Acessibilidade / Reportar erro

Horizontal transfer and the widespread presence of Galileo transposons in Drosophilidae (Insecta: Diptera)

Abstract

Galileo is a transposon notoriously involved with inversions in Drosophila buzzatii by ectopic recombination. Although widespread in Drosophila, little is known about this transposon in other lineages of Drosophilidae. Here, the abundance of the canonical Galileo and its evolutionary history in Drosophilidae genomes was estimated and reconstructed across genera within its two subfamilies. Sequences of this transposon were masked in these genomes and their transposase sequences were recovered using BLASTn. Phylogenetic analyses were employed to reconstruct their evolutionary history and compare it to that of host genomes. Galileo was found in nearly all 163 species, however, only 37 harbored nearly complete transposase sequences. In the remaining, Galileo was found highly fragmented. Copies from related species were clustered, however horizontal transfer events were detected between the melanogaster and montium groups of Drosophila, and between the latter and the Lordiphosa genus. The similarity of sequences found in the virilis and willistoni groups of Drosophila was found to be a consequence of lineage sorting. Therefore, the evolution of Galileo is primarily marked by vertical transmission and long-term inactivation, mainly through the deletion of open reading frames. The latter has the potential to lead copies of this transposon to become miniature inverted-repeat transposable elements.

Keywords:
DNA transposon; MITEs; P superfamily

Introduction

Transposable elements (TEs) belong to the repetitive fraction of genomes, and are linear sequences of DNA that have the ability to move within or between genomes (Wells and Feschotte, 2020Wells JN and Feschotte C (2020) A field guide to eukaryotic transposable elements. Annu Rev Genet 54:539-561.). Classifications divide these sequences firstly into two classes, based on the intermediate molecule in their transposition process (Finnegan, 1989Finnegan DJ (1989) Eukaryotic transposable elements and genome evolution. Trends Genet 5:103-107. ). Class I is composed of retrotransposons as their mobilization involves the synthesis of an RNA molecule, which are retrotranscribed into DNA and then inserted elsewhere in the genome (Wicker et al., 2007Wicker T, Sabot F, Hua-Van A, Bennetzen JL, Capy P, Chalhoub B, Flavell A, Leroy P, Morgante M, Panaud O, Paux E, SanMiguel P and Schulman AH (2007) A unified classification system for eukaryotic transposable elements. Nat Rev Genet 8:973-982.). On the other hand, the majority of Class II elements - or DNA transposons - are directly excised by their transposase (TPase), and then reinserted in another site in the genome (Wicker et al., 2007Wicker T, Sabot F, Hua-Van A, Bennetzen JL, Capy P, Chalhoub B, Flavell A, Leroy P, Morgante M, Panaud O, Paux E, SanMiguel P and Schulman AH (2007) A unified classification system for eukaryotic transposable elements. Nat Rev Genet 8:973-982.).

In addition, TEs can be either autonomous or nonautonomous (Wicker et al., 2007Wicker T, Sabot F, Hua-Van A, Bennetzen JL, Capy P, Chalhoub B, Flavell A, Leroy P, Morgante M, Panaud O, Paux E, SanMiguel P and Schulman AH (2007) A unified classification system for eukaryotic transposable elements. Nat Rev Genet 8:973-982.). The first are those that present their structures preserved, encoding all necessary enzymes to be transposed. The latter comprise defective TEs that no longer encode nor produce their own proteins, and move only if recognized by the enzymes of a closely related autonomous TE; such as the Miniature Inverted-repeat Transposable Elements (MITEs). MITEs are non-autonomous TEs, derived from autonomous Class II transposons, and present a few structural characteristics: (i) small size, ranging from 50 to 500 base pairs (bp); (ii) AT-rich sequences; and (iii) a lack of a functional TPase (Deprá et al., 2012Deprá M, Ludwig A, Valente VL and Loreto EL (2012) Mar, a MITE family of hAT transposons in Drosophila. Mob DNA 3:13. ; Fattash et al., 2013Fattash I, Rooke R, Wong A, Hui C, Luu T, Bhardwaj P and Yang G (2013) Miniature inverted-repeat transposable elements: Discovery, distribution, and activity. Genome 56:475-486. ).

Transposable elements are often referred to as “parasites” (Colonna Romano and Fanti, 2022Colonna Romano N and Fanti L (2022) Transposable elements: Major players in shaping genomic and evolutionary patterns. Cells 11:1048. ), given their ability to invade new genomes and increase their copy number (Loreto et al., 2008Loreto ELS, Carareto CMA and Capy P (2008) Revisiting horizontal transfer of transposable elements in Drosophila. Heredity 100:545-554. ). Horizontal transposon transfer (HTT) is the phenomenon in which a given TE “jumps” to the genome of a non-closely related species, i.e., sexually isolated organisms (Panaud, 2016Panaud O (2016) Horizontal transfers of transposable elements in eukaryotes: The flying genes. C R Biol 7-8: 296-299.). The role of HTT in shaping diversity as an endogenous source of evolution is widely recognized (Pace et al., 2008Pace JK, Gilbert C, Clark MS and Feschotte C (2008) Repeated horizontal transfer of a DNA transposon in mammals and other tetrapods. Proc Natl Acad Sci U S A 105:17023-17028.; Gilbert and Feschotte, 2018Gilbert C and Feschotte C (2018) Horizontal acquisition of transposable elements and viral sequences: Patterns and consequences. Curr Opin Genet Dev 49:15-24. ; Carvalho et al., 2023Carvalho TL, Cordeiro J, Vizentin-Bugoni J, Fonseca PM, Loreto ELS and Robe LJ (2023) Horizontal transposon transfer and their ecological drivers: The case of flower-breeding Drosophila. Genome Biol Evol 15:evad068.), and its frequency is much higher than previously thought (Schaack et al., 2010Schaack S, Gilbert C and Feschotte C (2010) Promiscuous DNA: Horizontal transfer of transposable elements and why it matters for eukaryotic evolution. Trends Ecol Evol 25:537-546. ; Panaud, 2016Panaud O (2016) Horizontal transfers of transposable elements in eukaryotes: The flying genes. C R Biol 7-8: 296-299.; Peccoud et al., 2017Peccoud J, Loiseau V, Cordaux R and Gilbert C (2017) Massive horizontal transfer of transposable elements in insects. Proc Natl Acad Sci U S A 114:4721-4726.; Melo and Wallau, 2020Melo ES and Wallau GL (2020) Mosquito genomes are frequently invaded by transposable elements through horizontal transfer. PLOS Genetics 16:e1008946.).

