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

Large introns in relation to alternative splicing and gene evolution: a case study of Drosophila bruno-3

Abstract

Background

Alternative splicing (AS) of maturing mRNA can generate structurally and functionally distinct transcripts from the same gene. Recent bioinformatic analyses of available genome databases inferred a positive correlation between intron length and AS. To study the interplay between intron length and AS empirically and in more detail, we analyzed the diversity of alternatively spliced transcripts (ASTs) in the Drosophila RNA-binding Bruno-3 (Bru-3) gene. This gene was known to encode thirteen exons separated by introns of diverse sizes, ranging from 71 to 41,973 nucleotides in D. melanogaster. Although Bru-3' s structure is expected to be conducive to AS, only two ASTs of this gene were previously described.

Results

Cloning of RT-PCR products of the entire ORF from four species representing three diverged Drosophila lineages provided an evolutionary perspective, high sensitivity, and long-range contiguity of splice choices currently unattainable by high-throughput methods. Consequently, we identified three new exons, a new exon fragment and thirty-three previously unknown ASTs of Bru-3. All exon-skipping events in the gene were mapped to the exons surrounded by introns of at least 800 nucleotides, whereas exons split by introns of less than 250 nucleotides were always spliced contiguously in mRNA. Cases of exon loss and creation during Bru-3 evolution in Drosophila were also localized within large introns. Notably, we identified a true de novo exon gain: exon 8 was created along the lineage of the obscura group from intronic sequence between cryptic splice sites conserved among all Drosophila species surveyed. Exon 8 was included in mature mRNA by the species representing all the major branches of the obscura group. To our knowledge, the origin of exon 8 is the first documented case of exonization of intronic sequence outside vertebrates.

Conclusion

We found that large introns can promote AS via exon-skipping and exon turnover during evolution likely due to frequent errors in their removal from maturing mRNA. Large introns could be a reservoir of genetic diversity, because they have a greater number of mutable sites than short introns. Taken together, gene structure can constrain and/or promote gene evolution.

Background

Alternative splicing (AS) and other post-transcriptional modifications of the same pre-mRNA generate structurally and functionally distinct transcripts from the same gene and thus could account for the origin of high protein diversity from the limited number of coding genes (e.g., [1–4]). The genome of Drosophila melanogaster [5] contains 15,182 annotated genes (Release 5.2 [6]) that encode 43,129 distinct transcripts (UniGene build #64 [7]). Therefore, on average, one Drosophila gene encodes 2.8 known or predicted transcripts. The ratio of genes to transcripts is even higher in humans, 1:9.5; 20,500 genomic genes [8] and 195,727 distinct transcripts (UniGene build #219 [7]).

The complex molecular structure of genes might be especially conducive to AS. In particular, exon number and intron length correlate with AS [e.g., [9–12]]. Genes with greater numbers of exons can generate more alternatively spliced transcripts (ASTs) just by possible combinations of a gene's exons [e.g., [10]]. One of the most extreme examples of Drosophila multi-exonic genes is Dscam. This gene potentially encodes 38,016 distinct axon guidance receptors via differential inclusion of its 95 exons [13]. Multiple recent bioinformatics analyses of genome databases inferred that exons surrounded by long introns (i.e., >250 bp) are skipped more often in gene transcripts than exons flanked by short introns [11, 12, 14, 15]. In addition, recursive splicing of extremely large introns [16, 17] may be prone to errors, which may lead to both exon-skipping and intron retention in transcripts. Therefore, the presence of many exons separated by long introns in a gene can promote, or at least associate with, the AS of a gene.

We studied the potential interplay between the intron length and AS using the Drosophila Bruno-3 (Bru-3). This gene has a genetic structure unusual for Drosophila. While the average Drosophila genes tend to have 4 exons and 487 bp-long introns [18], Bru-3 is encoded by 13 exons that spread along 129 kb of genomic sequence and its longest known mature mRNA is only 2.6 kb long. The lengths of the twelve introns in this gene are extremely diverse and range from 71 bp up to 41,973 bp (FlyBase.org, D. melanogaster release 5.13). Therefore, the inclusion frequency of exons surrounded by differently sized introns can be examined within this single gene. Only two alternatively spliced transcripts (ASTs) of Bru-3 - Bru-3-RA (2606 bp) and Bru-3-RB (2429 bp) - were identified and confirmed in D. melanogaster (UniGene Dm.13624). One additional alternative transcript can be inferred from 17 ESTs available at UniGene Database [7] that match the Bru-3 genomic sequence. However, this gene might have a greater diversity of ASTs than previously documented. By focusing our analysis to a single interesting gene, we sought to more thoroughly sample AST diversity directly and study the interplay between intron length and AS.

