Skip to main content
  • Research article
  • Open access
  • Published:

Evolutionary dynamics of the LTR retrotransposons roo and rooA inferred from twelve complete Drosophila genomes

Abstract

Background

Roo is the most abundant retrotransposon in the fruit fly Drosophila melanogaster. Its evolutionary origins and dynamics are thus of special interest for understanding the evolutionary history of Drosophila genome organization. We here study the phylogenetic distribution and evolution of roo, and its highly diverged relative rooA in 12 completely sequenced genomes of the genus Drosophila.

Results

We identify a total of 164 roo copies, 57 of which were previously unidentified copies that occur in 9 of the 12 genomes. Additionally we find 66 rooA copies in four genomes and remnants of this element in two additional genomes. We further increased the number of elements by searching for individual roo/rooA sequence domains. Most of our roo and rooA elements have been recently inserted. Most elements within a genome are highly similar. A comparison of the phylogenetic tree of our roo and rooA elements shows that the split between roo and rooA took place early in Drosophila evolution. Furthermore there is one incongruency between the species tree and the phylogenetic tree of the roo element. This incongruency regards the placement of elements from D. mojavensis, which are more closely related to D. melanogaster than elements from D. willistoni.

Conclusion

Within genomes, the evolutionary dynamics of roo and rooA range from recent transpositional activity to slow decay and extinction. Among genomes, the balance of phylogenetic evidence, sequence divergence distribution, and the occurrence of solo-LTR elements suggests an origin of roo/rooA within the Drosophila clade. We discuss the possibility of a horizontal gene transfer of roo within this clade.

Background

Transposable elements (TEs) are DNA sequences that have the ability to replicate within a genome using a variety of mechanisms [1]. They are present in almost all eukaryotic genomes, and they play an important role in genome evolution by creating genetic variation through their mobility. Most new transposable element insertions have a negative effect on the host's fitness through insertions in genes or regulatory regions [2, 3], or through ectopic recombination between two copies of an element [4]. For this reason, their transpositial activity is regulated and suppressed by various host encoded mechanisms, like cytosine methylation [5], repeat induced point mutations [6], and expression silencing by RNA interference [7, 8]. The number of active copies also decreases due to random mutations, excision, purifying selection, and stochastic loss [2].

TEs can be divided into two classes (I and II) based on their replication mechanism. Class I elements use an RNA intermediate for transposition and are called retrotransposons. They can be further subdivided into LTR and non-LTR elements, named after the presence or absence of long terminal repeat (LTR) sequences in the element. Class II elements use a DNA intermediate and are therefore called DNA transposons [1]. Although TEs are mainly transmitted vertically from parent to offspring, they also have the ability to invade new genomes and even cross species boundaries through horizontal gene transfer. In D. melanogaster class I and II TEs occupy around 5.35 percent of the euchromatin part of the genome [9], with the LTR element roo being the most abundant element [10, 11]. Roo differs from most other LTR elements in that it encodes an envelope (env) gene in addition to the gag and pol genes encoded by most LTR elements [12]. The origin of roo's env gene is not completely clear. It might have been captured from retroviruses, where the gene encodes an envelope glycoprotein that is important for the infectivity of the virus, or it might have been lost from most LTR elements other than roo. Aside from roo, only a few other LTR elements harbor an env protein. With the exception of the env gene in the LTR element gypsy, the proteins encoded by env in LTR retrotransposons are thought to be non-functional. Roo was first identified by Scherer et al. under the name B104 [13] and by Meyerowitz and Hogness as roo [14]. The element is 9092 basepairs (bp) long. It contains one open reading frame (ORF) with 7083 nucleotides that comprise the gag, pol and env genes. The structure of the element is shown in Figure 1a. The element is strongly expressed during embryogenesis and is responsible for gene mutations [15].

Figure 1
figure 1

Roo element structure and phylogenetic tree of the 12 sequenced Drosophila species. a) The canonical roo element is 9092 basepairs (bp) long, has one 7083 bps long ORF (light gray box), and is flanked by 428/429 bps long LTR sequences (dark gray boxes). The ORF encodes the gag, pol and env genes. b) The phylogenetic tree of the Drosophila species is taken from [43]. The number of identified roo and rooA elements is listed for each species. Roo and rooA elements are highlighted with red and blue bars, respectively.

Frequent ectopic recombinations between roo elements cause chromosome rearrangements [4]. So far, roo has only been identified in a few species belonging to the melanogaster subgroup [16], and one horizontal transfer event between D. melanogaster and D. simulans has been suggested [17]. An anciently diverged relative of roo is rooA, which also has members in D. melanogaster. In this subfamily, the internal coding region is only ~62 percent identical to the coding region of roo. The rooA subfamily was probably active in the D. melanogaster genome ~2.5 myr ago [18]. In contrast to the high copy number of roo, only a few rooA elements were previously identified in the D. melanogaster genome [11]. These elements are mostly short fragments and up to now, no complete rooA element has been identified.

Roo's high abundance makes it a prominent constituent of mobile DNA in Drosophila. This fact, and the existence of a diverged subfamily renders the element worthy of study. We here use the 12 sequenced Drosophila genomes [9] to study the evolutionary dynamics of roo and rooA on different timescales. Specifically, we explore these evolutionary dynamics within and among genomes, including likely transmission routes, and we examine the origin of the common ancestor of all roo/rooA elements.

Results

Distribution of Roo elements

We first used an iterative approach to identify all roo and rooA elements in the currently available 12 sequenced Drosophila genomes [9]. This approach (see Methods for details) used the canonical Drosophila roo and rooA elements as query sequences to identify a series of additional elements in the Drosophila genomes. These additional elements then served as queries in further search rounds, which were continued until no further roo/rooA elements could be identified. To identify more diverged elements, we used the 157 high confidence elements identified through this iterative search as initial queries for a second search with less stringent search parameters. Because roo and rooA query sequences can identify members of both families, we separated all identified elements based on their similarity to the roo and rooA canonical element. While most elements in D. mojavensis correspond to roo elements, one element shared a higher similarity to the canonical rooA element. All other identified rooA elements, however, are much more similar to the rooA canonical sequence than this one element. We mention this element, because its dubious affiliation affects the evolutionary scenarios we discuss below.

Our combined search approach (round one and two) yielded a total of 230 elements, 164 roo and 66 rooA elements, that occurred in nine out of the 12 genomes (see Additional file 1). Columns 2 and 5 of table 1 display the number of roo and rooA elements, respectively, in each genome, as does Figure 1b in the context of the Drosophila phylogeny. Table 1 also lists the number of probably active elements (full-length, no frameshift, no stop-codon). The distribution of elements is different between the two subfamilies. For example, while the D. melanogaster genome is the most roo-rich genome, harboring 107 elements (two thirds of all identified roo elements), it is D. erecta that harbors most rooA elements (44 copies). Overall, we identified roo elements in nine out of the 12 genomes, but found rooA elements in only four genomes. We now briefly highlight some features of roo's distribution. The two species most closely related to D. melanogaster are the sister species D. simulans and D. sechellia. In contrast to the large number of roo copies in D. melanogaster, we find only 16 roo copies in D. sechellia, and only one fragmented copy in D. simulans. We want to note that the genome of D. sechellia is not assembled and that the copy number might, therefore, be lower.

Table 1 Element distribution and insertion time

Like D. simulans and D. melanogaster, D. pseudoobscura and D. persimilis are very closely related (Figure 1b). We found no roo elements in D. pseudoobscura, and only one in D. persimilis. This sole element is highly diverged, contains multiple small ORFs as well as a frameshift mutation, and has only 63 percent of the length of the canonical roo element. The element may thus be a remnant of a functional roo element. The absence of additional roo copies in this genome indicates that roo is in the process of being eliminated from the genome.

The roo copies in D. mojavensis deserve special mention. This species is, together with D. virilis and D. grimshawi, one of the most distantly related species to D. melanogaster. The latter two species harbor no roo elements, but D. mojavensis harbors 12 copies, of which at least one copy may be functional. This tentative assessment is based on the element's complete coding sequence and the absence of frameshifts and stop codons.