In this sense, several evolutionary events have been proposed as a direct consequence of TEs mobilization and/or recombination. For instance, in several taxa the variation and evolution of genome size are directly related to the amplification or contraction in TEs copy number (Canapa et al., 2015Canapa A, Barucca M, Biscotti MA, Forconi M and Olmo E (2015) Transposons, genome size, and evolutionary insights in animals. Cytogenet Genome Res 147:217-239. ; Antoniolli et al., 2023Antoniolli HRM, Deprá M and Valente VLS (2023) Patterns of genome size evolution versus fraction of repetitive elements in statu nascendi species: the case of the willistoni subgroup of Drosophila (Diptera, Drosophilidae). Genome 66:193-201. ). Nucleotide polymorphisms are also frequently produced after transposition events (Bourque et al., 2018Bourque G, Burns KH, Gehring M, Gorbunova V, Seluanov A, Hammell M, Imbeault M, Izsvák Z, Levin HL, Macfarlan TS, Mager DL and Feschotte C (2018) Ten things you should know about transposable elements. Genome Biol 19:199. ). Transposable elements are also known to be related to changes in gene expression, either by silencing or enhancing them (Finnegan, 1989Finnegan DJ (1989) Eukaryotic transposable elements and genome evolution. Trends Genet 5:103-107. ), and chromosomal rearrangements - i.e., deletions, duplications, translocations and inversions by ectopic recombination (Kidwell and Lisch, 1997Kidwell MG and Lisch D (1997) Transposable elements as sources of variation in animals and plants. Proc Natl Acad Sci U S A94:7704-7711. ). In the latter, distant loci in a genome carry highly similar TE copies, which allows homologous recombination to occur (see review in Bourque et al., 2018Bourque G, Burns KH, Gehring M, Gorbunova V, Seluanov A, Hammell M, Imbeault M, Izsvák Z, Levin HL, Macfarlan TS, Mager DL and Feschotte C (2018) Ten things you should know about transposable elements. Genome Biol 19:199. ), thus resulting in a drastic modification in the chromosome architecture (Ren et al., 2018Ren L, Huang W, Cannon EKS, Bertioli DJ and Cannon SB (2018) A mechanism for genome size reduction following genomic rearrangements. Front Genet 9:454. ). Documented cases of a TE as a mediator of ectopic recombination include the families of retrotransposons Bel-Pao, Doc, I element and roo, as well as the transposons foldback, Galileo and hobo (Lim and Simmons, 1994Lim JK and Simmons MJ (1994) Gross chromosome rearrangements mediated by transposable elements in Drosophila melanogaster. Bioessays 16:269-275.; Delprat et al., 2009Delprat A, Negre B, Puig M and Ruiz A (2009) The transposon Galileo generates natural chromosomal inversions in Drosophila by ectopic recombination. PLoS One 4:e7883. ).

Galileo is a family of Class II transposons, and encodes its own TPase flanked by terminal inverted repeats (TIRs). Initially described as a foldback-like element, its TIRs and THAP domains exhibit similarities with those of the P element, leading to the classification of Galileo within the P superfamily (Marzo et al., 2008Marzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962. ). However, unlike the P element, Galileo does not present introns (Marzo et alMarzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962. ., 2008). Galileo was discovered by Cáceres et al. (1999Cáceres M, Ranz JM, Barbadilla A, Long M and Ruiz A (1999) Generation of a widespread Drosophila inversion by a transposable element. Science 285:415-418. ) due to its association with the breakpoints of the 2j inversion in wild specimens of Drosophila buzzatii. In fact, Galileo is the only TE known to induce chromosomal rearrangements in natural populations of Drosophila (Marzo et al., 2008Marzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962. ), as most others have been observed in laboratory populations (Lim and Simmons, 1994Lim JK and Simmons MJ (1994) Gross chromosome rearrangements mediated by transposable elements in Drosophila melanogaster. Bioessays 16:269-275.). Besides the 2j inversion, Galileo was involved with two other rearrangements described in D. buzzatii (Casals et al., 2003Casals F, Cáceres M and Ruiz A (2003) The foldback-like transposon Galileo is involved in the generation of two different natural chromosomal inversions of Drosophila buzzatii. Mol Biol Evol 20:674-685. ; Delprat et al., 2009Delprat A, Negre B, Puig M and Ruiz A (2009) The transposon Galileo generates natural chromosomal inversions in Drosophila by ectopic recombination. PLoS One 4:e7883. ). This makes this transposon as one of the most well-documented examples of a natural TE-induced chromosomal rearrangement.

Studies have shown the widespread presence of Galileo in the Drosophila genus (Marzo et al., 2008Marzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962. ; Acurio, 2015Acurio AE (2015) Coevolutionary analysis of the transposon Galileo in the genus Drosophila. M. Sc. Thesis, Autonomous University of Barcelona, 219 p.). The main focus of the present study was to characterize the evolutionary history of the Galileo family and evaluate its main transmission mode in Drosophilidae. This transposon was masked in genome assemblies of 163 species available at online databases, and TPase sequences found were employed for reconstructing a phylogeny and testing putative cases of HTT.

Material and Methods

Masking Galileo in the genome assemblies

Representative genome assemblies of 163 Drosophilidae species (see details on taxonomy and accession numbers in Table S1 Table S1 - List and taxonomy of Drosophilidae species included in this study, including positive results from BLASTn searches of the Galileo transposase sequences and accession numbers to the genome assembly and short-read sequencing data on NCBI. ) were retrieved from GenBank (NCBI) with a Python package written by Blin (2021Blin K (2021) NCBI Genome Downloading Scripts, Blin K (2021) NCBI Genome Downloading Scripts, https://github.com/kblin/ncbi-genome-download/ (accessed 10 November 2022)
https://github.com/kblin/ncbi-genome-dow...
). These species belong to the Chymomyza, Drosophila, Lordiphosa, Scaptodrosophila, Scaptomyza, and Zaprionus genera of the Drosophilinae subfamily; and Leucophenga and Phortica of Steganinae subfamily (Table S1 Table S1 - List and taxonomy of Drosophilidae species included in this study, including positive results from BLASTn searches of the Galileo transposase sequences and accession numbers to the genome assembly and short-read sequencing data on NCBI. ). BUSCO v.5 (Manni et al., 2021Manni M, Berkeley MR, Seppey M, Simão FA and Zdobnov EM (2021) BUSCO Update: Novel and streamlined workflows along with broader and deeper phylogenetic coverage for scoring of eukaryotic, prokaryotic, and viral genomes. Mol Biol Evol 38(10):4647-4654. ) was employed to assess the completeness of each assembly with the Diptera orthologous database.

The nucleotide sequence of seven Galileo copies characterized by Marzo et al. (2008Marzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962. ) in D. ananassae (Dana\Galileo - BK006363), D. buzzatii (Dbuz\Galileo - EU334682 and EU334685), D. mojavensis (Dmoj\Galileo - BK006357), D. persimilis (Dper\Galileo - BK006361), D. virilis (Dvir\Galileo - BK006359) and D. willistoni (Dwil\Galileo - BK006360) were downloaded from GenBank, and used as queries in our workflow. Firstly, the queries were input as the repeat library in RepeatMasker (Smit et al., 2023Smit A, Hubley R and Green P (2023) RepeatMasker 4.0 , 0 , http://www.repeatmasker.org/ (accessed 10 November 2022)
http://www.repeatmasker.org/...
) for masking canonical Galileo sequences in each genome assembly. The script ‘One code to find them all’ (Bailly-Bechet et al., 2014Bailly-Bechet M, Haudry A and Lerat E (2014) “One code to find them all”: A perl tool to conveniently parse RepeatMasker output files. Mob DNA 5:13. ) was then employed to parse the output, recovering the nucleotide sequence of each identified copy in an assembly with at least 80% identity to its best query and a minimum length of 80 base pairs.

