Next Article in Journal
Determining Ancestry between Rodent- and Human-Derived Virus Sequences in Endemic Foci: Towards a More Integral Molecular Epidemiology of Lassa Fever within West Africa
Previous Article in Journal
High-Fat Diet Propelled AOM/DSS-Induced Colitis-Associated Colon Cancer Alleviated by Administration of Aster glehni via STAT3 Signaling Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

“What You Need, Baby, I Got It”: Transposable Elements as Suppliers of Cis-Operating Sequences in Drosophila

by
Roberta Moschetti
1,†,
Antonio Palazzo
2,†,
Patrizio Lorusso
1,
Luigi Viggiano
1 and
René Massimiliano Marsano
1,*
1
Dipartimento di Biologia, Università degli Studi di Bari “Aldo Moro”, Via Orabona 4, 70125 Bari, Italy
2
Laboratory of Translational Nanotechnology, “Istituto Tumori Giovanni Paolo II” I.R.C.C.S, Viale Orazio Flacco 65, 70125 Bari, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 26 December 2019 / Revised: 27 January 2020 / Accepted: 30 January 2020 / Published: 3 February 2020
(This article belongs to the Section Genetics and Genomics)

Abstract

:
Transposable elements (TEs) are constitutive components of both eukaryotic and prokaryotic genomes. The role of TEs in the evolution of genes and genomes has been widely assessed over the past years in a variety of model and non-model organisms. Drosophila is undoubtedly among the most powerful model organisms used for the purpose of studying the role of transposons and their effects on the stability and evolution of genes and genomes. Besides their most intuitive role as insertional mutagens, TEs can modify the transcriptional pattern of host genes by juxtaposing new cis-regulatory sequences. A key element of TE biology is that they carry transcriptional control elements that fine-tune the transcription of their own genes, but that can also perturb the transcriptional activity of neighboring host genes. From this perspective, the transposition-mediated modulation of gene expression is an important issue for the short-term adaptation of physiological functions to the environmental changes, and for long-term evolutionary changes. Here, we review the current literature concerning the regulatory and structural elements operating in cis provided by TEs in Drosophila. Furthermore, we highlight that, besides their influence on both TEs and host genes expression, they can affect the chromatin structure and epigenetic status as well as both the chromosome’s structure and stability. It emerges that Drosophila is a good model organism to study the effect of TE-linked regulatory sequences, and it could help future studies on TE–host interactions in any complex eukaryotic genome.

1. Introduction

Transposable elements (TEs), also known as “jumping genes”, are exceptional modifiers of the genome structure and gene expression. Since their discovery and characterization in eukaryotic genomes in the 1940s [1], TEs have long been regarded as junk DNA, useless and harmful sequences that replicate in the genome with no advantage conferred to the host [2,3]. Nowadays, there is a considerable amount of evidence against the junk DNA hypothesis. With few known exceptions, eukaryotic genomes are densely wrapped with TEs [4] that contribute to their adaptation and evolution [5].
TEs move around in the host genome using self-encoded enzymes that catalyze the transposition reaction. Members of Class I adopt a replicative transposition, directly resulting in an effective increase of the TE copies per genome after the completion of the transposition, while Class II elements perform a conservative transposition, i.e., yield an identical copy number after the transposition.
Members of Class I (retrotransposable elements, RTE) transpose via reverse transcription of an RNA intermediate that is afterward integrated into a new genomic locus. The two main subclasses are represented by the LTR-containing retrotransposons and the non-LTR retrotransposons. Besides their main structural difference (i.e. the presence of LTR, long terminal repeats, terminal directly repeated sequences), they are extremely different in their mechanism of transposition. LTR-retrotransposons perform transposition in a way comparable to that of retroviruses, priming the reverse transcription process with the 3′ end of an endogenous tRNA molecule and two distinct template jumps that allow the completion of the cDNA synthesis [6]. The main enzymatic activities that take part in the replication process (reverse transcriptase and RNAseH), are RTE-encoded. The integrase enzymatic activity completes the retrotransposition with the integration of the new copy. The retrotransposition of non-LTR retrotransposons also relies on the reverse transcriptase activity that in this case is primed by a free 3′ single-strand DNA end at the cleaved insertion site, a mechanism known as target primed reverse transcription (TPRT) [7].
TEs belonging to Class II are also called DNA transposons. Usually, they contain terminal inverted repeats (TIRs), flanking the transposase gene that encodes an integrase essential to perform the transposition step, known as the cut-and-paste mechanism. The transposase excises the donor transposon and inserts it into a new locus through a TE-specific recognition of the TIRs.
For many years, Drosophila melanogaster has been considered as a warhorse for genetic studies. Indeed, the fly cultures’ low management costs together with its short life-cycle, the support earned from more than a century-long story in genetics studies, and the availability of sophisticated toolkits and protocols for genetic investigations [8,9] have strongly consolidated this model organism, making it unparalleled compared to other animal models. Also, the genome of D. melanogaster was one of the earliest sequenced animal genome [10], even in its heterochromatic compartment [11,12]. D. melanogaster is currently widely used as an animal model to study the most diverse aspects of genetics, from basic inheritance to cancer [9], but additional genomic resources are continuously developed for other species of the Drosophila genus that will soon become model species in specific fields of investigation [13,14,15,16,17].

2. Drosophila TEs: A Brief Overview

The earliest hypothesis on the presence of TEs in the genome of D. melanogaster date back to the late 70s, with the observation that repetitive sequences inserted at new sites in vitro [18] and in vivo [19]. During the same years, the “P factor hypothesis” [20]-a transposon-linked explanation of the Drosophila P-M hybrid dysgenesis-was confirmed [21,22]. Shortly after, in the early 1980s, the instability of an eye-color phenotype was associated with the presence of extra DNA inserted in the proximity of the white locus [19,23,24].
After that, many known repetitive sequences proved to be TEs. The molecular characterization of some of them led in the following years to the development of powerful insertional mutagenesis tools such as P-element from D. melanogaster [25,26] and Minos from D. hydei [27].
The D. melanogaster genome sequence draft offers the opportunity to annotate a reference mobilome [28]. Afterwards, few additional transposon families, mainly residing in the heterochromatin or absent in the reference strain, were discovered and characterized [29,30,31]. This set of information has been complemented with the genome sequencing of 69 additional species of the Drosophila genus https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/genome/?term = drosophila-last (accessed on 24 December 2019) and the TE characterization in non-model Drosophila species, leading to the possibility to perform large comparative and evolutionary studies [32,33,34,35]. Figure 1A summarizes the main structural features of the TE types in the genome. The number of families currently annotated in the genome of D. melanogaster as well as in other Drosophila species is reported in Figure 1B.
TEs of both classes occupy roughly 20% of the genome of D. melanogaster. It has been estimated that nearly 30% of the TE complement (20% of the DNA transposons, 21% of non-LTR retrotransposons and 45% of LTR retrotransposons respectively) in D. melanogaster consists of full length and potentially active elements [28].
Usually, all TE families consist of a non-autonomous element in addition to transposition-competent elements. At least a fraction of non-autonomous elements, that usually exceed in number the autonomous one, could be still mobilized by the in trans action of the wild type transposition machinery expressed from autonomous elements. It is believed that trans-mobilized non-autonomous elements are the principal contributors of the dissemination of cis-acting regulatory sequences throughout the genome, inducing transcriptional network rewiring and the alteration of wild type transcriptional patterns [36].
Few Helitron families are also annotated in the reference genomes of sequenced Drosophila species. Helitrons encode a 5′-to-3′ DNA helicase and nuclease/ligase similar to those encoded by rolling-circle replicons, and process a single stranded DNA intermediate that replicate using the rolling-circle replication mechanism.
It is remarkable that no active DNA transposons have been identified in humans and mice [37] nor in the vast majority of mammals, with the exception of some bat species [38,39,40,41], thus limiting the possibility to investigate in these species the short-term effect of insertions mediated by this group of TEs. D. melanogaster as well as other Drosophila species, are therefore promising model organisms for studying the contribution to regulatory sequences by eukaryotic TEs.
Here, we will review the current knowledge on the cis-acting sequences identified in TEs and their effects on gene expression and genome architecture. Their contribution is indeed not limited to sequences that affect (either positively or negatively) the transcription of genes, but extends to sequences with important structural functions given their ability to recruit chromatin proteins. A list of known TE-insertions contributing cis-acting sequences in Drosophila is reported in Table 1.
The contribution in cis-acting sequences provided by TEs is described in below and is summarized in Figure 2.

