in the brain [155]. Ectopic expression of calcium indicators, such as GCaMP, in a target chemosensory neuron has enabled the monitoring of neuronal activities stimulated by ligands. Although these techniques are limited to D. melanogaster, the utilization of a binary expression system, the QF/QUAS system [169], in Anopheles mosquitoes has successfully provided insights into olfactory innervations in the brain and olfactory system underlying the mode of action for insect repellents [159,170]. The successful generation of transgenic strains for the QF/QUAS system has enabled robust expression of reporters in chemosensory neurons. The labeling of ORCO-expressing neurons with GFP represents detailed anatomical maps of olfactory innervations in the brain [159]. The ectopic expression of GCaMP in the ORCO-expressing neurons allowed the direct visualization of olfactory responses to odorants in olfactory neurons, which demonstrated the different modes of action between natural repellents and synthetic repellents [170]. Furthermore, the advent of genome editing has made transgenic techniques applicable to a wide range of non-model insects [171]. For example, a CRISPR/Cas9-based knock-in technique has been developed to elucidate the roles of an ion channel, ppk301, in the control of egg-laying initiation and choice in the mosquito, Ae. aegypti [172]. The generation of a transgenic reporter strain in which a transcriptional activator was inserted into the ppk301 locus led to a loss-of-function mutation. This also enabled ectopic expression of the transcriptional activator in ppk301-expressing cells. GFP expression using this ppk301-activator revealed projections of ppk301-expressing sensory neurons into central gustatory centers, the subesophageal zone, in the brain. Furthermore, the genetically introduced calcium sensor GCaMP6s in ppk301-expressing sensory neurons showed obvious responses to water using calcium imaging.
Thus, useful genetic tools developed in Drosophila could also be adopted for similar uses to elucidate chemosensory mechanisms in Dacini flies, as shown by the mosquito studies. A visualization of the innervation patterns of ORCO- and IR-expressing neurons provides information on chemosensory processing in the brain. Generation of transgenic strains, in which a transcriptional activator is regulated by a promoter of chemosensory receptors, render various analyses possible. For example, ectopic expression of GCaMP in cells of a chemosensory receptor would enable the identification of ligands for the receptor. These new approaches to study Dacini flies will not only broaden our knowledge of the chemosensory mechanisms underlying evolutionary adaptation to environmental changes, but also provide novel techniques that are useful for pest management.

7. Conclusions

Herein, we have reviewed chemoreception and related topics in Dacini fruit flies from various aspects of chemical ecology, physiology, neuroscience, and molecular biology. Due to the importance of pest management and the intriguing life cycle of Dacini fruit flies, a large variety of semiochemicals intimately associated with the life cycles of various species have been well identified since the discovery of a male lure for Dacini fruit flies in 1912 [16]. In addition, only a few molecular components required for chemoreception have been characterized for a limited number of species. In particular, identification of ligands for ORs has been achieved in the past few years. Molecular and cellular understanding of chemosensory mechanisms is much more advanced in the model insect D. melanogaster owing to the availability of various genetic tools. Currently, several genetic tools developed in D. melanogaster have become available for non-model species using CRISPR/Cas9 techniques. Thus, the mechanisms underlying the perception by peripheral organs and information processing in the brain may be elucidated in Dacini fruit flies by utilizing these new tools. Numerous studies on the unique behavioral and physiological traits associated with chemosensory perception have been conducted in Dacini fruit flies; therefore, understanding the molecular and physiological mechanisms of the chemosensory system provides insights into how Dacini fruit flies have optimally adapted to their surrounding environments via chemoreception at the molecular level. Such detailed understanding of chemical communication strategies also provides cues for the development of new biocontrol methods.

Author Contributions

Writing—original draft preparation, H.O., A.K.-W.H., H.J.; writing—review and editing, H.O., A.K.-W.H., H.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partly supported by JSPS KAKENHI Grant Numbers 26450466 and 18H02214 to HO, by the IAEA Research Contract No. 20377 to AKWH, and funding from the National Natural Science Foundation of China (3177110496, 32072491), and Chongqing Natural Science Foundation (cstc2019jcyj-zdxmX0034) to HJ.

Conflicts of Interest

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

References

  1. Krosch, M.N.; Schutze, M.K.; Armstrong, K.F.; Graham, G.C.; Yeates, D.K.; Clarke, A.R. A molecular phylogeny for the Tribe Dacini (Diptera: Tephritidae): Systematic and biogeographic implications. Mol. Phylogenet. Evol. 2012, 64, 513–523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Bernays, E.A.; Chapman, R.F. Host-Plant Selection by Phytophagous Insects; Chapman & Hall: New York, NY, USA, 1994; ISBN 9780585304557. [Google Scholar]
  3. Tan, K.H.; Nishida, R. Methyl eugenol: Its occurrence, distribution, and role in nature, especially in relation to insect behavior and pollination. J. Insect Sci. 2012, 12, 56. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Steiner, L.F.; Mitchell, W.C.; Harris, E.J.; Kozuma, T.T.; Fujimoto, M.S. Oriental fruit fly eradication by male annihilation. J. Econ. Entomol. 1965, 58, 961–964. [Google Scholar] [CrossRef]
  5. Steiner, L.F.; Hart, W.G.; Harris, E.J.; Cunningham, R.T.; Ohinata, K.; Kamakahi, D.C. Eradication of the oriental fruit fly from the Mariana Islands by the methods of male annihilation and sterile insect release. J. Econ. Entomol. 1970, 63, 131–135. [Google Scholar] [CrossRef]
  6. Koyama, J.; Teruya, T.; Tanaka, K. Eradication of the oriental fruit fly (Diptera: Tephritidae) from the Okinawa Islands by a male annihilation method. J. Econ. Entomol. 1984, 77, 468–472. [Google Scholar] [CrossRef]
  7. Siderhurst, M.; Jang, E. Female-biased attraction of oriental fruit fly, Bactrocera dorsalis (Hendel), to a blend of host fruit volatiles from Terminalia catappa L. J. Chem. Ecol. 2006, 32, 2513–2524. [Google Scholar] [CrossRef]