Phylogenetic analysis of Galileo potentially autonomous copies

The complete nucleotide sequence encoding the transposase (TPase) of six copies (Dana\Galileo, Dbuz\Galileo, Dmoj\Galileo, Dper\Galileo, Dvir\Galileo, and Dwil\Galileo) served as queries for local BLASTn searches in each FASTA file containing the Galileo copies of each genome. Hits with at least 80% identity and coverage of at least 70% for any of the queries were used in downstream analyses. Additionally, a P element from the genome of Drosophila buzzatii (GenBank accession No. KC690135) and two copies of the 1360 element (GenBank accession Nos. AF533772 and AY138841) were included in the nucleotide matrix as outgroups. This matrix was aligned with MACSE v2 (Ranwez et al., 2018Ranwez V, Douzery EJP, Cambon C, Chantret N and Delsuc F (2018) MACSE v2: Toolkit for the alignment of coding sequences accounting for frameshifts and stop codons. Mol Biol Evol 35:2582-2584. ) in two steps: (i) using the option alignSequences, which aligns nucleotide sequences based on their underlying codon structure, accounting for frameshifts and stop codons; (ii) the resulting alignment was edited with the option exportAlignment, replacing codons containing frameshifts and internal stop codons with “N” (e.g., TG! was replaced by NNN). The codon alignment was then processed with Gblocks (Castresana, 2000Castresana J (2000) Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol 17:540-552. ) to remove poorly aligned regions, allowing the presence of gaps.

The final codon alignment was translated to amino acids and used for a Bayesian phylogenetic inference (BI) analysis, performed in MrBayes 3.2.7 (Ronquist et al., 2012Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Höhna S, Larget B, Liu L, Suchard MA and Huelsenbeck JP (2012) MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol 61:539-542. ). The majority-rule consensus tree was built under the best amino acid substitution model, as estimated by ModelTest-NG (Darriba et al., 2020Darriba D, Posada D, Kozlov AM, Stamatakis A, Morel B and Flouri T (2020) ModelTest-NG: A new and scalable tool for the selection of DNA and protein evolutionary models. Mol Biol Evol 37:291-294. ). Metropolis-coupled Markov chain Monte Carlo (MCMCMC) analysis was run with two parallel runs with four chains each for 1,000,000 generations, sampling every 100. Convergence was reached when the average standard deviation of split frequencies was below 1%. A burn-in of 25% was applied to the sampled trees before obtaining the consensus tree. The tree was visualized and edited in FigTree (Rambaut, 2018Rambaut A (2018) FigTree v1.4.4 , 4 , https://github.com/rambaut/figtree/ (accessed 9 April 2023)
https://github.com/rambaut/figtree/...
).

Analysis of abundance and repeat profile

Forward short-reads of high-throughput whole genome sequencing were downloaded from the Sequence Read Archive of NCBI (see SRA accession No. in Table S1 Table S1 - List and taxonomy of Drosophilidae species included in this study, including positive results from BLASTn searches of the Galileo transposase sequences and accession numbers to the genome assembly and short-read sequencing data on NCBI. ) for those species with positive hits for the TPase queries. These were submitted to the RepeatProfiler pipeline (Negm et al., 2021Negm S, Greenberg A, Larracuente AM and Sproul JS (2021) RepeatProfiler: A pipeline for visualization and comparative analysis of repetitive DNA profiles. Mol Ecol Resour 21:969-981. ), an analysis in which sequencing reads are mapped against queries to build coverage graphs, allowing to infer which regions of a given query have a higher or lower abundance.

Quality trim was performed with fastp (Chen et al., 2018Chen S, Zhou Y, Chen Y and Gu J(2018) fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 34:i884-i890. ), when reads had their adaptor removed while keeping only reads with no N base. The total reads were downsampled to 3 million, achieving near 1x coverage for all genomes (assuming a genome size mean of 200 megabases for species of the family Drosophilidae). In addition, five single-copy genes were randomly selected in the Diptera orthologous genes dataset of BUSCO 5 (Manni et al., 2021Manni M, Berkeley MR, Seppey M, Simão FA and Zdobnov EM (2021) BUSCO Update: Novel and streamlined workflows along with broader and deeper phylogenetic coverage for scoring of eukaryotic, prokaryotic, and viral genomes. Mol Biol Evol 38(10):4647-4654. ) to normalize the results (Table S2 Table S2 - List of genes used to normalize the results of profile and abundance of Galileo across the analyzed genomes in this study. ). The six complete copies of Galileo used in BLASTn searches were used as queries (Dbuz\Galileo - EU334685 was excluded because it was shorter than Dbuz\Galileo - EU334682). RepeatProfiler (Negm et al., 2021Negm S, Greenberg A, Larracuente AM and Sproul JS (2021) RepeatProfiler: A pipeline for visualization and comparative analysis of repetitive DNA profiles. Mol Ecol Resour 21:969-981. ) was executed with default parameters.

Inference of HTT events

Possible cases of HTT were determined based on incongruences between the phylogeny of host genomes and the phylogeny of Galileo. Validation of such cases was performed with the vhica R package (Wallau et al., 2015Wallau GL, Capy P, Loreto E, Le-Rouzic A and Hua-Van A (2015) VHICA, a new method to discriminate between vertical and horizontal transposon transfer: Application to the Mariner Family within Drosophila. Mol Biol Evol 33:1094-1109.), implemented on the HTT-DB platform (Dotto et al., 2015Dotto BR, Carvalho EL, Silva AF, Duarte Silva LF, Pinto PM, Ortiz MF and Wallau GL (2015) HTT-DB: Horizontally transferred transposable elements database. Bioinformatics 31:2915-2917. ). This method relies on discrepancies in the evolutionary rates of synonymous positions (dS), which considers codon usage bias (CUB), between nuclear genes (vertically transmitted) and transposable elements (TEs). Wallau et al. (2015) demonstrated that dS and CUB are correlated, and low values for both are indicative of inconsistencies with vertical transmission.