3. TEs as Promoter Suppliers

The promoter region is defined as a cis-regulatory sequence that assembles the pre-initiation complex (PIC) [80] to recruit the RNA polymerase, which starts the transcription process. Promoters are modular sequences containing transcription factor (TF) binding sites (TFBSs), consisting of short sub-sequences that are recognized, more or less specifically, by TFs. Just like non-mobile genes, TEs need promoters to start transcription. TE-associated promoters are recognized by the same RNA polymerases that operate in the nucleus and thus must contain species-specific promoter motifs in order to assemble the PIC and start transcription. The exceptional mobile ability of these sequences allows the incorporation of new TFBSs in the proximity of promoter-less coding sequences or their juxtaposition to existing promoters. In the first case, new transcripts can be generated from previously non-expressed sequences, such as retroposed pseudogenes, leading to the birth of new genes, a relevant event in the evolution of genomes.
Many cases of transposition-mediated promoter acquisition have been described in Drosophila (Table 1). Elements belonging to both classes of TEs can provide promoter sequences to resident genes, thus originating relevant phenotypes. Besides specific studies demonstrating that individual TE insertions modify the expression of nearby genes, a systematic study by Batut and colleagues suggested that TEs contribute large number of developmentally expressed transcriptional start sites and can distribute pre-assembled cis-regulatory modules in the genome [64].
Furthermore, the promoters of elements belonging to the Bari family [34] have been recently tested for their ability to drive a reporter gene expression in expression vectors [81,82]. While the promoter of LTR retrotransposons such as copia, ZAM and Tirant strongly supported the reporter transcription, the promoter of two DNA transposons, Bari1 and Bari3, turned out to be weak promoters [81]. Surprisingly, the promoters of the Bari transposons show an inter-Domain transcriptional activation [81], which is not displayed by other elements, suggesting that they evolved special features enabling their spread in other genomes. Interestingly, this feature seems to be conserved among the members of the Tc1/mariner superfamily [82].
It has been also shown that many retrotransposons and a few TIR elements are transcribed bi-directionally, starting from internal canonical RNA polymerase II promoters, an observation that deals with their regulation through the RNA interference pathway both in somatic and in germline tissues [83,84].
A recent study performed using bioinformatic prediction coupled with Chip-seq data has revealed a significant enrichment of stress-related TFBSs in TEs [85], definitively supporting the idea that TEs are involved in stress responses.
TEs turned out to be an important source of promoters also in the heterochromatin. Heterochromatic genes are regularly transcribed in Drosophila [86] and their promoters have peculiar structural and functional features compared to euchromatic gene promoters [87]. As proposed by Yasuhara et al. [87] “an attractive possibility is an acquisition of TE-derived promoters given the predominance of TE-like sequences in heterochromatin and the finding that some TE promoters are transcribed in heterochromatin”.

3.1. Enhancers, Silencers and Insulators within TEs

Repeated DNA in the form of a simple or complex minisatellite has been frequently observed in the UTRs of many retrotransposons. This apparently unusual feature has been associated with the ability to form complexes with DNA binding proteins, thus interfering with the transcription of nearby genes by modifying the chromatin status of the locus. Indeed, the tandem repeat DNA within the UTRs of retrotransposons contains arrays of protein binding sites that are associated with either enhancer or insulator functions. An analysis of the UTRs of retrotransposons performed by Minervini and colleagues [50] provided evidence that repeats are commonly found in the Drosophila RTEs.
The first well-characterized function associated with a tandem repeat within a TE was found in the 5′UTR of the copia retrotransposon [88]. Later on, a potent insulator was characterized in the 5′UTR of the gypsy retrotransposon [89]. This is a 350 bp-long sequence consisting of an array of 12 degenerated binding sites for the su(Hw) gene product, a DNA binding protein that determines the insulator function. The potency of the gypsy insulator depends on the amount of su(Hw) binding sites [90]. Another efficient insulator has been characterized in the LTR of Idefix [48].
Repeated DNA sequences within the 5′UTR of some retrotransposons may also act as transcriptional enhancers. The first well-characterized retrotransposon-associated enhancer in D. melanogaster is ZAM [48]. ZAM was formerly discovered in a fly strain displaying an unstable eye-color phenotype over time [31].
The LTR-retrotransposon Accord provides an additional example of retrotransposon-associated enhancer. Indeed, fly populations carrying an Accord insertion upstream the cyp6g1 gene are resistant to DDT [91] and nicotine [92] due to the augmented expression of the cyp6g1 gene.
Silencers associated with TEs are poorly described in the scientific literature. However, a silencer has been recently identified and characterized in the D. melanogaster Mos1 element, which belongs to the Tc1/mariner superfamily [93]. This was a bit surprising given the simple and compact structure of the mariner-like elements, which is expected to contain minimal cis-regulatory sequences (e.g., promoters). Also surprising is the evolutionary conservation of silencers in a homologous region of other animal mariner-like elements suggesting that either the silencer function is very ancient, or it might have been raised several times in the mariner elements during animal evolution [93]. Interestingly, the gypsy insulator behaves as a silencer depending on the genetic background [62], a situation that clearly shows the versatility of some TE-linked regulatory sequences. A similar duality has been also highlighted for the ZAM 5′UTR, which behaves as an enhancer when tested in vivo [48] while it acts as an insulator when tested in cultured cells [94].

3.2. Additional Cis-Regulatory Transcriptional Signals within TE

In addition to the above-described functions, TEs are also a source of cis-acting sequences involved in the transcription termination, splicing, and mRNA stability. TE insertions within genes could alter the splicing pattern of primary RNAs depending on the strength of the splicing consensus introduced upon insertion, further increasing the transcriptome variability of the host genome.
POGON1 and Bari1 supply poly-adenylation signals that increase the expression of the gene located upstream their insertion sites, conferring a relevant xenobiotic-resistance phenotype to population bearing such insertions [76] [72]. A transcription termination site has been also described in the 5′UTR of MDG1 element [68].
TEs also provide splicing sites. TEs can modify the exon/intron structure with the introduction of splicing consensus sequences, allowing the incorporation of TE sequence into the mRNA. This phenomenon, called TE exonization [95], has been recently observed in the brain of D. melanogaster in which newly inserted copies of TEs are expressed in a way directly correlates with that of neighboring genes [96].
While splicing is a common post-transcriptional modification in retrotransposons, it is less frequent in members of Class II TEs. P-element is a DNA transposon of D. melanogaster that possess introns that are spliced out with a tissue-specific pattern [97]. Interestingly, spliced RNA isoforms have been described in two active Tc1/mariner elements. While these elements contain intron-less transposase gene, their transcripts are spliced following the canonical (Bari3 [98]) or the unconventional (Bari1 [99]) splicing when over-expressed in experimental model systems.