  8. Siderhurst, M.; Jang, E. Attraction of female oriental fruit fly, Bactrocera dorsalis, to Terminalia catappa fruit extracts in wind tunnel and olfactometer tests. Formos. Entomol. 2006, 26, 45–55. [Google Scholar]
  9. Kamala Jayanthi, P.D.; Woodcock, C.M.; Caulfield, J.; Birkett, M.A.; Bruce, T.J.A. Isolation and identification of host cues from mango, Mangifera indica, that attract gravid female oriental fruit fly, Bactrocera dorsalis. J. Chem. Ecol. 2012, 38, 361–369. [Google Scholar] [CrossRef]
  10. Kamala Jayanthi, P.D.; Kempraj, V.; Aurade, R.M.; Venkataramanappa, R.K.; Nandagopal, B.; Verghese, A.; Bruce, T.J.A. Specific volatile compounds from mango elicit oviposition in gravid Bactrocera dorsalis females. J. Chem. Ecol. 2014, 40, 259–266. [Google Scholar] [CrossRef]
  11. Biasazin, T.D.; Karlsson, M.F.; Hillbur, Y.; Seyoum, E.; Dekker, T. Identification of host blends that attract the african invasive fruit fly, Bactrocera invadens. J. Chem. Ecol. 2014, 40, 966–976. [Google Scholar] [CrossRef]
  12. Nishida, R.; Tan, K.H. Search for new fruit fly attractants from plants: A review. In Proceedings of the 9th International Symposium on Fruit Flies of Economic Importance, Bangkok, Thailand, 12 May 2014; pp. 249–262. [Google Scholar]
  13. Tan, K.H.; Nishida, R.; Jang, E.B.; Shelly, T.E. Pheromones, Male Lures, and Trapping of Tephritid Fruit Flies; Springer: Dordrecht, The Netherlands, 2014; ISBN 9789401791922. [Google Scholar]
  14. Shelly, T. Effects of methyl eugenol and raspberry ketone/cue lure on the sexual behavior of Bactrocera species (Diptera: Tephritidae). Appl. Entomol. Zool. 2010, 45, 349–361. [Google Scholar] [CrossRef] [Green Version]
  15. Howlett, F.M. VII. The effect of oil of Citronella on two species of Dacus. Trans. R. Entomol. Soc. Lond. 1912, 60, 412–418. [Google Scholar] [CrossRef]
  16. Howlett, F. Chemical reactions of fruit-flies. Bull. Entomol. Res. 1915, 6, 297–305. [Google Scholar] [CrossRef]
  17. Royer, J.E.; Khan, M.; Mayer, D.G. Methyl-isoeugenol, a highly attractive male lure for the cucurbit flower pest Zeugodacus diversus (coquillett) (syn. Bactrocera diversa) (Diptera: Tephritidae: Dacinae). J. Econ. Entomol. 2018, 111, 1197–1201. [Google Scholar] [CrossRef]
  18. Drew, R.A.I.; Hooper, G.H.S. The responses of fruit fly species (Diptera: Tephritidae) in Australia to various attractants. Aust. J. Entomol. 1981, 20, 201–205. [Google Scholar] [CrossRef]
  19. Drew, R.A.I. The responses of fruit fly species (Diptera: Tephritidae) in the South Pacific area to male attractants. Aust. J. Entomol. 1974, 13, 267–270. [Google Scholar] [CrossRef]
  20. Tan, K.H.; Nishida, R. Synomone or kairomone?—Bulbophyllum apertum flower releases raspberry ketone to attract Bactrocera fruit flies. J. Chem. Ecol. 2005, 31, 497–507. [Google Scholar] [CrossRef]
  21. Tan, K.; Nishida, R.; Toong, Y. Floral synomone of a wild orchid, Bulbophyllum cheiri, lures Bactrocera fruit flies for pollination. J. Chem. Ecol. Ecol. 2002, 28, 1161–1172. [Google Scholar] [CrossRef]
  22. Beroza, M.; Alexanderl, B.H.; Steiner, F.; Mitchell, W.C.; Miyashita, D.H. New synthetic lures for the male melon fly. Science 1960, 131, 1044–1045. [Google Scholar] [CrossRef]
  23. Tan, K.H.; Nishida, R. Mutual reproductive benefits between a wild orchid, Bulbophyllum patens, and Bactrocera fruit flies via a floral synomone. J. Chem. Ecol. 2000, 26, 533–546. [Google Scholar] [CrossRef]
  24. Nishida, R.; Tan, K.H.; Serit, M.; Lajis, N.H.; Sukari, A.M.; Takahashi, S.; Fukami, H. Accumulation of phenylpropanoids in the rectal glands of males of the oriental fruit fly, Dacus dorsalis. Experientia 1988, 44, 534–536. [Google Scholar] [CrossRef]
  25. Fletcher, B.S. Storage and release of a sex pheromone by the Queensland fruit fly, Dacus tryoni (Diptera: Trypetidae). Nature 1968, 219, 631–632. [Google Scholar] [CrossRef] [PubMed]
  26. Fletcher, B.S. The structure and function of the sex pheromone glands of the male Queensland fruit fly, Dacus tryoni. J. Insect Physiol. 1969, 15, 1309–1322. [Google Scholar] [CrossRef]
  27. Nishida, R.; Iwahashi, O.; Tan, K.H. Accumulation of Dendrobium superbum (Orchidaceae) fragrance in the rectal glands by males of the melon fly, Dacus cucurbitae. J. Chem. Ecol. 1993, 19, 713–722. [Google Scholar] [CrossRef] [PubMed]
  28. Tan, K.H.; Nishida, R. Incorporation of raspberry ketone in the rectal glands of males of the Queensland fruit fly, Bactrocera tryoni Froggatt (Diptera: Tephritidae). Appl. Entomol. Zool. 1995, 30, 494–497. [Google Scholar] [CrossRef] [Green Version]
  29. Kobayashi, R.M.; Ohinata, K.; Chambers, D.L.; Fujimoto, M.S. Sex pheromones of the oriental fruit fly and the melon fly: Mating behavior, bioassay method, and attraction of females by live males and by suspected pheromone glands of males. Environ. Entomol. 1978, 7, 107–112. [Google Scholar] [CrossRef]
  30. Khoo, C.C.H.; Yuen, K.H.; Tan, K.H. Attraction of female Bactrocera papayae to sex pheromone components with two different release devices. J. Chem. Ecol. 2000, 26, 2487–2496. [Google Scholar] [CrossRef]
  31. Hee, A.K.W.; Tan, K.H. Attraction of female and male Bactrocera papayae to conspecific males fed with methyl eugenol and attraction of females to male sex pheromone components. J. Chem. Ecol. 1998, 24, 753–764. [Google Scholar] [CrossRef]
  32. Ohinata, K.; Jacobson, M.; Kobayashi, R.M.; Chambers, D.L.; Fujimoto, M.S.; Higa, H.H. Oriental fruit fly and melon fly; biological and chemical studies of smoke produced by males. J. Environ. Sci. Health Part A Environ. Sci. Eng. 1982, 17, 197–216. [Google Scholar] [CrossRef]
  33. Khoo, C.C.H.; Tan, K.H. Attraction of both sexes of melon fly, Bactrocera cucurbitae to conspecific males—A comparison after pharmacophagy of cue-lure and a new attractant—Zingerone. Entomol. Exp. Appl. 2000, 97, 317–320. [Google Scholar] [CrossRef]
  34. Haniotakis, G.E.; Mazomenos, B.E.; Tumlinson, J.H. A sex attractant of the olive fruit fly, Dacus oleae and its biological activity under laboratory and field conditions. Entomol. Exp. Appl. 1977, 21, 81–87. [Google Scholar] [CrossRef]
  35. Baker, R.; Herbert, R.; Howse, P.E.; Jones, O.T.; Francke, W.; Reith, W. Identification and synthesis of the major sex pheromone of the olive fly (Dacus oleae). J. Chem. Soc. Chem. Commun. 1980, 2, 52–53. [Google Scholar] [CrossRef]
  36. Mazomenos, B.E.; Haniotakis, G.E. Male olive fruit fly attraction to synthetic sex pheromone components in laboratory and field tests. J. Chem. Ecol. 1985, 11, 397–405. [Google Scholar] [CrossRef] [PubMed]