With respect to the distribution of rooA, we note that the sister species D. yakuba and D. erecta contain the vast majority of rooA elements. Before our analysis, no rooA element had been identified outside the D. melanogaster genome, and none of the previously identified elements are full length [11]. While none of the elements in D. yakuba are full length, D. erecta contains at least one full length and probably functional element. Compared to other elements, this element has a 271 amino acid duplication in its coding region, but has only one open reading frame (ORF), and shows no frameshift mutations or internal stop codon. We also note that we did not identify any rooA elements in D. melanogaster, the species where it was first identified. The flybase annotation of the D. melanogaster genome (release 4.1) lists five rooA elements. Four of these copies are, however, very short and one element has an extremely diverged coding region, so that our search procedure would not identify it.

Solo-LTRs indicate loss of roo/rooA in some species

An individual LTR retrotransposon can most easily get lost from a genome via a recombination event between the two LTR repeat units. The remnant of such a recombination event is a solo-LTR, a single remaining LTR repeat unit [19]. Instances of solo-LTRs have been reported for roo elements [20]. In order to find out whether some genomes may have lost all their roo or rooA elements, we searched for roo and rooA solo-LTR elements in the twelve genomes.

To identify these elements, we performed a stringent blast search [21], using each LTR sequence of all previously identified roo and rooA elements as queries. We only excluded LTR sequences as queries that were shorter than 200 bp and that contained an "N" (unidentified nucleotide) at more than half of their nucleotides in the genome sequence. This procedure led us to exclude 26 (out of 328) roo LTRs and 22 (out of 132) rooA LTR sequences as queries. Each identified solo-LTR had to show at least 70 percent identity to the query sequence, and deviate in its length by no more than 10 percent from it. Table 1 (columns 3 and 6) shows the number of identified solo-LTR sequences for each genome.

In addition to solo-LTRs in the genomes that contained roo elements, we found 33 solo-LTRs in D. pseudoobscura. These observations validate the above assertion that roo has become extinct from the genome of D. pseudoobscura. A similar extinction process may still be ongoing in the highly mutated sole remaining element of D. persimilis.

In contrast to roo solo-LTRs, rooA solo-LTRs occur only in species within the melanogaster subgroup, with one exception: Solo-LTRs in D. mojavensis, which were all identified by the two LTR sequences of the single rooA element in this species. The number of rooA solo-LTRs varies significantly between the different species in the melanogaster subgroup: Whereas D. melanogaster only harbors 55 copies, D. yakuba harbors more than 3000 solo-LTRs.

The only two genomes that did not contain any solo-LTRs for either roo or rooA are D. virilis and D. grimshawi.

LTR divergence indicates that most roo elements have been recently inserted

Because the two LTR sequences of a retrotransposon are identical at the time of insertion, their divergence can be used to estimate an element's age, the time since insertion [22]. Highly similar LTR sequences suggest that an element was only recently inserted, because too little time elapsed for many mutations to accumulate between them. From alignments of the LTR sequences of each roo/rooA element, excluding elements with unidentified nucleotides in their sequences, we estimated the intra-element LTR divergence. We then used this information, together with the synonymous substitution rate of 0.016 substitutions per site per million years (myr) for Drosophila [23], to estimate the age distribution of roo/rooA elements. The median estimated time since insertion for the elements is shown in Table 1 (column 4 and 7 for roo and rooA elements, respectively) for each genome, and the full age distribution is shown in Figure 2 for roo and rooA in red and blue bars, respectively. These insertion times have to be considered with caution, because it is possible that transposable elements may experience gene conversion [24, 25], which compromise this molecular clock. (For instance, the LTR sequences of D. ananassae are identical despite high divergence in their coding regions, a disagreement that might be explained by gene conversion of the LTR sequences.) We, therefore, also show the distance distribution itself (Additional file 2). We want to point out that the insertion time of an element is not the same as the age of the element in the species. Even if all elements in one species appear recently inserted, the element may have resided in the genome for a long time, for example if there is a high turnover of elements within a genome. The high number of solo-LTRs we observe is evidence for such high turnover.

Figure 2
figure 2

Insertion time distribution. The histograms show the estimated insertion time distribution of roo and rooA (in red and blue bars, respectively), based on the intra-element LTR divergence. Notice the different scales on the y-axis. The vertical red and blue lines indicate the median insertion time in each genome for roo and rooA, respectively. Two roo elements of D. mojavensis had to be excluded because their LTR sequences were falsely identified. The histogram in the bottom right corner shows the insertion times of all roo and rooA elements. myr: million years.

The median age of insertion of all roo elements in this analysis is below the detection limit of 0.05 myr obtainable from divergence data, indicating that most roo elements were inserted less than a million years ago, as indicated in the bottom right panel of Figure 2. However, there are significant differences in the roo age distribution among species. In D. melanogaster, the species with most roo elements, the vast majority of elements are young, a pattern that is most striking with its more than 100 roo elements. Specifically, more than half of the D. melanogaster elements have identical LTR sequences, indicating very recent insertion. The median element age is below the detection limit of 0.05 myr. The oldest identified element has inserted only approximately 0.6 myr ago. The Drosophila melanogaster roo elements thus underwent many recent transpositions. The median age is even lower than that estimated in previous work [10]. The high LTR identity of the roo element is not unusual for LTR elements in D. melanogaster. It also occurs for other LTR elements in Drosophila melanogaster like copia, blood or 412 [10, 20]. In D. melanogaster the median insertion time for every analyzed LTR element family was less than 0.11 myr ago.

In species other than D. melanogaster, the insertion time distributions for roo are highly variable. For example, all five roo elements from D. ananassae have identical LTR sequences. The roo elements in D. sechellia are very similar to those in D. melanogaster, but here only two elements have identical LTR sequences, and the insertion times are more diverse, which suggests that no recent replication activity took place. Finally, in D. mojavensis the median roo insertion age is 0.09 myr.

The age distribution of rooA elements is also quite diverse. D. erecta, the species with the highest abundance of 44 rooA elements, also shows many young elements but the insertion time distribution is broader than for roo in D. melanogaster. For example, ten out of the 44 identified rooA elements in D. erecta have identical LTR sequence, and may have, therefore, been inserted recently, but the oldest element is already approximately 4.1 myr old. The median insertion time, however, is only 0.16 million years ago, which indicates that rooA was recently active in this species, and that it might still be spreading. Despite the young age of most elements, the majority of elements may be non-functional, as suggested by the fact that they have only 25 percent of the length of the canonical element, and that some elements also carry stop codons in the coding region. These observations suggest that passive spreading via intact helper elements plays a role in rooA proliferation.

In contrast to D. erecta, the rooA elements in D. yakuba have a much higher insertion age (median: 1.02 myr). Most D. yakuba rooA elements also have frameshift mutations or stop codons in their coding region. Because D. yakuba also shows no signs of recent transposition activity (Figure 2), we speculate that the rooA element may be in the process of being deleted from this genome. The putative rooA element in D. mojavensis is already 1.5 myr old, in contrast to this genome's roo elements, which have been recently inserted.

Taken together, the age distribution of rooA is broader than that of roo.

Intra- and inter species divergence

We next compared the sequence divergence of different elements within and among species. Such a comparison may yield additional information about the evolutionary dynamics of the roo/rooA elements. Specifically, we estimated the percent similarity between their protein coding sequences using a maximum likelihood approach implemented in the package Phylip[26]. Because similarity estimation requires coding sequence information, we only used the 134 elements (123 roo and 11 rooA elements) with a coding sequence length spanning at least half the coding sequence of the canonical element (1180 aa). Most elements with shorter coding sequence are highly diverged and contain several short ORFs interrupted by stop codons and/or frameshift mutations. As a result of this length threshold, we were able to include 10 (of 16) roo elements from D. sechellia, 97 (of 107) from D. melanogaster, 1 (of 7) from D. yakuba, all from D. ananassae, 3 (of 14) from D. willistoni and 7 (of 12) from D. mojavensis, and no elements from D. simulans, erecta, and D. persimilis. For rooA, we were able to use 2 (of 18) rooA elements from D. yakuba, and 8 (of 44) elements from D. erecta, the single element in D. mojavensis.