3.3. Structural Role of Cis-Operating Sequences within TEs

Besides the transcriptional control elements, TEs contain cis-acting sequences that might influence the epigenetic status of the insertion locus. It has been experimentally demonstrated that arrays of three or more P-elements carrying a white reporter gene produce a variegated eye phenotype [100] similar to the classical heterochromatin-induced position-effect variegation [101]. This was the first experimental demonstration of the ability of TEs to seed heterochromatin in virtually every genomic site. Similar behavior was observed for the 1360 transposon and for the invader4 retrotransposon [77] suggesting a broad ability of TEs to induce heterochromatin formation. This ability is granted by the recruitment of heterochromatic proteins such as HP1 at the site of insertion [77]. HP1a, and to some extent HP1b, are key heterochromatin-associated proteins that can interact with a plethora of additional chromatin proteins [102] that can mediate the establishment of repressive chromatin marks. HP1 binding ability has been observed for several TE families [103] [50].
The ability of TEs to introduce new chromatin protein binding sites upon insertion is also relevant in the context of the rewiring of pre-existing transcriptional circuits. An amazing example is the evolutionarily new X chromosome in D. miranda that has accumulated hundreds of MSL complex (male-specific lethal complex) binding sites provided by reiterated insertion of ISY [104], a Helitron element. The MSL complex is recruited to high-affinity chromatin entry sites on the Drosophila male X chromosome and spreads in cis to coordinate the expression of X-linked genes, thus achieving dosage compensation. In D. miranda, the accumulation of ISY has led to switching off the dosage compensation system on the old X chromosome, rewiring it to the newly emerged (neo-X) sex chromosome [71].
In this context, the role of TEs in maintaining the centromeres and the telomeres in Drosophila is well known. A profound cooperation between three LINE-like elements (HeT-A, TAHRE, and TART) allows both their transposition and stability of host chromosomes [105]. In addition, a suggestive hypothesis has been proposed that directly links the organization and function of centromeres of D. melanogaster and D. simulans to the ability of the G2/Jockey-3 transposon to recruit the centromeric protein CENP-A [106].
piRNA clusters (or piRNA loci) are TE-dense heterochromatic loci from which piRNAs are produced to defend the host genome from transposition in the germline [107]. It has been demonstrated that sequences sharing homology to piRNAs operate as cis-acting targets for heterochromatin assembly, which is usually associated with HP1a and H3K9me2/3 [77]. In this context, many TEs can aid in establishing the epigenetic organization of the piRNA loci in Drosophila as well as in other organisms.
Notably, the same gypsy sub-sequence that contains the insulator/silencer function (described above), also functions as MAR/SAR (matrix attachment region/scaffold attachment region), connecting a transcriptional role to a structural role of the gypsy retrotransposon [69]. A sequence displaying MAR function was also identified and characterized in the roo element [70]. Although this aspect is not deeply investigated, these results highlight a cis-structural role of TEs, whose importance is comparable to the role of TEs in centromeres and telomeres.

4. Conclusions and Future Directions: What Can We Still Learn from Drosophila?

Many phenotypes that have been partially characterized in D. melanogaster might be due to the introduction of new cis-regulatory elements resident in transposons.
As an example, there is evidence suggesting that Tirant, an LTR-retrotransposon, could also carry an insulator. The insertion of a defective copy of Tirant in the 21B region, upstream of the GS1 gene (fs(2)PM11-19 mt-gs), has been previously reported to cause a hypomorphic mutation that raises a female-sterile phenotype [49]. Notably, upon insertion, the Tirant-21B element acquired a transcriptional pattern that is the perfect merge of the GS1 and the wild type Tirant patterns (Figure 3).
This situation is compatible with the presence of an insulator function within Tirant that in turn focuses the GS1 enhancer action on its own promoter. From an evolutionary point of view, this could be a strategy that increases the expression of RTE-related genes in specific tissues, such as the germline.
Little is known about the role of TE insertions into the Y chromosome. This entirely heterochromatic chromosome, while dispensable for male fly viability, is essential for male fertility, since it carries genes that are important in spermatogenesis. These are among the largest genes known in D. melanogaster and their transcription mechanism has been recently disclosed [108]. If possible, less is known about the cis-effect exerted by transposon islands that populate the centromeric DNA of D. melanogaster chromosomes. The organization of the centromeric DNA has been determined using mini-chromosomes obtained by progressive deletions, which confirmed previous data on the satellite and transposon islands populating the centromeric DNA in Drosophila. TEs are responsible for neuronal mosaicism in the mushroom bodies of D. melanogaster [109,110]. A similar transposition-based genetic mosaicism was described in the hippocampal neurons of the human brain [111,112], suggesting a conserved role of TEs as the basis of the genetic and functional diversification in the cells of particular neuronal districts in the animal’s brain. Additional effort will be necessary to fully understand how TEs modify the transcriptional profile at the single neuron level and the impact at a larger scale neurological level.
TEs densely populate the centromeric and pericentromeric heterochromatin of D. melanogaster. Their arrangement, in combination with simple and complex satellites, is a feature of the centromeric DNA whose importance is still undeciphered relative to centromere function. An interesting aspect of the presence of TEs in the pericentric heterochromatin is the presence of TE clusters. One of them has a peculiar feature. The Bari1 cluster maps in the h39 region of the second chromosome of D. melanogaster, adjacent to the XbaI repeat that identify the Responder (Rsp) locus [34]. While apparently devoid of functional significance, this region has been proven to be important for some fitness-related performance of the species [113]. However, while the Rsp cluster is highly polymorphic, the Bari1 cluster shows high structural conservation, in terms of copy/number, in many populations tested so far [114]. This could be the result of an unexplored cis-effect on the centromere or on the whole chromosome. In vitro and in vivo studies using DNA adenine methyltransferase identified HP1 binding sites within the Bari1 cluster [75], reinforcing its structural role in the establishment of the heterochromatin domain in the h39 region. New methodological approaches, such as genome editing [115], could enable the discovery of new functions associated to heterochromatic TE-clusters.

Author Contributions

A.P., R.M., P.L., L.V. and R.M.M. collected data. R.M.M. drafted the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding

Acknowledgments

We the Reviewers for the careful and insightful review of our manuscript. We are also grateful to Ruggiero Caizzi for stimulating and constructive discussion on the mobile genetic element universe.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. McClintock, B. The origin and behavior of mutable loci in maize. Proc. Natl. Acad. Sci. USA 1950, 36, 344–355. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Orgel, L.E.; Crick, F.H. Selfish DNA: The ultimate parasite. Nature 1980, 284, 604–607. [Google Scholar] [CrossRef] [PubMed]
  3. Doolittle, W.F.; Sapienza, C. Selfish genes, the phenotype paradigm and genome evolution. Nature 1980, 284, 601–603. [Google Scholar] [CrossRef] [PubMed]
  4. Britten, R.J.; Kohne, D.E. Repeated sequences in DNA. Hundreds of thousands of copies of DNA sequences have been incorporated into the genomes of higher organisms. Science 1968, 161, 529–540. [Google Scholar] [CrossRef]
  5. Cosby, R.L.; Chang, N.C.; Feschotte, C. Host-transposon interactions: Conflict, cooperation, and cooption. Genes. Dev. 2019, 33, 1098–1116. [Google Scholar] [CrossRef] [Green Version]
  6. Arkhipova, I.R.; Mazo, A.M.; Cherkasova, V.A.; Gorelova, T.V.; Schuppe, N.G.; Llyin, Y.V. The steps of reverse transcription of Drosophila mobile dispersed genetic elements and U3-R-U5 structure of their LTRs. Cell 1986, 44, 555–563. [Google Scholar] [CrossRef]
  7. Luan, D.D.; Korman, M.H.; Jakubczak, J.L.; Eickbush, T.H. Reverse transcription of R2Bm RNA is primed by a nick at the chromosomal target site: A mechanism for non-LTR retrotransposition. Cell 1993, 72, 595–605. [Google Scholar] [CrossRef]
  8. St Johnston, D. The art and design of genetic screens: Drosophila melanogaster. Nat. Rev. Genet. 2002, 3, 176–188. [Google Scholar] [CrossRef]
  9. Hales, K.G.; Korey, C.A.; Larracuente, A.M.; Roberts, D.M. Genetics on the Fly: A Primer on the Drosophila Model System. Genetics 2015, 201, 815–842. [Google Scholar] [CrossRef] [Green Version]
  10. Adams, M.D.; Celniker, S.E.; Holt, R.A.; Evans, C.A.; Gocayne, J.D.; Amanatides, P.G.; Scherer, S.E.; Li, P.W.; Hoskins, R.A.; Galle, R.F.; et al. The genome sequence of Drosophila melanogaster. Science 2000, 287, 2185–2195. [Google Scholar] [CrossRef] [Green Version]
  11. Hoskins, R.A.; Smith, C.D.; Carlson, J.W.; Carvalho, A.B.; Halpern, A.; Kaminker, J.S.; Kennedy, C.; Mungall, C.J.; Sullivan, B.A.; Sutton, G.G.; et al. Heterochromatic sequences in a Drosophila whole-genome shotgun assembly. Genome. Biol. 2002, 3, RESEARCH0085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Hoskins, R.A.; Carlson, J.W.; Wan, K.H.; Park, S.; Mendez, I.; Galle, S.E.; Booth, B.W.; Pfeiffer, B.D.; George, R.A.; Svirskas, R.; et al. The Release 6 reference sequence of the Drosophila melanogaster genome. Genome Res. 2015, 25, 455–458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Signor, S.; Seher, T.; Kopp, A. Genomic resources for multiple species in the Drosophila ananassae species group. Fly 2013, 7, 47–57. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Gilbert, D.G. DroSpeGe: Rapid access database for new Drosophila species genomes. Nucleic Acids. Res. 2007, 35, D480–D485. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Celniker, S.E.; Dillon, L.A.L.; Gerstein, M.B.; Gunsalus, K.C.; Henikoff, S.; Karpen, G.H.; Kellis, M.; Lai, E.C.; Lieb, J.D.; MacAlpine, D.M.; et al. Unlocking the secrets of the genome. Nature 2009, 459, 927–930. [Google Scholar] [CrossRef] [Green Version]
  16. Agapite, J.; Albou, L.P.; Aleksander, S.; Argasinska, J.; Arnaboldi, V.; Attrill, H.; Bello, S.M.; Blake, J.A.; Blodgett, O.; Bradford, Y.M. Alliance of Genome Resources C. Alliance of Genome Resources Portal: Unified model organism research platform. Nucleic Acids Res. 2020, 48, D650–D658. [Google Scholar]
  17. Thurmond, J.; Goodman, J.L.; Strelets, V.B.; Attrill, H.; Gramates, L.S.; Marygold, S.J.; Matthews, B.B.; Millburn, G.; Antonazzo, G.; Trovisco, V.; et al. FlyBase 2.0: The next generation. Nucleic Acids Res. 2018, 47, D759–D765. [Google Scholar] [CrossRef] [Green Version]
  18. Potter, S.S.; Brorein, W.J., Jr.; Dunsmuir, P.; Rubin, G.M. Transposition of elements of the 412, copia and 297 dispersed repeated gene families in Drosophila. Cell 1979, 17, 415–427. [Google Scholar] [CrossRef]
  19. Strobel, E.; Dunsmuir, P.; Rubin, G.M. Polymorphisms in the chromosomal locations of elements of the 412, copia and 297 dispersed repeated gene families in Drosophila. Cell 1979, 17, 429–439. [Google Scholar] [CrossRef]
  20. Engels, W.R. Extrachromosomal control of mutability in Drosophila melanogaster. Proc. Natl. Acad. Sci. USA 1979, 76, 4011–4015. [Google Scholar] [CrossRef] [Green Version]
  21. Rubin, G.M.; Kidwell, M.G.; Bingham, P.M. The molecular basis of P-M hybrid dysgenesis: The nature of induced mutations. Cell 1982, 29, 987–994. [Google Scholar] [CrossRef]
  22. Bingham, P.M.; Kidwell, M.G.; Rubin, G.M. The molecular basis of P-M hybrid dysgenesis: The role of the P element, a P-strain-specific transposon family. Cell 1982, 29, 995–1004. [Google Scholar] [CrossRef]
  23. Rasmuson, B.; Montell, I.; Rasmuson, A.; Svahlin, H.; Westerberg, B.M. Genetic instability in Drosophila melanogaster: Evidence for regulation, excision and transposition at the white locus. Mol. Gen. Genet. 1980, 177, 567–570. [Google Scholar] [CrossRef] [PubMed]
  24. Ilyin, Y.V.; Tchurikov, N.A.; Ananiev, E.V.; Ryskov, A.P.; Yenikolopov, G.N.; Limborska, S.A.; Maleeva, N.E.; Gvozdev, V.A.; Georgiev, G.P. Studies on the DNA fragments of mammals and Drosophila containing structural genes and adjacent sequences. Cold Spring Harb. Symp. Quant. Biol. 1978, 42, 959–969. [Google Scholar] [CrossRef] [PubMed]
  25. Spradling, A.C.; Rubin, G.M. Transposition of cloned P elements into Drosophila germ line chromosomes. Science 1982, 218, 341–347. [Google Scholar] [CrossRef] [Green Version]
  26. Rubin, G.M.; Spradling, A.C. Genetic transformation of Drosophila with transposable element vectors. Science 1982, 218, 348–353. [Google Scholar] [CrossRef] [Green Version]
  27. Pavlopoulos, A.; Oehler, S.; Kapetanaki, M.G.; Savakis, C. The DNA transposon Minos as a tool for transgenesis and functional genomic analysis in vertebrates and invertebrates. Genome Biol. 2007, 8 (Suppl. 1), S2. [Google Scholar] [CrossRef] [Green Version]
  28. Kaminker, J.S.; Bergman, C.M.; Kronmiller, B.; Carlson, J.; Svirskas, R.; Patel, S.; Frise, E.; Wheeler, D.A.; Lewis, S.E.; Rubin, G.M.; et al. The transposable elements of the Drosophila melanogaster euchromatin: A genomics perspective. Genome Biol. 2002, 3, RESEARCH0084. [Google Scholar] [CrossRef] [Green Version]