  37. Haniotakis, G.; Francke, W.; Mori, K.; Redlich, H.; Schurig, V. Sex-specific activity of (R)-(-)- and (S)- (+)-1,7-dioxaspiro[5.5]undecane, the major pheromone of Dacus oleae. J. Chem. Ecol. 1986, 12, 1559–1568. [Google Scholar] [CrossRef] [PubMed]
  38. Fletcher, M.T.; Kitching, W. Chemistry of fruit flies. Chem. Rev. 1995, 95, 789–828. [Google Scholar] [CrossRef]
  39. Quilici, S.; Atiama-Nurbel, T.; Brévault, T. Plant odors as fruit fly attractants. In Trapping and the Detection, Control, and Regulation of Tephritid Fruit Flies; Shelly, T.E., Epsky, N., Jang, E.B., Reyes-Flores, J., Vargas, R.I., Eds.; Springer: Dordrecht, The Netherlands, 2014; pp. 119–144. ISBN 978-94-017-9192-2. [Google Scholar]
  40. Light, D.M.; Jang, E.B. Electroantennogram responses of the oriental fruit fly, Dacus dorsalis, to a spectrum of alcohol and aldehyde plant volatiles. Entomol. Exp. Appl. 1987, 45, 55–64. [Google Scholar] [CrossRef]
  41. Damodaram, K.J.P.; Kempraj, V.; Aurade, R.M.; Venkataramanappa, R.K.; Nandagopal, B.; Verghese, A.; Bruce, T. Oviposition site-selection by Bactrocera dorsalis is mediated through an innate recognition template tuned to γ-octalactone. PLoS ONE 2014, 9, e85764. [Google Scholar] [CrossRef]
  42. Liu, Z.; Smagghe, G.; Lei, Z.; Wang, J.J. Identification of male- and female-specific olfaction genes in antennae of the oriental fruit fly (Bactrocera dorsalis). PLoS ONE 2016, 11, e0147783. [Google Scholar] [CrossRef] [Green Version]
  43. Tokushima, I.; Orankanok, W.; Tan, K.H.; Ono, H.; Nishida, R. Accumulation of phenylpropanoid and sesquiterpenoid volatiles in male rectal pheromonal glands of the guava fruit fly, Bactrocera correcta. J. Chem. Ecol. 2010, 36, 1327–1334. [Google Scholar] [CrossRef]
  44. Wee, S.L.; Chinvinijkul, S.; Tan, K.H.; Nishida, R. A new and highly effective male lure for the guava fruit fly Bactrocera correcta. J. Pest. Sci. 2018, 91, 691–698. [Google Scholar] [CrossRef]
  45. Kamiji, T.; Kaneda, M.; Sasaki, M.; Ohto, K. Sexual maturation of male Bactrocera correcta (Diptera: Tephritidae) and age-related responses to β-caryophyllene and methyl eugenol. Appl. Entomol. Zool. 2018, 53, 41–46. [Google Scholar] [CrossRef]
  46. Jaleel, W.; He, Y.; Lü, L. The response of two Bactrocera species (Diptera: Tephritidae) to fruit volatiles. J. Asia. Pac. Entomol. 2019, 22, 758–765. [Google Scholar] [CrossRef]
  47. Flath, R.A.; Cunningham, R.T.; Liquido, N.; McGovern, T.P. Alpha-ionol as attractant for trapping Bactrocera latifrons (Diptera: Tephritidae). J. Econ. Entomol. 1994, 87, 1470–1476. [Google Scholar] [CrossRef]
  48. Ishida, T.; Enomoto, H.; Nishida, R. New attractants for males of the solanaceous fruit fly Bactrocera latifrons. J. Chem. Ecol. 2008, 34, 1532–1535. [Google Scholar] [CrossRef] [PubMed]
  49. Nishida, R.; Enomoto, H.; Shelly, T.E.; Ishida, T. Sequestration of 3-oxygenated α-ionone derivatives in the male rectal gland of the solanaceous fruit fly, Bactrocera latifrons. Entomol. Exp. Appl. 2009, 131, 85–92. [Google Scholar] [CrossRef]
  50. Enomoto, H.; Ishida, T.; Hamagami, A.; Nishida, R. 3-Oxygenated α-ionone derivatives as potent male attractants for the solanaceous fruit fly, Bactrocera latifrons (Diptera: Tephritidae), and sequestered metabolites in the rectal gland. Appl. Entomol. Zool. 2010, 45, 551–556. [Google Scholar] [CrossRef] [Green Version]
  51. Vosshall, L.B.; Stocker, R.F. Molecular architecture of smell and taste in Drosophila. Annu. Rev. Neurosci. 2007, 30, 505–533. [Google Scholar] [CrossRef] [Green Version]
  52. Chieng, A.C.T.; Hee, A.K.W.; Wee, S.L. Involvement of the antennal and maxillary palp structures in detection and response to methyl eugenol by male Bactrocera dorsalis (Diptera: Tephritidae). J. Insect Sci. 2018, 18, 19. [Google Scholar] [CrossRef]
  53. Verschut, T.A.; Farnier, K.; Cunningham, J.P.; Carlsson, M.A. Behavioral and physiological evidence for palp detection of the male-specific attractant cuelure in the Queensland fruit fly (Bactrocera tryoni). Front. Physiol. 2018, 9, 990. [Google Scholar] [CrossRef]
  54. Park, K.C.; Jeong, S.A.; Kwon, G.; Oh, H.W. Olfactory attraction mediated by the maxillary palps in the striped fruit fly, Bactrocera scutellata: Electrophysiological and behavioral study. Arch. Insect Biochem. Physiol. 2018, 99, 1–13. [Google Scholar] [CrossRef]
  55. Shiraiwa, T. Multimodal chemosensory integration through the maxillary palp in Drosophila. PLoS ONE 2008, 3, e2191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Maeda, T.; Tamotsu, S.; Iwasaki, M.; Nisimura, T.; Shimohigashi, M.; Hojo, M.K.; Ozaki, M. Neuronal projections and putative interaction of multimodal inputs in the subesophageal ganglion in the blowfly, Phormia regina. Chem. Senses 2014, 39, 391–401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Hodgson, E.S.; Lettvin, J.Y.; Roeder, K.D. Physiology of a primary chemoreceptor unit. Science 1955, 122, 417–418. [Google Scholar] [CrossRef] [PubMed]
  58. Díaz-Fleischer, F.; Piñero, J.; Shelly, T. Interactions between tephritid fruit fly physiological state and stimuli from baits and traps: Looking for the pied piper of hamelin to lure pestiferous fruit flies. In Trapping and the Detection, Control, and Regulation of Tephritid Fruit Flies; Shelly, T., Epsky, N., Jang, E., Reyes-Flores, J., Vargas, R., Eds.; Springer: Dordrecht, The Netherlands, 2014; pp. 145–172. [Google Scholar]
  59. Gadenne, C.; Barrozo, R.B.; Anton, S. Plasticity in insect olfaction: To smell or not to smell? Annu. Rev. Entomol. 2016, 61, 317–333. [Google Scholar] [CrossRef] [PubMed]
  60. Manoukis, N.C.; Jang, E.B. The diurnal rhythmicity of Bactrocera cucurbitae (Diptera: Tephritidae) attraction to cuelure: Insights from an interruptable lure and computer vision. Ann. Entomol. Soc. Am. 2013, 106, 136–142. [Google Scholar] [CrossRef]
  61. Raghu, S.; Clarke, A.R. Spatial and temporal partitioning of behaviour by adult dacines: Direct evidence for methyl eugenol as a mate rendezvous cue for Bactrocera cacuminata. Physiol. Entomol. 2003, 28, 175–184. [Google Scholar] [CrossRef]
  62. Fitt, G.P. The influence of age, nutrition and time of day on the responsiveness of male Dacus opiliae to the synthetic lure, methyl eugenol. Entomol. Exp. Appl. 1981, 30, 83–90. [Google Scholar] [CrossRef]
  63. Wee, S.L.; Hee, A.K.W. Diurnal attraction of fruit flies (Diptera: Tephritidae) to methyl eugenol in a village ecosystem in Tanjung Bungah, Penang, Malaysia. Serangga 2018, 23, 83–91. [Google Scholar]
  64. Hiap, W.W.; Wee, S.L.; Tan, K.H.; Hee, A.K.W. Phenylpropanoid sex pheromone component in hemolymph of male Carambola fruit fly, Bactrocera carambolae (Diptera: Tephritidae). Chemoecology 2019, 29, 25–34. [Google Scholar] [CrossRef]