Drosophila Bru-3 belongs to a complex and diverse group of RNA binding proteins [e.g., [19]]. It encodes two RNA recognition motifs (RRMs), one at the N-terminus and one at the C-terminus of the gene, and a linker region between the RRMs [20, 21]. Bru-3 was classified as the Drosophila orthologue of human and mouse CUG-BP (also called as TNRC4) and Xenopus EDEN-BP [i.e., Bluno-like proteins, [20, 22]]. While it was shown that Bru-3 represses the translation of target genes by binding to the 3' UTR of their mRNA in Drosophila, [e.g., [20]], it may also regulate pre-mRNA AS and be involved in mRNA editing, as done by other members of this protein family [e.g., [19, 23, 24]]. The importance of these regulatory functions of Bru-3 is likely reflected in high sequence conservation at both coding and non-coding regions of this gene [but see also, [25]]. Bru-3 has the highest number of ultra-conserved elements per gene between D. melanogaster and D. pseudoobscura, separated 55 million years (MY) ago [26]. These elements comprise 11.2 kb [27], while the longest mature mRNA of D. melanogaster Bru-3 is only 2.6 kb (FlyBase.org, release 5.13).

Common computational methods rely on expression sequence tags [EST, [28]] to infer ASTs. As a result, inherent biases of EST recovery [for more detail refer to: [3, 14, 29, 30]] lead to underestimation of the true number of alternatively spliced genes and especially the diversity of alternative transcripts of individual genes [e.g., [31, 32]]. Direct bioinformatic identification of alternative splicing events, including cassette exons, alternative 5' and 3' splice sites, is still not totally accurate and comprehensive [31, 32]. Application of microarrays for probing the diversity of alternative transcripts requires knowing the locations of exon-exon and exon-intron junctions [29, 33]. In addition, hybridization-based methods tend to have low sensitivity and specificity [30, 34, 35]. Finally, the above methods rely on short sequences (i.e., perhaps 500-800 bp maximum) to infer particular exon combinations, and long-range contiguity of splice choices along a maturing mRNA cannot be extracted from these data. Therefore, detailed study of candidate genes using classic molecular methods, like RT-PCR of an entire transcript, is the method of choice when the entire diversity of alternative transcripts is investigated.

Here, the diversity of Bru-3 ASTs was analyzed via RT-PCR in four Drosophila species - D. melanogaster, D. persimilis, D. pseudoobscura and D. virilis. These species represent three distantly related Drosophila groups that were separated by 55 million years of independent evolution [26]. Thus the evolutionary conservation of ASTs of Bru-3 along such divergence would suggest their functional significance [e.g., [36]]. We also studied the molecular evolution of the translated part of Bru-3, including patterns of exon gains and losses, using the published genome sequences of twelve Drosophila species [37]. Roy et al. [15] found that long introns were enriched with newly created exons compared to short introns. Our data presented an exciting opportunity to analyze the patterns of exon gains and losses in regards to intron lengths. Finally, we tested the effect of splicing constraint (i.e., constitutive versus alternative splicing) on the evolution of homologous RRM domains in Bru-3 among twelve different species.

Results

Molecular structure and AS of Bru-3 in Drosophila species

The structure of Bru-3

The highest number of exons in Bru-3 was found in D. pseudoobscura and D. persimilis. A total of sixteen exons encode Bru-3 in these species (Figure 1A and Additional file 1). Of these, thirteen exons encode parts of the open reading frame. The remaining three exons either entirely form UTRs (i.e., exons 1 and 15) or are only occasionally retained in the mature mRNA and do not have a consistent, long ORF (i.e., 'nonsense' exon 4a in Figure 1A). Single PCR products were amplified from the 5'- and 3'-end UTRs of Bru-3 in D. pseudoobscura and D. persimilis, and thus the exons encoding the UTRs - exons 1, 2, 14 and 15 - were constitutively spliced in both species. Only single synonymous substitutions in the ORF and a few short insertions/deletions and substitutions in the 5'- and 3'-end UTRs distinguish both D. pseudoobscura isolates and D. persimilis MSH93. We identified a single nucleotide insertion in the published genomic sequence of D. persimilis at exon 10, which introduces two premature termination codons (PTCs) and thus was likely a sequencing or assembly mistake (scaffold 36 at position 405678, initial assembly; Flybase.org). The longest ORF of 1281 nucleotides was identified in D. pseudoobscura, transcript #1 in Table 1. Still, exon 3 was skipped in this transcript, and thus the longest potential ORF of Bru-3 in D. pseudoobscura and D. persimilis is 1323 nucleotides. The shortest ORF was identified in D. persimilis and D. melanogaster, and was only 414 nucleotides (transcript 33 in Table 1). Four translated exons were included in the ORF of this transcript, exons 2, 4, 13 and 14. Only two exons from the ORF of Bru-3 (13 and 14) were constitutively spliced and translated in all tested species (Table 1). The other exons of the ORF were skipped in at least one AST and thus were alternatively spliced exons sensu stricto.