  29. Marsano, R.M.; Marconi, S.; Moschetti, R.; Barsanti, P.; Caggese, C.; Caizzi, R. MAX, a novel retrotransposon of the BEL-Pao family, is nested within the Bari1 cluster at the heterochromatic h39 region of chromosome 2 in Drosophila melanogaster. Mol. Genet. Genomics 2004, 270, 477–484. [Google Scholar] [CrossRef]
  30. Zanni, V.; Eymery, A.; Coiffet, M.; Zytnicki, M.; Luyten, I.; Quesneville, H.; Vaury, C.; Jensen, S. Distribution, evolution, and diversity of retrotransposons at the flamenco locus reflect the regulatory properties of piRNA clusters. Proc. Natl. Acad. Sci. USA 2013, 110, 19842–19847. [Google Scholar] [CrossRef] [Green Version]
  31. Desset, S.; Conte, C.; Dimitri, P.; Calco, V.; Dastugue, B.; Vaury, C. Mobilization of two retroelements, ZAM and Idefix, in a novel unstable line of Drosophila melanogaster. Mol. Biol. Evol. 1999, 16, 54–66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Drosophila 12 Genomes Consortium. Evolution of genes and genomes on the Drosophila phylogeny. Nature 2007, 450, 203–218. [Google Scholar] [CrossRef] [PubMed]
  33. Seetharam, A.S.; Stuart, G.W. Whole genome phylogeny for 21 Drosophila species using predicted 2b-RAD fragments. PeerJ 2013, 1, e226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Palazzo, A.; Lovero, D.; D’Addabbo, P.; Caizzi, R.; Marsano, R.M. Identification of Bari Transposons in 23 Sequenced Drosophila Genomes Reveals Novel Structural Variants, MITEs and Horizontal Transfer. PLoS ONE 2016, 11, e0156014. [Google Scholar] [CrossRef]
  35. Bargues, N.; Lerat, E. Evolutionary history of LTR-retrotransposons among 20 Drosophila species. Mob. DNA 2017, 8, 7. [Google Scholar] [CrossRef] [Green Version]
  36. Loreto, E.L.S.; Depra, M.; Diesel, J.F.; Panzera, Y.; Valente-Gaiesky, V.L.S. Drosophila relics hobo and hobo-MITEs transposons as raw material for new regulatory networks. Genet Mol. Biol. 2018, 41, 198–205. [Google Scholar] [CrossRef] [Green Version]
  37. Gagnier, L.; Belancio, V.P.; Mager, D.L. Mouse germ line mutations due to retrotransposon insertions. Mob. DNA 2019, 10, 15. [Google Scholar] [CrossRef] [Green Version]
  38. Pritham, E.J.; Feschotte, C. Massive amplification of rolling-circle transposons in the lineage of the bat Myotis lucifugus. Proc. Natl. Acad. Sci. USA 2007, 104, 1895–1900. [Google Scholar] [CrossRef] [Green Version]
  39. Ray, D.A.; Feschotte, C.; Pagan, H.J.; Smith, J.D.; Pritham, E.J.; Arensburger, P.; Atkinson, P.W.; Craig, N.L. Multiple waves of recent DNA transposon activity in the bat, Myotis lucifugus. Genome Res. 2008, 18, 717–728. [Google Scholar] [CrossRef] [Green Version]
  40. Ray, D.A.; Pagan, H.J.; Thompson, M.L.; Stevens, R.D. Bats with hATs: Evidence for recent DNA transposon activity in genus Myotis. Mol. Biol. Evol. 2007, 24, 632–639. [Google Scholar] [CrossRef]
  41. Pagan, H.J.; Macas, J.; Novak, P.; McCulloch, E.S.; Stevens, R.D.; Ray, D.A. Survey sequencing reveals elevated DNA transposon activity, novel elements, and variation in repetitive landscapes among vesper bats. Genome Biol. Evol. 2012, 4, 575–585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Chung, H.; Bogwitz, M.R.; McCart, C.; Andrianopoulos, A.; Ffrench-Constant, R.H.; Batterham, P.; Daborn, P.J. Cis-regulatory elements in the Accord retrotransposon result in tissue-specific expression of the Drosophila melanogaster insecticide resistance gene Cyp6g1. Genetics 2007, 175, 1071–1077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Schlenke, T.A.; Begun, D.J. Strong selective sweep associated with a transposon insertion in Drosophila simulans. Proc. Natl. Acad. Sci. USA 2004, 101, 1626–1631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Clemmons, A.W.; Wasserman, S.A. Combinatorial effects of transposable elements on gene expression and phenotypic robustness in Drosophila melanogaster development. G3 Genes Genomes Genet. 2013, 3, 1531–1538. [Google Scholar]
  45. Zhang, L.; Beaucher, M.; Cheng, Y.; Rong, Y.S. Coordination of transposon expression with DNA replication in the targeting of telomeric retrotransposons in Drosophila. EMBO J. 2014, 33, 1148–1158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. George, J.A.; Traverse, K.L.; DeBaryshe, P.G.; Kelley, K.J.; Pardue, M.L. Evolution of diverse mechanisms for protecting chromosome ends by Drosophila TART telomere retrotransposons. Proc. Natl. Acad. Sci. USA 2010, 107, 21052–21057. [Google Scholar] [CrossRef] [Green Version]
  47. Abad, J.P.; De Pablos, B.; Osoegawa, K.; De Jong, P.J.; Martin-Gallardo, A.; Villasante, A. TAHRE, a novel telomeric retrotransposon from Drosophila melanogaster, reveals the origin of Drosophila telomeres. Mol. Biol. Evol. 2004, 21, 1620–1624. [Google Scholar] [CrossRef]
  48. Conte, C.; Dastugue, B.; Vaury, C. Coupling of enhancer and insulator properties identified in two retrotransposons modulates their mutagenic impact on nearby genes. Mol. Cell Biol. 2002, 22, 1767–1777. [Google Scholar] [CrossRef] [Green Version]
  49. Caggese, C.; Barsanti, P.; Viggiano, L.; Bozzetti, M.P.; Caizzi, R. Genetic, molecular and developmental analysis of the glutamine synthetase isozymes of Drosophila melanogaster. Genetica 1994, 94, 275–281. [Google Scholar] [CrossRef]
  50. Minervini, C.F.; Marsano, R.M.; Casieri, P.; Fanti, L.; Caizzi, R.; Pimpinelli, S.; Rocchi, M.; Viggiano, L. Heterochromatin protein 1 interacts with 5’UTR of transposable element ZAM in a sequence-specific fashion. Gene 2007, 393, 1–10. [Google Scholar] [CrossRef]
  51. Wilson, S.; Matyunina, L.V.; McDonald, J.F. An enhancer region within the copia untranslated leader contains binding sites for Drosophila regulatory proteins. Gene 1998, 209, 239–246. [Google Scholar] [CrossRef]
  52. Ding, D.; Lipshitz, H.D. Spatially regulated expression of retrovirus-like transposons during Drosophila melanogaster embryogenesis. Genet. Res. 1994, 64, 167–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Zatsepina, O.G.; Velikodvorskaia, V.V.; Molodtsov, V.B.; Garbuz, D.; Lerman, D.N.; Bettencourt, B.R.; Feder, M.E.; Evgenev, M.B. A Drosophila Melanogaster Strain From Sub-Equatorial Africa Has Exceptional Thermotolerance But Decreased Hsp70 Expression. J. Exp. Biol. 2001, 204, 1869. [Google Scholar] [PubMed]
  54. Lerman, D.N.; Feder, M.E. Naturally occurring transposable elements disrupt hsp70 promoter function in Drosophila melanogaster. Mol. Biol. Evol. 2005, 22, 776–783. [Google Scholar] [CrossRef] [PubMed]
  55. Kimbrell, D.A.; Tojo, S.J.; Alexander, S.; Brown, E.E.; Tobin, S.L.; Fristrom, J.W. Regulation of larval cuticle protein gene expression in Drosophila melanogaster. Dev. Genet. 1989, 10, 198–209. [Google Scholar] [CrossRef] [PubMed]