Tables 2 and 3 show the percent similarity for roo and rooA elements, respectively, between elements within the same species (diagonal), and among species (off-diagonal). As expected from our LTR analysis, within-species divergence is generally very low, suggesting either recent acquisition or high turnover of roo elements within genomes.

Table 2 Average protein similarity between roo elements
Table 3 Average protein similarity between rooA elements

One conspicuous pattern is that roo elements in D. mojavensis are 10 percent more similar to elements in D. melanogaster than to elements from the more closely related D. ananassae and D. willistoni. This observation is strictly speaking inconsistent with the Drosophila species tree in Figure 1b. However, its weight is reduced if one considers that the similarities between roo elements in D. melanogaster on one hand, and D. ananassae, D. willistoni and D. mojavensis on the other hand, are of comparable magnitude.

Phylogenetic incongruence between roo and species tree

We next constructed and manually inspected a multiple alignment of the protein sequences of roo's and rooA's coding regions. To create a maximum likelihood phylogenetic tree of the roo/rooA elements from this alignment, we used PhyML_aLRT[27], a version of PhyML[28] that incorporates an approximate likelihood ratio test to estimate the statistical support of the tree topology. Furthermore we used the canonical sequence of the BEL element [GenBank:U23420] from D. melanogaster as outgroup. The phylogenetic tree of the roo/rooA elements is shown in Figure 3. Because of the high number of elements the terminal clades are collapsed. The black triangle at the end of each terminal clade indicates the divergence of the elements within a clade or species. Additional file 3 shows the complete tree. Roo and rooA elements are clearly distinguishable in the tree of Figure 3, because they form two distinct clades (highlighted in red and blue, respectively). Three major patterns are evident from this tree. First, roo elements from distantly related genomes are separated in different clades, as one might expect if the mode of transmission is predominantly vertical. Second, elements within one species are usually more closely related to each other than to elements in other species. The only exception are roo and rooA elements in species of the melanogaster subgroup. In these species, the elements are so closely related that the divergence of the genomes that harbor them is not reflected in the elements' divergence. Third, elements within a genome vary greatly in their divergence. For example, the roo elements in D. mojavensis and D. ananassae, show a low divergence (small triangle), the roo elements in D. willistoni show the highest divergence (large triangle). Roo elements in the melanogaster subgroup also seem to be quite diverged. This divergence can be traced back to only 7 (out of the 132) elements. All other elements are very similar, as can be seen in Additional file 3. High divergence of elements can point to different subclasses of elements, such as functional and non-functional elements.

Figure 3
figure 3

Phylogenetic tree of the 230 identified roo and rooA elements. The tree is based on the protein coding sequence of roo and rooA. It was constructed using PhyML with an approximate likelihood ratio test to estimate the statistical support of the tree topology [27], as shown by the numbers at the branches. All branches have a very high support. There is a clear division between roo (red background) and rooA (blue background) elements on the tree. Specifically, all roo elements from one species form a clade and the same holds for rooA. The black triangles at some leaves indicate the divergence of the elements associated with this leaf, with long triangles indicating great divergence. The number in brackets behind each species name indicates how many elements are present in the respective species.

Despite these general patterns, there is one major incongruency between the species tree (Figure 1b) and the element tree (Figure 3). It regards the roo elements. This incongruency occurs in the D. mojavensis genome. This species is one of the species most distantly related to D. melanogaster. Its common ancestor split from the Drosophila melanogaster lineage around 40 myr ago, but the element tree shows that the roo elements in D. mojavensis are more closely related to D. melanogaster than the elements in D. willistoni. An error in tree estimation is made unlikely by the high statistical support of both the D. mojavensis and D. willistoni branches in the roo element tree (Figure 3). The observation raises the possibility of a horizontal transfer of the roo element into D. mojavensis from a Drosophila species after the split from D. willistoni, but before the split from D. persimilis.

Additional fragments validate our earlier findings

The number of roo and rooA copies we identified is probably an underestimate of the real copy number in these genomes caused by our high stringency search. We, therefore, performed two additional searches to get a better copy number estimate.

While the coverage and assembly quality of the D. melanogaster genome is very good, the quality of most of the other genomes is much lower. Even though the sequence coverage is high, the sequence is divided into many small stretches of contiguous sequence (contigs). The length of many contigs is smaller than 10 kb and therefore smaller than the roo/rooA element. Identification of complete elements in these genomic regions is, therefore, not possible. Furthermore, the assembly of repeated regions is challenging. To overcome this limitation, we identified the reverse transcriptase (RT), integrase (IN), and protease (PR) domains in the pol gene of our high stringency sequences using Pfam domains [29]. The amino acid sequences of these domains, which vary in length from 153 to 210 amino acid, were then taken as queries for a tblastn[21] search. The numbers of identified domains, that do not belong to previously identified roo or rooA elements, are shown in Table 4.

Table 4 Low stringency search and domains

This search identified roo domains in all genomes with previously identified roo elements, and additional domains in D. pseudoobscura, but still did not identify any elements in D. virilis and D. grimshawi. This leads to the conclusion that the roo element is not present in these two genomes, or that it was deleted a long time ago.

The same search identified rooA domains only in genomes belonging to the melanogaster subgroup and in D. mojavensis (Table 4).

For each of the three domains, our search identified more than 100 domains that were not part of previously identified roo or rooA elements. We calculated the amino acid similarity between domains within and between elements from different species. The results mainly validate our whole element comparison (see Additional files 4, 5, and 6 for results from the reverse transcriptase, integrase and protease domain, respectively). An exception is the roo integrase domain. The D. pseudoobscura genome, where we had not identified any element previously, contains a few (three) integrase domains (Table 4). These domains are very similar (92.4% identity) to the domains in the sister species D. persimilis (Additional file 5).

Many transposable elements are found in the heterochromatin part of the genome. These regions are usually poorly assembled, which means that we possibly missed some of the roo and rooA elements. We therefore blasted the amino acid sequence of all roo and rooA elements against the sequence trace archives of all genomes, with the exception of D. pseudoobscura, where no trace archive was available, using a similarity threshold of 70 percent. Table 4 indicates, whether part of a roo (column 6) or rooA (column 11) element was found. The results are largely confirmatory of our previous analysis. No roo or rooA elements occur in D. virilis or D. grimshawi. Roo elements occur in the remaining nine genomes. RooA elements occur in all genomes of the melanogaster subgroup and in D. mojavensis.

No likely origin of roo/rooA outside Drosophila

To detect if the roo or rooA element is also present in species outside the Drosophila group, we searched for the protein sequence of the canonical element in all 162 available, fully sequenced eukaryotic genomes using genomic blast [30]. We found no complete roo/rooA elements, and only short fragments covering less than 65 percent of the coding sequence length of roo. Specifically, we found similar sequences of length between 1000 and 1540 amino acids that overlapped the complete pol region and part of the gag regions. The longest of these sequences occurred in insect genomes from mosquitoes Aedes aegypti, Anopheles gambiae, and Culex quinquefasciatus, the silkworm Bombyx mori, the parasitoid wasp Nasonia vitripennis, and the beetle Tribolium castaneum. Sequences of around the same length were also found in the teleost fishes Danio rerio, Gasterosteus aculeatus, and Oryzias latipes, the sea squirt Ciona savignyi, the lizard Anolis carolinensis, and the sea urchin Strongylocentrotus purpuratus. Shorter sequences that still covered the whole pol region were found in nematode genomes. No similar sequences were identified in mammals. We next asked whether these sequences correspond to roo elements outside the Drosophila clade, or if they derive from other retrotransposons which also share pol and gag regions. To this end, we used our relaxed search procedure (see Methods) on all roo/rooA elements from Drosophila. This relaxed search did not identify any roo or rooA elements in any of the genomes from the previous paragraph, which suggests that the sequences found by blast stem from other retrotransposon families.

