Acessibilidade / Reportar erro

Characterization of the chemoreceptor repertoire of a highly specialized fly with comparisons to other Drosophila species

Abstract

To explore the diversity of scenarios in nature, animals have evolved tools to interact with different environmental conditions. Chemoreceptors are an important interface component and among them, olfactory receptors (ORs) and gustatory receptors (GRs) can be used to find food and detect healthy resources. Drosophila is a model organism in many scientific fields, in part due to the diversity of species and niches they occupy. The contrast between generalists and specialists Drosophila species provides an important model for studying the evolution of chemoreception. Here, we compare the repertoire of chemoreceptors of different species of Drosophila with that of D. incompta, a highly specialized species whose ecology is restricted to Cestrum flowers, after reporting the preferences of D. incompta to the odor of Cestrum flowers in olfactory tests. We found evidence that the chemoreceptor repertoire in D. incompta is smaller than that presented by species in the Sophophora subgenus. Similar patterns were found in other non-Sophophora species, suggesting the presence of underlying phylogenetic trends. Nevertheless, we also found autapomorphic gene losses and detected some genes that appear to be under positive selection in D. incompta, suggesting that the specific lifestyle of these flies may have shaped the evolution of individual genes in each of these gene families.

Keywords:
Olfaction; taste; odorant receptor; gustatory receptor

Introduction

Nature consists of many environments with different conditions and particularities. Animals living in each environment are in constant contact with many chemical compounds. Some of these compounds come from other living forms and are received and interpreted for many different purposes, such as finding sexual partners (Wyatt, 2003Wyatt TD (2003) Pheromones and animal behaviour: Communication by smell and taste. Cambridge University Press, Cambridge.; Davis, 2007Davis RL (2007) The scent of Drosophila sex. Neuron 54:14-16.; Kurtovic et al., 2007Kurtovic A, Widmer A and Dickson BJ (2007) A single class of olfactory neurons mediates behavioural responses to a Drosophila sex pheromone. Nature 446:542-546.), identifying predators or parasites (Ebrahim et al., 2015Ebrahim SA, Dweck HK, Stökl J, Hofferberth JE, Trona F, Weniger K, Rybak J, Seki Y, Stensmyr MC, Sachse S et al. (2015) Drosophila avoids parasitoids by sensing their semiochemicals via a dedicated olfactory circuit. PLoS Biol 13:e1002318.; Joseph and Carlson, 2015Joseph RM and Carlson JR (2015) Drosophila chemoreceptors: A molecular interface between the chemical world and the brain. Trends Genet 31:683-695.), or even finding and tasting food (Dahanukar et al., 2005Dahanukar A, Hallem EA and Carlson JR (2005) Insect chemoreception. Curr Opin Neurobiol 15:423-430.; Hallem et al., 2006Hallem EA, Dahanukar A and Carlson JR (2006) Insect odor and taste receptors. Annu Rev Entomol 51:113-135.). Regardless of the nature of the chemical signals, animals usually interpret them through chemoreceptors present in their cells. Because chemoreceptors are directly involved in this first contact with nature, they are subject to differential selection pressures during animal evolution (Whiteman and Pierce, 2008Whiteman NK and Pierce NE (2008) Delicious poison: Genetics of Drosophila host plant preference. Trends Ecol Evol 23:473-478.). In this sense, chemoreceptor genes generally evolve fast, leading to great diversification in genic repertoire (McBride, 2007McBride CS (2007) Rapid evolution of smell and taste receptor genes during host specialization in Drosophila sechellia. Proc Nat Acad Sci U S A 104:4996-5001.; McBride and Arguello, 2007McBride CS and Arguello JR (2007) Five Drosophila genomes reveal nonneutral evolution and the signature of host specialization in the chemoreceptor superfamily. Genetics 177:1395-1416.; Vieira and Rozas, 2011Vieira FG and Rozas J (2011) Comparative genomics of the odorant-binding and chemosensory protein gene families across the Arthropoda: Origin and evolutionary history of the chemosensory system. Genome Biol Evol 3:476-490.; Cande et al., 2013Cande J, Prud’homme B and Gompel N (2013) Smells like evolution: The role of chemoreceptor evolution in behavioral change. Curr Opin Neurobiol 23:152-158.). These patterns generally lead to an association between odor detection and individual ecological needs (Carey et al., 2010Carey AF, Wang G, Su CY, Zwiebel LJ and Carlson JR (2010) Odorant reception in the malaria mosquito Anopheles gambiae. Nature 464:66-71.).

Since the early 20th century, many Drosophila species have been used as model organisms in various scientific fields, among which we can highlight Drosophila melanogaster Meigen, 1830 (Markow, 2015Markow TA (2015) The natural history of model organisms: The secret lives of Drosophila flies. Elife 4:e06793.). Nevertheless, the genus includes 1,644 described species with many different traits, niches, and distributions (O’Grady and DeSalle, 2018O’Grady PM and DeSalle R (2018) Phylogeny of the genus Drosophila. Genetics 209:25.; TaxoDrosThe database on taxonomy of Drosophilidae (TaxoDros), The database on taxonomy of Drosophilidae (TaxoDros), https://www.taxodros.uzh.ch/ (accessed 24 September 2023).
https://www.taxodros.uzh.ch/ ...
). Many of these species have provided excellent models for evolutionary genomics, and there is a wealth of phylogenetic information spanning hundreds of species and at least thirty well-annotated genomes (Drosophila 12 Genomes Consortium et al., 2007Drosophila 12 Genomes Consortium; Clark AG, Eisen MB, Smith DR, Bergman CM, Oliver B, Markow TA, Kaufman TC, Kellis M, Gelbart W, Iyer VN et al. (2007) Evolution of genes and genomes on the Drosophila phylogeny. Nature 450:203-218.; Chen et al., 2014Chen ZX, Sturgill D, Qu J, Jiang H, Park S, Boley N, Suzuki AM, Fletcher AR, Plachetzki DC, FitzGerald PC et al. (2014) Comparative validation of the D. melanogaster modENCODE transcriptome annotation. Genome Res 24:1209-1223.; O’Grady and DeSalle, 2018O’Grady PM and DeSalle R (2018) Phylogeny of the genus Drosophila. Genetics 209:25.). Drosophila also provides several advantages while studying adaptations associated with different ecological habits, as it includes species with contrasting behaviors and niche breaths. Among these species we can identify generalists such as D. melanogaster, Drosophila ananassaeDoleschall (1858Doleschall CL (1858) Derde leijdrage tot de kennis der dipteran fauna van Nederlandsh Indie. Natrk. Tijdschr. Nederland Indie 17:73-128.), and Drosophila simulansSturtevant (1919Sturtevant AH (1919) A new species closely resembling Drosophila melanogaster. Psyche 26:153-155.) and other specialists such as Drosophila grimshawi Oldenberg, 1914, Drosophila erecta Tsacas and Lachaise (1974Tsacas L and Lachaise D (1974) Quatre nouvelles especes de la Cote-d’Ivoire du genre Drosophila, groupe melanogaster et discussion de l’origine du sous-groupe melanogaster (Diptera: Drosophilidae). Ann Univ Abidjan 7:193-211.) and Drosophila sechellia Tsacas and Bächli, 1981. Generalist habits allow exploring a variety of resources for feeding and oviposition, making such species resilient under conditions of limited resource availability or competition (Anholt, 2020Anholt RR (2020) Chemosensation and evolution of Drosophila host plant selection. Science 23:100799.). Conversely, specialists benefit from low competition for food or oviposition sites once they can overcome toxic or extreme niches (Anholt, 2020Anholt RR (2020) Chemosensation and evolution of Drosophila host plant selection. Science 23:100799.).

Among the different families of chemoreceptor genes, those coding for gustatory receptors (GR) and olfactory receptors (OR) are responsible for the detection of flavors or odorants, respectively (Clyne et al., 1999Clyne PJ, Warr CG, Freeman MR, Lessing D, Kim J and Carlson JR (1999) A novel family of divergent seven-transmembrane proteins: Candidate odorant receptors in Drosophila. Neuron 22:327-338.; Gao and Chess, 1999Gao Q and Chess A (1999) Identification of candidate Drosophila olfactory receptors from genomic DNA sequence. Genomics 60:31-39.; Scott et al., 2001Scott K, Brady Jr R, Cravchik A, Morozov P, Rzhetsky A, Zuker C and Axel R (2001) A chemosensory gene family encoding candidate gustatory and olfactory receptors in Drosophila. Cell 104:661-673.; Vosshall and Stocker, 2007Vosshall LB and Stocker RF (2007) Molecular architecture of smell and taste in Drosophila. Annu Rev Neurosci 30:505-533.; Martin et al., 2013Martin F, Boto T, Gomez‐Diaz C and Alcorta E (2013) Elements of olfactory reception in adult Drosophila melanogaster. Anat Rec 296:1477-1488.). In D. melanogaster, each of these two gene families harbors 60 genes (Robertson et al., 2003Robertson HM, Warr CG and Carlson JR (2003) Molecular evolution of the insect chemoreceptor gene superfamily in Drosophila melanogaster. Proc Nat Acad Sci U S A 100:14537-14542.). These genes are commonly involved in adaptations to different ecological contexts (Depetris-Chauvin et al., 2015Depetris-Chauvin A, Galagovsky D and Grosjean Y (2015) Chemicals and chemoreceptors: Ecologically relevant signals driving behavior in Drosophila. Front Ecol Evol 3:41.; Diaz et al., 2018Diaz F, Allan CW and Matzkin LM (2018) Positive selection at sites of chemosensory genes is associated with the recent divergence and local ecological adaptation in cactophilic Drosophila. BMC Evol Biol 18:144.). Such a pattern has been demonstrated, for example, for the pestiferous species Drosophila suzukii Matsumura that differs from other species of the D. melanogaster group due to its characteristic of exploring fresh fruit rather than rotting ones, which is reflected in the OR and GR repertoire (Hickner et al., 2016Hickner PV, Rivaldi CL, Johnson CM, Siddappaji M, Raster GJ and Syed Z (2016) The making of a pest: Insights from the evolution of chemosensory receptor families in a pestiferous and invasive fly, Drosophila suzukii. BMC Genom 17:648.). Moreover, specialist Drosophila species can present a fivefold higher dN/dS ratio for OR and GR genes (McBride, 2007McBride CS (2007) Rapid evolution of smell and taste receptor genes during host specialization in Drosophila sechellia. Proc Nat Acad Sci U S A 104:4996-5001.) and may also present faster rates of gene loss for these two gene families (McBride and Arguello, 2007McBride CS and Arguello JR (2007) Five Drosophila genomes reveal nonneutral evolution and the signature of host specialization in the chemoreceptor superfamily. Genetics 177:1395-1416.). Nevertheless, according to Gardiner et al. (2008Gardiner A, Barker D, Butlin RK, Jordan WC and Ritchie MG (2008) Drosophila chemoreceptor gene evolution: Selection, specialization and genome size. Mol Ecol 17:1648-1657.), endemism rather than niche specialization may account for much of this straightforward chemosensory gene loss.