  65. Hee, A.K.; Tan, K.H. Transport of methyl eugenol-derived sex pheromonal components in the male fruit fly, Bactrocera dorsalis. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2006, 143, 422–428. [Google Scholar] [CrossRef]
  66. Hee, A.K.W.; Tan, K.H. Bioactive fractions containing methyl eugenol-derived sex pheromonal components in haemolymph of the male fruit fly Bactrocera dorsalis (Diptera: Tephritidae). Bull. Entomol. Res. 2005, 95, 615–620. [Google Scholar] [CrossRef] [PubMed]
  67. Hee, A.K.; Tan, K. Male sex pheromonal components derived from methyl eugenol in the hemolymph of the fruit fly Bactrocera papayae. J. Chem. Ecol. 2004, 30, 2127–2138. [Google Scholar] [CrossRef] [PubMed]
  68. Tan, K.; Kirton, L.; Serit, M. Age response of Dacus dorsalis (Hendel) to methyl eugenol in A) a wind tunnel and B) traps set in a village, and its implication in population estimation. In Fruit Flies: Proceedings of the Second International Symposium; Economopoulos, A.P., Ed.; Elsevier: New York, NY, USA, 1987; pp. 425–432. ISBN 978-0444989468. [Google Scholar]
  69. Wong, T.T.Y.; McInnis, D.O.; Nishimoto, J.I. Relationship of sexual maturation rate to response of oriental fruit fly strains (Diptera: Tephritidae) to methyl eugenol. J. Chem. Ecol. 1989, 15, 1399–1405. [Google Scholar] [CrossRef] [PubMed]
  70. Shelly, T.E.; Edu, J.; Pahio, E.; Wee, S.L.; Nishida, R. Re-examining the relationship between sexual maturation and age of response to methyl eugenol in males of the oriental fruit fly. Entomol. Exp. Appl. 2008, 128, 380–388. [Google Scholar] [CrossRef]
  71. Wee, S.L.; Tan, K.H. Sexual maturity and intraspecific mating success of two sibling species of the Bactrocera dorsalis complex. Entomol. Exp. Appl. 2000, 94, 133–139. [Google Scholar] [CrossRef]
  72. Wong, T.T.Y.; McInnis, D.O.; Ramadan, M.M.; Nishimoto, J.I. Age-related response of male melon flies Dacus cucurbitae (Diptera: Tephritidae) to cue-lure. J. Chem. Ecol. 1991, 17, 2481–2487. [Google Scholar] [CrossRef] [PubMed]
  73. Wee, S.L.; Abdul Munir, M.Z.; Hee, A.K.W. Attraction and consumption of methyl eugenol by male Bactrocera umbrosa Fabricius (Diptera: Tephritidae) promotes conspecific sexual communication and mating performance. Bull. Entomol. Res. 2018, 108, 116–124. [Google Scholar] [CrossRef]
  74. Wee, S.L.; Peek, T.; Clarke, A.R. The responsiveness of Bactrocera jarvisi (Diptera: Tephritidae) to two naturally occurring phenylbutaonids, zingerone and raspberry ketone. J. Insect Physiol. 2018, 109, 41–46. [Google Scholar] [CrossRef]
  75. Roan, C.C.; Flitters, N.E.; Davis, C.J. Light intensity and temperature as factors limiting the mating of the oriental fruit fly. Ann. Entomol. Soc. Am. 1954, 47, 593–594. [Google Scholar] [CrossRef]
  76. Arakaki, N.; Kuba, H.; Soemori, H. Mating behavior of the oriental fruit fly, Dacus dorsalis Hendel (Diptera: Tephritidae). Appl. Entomol. Zool. 1984, 19, 42–51. [Google Scholar] [CrossRef] [Green Version]
  77. Kokkari, A.I.; Pliakou, O.D.; Floros, G.D.; Kouloussis, N.A.; Koveos, D.S. Effect of fruit volatiles and light intensity on the reproduction of Bactrocera (Dacus) oleae. J. Appl. Entomol. 2017, 141, 841–847. [Google Scholar] [CrossRef]
  78. Jang, E.B. Effects of mating and accessory gland injections on olfactory-mediated behavior in the female Mediterranean fruit fly, Ceratitis capitata. J. Insect Physiol. 1995, 41, 705–710. [Google Scholar] [CrossRef]
  79. Cornelius, M.L.; Nergel, L.; Duan, J.J.; Messing, R.H. Responses of female oriental fruit flies (Diptera: Tephritidae) to protein and host fruit odors in field cage and open field tests. Environ. Entomol. 2000, 29, 14–19. [Google Scholar] [CrossRef] [Green Version]
  80. Hagen, K.S.; Finney, G.L. A food supplement for effectively increasing the fecundity of certain Tephritid species. J. Econ. Entomol. 1950, 43, 735. [Google Scholar] [CrossRef]
  81. Perez-Staples, D.; Prabhu, V.; Taylor, P.W. Post-teneral protein feeding enhances sexual performance of Queensland fruit flies. Physiol. Entomol. 2007, 32, 225–232. [Google Scholar] [CrossRef]
  82. Shelly, T.E.; Edu, J.; Pahio, E. Influence of diet and methyl eugenol on the mating success of males of the oriental fruit fly, Bactrocera dorsalis (Diptera: Tephritidae). Fla. Entomol. 2005, 88, 307–313. [Google Scholar] [CrossRef]
  83. Shelly, T.E.; Edu, J.; Pahio, E. Condition-dependent mating success in male fruit flies: Ingestion of a pheromone precursor compensates for a low-quality diet. J. Insect Behav. 2007, 20, 347–365. [Google Scholar] [CrossRef]
  84. Weldon, C.W.; Perez-Staples, D.; Taylor, P.W. Feeding on yeast hydrolysate enhances attraction to cue-lure in Queensland fruit flies, Bactrocera tryoni. Entomol. Exp. Appl. 2008, 129, 200–209. [Google Scholar] [CrossRef]
  85. Shelly, T.E.; Dewire, A.L.M. Chemically mediated mating success in male oriental fruit flies (Diptera: Tephritidae). Ann. Entomol. Soc. Am. 1994, 87, 375–382. [Google Scholar] [CrossRef]
  86. Khan, M.A.M.; Shuttleworth, L.A.; Osborne, T.; Collins, D.; Gurr, G.M.; Reynolds, O.L. Raspberry ketone accelerates sexual maturation and improves mating performance of sterile male Queensland fruit fly, Bactrocera tryoni (Froggatt). Pest Manag. Sci. 2019, 75, 1942–1950. [Google Scholar] [CrossRef]
  87. Akter, H.; Mendez, V.; Morelli, R.; Pérez, J.; Taylor, P.W. Raspberry ketone supplement promotes early sexual maturation in male Queensland fruit fly, Bactrocera tryoni (Diptera: Tephritidae). Pest Manag. Sci. 2017, 73, 1764–1770. [Google Scholar] [CrossRef] [PubMed]
  88. Khan, M.A.M.; Manoukis, N.C.; Osborne, T.; Barchia, I.M.; Gurr, G.M.; Reynolds, O.L. Semiochemical mediated enhancement of males to complement sterile insect technique in management of the tephritid pest Bactrocera tryoni (Froggatt). Sci. Rep. 2017, 7, 13366. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Akter, H.; Adnan, S.; Morelli, R.; Rempoulakis, P.; Taylor, P.W. Suppression of cuelure attraction in male Queensland fruit flies provided raspberry ketone supplements as immature adults. PLoS ONE 2017, 12, e0184086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Shelly, T.E.; Villalobos, E.M. Cue lure and the mating behavior of male melon flies (Diptera: Tephritidae). Fla. Entomol. 1995, 78, 473–482. [Google Scholar] [CrossRef]
  91. Fezza, T.J.; Shelly, T.E. Raspberry ketone-supplemented diet has no effect on fitness parameters or lure responsiveness in male melon flies, Zeugodacus cucurbitae (Diptera: Tephritidae). J. Asia Pac. Entomol. 2018, 21, 1384–1388. [Google Scholar] [CrossRef]