  56. Kimbrell, D.A.; Berger, E.; King, D.S.; Wolfgang, W.J.; Fristrom, J.W. Cuticle protein gene expression during the third instar of Drosophila melanogaster. Insect Biochem. 1988, 18, 229–235. [Google Scholar] [CrossRef]
  57. Gruber, J.D.; Genissel, A.; Macdonald, S.J.; Long, A.D. How repeatable are associations between polymorphisms in achaete-scute and bristle number variation in Drosophila? Genetics 2007, 175, 1987–1997. [Google Scholar] [CrossRef] [Green Version]
  58. Long, A.D.; Lyman, R.F.; Morgan, A.H.; Langley, C.H.; Mackay, T.F. Both naturally occurring insertions of transposable elements and intermediate frequency polymorphisms at the achaete-scute complex are associated with variation in bristle number in Drosophila melanogaster. Genetics 2000, 154, 1255–1269. [Google Scholar]
  59. Gonzalez, J.; Lenkov, K.; Lipatov, M.; Macpherson, J.M.; Petrov, D.A. High rate of recent transposable element-induced adaptation in Drosophila melanogaster. PLoS Biol. 2008, 6, e251. [Google Scholar] [CrossRef] [Green Version]
  60. Spana, C.; Harrison, D.A.; Corces, V.G. The Drosophila melanogaster suppressor of Hairy-wing protein binds to specific sequences of the gypsy retrotransposon. Genes. Dev. 1988, 2, 1414–1423. [Google Scholar] [CrossRef] [Green Version]
  61. Roseman, R.R.; Swan, J.M.; Geyer, P.K. A Drosophila insulator protein facilitates dosage compensation of the X chromosome min-white gene located at autosomal insertion sites. Development 1995, 121, 3573–3582. [Google Scholar] [PubMed]
  62. Cai, H.N.; Levine, M. The gypsy insulator can function as a promoter-specific silencer in the Drosophila embryo. EMBO J. 1997, 16, 1732–1741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Brönner, G.; Taubert, H.; Jäckle, H. Mesoderm-specific B104 expression in the Drosophila embryo is mediated by internal cis-acting elements of the transposon. Chromosoma 1995, 103, 669–675. [Google Scholar] [PubMed]
  64. Batut, P.; Dobin, A.; Plessy, C.; Carninci, P.; Gingeras, T.R. High-fidelity promoter profiling reveals widespread alternative promoter usage and transposon-driven developmental gene expression. Genome Res. 2013, 23, 169–180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Ding, Y.; Berrocal, A.; Morita, T.; Longden, K.D.; Stern, D.L. Natural courtship song variation caused by an intronic retroelement in an ion channel gene. Nature 2016, 536, 329–332. [Google Scholar] [CrossRef]
  66. Merenciano, M.; Ullastres, A.; de Cara, M.A.; Barron, M.G.; Gonzalez, J. Multiple Independent Retroelement Insertions in the Promoter of a Stress Response Gene Have Variable Molecular and Functional Effects in Drosophila. PLoS Genet. 2016, 12, e1006249. [Google Scholar] [CrossRef]
  67. Tanda, S.; Corces, V.G. Retrotransposon-induced overexpression of a homeobox gene causes defects in eye morphogenesis in Drosophila. EMBO J. 1991, 10, 407–417. [Google Scholar] [CrossRef]
  68. Cherkasova, V.A.; IuV, I. The leader region of the Drosophila transposon MDG1 contains transcription termination sites. Genetika 1990, 26, 1893–1904. [Google Scholar]
  69. Nabirochkin, S.; Ossokina, M.; Heidmann, T. A nuclear matrix/scaffold attachment region co-localizes with the gypsy retrotransposon insulator sequence. J. Biol. Chem. 1998, 273, 2473–2479. [Google Scholar] [CrossRef] [Green Version]
  70. Mamillapalli, A.; Pathak, R.U.; Garapati, H.S.; Mishra, R.K. Transposable element ‘roo’ attaches to nuclear matrix of the Drosophila melanogaster. J. Insect. Sci. 2013, 13, 111. [Google Scholar] [CrossRef]
  71. Ellison, C.E.; Bachtrog, D. Dosage compensation via transposable element mediated rewiring of a regulatory network. Science 2013, 342, 846–850. [Google Scholar] [CrossRef] [Green Version]
  72. Mateo, L.; Ullastres, A.; Gonzalez, J. A transposable element insertion confers xenobiotic resistance in Drosophila. PLoS Genet. 2014, 10, e1004560. [Google Scholar] [CrossRef] [Green Version]
  73. Guio, L.; Barron, M.G.; Gonzalez, J. The transposable element Bari-Jheh mediates oxidative stress response in Drosophila. Mol Ecol. 2014, 23, 2020–2030. [Google Scholar] [CrossRef] [PubMed]
  74. Guio, L.; Vieira, C.; Gonzalez, J. Stress affects the epigenetic marks added by natural transposable element insertions in Drosophila melanogaster. Sci. Rep. 2018, 8, 12197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Van Steensel, B.; Delrow, J.; Henikoff, S. Chromatin profiling using targeted DNA adenine methyltransferase. Nat. Genet. 2001, 27, 304–308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Marsano, R.M.; Caizzi, R.; Moschetti, R.; Junakovic, N. Evidence for a functional interaction between the Bari1 transposable element and the cytochrome P450 cyp12a4 gene in Drosophila melanogaster. Gene 2005, 357, 122–128. [Google Scholar] [CrossRef]
  77. Sentmanat, M.F.; Elgin, S.C. Ectopic assembly of heterochromatin in Drosophila melanogaster triggered by transposable elements. Proc. Natl. Acad. Sci. USA 2012, 109, 14104–14109. [Google Scholar] [CrossRef] [Green Version]
  78. Deprá, M.; da Silva Valente, V.L.; Margis, R.; Loreto, E.L.S. The hobo transposon and hobo-related elements are expressed as developmental genes in Drosophila. Gene 2009, 448, 57–63. [Google Scholar] [CrossRef]
  79. Maside, X.; Bartolome, C.; Charlesworth, B. S-element insertions are associated with the evolution of the Hsp70 genes in Drosophila melanogaster. Curr. Biol. 2002, 12, 1686–1691. [Google Scholar] [CrossRef] [Green Version]
  80. Schilbach, S.; Hantsche, M.; Tegunov, D.; Dienemann, C.; Wigge, C.; Urlaub, H.; Cramer, P. Structures of transcription pre-initiation complex with TFIIH and Mediator. Nature 2017, 551, 204–209. [Google Scholar] [CrossRef]
  81. Palazzo, A.; Caizzi, R.; Viggiano, L.; Marsano, R.M. Does the Promoter Constitute a Barrier in the Horizontal Transposon Transfer Process? Insight from Bari Transposons. Genome Biol. Evol. 2017, 9, 1637–1645. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Palazzo, A.; Lorusso, P.; Miskey, C.; Walisko, O.; Gerbino, A.; Marobbio, C.M.T.; Ivics, Z.; Marsano, R.M. Transcriptionally promiscuous “blurry” promoters in Tc1/mariner transposons allow transcription in distantly related genomes. Mob. DNA 2019, 10, 13. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Russo, J.; Harrington, A.W.; Steiniger, M. Antisense Transcription of Retrotransposons in Drosophila: An Origin of Endogenous Small Interfering RNA Precursors. Genetics 2016, 202, 107–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Harrington, A.W.; Steiniger, M. Bioinformatic analyses of sense and antisense expression from terminal inverted repeat transposons in Drosophila somatic cells. Fly 2016, 10, 1–10. [Google Scholar] [CrossRef] [Green Version]