Within Drosophila, the D. flavopilosa species group draws attention for its close association with flowers of Cestrum (Solanaceae), using this substrate as a unique resource for oviposition, larval development, and adult feeding (Brncic, 1966Brncic D (1966) Ecological and cytogenetic studies of Drosophila flavopilosa, a neotropical species living in Cestrum flowers. Evolution 1:16-29.; Robe et al., 2013Robe LJ, De Ré FC, Ludwig A and Loreto EL (2013) The Drosophila flavopilosa species group (Diptera, Drosophilidae) An array of exciting questions. Fly 7:59-69.). This group consists of 18 species (TaxoDros) divided into two subgroups and is considered part of the virilis-repleta radiation (Throckmorton, 1975Throckmorton LH (1975) The phylogeny, ecology and geography of Drosophila. In: King RC (ed). Handbook of genetics, vol 3, Plenum Press, New York, pp 421-469. ), which is currently considered a subgenus of the Drosophila genus (the subgenus Siphlodora) (Yassin, 2013Yassin A (2013) Phylogenetic classification of the Drosophilidae Rondani (Diptera): The role of morphology in the postgenomic era. Syst Entomol 38:349-364.). Nevertheless, the exact phylogenetic position of the group within the subgenus remains controversial, as it can be considered either a sister group of the D. annulimana species group (Robe et al., 2010Robe LJ, Loreto EL and Valente VL (2010) Radiation of the “Drosophila” subgenus (Drosophilidae, Diptera) in the Neotropics. J Zool Syst Evol Res 48:310-321.) or a sister group of the D. virilis or D. repleta species groups (de Ré et al., 2017De Ré FC, Robe LJ, Wallau GL and Loreto EL (2017) Inferring the phylogenetic position of the Drosophila flavopilosa group: Incongruence within and between mitochondrial and nuclear multilocus datasets. J Zool Syst Evol Res 55:208-221.). The flies of the flavopilosa group share some strict morphological and physiological adaptations for the use of Cestrum flowers: the pale yellow color of the body, which is cryptic in many Cestrum flowers; the medium to small size of the flies; oviposition at advanced stages of embryonic development, which matches with the ephemeral nature of Cestrum flowers; and the robust and full sting characteristic of ovipositors used to scarify the flower surface (Wheeler et al., 1962Wheeler MR, Takada H and Brncic D (1962) The flavopilosa species group of Drosophila. Univ Texas Publ Genetics 6205:395-414.; Pipkin et al., 1966Pipkin SB, Rodriguez RL and Leon J (1966) Plant host specificity among flower-feeding neotropical Drosophila (Diptera: Drosophilidae). Am Nat 100:135-156.; Brncic, 1983Brncic D (1983) The flavopilosa group of species as an example of flower-breeding species. In: Ashburner M, Carson HL and Thompson JN (eds) The genetics and biology of Drosophila. 3rd edition. Academic Press, New York, pp 360-377. ; Ludwig et al., 2002Ludwig A, Vidal NM, Loreto EL and Sepel LM (2002) Drosophila incompta development without flowers. Drosoph Inf Ser 85:40-41.; Robe et al., 2013Robe LJ, De Ré FC, Ludwig A and Loreto EL (2013) The Drosophila flavopilosa species group (Diptera, Drosophilidae) An array of exciting questions. Fly 7:59-69.). Although the occurrence of this group covers a large portion of the Neotropical region, most species are endemic to small patches (Robe et al., 2013Robe LJ, De Ré FC, Ludwig A and Loreto EL (2013) The Drosophila flavopilosa species group (Diptera, Drosophilidae) An array of exciting questions. Fly 7:59-69.). The most widespread species in the group is D. incompta, which occurs from Argentina to Mexico (TaxoDros).

The large number of chemoreceptors that can be found in Drosophila possibly enables the exploration of different resources, providing means to the establishment of different niches. To relate the preference of a species to its specific resource, choice tests are usually performed comparing different substrates. In these tests, Drosophila usually prefers citric substrates for oviposition (Dweck et al., 2013Dweck HK, Ebrahim SA, Kromann S, Bown D, Hillbur Y, Sachse S, Hansson BS and Stensmyr MC (2013) Olfactory preference for egg laying on citrus substrates in Drosophila. Curr Biol 23:2472-2480.), but this choice can vary depending on the niche of each species. Mansourian et al. (2018Mansourian S, Enjin A, Jirle EV, Ramesh V, Rehermann G, Becher PG, Pool JE and Stensmyr MC (2018) Wild African Drosophila melanogaster are seasonal specialists on marula fruit. Curr Biol 28:3960-3968.) demonstrated that D. melanogaster specimens are seasonal specialist of marula fruit (Sclerocarya birrea), preferring volatile marula odors over orange scents, although they preferred orange when tested with other fruit scents such as banana. Host switching between populations can also be assessed by comparing preferences for different odors. For example, for D. mojavensis, a cactophilic species that feeds on and breeds on various cactus species, population-specific preferences were previously detected (Date et al., 2013Date P, Dweck HK, Stensmyr MC, Shann J, Hansson BS and Rollmann SM (2013) Divergence in olfactory host plant preference in D. mojavensis in response to cactus host use. PLoS One 8:e70027.).

The aim of this study is to contribute to the knowledge of the patterns of molecular evolution associated with the chemoreceptor repertoire in different ecological contexts. To this end, we first examined the olfactory preferences of D. incompta to test if there are preferences for Cestrum flowers extracts over other odors. We then listed OR and GR genes that are part of the chemoreceptor repertoire in this species and compared them to homologous sequences in other Drosophila species to gain insight into the forces that might be shaping the evolution of chemoreceptor genes in specialized species.

Material and Methods

Sampling

Specimens of the Drosophila flavopilosa group were obtained from flowers of Cestrum strigilatum (Solanaceae) collected in the city of Santa Maria, southern Brazil (-29.710553 latitude and -53.717070 longitude). The sampled flowers were kept in the laboratory until the adult flies emerged. Drosophila incompta was distinguished from the other species of the group by its external morphology and male genitalia patterns, according to Wheeler et al. (1962Wheeler MR, Takada H and Brncic D (1962) The flavopilosa species group of Drosophila. Univ Texas Publ Genetics 6205:395-414.).

Olfactory choice tests

To test the flies’ preference for different odors, we performed an olfactory choice test (Table S1 Table S1 - Sample design of the five olfactory choice tests performed with Drosophila incompta (A, B, and C) and D. melanogaster (D and E), comparing odors of Cestrum × Banana (replicates A and D), Cestrum × Orange (replicates B), Cestrum × Brunfelsia uniflora (replicates C), and Banana × Orange (replicates E). ) with adult flies of D. incompta collected from the field, as described above. The tests were also performed with adults of D. melanogaster from the Oregon-R strain maintained in our laboratory, as a control to test if the olfactometer was working well. Flies were tested up to two days after emergence, until the total number of 10 flies was achieved per test. The elected odors were: 1) extract of Cestrum strigilatum L. flowers [1 ml of a solution of non-rotting flowers, prepared with five entire flowers (approximately 5 g) macerated in 5 ml of distilled water]; 2) extract of Brunfelsia uniflora (pohl) (Solanaceae) [1 ml of a solution of non-rotting flowers, prepared with two entire flowers (approximately 5 g) macerated in 5 ml of distilled water]; 3) extract of banana fruits (1 ml of a solution prepared with 5 g of a mature banana macerated in 5 ml of distilled water); and 4) extract of orange fruits (1 ml of a solution prepared with 5 g of a mature orange with peel macerated in 5 ml of distilled water). These odors were chosen considering the putatively strict adaptation of D. incompta to Cestrum spp. (Hofmann and Napp, 1984Hofmann PR and Napp M (1984) Genetic-environmental relationships in Drosophila incompta, a species of restricted ecology. Rev Bras Genet 7:21-39.; Santos and Vilela, 2005Santos RD and Vilela CR (2005) Breeding sites of Neotropical Drosophilidae (Diptera): IV. Living and fallen flowers of Sessea brasiliensis and Cestrum spp.(Solanaceae). Rev Bras Entom 49:544-551.), against the general profile of exploring different fermenting fruits presented by frugivorous Drosophila species (Sevenster and Van Alphen, 1993Sevenster JG and Van Alphen JJ (1993) A life history trade-off in Drosophila species and community structure in variable environments. J Anim Ecol 62:720-736.; Markow, 2015Markow TA (2015) The natural history of model organisms: The secret lives of Drosophila flies. Elife 4:e06793.). Conversely, Brunfelsia flowers were elected because they blossom concomitantly with Cestrum and have records of interaction with other Drosophila species (Frota-Pessoa, 1952Frota-Pessoa O (1952) Flower-feeding Drosophilidae. Drosoph Inf Ser 26:101-102.; Malogolowkin, 1953Malogolowkin C (1953) Sobre a genitália dos drosofilídeos. IV. A genitália masculina no subgênero Drosophila (Diptera, Drosophilidae). Rev Bras Biol 13:245-264.; Sepel et al., 2000Sepel LM, Golombieski RM, Napp M and Loreto EL (2000) Seasonal fluctuations of D. cestri and D. Incompta, two species of the flavopilosa group. Drosoph Inf Serv 83:122-126.; Schmitz and Valente, 2019Schmitz HJ and Valente VL (2019) The flower flies and the unknown diversity of Drosophilidae (Diptera): A biodiversity inventory in the Brazilian fauna. Pap Avulsos Zool 59:e20195945.).

In each case, the extracts were combined in pairs and dropped onto a new piece of cotton (Figure 1). Ten flies of the same sex were placed in the center of the olfactometer (following Fuyama, 1976Fuyama Y (1976) Behavior genetics of olfactory responses in Drosophila. I. Olfactometry and strain differences in Drosophila melanogaster. Behav Genet 6:407-420.) with the gates closed for two minutes for acclimation. After, gates were opened, and flies were allowed to choose a side. After five minutes, flies located on each side of the artifact were counted. All tests were conducted in a dark room with only red light available, in the morning. Each test was repeated at least six times per treatment, considering replicates performed either for males or with females, as both sexes showed similar patterns. Analysis of variance (ANOVA) was performed to assess the significance of the preferences.

Figure 1 -
Olfactometer model used to perform the olfactory preference tests. Legend: A, Extracts scent vessels; B, Flies waiting box; C, Preference tubes; D, Air inlet pipes; E, Air suction pipe; and F, Gates separating the flies from scents.