Sequences of single-copy orthologous genes were searched in the assemblies with positive hits of Galileo using BUSCO 5 (Manni et al., 2021Manni M, Berkeley MR, Seppey M, Simão FA and Zdobnov EM (2021) BUSCO Update: Novel and streamlined workflows along with broader and deeper phylogenetic coverage for scoring of eukaryotic, prokaryotic, and viral genomes. Mol Biol Evol 38(10):4647-4654. ) and the Diptera orthologous database. Nucleotide sequences of 30 randomly selected genes (see Table S3 Table S3 - List of genes used for Codon Usage Bias (CUB) comparisons in the analysis of horizontal transposon transfer (HTT) in vhica R package. ) were aligned based on codons using the ClustalW 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.) implemented in MEGA 11 (Tamura et al., 2021Tamura K, Stecher G and Kumar S (2021) MEGA11: Molecular Evolutionary Genetics Analysis Version 11. Mol Biol Evol 38:3022-3027. ). These alignments were used to compare the dS-CUB between the host nuclear genome and Galileo sequences. A substitution rate of 0.016 per million years (Sharp and Li, 1989Sharp PM and Li WH (1989) On the rate of DNA sequence evolution in Drosophila. J Mol Evol 28: 398-402. ) was applied to estimate the time of divergence between Galileo sequences.

To provide an evolutionary context for the results, a phylogenetic tree of the 37 host genomes was reconstructed using the entire set of BUSCO genes shared among them. Their amino acid sequences were aligned with MUSCLE (Edgar, 2004Edgar RC (2004) MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 32:1792-1797. ) and refined with trimAl (Capella-Gutiérrez et al., 2009Capella-Gutiérrez S, Silla-Martínez JM and Gabaldón T (2009) trimAl: A tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25:1972-1973.), implemented in a pipeline written by McGowan (2020McGowan J (2020) jamiemcg/BUSCO_phylogenomics: BUSCO v4 , McGowan J (2020) jamiemcg/BUSCO_phylogenomics: BUSCO v4 , https://zenodo.org/records/7334954 (accessed 8 April 2023)
https://zenodo.org/records/7334954...
). Scaptodrosophila lebanonensis was included in this analysis as an outgroup. Their phylogenetic relationships were reconstructed under maximum likelihood in IQ-TREE 2 (Minh et al., 2020Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A and Lanfear R (2020) IQ-TREE 2: New models and efficient methods for phylogenetic inference in the genomic era. Mol Biol Evol 37:1530-1534. ), with the best substitution model selected based on AIC scores (flags --m and --merit). Branch supports were estimated by applying 1,000 replicates of ultrafast bootstrap.

Results

Search for canonical Galileo copies

Sequences of Galileo were masked in all analyzed genomes (Table S1 Table S1 - List and taxonomy of Drosophilidae species included in this study, including positive results from BLASTn searches of the Galileo transposase sequences and accession numbers to the genome assembly and short-read sequencing data on NCBI. ), except for D. ercepeae and D. nannoptera - which belong to the melanogaster and nannoptera groups, respectively. Assemblies showed satisfactory levels of completeness, with the majority having more than 90% of single-copy orthologous genes (S). The exception was eight species, with S percentages ranging from 70% to 90% (see Table S4 Table S4 - Statistics of assembly completeness for each analyzed genome. ). In the second round of searches, conducted using local BLASTn with TPases as queries, 37 species yielded positive hits after the filtering process (Table S1 Table S1 - List and taxonomy of Drosophilidae species included in this study, including positive results from BLASTn searches of the Galileo transposase sequences and accession numbers to the genome assembly and short-read sequencing data on NCBI. ). The positive results in the BLASTn search were limited to species within the Drosophila and Lordiphosa genera (Drosophilinae subfamily, Drosophilini tribe). All identified TPase sequences exhibited mutations, including stop codons, coding frame shifts, or both.

Phylogenetic analysis and abundance of Galileo sequences

The final sizes of nucleotide and amino acid alignments were 1,035 bp and 345 amino acids, respectively. The best amino acid substitution model was JTT+G4+F, based on the Akaike Information Criterion (AIC). Every copy of Galileo found in the genomes was placed in the same clade as its query. Major clades exhibited strong node support (PP > 0.95), with exceptions mainly observed among intraspecific sequences.

The query Dana\Galileo recovered three clades: the first two (yellow, Figure S1 Figure S1 - Phylogenetic relationships between sequences of Galileo found across genomes of Drosophilidae, reconstructed through Bayesian Inference. ) containing sequences found in genomes of the melanogaster group (in which D. ananassae is phylogenetically placed); and the third (orange clade, Figure S1 Figure S1 - Phylogenetic relationships between sequences of Galileo found across genomes of Drosophilidae, reconstructed through Bayesian Inference. ) containing sequences found in species of the montium group, along with Lordiphosa collinella and L. stackelbergi (pink sequences, Figure S1 Figure S1 - Phylogenetic relationships between sequences of Galileo found across genomes of Drosophilidae, reconstructed through Bayesian Inference. ). Dwil\Galileo clustered homologous sequences found in the willistoni group (light pink sequences, Figure S1 Figure S1 - Phylogenetic relationships between sequences of Galileo found across genomes of Drosophilidae, reconstructed through Bayesian Inference. ), along with its sister saltans group (blue sequences, Figure S1 Figure S1 - Phylogenetic relationships between sequences of Galileo found across genomes of Drosophilidae, reconstructed through Bayesian Inference. ). On the other hand, the sequences of Galileo found by Dvir\Galileo (green clade, Figure S1 Figure S1 - Phylogenetic relationships between sequences of Galileo found across genomes of Drosophilidae, reconstructed through Bayesian Inference. ) in species of the virilis group formed a sister clade (PP = 1.0) to those of the willistoni and saltans groups. Finally, Dper\Galileo recovered Galileo from species belonging to the obscura group (red clade, Figure S1 Figure S1 - Phylogenetic relationships between sequences of Galileo found across genomes of Drosophilidae, reconstructed through Bayesian Inference. ), and Dmoj\Galileo retrieved sequences in D. mojavensis (purple clade, Figure S1 Figure S1 - Phylogenetic relationships between sequences of Galileo found across genomes of Drosophilidae, reconstructed through Bayesian Inference. ). The abundance of Galileo sequences in these species, as assessed by the coverage analysis in RepeatProfiler, showed that the TPase region had lower coverage than that of TIRs in all cases (Figures 1 and S2-S7 Figure S2 - Normalized coverage graphs for Dana\Galileo used for searching transposase sequences across genomes of Drosophilidae. ).

Figure 1 -
Coverage graphs for six queries of Galileo against its corresponding species: (A) Dana\Galileo in Drosophila ananassae; (B) Dbuz\Galileo in D. buzzatii; (C) Dmoj\Galileo in D. mojavensis; (D) Dper\Galileo in D. persimilis; (E) Dvir\Galileo in D. virilis; and (F) Dwil\Galileo in D. willistoni. Colors correspond to the coverage scale on the right side of each graph. Axis X corresponds to base pairs positions.

Inference of HTT events

Two major incongruences were found between host species (Figure 2A) and Galileo phylogenies. The first is the similarity of elements found in Lordiphosa collinella and Lordiphosa stackelbergi with species of the montium group (Figure 2B). The second incongruence (Figure 2C) is the clade formed by virilis (Drosophila subgenus) and willistoni plus saltans groups (Sophophora subgenus). No signals of HTT events were detected (p-value > 0.05) between the species of the virilis group and the willistoni and saltans groups (Figure 2D). However, HTT was detected (p-value < 0.05) between L. collinella and L. stackelbergi and species of the montium group. Signals were also detected between the melanogaster and montium groups, both belonging to the Sophophora subgenus (Figure 2E). Estimates of divergence times (Table S5 Table S5 - Statistically significant results of pairwise comparisons of horizontal transposon transfer performed with vhica at HTT-DB. ) span from ~679 thousand years ago (D. auraria x L. stackelbergi) to ~6 million years ago (D. carrolli × D. watanabei).