  85. Villanueva-Canas, J.L.; Horvath, V.; Aguilera, L.; Gonzalez, J. Diverse families of transposable elements affect the transcriptional regulation of stress-response genes in Drosophila melanogaster. Nucleic Acids. Res. 2019, 47, 6842–6857. [Google Scholar] [CrossRef] [Green Version]
  86. Marsano, R.M.; Giordano, E.; Messina, G.; Dimitri, P. A New Portrait of Constitutive Heterochromatin: Lessons from Drosophila melanogaster. Trends Genet. 2019, 35, 615–631. [Google Scholar] [CrossRef]
  87. Yasuhara, J.C.; DeCrease, C.H.; Wakimoto, B.T. Evolution of heterochromatic genes of Drosophila. Proc. Natl. Acad. Sci. USA 2005, 102, 10958–10963. [Google Scholar] [CrossRef] [Green Version]
  88. Sneddon, A.; Flavell, A.J. The transcriptional control regions of the copia retrotransposon. Nucleic Acids Res. 1989, 17, 4025–4035. [Google Scholar] [CrossRef] [Green Version]
  89. Holdridge, C.; Dorsett, D. Repression of hsp70 heat shock gene transcription by the suppressor of hairy-wing protein of Drosophila melanogaster. Mol Cell Biol. 1991, 11, 1894–1900. [Google Scholar] [CrossRef] [Green Version]
  90. Scott, K.C.; Taubman, A.D.; Geyer, P.K. Enhancer blocking by the Drosophila gypsy insulator depends upon insulator anatomy and enhancer strength. Genetics 1999, 153, 787–798. [Google Scholar]
  91. Daborn, P.J.; Yen, J.L.; Bogwitz, M.R.; Le Goff, G.; Feil, E.; Jeffers, S.; Tijet, N.; Perry, T.; Heckel, D.; Batterham, P.; et al. A single p450 allele associated with insecticide resistance in Drosophila. Science 2002, 297, 2253–2256. [Google Scholar] [CrossRef] [PubMed]
  92. Li, X.; Bai, S.; Cass, B.N. Accord insertion in the 5’ flanking region of CYP6G1 confers nicotine resistance in Drosophila melanogaster. Gene 2012, 502, 1–8. [Google Scholar] [CrossRef]
  93. Bire, S.; Casteret, S.; Piegu, B.; Beauclair, L.; Moire, N.; Arensbuger, P.; Bigot, Y. Mariner Transposons Contain a Silencer: Possible Role of the Polycomb Repressive Complex 2. PLoS Genet. 2016, 12, e1005902. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Minervini, C.F.; Ruggieri, S.; Traversa, M.; D’Aiuto, L.; Marsano, R.M.; Leronni, D.; Centomani, I.; Giovanni, D.C.; Viggiano, L. Evidences for insulator activity of the 5’UTR of the Drosophila melanogaster LTR-retrotransposon ZAM. Mol. Genet. Genom. 2010, 283, 503–509. [Google Scholar] [CrossRef] [PubMed]
  95. Sorek, R. The birth of new exons: Mechanisms and evolutionary consequences. RNA 2007, 13, 1603–1608. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Treiber, C.D.; Waddell, S. Transposon expression in the Drosophila brain is driven by neighboring genes and diversifies the neural transcriptome. bioRxiv 2019, 838045. [Google Scholar] [CrossRef] [Green Version]
  97. Laski, F.A.; Rio, D.C.; Rubin, G.M. Tissue specificity of Drosophila P element transposition is regulated at the level of mRNA splicing. Cell 1986, 44, 7–19. [Google Scholar] [CrossRef]
  98. Palazzo, A.; Moschetti, R.; Caizzi, R.; Marsano, R.M. The Drosophila mojavensis Bari3 transposon: Distribution and functional characterization. Mob. DNA 2014, 5, 21. [Google Scholar] [CrossRef] [Green Version]
  99. Palazzo, A.; Marconi, S.; Specchia, V.; Bozzetti, M.P.; Ivics, Z.; Caizzi, R.; Massimiliano, R.M. Functional characterization of the Bari1 transposition system. PLoS ONE 2013, 8, e79385. [Google Scholar] [CrossRef] [Green Version]
  100. Dorer, D.R.; Henikoff, S. Expansions of transgene repeats cause heterochromatin formation and gene silencing in Drosophila. Cell 1994, 77, 993–1002. [Google Scholar] [CrossRef]
  101. Tartof, K.D.; Hobbs, C.; Jones, M. A structural basis for variegating position effects. Cell 1984, 37, 869–878. [Google Scholar] [CrossRef]
  102. Ryu, H.W.; Lee, D.H.; Florens, L.; Swanson, S.K.; Washburn, M.P.; Kwon, S.H. Analysis of the heterochromatin protein 1 (HP1) interactome in Drosophila. J. Proteom. 2014, 102, 137–147. [Google Scholar] [CrossRef] [PubMed]
  103. Greil, F.; van der Kraan, I.; Delrow, J.; Smothers, J.F.; de Wit, E.; Bussemaker, H.J.; van Driel, R.; Henikoff, S.; van Steensel, B. Distinct HP1 and Su(var)3-9 complexes bind to sets of developmentally coexpressed genes depending on chromosomal location. Genes Dev. 2003, 17, 2825–2838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Steinemann, M.; Steinemann, S. Degenerating Y chromosome of Drosophila miranda: A trap for retrotransposons. Proc. Natl. Acad. Sci. USA 1992, 89, 7591–7595. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Frydrychova, R.C.; Biessmann, H.; Mason, J.M. Regulation of telomere length in Drosophila. Cytogenet. Genome Res. 2008, 122, 356–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Chang, C.H.; Chavan, A.; Palladino, J.; Wei, X.; Martins, N.M.C.; Santinello, B.; Chen, C.C.; Erceg, J.; Wu, C.T.; Larracuente, A.M.; et al. Islands of retroelements are major components of Drosophila centromeres. PLoS Biol. 2019, 17, e3000241. [Google Scholar] [CrossRef] [Green Version]
  107. Brennecke, J.; Aravin, A.A.; Stark, A.; Dus, M.; Kellis, M.; Sachidanandam, R.; Hannon, G.J. Discrete small RNA-generating loci as master regulators of transposon activity in Drosophila. Cell 2007, 128, 1089–1103. [Google Scholar] [CrossRef] [Green Version]
  108. Fingerhut, J.M.; Moran, J.V.; Yamashita, Y.M. Satellite DNA-containing gigantic introns in a unique gene expression program during Drosophila spermatogenesis. PLoS Genet. 2019, 15, e1008028. [Google Scholar] [CrossRef] [Green Version]
  109. Perrat, P.N.; DasGupta, S.; Wang, J.; Theurkauf, W.; Weng, Z.; Rosbash, M.; Waddell, S. Transposition-driven genomic heterogeneity in the Drosophila brain. Science 2013, 340, 91–95. [Google Scholar] [CrossRef] [Green Version]
  110. Treiber, C.D.; Waddell, S. Resolving the prevalence of somatic transposition in Drosophila. Elife 2017, 6, e28297. [Google Scholar] [CrossRef] [Green Version]
  111. Baillie, J.K.; Barnett, M.W.; Upton, K.R.; Gerhardt, D.J.; Richmond, T.A.; De Sapio, F.; Bernnan, P.M.; Rizzu, P.; Smith, S.; Fell, M.; et al. Somatic retrotransposition alters the genetic landscape of the human brain. Nature 2011, 479, 534–537. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Upton, K.R.; Gerhardt, D.J.; Jesuadian, J.S.; Richardson, S.R.; Sanchez-Luque, F.J.; Bodea, G.O.; Ewing, A.D.; Salvador-Palomeque, C.; van der Knaap, M.S.; Brennan, P.M.; et al. Ubiquitous L1 mosaicism in hippocampal neurons. Cell 2015, 161, 228–239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Wu, C.I.; Lyttle, T.W.; Wu, M.L.; Lin, G.F. Association between a satellite DNA sequence and the Responder of Segregation Distorter in D. melanogaster. Cell 1988, 54, 179–189. [Google Scholar] [CrossRef]