Discussion

Element distribution

We identified 164 copies of the roo element in nine out of the 12 sequenced Drosophila genomes, and 66 copies of rooA in four of the genomes. An additional search for roo and rooA fragments using sequence domains (see Methods), and a search against the trace archives of the genomes found additional elements. These additional searches identified roo elements or remnants of roo elements in the genome of D. pseudoobscura, where the high stringency search had not found any elements. This observation, together with our observation of solo-LTRs in this species, suggests that D. pseudoobscura contains remnants of these elements. In sum, ten of the 12 Drosophila species contain roo elements. The additional searches also revealed new rooA elements or fragments in the two genomes of D. sechellia and D. melanogaster that did not contain high stringency elements. RooA elements are thus present in six out of the 12 genomes, i.e., in the five genomes belonging to the melanogaster subgroup, and in the genome of D. mojavensis. Overall, the number of identified elements is probably still an underestimate because not all parts of every genome are sequenced. Especially the high number of genomes which contain only non-functional elements (see Table 1 and Figure 4) is surprising, as it is not likely that these elements were still active recently (as indicated by LTR similarity), and then were lost simultaneously in many genomes. Active elements might still be present in non-sequenced parts of some genomes.

Figure 4
figure 4

Three hypotheses for the evolutionary dynamics of the roo and rooA element in the 12 Drosophila genomes. Red and blue branches indicate species harboring roo and rooA, respectively. Black branches indicate the loss of both elements. a) Roo/rooA may have originated in the common ancestor of all 12 Drosophila genomes with the split between the two elements shortly thereafter. Both elements were then vertically transmitted, as shown by the red and blue branches. Roo was lost in 3 species and rooA in six or seven (depending on the correctness of the rooA element in D. mojavensis (see text for details)). b) Roo/rooA may have arisen in the common ancestor of the Sophophora group and split shortly thereafter into separate lineages. Both elements were vertically transmitted to all descendant genomes, with a loss of roo in one genome and of rooA in four genomes. In this scenario, a horizontal transfer of roo to the genome of D. mojavensis took place after the split from D. willistoni (red arrow), but the exact time point of transmission cannot be determined. c) Roo/rooA might have originated in the common ancestor of all 12 genomes. With the split of the Sophophora group the element evolved into the roo element in the Sophophora group, and into the rooA element in the other three genomes. While roo was vertically transferred to all genomes, rooA is today only present in the D. mojavensis member of this clade. A horizontal transfer of roo to the genome of D. mojavensis took place after the split from D. willistoni. Later, another horizontal transfer, this time of rooA, took place from D. mojavensis to the common ancestor of the melanogaster subgroup where it then spread in all five genomes.

The single rooA element in D. mojavensis remains a mystery. The solo-LTRs and additional integrase domains we find in this species, were all identified by the sequence of the complete element in this species, and not by any of the element sequences in D. yakuba or D. erecta. We did not find any traces of rooA elements in any other genome outside the melanogaster subgroup. It is possible that this single element belongs to a different LTR element family. Below we will discuss the implications of this element on the evolutionary dynamics of the roo and rooA families.

Comparison to previous work

Our analysis can best be assessed by comparison to previous work focusing only on the D. melanogaster genome, and to the annotation of the D. melanogaster genome. Previously, Sánchez-Gracia et al. (2005) reported 54 roo copies based on PCR amplification [17]. This study, however, used only a 1.5 kb fragment of coding DNA to identify a roo element. This is much shorter than our thresholds for the high and low stringency search. Vieira et al. (1999) used in situ hybridization to identify roo insertion sites in D. melanogaster and D. simulans [31]. Here again, the probe DNA had to be short (less then 300 bp), which means that no study to date identified complete and active roo elements in D. simulans. This shortcoming was also pointed out by a recent study by Bartolomé et al. (2009) [32]. Lerat et al. (2003) [20] report 103 roo elements, including 30 complete elements. These numbers are similar to the 107 elements we found in Drosophila melanogaster. In contrast, Kaminker et al. (2002) [11] report 146 roo elements. However, the latter authors identified a sequence as a roo element if it was only 50 bp long, a low stringency that is not appropriate for our analysis. At higher stringency, our analysis may be more comprehensive. We note that both previous searches were based on release 3 of the D. melanogaster genome, whereas our analysis uses release 4. In release 4, additional gaps in the sequence were closed, adding an additional 1.4 Mb to the genome sequence [33]. The flybase annotation of release 4.1 of the D. melanogaster genome still contains 146 roo elements. Thirty-six of them were not identified with our search strategy. All but four are truncated elements, i.e., they are solo-LTRs or fragments with a length shorter than our search threshold. Two of the remaining four long elements are not flanked by recognizable LTR sequences; one element has very short LTR sequences; the last element has no identifiable ORF. This comparison shows that our search strategy, together with a search for solo-LTRs and domains, can identify the vast majority of previously known elements. Only short and/or highly diverged element fragments are not identifiable by our search strategy as seen, for example, for the previously identified rooA fragments in D. melanogaster.

Roo and rooA are in various states of activity

LTR elements in eukaryotic genomes undergo different waves of activation and inactivation [34, 35]. The roo and rooA elements we observed are in various states of activity. In some species, such as D. persimilis, roo may be on its way to extinction, as indicated by the sole element in this species, its truncation, its multiple stop codons, its frameshift mutation, and its high divergence. In other species, roo elements are numerous and active. Species with abundant roo or rooA elements have experienced recent transpositional activity. They include Drosophila melanogaster for the roo element and D. erecta for the rooA element. In these species, the LTR sequences of any one element are generally very similar to one another, indicating recent insertion. Within-genome amino acid similarity is generally high, also indicating recent transposition. For D. melanogaster, these observations are consistent with evidence suggesting that LTR elements have recently become activated [10, 36].

The typically young age of most roo LTR elements within a genome suggests three possible scenarios for roo's evolutionary dynamics. First, roo may have entered any one genome only recently. This possibility apparently contradicts the observation that roo elements are moderately diverged within the Drosophila clade as a whole, suggesting that they have segregated in the clade for some time. This contradiction could be resolved if roo elements that enter a genome cause the host's extinction on short evolutionary time scales, but before they do so, become horizontally transferred to another host. In this case, roo elements would persist in a clade largely through horizontal transfer. However, although horizontal transfer has been suggested by Sánchez-Gracia et al. (2005) between roo elements from D. simulans and D. melanogaster [17], and although our data is suggestive of another horizontal transfer event, it contains no indication that such transfer is rampant. For example, the roo element tree is largely congruent with the species tree.

The second scenario is that roo and rooA elements simply show high turnover within a genome, with elements continually becoming deleted (leaving solo-LTRs behind) and transposing into new locations. D. melanogaster has indeed a high rate of DNA loss [37]. Compared to other invertebrate genomes, the D. melanogaster genome is much smaller and contains less non-coding DNA. A high rate of transposable element deletion may be part of the process maintaining genome compactness. Although most other Drosophila genomes are larger than that of D. melanogaster, they also have a small genome compared to other invertebrates. The many solo-LTRs we find confirm that the Drosophila genomes we study may have a high rate of mobile DNA loss.

A third scenario is that non-autonomous elements are under strong selection to maintain their LTR sequences, because LTR sequences are important for transposition and DNA interaction. Having identical or highly similar LTRs might enable non-autonomous elements to use the transcription mechanism of autonomous copies. This may explain the high coding sequence diversity despite very similar LTR sequences observed for some elements. Examples include the five elements in D. ananassae. If this possibility is correct, than many seemingly young elements might not be that young in reality.

In balance, the lack of rampant horizontal transfer, and the frequency with which roo elements become excised, support the second and third scenario, which are not mutually exclusive.