Molecular biology procedures

Total DNA was isolated from 20 males of D. incompta using a NucleoSpin Tissue XS kit (Macherey-Nagel, Düren, Germany) according to the manufacturer’s protocol. Only males were employed at this stage due to the cryptic nature of D. incompta regarding other closed related species, which hampers females’ identification (Wheeler et al., 1962Wheeler MR, Takada H and Brncic D (1962) The flavopilosa species group of Drosophila. Univ Texas Publ Genetics 6205:395-414.; Robe et al., 2013Robe LJ, De Ré FC, Ludwig A and Loreto EL (2013) The Drosophila flavopilosa species group (Diptera, Drosophilidae) An array of exciting questions. Fly 7:59-69.). Two approaches were used to obtain shotgun sequences of the whole genome. First, 100-bp single-end reads were obtained using a Solexa-Illumina HiSeq 2000 Next Generation Sequencing (NGS) instrument (Fasteris DNA Sequencing Service). Furthermore, we also generated 100-bp paired end reads in a Solexa-Illumina HiSeq 2000 instrument (Illumina Inc., San Diego, USA) at Macrogen Sequencing Service (Korea) using the provider’s protocols.

Other 20 adult males of D. incompta were directly collected from the medium with sampled flowers up to two hours after eclosion and further employed to isolate total adult RNA using TRIzol reagent (ThermoFisher Scientific), according to the protocol of Rio et al. (2010Rio DC, Ares M, Hannon GJ and Nilsen TW (2010) Purification of RNA using TRIzol (TRI reagent). Cold Spring Harb Protoc 2010:pdb-rot5439.). The mRNA was isolated using the mRNA Dynabeads C kit (Life Technologies) and libraries were prepared using the Seq RNA Total Ion V2 kit (ThermoFisher Scientific). Sequencing was performed using a Ion S5 sequencer (ThermoFisher Scientific).

The genome of D. incompta was assembled using a hybrid assembly approach for both paired-end and single-end libraries in hybridSPAdes (Antipov et al., 2016Antipov D, Korobeynikov A, McLean JS and Pevzner PA (2016) hybridSPAdes: An algorithm for hybrid assembly of short and long reads. Bioinformatics 32:1009-1015.). The performance of the de novo assembly was evaluated using Quast (Gurevich et al., 2013Gurevich A, Saveliev V, Vyahhi N and Tesler G (2013) QUAST: Quality assessment tool for genome assemblies. Bioinformatics 29:1072-1075.) and Busco 5.4.6 software (Simão et al., 2015Simão FA, Waterhouse RM, Ioannidis P, Kriventseva EV and Zdobnov EM (2015) BUSCO: Assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics 31:3210-3212.). In the last case, the Diptera OrthoDBv.10 (odb10) dataset was employed, with default parameters (Table S2 Table S2 - Metrics and statistics of Drosophila incompta genome and transcriptome draft assemblies. ). Transcriptome de novo assembly was performed using Trinity (Grabherr et al., 2011Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, Adiconis X, Fan L, Raychowdhury R, Zeng Q et al. (2011) Trinity: Reconstructing a full-length transcriptome without a genome from RNA-Seq data. Nat Biotechnol 29:644.) under default parameters in the Galaxy platform (Afgan et al., 2018Afgan E, Baker D, Batut B, van den Beek M, Bouvier D, Cech M, Chilton J, Clements D, Coraor N, Grüning BA, Guerler A et al. (2018) The Galaxy platform for accessible, reproducible and collaborative biomedical analyses: 2018 update. Nucleic Acids Res 46:W537-W544.).

Predicted coding sequences (cds) and protein genes for both genome and transcriptome drafts were identified using the Augustus 3.4.0 software (Stanke et al., 2006Stanke M, Keller O, Gunduz I, Hayes A, Waack S and Morgenstern B (2006) AUGUSTUS: Ab initio prediction of alternative transcripts. Nucleic Acids Res 34:W435-W439.). Both cds datasets were merged, and redundant sequences were removed using the cd-hit-2d program (Fu et al., 2012Fu L, Niu B, Zhu Z, Wu S and Li W (2012) CD-HIT: Accelerated for clustering the next-generation sequencing data. Bioinformatics 28:3150-3152.). This resulted in a non-redundant hybrid annotation of predicted cds, including genome and transcriptome sequences.

Recovery of OR and GR gene sequences in D. incompta

To retrieve sequences related to OR and GR chemoreceptors in the genome of D. incompta, at first, the repertory of olfactory and gustatory receptor genes of D. melanogaster (Robertson et al., 2003Robertson HM, Warr CG and Carlson JR (2003) Molecular evolution of the insect chemoreceptor gene superfamily in Drosophila melanogaster. Proc Nat Acad Sci U S A 100:14537-14542.) and D. virilis (Drosophila 12 Genomes Consortium et al., 2007) were downloaded from FlyBase (Thurmond et al., 2019Thurmond J, Goodman JL, Strelets VB, Attrill H, Gramates LS, Marygold SJ, Matthews BB, Millburn G, Antonazzo G, Trovisco V et al. (2019) FlyBase 2.0: The next generation. Nucleic Acids Res 47:59-65.), totaling 121 and 103 different genes sequences, respectively. Drosophila melanogaster was chosen because it is a model organism that has a well-annotated and well-studied repertoire of chemoreceptors. In turn, D. virilis entered as a query in our searches because it is closely related to D. incompta (Robe et al., 2010; De Ré et al., 2017De Ré FC, Robe LJ, Wallau GL and Loreto EL (2017) Inferring the phylogenetic position of the Drosophila flavopilosa group: Incongruence within and between mitochondrial and nuclear multilocus datasets. J Zool Syst Evol Res 55:208-221.). These sequences were arranged into two separate matrices subdivided by chemoreceptor gene family.

The pipeline employed in the InsectOR tool (Karpe et al., 2021Karpe SD, Tiwari V and Ramanathan S (2021) InsectOR-Webserver for sensitive identification of insect olfactory receptor genes from non-model genomes. PLoS One 16:e0245324.) was then used to predict and validate OR and GR sequences in D. incompta genome. For this task, Exonerate was first used to align the protein sequences recovered for D. melanogaster and D. virilis against the draft genome of D. incompta assembled by HybridSpades. These alignment files, the draft genome, and the protein queries were then submitted as input to the InsectOR web server. As a result, detailed information on each gene prediction and cds and protein fasta sequences were obtained. Finally, for an exhaustive approach, three additional strategies were employed: 1) OR and GR gene sequences that were not found with InsectOR were searched against the Augustus annotation using protein sequences of D. melanogaster and D. virilis as queries in BLASTp and tBLASTn searches; 2) these sequences were also used as seed against the raw reads of D. incompta genome and transcriptome, through the use of aTRAM 2.0 software (Allen et al., 2018Allen JM, LaFrance R, Folk RA, Johnson KP and Guralnick RP (2018) aTRAM 2.0: An improved, flexible locus assembler for NGS data. Evol Bioinform Online 14:1176934318774546.), which simultaneously searches and assembles orthologous gene sequences; 3) eggNOG-mapper (Huerta-Cepas et al., 2017Huerta-Cepas J, Forslund K, Coelho LP, Szklarczyk D, Jensen LJ, Von Mering C and Bork P (2017) Fast genome-wide functional annotation through orthology assignment by eggNOG-mapper. Mol Biol Evol 34:2115-2122.) was employed to functionally annotate the entire genome of D. incompta, using the paired-end Solexa-Illumina raw reads with default parameters.

Repertoire of OR and GR gene sequences in other Drosophila species

To recover OR and GR gene sequences from other species, we relied on the repertoires found by Gardiner et al. (2008Gardiner A, Barker D, Butlin RK, Jordan WC and Ritchie MG (2008) Drosophila chemoreceptor gene evolution: Selection, specialization and genome size. Mol Ecol 17:1648-1657.) for the first 12 Drosophila sequenced genomes (Drosophila 12 Genomes Consortium et al., 2007) (D. ananassae, D. erecta, D. grimshawi, D. mojavensis, D. melanogaster, D. persimilis, D. pseudoobscura, D. sechellia, D. simulans, D. virilis, D. willistoni, and D. yakuba). To complement the matrices and provide a broader glimpse about species with different levels of niche specialization, we also retrieved OR and GR genes from the genomes of other 19 Drosophila species (D. arizonae, D. biarmipes, D. bipectinata, D. busckii, D. buzzatii, D. elegans, D. eugracilis, D. ficusphila, D. hydei, D. kikkawai, D. mauritiana, D. miranda, D. navojoa, D. novamexicana, D. obscura, D. rhopaloa, D. serrate, D. suzukii, and D. takahashii) and outgroups (see below). These searches were performed through BLASTn in the NCBI database using OR and GR genes of D. virilis and D. melanogaster as queries. In each case, only the main copy of each gene was considered, and duplicated copies were not counted. Conversely, incomplete copies were eventually included in further analyses if they were the only available copies of the respective genes. We considered this approach a strategy to uncover the general evolutionary pattern of the two different gene families since it is not strongly affected by recent duplications or gene losses in other Drosophila species. Moreover, this approach is more robust to problems related to sequencing or assembling strategies.

Orthology assignments

To correctly assign orthologous relationships, the whole set of amino acid sequences of OR and GR gene families isolated for D. melanogaster and D. virilis were aligned against the putative orthologous copies retrieved for D. incompta. This alignment was performed under a PAM matrix, using the Clustal W algorithm (Thompson et al., 1994Thompson JD, Higgins DG and Gibson TJ (1994) CLUSTAL W: Improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res 22:4673-4680.), as available in Mega 7.0.26 software (Kumar et al., 2018Kumar S, Stecher G, Li M, Knyaz C and Tamura K (2018) MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol 35:1547-1549.). OR and GR matrixes were then employed to reconstruct a phylogenetic tree, under maximum likelihood (ML) in the IQTREE v.1.6.12 software (Nguyen et al., 2015Nguyen LT, Schmidt HA, Von Haeseler A and Minh BQ (2015) IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol Biol Evol 32:268-274.), using 10,000 ultra-fast bootstrap replicates. The ML tree was finally visualized in FigTree v1.4 (2018FigTree v1.4 (2018), Software, 4 (2018), Software, http://tree.bio.ed.ac.uk/software/figtree/ (accessed 4 March 2024).
http://tree.bio.ed.ac.uk/software/figtre...
) and rooted at the midpoint.