Table 1 Diversity of alternatively spliced transcripts of Bruno-3 identified in four different species of Drosophila
Figure 1
figure 1

Intron length and alternative splicing of Bru-3 in distantly related Drosophila species. (A) Splice choices identified in D. melanogaster, D. pseudoobscura, D. persimilis and D. virilis. Only lengths of exons (boxes), but not introns (lines connecting exons), are drawn to scale. The ORF is shaded. The untranslated regions (UTRs) of Bru-3 were not analyzed in D. virilis. For simplicity, exon skipping and alternative acceptor sites, but not intron splicing per se, are depicted. While many splice choices in Bru-3 were found conserved between tested species, the gene underwent evolution at the exon level. Exon 8 were created de nove from intronic sequence in D. pseudoobscura and D. persimilis sibling species pair, and exon 6 lost coding potential in D. melanogaster. (B) The gene structure of Bru-3 in D. pseudoobscura. All exons and the introns that are shorter than 2 kb are drawn to scale. The lengths of the other introns are presented. (C) The intron length and splicing of adjacent exons in Bru-3. The column heights are proportional to the natural logarithm of intron lengths in D. melanogaster, D. pseudoobscura and D. virilis. The columns are aligned between corresponding exons. The circles enclose the number of exons that included into the majority of ASTs. The horizontal line depicts the suggested cut-off, 250 nucleotides, between short- and long- introns [12, 54]. The shorter introns are likely spliced out through the intron definition mechanism whereas the longer introns relay on the exon definition mechanism for their splicing [12]. The intron definition mechanism of splicing was shown to be more precise and efficient than the exon definition mechanism [12]. We found that the exons adjacent to long-introns were frequently skipped in ASTs of Bru-3, whereas the exons connected by short-introns - exons 9, 10 and 11, and exons 13 and 14 (black bars) - were always spliced together.

Complex AS of Bru-3

Multiple ASTs of Bru-3 are expressed in adult flies of D. melanogaster, D. pseudoobscura, D. persimilis and D. virilis. We identified 22 distinct ASTs from the alignment of 49 sequenced clones of the entire ORF from D. pseudoobscura female and male flies [GenBank: GQ497346-GQ497396]. A smaller number of clones were sequenced for the other Drosophila species, and thus fewer distinct ASTs of Bru-3 were found for these species. 15 different ASTs were sampled for D. melanogaster, 8 ASTs for D. virilis, and 6 ASTs for D. persimilis (Additional file 2, GenBank: GQ497397-GQ497445). In addition, a few new exon combinations that represent new ASTs were inferred using primers bound to specific exons, and further increased the number of ASTs found in each Drosophila species (Additional file 2). Altogether, 35 distinct ASTs of Bru-3 have been identified across four Drosophila species (Table 1).

Multiple splicing events - exon skipping, alternative start codons, and alternative 3' splice sites (SS) - contributed to the diversity of transcripts of this gene. Although the diversity of ASTs of Bru-3 is likely underestimated for species other than D. pseudoobscura, the evolutionary conservation of individual splicing events and multiple ASTs along tens of MY of evolution is obvious (Table 1 and Figure 1A). Nine ASTs and two specific exon combinations - exons 3, 4, 5 and exons 5, 6, 7 - have been conserved in at least two distantly related Drosophila species, separated by 55 MY [transcripts with stars in Table 1, [26]]. For example, one distinct transcript that contained 'nonsense' exon 4a was identified in D. melanogaster and D. pseudoobscura (Table 1: transcript 29). The same two alternative 5' SS in exons 4 and 13 were used in D. pseudoobscura and D. melanogaster, and thus 18 bp-long truncations at the 5'-ends of these exons may be functionally significant in both species. The latter 5' SS is also used by D. virilis. While the alternative 5' SS in exon 10 appears to be an apomorphy of the melanogaster group (see below), the sequence at the alternative 5' SS in exon 11 is conserved among the twelve Drosophila species [37] and may be used as a SS in other Drosophila species (Figure 1A). Overall, many ASTs as well as specific splicing events of Bru-3 are conserved between distantly related Drosophila species.