  114. Caggese, C.; Pimpinelli, S.; Barsanti, P.; Caizzi, R. The distribution of the transposable element Bari-1 in the Drosophila melanogaster and Drosophila simulans genomes. Genetica 1995, 96, 269–283. [Google Scholar] [CrossRef]
  115. Port, F.; Strein, C.; Stricker, M.; Rauscher, B.; Heigwer, F.; Zhou, J.; Beyersdörffer, C.; Frei, J.; Hess, A.; Kern, K.; et al. A large-scale resource for tissue-specific CRISPR mutagenesis in Drosophila. bioRxiv. 2019, 636076. [Google Scholar] [CrossRef]
Figure 1. (A) Structural features of the TEs identified and described in Drosophila. The symbols used are described on the left part of the panel. (B) Overview of the number of TE families belonging to each of the main groups of TEs in D. melanogaster and in other Drosophila species (source: http://flybase.org, last access December 2019).
Figure 1. (A) Structural features of the TEs identified and described in Drosophila. The symbols used are described on the left part of the panel. (B) Overview of the number of TE families belonging to each of the main groups of TEs in D. melanogaster and in other Drosophila species (source: http://flybase.org, last access December 2019).
Biology 09 00025 g001
Figure 2. Schematic representation of the key effects produced by the cis-operating sequences upon TE insertion. Symbols are explained in the box.
Figure 2. Schematic representation of the key effects produced by the cis-operating sequences upon TE insertion. Symbols are explained in the box.
Biology 09 00025 g002
Figure 3. In situ hybridization performed on wild type (panels A and B) or mutant (panel C) ovaries using GS1 (panel A) or Tirant (panels B and C) specific probes. The organization of the relevant locus in the 21B region of the polytene chromosomes of D. melanogaster is provided.
Figure 3. In situ hybridization performed on wild type (panels A and B) or mutant (panel C) ovaries using GS1 (panel A) or Tirant (panels B and C) specific probes. The organization of the relevant locus in the 21B region of the polytene chromosomes of D. melanogaster is provided.
Biology 09 00025 g003
Table 1. List of reported cis-regulatory elements provided by TEs in Drosophila species. Species are indicated with a four-letter code (the first letter specifies the Genus, the following three specify the species).
Table 1. List of reported cis-regulatory elements provided by TEs in Drosophila species. Species are indicated with a four-letter code (the first letter specifies the Genus, the following three specify the species).
SpeciesAffected Gene/LocusTETransposon TypeCis-Regulatory ActivityEffectEvidencesReference
DmelCyp6g1AccordLTREnhancerIncreased xenobiotic resistanceReporter Assay[42]
DsimCyp6g1Docnon-LTREnhancer(?)Increased xenobiotic resistanceDDT resistance and gene over-expression[43]
Dmelbxd1gypsyLTRenhancerDevelopment of thoracic segmentPhenotype assay[44]
DmelTelomeresHeT-Anon-LTRTelomere elongationTelomere maintainanceIn vivo assay[45]
Dmel DvirTelomeresTARTnon-LTRTelomere elongationTelomere maintainanceIn vivo assay[46]
DmelTelomeresTAHREnon-LTRTelomere elongationTelomere maintainanceDNA sequencing[47]
DmelwhiteIdefixLTRinsulatorEye pigmentationPhenotype assay[48]
DmelGS1TirantLTRNDND Direct assay [49]
DmelNAZAMLTRHP1 bindingChromatin state determinationIn vitro assay[50]
DmelNAcopiaLTREnhancerReporter expression Direct assay[51]
Dmeldevelopment genes17.6, 297, 412, 1731, 3S18, blood, copia, gypsy, HMS Beagle
Kermit/fleamdg1
mdg3
opus
B104/roo springer
LTRCis-regulatory sequencesalterations of gene expression during embryogenesisExpression analyses[52]
DmelHsp70BajockeyLTRCis-regulatory sequencessuppression of the deleterious phenotypes of Hsp70.Phenotypic assay, expression analysis[53,54]
Dmel87A7 hsp70 H.M.S. BeagleLTR Unknownsuppression of the deleterious phenotypes of Hsp70.Phenotypic assay, expression analysis [53]
DmelLCP-1
LCP-f2
H.M.S. BeagleLTR enhancer-like elementsTranscriptional activation of LCP genesGenetic variants analyses[55]
[56]
Dmelachaete-scute complextranspacLTRenhancer-like elements variation in bristle numberGenetic variants analyses[57]
[58]
DmelkuzF-elementnon-LTRcis-regulatory Gene up- regulationPopulation analyses[59]
DmelyellowgypsyLTRinsulatorYellow phenotype In vivo analyses[60]
[61]
[62]
DmelNAB104/rooLTRpromoterNAInferred from in vivo assay[63]
DmelTM4SF297LTRpromoterNARNA ligase-mediated 5′-RACE[64]
Dmel152 annotated genesroo
gypsy
Pao
LTRPromoterNARNA ligase-mediated 5′-RACE[64]
DsimSlowpokeShellderLTRAltered splicingCourtship song variationTrait mapping, in vivo CRISPR knockout[65]
DmelCG18446rooLTRalternative transcription start siteincreased expression5′-RACE[66]
DanaOm(10)TOMLTRenhancersEye morphogenesisIn vivo assay[67]
DmelNAMDG1LTRTranscription terminationNATranscriptional analysis[68]
DmelNAgypsyLTRMARNAIn vivo assay[69]
DmelNArooLTRMARNAIn vivo assay[70]
DmirNeo XISYHelitronMSL binding siteDosage compensationDirect assay[71]
DmelHSP70BAP-elementDNASilencerReduction of Hsp70 expression level.Direct phenotypic assay[54]
[53]
DmelCG11699POGON1TIRPoly-A signalIncreased xenobiotic resistance3′ RACE[72]
DmelJheh1, Jheh2Bari1TIRHP1 seedingAntioxidant responsePhenotypic assay[73]
[74]
Dmelh39 regionBari1TIRHP1 bindingChromatin state determinationDirect assay[75]
DmelCyp12a4Bari1TIRpolyA signaldetoxification3′ RACE[76]
DmelNA1360/hoppelTIRHp1 recruitmentHeterochromatin formationIn vivo assay[77]
Dsimhunchback even-skippedhoboVATIRPromoter, transcription factor binding sites (TFBSs)new phenotypeExpression and in situ analyses[36,78]
DmelHsp70BbS-elementTIRcis-regulatoryNApopulation genetics study[79]
DmelrdxS-elementTIRcis-regulatorydown-regulationPopulation analysis[59]
Dmel152 annotated genesTc1
P
hAT Helitron
TIR
Helitron
Promoter
TSS clusters
NARNA ligase-mediated 5′-RACE[64]

Share and Cite

MDPI and ACS Style

Moschetti, R.; Palazzo, A.; Lorusso, P.; Viggiano, L.; Massimiliano Marsano, R. “What You Need, Baby, I Got It”: Transposable Elements as Suppliers of Cis-Operating Sequences in Drosophila. Biology 2020, 9, 25. https://0-doi-org.brum.beds.ac.uk/10.3390/biology9020025

AMA Style

Moschetti R, Palazzo A, Lorusso P, Viggiano L, Massimiliano Marsano R. “What You Need, Baby, I Got It”: Transposable Elements as Suppliers of Cis-Operating Sequences in Drosophila. Biology. 2020; 9(2):25. https://0-doi-org.brum.beds.ac.uk/10.3390/biology9020025

Chicago/Turabian Style

Moschetti, Roberta, Antonio Palazzo, Patrizio Lorusso, Luigi Viggiano, and René Massimiliano Marsano. 2020. "“What You Need, Baby, I Got It”: Transposable Elements as Suppliers of Cis-Operating Sequences in Drosophila" Biology 9, no. 2: 25. https://0-doi-org.brum.beds.ac.uk/10.3390/biology9020025

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

Article Metrics

Back to TopTop