Figure 2 -
(A) Ultrametric tree showing the phylogenetic relationships between species harboring nearly complete transposases, assessed through maximum likelihood. Ultrafast bootstrap (UFboot) not shown, as for all nodes UFboot = 100. (B and C) Majority-rule consensus tree showing the phylogenetic relationships between sequences of Galileo, (B) found in genomes of the montium group of Drosophila and species of the Lordiphosa genus, and (C) found in genomes of the saltans, virilis and willistoni groups of Drosophila; numbers next to each node reflect its posterior probability support. (D and E) Results of the horizontal transposon transfer (HTT) analysis in vhica, between (D) saltans, virilis and willistoni groups of Drosophila; and (E) Lordiphosa genus and melanogaster and montium groups of Drosophila. (D and E) Red squares represent statistically significant (P < 0.05) pairwise comparisons between sequences of Galileo, indicating a HTT event. Phylogenetic relationships between host genomes are shown by ultrametric trees drawn on the external sides of each graph.

Discussion

The 163 genomes analyzed in this study provided a broader sampling across Drosophilidae when compared to previous studies (Marzo et al., 2008Marzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962. ; Acurio, 2015Acurio AE (2015) Coevolutionary analysis of the transposon Galileo in the genus Drosophila. M. Sc. Thesis, Autonomous University of Barcelona, 219 p.), including many different taxonomic levels. We were able to search for Galileo in the genomes of the two subfamilies - Drosophilinae and, for the first time, Steganinae. Furthermore, our sampling included two tribes of the first (Colocasiomyini and Drosophilini) and two tribes of the latter (Gitonini and Steganini). Indeed, the Drosophila genus is a paraphyletic lineage due to the offshoot of several genera within its phylogenetic tree (Suvorov et al., 2022Suvorov A, Kim BY, Wang J, Armstrong EE, Peede D, D’Agostino ERR, Price DK, Waddell PJ, Lang M, Courtier-Orgogozo V et al. (2022) Widespread introgression across a phylogeny of 155 Drosophila genomes. Curr Biol 32:111-123. ); e.g., the Lordiphosa genus is placed within the Sophophora subgenus as a sister lineage to the Neotropical clade, which includes the saltans and willistoni groups (Figure 2D).

Galileo is fragmentally widespread in Drosophilidae

The majority of Galileo sequences recovered in our study consisted of fragments. Indeed, high levels of structural dynamism in Galileo have been described both within and between genomes, as TIRs presented variable sizes (see review in Marzo et al., 2008Marzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962. ). Therefore, our results suggest that the canonical Galileo is widespread and abundant in the genomes of Drosophilidae, although its copies are potentially defective. Given the lack of coding for a transposase, these copies would be incapable of autonomous transposition, remaining as relics - as in the case of Miniature Inverted-repeat Transposable Elements (MITEs).

The hypothesis of classifying these fragmented copies as MITEs of Galileo in D. mojavensis was considered by Marzo et al. (2013aMarzo M, Bello X, Puig M, Maside X and Ruiz A (2013a) Striking structural dynamism and nucleotide sequence variation of the transposon Galileo in the genome of Drosophila mojavensis. Mob DNA 4:6. ), but was discarded by those authors because the sequences were longer and had a lower copy number compared to typical MITEs. However, our analysis of normalized coverage suggested the opposite; highly amplified short segments of Galileo TIRs were detected (Figures 1 and S2-S7 Figure S2 - Normalized coverage graphs for Dana\Galileo used for searching transposase sequences across genomes of Drosophilidae. ), consistent with the size of MITEs. In D. virilis, for example, the TPase segment of Dvir\Galileo had a low coverage (~10X) while its TIRs had a coverage of < 200X (Figure 1E). Although strong evidence was found, further characterization is still needed to assist in the classification of these short canonical sequences as MITEs.

Interestingly, Galileo seems to be highly amplified in Neotropical species. Among the 15 species with the highest copy number (Figure 3; Table S1 Table S1 - List and taxonomy of Drosophilidae species included in this study, including positive results from BLASTn searches of the Galileo transposase sequences and accession numbers to the genome assembly and short-read sequencing data on NCBI. ), eight are endemic to the Neotropical region: D. mojavensis, D. sturtevanti, D. willistoni, D. paulistorum, D. navojoa, and D. buzzatii, D. tropicalis, and D. montana (listed from the highest to the lowest copy number). In fact, the heterogeneity found across the Neotropical region provides innumerous distinct environments, challenging the survival of species (Miranda et al., 2022Miranda FR, Machado AF, Clozato CL and Silva SM (2022) Nine biomes and nine challenges for the conservation genetics of Neotropical species, the case of the vulnerable giant anteater (Myrmecophaga tridactyla). Biodivers Conserv 31:2515-2541.). Such environments also impact genomes, as expanding into new areas may relieve the epigenetic silencing or control of TEs, leading to their mobilization and amplification (Gregory, 2001Gregory TR (2001) Coincidence, coevolution, or causation? DNA content, cell size, and the C-value enigma. Biol Rev Camb Philos Soc 76:65-101. ; Rebollo et al., 2010Rebollo R, Horard B, Hubert B and Vieira C (2010) Jumping genes and epigenetics: Towards new species. Gene 454:1-7. ; Antoniolli et al., 2023Antoniolli HRM, Deprá M and Valente VLS (2023) Patterns of genome size evolution versus fraction of repetitive elements in statu nascendi species: the case of the willistoni subgroup of Drosophila (Diptera, Drosophilidae). Genome 66:193-201. ). For instance, D. willistoni - which harbors an exceptional diversity of Galileo (Gonçalves et al., 2014Gonçalves JW, Valiati VH, Delprat A, Valente VL and Ruiz A (2014) Structural and sequence diversity of the transposon Galileo in the Drosophila willistoni genome. BMC Genomics 15:792. ) - is distributed throughout the Neotropical region, and TEs differentially populate its genomes (Bertocchi et al., 2022Bertocchi NA, Oliveira TD, Deprá M, Goñi B and Valente VLS (2022) Interpopulation variation of transposable elements of the hAT superfamily in Drosophila willistoni (Diptera: Drosophilidae): in-situ approach. Genet Mol Biol 45:e20210287. ).

Figure 3 -
Figure 3 - Number of sequences (X axis) masked as Galileo elements by RepeatMasker for the top 15 species (Y axis) with the highest number of sequences. Species highlighted in bold are endemic to the Neotropical region.