Roo and rooA split a long time ago

We next discuss the relationship of roo and rooA, whose element trees are clearly separable (Figure 3). The split between roo and rooA probably took place before the branch leading to D. willistoni or even before the branch leading to D. mojavensis originated, because otherwise the elements in these genomes would belong to the ancestral roo/rooA family and they would be joined to the outgroup in the element tree. The rooA family probably consisted of only one or very few copies before the melanogaster subgroup arose, because we did only find one putative occurrence outside this subgroup. The rooA copies were probably activated and experienced a transposition activity after the split from D. ananassae and then spread through all species, as indicated by the blue lines in Figure 4. One putative rooA element, however, occurs in the genome of D. mojavensis. This suggests that the split between roo and rooA indeed occurred in the common ancestor of all 12 species.

Incongruency between species and roo element tree

We now turn to an incongruency in the subtree for the roo elements (Figure 3, red part) that involves the D. mojavensis genome. D. mojavensis is one of the most distantly related species to D. melanogaster, but the roo elements in D. mojavensis are more closely related to D. melanogaster than the elements in D. melanogaster's closer relative D. willistoni. The element tree topology (including the branch in question) is well supported, making a tree estimation error unlikely. Incongruencies like these raise the possibility of horizontal gene transfer events, but before one can conclude that such an event took place, a number of potentially confounding factors need to be considered. First, the genomes of D. willistoni and D. saltans, a sister group of D. willistoni with no sequenced genome, are evolving faster than the other genomes of the Drosophila genus [38]. This could have influenced the evolution of the roo element and explain the position of the D. willistoni elements in the element tree. Second, an error in the multiple alignment could lead to this incongruency. Multiple alignments based on many, very long sequences are error prone, especially if the sequences differ in their length and are not all closely related. To exclude this possibility, we examined the average amino acid similarity between elements based on pairwise alignments. This measure of divergence indicates a higher similarity between roo elements from D. mojavensis and Drosophila melanogaster than between elements from D. willistoni and Drosophila melanogaster. While the difference in similarity is small (61.3 percent versus 59 percent) the result still supports the element tree's structure. A third confounding factor is the species tree itself. The current placement of D. willistoni is close to the split of all 12 Drosophila genomes; this placement, however, is ambiguous. Many gene trees do not match the current phylogeny [9] and place D. willistoni as an outgroup to the other genomes. If so, the species tree would be congruent with the roo element tree. In sum, a horizontal gene transfer event involving D. mojavensis is possible but does have mixed support in the data.

High similarity between roo/rooA elements in melanogaster subgroup

Sánchez-Gracia et al. (2005) suggested a horizontal transfer event of roo between D. melanogaster and D. simulans based on synonymous and non-synonymous divergence data. We also find a high similarity in roo between these two species. However, we can not unambiguously explain this low divergence by a horizontal transfer event, because the similarity for elements within one genome is higher than the similarity between elements from different genomes. Also, we find high similarity not only between elements in these two genomes but also between all roo and rooA elements in species of the melanogaster subgroup. This observation might lead to the speculation that the roo and the rooA element were repeatedly "jumping" from one species to another. However, Sánchez-Gracia et al. showed that other LTR, non-LTR and DNA transposons also show low divergence, raising the possibility that these patterns are consistent with vertical transfer.

Three possibilities for the evolutionary dynamics of roo and rooA

Taken together, all our observations suggest three alternative hypotheses for the evolutionary dynamics of the roo and rooA elements among the 12 Drosophila genomes. These hypotheses are based on the assumption that the roo and rooA subfamilies diverged shortly after their common ancestor originated, because they are highly diverged. The first two hypotheses differ in assigning a time for the first occurrence of a roo/rooA element in the Drosophila species. The first hypothesis (Figure 4a) assumes that the roo/rooA element was already present in the common ancestor of all 12 species, and that both the roo and the rooA element were vertically transferred to all species. This would explain the absence of roo elements in some species by stochastic loss (black branches in Figure 4). Wherever element numbers are low in a genome, such stochastic loss is likely. The closest relatives of genomes with no roo elements have small to modest element numbers, which supports the loss hypothesis. For example, the closest relative of D. pseudoobscura, which has no roo elements, is D. persimilis, which has only one roo element. According to this hypothesis roo was lost from the genomes of D. virilis and D. grimshawi, but is still present in D. mojavensis. The split between roo and rooA also took place in the common ancestor of all 12 genomes, and rooA was transmitted vertically to all genomes. It might have been lost from the common ancestor of the three species outside the Sophophora group soon thereafter. Alternatively, if the single putative rooA element in D. mojavensis is a true rooA element, it might have been lost in each non-Sophophora genome separately (as indicated by the dashed blue lines in Figure 4a).

The second hypothesis addresses the key shortcoming of the first hypothesis, which cannot explain the incongruence between the species (Figure 1b) and the element tree (Figure 3). This second hypothesis places the first occurrence of the roo/rooA element in the ancestor of the Sophophora group and the split between the two elements shortly thereafter, as shown in Figure 4b by the red and blue branches. The elements were then vertically transmitted to all descendant species. While roo elements are still present in all of the nine genomes, rooA elements are only found in species belonging to the melanogaster subgroup and were lost from the other genomes. This hypothesis also assumes that the single rooA element in D. mojavensis belongs to a different LTR family and was wrongly identified. The roo element in this genome, however, is explained by a horizontal transfer of a roo element after the split of D. willistoni to the genome of D. mojavensis. This hypothesis can also explain the element tree in Figure 3 (if we assume that the species tree in Figure 1 is correct), the lack of solo LTRs and domains in D. virilis and D. grimshawi, and the amino acid similarity between the coding regions of the elements. This transfer would have been an ancient transfer, because the elements from D. mojavensis share little similarity with elements from other species. An origin of roo within the Drosophila clade is further supported by the fact that no roo elements can be found outside Drosophila. Some of the RETRO elements in A. gambiae show a similarity to the roo element [39] but they may not belong to the same family. Furthermore, only LTR sequences of RETRO elements have been identified so far, which makes a comparison to the roo/rooA elements difficult. Previous phylogenetic studies identified the Tinker and BEL elements in D. melanogaster as closest relative of the roo element [12, 40], but when the split between these elements may have occurred is unclear.

This second hypothesis however, does not explain the single rooA element in D. mojavensis, if we assume it was not falsely identified. Our third hypothesis (Figure 4c) addresses this shortcoming. As in the first hypothesis, we assume that the roo/rooA element arose in the common ancestor of all 12 genomes. In contrast to the first hypothesis, the third hypothesis posits that the split between roo and rooA corresponds to the split between the genomes of the Sophophora group and the other three genomes. In the Sophophora group the ancestral roo/rooA element evolved into today's roo elements, as indicated in Figure 4c (red lines). In the three genomes of D. mojavensis, D. virilis and D. grimshawi the ancestral element evolved into the rooA element. But in contrast to the roo elements in the Sophophora group, the rooA element did not replicate often and was lost early from D. virilis and D. grimshawi. Only in D. mojavensis did the element start to spread. At one point a roo element was horizontally transferred into D. mojavensis, and later a second horizontal transfer took place, this time of rooA from D. mojavensis to the common ancestor of the genomes of the melanogaster subgroup. In this group rooA transposed often, especially in the genomes of D. yakuba and D. erecta. As no functional copy is left in D. mojavensis, this horizontal transfer prevented the extinction of the rooA element.

In sum, each of these three hypotheses explains some aspect of the available data, but this data is currently not sufficient to decide equivocally for or against one of them.

Conclusion

Within genomes, the evolutionary dynamics of roo and rooA shows a broad spectrum ranging from recent intense transpositional activity to slow decay and extinction. Among genomes, the balance of evidence suggests an origin of roo/rooA in the common ancestor of all 12 genomes or within the Drosophila clade, and the separation between the two families shortly afterwards. Furthermore, one horizontal transfer of roo and one of rooA may have occurred, but the evidence for such transfers is equivocal. Our analysis is based on a limited number of genomes, and conclusive proof of the evolutionary dynamics of roo and rooA, as well as dating of their origin would need sequence information from additional species. For example, horizontal transfer of roo and rooA could be confirmed if future genome sequencing showed that all roo elements occur in species closely related to the Sophophora group and to D. mojavensis, whereas rooA elements were found to occur only in genomes of the melanogaster subgroup and outside the Sophophora group.