  92. Wolfson, J.L. Bioassay techniques—An ecological perspective. J. Chem. Ecol. 1988, 14, 1951–1963. [Google Scholar] [CrossRef]
  93. Miyatake, T. Genetic changes of life history and behavioral traits during mass-rearing in the melon fly, Bactrocera cucurbitae (Diptera: Tephritidae). Res. Popul. Ecol. 1998, 40, 301–310. [Google Scholar] [CrossRef]
  94. Koyama, J.; Kakinohana, H.; Miyatake, T. Eradication of the melon fly, Bactrocera cucurbitae, in Japan: Importance of behavior, ecology, genetics, and evolution. Annu. Rev. Entomol. 2004, 49, 331–349. [Google Scholar] [CrossRef]
  95. Fleischer, J.; Pregitzer, P.; Breer, H.; Krieger, J. Access to the odor world: Olfactory receptors and their role for signal transduction in insects. Cell. Mol. Life Sci. 2018, 75, 485–508. [Google Scholar] [CrossRef]
  96. Montell, C. A taste of the Drosophila gustatory receptors. Curr. Opin. Neurobiol. 2009, 19, 345–353. [Google Scholar] [CrossRef] [Green Version]
  97. Rytz, R.; Croset, V.; Benton, R. Ionotropic Receptors (IRs): Chemosensory ionotropic glutamate receptors in Drosophila and beyond. Insect Biochem. Mol. Biol. 2013, 43, 888–897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Touhara, K.; Vosshall, L.B. Sensing odorants and pheromones with chemosensory receptors. Annu. Rev. Physiol. 2009, 71, 307–332. [Google Scholar] [CrossRef] [PubMed]
  99. Vogt, R.G.; Riddiford, L.M. Pheromone binding and inactivation by moth antennae. Nature 1981, 293, 161–163. [Google Scholar] [CrossRef] [PubMed]
  100. Brito, N.F.; Moreira, M.F.; Melo, A.C.A. A look inside odorant-binding proteins in insect chemoreception. J. Insect Physiol. 2016, 95, 51–65. [Google Scholar] [CrossRef]
  101. Butterwick, J.A.; del Mármol, J.; Kim, K.H.; Kahlson, M.A.; Rogow, J.A.; Walz, T.; Ruta, V. Cryo-EM structure of the insect olfactory receptor Orco. Nature 2018, 560, 447–452. [Google Scholar] [CrossRef]
  102. Abuin, L.; Bargeton, B.; Ulbrich, M.H.; Isacoff, E.Y.; Kellenberger, S.; Benton, R. Functional architecture of olfactory ionotropic glutamate receptors. Neuron 2011, 69, 44–60. [Google Scholar] [CrossRef] [Green Version]
  103. Benton, R.; Vannice, K.S.; Gomez-Diaz, C.; Vosshall, L.B. Variant ionotropic glutamate receptors as chemosensory receptors in Drosophila. Cell 2009, 136, 149–162. [Google Scholar] [CrossRef] [Green Version]
  104. Sato, K.; Pellegrino, M.; Nakagawa, T.; Nakagawa, T.; Vosshall, L.B.; Touhara, K. Insect olfactory receptors are heteromeric ligand-gated ion channels. Nature 2008, 452, 1002–1006. [Google Scholar] [CrossRef]
  105. Wicher, D.; Schäfer, R.; Bauernfeind, R.; Stensmyr, M.C.; Heller, R.; Heinemann, S.H.; Hansson, B.S. Drosophila odorant receptors are both ligand-gated and cyclic-nucleotide- activated cation channels. Nature 2008, 452, 1007–1011. [Google Scholar] [CrossRef]
  106. Sato, K.; Tanaka, K.; Touhara, K. Sugar-regulated cation channel formed by an insect gustatory receptor. Proc. Natl. Acad. Sci. USA 2011, 108, 11680–11685. [Google Scholar] [CrossRef] [Green Version]
  107. Wu, Z.; Zhang, H.; Wang, Z.; Bin, S.; He, H.; Lin, J. Discovery of chemosensory genes in the oriental fruit fly, Bactrocera dorsalis. PLoS ONE 2015, 10, e0129794. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Elfekih, S.; Chen, C.; Hsu, J.; Belcaid, M.; Haymer, D. Identification and preliminary characterization of chemosensory perception-associated proteins in the melon fly Bactrocera cucurbitae using RNA-seq. Sci. Rep. 2016, 6, 19112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Cheng, J.-F.; Yu, T.; Chen, Z.-J.; Chen, Y.-P.; Gao, L.; Zhang, W.-H.; Jiang, B.; Bai, X.; Liu, J.; Lu, Y.-Y. Identification and expression analysis of chemosensory genes in the citrus fruit fly Bactrocera (Tetradacus) minax. PeerJ Prepr. 2018, 6, e27297v1. [Google Scholar] [CrossRef]
  110. Wu, Z.; Kang, C.; Qu, M.; Chen, J.; Chen, M.; Bin, S.; Lin, J. Candidates for chemosensory genes identified in the Chinese citrus fly, Bactrocera minax, through a transcriptomic analysis. BMC Genom. 2019, 20, 646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Wu, Z.; Cui, Y.; Ma, J.; Qu, M.; Lin, J. Analyses of chemosensory genes provide insight into the evolution of behavioral differences to phytochemicals in Bactrocera species. Mol. Phylogenet. Evol. 2020, 151, 106858. [Google Scholar] [CrossRef] [PubMed]
  112. Kaissling, K.E. Kinetics of olfactory responses might largely depend on the odorant-receptor interaction and the odorant deactivation postulated for flux detectors. J. Comp. Physiol. A 2013, 199, 879–896. [Google Scholar] [CrossRef] [PubMed]
  113. Larter, N.K.; Sun, J.S.; Carlson, J.R. Organization and function of Drosophila odorant binding proteins. eLife 2016, 5, e20242. [Google Scholar] [CrossRef]
  114. Zhu, G.H.; Zheng, M.Y.; Sun, J.B.; Khuhro, S.A.; Yan, Q.; Huang, Y.; Syed, Z.; Dong, S.L. CRISPR/Cas9 mediated gene knockout reveals a more important role of PBP1 than PBP2 in the perception of female sex pheromone components in Spodoptera litura. Insect Biochem. Mol. Biol. 2019, 115, 103244. [Google Scholar] [CrossRef]
  115. Chen, X.F.; Xu, L.; Zhang, Y.X.; Wei, D.; Wang, J.J.; Jiang, H.B. Genome-wide identification and expression profiling of odorant-binding proteins in the oriental fruit fly, Bactrocera dorsalis. Comp. Biochem. Physiol. Part D Genom. Proteom. 2019, 31, 100605. [Google Scholar] [CrossRef]
  116. Liu, H.; Zhao, X.F.; Fu, L.; Han, Y.Y.; Chen, J.; Lu, Y.Y. BdorOBP2 plays an indispensable role in the perception of methyl eugenol by mature males of Bactrocera dorsalis (Hendel). Sci. Rep. 2017, 7, 15894. [Google Scholar] [CrossRef] [Green Version]
  117. Wu, Z.; Lin, J.; Zhang, H.; Zeng, X. BdorOBP83a-2 mediates responses of the oriental fruit fly to semiochemicals. Front. Physiol. 2016, 7, 452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Jang, E.B.; Light, D.M.; Flath, R.A.; Nagata, J.T.; Mon, T.R. Electroantennogram responses of the Mediterranean fruit fly, Ceratitis capitata, to the volatile constituents of nectarines. Entomol. Exp. Appl. 1989, 50, 7–19. [Google Scholar] [CrossRef]