A few mature mRNA included 'nonsense' exon 4a

Six clones of the entire ORF of Bru-3 from D. pseudoobscura and D. persimilis included an additional sequence, called here 'nonsense' exon 4a. These clones were derived from four independent cDNA pools of both D. pseudoobscura isolates and D. persimilis MSH93, and represent distinct ASTs (transcripts 10, 18, 22, and 29 in Table 1). The inclusion of exon 4a into transcripts introduced multiple PTCs and thus possibly rendered them non-functional for protein translation (aka 'nonsense'). The sequence of exon 4a is highly conserved between D. pseudoobscura and D. persimilis. Only two insertion/deletions of a single nucleotide and 5 substitutions out of 454 nucleotides distinguish both species. Exon 4a derives from the longest intron in Bru-3 - 40,652 bp - between exons 4 and 5; 6,664 nucleotides upstream of exon 5 in D. pseudoobscura (r2.3, FlyBase.org). Interestingly, the sequences that appeared homologous to 'nonsense' exon 4a in D. pseudoobscura were also identified in two ASTs of D. melanogaster (transcripts 29 and 34 in Table 1). D. melanogaster exon 4a had the same length, introduced PTCs, and originated from the extremely long intron of Bru-3 - 36,042 bp - between exons 4 and 5; 5,918 nucleotides upstream of exon 5 (r5.13, FlyBase.org). However, while exon 4a originated from the homologous positions in D. pseudoobscura and D. melanogaster genomes, exon 4a was not conserved between two Drosophila species at the sequence level. In fact, both sequences appeared totally unrelated. Because exon 4a does not have the splice sites matching both U2- and U12-classes of spliceosomes [38] in both Drosophila species: once retained it cannot be spliced out individually from maturing mRNAs to re-establish their protein-encoding potential.

Alternative start codon of Bru-3

We inferred an alternative start codon for Bru-3 located in exon 7. Exons 2 and 7 were spliced to each other in D. melanogaster transcript 31 (Table 1). Although skipping exon 4 offset the reading frame starting from exon 2, the alternative start codon located at the 3'-end of exon 7 could be used in transcript 31. The resulted polypeptide chain would skip the entire N-terminal RRM, but it would include the other functionally important domains of Bru-3, like the linker region and C-terminal RRM [20]. Therefore, transcript 31 may be a valid, albeit rare, transcript encoding a functional protein in D. melanogaster.

Exon 6 can be spliced in the middle of the N-terminal RRM

Each Bru-3 RRM is formed by two exons. Exons 5 and 7 together make the N-terminal RRM. We identified exon 6 in D. pseudoobscura that was located between these functionally linked exons in the genomic sequence and was spliced together with exon 7 (Table 1). However, none of the sequenced clones of Bru-3 had exon 6 retained between exons 5 and 7. To test whether the constraints acting on the functional domain would prevent the production of Bru-3 transcripts with exon 6 disrupting the RRM, we screened cDNA pools from D. pseudoobscura, and D. persimilis for transcripts with consecutive exons 5, 6 and 7. Amplicons with exons 5, 6 and 7, as well as exons 5 and 6 spliced together were found in both species (Additional file 3). Because exon 6 does not have a 3' SS, once retained between exon 5 and 7, it cannot not be spliced out by itself.

Intron size and AS in Bru-3

In the translated segment of Bru-3, exons separated by large introns were skipped more often than exons divided by small introns, less than 250 nucleotides. Furthermore, the consecutive exons that were always spliced together in ASTs were divided by three shortest introns in the gene - all less that 160 bp (Figure 1B, C) in the tested Drosophila species. The two shortest introns - introns 9 and 10 - divide exons 9, 10 and 11. These exons were always spliced together: if any of these three exons was included in the transcript the others were also included (Table 1). The third shortest intron - intron 13 - divides exons 13 and 14 (Figure 1B). These exons were included in every ASTs of Bru-3 identified in four Drosophila species (Table 1). The next shortest intron separating the translated exons in Bru-3 - intron 11 - was already unusually large for Drosophila (Figure 35 in [18]): 847 bp in D. melanogaster, and 1024 bp in D. pseudoobscura. Less than 9% of all D. melanogaster and D. pseudoobscura introns have the same or longer sizes, and 86 and 69 nucleotides are the median intron sizes is both species, respectively. Intron 11 divides exon 12 from the group of contiguously spliced exons 9, 10 and 11 (Figure 1B). Perhaps not surprisingly, exon 12 was skipped in more that half of the transcripts that included exons 9, 10 and 11 (Table 1). The 5'-end part of Bru-3 ORF - exons from 2 through 9 - has larger introns that the 3'-end part of ORF, and overall the higher number of AS events were identified in the 5'-end part of the ORF (Table 1). Nevertheless, exons 4 and 7, each located adjacent to large introns, were faithfully included into almost all surveyed transcripts, precluding statistically significance in a simple association between intron size and exon skipping.