Methods

Element Identification

To identify roo elements in the 12 focal Drosophila genomes, we carried out a high stringency and a low stringency search. Both of these searches use an iterative procedure that has as its core the tool IScan[41], which was originally designed to identify insertion sequences, simple transposable elements in bacteria. To identify a transposable element's coding region, IScan performs a tblastn search (using WUBLAST[42]) to identify matches of the amino acid query sequence in a genome sequence of interest. After having identified the coding region, IScan also allows the identification of indirect repeat regions characteristic of some transposable elements, and associated with the coding regions. For our purpose, we modified IScan to allow us to identify the direct repeats associated with roo elements.

In our high stringency search, we used the roo canonical sequence [GenBank:AY180917] and the rooA canonical sequence [18] as a first query, and searched for matches in the 12 Drosophila genomes which we obtained from the Comparative Assembly Freeze 1 (CAF1) [43]. In the high stringency search, we considered further only hits with an evalue ≤ 0.0001 and at least 50 percent identical residues between the hit and query sequence. We only considered hits with an LTR sequence length between 100 bp and 1000 bp, and allowed no unidentified ('N') nucleotides in the coding and LTR sequences.

After having identified a series of roo sequences that fulfill these criteria, we used these sequences as new queries in a new search round, and reiterated this procedure until no new roo sequences were found. Between each iteration, we verified that no previously identified sequence was used more than once as a query. We then manually inspected all roo elements for spurious hits.

For the low stringency search, we used those 118 elements previously identified in the high stringency search where the predicted protein length was longer than 1180 aa, that is, at least half of the canonical protein length. For this low stringency search, we used an evalue ≤ 0.01, required at least a 40 percent match between the query sequences and the hit, and permitted sequences with unidentified nucleotides ('N') in their coding and LTR sequences. This search yielded additional roo and rooA fragments in only one genome (D. simulans) where the previous search had failed. However, these elements are of spurious validity, as they contain many sequence fragments annotated with unknown nucleotides ('N'). In all other genomes except D. persimilis where the high stringency search had identified one roo element, the low stringency search yielded additional roo and rooA copies. Because both canonical sequences identify roo and rooA elements we calculated the similarity of each identified element to both canonical elements and added the element to the family with the higher similarity.

Identification of complete coding region

Our search strategy described above may identify candidate coding regions only incompletely. Specifically, its results may not include the upstream-most and downstream-most parts of coding regions, such as start codons and stop codons. To help us to get complete coding region information, we applied the getorf tool from the EMBOSS package [44] to the complete element sequence, from the first nucleotide of the 5' LTR sequence through the last nucleotide of the 3' LTR.

Specifically, this processing step used the coding region C identified in the initial analysis, and used getorf to identify all open reading frames (ORFs) in the same reading frame as C and overlapping C. For most elements (176 elements), this analysis found only one ORF, which covered C completely, which we then used as the final coding sequence. However, in 54 roo and rooA elements we found two (or more) ORFs overlapping C, which may be caused by nonsense mutations that insert stop codons into a gene. For these elements, we designated the final coding sequence as extending from the first (upstream-most) to the last (downstream-most) ORF overlapping C. The potential non-coding region between the ORFs, if any, were also included into this final region. With their median length of 270 bp they comprise a small average fraction of 5.6 percent of the coding region, and are thus not likely to significantly alter our results. In fewer than 5 cases additional ORFs in the reading frame but not overlapping C showed a high similarity to the canonical roo sequence, in which case we included these ORFs into the coding region as well.

Sequence alignment and phylogenetic analysis

The alignments we used are based on the protein coding sequences of all roo and rooA elements. Our dataset contained 230 roo and rooA elements with coding sequences as long as 2615 amino acids, which renders accurate multiple alignment identification challenging. We compared different alignment algorithms (ClustalW[45], Mafft[46], Muscle[47], Prank[48] and ProbCons[49], T-Coffee[50]) to find the best algorithm for our dataset. Because of its high speed and low number of artefacts evident upon visual inspection, we chose Mafft version 5, which refines alignments iteratively using pairwise alignment information.

Based on the multiple alignment, we computed phylogenetic trees of elements using PhyML_aLRT[27] a version of PhyML[28] which incorporates an approximate likelihood ratio test to estimate the statistical support of the tree topology. This approach is superior to a bootstrap calculation with respect to accuracy and power, and it is computationally much more efficient [27]. The method assigns to each branch a statistical significance ranging from 0 (least significant) to 1 (highly significant). We used the default options of PhyML_aLRT with the JTT matrix for amino acid substitutions, the proportion of invariable sites set to zero, and with only one category of substitution rate [28]. We chose the χ2-based parametric branch support for approximate likelihood ratio tests [27]. As outgroup we used the canonical sequence of the BEL element in D. melanogaster to root the tree.

Age calculation

We estimate the intra-element pairwise nucleotide identity between LTR sequences using the dnadist program from the PHYLIP package [51] with the Felsenstein F84 parameter method. We used the formula T = K/2r where T = time to most recent common ancestry, K = sequence divergence, and r = substitution rate [23] to calculate the age of an element, as described by Bowen and McDonald (2001) [10]. As a substitution rate we used 0.016 substitutions per site per million years, as estimated for Drosophila [23].

Calculation of amino acid similarity

The calculation is based on the coding region of the element. To avoid biases associated with truncated copies, which are usually also highly diverged, we used only the 134 roo and rooA elements with an encoded protein sequence length longer than 1180 amino acids, which is half the length of the canonical element's coding region. Using the protdist program from the PHYLIP package [51] we calculated the percent similarity between pairwise Mafft alignments of element pairs.

Domain identification

The pol gene of all autonomous LTR retrotransposons contains a reverse transcriptase (RT), an integrase (IN), and a protease (PR) domain. We obtained the following hidden markov models [52] from Pfam [53]: "RVT_1" for reverse transcriptase, "rve" for integrase, and "Peptidase_A17" for the protease domain. Using Hmmer [54] to search for these domains in our high stringency hits, we then used the identified domains in a tblastn search against all twelve Drosophila genomes. Only hits which share at least 99 percent of the length to their query sequence and a sequence identity of 90 percent were considered for further analysis. The amino acid similarity between domains was estimated as described above.

Search in trace archives

We used the protein sequences of all roo and rooA elements in a tblastn search against the traces archives of all genomes, except that of D. pseudoobscura, where no trace archive was available. A hit required at least 70 percent similarity to a roo or rooA sequence, and it had to overlap at least half of the trace archive sequence.

References

  1. Craig NL, Craigie R, Gellert M, Lambowitz AM, Eds: Mobile DNA II. 2002, Washington: ASM Press

    Google Scholar 

  2. Charlesworth B, Sniegowski P, Stepha W: The evolutionary dynamics of repetitive DNA in eukaryotes. Nature. 1994, 371 (6494): 215-220. 10.1038/371215a0.

    Article  CAS  PubMed  Google Scholar 

  3. Miki Y: Retrotransposal integration of mobile genetic elements in human diseases. J Hum Genet. 1998, 43 (2): 77-84. 10.1007/s100380050045.

    Article  CAS  PubMed  Google Scholar 

  4. Montgomery EA, Huang SM, Langley CH, Judd BH: Chromosome rearrangement by ectopic recombination in Drosophila melanogaster: genome structure and evolution. Genetics. 1991, 129 (4): 1085-1098.

    PubMed Central  CAS  PubMed  Google Scholar 

  5. Yoder JA, Walsh CP, Bestor TH: Cytosine methylation and the ecology of intragenomic parasites. Trends Genet. 1997, 13 (8): 335-340. 10.1016/S0168-9525(97)01181-5.