  119. Siciliano, P.; He, X.L.; Woodcock, C.; Pickett, J.A.; Field, L.M.; Birkett, M.A.; Kalinova, B.; Gomulski, L.M.; Scolari, F.; Gasperi, G.; et al. Identification of pheromone components and their binding affinity to the odorant binding protein CcapOBP83a-2 of the Mediterranean fruit fly, Ceratitis capitata. Insect Biochem. Mol. Biol. 2014, 48, 51–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Papanicolaou, A.; Schetelig, M.F.; Arensburger, P.; Atkinson, P.W.; Benoit, J.B.; Bourtzis, K.; Castañera, P.; Cavanaugh, J.P.; Chao, H.; Childers, C.; et al. The whole genome sequence of the Mediterranean fruit fly, Ceratitis capitata (Wiedemann), reveals insights into the biology and adaptive evolution of a highly invasive pest species. Genome Biol. 2016, 17, 192. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Falchetto, M.; Ciossani, G.; Scolari, F.; Di Cosimo, A.; Nenci, S.; Field, L.M.; Mattevi, A.; Zhou, J.J.; Gasperi, G.; Forneris, F. Structural and biochemical evaluation of Ceratitis capitata odorant-binding protein 22 affinity for odorants involved in intersex communication. Insect Mol. Biol. 2019, 28, 431–443. [Google Scholar] [CrossRef]
  122. Li, Y.; Zhou, P.; Zhang, J.; Yang, D.; Li, Z.; Zhang, X.; Zhu, S.; Yu, Y.; Chen, N. Identification of odorant binding proteins in Carpomya vesuviana and their binding affinity to the male-borne semiochemicals and host plant volatiles. J. Insect Physiol. 2017, 100, 100–107. [Google Scholar] [CrossRef]
  123. Li, H.; Ren, L.; Xie, M.; Gao, Y.; He, M.; Hassan, B.; Lu, Y.; Cheng, D. Egg-surface bacteria are indirectly associated with oviposition aversion in Bactrocera dorsalis. Curr. Biol. 2020, 30, 4432–4440. [Google Scholar] [CrossRef]
  124. Yao, R.; Zhao, M.; Zhong, L.; Li, Y.; Li, D.; Deng, Z.; Ma, X. Characterization of the binding ability of the odorant binding protein BminOBP9 of Bactrocera minax to citrus volatiles. Pest Manag. Sci. 2020. [Google Scholar] [CrossRef]
  125. Nakagawa, T. Insect sex-pheromone signals mediated by specific combinations of olfactory receptors. Science 2005, 307, 1638–1642. [Google Scholar] [CrossRef]
  126. Miyazaki, H.; Otake, J.; Mitsuno, H.; Ozaki, K.; Kanzaki, R.; Chui-Ting Chieng, A.; Kah-Wei Hee, A.; Nishida, R.; Ono, H. Functional characterization of olfactory receptors in the oriental fruit fly Bactrocera dorsalis that respond to plant volatiles. Insect Biochem. Mol. Biol. 2018, 101, 32–46. [Google Scholar] [CrossRef] [Green Version]
  127. Liu, Y.; Liu, Y.; Jiang, X.; Wang, G. Cloning and functional characterization of three new pheromone receptors from the diamondback moth, Plutella xylostella. J. Insect Physiol. 2018, 107, 14–22. [Google Scholar] [CrossRef] [PubMed]
  128. Miura, N.; Nakagawa, T.; Tatsuki, S.; Touhara, K.; Ishikawa, Y. A male-specific odorant receptor conserved through the evolution of sex pheromones in Ostrinia moth species. Int. J. Biol. Sci. 2009, 5, 319–330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Miura, N.; Nakagawa, T.; Touhara, K.; Ishikawa, Y. Broadly and narrowly tuned odorant receptors are involved in female sex pheromone reception in Ostrinia moths. Insect Biochem. Mol. Biol. 2010, 40, 64–73. [Google Scholar] [CrossRef] [PubMed]
  130. Mitsuno, H.; Sakurai, T.; Murai, M.; Yasuda, T.; Kugimiya, S.; Ozawa, R.; Toyohara, H.; Takabayashi, J.; Miyoshi, H.; Nishioka, T. Identification of receptors of main sex-pheromone components of three Lepidopteran species. Eur. J. Neurosci. 2008, 28, 893–902. [Google Scholar] [CrossRef] [PubMed]
  131. Forstner, M.; Breer, H.; Krieger, J. A receptor and binding protein interplay in the detection of a distinct pheromone component in the silkmoth Antheraea polyphemus. Int. J. Biol. Sci. 2009, 5, 745–757. [Google Scholar] [CrossRef] [Green Version]
  132. Sakurai, T.; Nakagawa, T.; Mitsuno, H.; Mori, H.; Endo, Y.; Tanoue, S.; Yasukochi, Y.; Touhara, K.; Nishioka, T. Identification and functional characterization of a sex pheromone receptor in the silkmoth Bombyx mori. Proc. Natl. Acad. Sci. USA 2004, 101, 16653–16658. [Google Scholar] [CrossRef] [Green Version]
  133. Wanner, K.W.; Nichols, A.S.; Allen, J.E.; Bunger, P.L.; Garczynski, S.F.; Linn, C.E.; Robertson, H.M.; Luetje, C.W. Sex pheromone receptor specificity in the european corn borer moth, Ostrinia nubilalis. PLoS ONE 2010, 5, e8685. [Google Scholar] [CrossRef] [Green Version]
  134. Zheng, W.; Zhu, C.; Peng, T.; Zhang, H. Odorant receptor co-receptor Orco is upregulated by methyl eugenol in male Bactrocera dorsalis (Diptera: Tephritidae). J. Insect Physiol. 2012, 58, 1122–1127. [Google Scholar] [CrossRef]
  135. Liu, H.; Chen, Z.S.; Zhang, D.J.; Lu, Y.Y. BdorOR88a modulates the responsiveness to methyl eugenol in mature males of Bactrocera dorsalis (Hendel). Front. Physiol. 2018, 9, 987. [Google Scholar] [CrossRef]
  136. Tan, K.H.; Nishida, R. Sex pheromone and mating competition after methyl eugenol consumption in the Bactrocera dorsalis complex. In Fruit Fly Pests: A World Assessment of Their Biology and Management; McPheron, B.A., Steck, G.J., Eds.; St. Lucie Press: Delray Beach, FL, USA, 1996; pp. 147–153. ISBN 978-1574440140. [Google Scholar]
  137. Tan, K.H.; Tokushima, I.; Ono, H.; Nishida, R. Comparison of phenylpropanoid volatiles in male rectal pheromone gland after methyl eugenol consumption, and molecular phylogenetic relationship of four global pest fruit fly species: Bactrocera invadens, B. dorsalis, B. correcta and B. zonata. Chemoecology 2011, 21, 25–33. [Google Scholar] [CrossRef]
  138. Wee, S.; Tan, K. Male endogenous pheromonal component of Bactrocera carambolae (Diptera: Tephritidae) deterred gecko predation. Chemoecology 2005, 203, 199–203. [Google Scholar] [CrossRef]
  139. Wee, S.; Tan, K.; Nishida, R. Pharmacophagy of methyl eugenol by males enhances sexual selection of Bactrocera carambolae. J. Chem. Ecol. 2007, 33, 1272–1282. [Google Scholar] [CrossRef] [PubMed]
  140. Wee, S.L.; Tan, K.H. Female sexual response to male rectal volatile constituents in the fruit fly, Bactrocera carambolae (Diptera: Tephritidae). Appl. Entomol. Zool. 2005, 40, 365–372. [Google Scholar] [CrossRef] [Green Version]
  141. Tan, K. Sex pheromone components in defense of melon fly, Bactrocera cucurbitae against Asian house gecko, Hemidactylus frenatus. J. Chem. Ecol. 2000, 26, 697–704. [Google Scholar] [CrossRef]
  142. Ono, H.; Miyazaki, H.; Mitsuno, H.; Ozaki, K.; Kanzaki, R.; Nishida, R. Functional characterization of olfactory receptors in three Dacini fruit flies (Diptera: Tephritidae) that respond to 1-nonanol analogs as components in the rectal glands. Comp. Biochem. Physiol. Part B Biochem. Mol. Biol. 2020, 239, 110346. [Google Scholar] [CrossRef] [PubMed]
  143. Lu, T.; Qiu, Y.T.; Wang, G.; Kwon, J.Y.; Rutzler, M.; Kwon, H.W.; Pitts, R.J.; van Loon, J.J.A.; Takken, W.; Carlson, J.R.; et al. Odor coding in the maxillary palp of the malaria vector mosquito Anopheles gambiae. Curr. Biol. 2007, 17, 1533–1544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Kreher, S.A.; Mathew, D.; Kim, J.; Carlson, J.R. Translation of sensory input into behavioral output via an olfactory system. Neuron 2008, 59, 110–124. [Google Scholar] [CrossRef] [Green Version]