Evolution of Bru-3 in Drosophila

Exon turnover during Bru-3 evolution

Two previously unknown exons (6 and 8) of Bru-3 were identified in D. pseudoobscura. These exons encode short polypeptide chains in Bru-3 protein - 17 and 19 amino acids, respectively. No previously sampled ASTs from Drosophila species other than D. pseudoobscura included either of these exons (Table 1). However, we confirmed via RT-PCR that both exons are included in mature mRNA of D. persimilis, and exon 6 is included in mature mRNA of D. virilis (Figure 2A, B and Additional file 2). The location of each exon in the twelve published genomes of Drosophila species is indicated in Additional file 1. We have not found sequences homologous to exon 6 in species of the melanogaster group whose genomes were sequenced [37]. Likewise, the sequences homologous to exon 8 could not be found in Drosophila species other than D. pseudoobscura and D. persimilis. When we mapped the presence of exons 6 and 8 onto the phylogeny of twelve Drosophila species, we found that exon 6 was lost along the branch leading to the melanogaster group, while exon 7 appeared to be an apomorphy of the obscura group (Figure 3).

Figure 2
figure 2

cDNA screens for specific splice choices of Bru-3. In the diagrams, arrows indicate the position and direction of primers in regard to the exons (e.g., E6, E7, E8 and E9) used for PCR. The expected sizes of PCR products (in nucleotides) are presented below the diagram. In the gel images, fly lines are indicated for each track. The long dash marks the track of a negative control (no template). Sizes of DNA ladder (ML) are shown to the right of the gel image. (A) Screening cDNA pools for the splice variant containing exon 6. The exon was included in mature mRNAs of D. pseudoobscura (pse), D. persimilis (per) and D. virilis (vir), whereas it was absent in D. melanogaster (mel) cDNA pool. (B) Screening cDNA pools for the splice variant containing exon 8. Although exon 8 was included in mature mRNAs of all tested species of the obscura group - D. obscura (obs), D. affinis (aff), D. bifasciata (bif), D. subobscura (sub) - it was not found in cDNA pools of the species outside of the obscura group, in D. melanogaster (mel) and D. virilis (vir).

Figure 3
figure 3

Phylogenetic positions of exon gain and loss in Bru-3. The coding sequences of Bru-3 were inferred for 8 Drosophila species from their genomic sequenced available at FlyBase.org [37]. A distance tree was calculated in PAUP 4.0b10 [99] from the coding parts of eleven exons of Bru-3 (exons 2-5, 7, 9-14) shared by twelve Drosophila species. Applying nucleotide BLASTing [90], we found that exon 6 was lost along the branch leading to the melanogaster group (the top shaded box), whereas exon 8 was de novo created in the obscura group (the bottom shaded box). We verified both events for Drosophila species in bold by amplifying and aligning the orthologous genomic regions that include exons 6 and 8 (supplementary files S1 and S2). In addition cDNA pools from the species in bold, except for D. simulans, were screened for inclusion of conserved exon 6 and 8 in mature mRNA (fig. 2). The status of exon 6 indicated with a circle; a black circle, the exon was identified in cDNA pool; a white circle, the exon was absent. A square was used to indicate the status of exon 8.

The proposed cases of exon loss and gain are not simple mistakes of genome assembly. We sequenced the genomic regions homologous by their positions to the regions that encompass exons 6 and 8 in D. pseudoobscura for D. simulans, D. melanogaster, D. erecta, D. ananassae, D. pseudoobscura, D. persimilis, D. mojavensis and D. virilis, and compared to their genomic counterparts. We have not found any mistakes of genome assembly that could explain the exon loss and gain in Bru-3, and thus the exon turnover is genuine. Screening of cDNA pools from D. melanogaster, D. erecta, D. ananassae, D. pseudoobscura, D. willistoni, D. mojavensis, and D. virilis for the presence of exons 6 and 8 further supported our conclusion. Exon 6 was identified in cDNA pools from D. pseudoobscura, D. mojavensis, and D. virilis, whereas exon 8 was found only in D. pseudoobscura cDNA pool (Figure 2A, B).

To investigate focal changes that could lead to the loss and gain of respective exons, we aligned the sequenced fragments encompassing exons 6 and 8 in D. simulans, D. melanogaster, D. erecta, D. ananassae, D. pseudoobscura, D. persimilis, D. willistoni, D. mojavensis and D. virilis (Additional files 4 and 5). Sequences clearly homologous to exon 6 were found in D. willistoni, D. mojavensis, and D. virilis. These sequences were conserved at both the amino acid and nucleotide levels. In contrast, although the tested species from the melanogaster group had the conserved 3' SS of the intron upstream of exon 6, the sequences homologous by position to exon 6 were not conserved in these species (Additional file 4). The region corresponding to exon 6 had an in-frame PTC in D. melanogaster, D. simulans and D. ananassae. In addition, D. ananassae had an insertion in this region. Therefore, the local deletions/insertions and/or substitutions at the 5' SS flaking exon 6 likely led to the loss of coding potential and subsequent decay of exon 6 in the melanogaster lineage (Figure 3). Higher sequence conservation was detected in the genomic fragments encompassing exon 8. The location of two 30 nucleotide-long ultraconserved regions near 5' SS facilitated the alignment of these fragments. Although all tested Drosophila species shared both 3' and 5' SSs flanking exon 8, the region corresponding to exon 8 in D. pseudoobscura and D. persimilis was not conserved at either amino acid or nucleotide levels in the other tested species (Additional file 5). We also found that D. simulans, D. melanogaster, D. ananassae, D. mojavensis and D. virilis all had at least one in-frame PTC, while D. willistoni had a 47 nucleotides-long insertion at the region corresponding to exon 8. Therefore, this region does not encode a polypeptide chain in any tested species except D. pseudoobscura and D. persimilis, and the ORF of exon 8 was likely created in situ from intron sequence in the obscura group.