Evolutionary tests

For each individual gene, matrices were constructed containing all homologous sequences found in the different Drosophila species and outgroups (Table S3 Table S3 - Accession numbers (GenBank - NCBI) of the sequences used in evolutionary analysis matrices. ). Scaptodrosophila lebanonensis was generally used as an outgroup except in cases where no orthologous sequences were found for this species. In these cases, another species closely related to Drosophilidae for which orthologous sequences were available in NCBI was used as the outgroup: Calliphora stygia, Ctenopseustis obliquana, or Rhopalosiphum maidis for ORs; and Bactrocera latifrons, Musca domestica, or Ceratitis capitata for GRs. Codon-based alignments were then performed individually for each gene of the GR or OR gene families for which orthologous sequences of at least 441 bp were found for D. incompta using MACSE (Ranwez et al., 2011Ranwez V, Harispe S, Delsuc F and Douzery EJ (2011) MACSE: Multiple Alignment of Coding SEquences accounting for frameshifts and stop codons. PLoS One 6:e22594.).

Subsequently, two different selection tests were performed to evaluate the presence of negative or positive selection in the OR and GR genes of D. incompta: A) the adaptive branch-site REL test for episodic diversification (aBSREL; Smith et al., 2015Smith MD, Wertheim JO, Weaver S, Murrell B, Scheffler K and Kosakovsky Pond SL (2015) Less is more: An adaptive branch-site random effects model for efficient detection of episodic diversifying selection. Mol Biol Evol 32:1342-1353.) was performed to test whether the branch leading to D. incompta evolved under positive selection; and B) the fixed effects likelihood test (FEL; Pond and Frost, 2005Pond SLK and Frost SD (2005) Not so different after all: A comparison of methods for detecting amino acid sites under selection. Mol Biol Evol 22:1208-1222.) was used to derive the nonsynonymous substitution (α) and synonymous substitution (β) rates, testing whether alpha is significantly higher than beta at each site in the sequence of D. incompta. The latter test was applied only to matrices that showed a signal of positive selection for D. incompta genes in the aBSREL test. In both cases, individual gene codon alignments were submitted to the Datamonkey web platform (Weaver et al., 2018Weaver S, Shank SD, Spielman SJ, Li M, Muse SV and Kosakovsky Pond SL (2018) Datamonkey 2.0: A modern web application for characterizing selective and other evolutionary processes. Mol Biol Evol 35:773-777.), and analysis were performed with default parameters.

Results

Olfactory preference tests

Despite exploring a limited variety of odors, the olfactory preference tests showed a clear preference of D. incompta for Cestrum flowers compared to fermented banana extract (p < 0.0001). Drosophila incompta also preferred Cestrum flowers compared to fermented orange extract (p < 0.0001) and Brunfelsia uniflora (Pohl.) flowers (p = 0.0005). For D. melanogaster, the test of preference for Cestrum versus fermented banana extract showed no statistically significant difference. On the other hand, our results corroborated the results found by Mansourian et al. (2018Mansourian S, Enjin A, Jirle EV, Ramesh V, Rehermann G, Becher PG, Pool JE and Stensmyr MC (2018) Wild African Drosophila melanogaster are seasonal specialists on marula fruit. Curr Biol 28:3960-3968.) about the significant preference of D. melanogaster for orange over banana extracts (Figure 2, Table S1 Table S1 - Sample design of the five olfactory choice tests performed with Drosophila incompta (A, B, and C) and D. melanogaster (D and E), comparing odors of Cestrum × Banana (replicates A and D), Cestrum × Orange (replicates B), Cestrum × Brunfelsia uniflora (replicates C), and Banana × Orange (replicates E). ), showing that our olfactometer provides an accurate description of smell perception and preference for the tested extracts in the target species.

Figure 2 -
Results of ANOVAs performed to test the preference of D. incompta (A, B, and C) and D. melanogaster (D and E) for each pair of extract scents. A and D: Cestrum strigilatum against banana. F-Stat = 267.0286, p = 1.4772e-22 and F-Stat = 0.03866, p = 0.84555, respectively; B: C. strigilatum against orange. F-Stat = 80.2002, p = 1.246e-7; C: C. strigilatum against Brunfelsia uniflora. F-Stat = 25.71407, p = 0.0005; E: Orange against banana. F-Stat = 74.88571, p = 1.9737e-7. Y: axis represents the percentage of the preference of flies for both compared extracts at every paired comparison.

Genome-wide analyses

The de novo genome assembly on hybridSPAdes resulted in 187,091 contigs from 289,083,143 raw reads (including both, single and paired end sequences), which resulted in 2,437 matches in Busco (Table S2). This was contrasted by the 11,209 contigs recovered by Trinity for the transcriptome. The hybrid dataset including genome and transcriptome nonredundant cds recovered by Augustus resulted in 2,660 matches in Busco (Table S2 Table S2 - Metrics and statistics of Drosophila incompta genome and transcriptome draft assemblies. ).

Repertoire of genes related to OR and GR gene families

After confirming orthology in a ML tree reconstructed for each family (Figs. 1S Figure S1 - Phylogenetic tree reconstructed with amino acid sequences related to the odorant receptor (OR) gene family in D. incompta, D. melanogaster and D. virilis. and 2S Figure S2 - Phylogenetic tree reconstructed with amino acid sequences related to the gustatory receptor (GR) gene family in D. incompta, D. melanogaster and D. virilis. ), we recovered a total of 84 sequences related to different chemoreceptors genes in D. incompta, of which 40 and 44 were identified as GRs and ORs, respectively (Tables 1 and 2). For ORs, 42 were recovered by InsectOR, and 11 and 37 were found by similarity searches using D. melanogaster and D. virilis genes as queries, respectively (Table S4 Table S4 - Olfactory (OR, sheet A) and gustatory receptor (GR, sheet B) orthologues recovered for Drosophila incompta by each tool. ). For GRs, all 40 genes were recovered by InsectOR, whereas 6 and 37 were found by similarity searches using D. melanogaster and D. virilis genes as queries, respectively (Table S4 Table S4 - Olfactory (OR, sheet A) and gustatory receptor (GR, sheet B) orthologues recovered for Drosophila incompta by each tool. ). Among the genes not recovered by InsectOR, EggNog mapper and aTRAM were able to recover incomplete copies of two ORs.

Table 1 -
Number of olfactory receptors (OR) and gustatory receptors (GR) related genes retrieved for D. incompta and each of the other 30 Drosophila species.
Table 2 -
Repertoire of olfactory receptors (OR) and gustatory receptors (GR) found in D. incompta, with their respective CDS fragment size and evidence of positive selection.

Notwithstanding, results for the other Drosophila species were based on alternative methodologies, the number of chemoreceptor genes associated with ORs genes varied from 38 (in D. busckii) to 60 (in D. melanogaster, D. sechellia and D. simulans) (Table 1, Table S5 Table S5 - List of presence (X on the table) or absence (- on the table) of olfactory (OR, sheet A) and gustatory receptor (GR, sheet B) orthologues retrieved for each of the analyzed Drosophila species. ). GRs varied from minimum values of 34 (in D. busckii) to maximum values of 60 (in D. melanogaster, D. sechellia, and D. simulans).

Regarding the distribution of gene losses or gains, evidence of phylogenetic signal was found for both ORs and GRs (Figures 3 and 4). In this context, it was interesting to see that the number of sequences found for both GR and OR was generally lower in non-Sophophora species. Nevertheless, two specific gene losses that are either autapomorphic or homoplastic for D. incompta were found for the genes ORs Or1a, and Or94b (Fig. 3).

Figure 3 -
Presence (blue boxes) and absence (red boxes) of 38 olfactory genes (OR) analyzed across the Drosophila phylogeny, modified from Robe et al. (2010Robe LJ, Loreto EL and Valente VL (2010) Radiation of the “Drosophila” subgenus (Drosophilidae, Diptera) in the Neotropics. J Zool Syst Evol Res 48:310-321.), Yang et al. (2012Yang Y, Hou ZC, Qian YH, Kang H and Zeng QT (2012) Increasing the data size to accurately reconstruct the phylogenetic relationships between nine subgroups of the Drosophila melanogaster species group (Drosophilidae, Diptera). Mol Phylogenet Evol 62:214-223.), Yassin (2013Yassin A (2013) Phylogenetic classification of the Drosophilidae Rondani (Diptera): The role of morphology in the postgenomic era. Syst Entomol 38:349-364.), Hjelmen and Johnston (2017Hjelmen CE and Johnston JS (2017) The mode and tempo of genome size evolution in the subgenus Sophophora. PLoS One 12:e0173505.), and O’Grady and DeSalle (2018O’Grady PM and DeSalle R (2018) Phylogeny of the genus Drosophila. Genetics 209:25.). Orthologues that were found in all analyzed species were not depicted on this graph.

Figure 4 -
Presence (blue boxes) and absence (red boxes) of 31 gustatory genes (GR ) analyzed across the Drosophila phylogeny, modified from Robe et al. (2010Robe LJ, Loreto EL and Valente VL (2010) Radiation of the “Drosophila” subgenus (Drosophilidae, Diptera) in the Neotropics. J Zool Syst Evol Res 48:310-321.), Yang et al. (2012Yang Y, Hou ZC, Qian YH, Kang H and Zeng QT (2012) Increasing the data size to accurately reconstruct the phylogenetic relationships between nine subgroups of the Drosophila melanogaster species group (Drosophilidae, Diptera). Mol Phylogenet Evol 62:214-223.), Yassin (2013Yassin A (2013) Phylogenetic classification of the Drosophilidae Rondani (Diptera): The role of morphology in the postgenomic era. Syst Entomol 38:349-364.), Hjelmen and Johnston (2017Hjelmen CE and Johnston JS (2017) The mode and tempo of genome size evolution in the subgenus Sophophora. PLoS One 12:e0173505.), and O’Grady and DeSalle (2018O’Grady PM and DeSalle R (2018) Phylogeny of the genus Drosophila. Genetics 209:25.). Orthologs that were found in all analyzed species were not depicted on this graph.

Selection tests

Of the 84 chemoreceptor genes found for D. incompta, 81 had a continuous coding sequence of at least 441 bp (147 aa) and were included in further evolutionary analysis. Three ORs were not included due to their fragmentary sequences.

Of the 81 chemoreceptors matrices used for selection tests, six presented signals of positive selection on the aBSREL approach (Table 2): OR19a, OR42a, OR63a, GR57a, GR64e and GR77a (p < 0.05). These six matrices were also tested on FEL analysis and presented signals of positive selection ranging from one site for GR64e to 17 sites for GR63a, according to the following distribution: four sites for GR77a, five sites for OR19a, six sites for GR57a and eight sites for OR42a.