  145. Bohbot, J.D.; Dickens, J.C. Characterization of an enantioselective odorant receptor in the yellow fever mosquito Aedes aegypti. PLoS ONE 2009, 4, e7032. [Google Scholar] [CrossRef] [Green Version]
  146. Hallem, E.A.; Nicole Fox, A.; Zwiebel, L.J.; Carlson, J.R. Mosquito receptor for human-sweat odorant. Nature 2004, 427, 212–213. [Google Scholar] [CrossRef]
  147. Wang, G.; Carey, A.F.; Carlson, J.R.; Zwiebel, L.J. Molecular basis of odor coding in the malaria vector mosquito Anopheles gambiae. Proc. Natl. Acad. Sci. USA 2010, 107, 4418–4423. [Google Scholar] [CrossRef] [Green Version]
  148. Liu, Y.; Cui, Z.; Si, P.; Liu, Y.; Zhou, Q.; Wang, G. Characterization of a specific odorant receptor for linalool in the Chinese citrus fly Bactrocera minax (Diptera: Tephritidae). Insect Biochem. Mol. Biol. 2020, 122, 103389. [Google Scholar] [CrossRef] [PubMed]
  149. Liu, Y.; Cui, Z.; Wang, G.; Zhou, Q.; Liu, Y. Cloning and functional characterization of three odorant receptors from the Chinese citrus fly Bactrocera minax (Diptera: Tephritidae). Front. Physiol. 2020, 11, 246. [Google Scholar] [CrossRef] [PubMed]
  150. Liu, L.; Zhou, Q. Olfactory response of female Bactrocera minax to chemical components of the preference host citrus volatile oils. J. Asia. Pac. Entomol. 2016, 19, 637–642. [Google Scholar] [CrossRef]
  151. Ai, M.; Blais, S.; Park, J.Y.; Min, S.; Neubert, T.A.; Suh, G.S.B. Ionotropic glutamate receptors IR64a and IR8a form a functional odorant receptor complex in vivo in Drosophila. J. Neurosci. 2013, 33, 10741–10749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Choo, A.; Crisp, P.; Saint, R.; O’Keefe, L.V.; Baxter, S.W. CRISPR/Cas9-mediated mutagenesis of the white gene in the tephritid pest Bactrocera tryoni. J. Appl. Entomol. 2018, 142, 52–58. [Google Scholar] [CrossRef]
  153. Koh, T.W.; He, Z.; Gorur-Shandilya, S.; Menuz, K.; Larter, N.K.; Stewart, S.; Carlson, J.R. The Drosophila IR20a clade of ionotropic receptors are candidate taste and pheromone receptors. Neuron 2014, 83, 850–865. [Google Scholar] [CrossRef] [Green Version]
  154. Hanssen, B.L.; Park, S.J.; Royer, J.E.; Jamie, J.F.; Taylor, P.W.; Jamie, I.M. Systematic modification of zingerone reveals structural requirements for attraction of Jarvis’s Fruit Fly. Sci. Rep. 2019, 9, 19332. [Google Scholar] [CrossRef] [Green Version]
  155. Vosshall, L.B.; Laissue, P.P. The olfactory sensory map in Drosophila. Adv. Exp. Med. Biol. 2008, 628, 102–114. [Google Scholar] [CrossRef]
  156. Gao, Q.; Yuan, B.; Chess, A. Convergent projections of Drosophila olfactory neurons to specific glomeruli in the antennal lobe. Nat. Neurosci. 2000, 3, 780–785. [Google Scholar] [CrossRef]
  157. Couto, A.; Alenius, M.; Dickson, B.J. Molecular, anatomical, and functional organization of the Drosophila olfactory system. Curr. Biol. 2005, 15, 1535–1547. [Google Scholar] [CrossRef] [Green Version]
  158. Lin, T.; Li, C.; Liu, J.; Smith, B.H.; Lei, H.; Zeng, X. Glomerular organization in the antennal lobe of the oriental fruit fly Bactrocera dorsalis. Front. Neuroanat. 2018, 12, 71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Riabinina, O.; Task, D.; Marr, E.; Lin, C.C.; Alford, R.; O’Brochta, D.A.; Potter, C.J. Organization of olfactory centres in the malaria mosquito Anopheles gambiae. Nat. Commun. 2016, 7, 13010. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Teal, P.E.A.; Gomez-Simuta, Y.; Proveaux, A.T. Mating experience and juvenile hormone enhance sexual signaling and mating in male Caribbean fruit flies. Proc. Natl. Acad. Sci. USA 2000, 97, 3708–3712. [Google Scholar] [CrossRef] [PubMed]
  161. Adnan, S.M.; Mendez, V.; Morelli, R.; Akter, H.; Farhana, I.; Taylor, P.W. Dietary methoprene supplement promotes early sexual maturation of male Queensland fruit fly Bactrocera tryoni. J. Pest Sci. 2018, 91, 1441–1454. [Google Scholar] [CrossRef]
  162. Haq, I.U.; Cáceres, C.; Hendrichs, J.; Teal, P.E.A.; Stauffer, C.; Robinson, A.S. Methoprene modulates the effect of diet on male melon fly, Bactrocera cucurbitae, performance at mating aggregations. Entomol. Exp. Appl. 2010, 136, 21–30. [Google Scholar] [CrossRef]
  163. Shelly, T.E.; Nishimoto, J.I.; Edu, J. Influence of methoprene on the age-related mating propensity of males of the oriental fruit fly and the Mediterranean fruit fly (Diptera: Tephritidae). Fla. Entomol. 2009, 92, 193–198. [Google Scholar] [CrossRef]
  164. Jindra, M.; Palli, S.R.; Riddiford, L.M. The Juvenile hormone signaling pathway in insect development. Annu. Rev. Entomol. 2013, 58, 181–204. [Google Scholar] [CrossRef]
  165. Li, K.; Jia, Q.Q.; Li, S. Juvenile hormone signaling—A mini review. Insect Sci. 2019, 26, 600–606. [Google Scholar] [CrossRef]
  166. Akami, M.; Andongma, A.A.; Zhengzhong, C.; Nan, J.; Khaeso, K.; Jurkevitch, E.; Niu, C.Y.; Yuval, B. Intestinal bacteria modulate the foraging behavior of the oriental fruit fly Bactrocera dorsalis (Diptera: Tephritidae). PLoS ONE 2019, 14, e0210109. [Google Scholar] [CrossRef] [Green Version]
  167. Akami, M.; Ren, X.M.; Qi, X.; Mansour, A.; Gao, B.; Cao, S.; Niu, C.Y. Symbiotic bacteria motivate the foraging decision and promote fecundity and survival of Bactrocera dorsalis (Diptera: Tephritidae). BMC Microbiol. 2019, 19, 229. [Google Scholar] [CrossRef] [Green Version]
  168. Jose, P.A.; Ben-Yosef, M.; Jurkevitch, E.; Yuval, B. Symbiotic bacteria affect oviposition behavior in the olive fruit fly Bactrocera oleae. J. Insect Physiol. 2019, 117, 103917. [Google Scholar] [CrossRef]
  169. Potter, C.J.; Tasic, B.; Russler, E.V.; Liang, L.; Luo, L. Resource the Q system: A repressible binary system for transgene expression, lineage tracing, and mosaic analysis. Cell 2010, 141, 536–548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Ali, A.; Betz, J.F.; Riabinina, O.; Lahondére, C.; Potter, C.J.; Afify, A.; Betz, J.F.; Riabinina, O.; Lahondère, C.; Potter, C.J. Commonly used insect repellents hide human odors from Anopheles mosquitoes. Curr. Biol. 2019, 29, 3669–3680. [Google Scholar] [CrossRef]
  171. Matthews, B.J.; Vosshall, L.B. How to turn an organism into a model organism in 10 “easy” steps. J. Exp. Biol. 2020, 223 (Suppl. 1). [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Matthews, B.J.; Younger, M.A.; Vosshall, L.B. The ion channel ppk301 controls freshwater egg-laying in the mosquito Aedes aegypti. eLife 2019, 8, e43963. [Google Scholar] [CrossRef]
Figure 1. Structures of representative attractants and sex pheromones in Dacini fruit flies. A: Attractant; M-P: Male sex pheromone; F-P: Female sex pheromone.