Signals of HTT in the Sophophora subgenus

The overall congruence between the phylogeny of Galileo and that of its host genomes, in terms of clustering species of the same group into the same clade (Figure S1 Figure S1 - Phylogenetic relationships between sequences of Galileo found across genomes of Drosophilidae, reconstructed through Bayesian Inference. ), may be explained by vertical transmission (Acurio, 2015Acurio AE (2015) Coevolutionary analysis of the transposon Galileo in the genus Drosophila. M. Sc. Thesis, Autonomous University of Barcelona, 219 p.). However, the observed incongruence involving copies found in Lordiphosa and species of the montium group (Figure 2B) was confirmed as horizontal transfer (HTT) event (Figure 2E). The Lordiphosa genus is actually a sister lineage to the willistoni group, and its MRCA with the montium group diverged around 40 million years ago (Mya) (Suvorov et al., 2022Suvorov A, Kim BY, Wang J, Armstrong EE, Peede D, D’Agostino ERR, Price DK, Waddell PJ, Lang M, Courtier-Orgogozo V et al. (2022) Widespread introgression across a phylogeny of 155 Drosophila genomes. Curr Biol 32:111-123. ). In this case, the oldest HTT event between them (L. stacklbergi × D. punjabiensis) is estimated to have occurred at around 2.2 Mya; much more recent than their MRCA.

Other cases of HTT involved the melanogaster and montium groups, whose MRCA diverged around 20 Mya (Suvorov et al., 2022Suvorov A, Kim BY, Wang J, Armstrong EE, Peede D, D’Agostino ERR, Price DK, Waddell PJ, Lang M, Courtier-Orgogozo V et al. (2022) Widespread introgression across a phylogeny of 155 Drosophila genomes. Curr Biol 32:111-123. ); also much older than the oldest HTT detected between them (around 6 Mya for D. carrolli × D. watanabei). The species involved with HTT events occur in sympatry, mainly in the Palearctic region of Asia (TaxoDros v1.04TaxoDros. The database on Taxonomy of Drosophilidae v1.04 , 04 , https://www.taxodros.uzh.ch/ (accessed 10 April 2023)
https://www.taxodros.uzh.ch/...
) - which permits a niche overlap. Addittionaly, Galileo exhibits a patchy distribution both in the Lordiphosa genus and the melanogaster and montium groups (Table S1 Table S1 - List and taxonomy of Drosophilidae species included in this study, including positive results from BLASTn searches of the Galileo transposase sequences and accession numbers to the genome assembly and short-read sequencing data on NCBI. ); in this case, the TE is present in some species but absent in another closely related one(s).

Furthermore, a specific THAP binding site for the Galileo transposase was identified at the 3’ end TIRs (Marzo et al., 2013bMarzo M, Liu D, Ruiz A and Chalmers R (2013b) Identification of multiple binding sites for the THAP domain of the Galileo transposase in the long terminal inverted-repeats. Gene 525:84-91. ). The sequences of Galileo found in these species involved in HTT cases presented highly conserved and amplified 3’ TIRs (Figures 1 and S2-S7 Figure S2 - Normalized coverage graphs for Dana\Galileo used for searching transposase sequences across genomes of Drosophilidae. ), providing further support for the plausibility of such HTT events. Nonetheless, the successful establishment of a TE in new genomes is highly dependent on its transposition rate (Le Rouzic and Capy, 2005Le Rouzic A and Capy P (2005) The first steps of transposable elements invasion: Parasitic strategy vs. genetic drift. Genetics 169:1033-1043. ), as it must avoid being lost in the population due to genetic drift (Blumenstiel, 2019Blumenstiel JP (2019) Birth, school, work, death, and resurrection: The life stages and dynamics of transposable element proliferation. Genes 10:336. ). While L. stackelbergi presented a low number of sequences (49 sequences), L. collinella harbors more than 680 sequences (Table S1 Table S1 - List and taxonomy of Drosophilidae species included in this study, including positive results from BLASTn searches of the Galileo transposase sequences and accession numbers to the genome assembly and short-read sequencing data on NCBI. ), similar to D. buzzatii (604 sequences), in which Galileo was first described. Many other cases of low copy number were also detected (Table S1 Table S1 - List and taxonomy of Drosophilidae species included in this study, including positive results from BLASTn searches of the Galileo transposase sequences and accession numbers to the genome assembly and short-read sequencing data on NCBI. ), and the smallest include D. ambigua (10), D. punjabiensis (37). and D. watanabei (58). The process of stochastic loss of an element may explain both its patchy distribution and low copy number (Blumenstiel, 2019Blumenstiel JP (2019) Birth, school, work, death, and resurrection: The life stages and dynamics of transposable element proliferation. Genes 10:336. ), as observed in mariner-like elements in Drosophila (Lohe et al., 1995Lohe AR, Moriyama EN, Lidholm DA, Hartl DL (1995) Horizontal transmission, vertical inactivation, and stochastic loss of mariner-like transposable elements. Mol Biol Evol 12:62-72.) and Rex elements in the ray-finned fish Characidium (Pucci et al., 2018Pucci MB, Nogaroto V, Moreira-Filho O and Vicari R (2018) Dispersion of transposable elements and multigene families: Microstructural variation in Characidium (Characiformes: Crenuchidae) genomes. Genet Mol Biol 41:585-592.).

Lineage sorting explains the similarity between the saltans, virilis and willistoni groups

Marzo et al. (2008Marzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962. ) described a high similarity between the copies found in the genomes of D. virilis and D. willistoni. Interestingly, the first belongs to the Drosophila subgenus, while the latter belongs to the Sophophora subgenus - their MRCA diverged around 49.9 Mya (Suvorov et al., 2022Suvorov A, Kim BY, Wang J, Armstrong EE, Peede D, D’Agostino ERR, Price DK, Waddell PJ, Lang M, Courtier-Orgogozo V et al. (2022) Widespread introgression across a phylogeny of 155 Drosophila genomes. Curr Biol 32:111-123. ). Acurio (2015Acurio AE (2015) Coevolutionary analysis of the transposon Galileo in the genus Drosophila. M. Sc. Thesis, Autonomous University of Barcelona, 219 p.) later confirmed this close relationship, identifying it along with the guarani and tripunctata groups (Drosophila subgenus). Our results further corroborate both studies by expanding the sample size to include D. littoralis and D. novamexicana (virilis group).

Interestingly, Galileo sequences found in each of these two groups clustered into sister clades that corresponded to their host species, with the addition of sequences from the saltans group in the latter. This clade (virilis + saltans + willistoni) was the first to split in the evolution of Galileo - also congruent with Marzo et al. (2008Marzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962. ). These authors also proposed two explanations for the incongruence between the phylogenies of Galileo and its host genomes: lineage sorting with ancestral HTT (Acurio, 2015Acurio AE (2015) Coevolutionary analysis of the transposon Galileo in the genus Drosophila. M. Sc. Thesis, Autonomous University of Barcelona, 219 p.); or horizontal transfer itself. As no signal of HTT was detected between or within these three species groups (Figure 2A), lineage sorting is a plausible explanation (Cummings, 1994Cummings MP (1994) Transmission patterns of eukaryotic transposable elements: Arguments for and against horizontal transfer. Trends Ecol Evol 9:141-145.). In this case, the transposon is vertically transmitted, but its copies coalesce prior to the split between the host species (Tenaillon et al., 2010Tenaillon MI, Hollister JD and Gaut BS (2010) A triptych of the evolution of plant transposable elements. Trends Plant Sci 15:471-478.) or are differentially lost along the branches of the species tree (Marzo et al., 2008Marzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962. ).