    Article  CAS  PubMed  Google Scholar 

  6. Galagan JE, Selker EU: RIP: the evolutionary cost of genome defense. Trends Genet. 2004, 20 (9): 417-423. 10.1016/j.tig.2004.07.007.

    Article  CAS  PubMed  Google Scholar 

  7. Jensen S, Gassama MP, Heidmann T: Taming of transposable elements by homology-dependent gene silencing. Nat Genet. 1999, 21 (2): 209-212. 10.1038/5997.

    Article  CAS  PubMed  Google Scholar 

  8. Vastenhouw NL, Plasterk RHA: RNAi protects the Caenorhabditis elegans germline against transposition. Trends Genet. 2004, 20 (7): 314-319. 10.1016/j.tig.2004.04.011.

    Article  CAS  PubMed  Google Scholar 

  9. Consortium DG: Evolution of genes and genomes on the Drosophila phylogeny. Nature. 2007, 450 (7167): 203-218. 10.1038/nature06341.

    Article  Google Scholar 

  10. Bowen NJ, McDonald JF: Drosophila euchromatic LTR retrotransposons are much younger than the host species in which they reside. Genome Res. 2001, 11 (9): 1527-1540. 10.1101/gr.164201.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  11. Kaminker JS, Bergman CM, Kronmiller B, Carlson J, Svirskas R, Patel S, Frise E, Wheeler DA, Lewis SE, Rubin GM, Ashburner M, Celniker SE: The transposable elements of the Drosophila melanogaster euchromatin: a genomics perspective. Genome Biol. 2002, 3 (12): RESEARCH0084-10.1186/gb-2002-3-12-research0084.

    Article  PubMed Central  PubMed  Google Scholar 

  12. Frame IG, Cutfield JF, Poulter RT: New BEL-like LTR-retrotransposons in Fugu rubripes, Caenorhabditis elegans, and Drosophila melanogaster. Gene. 2001, 263 (1–2): 219-230. 10.1016/S0378-1119(00)00567-9.

    Article  CAS  PubMed  Google Scholar 

  13. Scherer G, Tschudi C, Perera J, Delius H, Pirrotta V: B104, a new dispersed repeated gene family in Drosophila melanogaster and its analogies with retroviruses. J Mol Biol. 1982, 157 (3): 435-451. 10.1016/0022-2836(82)90470-3.

    Article  CAS  PubMed  Google Scholar 

  14. Meyerowitz EM, Hogness DS: Molecular organization of a Drosophila puff site that responds to ecdysone. Cell. 1982, 28: 165-176. 10.1016/0092-8674(82)90386-5.

    Article  CAS  PubMed  Google Scholar 

  15. Brunner E, Brunner D, Fu W, Hafen E, Basler K: The dominant mutation Glazed is a gain-of-function allele of wingless that, similar to loss of APC, interferes with normal eye development. Dev Biol. 1999, 206 (2): 178-188. 10.1006/dbio.1998.9136.

    Article  CAS  PubMed  Google Scholar 

  16. Biémont C, Cizeron G: Distribution of transposable elements in Drosophila species. Genetica. 1999, 105: 43-62. 10.1023/A:1003718520490.

    Article  PubMed  Google Scholar 

  17. Sánchez-Gracia A, Maside X, Charlesworth B: High rate of horizontal transfer of transposable elements in Drosophila. Trends Genet. 2005, 21 (4): 200-203. 10.1016/j.tig.2005.02.001.

    Article  PubMed  Google Scholar 

  18. Kapitonov VV, Jurka J: RooA, an ancient family of endogenous retrovirus ancestral to the modern ROO family. Repbase Reports. 2002, 2 (3): 8-

    Google Scholar 

  19. Jordan IK, McDonald JF: Evidence for the role of recombination in the regulatory evolution of Saccharomyces cerevisiae Ty elements. J Mol Evol. 1998, 47: 14-20. 10.1007/PL00006358.

    Article  CAS  PubMed  Google Scholar 

  20. Lerat E, Rizzon C, Biémont C: Sequence divergence within transposable element families in the Drosophila melanogaster genome. Genome Res. 2003, 13 (8): 1889-1896.

    PubMed Central  CAS  PubMed  Google Scholar 

  21. BLAST. [http://blast.wustl.edu]

  22. SanMiguel P, Gaut BS, Tikhonov A, Nakajima Y, Bennetzen JL: The paleontology of intergene retrotransposons of maize. Nat Genet. 1998, 20: 43-45. 10.1038/1695.

    Article  CAS  PubMed  Google Scholar 

  23. Li WH: Molecular Evolution. 1997, Sunderland: Sinauer Associates

    Google Scholar 

  24. Salenko V, Kotnova A, Karpova N, Lyubomirskaya N, Ilyin Y: Polymorphism of canonical and noncanonical gypsy sequences in different species of Drosophila melanogaster subgroup: possible evolutionary relations. Mol Genet Genomics. 2008, 279: 463-472. 10.1007/s00438-008-0325-6.

    Article  CAS  PubMed  Google Scholar 

  25. Roeder GS, Fink GR: Movement of yeast transposable elements by gene conversion. Proc Natl Acad Sci USA. 1982, 79 (18): 5621-5625. 10.1073/pnas.79.18.5621.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  26. Yang Z: PAML 4: phylogenetic analysis by maximum likelihood. Mol Biol Evol. 2007, 24 (8): 1586-1591. 10.1093/molbev/msm088.

    Article  CAS  PubMed  Google Scholar 

  27. Anisimova M, Gascuel O: Approximate likelihood-ratio test for branches: A fast, accurate, and powerful alternative. Syst Biol. 2006, 55 (4): 539-552. 10.1080/10635150600755453. [http://sysbio.oxfordjournals.org/cgi/content/abstract/55/4/539]

    Article  PubMed  Google Scholar 

  28. Guindon S, Gascuel O: A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst Biol. 2003, 52 (5): 696-704. 10.1080/10635150390235520.

    Article  PubMed  Google Scholar 

  29. Finn RD, Tate J, Mistry J, Coggill PC, Sammut SJ, Hotz HR, Ceric G, Forslund K, Eddy SR, Sonnhammer ELL, Bateman A: The Pfam protein families database. Nucleic Acids Res. 2008, D281-D288. 36 Database

  30. Cummings L, Riley L, Black L, Souvorov A, Resenchuk S, Dondoshansky I, Tatusova T: Genomic BLAST: custom-defined virtual databases for complete and unfinished genomes. FEMS Microbiol Lett. 2002, 216 (2): 133-138. 10.1111/j.1574-6968.2002.tb11426.x.

    Article  CAS  PubMed  Google Scholar 

  31. Vieira C, Lepetit D, Dumont S, Biémont C: Wake up of transposable elements following Drosophila simulans worldwide colonization. Mol Biol Evol. 1999, 16 (9): 1251-1255.

    Article  CAS  PubMed  Google Scholar 

  32. Bartolomé C, Bello X, Maside X: Widespread evidence for horizontal transfer of transposable elements across Drosophila genomes. Genome Biol. 2009, 10 (2): R22-10.1186/gb-2009-10-2-r22.

    Article  PubMed Central  PubMed  Google Scholar 

  33. BDGP: Drosophila melanogaster release 4. [http://www.fruitfly.org/annot/release4.html]

  34. Bennetzen JL: Mechanisms and rates of genome expansion and contraction in flowering plants. Genetica. 2002, 115: 29-36. 10.1023/A:1016015913350.

    Article  CAS  PubMed  Google Scholar 

  35. Volff JN: Genome evolution and biodiversity in teleost fish. Heredity. 2005, 94 (3): 280-294. 10.1038/sj.hdy.6800635.

    Article  CAS  PubMed  Google Scholar 

  36. Mugnier N, Gueguen L, Vieira C, Biémont C: The heterochromatic copies of the LTR retrotransposons as a record of the genomic events that have shaped the Drosophila melanogaster genome. Gene. 2008, 411 (1–2): 87-93. 10.1016/j.gene.2008.01.010.