Figure 1. Structures of representative attractants and sex pheromones in Dacini fruit flies. A: Attractant; M-P: Male sex pheromone; F-P: Female sex pheromone.
Insects 12 00106 g001
Figure 2. Structures of ligands for odorant biding proteins and olfactory receptors presented in Table 1 and Table 2. Stereocenters are shown by asterisks.
Figure 2. Structures of ligands for odorant biding proteins and olfactory receptors presented in Table 1 and Table 2. Stereocenters are shown by asterisks.
Insects 12 00106 g002
Figure 3. Metabolites of methyl eugenol in three Bactrocera-related species. The phylogenetic tree was constructed from mtDNA sequences by the maximum likelihood method. The accession numbers of the mtDNA sequences are indicated in the parentheses. ME: methyl eugenol; DMP: 2-allyl-4,5-dimethoxyphenol; Z-DMC: (Z)-3,4-dimethoxycinnamyl alcohol; E-CF: (E)-coniferyl alcohol; Z-CF: (Z)-coniferyl alcohol (Z-CF).
Figure 3. Metabolites of methyl eugenol in three Bactrocera-related species. The phylogenetic tree was constructed from mtDNA sequences by the maximum likelihood method. The accession numbers of the mtDNA sequences are indicated in the parentheses. ME: methyl eugenol; DMP: 2-allyl-4,5-dimethoxyphenol; Z-DMC: (Z)-3,4-dimethoxycinnamyl alcohol; E-CF: (E)-coniferyl alcohol; Z-CF: (Z)-coniferyl alcohol (Z-CF).
Insects 12 00106 g003
Table 1. OBPs and their high affinity odorants.
Table 1. OBPs and their high affinity odorants.
SpeciesNameAccession NumberLigandRef.
Bactrocera dorsalisBdorOBP56f-2 (BdorOBP2)KC559113Methyl eugenol[112]
BdorOBP83b (BdorOBP83a-2)KP743700Methyl eugenol[113]
BdorOBP99b (BdorOBP99a)KP743706Limonene[120]
Ocimene
BdorOBP84a-1MT4746231-Octen-3-ol[113]
Bactrocera minaxBminOBP9 (BminGOBP99a)KY463449β-Caryophyllene[120]
Ocimene
Myrcene
Ceratitis capitataCcapOBP22 (CcapOBP69a)NP_001295335.1(E,E)-α-Farnesene[117]
CcapOBP83a-2 XM_004523387.1(E,E)-α-Farnesene[115]
Carpomya vesuvianaCvesOBP1 KU975053β-Caryophyllene[118]
CvesOBP2KU975054(Z)-3-Hexenyl acetate[118]
CvesOBP3KU975055α-Farnesene[118]
CvesOBP4KX059394β-Caryophyllene[118]
CvesOBP5 KU975056Nonanal[118]
CvesOBP6KX059395Nonanal[118]
Table 2. Olfactory receptors that respond to semiochemicals in Dacini fruit flies.
Table 2. Olfactory receptors that respond to semiochemicals in Dacini fruit flies.
SpeciesNameAccession NumberDrosophila Homolog (1)E-ValueLigand (2)Ref.
Batrocera dorsalisBdorOR13aFX985890OR13a3e−1221-Octen-3-ol (+++)[122]
Batrocera dorsalisBdorOR63a-1KP743726OR63a1e−95Methyl eugenol (+)[131]
Batrocera dorsalisBdorOR74aFX985927OR74a5e−631-Nonanol (+++)[138]
6-Oxo-1-nonanol (++)
Batrocera dorsalisBdorOR82aFX985928OR82a2e−83Geranyl acetate (+++)[122]
Farnesenes (3) (+)
Linalyl acetate (++)
Batrocera dorsalisBdorOR88aKP743732OR88a1e−35Methyl eugenol (+++)[131]
Batrocera latifronsBlatOR74aFX986103OR74a2e−631-Nonanol (++)[138]
Bactrocera minaxBminOR3 (4)MN537976OR13a5e−1191-Octen-3-ol (+++)[145]
Bactrocera minaxBminOR12MN855530OR63a4e−109Methyl salicylate (+++)[145]
Benzaldehyde (++)
(Z)-3-Hexenyl acetate (++)
Butyl acrylate (++)
Butyl propionate (++)
1-Octanol (+)
(S)-(+)-Carvone (+)
Benzyl alcohol (+)
Bactrocera minaxBminOR16MN537977OR67c9e−331-Undecanol (++)[145]
Bactrocera minaxBminOR24MN537978OR7a1e−50Linalool (+++)[144]
Methyl eugenol (+)
Zeugodacus cucurbitaeZcucOR74aFX986226OR74a7e−641-Nonanol (+++)[138]
6-Oxo-1-nonanol (+)
(1) Homolog of Drosophila melanogaster was searched using the flybase (http://flybase.org). (2) Responses are indicated in the parentheses as the following currents responded to ligands at 100 µM. +++: ≥100 nA; ++: ≥50 nA; +: <50 nA. (3) Two typical natural products are shown in Figure 2. (4) This OR is registered as BminOR5 in the NCBI website.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ono, H.; Hee, A.K.-W.; Jiang, H. Recent Advancements in Studies on Chemosensory Mechanisms Underlying Detection of Semiochemicals in Dacini Fruit Flies of Economic Importance (Diptera: Tephritidae). Insects 2021, 12, 106. https://doi.org/10.3390/insects12020106

AMA Style

Ono H, Hee AK-W, Jiang H. Recent Advancements in Studies on Chemosensory Mechanisms Underlying Detection of Semiochemicals in Dacini Fruit Flies of Economic Importance (Diptera: Tephritidae). Insects. 2021; 12(2):106. https://doi.org/10.3390/insects12020106

Chicago/Turabian Style

Ono, Hajime, Alvin Kah-Wei Hee, and Hongbo Jiang. 2021. "Recent Advancements in Studies on Chemosensory Mechanisms Underlying Detection of Semiochemicals in Dacini Fruit Flies of Economic Importance (Diptera: Tephritidae)" Insects 12, no. 2: 106. https://doi.org/10.3390/insects12020106

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