Conclusions

The evolutionary history of Galileo in Drosophilidae is marked mostly by vertical and possibly ancient horizontal transmissions, as identified by Acurio (2015Acurio AE (2015) Coevolutionary analysis of the transposon Galileo in the genus Drosophila. M. Sc. Thesis, Autonomous University of Barcelona, 219 p.), with stochastic loss through genetic drift occurring while species diverged. In addition, its high fragmentation level is compatible with the characteristics of MITEs, although a thorough characterization is still needed to confirm this. Galileo found favorable conditions for its amplification in the heterogeneous Neotropical region, with an astounding copy number detected in Drosophilidae species inhabiting this area. Finally, considering the potential of Galileo to induce chromosomal rearrangements and their evolutionary implications, the HTT described between Lordiphosa and the montium group, and between the latter and the melanogaster group, these results raise an intriguing question (Alfredo Ruiz, personal communication): could evolution be infectious?

Acknowledgements

The authors thank Dr. Arnaud Le Rouzic and Dr. Gabriel L. Wallau for technical assistance with the vhica analysis and Dr. Alfredo Ruiz for inspiring comments in early stages of this study. This study was supported by the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), under scholarship No. 141319/2020-8, and research productivity grant No. 312781/2018-0.

References

  • Acurio AE (2015) Coevolutionary analysis of the transposon Galileo in the genus Drosophila M. Sc. Thesis, Autonomous University of Barcelona, 219 p.
  • Antoniolli HRM, Deprá M and Valente VLS (2023) Patterns of genome size evolution versus fraction of repetitive elements in statu nascendi species: the case of the willistoni subgroup of Drosophila (Diptera, Drosophilidae). Genome 66:193-201.
  • Bailly-Bechet M, Haudry A and Lerat E (2014) “One code to find them all”: A perl tool to conveniently parse RepeatMasker output files. Mob DNA 5:13.
  • Bertocchi NA, Oliveira TD, Deprá M, Goñi B and Valente VLS (2022) Interpopulation variation of transposable elements of the hAT superfamily in Drosophila willistoni (Diptera: Drosophilidae): in-situ approach. Genet Mol Biol 45:e20210287.
  • Blumenstiel JP (2019) Birth, school, work, death, and resurrection: The life stages and dynamics of transposable element proliferation. Genes 10:336.
  • Bourque G, Burns KH, Gehring M, Gorbunova V, Seluanov A, Hammell M, Imbeault M, Izsvák Z, Levin HL, Macfarlan TS, Mager DL and Feschotte C (2018) Ten things you should know about transposable elements. Genome Biol 19:199.
  • Cáceres M, Ranz JM, Barbadilla A, Long M and Ruiz A (1999) Generation of a widespread Drosophila inversion by a transposable element. Science 285:415-418.
  • Canapa A, Barucca M, Biscotti MA, Forconi M and Olmo E (2015) Transposons, genome size, and evolutionary insights in animals. Cytogenet Genome Res 147:217-239.
  • Capella-Gutiérrez S, Silla-Martínez JM and Gabaldón T (2009) trimAl: A tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25:1972-1973.
  • Carvalho TL, Cordeiro J, Vizentin-Bugoni J, Fonseca PM, Loreto ELS and Robe LJ (2023) Horizontal transposon transfer and their ecological drivers: The case of flower-breeding Drosophila. Genome Biol Evol 15:evad068.
  • Casals F, Cáceres M and Ruiz A (2003) The foldback-like transposon Galileo is involved in the generation of two different natural chromosomal inversions of Drosophila buzzatii Mol Biol Evol 20:674-685.
  • Castresana J (2000) Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol 17:540-552.
  • Chen S, Zhou Y, Chen Y and Gu J(2018) fastp: an ultra-fast all-in-one FASTQ preprocessor. Bioinformatics 34:i884-i890.
  • Cummings MP (1994) Transmission patterns of eukaryotic transposable elements: Arguments for and against horizontal transfer. Trends Ecol Evol 9:141-145.
  • Colonna Romano N and Fanti L (2022) Transposable elements: Major players in shaping genomic and evolutionary patterns. Cells 11:1048.
  • Darriba D, Posada D, Kozlov AM, Stamatakis A, Morel B and Flouri T (2020) ModelTest-NG: A new and scalable tool for the selection of DNA and protein evolutionary models. Mol Biol Evol 37:291-294.
  • Delprat A, Negre B, Puig M and Ruiz A (2009) The transposon Galileo generates natural chromosomal inversions in Drosophila by ectopic recombination. PLoS One 4:e7883.
  • Deprá M, Ludwig A, Valente VL and Loreto EL (2012) Mar, a MITE family of hAT transposons in Drosophila Mob DNA 3:13.
  • Dotto BR, Carvalho EL, Silva AF, Duarte Silva LF, Pinto PM, Ortiz MF and Wallau GL (2015) HTT-DB: Horizontally transferred transposable elements database. Bioinformatics 31:2915-2917.
  • Edgar RC (2004) MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 32:1792-1797.
  • Fattash I, Rooke R, Wong A, Hui C, Luu T, Bhardwaj P and Yang G (2013) Miniature inverted-repeat transposable elements: Discovery, distribution, and activity. Genome 56:475-486.
  • Finnegan DJ (1989) Eukaryotic transposable elements and genome evolution. Trends Genet 5:103-107.
  • Gilbert C and Feschotte C (2018) Horizontal acquisition of transposable elements and viral sequences: Patterns and consequences. Curr Opin Genet Dev 49:15-24.
  • Gonçalves JW, Valiati VH, Delprat A, Valente VL and Ruiz A (2014) Structural and sequence diversity of the transposon Galileo in the Drosophila willistoni genome. BMC Genomics 15:792.
  • Gregory TR (2001) Coincidence, coevolution, or causation? DNA content, cell size, and the C-value enigma. Biol Rev Camb Philos Soc 76:65-101.
  • Kidwell MG and Lisch D (1997) Transposable elements as sources of variation in animals and plants. Proc Natl Acad Sci U S A94:7704-7711.
  • Le Rouzic A and Capy P (2005) The first steps of transposable elements invasion: Parasitic strategy vs. genetic drift. Genetics 169:1033-1043.
  • Lim JK and Simmons MJ (1994) Gross chromosome rearrangements mediated by transposable elements in Drosophila melanogaster Bioessays 16:269-275.
  • Lohe AR, Moriyama EN, Lidholm DA, Hartl DL (1995) Horizontal transmission, vertical inactivation, and stochastic loss of mariner-like transposable elements. Mol Biol Evol 12:62-72.
  • Loreto ELS, Carareto CMA and Capy P (2008) Revisiting horizontal transfer of transposable elements in Drosophila Heredity 100:545-554.
  • Marzo M, Puig M and Ruiz A (2008) The Foldback-like element Galileo belongs to the P superfamily of DNA transposons and is widespread within the Drosophila genus. Proc Natl Acad Sci U S A 105:2957-2962.
  • Marzo M, Bello X, Puig M, Maside X and Ruiz A (2013a) Striking structural dynamism and nucleotide sequence variation of the transposon Galileo in the genome of Drosophila mojavensis Mob DNA 4:6.
  • Marzo M, Liu D, Ruiz A and Chalmers R (2013b) Identification of multiple binding sites for the THAP domain of the Galileo transposase in the long terminal inverted-repeats. Gene 525:84-91.
  • Manni M, Berkeley MR, Seppey M, Simão FA and Zdobnov EM (2021) BUSCO Update: Novel and streamlined workflows along with broader and deeper phylogenetic coverage for scoring of eukaryotic, prokaryotic, and viral genomes. Mol Biol Evol 38(10):4647-4654.
  • Melo ES and Wallau GL (2020) Mosquito genomes are frequently invaded by transposable elements through horizontal transfer. PLOS Genetics 16:e1008946.
  • Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A and Lanfear R (2020) IQ-TREE 2: New models and efficient methods for phylogenetic inference in the genomic era. Mol Biol Evol 37:1530-1534.
  • Miranda FR, Machado AF, Clozato CL and Silva SM (2022) Nine biomes and nine challenges for the conservation genetics of Neotropical species, the case of the vulnerable giant anteater (Myrmecophaga tridactyla). Biodivers Conserv 31:2515-2541.
  • Negm S, Greenberg A, Larracuente AM and Sproul JS (2021) RepeatProfiler: A pipeline for visualization and comparative analysis of repetitive DNA profiles. Mol Ecol Resour 21:969-981.
  • Pace JK, Gilbert C, Clark MS and Feschotte C (2008) Repeated horizontal transfer of a DNA transposon in mammals and other tetrapods. Proc Natl Acad Sci U S A 105:17023-17028.
  • Peccoud J, Loiseau V, Cordaux R and Gilbert C (2017) Massive horizontal transfer of transposable elements in insects. Proc Natl Acad Sci U S A 114:4721-4726.
  • Panaud O (2016) Horizontal transfers of transposable elements in eukaryotes: The flying genes. C R Biol 7-8: 296-299.
  • Pucci MB, Nogaroto V, Moreira-Filho O and Vicari R (2018) Dispersion of transposable elements and multigene families: Microstructural variation in Characidium (Characiformes: Crenuchidae) genomes. Genet Mol Biol 41:585-592.
  • Ranwez V, Douzery EJP, Cambon C, Chantret N and Delsuc F (2018) MACSE v2: Toolkit for the alignment of coding sequences accounting for frameshifts and stop codons. Mol Biol Evol 35:2582-2584.
  • Rebollo R, Horard B, Hubert B and Vieira C (2010) Jumping genes and epigenetics: Towards new species. Gene 454:1-7.
  • Ren L, Huang W, Cannon EKS, Bertioli DJ and Cannon SB (2018) A mechanism for genome size reduction following genomic rearrangements. Front Genet 9:454.
  • Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Höhna S, Larget B, Liu L, Suchard MA and Huelsenbeck JP (2012) MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol 61:539-542.
  • Schaack S, Gilbert C and Feschotte C (2010) Promiscuous DNA: Horizontal transfer of transposable elements and why it matters for eukaryotic evolution. Trends Ecol Evol 25:537-546.
  • Sharp PM and Li WH (1989) On the rate of DNA sequence evolution in Drosophila J Mol Evol 28: 398-402.
  • Suvorov A, Kim BY, Wang J, Armstrong EE, Peede D, D’Agostino ERR, Price DK, Waddell PJ, Lang M, Courtier-Orgogozo V et al (2022) Widespread introgression across a phylogeny of 155 Drosophila genomes. Curr Biol 32:111-123.
  • Tamura K, Stecher G and Kumar S (2021) MEGA11: Molecular Evolutionary Genetics Analysis Version 11. Mol Biol Evol 38:3022-3027.
  • Tenaillon MI, Hollister JD and Gaut BS (2010) A triptych of the evolution of plant transposable elements. Trends Plant Sci 15:471-478.
  • 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.
  • Wallau GL, Capy P, Loreto E, Le-Rouzic A and Hua-Van A (2015) VHICA, a new method to discriminate between vertical and horizontal transposon transfer: Application to the Mariner Family within Drosophila Mol Biol Evol 33:1094-1109.
  • Wells JN and Feschotte C (2020) A field guide to eukaryotic transposable elements. Annu Rev Genet 54:539-561.
  • Wicker T, Sabot F, Hua-Van A, Bennetzen JL, Capy P, Chalhoub B, Flavell A, Leroy P, Morgante M, Panaud O, Paux E, SanMiguel P and Schulman AH (2007) A unified classification system for eukaryotic transposable elements. Nat Rev Genet 8:973-982.