    Article  CAS  PubMed  Google Scholar 

  37. Petrov DA: DNA loss and evolution of genome size in Drosophila. Genetica. 2002, 115: 81-91. 10.1023/A:1016076215168.

    Article  CAS  PubMed  Google Scholar 

  38. Rodríguez-Trelles F, Tarrío R, Ayala FJ: Fluctuating mutation bias and the evolution of base composition in Drosophila. J Mol Evol. 2000, 50: 1-10. 10.1007/s002399910001.

    Article  PubMed  Google Scholar 

  39. Jurka J, Drazkiewicz A: Anopheles gambiae long terminal repeat from RETRO931_AG retrotransposon – a consensus. Repbase Reports. 2002, 2 (12): 18-

    Google Scholar 

  40. Marsano RM, Caizzi R: A genome-wide screening of BEL-Pao like retrotransposons in Anopheles gambiae by the LTR_STRUC program. Gene. 2005, 357 (2): 115-121. 10.1016/j.gene.2005.06.005.

    Article  PubMed  Google Scholar 

  41. Wagner A, Lewis C, Bichsel M: A survey of bacterial insertion sequences using IScan. Nucleic Acids Res. 2007, 35 (16): 5284-5293. 10.1093/nar/gkm597.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  42. Altschul SF, Madden TL, Schäffer AA, Zhang J, Zhang Z, Miller W, Lipman DJ: Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997, 25 (17): 3389-3402. 10.1093/nar/25.17.3389.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  43. Assembly/Alignment/Annotation of 12 related Drosophila species. [http://rana.lbl.gov/drosophila/index.html]

  44. Rice P, Longden I, Bleasby A: EMBOSS: the European Molecular Biology Open Software Suite. Trends Genet. 2000, 16 (6): 276-277. 10.1016/S0168-9525(00)02024-2.

    Article  CAS  PubMed  Google Scholar 

  45. Thompson JD, Higgins DG, Gibson TJ: CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Research. 1994, 22 (22): 4673-4680. 10.1093/nar/22.22.4673. [http://nar.oxfordjournals.org/cgi/content/abstract/22/22/4673]

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  46. Katoh K, Kuma K, Toh H, Miyata T: MAFFT version 5: improvement in accuracy of multiple sequence alignment. Nucleic Acids Res. 2005, 33 (2): 511-518. 10.1093/nar/gki198.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  47. Edgar RC: MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res. 2004, 32 (5): 1792-1797. 10.1093/nar/gkh340.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  48. Löytynoja A, Goldman N: Phylogeny-aware gap placement prevents errors in sequence alignment and evolutionary analysis. Science. 2008, 320 (5883): 1632-1635. 10.1126/science.1158395.

    Article  PubMed  Google Scholar 

  49. Do CB, Mahabhashyam MSP, Brudno M, Batzoglou S: ProbCons: Probabilistic consistency-based multiple sequence alignment. Genome Res. 2005, 15 (2): 330-340. 10.1101/gr.2821705.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  50. Notredame C, Higgins DG, Heringa J: T-Coffee: A novel method for fast and accurate multiple sequence alignment. J Mol Biol. 2000, 302: 205-217. 10.1006/jmbi.2000.4042.

    Article  CAS  PubMed  Google Scholar 

  51. PHYLIP package. [http://evolution.genetics.washington.edu/phylip.html]

  52. Krogh A, Brown M, Mian IS, Sjölander K, Haussler D: Hidden Markov models in computational biology. Applications to protein modeling. J Mol Biol. 1994, 235 (5): 1501-1531. 10.1006/jmbi.1994.1104.

    Article  CAS  PubMed  Google Scholar 

  53. Finn RD, Mistry J, Schuster-Böckler B, Griffiths-Jones S, Hollich V, Lassmann T, Moxon S, Marshall M, Khanna A, Durbin R, Eddy SR, Sonnhammer ELL, Bateman A: Pfam: clans, web tools and services. Nucleic Acids Res. 2006, D247-D251. 10.1093/nar/gkj149. 34 Database

  54. Hmmer. [http://hmmer.janelia.org]

Download references

Acknowledgements

We would like to acknowledge support from Swiss National Science Foundation grants 315200-116814 and 315200-119697.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nicole de la Chaux.

Additional information

Authors' contributions

NC carried out the research. NC and AW designed the study and wrote the manuscript. All authors read and approved the final manuscript.

Electronic supplementary material

12862_2009_1119_MOESM1_ESM.xls

Additional file 1: List of all full length roo and rooA elements. A list of all 230 full length roo and rooA elements including species name, accession number, start and end position of the complete element, start and end positions for both LTR sequences and the open reading frame (ORF), the elements age in million years, and the coding strand. The last column indicates whether the element is a roo element (0) or a rooA element (1). (XLS 50 KB)

12862_2009_1119_MOESM2_ESM.pdf

Additional file 2: LTR sequence distance distribution. The histograms show the estimated distance between the LTR sequences of roo (red bars) and rooA (blue bars). Notice the different scales on the y-axis. The vertical red and blue lines indicate the median distance in each genome for roo and rooA, respectively. Two roo elements of D. mojavensis had to be excluded because their LTR sequences were falsely identified. The histogram in the bottom right corner shows the distance of all roo and rooA elements. myr: million years. (PDF 331 KB)

12862_2009_1119_MOESM3_ESM.pdf

Additional file 3: Complete phylogenetic tree of roo and rooA elements. The complete phylogenetic tree of roo and rooA elements is shown. Each element is named by the first four letters of their genome and a consecutive number, e. g., Dmel3 is the third element from D. melanogaster. RooA elements have an additional "A" at the end. The tree is rooted by the BEL element, a LTR element from D. melanogaster. Dsim: D. simulans; Dsec: D. sechellia; Dmel: D. melanogaster; Dyak: D. yakuba; Dere: D. erecta; Dana: D. ananassae; Dper: D. persimilis; Dwil: D. willistoni; Dmoj: D. mojavensis. (PDF 20 KB)

12862_2009_1119_MOESM4_ESM.pdf

Additional file 4: Average amino acid similarity between reverse transcriptase domains. All reverse transcriptase domains were used to estimate the amino acid similarity for every pair of these elements using protdist from the PHYLIP package [51]. Dsim: D. simulans; Dsec: D. sechellia; Dmel: D. melanogaster; Dyak: D. yakuba; Dere: D. erecta; Dwil: D. willistoni; Dmoj: D. mojavensis. (PDF 34 KB)

12862_2009_1119_MOESM5_ESM.pdf

Additional file 5: Average amino acid similarity between integrase domains. All integrase domains were used to estimate the amino acid similarity for every pair of these elements using protdist from the PHYLIP package [51]. Dsim: D. simulans; Dsec: D. sechellia; Dmel: D. melanogaster; Dyak: D. yakuba; Dere: D. erecta; Dana: D. ananassae; Dpse: D. pseudoobscura; Dper: D. persimilis; Dwil: D. willistoni; Dmoj: D. mojavensis. (PDF 36 KB)

12862_2009_1119_MOESM6_ESM.pdf

Additional file 6: Average amino acid similarity between the protease domains. All protease domains were used to estimate the amino acid similarity for every pair of these elements using protdist from the PHYLIP package [51]. Dsim: D. simulans; Dsec: D. sechellia; Dmel: D. melanogaster; Dyak: D. yakuba; Dere: D. erecta; Dana: D. ananassae; Dwil: D. willistoni; Dmoj: D. mojavensis. (PDF 34 KB)

Authors’ original submitted files for images

Rights and permissions

This article is published under license to BioMed Central Ltd. This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/2.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Reprints and permissions

About this article

Cite this article

de la Chaux, N., Wagner, A. Evolutionary dynamics of the LTR retrotransposons roo and rooA inferred from twelve complete Drosophila genomes. BMC Evol Biol 9, 205 (2009). https://doi.org/10.1186/1471-2148-9-205

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2148-9-205

Keywords