Exon 8 originated early in the evolution of the obscura group

To study the origin of exon 8 in further detail, the genomic fragments encompassing exon 8 were sequenced for the species representing all major lineages in the obscura group [e.g., [39, 40]]. All tested species from this group - D. pseudoobscura bogotana, D. miranda, D. obscura, D. bifasciata, D. subobscura and D. affinis - possessed the conserved exon 8. The last four of the above species each had one non-synonymous substitution in the 3'-end of exon 8 and a few substitutions at 3' and 5' SS around this exon (Additional file 6). To test whether these substitutions at the 3' and 5' SS could affect the inclusion of exon 8 into Bru-3 mRNA, we screened cDNA pools of D. pseudoobscura and D. persimilis, D. obscura, D. bifasciata, D. subobscura and D. affinis for its presence. Exon 8 was identified in the RNA pools of males and females from each tested obscura species, but not in D. melanogaster and D. virilis (Figure 2B). Therefore, exon 8 originated early in the lineage of the obscura group before the major branches [39, 40] within this group appeared.

New 3' alternative splice site in exon 10

The longer version of exon 10 was identified in four ASTs of D. melanogaster (Table 1: transcripts 3, 7, 8 and 26). This version of exon 10 was 24 nucleotides longer that the previously known size of this exon in D. melanogaster (i.e., 277 nucleotides, FlyBase.org: r5.13). We analyzed the origin of the alternative 3' SS of exon 10 using the published genomes of Drosophila species [37]. While all twelve Drosophila species share the common 3' SS upstream of exon 10, an additional 3' SS was created along the lineage leading to the melanogaster group (Figure 3). The splicing at this new 3' SS would add 30 bp to D. erecta exon 10, 18 bp to D. ananassae exon 10, and 24 bp to exon 10 in the other four species in the melanogaster group (Additional file 7). While no potential 3' SS could be inferred in the homologous intron for D. grimshawi, D, mojavensis, D. virilis, and D. willistoni, three such sites were identified upstream of exon 10 in D. pseudoobscura and D. persimilis. However, we doubt that these additional 3' SS are active in both species. The minimum intron length required for successful splicing is around 45 nucleotides in Drosophila species [e.g., [41, 42]], whereas these 3' SS upstream of exon 10 were only 15, 21, and 27 nucleotides apart from the 5' SS downstream of exon 9, and exons 9 and 10 were always spliced together (Tables 1). Therefore, the alternative 3' SS in exon 10 is a true apomorphy of the melanogaster group.

Splicing constraints affect the rate of sequence evolution

Splicing constraints (i.e., alternative or constitutive splicing) affect the evolutionary potential of a gene region. Constitutively spliced regions are expected to experience stronger functional constraints than alternatively spliced regions of the same gene [e.g., [10, 43]]. Bru-3 allows the comparison of substitution patterns between two homologous domains present in the same gene that experience different splicing constraints: AS RRM versus CS RRM. Because these RRM were paralogous domains and thus potentially shared the common ancestry, this comparison controlled better for the variables that could bias the analysis. The AS RRM had more differences (i.e., accelerated evolution) in comparison to the CS RRM (i.e., AS RRM: 192 bp, 19.8% sites with substitution, across species = 0.0744 ± 0.0095, KS = 0.4034 and KA = 0.0340; CS RRM: 219 bp, 13.7% sites with substitution, across species = 0.0489 ± 0.0067, KS = 0.2047 and KA = 0.0), and this difference was marginally significant (G2 = 2.75, p = 0.097).

Discussion

Recent bioinformatic analyses of genome databases pointed to a correlation between intron length and alternative splicing [AS, e.g., [11, 14, 15]]. We studied in-depth the interplay between intron length and AS within the Drosophila Bruno-3 (Bru-3) gene. The molecular structure of this gene, multiple exons separated by large introns, was expected to be highly conducive to AS. Further, multiple Bru-3 exons are split by introns of remarkably diverse lengths (Figure 1B, Additional file 8), and thus presented a representative sample of intron sizes. Although only a single gene, extensive sampling of the diversity of Bru-3 ASTs across distantly related Drosophila species provided a case-study for several fundamental questions dealing with the interaction between gene structure, AS and gene evolution.