Internet Resources

Supplementary material

The following online material is available for this article:

Figure S1 - Phylogenetic relationships between sequences of Galileo found across genomes of Drosophilidae, reconstructed through Bayesian Inference.

Figure S2 - Normalized coverage graphs for Dana\Galileo used for searching transposase sequences across genomes of Drosophilidae.

Figure S3 - Normalized coverage graphs for Dbuz\Galileo used for searching transposase sequences across genomes of Drosophilidae.

Figure S4 - Normalized coverage graphs for Dmoj\Galileo used for searching transposase sequences across genomes of Drosophilidae.

Figure S5 - Normalized coverage graphs for Dper\Galileo used for searching transposase sequences across genomes of Drosophilidae.

Figure S6 - Normalized coverage graphs for Dvir\Galileo used for searching transposase sequences across genomes of Drosophilidae.

Figure S7 - Normalized coverage graphs for Dwil\Galileo used for searching transposase sequences across genomes of Drosophilidae.

Table S1 - List and taxonomy of Drosophilidae species included in this study, including positive results from BLASTn searches of the Galileo transposase sequences and accession numbers to the genome assembly and short-read sequencing data on NCBI.

Table S2 - List of genes used to normalize the results of profile and abundance of Galileo across the analyzed genomes in this study.

Table S3 - List of genes used for Codon Usage Bias (CUB) comparisons in the analysis of horizontal transposon transfer (HTT) in vhica R package.

Table S4 - Statistics of assembly completeness for each analyzed genome.

Table S5 - Statistically significant results of pairwise comparisons of horizontal transposon transfer performed with vhica at HTT-DB.

Edited by

Associate Editor:

Loreta Brandão de Freitas

Publication Dates

  • Publication in this collection
    29 Mar 2024
  • Date of issue
    2023

History

  • Received
    08 May 2023
  • Accepted
    30 Jan 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