Discussion

Drosophila has been widely used as a model to study the forces shaping the evolution of chemoreceptor gene families in different ecological contexts. Nevertheless, the role that selection and genetic drift have played in the evolution of these genes in specialized species is still controversial (McBride, 2007Nguyen LT, Schmidt HA, Von Haeseler A and Minh BQ (2015) IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol Biol Evol 32:268-274.; McBride and Arguello, 2007McBride CS and Arguello JR (2007) Five Drosophila genomes reveal nonneutral evolution and the signature of host specialization in the chemoreceptor superfamily. Genetics 177:1395-1416.; Gardiner et al., 2008Gardiner A, Barker D, Butlin RK, Jordan WC and Ritchie MG (2008) Drosophila chemoreceptor gene evolution: Selection, specialization and genome size. Mol Ecol 17:1648-1657.). This study contributes to the understanding of the patterns of molecular evolution exhibited by the OR and GR gene families in a species with a strict niche of the D. flavopilosa group, D. incompta, and provides several insights into understanding the general patterns of the entire Drosophila genus.

Molecular evolution of OR and GR gene families in D. incompta

Although it has been previously demonstrated that D. incompta can only explore Cestrum flowers as feeding and breeding resources (Sepel et al., 2000Sepel LM, Golombieski RM, Napp M and Loreto EL (2000) Seasonal fluctuations of D. cestri and D. Incompta, two species of the flavopilosa group. Drosoph Inf Serv 83:122-126.; Santos and Vilela, 2005Santos RD and Vilela CR (2005) Breeding sites of Neotropical Drosophilidae (Diptera): IV. Living and fallen flowers of Sessea brasiliensis and Cestrum spp.(Solanaceae). Rev Bras Entom 49:544-551.; Robe et al., 2013Robe LJ, De Ré FC, Ludwig A and Loreto EL (2013) The Drosophila flavopilosa species group (Diptera, Drosophilidae) An array of exciting questions. Fly 7:59-69.), the ethological or physiological characteristics associated with this restricted ecology have not yet been recognized. Our results regarding the olfactory preferences of D. incompta suggest a chemosensory basis for this condition, as indicated by the clear preference for extracts of Cestrum flowers compared to other scents (Brunfelsia, banana, or orange extract scents). It is generally expected that at least part of this specialization is related to chemosensory changes, ultimately linked to genetic alterations. This hypothesis is supported by some of the results presented here for the OR and GR gene families in D. incompta, such as some autapomorphic or homoplastic gene losses suggested for some genes, in contrast to the signals of positive selection detected for others.

It is well known that both OR and GR gene families frequently evolve through gene duplication, gene loss, and pseudogenization (Gardiner et al., 2008Gardiner A, Barker D, Butlin RK, Jordan WC and Ritchie MG (2008) Drosophila chemoreceptor gene evolution: Selection, specialization and genome size. Mol Ecol 17:1648-1657.; Benton, 2015Benton R (2015) Multigene family evolution: Perspectives from insect chemoreceptors. Trends Ecol Evol 30:590-600.). In particular, it has been suggested that during host specialization there may be a general trend toward loss of some chemosensory genes with concomitant positive selection on others (McBride, 2007McBride CS (2007) Rapid evolution of smell and taste receptor genes during host specialization in Drosophila sechellia. Proc Nat Acad Sci U S A 104:4996-5001.; McBride and Arguello, 2007McBride CS and Arguello JR (2007) Five Drosophila genomes reveal nonneutral evolution and the signature of host specialization in the chemoreceptor superfamily. Genetics 177:1395-1416.). For D. incompta, orthologous sequences were found for 84 of the 120 chemoreceptor genes examined in this study. Different strategies were employed for this task, among which InsectOR revealed the most sensitive, being able to recover 42 of the 44 ORs, and all the GRs, through the use of a mixed query matrix containing D. melanogaster and D. virlis chemoreceptor genes. Consistency of these results were further revealed by the ML tree reconstructed with amino acid sequences for each family, where most identifications performed with InsectOR recovered monophyletic groups. In fact, the only exceptions to this pattern were related to polytomous or poorly supported relationships, in which putatively orthologous sequences encompassed short branches that were located as adjacent in the three. Such results are probably an outcome related to the insufficient phylogenetic signal presented by some sequences, or to the positive selection detected for others (see below), since none D. incompta sequence appeared as nested within clades compounded by alternative orthogroups.

Although the small number of genes recovered for D. incompta seems to be consistent with the hypothesis suggesting a general trend of gene loss in specialized species (McBride and Arguello, 2007McBride CS and Arguello JR (2007) Five Drosophila genomes reveal nonneutral evolution and the signature of host specialization in the chemoreceptor superfamily. Genetics 177:1395-1416.; Croset et al., 2010Croset V, Rytz R, Cummins SF, Budd A, Brawand D, Kaessmann H, Gibson TJ and Benton R (2010) Ancient protostome origin of chemosensory ionotropic glutamate receptors and the evolution of insect taste and olfaction. PLoS Genet 6:e1001064.), this pattern should be interpreted with caution. At first, it is important to consider that at least some of these putative losses could be an outcome of methodological sources of error related to sequencing coverage, genome assembling or annotation. Moreover, a closer inspection to the general phylogenetic profile reveals that most of the missing chemoreceptor genes detected in D. incompta appear to have been lost in the ancestors (see below), reflecting a phylogenetic signal or a historical contingency. Indeed, only two of the 36 missing genes of D. incompta provided evidence of autapomorphic or homoplastic gene loss in the target species. This is true, for example, for the putative loss of OR1a and OR94b in the D. incompta genome, which were previously shown to respond to specific odorants, like the aldehyde (E)-2-hexenal (Fishilevich et al., 2005Fishilevich E, Domingos AI, Asahina K, Naef F, Vosshall LB and Louis M (2005) Chemotaxis behavior mediated by single larval olfactory neurons in Drosophila. Curr Biol 15:2086-2096. ) and the phenol 4-methylphenol (Kreher et al., 2005Kreher SA, Kwon JY and Carlson JR (2005) The molecular basis of odor coding in the Drosophila larvae. Neuron 46:445-456.), respectively. Interestingly, neither of these compounds seem to have been isolated in Cestrum yet (Alrabayah et al., 2024Alrabayah IN, El Hawary SS, El-Kadder EM, Essa A and Elraey M (2024) Genus Cestrum therapeutic potential: An updated review of its phytochemical, pharmacological, and morphological features. Egyptian J Chem 67:375-393.).

The effects of differential selection pressure in the repertoire of chemoreceptor genes of D. incompta could be directly tested in the positive selection tests, when tree genes of both the GR and the OR gene families showed signals of positive selection. Interestingly, two of the six genes exhibiting such signals (OR63a and GR57a) were found in all other species examined, suggesting that their products serve some important functions that are somehow conserved across the Drosophila phylogeny. Otherwise, GR64e seems to be missed only in D. busckii, and was previously shown to confer responsiveness to glycerol, a byproduct of yeast fermentation, affecting feeding preferences in Drosophila (Wisotsky et al., 2011Wisotsky Z, Medina A, Freeman E and Dahanukar A (2011) Evolutionary differences in food preference rely on Gr64e, a receptor for glycerol. Nat Neurosci 14:1534-1541.). Since yeasts are common inhabitants of many angiosperm nectars (Pozo et al., 2009Pozo MJ, de Vega C, Canto A and Herrera CM (2009) Presence of yeasts in floral nectar is consistent with the hypothesis of microbial-mediated signaling in plant-pollinator interactions. Plant Signal Behav 4:1102-1104.), this result raises the possibility that yeasts are the effective feeding resources used by D. incompta larvae and/or adults in Cestrum flowers, suggesting that the D. flavopilosa group may not have diverged so much from the ancestral Drosophilidae feeding habits (Throckmorton, 1975Throckmorton LH (1975) The phylogeny, ecology and geography of Drosophila. In: King RC (ed). Handbook of genetics, vol 3, Plenum Press, New York, pp 421-469. ).

Evolution of OR and GR genes in Drosophila

By increasing the number of species studied, we improved our understanding of the evolutionary patterns of the OR and GR gene families across the phylogeny of Drosophila. In this sense, we found a large variation in the number of chemoreceptor genes among species, with the higher and lower numbers of OR and GR genes generally found for species of the subgenus Sophophora, especially for species of the melanogaster subgroup, and for D. busckii, respectively. Because D. melanogaster was one of the species used as a query or seed in our searches, the higher number of chemoreceptor genes found for this species and other closely related taxa may be a methodological artifact. The lower number of orthologous sequences found for D. busckii could also be a biased result, as this species is not closely related to either of the two species used as a query or seed in our searches (Yassin, 2013Yassin A (2013) Phylogenetic classification of the Drosophilidae Rondani (Diptera): The role of morphology in the postgenomic era. Syst Entomol 38:349-364.; O’Grady and DeSalle, 2018O’Grady PM and DeSalle R (2018) Phylogeny of the genus Drosophila. Genetics 209:25.). Moreover, chemoreceptor annotation for this species has not yet received special attention. Nevertheless, such a pattern does not explain the results found for species closely related to D. virilis, since the last species not only presents a well-annotated genome but was also used as a query or seed in our searches. Even so, the number of orthologous sequences found in this species is close to the lower range for the OR and GR genes (50 and 40, respectively). Another point to consider is that we are only looking at one copy of each gene, which means that the scenario presented focuses on more ancient events of gene duplication. In this sense, the number of genes may be somewhat underestimated for some species. This is the case for D. grimshawi, which has already been shown to have multiple copies of some loci associated with recent gene duplications (Guo and Kim, 2007Guo S and Kim J (2007) Molecular evolution of Drosophila odorant receptor genes. Mol Biol Evol 24:1198-1207.; Nozawa and Nei, 2007Nozawa M and Nei M (2007) Evolutionary dynamics of olfactory receptor genes in Drosophila species. Proc Natl Acad Sci U S A 104:7122-7127.; Gardiner et al., 2008Gardiner A, Barker D, Butlin RK, Jordan WC and Ritchie MG (2008) Drosophila chemoreceptor gene evolution: Selection, specialization and genome size. Mol Ecol 17:1648-1657.). Even so, it is important to mention that our OR family annotation generally agrees with Ramasamy et al. (2016Ramasamy S, Ometto L, Crava CM, Revadi S, Kaur R, Horner DS, Pisani D, Dekker T, Anfora G and Rota-Stabelli O (2016) The evolution of olfactory gene families in Drosophila and the genomic basis of chemical-ecological adaptation in Drosophila suzukii. Genome Biol Evol 8:2297-2311.) concerning the number of genes for each species. The few divergences may be explained by the use of different annotation tools, albeit the main approach of both studies was the BLAST analyses.