The sensitivity and specificity of our analysis of ASTs was accomplished by reverse transcription coupled with subsequent PCR and cloning. We identified 35 distinct ASTs of Bru-3 transcribed in Drosophila adult flies (Table 1), 33 of these ASTs were previously unknown. Notably, only seventeen ESTs of the 544,789 total D. melanogaster ESTs map to Bru-3, whereas for example, 403 ESTs map to myosin light chain 2 (UniGene build #59). It is well known that genes poorly represented in EST collections are expressed at a low level. However, such genes are more likely to be regulatory in function and undergo AS [e.g., [3]], like Bru-3. Cloning and sequencing of the complete ORF allowed the analysis of long-range contiguity of splice choices along the entire translated region of Bru-3. The contiguity of splice choices along a mature mRNA could not be inferred from either EST collections or by currently popular high-throughput methods; including classic microarrays, tiling arrays [e.g., [34, 44, 45]] and pyrosequencing of entire transcriptomes [e.g., [30]]; because frequencies of usage of individual splice choices and/or exons along mRNAs cannot be easily concatenated into distinct splicing variants. Below, we elaborate on several specific findings from our study.

Conservation of alternative splice variants

High conservation of ASTs of a gene implies their functional significance [e.g., [36]]. We screened for Bru-3 ASTs in four Drosophila species. These species represented three distantly related species groups - melanogaster, obscura, and virilis - that were separated at least 55 million years (MY) ago [26]. D. pseudoobscura and D. persimilis are closely related species, which diverged less than 1 MY ago [26], and frequently share long haplotypes. Complex alternative splicing of Bru-3 was found in each tested species, and many individual ASTs and splicing events remained conserved along 55 MY of independent evolution (Table 1 and Figure 1A). For example, the short transcript of Bru-3 that included only 4 exons out of 13 translated exons (identified in D. melanogaster and D. persimilis) can be a functional transcript. Likewise, 18 nucleotide-long truncations in exons 4 and 13 may reflect either a function of the ASTs that includes these truncated exons or a common error-generating mechanism in D. melanogaster and D. pseudoobscura. The great diversity of conserved ASTs of Bru-3 uncovered in our study implied that this gene has a complex regulatory potential with many target genes. This spectrum of potentially distinct functions of Bru-3 is performed via ASTs of the same gene, not via gene duplications [46]. Therefore, many distinct transcripts need to be considered to study the function(s) of Bru-3 in Drosophila.

Alternative splicing and accelerated evolution

Alternatively spliced exons (ASEs) are expected to undergo faster divergence than constitutively spliced exons (CSEs) at non-synonymous sites, because they are not included in every transcript of the gene and thus experience weaker functional constraints at the protein level [e.g., [43, 47, 48]]. The relaxation of functional constraints in ASEs can indirectly accelerate their evolution and permit additional diversification. Thus ASEs tend to have a higher frequency of non-synonymous substitutions than CSEs [e.g., [43, 47–49]]. However, higher sequence conservation was previously found at synonymous sites in ASEs [e.g., [48, 50]] and the introns flanking ASEs [e.g., [49, 51]] than those in CSEs. To explain this counter-intuitive inference, Xing and Lee [43, 48] suggested that ASEs experienced the stronger "RNA-level selection pressure" than CSEs, and thus ASEs had stronger conservation at the synonymous sites than CSEs. Indeed, AS requires the involvement of more regulatory elements than constitutive splicing [51, 52]. However, it is unclear why "RNA-level selection pressure" would affect non-synonymous sites differently from synonymous sites. At the same time, ASEs tend to have smaller sizes than CSEs [e.g., [48, 49, 53]], and thus on average, splicing regulatory elements can occupy the higher proportion of the sequence in ASEs than those in CSEs. This bias was not always accounted for in bioinformatic analysis of exon collections [but see [49]].

The two paralogous RRMs in Bru-3 provided a system to test the effect of splicing constraints on the evolution, since their common ancestry supported the strongest control for potential biases. The C-terminal RRM was a CS domain, while the N-terminal RRM was an AS domain (Table 1). To account for the potential effect of exon size differences, we compared only the sequences of the RRM domains (AS RRM: 192 bp, CS RRM: 219 bp). The AS RRM experienced accelerated evolution compared with the CS RRM at both synonymous and non-synonymous sites (i.e., KS and KA). Therefore, after accounting for the different biases, we observed a trend indicating that the AS RRM experienced weaker purifying selection pressure at both synonymous and non-synonymous sites than the CS RRM in Bru-3, and thus the AS of the N-terminal RRMs could facilitate the evolution of an additional secondary specialization of this domain.