Despite some putative sources of bias, the general picture suggests that the distribution of OR and GR genes across the Drosophila phylogeny may reflect a phylogenetic signal related to gene gains or losses in ancestral populations that have been stably inherited over millions of years. These conclusions are further supported when the presence or absence of each of the 16 OR and 20 GR genes of D. melanogaster not found in D. incompta are evaluated in a phylogenetic context. Of the ORs not found in D. incompta, six were also not found in any non-Sophophora species (OR7a, OR33a, OR33b, OR43b, OR59c, and OR65a), and three were found in only one species of the subgenus Siphlodora (OR65c in D. arizonae, and OR85a in D. mojavensis) or the subgenus Dorsilopha (OR45a in D. busckii). This pattern also holds for the gene family GR, where most genes not found in D. incompta appear to be largely restricted to the Sophophora subgenus. Indeed, 16 GR genes (GR10b, GR22a, GB22b, GR22c, GR22d, GR22f, GR36a, GR36b, GR36c, GR59a, GR59c, GR92a, GR93b, GR93d, GR98c, and GR98d) were not found in any non-Sophophora species, and two of them (GR22c and GR93d) were present in only five Sophophora species. These absences may indicate ancestral gene losses in non-Sophophora species, gene gains in Sophophora species, or even high divergence compared to query sequences reaching the limit of sensitivity of BLAST. In fact, comparisons involving the mean number of chemoreceptor genes between Sophophora with 57.36 (σ = 2.44) for OR and 53.82 (σ = 3.96) for GR genes, and non-Sophophora species with 44.67 (σ = 3.53) for OR and 39 (σ = 2.31) for GR genes, further agrees with these putative phylogenetic trends (Figure 5). Nevertheless, there are interesting exceptions to this rule, like D. willistoni and D. busckii, which stands out as a Sophophora and a non-Sophophora species with a lower number of OR and GR genes respectively.

Figure 5 -
Box plot graph comparing the total number of orthologs found for each of the two gene families by different Drosophila lineages (A: Sophophora species and B: non-Sophophora species). The bold horizontal line in the boxes represents the median, while the dashed lines represent the mean with standard deviation. Dots represent the outliers.

Because species in the melanogaster group often have the greater number of OR and GR genes (60 for both in D. melanogaster), it is tempting to speculate that, in at least some cases, the number of chemoreceptor genes also defines evolutionary potentials related not only to niche breadth but also to the distribution range of individual species within each of the major lineages of Drosophila. Indeed, several Sophophora species are recognized as generalists and cosmopolitans, such as D. ananassae (Singh, 1996Singh BN (1996) Population and behaviour genetics of Drosophila ananassae. Genetica 97:321-329.), D. kikkawai (Ramniwas and Kajla, 2012Ramniwas S and Kajla B (2012) Divergent strategy for adaptation to drought stress in two sibling species of montium species subgroup: Drosophila kikkawai and Drosophila leontia. J Insect Physiol 58:1525-1533.), D. mauritiana (Tsacas and David, 1974Tsacas L and David J (1974) Drosophila mauritiana n. sp. du groupe melanogaster de l’Ile Maurice [Dipt. Drosophilidae]. Bull Soc Entomol 79:42-46.), D. melanogaster (David and Capy, 1988David JR and Capy P (1988) Genetic variation of Drosophila melanogaster natural populations. Trends Genet 4:106-111.; Lachaise and Silvain, 2004Lachaise D and Silvain JF (2004) How two Afrotropical endemics made two cosmopolitan human commensals: The Drosophila melanogaster-D. simulans palaeogeographic riddle. Genetica 120:17-39.; Markow, 2015Markow TA (2015) The natural history of model organisms: The secret lives of Drosophila flies. Elife 4:e06793.), D. obscura (Begon, 1975Begon M (1975) The relationships of Drosophila obscura Fallen and D. subobscura Collin to naturally-occurring fruits. Oecologia 20:255-277.; Markow and O’Grady, 2008Markow TA and O’Grady P (2008) Reproductive ecology of Drosophila. Funct Ecol 747-759.), D. persimilis (Wogaman and Seiger, 1983Wogaman DJ and Seiger MB (1983) Light intensity as a factor in the choice of an oviposition site by Drosophila pseudoobscura and Drosophila persimilis. Can J Genet Cytol 25:370-377.), D. pseudoobscura (Carson, 1971Carson HL (1971) Speciation and the founder principle. Stadler Genetics Symposium 3:51-70.), D. serrata (Kelemen and Moritz, 1999Kelemen L and Moritz C (1999) Comparative phylogeography of a sibling pair of rainforest Drosophila species (Drosophila serrata and D. birchii). Evolution 53:1306-1311.; Schiffer and McEvey, 2006Schiffer M and McEvey SF (2006) Drosophila bunnanda: A new species from northern Australia with notes on other Australian members of the montium subgroup (Diptera: Drosophilidae). Zootaxa 1333:1-23.), D. simulans (Lemeunier et al., 1986Lemeunier F, Tsacas L, David J and Ashburner M (1986) The melanogaster species group. In: Ashburner M, Thompson JR Jr and Carson HL (eds) The genetics and biology of Drosophila, vol. 3, Academic Press, New York , pp 147-256.), D. suzukii (Asplen et al., 2015Asplen MK, Anfora G, Biondi A, Choi DS, Chu D, Daane KM, Gibert P, Gutierrez AP, Hoelmer KA, Hutchison WD et al. (2015) Invasion biology of spotted wing Drosophila (Drosophila suzukii): A global perspective and future priorities. J Pest Sci 88:469-494.; Biondi et al., 2016Biondi A, Traugott M and Desneux N (2016) Special issue on Drosophila suzukii: From global invasion to sustainable control. J Pest Sci 89:603-604.), and D. willistoni (Spassky et al., 1971Spassky B, Richmond RC, Perez-Salas S, Pavlovsky O, Mourao CA, Hunter AS, Hoenigsberg H, Dobzhansky T and Ayala FJ (1971) Geography of the sibling species related to Drosophila willistoni, and of the semispecies of the Drosophila paulistorum complex. Evolution 25:129-143., Cordeiro and Winge, 1995Cordeiro AR and Winge H (1995) Levels of evolutionary divergence of Drosophila willistoni sibling species. In: Levine L (ed), Genetics of Natural Populations of Theodosius Dobzhansky. 1st edition. Columbia University Press, New York, pp. 241-261 ; Powell, 1997Powell JR (1997) Progress and prospects in evolutionary biology: The Drosophila model. Oxford University Press, Oxford, 576 p.), a pattern that appears to be less common in non-Sophophora species. Nevertheless, this hypothesis also has some important exceptions, as in the case of D. sechellia, a Sophophora species with 120 chemoreceptor genes which is endemic to the Seychelles islands in the Indian Ocean and is specialized to a single resource, the fruit of Morinda citrifolia (Tsacas and Bächli, 1981Tsacas L and Bächli G (1981) Drosophila sechellia, n. sp., huitieme espece du sous-groupe melanogaster des iles Sechelles (Diptera, Drosophilidae). Rev Fr Entomol 3:146-150.; Louis and David, 1986Louis J and David JR (1986) Ecological specialization in the Drosophila melanogaster species subgroup: A case study of D. sechellia. Acta Ecol Gener 7:215-229.; Gerlach, 2009Gerlach J (2009) Conservation of the Seychelles sheath-tailed bat Coleura seychellensis on Silhouette Island, Seychelles. Endanger Species Res 8:5-13.); and D. erecta, another species of this subgenus, which has 59 and 56 OR and GR genes, respectively, and is endemic to west-central Africa, and specialized on ripe fruit of Pandanus spp. (Lachaise and Tsacas, 1974Lachaise D and Tsacas L (1974) Les drosophilidae des savanes prèforestiéres de la region tropicale de Lamto (Cote-d’Ivoire). II. Le peuplement des fruits de Pandanus candelabrum (pandanaceae). Ann Univ Abidjan 7:153-192. ; Rio et al., 2010Rio DC, Ares M, Hannon GJ and Nilsen TW (2010) Purification of RNA using TRIzol (TRI reagent). Cold Spring Harb Protoc 2010:pdb-rot5439.; Linz et al., 2013Linz J, Baschwitz A, Strutz A, Dweck HK, Sachse S, Hansson BS and Stensmyr MC (2013) Host plant-driven sensory specialization in Drosophila erecta. Proc Royal Soc B 280:20130626.). This reinforces that, although several trends and potentials seem to have been inherited from ancestral populations, the patterns and processes behind the evolution of OR and GR gene families are certainly quite complex and need to be evaluated under different perspectives.

Conclusion

The evolution of specialization in the use of specific resources is an interesting topic that is not yet fully understood (Etges, 2019Etges WJ (2019) Evolutionary genomics of host plant adaptation: Insights from Drosophila. Curr Opin Insect Sci 36:96-102.; Markow, 2019Markow TA (2019) Host use and host shifts in Drosophila. Curr Opin Insect Sci 31:139-145.; Auer et al., 2020Auer TO, Khallaf MA, Silbering AF, Zappia G, Ellis K, Álvarez-Ocaña R, Arguello JR, Hansson BS, Jefferis GS, Caron SJ et al. (2020) Olfactory receptor and circuit evolution promote host specialization. Nature 579:402-408.). Here, we demonstrated the clear preference of D. incompta for the scent of macerated Cestrum flowers compared to other resources and characterized some genomic changes on the OR and GR genes that are possibly related to this behavior. In characterizing the chemoreceptor repertoire of D. incompta, we not only detected some autapomorphic or homoplastic gene losses, but also found several indications of positive selection on different components of the two major gene families that encompass some of the basic tools for interacting with nature (Benton, 2015Benton R (2015) Multigene family evolution: Perspectives from insect chemoreceptors. Trends Ecol Evol 30:590-600.). However, most gene losses in this species appear to have first occurred in its ancestors located at different nodes of the Drosophila phylogenetic tree. So, generally speaking, our results do not seem to corroborate the idea that specialized species have a more limited number of ORs and GRs. Nevertheless, the precise relationship between these changes and the ecological potential of each species remains largely unclear, and it is not currently possible to decipher the putative causes and consequences of ecological specialization.

Acknowledgements

This study was supported by research grants and fellowships from Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq). PMF received a research grant from CNPQ - GD. ELSL and LJR are research fellows of CNPq - PQ (305613/2020-0 and 314206/2021-3, respectively).

References

  • Afgan E, Baker D, Batut B, van den Beek M, Bouvier D, Cech M, Chilton J, Clements D, Coraor N, Grüning BA, Guerler A et al (2018) The Galaxy platform for accessible, reproducible and collaborative biomedical analyses: 2018 update. Nucleic Acids Res 46:W537-W544.
  • Allen JM, LaFrance R, Folk RA, Johnson KP and Guralnick RP (2018) aTRAM 2.0: An improved, flexible locus assembler for NGS data. Evol Bioinform Online 14:1176934318774546.
  • Alrabayah IN, El Hawary SS, El-Kadder EM, Essa A and Elraey M (2024) Genus Cestrum therapeutic potential: An updated review of its phytochemical, pharmacological, and morphological features. Egyptian J Chem 67:375-393.
  • Anholt RR (2020) Chemosensation and evolution of Drosophila host plant selection. Science 23:100799.
  • Antipov D, Korobeynikov A, McLean JS and Pevzner PA (2016) hybridSPAdes: An algorithm for hybrid assembly of short and long reads. Bioinformatics 32:1009-1015.
  • Asplen MK, Anfora G, Biondi A, Choi DS, Chu D, Daane KM, Gibert P, Gutierrez AP, Hoelmer KA, Hutchison WD et al (2015) Invasion biology of spotted wing Drosophila (Drosophila suzukii): A global perspective and future priorities. J Pest Sci 88:469-494.
  • Auer TO, Khallaf MA, Silbering AF, Zappia G, Ellis K, Álvarez-Ocaña R, Arguello JR, Hansson BS, Jefferis GS, Caron SJ et al (2020) Olfactory receptor and circuit evolution promote host specialization. Nature 579:402-408.
  • Begon M (1975) The relationships of Drosophila obscura Fallen and D. subobscura Collin to naturally-occurring fruits. Oecologia 20:255-277.
  • Benton R (2015) Multigene family evolution: Perspectives from insect chemoreceptors. Trends Ecol Evol 30:590-600.
  • Biondi A, Traugott M and Desneux N (2016) Special issue on Drosophila suzukii: From global invasion to sustainable control. J Pest Sci 89:603-604.
  • Brncic D (1966) Ecological and cytogenetic studies of Drosophila flavopilosa, a neotropical species living in Cestrum flowers. Evolution 1:16-29.
  • Brncic D (1983) The flavopilosa group of species as an example of flower-breeding species. In: Ashburner M, Carson HL and Thompson JN (eds) The genetics and biology of Drosophila 3rd edition. Academic Press, New York, pp 360-377.
  • Cande J, Prud’homme B and Gompel N (2013) Smells like evolution: The role of chemoreceptor evolution in behavioral change. Curr Opin Neurobiol 23:152-158.
  • Carey AF, Wang G, Su CY, Zwiebel LJ and Carlson JR (2010) Odorant reception in the malaria mosquito Anopheles gambiae Nature 464:66-71.
  • Carson HL (1971) Speciation and the founder principle. Stadler Genetics Symposium 3:51-70.
  • Chen ZX, Sturgill D, Qu J, Jiang H, Park S, Boley N, Suzuki AM, Fletcher AR, Plachetzki DC, FitzGerald PC et al (2014) Comparative validation of the D. melanogaster modENCODE transcriptome annotation. Genome Res 24:1209-1223.
  • Clyne PJ, Warr CG, Freeman MR, Lessing D, Kim J and Carlson JR (1999) A novel family of divergent seven-transmembrane proteins: Candidate odorant receptors in Drosophila Neuron 22:327-338.
  • Cordeiro AR and Winge H (1995) Levels of evolutionary divergence of Drosophila willistoni sibling species. In: Levine L (ed), Genetics of Natural Populations of Theodosius Dobzhansky. 1st edition. Columbia University Press, New York, pp. 241-261
  • Croset V, Rytz R, Cummins SF, Budd A, Brawand D, Kaessmann H, Gibson TJ and Benton R (2010) Ancient protostome origin of chemosensory ionotropic glutamate receptors and the evolution of insect taste and olfaction. PLoS Genet 6:e1001064.
  • Dahanukar A, Hallem EA and Carlson JR (2005) Insect chemoreception. Curr Opin Neurobiol 15:423-430.
  • Date P, Dweck HK, Stensmyr MC, Shann J, Hansson BS and Rollmann SM (2013) Divergence in olfactory host plant preference in D. mojavensis in response to cactus host use. PLoS One 8:e70027.
  • David JR and Capy P (1988) Genetic variation of Drosophila melanogaster natural populations. Trends Genet 4:106-111.
  • Davis RL (2007) The scent of Drosophila sex. Neuron 54:14-16.
  • De Ré FC, Robe LJ, Wallau GL and Loreto EL (2017) Inferring the phylogenetic position of the Drosophila flavopilosa group: Incongruence within and between mitochondrial and nuclear multilocus datasets. J Zool Syst Evol Res 55:208-221.
  • Depetris-Chauvin A, Galagovsky D and Grosjean Y (2015) Chemicals and chemoreceptors: Ecologically relevant signals driving behavior in Drosophila Front Ecol Evol 3:41.
  • Diaz F, Allan CW and Matzkin LM (2018) Positive selection at sites of chemosensory genes is associated with the recent divergence and local ecological adaptation in cactophilic Drosophila BMC Evol Biol 18:144.
  • Doleschall CL (1858) Derde leijdrage tot de kennis der dipteran fauna van Nederlandsh Indie. Natrk. Tijdschr. Nederland Indie 17:73-128.
  • Drosophila 12 Genomes Consortium; Clark AG, Eisen MB, Smith DR, Bergman CM, Oliver B, Markow TA, Kaufman TC, Kellis M, Gelbart W, Iyer VN et al (2007) Evolution of genes and genomes on the Drosophila phylogeny. Nature 450:203-218.
  • Dweck HK, Ebrahim SA, Kromann S, Bown D, Hillbur Y, Sachse S, Hansson BS and Stensmyr MC (2013) Olfactory preference for egg laying on citrus substrates in Drosophila Curr Biol 23:2472-2480.
  • Ebrahim SA, Dweck HK, Stökl J, Hofferberth JE, Trona F, Weniger K, Rybak J, Seki Y, Stensmyr MC, Sachse S et al (2015) Drosophila avoids parasitoids by sensing their semiochemicals via a dedicated olfactory circuit. PLoS Biol 13:e1002318.
  • Etges WJ (2019) Evolutionary genomics of host plant adaptation: Insights from Drosophila Curr Opin Insect Sci 36:96-102.
  • Fishilevich E, Domingos AI, Asahina K, Naef F, Vosshall LB and Louis M (2005) Chemotaxis behavior mediated by single larval olfactory neurons in Drosophila Curr Biol 15:2086-2096.
  • Frota-Pessoa O (1952) Flower-feeding Drosophilidae Drosoph Inf Ser 26:101-102.
  • Fu L, Niu B, Zhu Z, Wu S and Li W (2012) CD-HIT: Accelerated for clustering the next-generation sequencing data. Bioinformatics 28:3150-3152.
  • Fuyama Y (1976) Behavior genetics of olfactory responses in Drosophila I. Olfactometry and strain differences in Drosophila melanogaster Behav Genet 6:407-420.
  • Gao Q and Chess A (1999) Identification of candidate Drosophila olfactory receptors from genomic DNA sequence. Genomics 60:31-39.
  • Gardiner A, Barker D, Butlin RK, Jordan WC and Ritchie MG (2008) Drosophila chemoreceptor gene evolution: Selection, specialization and genome size. Mol Ecol 17:1648-1657.
  • Gerlach J (2009) Conservation of the Seychelles sheath-tailed bat Coleura seychellensis on Silhouette Island, Seychelles. Endanger Species Res 8:5-13.
  • Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, Adiconis X, Fan L, Raychowdhury R, Zeng Q et al (2011) Trinity: Reconstructing a full-length transcriptome without a genome from RNA-Seq data. Nat Biotechnol 29:644.
  • Guo S and Kim J (2007) Molecular evolution of Drosophila odorant receptor genes. Mol Biol Evol 24:1198-1207.
  • Gurevich A, Saveliev V, Vyahhi N and Tesler G (2013) QUAST: Quality assessment tool for genome assemblies. Bioinformatics 29:1072-1075.
  • Hallem EA, Dahanukar A and Carlson JR (2006) Insect odor and taste receptors. Annu Rev Entomol 51:113-135.
  • Hickner PV, Rivaldi CL, Johnson CM, Siddappaji M, Raster GJ and Syed Z (2016) The making of a pest: Insights from the evolution of chemosensory receptor families in a pestiferous and invasive fly, Drosophila suzukii BMC Genom 17:648.
  • Hjelmen CE and Johnston JS (2017) The mode and tempo of genome size evolution in the subgenus Sophophora PLoS One 12:e0173505.
  • Hofmann PR and Napp M (1984) Genetic-environmental relationships in Drosophila incompta, a species of restricted ecology. Rev Bras Genet 7:21-39.
  • Huerta-Cepas J, Forslund K, Coelho LP, Szklarczyk D, Jensen LJ, Von Mering C and Bork P (2017) Fast genome-wide functional annotation through orthology assignment by eggNOG-mapper. Mol Biol Evol 34:2115-2122.
  • Joseph RM and Carlson JR (2015) Drosophila chemoreceptors: A molecular interface between the chemical world and the brain. Trends Genet 31:683-695.
  • Karpe SD, Tiwari V and Ramanathan S (2021) InsectOR-Webserver for sensitive identification of insect olfactory receptor genes from non-model genomes. PLoS One 16:e0245324.
  • Kelemen L and Moritz C (1999) Comparative phylogeography of a sibling pair of rainforest Drosophila species (Drosophila serrata and D. birchii). Evolution 53:1306-1311.
  • Kreher SA, Kwon JY and Carlson JR (2005) The molecular basis of odor coding in the Drosophila larvae. Neuron 46:445-456.
  • Kumar S, Stecher G, Li M, Knyaz C and Tamura K (2018) MEGA X: Molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol 35:1547-1549.
  • Kurtovic A, Widmer A and Dickson BJ (2007) A single class of olfactory neurons mediates behavioural responses to a Drosophila sex pheromone. Nature 446:542-546.
  • Lachaise D and Tsacas L (1974) Les drosophilidae des savanes prèforestiéres de la region tropicale de Lamto (Cote-d’Ivoire). II. Le peuplement des fruits de Pandanus candelabrum (pandanaceae). Ann Univ Abidjan 7:153-192.
  • Lachaise D and Silvain JF (2004) How two Afrotropical endemics made two cosmopolitan human commensals: The Drosophila melanogaster-D. simulans palaeogeographic riddle. Genetica 120:17-39.
  • Lemeunier F, Tsacas L, David J and Ashburner M (1986) The melanogaster species group. In: Ashburner M, Thompson JR Jr and Carson HL (eds) The genetics and biology of Drosophila, vol. 3, Academic Press, New York , pp 147-256.
  • Linz J, Baschwitz A, Strutz A, Dweck HK, Sachse S, Hansson BS and Stensmyr MC (2013) Host plant-driven sensory specialization in Drosophila erecta Proc Royal Soc B 280:20130626.
  • Louis J and David JR (1986) Ecological specialization in the Drosophila melanogaster species subgroup: A case study of D. sechellia Acta Ecol Gener 7:215-229.
  • Ludwig A, Vidal NM, Loreto EL and Sepel LM (2002) Drosophila incompta development without flowers. Drosoph Inf Ser 85:40-41.
  • Malogolowkin C (1953) Sobre a genitália dos drosofilídeos. IV. A genitália masculina no subgênero Drosophila (Diptera, Drosophilidae). Rev Bras Biol 13:245-264.
  • Mansourian S, Enjin A, Jirle EV, Ramesh V, Rehermann G, Becher PG, Pool JE and Stensmyr MC (2018) Wild African Drosophila melanogaster are seasonal specialists on marula fruit. Curr Biol 28:3960-3968.
  • Markow TA and O’Grady P (2008) Reproductive ecology of Drosophila Funct Ecol 747-759.
  • Markow TA (2015) The natural history of model organisms: The secret lives of Drosophila flies. Elife 4:e06793.
  • Markow TA (2019) Host use and host shifts in Drosophila Curr Opin Insect Sci 31:139-145.
  • Martin F, Boto T, Gomez‐Diaz C and Alcorta E (2013) Elements of olfactory reception in adult Drosophila melanogaster Anat Rec 296:1477-1488.
  • McBride CS (2007) Rapid evolution of smell and taste receptor genes during host specialization in Drosophila sechellia Proc Nat Acad Sci U S A 104:4996-5001.
  • McBride CS and Arguello JR (2007) Five Drosophila genomes reveal nonneutral evolution and the signature of host specialization in the chemoreceptor superfamily. Genetics 177:1395-1416.
  • Nguyen LT, Schmidt HA, Von Haeseler A and Minh BQ (2015) IQ-TREE: A fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol Biol Evol 32:268-274.
  • Nozawa M and Nei M (2007) Evolutionary dynamics of olfactory receptor genes in Drosophila species. Proc Natl Acad Sci U S A 104:7122-7127.
  • O’Grady PM and DeSalle R (2018) Phylogeny of the genus Drosophila Genetics 209:25.
  • Pipkin SB, Rodriguez RL and Leon J (1966) Plant host specificity among flower-feeding neotropical Drosophila (Diptera: Drosophilidae). Am Nat 100:135-156.
  • Pond SLK and Frost SD (2005) Not so different after all: A comparison of methods for detecting amino acid sites under selection. Mol Biol Evol 22:1208-1222.
  • Powell JR (1997) Progress and prospects in evolutionary biology: The Drosophila model. Oxford University Press, Oxford, 576 p.
  • Pozo MJ, de Vega C, Canto A and Herrera CM (2009) Presence of yeasts in floral nectar is consistent with the hypothesis of microbial-mediated signaling in plant-pollinator interactions. Plant Signal Behav 4:1102-1104.
  • Ramasamy S, Ometto L, Crava CM, Revadi S, Kaur R, Horner DS, Pisani D, Dekker T, Anfora G and Rota-Stabelli O (2016) The evolution of olfactory gene families in Drosophila and the genomic basis of chemical-ecological adaptation in Drosophila suzukii Genome Biol Evol 8:2297-2311.
  • Ramniwas S and Kajla B (2012) Divergent strategy for adaptation to drought stress in two sibling species of montium species subgroup: Drosophila kikkawai and Drosophila leontia J Insect Physiol 58:1525-1533.
  • Ranwez V, Harispe S, Delsuc F and Douzery EJ (2011) MACSE: Multiple Alignment of Coding SEquences accounting for frameshifts and stop codons. PLoS One 6:e22594.
  • Rio DC, Ares M, Hannon GJ and Nilsen TW (2010) Purification of RNA using TRIzol (TRI reagent). Cold Spring Harb Protoc 2010:pdb-rot5439.
  • Robe LJ, De Ré FC, Ludwig A and Loreto EL (2013) The Drosophila flavopilosa species group (Diptera, Drosophilidae) An array of exciting questions. Fly 7:59-69.
  • Robe LJ, Loreto EL and Valente VL (2010) Radiation of the “Drosophila” subgenus (Drosophilidae, Diptera) in the Neotropics. J Zool Syst Evol Res 48:310-321.
  • Robertson HM, Warr CG and Carlson JR (2003) Molecular evolution of the insect chemoreceptor gene superfamily in Drosophila melanogaster Proc Nat Acad Sci U S A 100:14537-14542.
  • Santos RD and Vilela CR (2005) Breeding sites of Neotropical Drosophilidae (Diptera): IV. Living and fallen flowers of Sessea brasiliensis and Cestrum spp(Solanaceae). Rev Bras Entom 49:544-551.
  • Schiffer M and McEvey SF (2006) Drosophila bunnanda: A new species from northern Australia with notes on other Australian members of the montium subgroup (Diptera: Drosophilidae). Zootaxa 1333:1-23.
  • Schmitz HJ and Valente VL (2019) The flower flies and the unknown diversity of Drosophilidae (Diptera): A biodiversity inventory in the Brazilian fauna. Pap Avulsos Zool 59:e20195945.
  • Scott K, Brady Jr R, Cravchik A, Morozov P, Rzhetsky A, Zuker C and Axel R (2001) A chemosensory gene family encoding candidate gustatory and olfactory receptors in Drosophila Cell 104:661-673.
  • Sepel LM, Golombieski RM, Napp M and Loreto EL (2000) Seasonal fluctuations of D. cestri and D. Incompta, two species of the flavopilosa group. Drosoph Inf Serv 83:122-126.
  • Sevenster JG and Van Alphen JJ (1993) A life history trade-off in Drosophila species and community structure in variable environments. J Anim Ecol 62:720-736.
  • Simão FA, Waterhouse RM, Ioannidis P, Kriventseva EV and Zdobnov EM (2015) BUSCO: Assessing genome assembly and annotation completeness with single-copy orthologs. Bioinformatics 31:3210-3212.
  • Singh BN (1996) Population and behaviour genetics of Drosophila ananassae Genetica 97:321-329.
  • Smith MD, Wertheim JO, Weaver S, Murrell B, Scheffler K and Kosakovsky Pond SL (2015) Less is more: An adaptive branch-site random effects model for efficient detection of episodic diversifying selection. Mol Biol Evol 32:1342-1353.
  • Spassky B, Richmond RC, Perez-Salas S, Pavlovsky O, Mourao CA, Hunter AS, Hoenigsberg H, Dobzhansky T and Ayala FJ (1971) Geography of the sibling species related to Drosophila willistoni, and of the semispecies of the Drosophila paulistorum complex. Evolution 25:129-143.
  • Stanke M, Keller O, Gunduz I, Hayes A, Waack S and Morgenstern B (2006) AUGUSTUS: Ab initio prediction of alternative transcripts. Nucleic Acids Res 34:W435-W439.
  • Sturtevant AH (1919) A new species closely resembling Drosophila melanogaster Psyche 26:153-155.
  • Thompson JD, Higgins DG and Gibson TJ (1994) CLUSTAL W: Improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res 22:4673-4680.
  • Throckmorton LH (1975) The phylogeny, ecology and geography of Drosophila In: King RC (ed). Handbook of genetics, vol 3, Plenum Press, New York, pp 421-469.
  • Thurmond J, Goodman JL, Strelets VB, Attrill H, Gramates LS, Marygold SJ, Matthews BB, Millburn G, Antonazzo G, Trovisco V et al (2019) FlyBase 2.0: The next generation. Nucleic Acids Res 47:59-65.
  • Tsacas L and David J (1974) Drosophila mauritiana n. sp. du groupe melanogaster de l’Ile Maurice [Dipt. Drosophilidae]. Bull Soc Entomol 79:42-46.
  • Tsacas L and Lachaise D (1974) Quatre nouvelles especes de la Cote-d’Ivoire du genre Drosophila, groupe melanogaster et discussion de l’origine du sous-groupe melanogaster (Diptera: Drosophilidae). Ann Univ Abidjan 7:193-211.
  • Tsacas L and Bächli G (1981) Drosophila sechellia, n. sp., huitieme espece du sous-groupe melanogaster des iles Sechelles (Diptera, Drosophilidae). Rev Fr Entomol 3:146-150.
  • Vieira FG and Rozas J (2011) Comparative genomics of the odorant-binding and chemosensory protein gene families across the Arthropoda: Origin and evolutionary history of the chemosensory system. Genome Biol Evol 3:476-490.
  • Vosshall LB and Stocker RF (2007) Molecular architecture of smell and taste in Drosophila Annu Rev Neurosci 30:505-533.
  • Weaver S, Shank SD, Spielman SJ, Li M, Muse SV and Kosakovsky Pond SL (2018) Datamonkey 2.0: A modern web application for characterizing selective and other evolutionary processes. Mol Biol Evol 35:773-777.
  • Wheeler MR, Takada H and Brncic D (1962) The flavopilosa species group of Drosophila Univ Texas Publ Genetics 6205:395-414.
  • Whiteman NK and Pierce NE (2008) Delicious poison: Genetics of Drosophila host plant preference. Trends Ecol Evol 23:473-478.
  • Wisotsky Z, Medina A, Freeman E and Dahanukar A (2011) Evolutionary differences in food preference rely on Gr64e, a receptor for glycerol. Nat Neurosci 14:1534-1541.
  • Wogaman DJ and Seiger MB (1983) Light intensity as a factor in the choice of an oviposition site by Drosophila pseudoobscura and Drosophila persimilis Can J Genet Cytol 25:370-377.
  • Wyatt TD (2003) Pheromones and animal behaviour: Communication by smell and taste. Cambridge University Press, Cambridge.
  • Yang Y, Hou ZC, Qian YH, Kang H and Zeng QT (2012) Increasing the data size to accurately reconstruct the phylogenetic relationships between nine subgroups of the Drosophila melanogaster species group (Drosophilidae, Diptera). Mol Phylogenet Evol 62:214-223.
  • Yassin A (2013) Phylogenetic classification of the Drosophilidae Rondani (Diptera): The role of morphology in the postgenomic era. Syst Entomol 38:349-364.

Internet Resources

Edited by

Associate Editor:

Fabrício Rodrigues dos Santos

Publication Dates

  • Publication in this collection
    14 June 2024
  • Date of issue
    2024

History

  • Received
    28 Dec 2022
  • Accepted
    07 May 2024
Sociedade Brasileira de Genética Rua Cap. Adelmio Norberto da Silva, 736, 14025-670 Ribeirão Preto SP Brazil, Tel.: (55 16) 3911-4130 / Fax.: (55 16) 3621-3552 - Ribeirão Preto - SP - Brazil
E-mail: editor@gmb.org.br