Intron length can promote alternative splicing

The positive correlation between exon skipping and increasing lengths of the flanking introns [11, 14] has been linked to the switch from the intron definition mechanism of splicing to the exon definition mechanism [12]. The intron definition mechanism operates effectively across short introns, whereas the exon definition becomes dominant when flanking introns are longer that 200-250 nucleotides [12], and thus splice sites are recognized across exons [42, 54]. These mechanisms offer distinct predictions of splicing errors. Failures of the exon definition mechanism lead to exon skipping, while mistakes of splicing through the intron definition mechanism result in intron retentions [e.g., [11, 54]]. However, splicing errors of extremely large introns via recursive splicing may also result in intron retentions. Fox-Walsh et al. [12] showed that the intron definition mechanism was more efficient at recognition of weak splice sites than the exon definition mechanism. The results of our analysis of AS in Bru-3 support these conclusions. The majority of exon skipping events mapped to the region of Bru-3 that included exons separated by long introns (Figure 1). Exons 2 through 9 in twelve Drosophila species [37] were separated by introns of at least 8,000 nucleotides (Additional file 8). In contrast, the translated exons that were always spliced contiguously in the four tested Drosophila species - exons 9, 10 and 11, and exons 13 and 14 - were separated by the smallest introns in the gene, each less than 200 nucleotides (Figure 1C). Therefore, the splicing of these exons was expected to proceed through the intron definition mechanism and thus be more efficient than that of other exons in the ORF of Bru-3. In the gene region between exons 9 and 14, only exon 12 was flanked by large introns (at least 847 nucleotides), and perhaps not coincidentally, exon 12 was skipped in more than a half of ASTs (Table 1). However, this trend was not perfect: exons 4 and 7 were included in almost all transcripts surveyed despite adjacency to large introns. Nonetheless, intron length can be important causal factor orchestrating the inclusion level of exons.

Intron length can also indirectly promote AS in general, because longer introns can support a finer level of AS regulation than the short introns. However, the alternative 3' SSs were identified equally frequently in both small and long introns of Bru-3: introns 9 and 10 versus introns 3 and 12, respectively (Figure 1A). Extremely large introns influence gene expression via their sheer size. For example, transcription of intron 7 in Bru-3 requires approximately 40 min [assuming 1 kb per min, [55, 56]], and thus large introns can substantially delay mRNA maturation. Regulatory genes may use the interplay between transcript elongation rate and splicing kinetics to facilitate the regulation of AS [17, 56–58]. In addition, long introns can contain many regulatory elements that are necessary to control AS at the tissue and temporal level [59, 60]. Many regulatory genes known to have ASTs contain extremely large introns, including muscleblind, Ultrabithorax, fruitless, kuzbanian, and pumilio [UniGene and [17]].

Intron length and the origin of coding diversity

The preferential gains of new exons within long introns can also explain the connection between exon-skipping and long introns. AS is strongly associated with recent exon origin [e.g., [33, 43, 47, 61–64]] and the rate of exon gains is inversely related to exon inclusion level [e.g., [15, 63, 64]]. In other words, recently arisen exons tend to be included only in a few ASTs of a gene (i.e., be minor-form ASEs). Using phylogenetic analysis of 17 vertebrate genomes, Roy et al. [15] found that newly originated exons were more enriched within longer introns compared with short introns. Therefore, longer introns may have slightly higher probability of exon gains than shorter introns due to the higher number of mutable sites [15].

We identified one new putatively functional exon that was created from an intron sequence de novo, exon 8 in D. pseudoobscura. Although this exon is flanked by the 5' and 3' SSs which are conserved among Drosophila species separated by 60 MY of independent evolution [26], only the species of the obscura group have a true ORF between these SSs (Figure 2). The tested Drosophila species outside the obscura group all have a non-translated sequence at the homologous location to exon 8 (Additional file 5), in the largest intron of Bru-3 - 41,973 bp in D. melanogaster and 45,859 bp in D. virilis (Figure 1C). As expected for a newly originated exon, exon 8 is a minor-form ASE. Only one sequenced clone of ORF of Bru-3 included exon 8 (Table 1), albeit subsequent screening found that this exon was included in mature mRNAs in both males and females of D. pseudoobscura and D. persimilis, as well as in the other four tested species from the obscura group (Figure 2B). Because all tested Drosophila species share some level of conservation between these SSs, and exon 8 is only 57 bp long, the sequence change that capacitated exon 8 were highly focal. Duplication-translocation of existing exons [e.g., [65, 66]], retrotransposition [e.g., [67]] or exaptation of repetitive elements [e.g., Alu element, [62, 68]] cannot explain the origin of exon 8. The functional ORF of exon 8 was created de novo from intronic sequence between the existing SSs that might previously function in recursive splicing of a huge intron [16, 17]. The origin of exon 8 is especially remarkable, since all well-studied cases of de novo exon creation from intronic sequence are known from bioinformatic analyses of vertebrate genomes [e.g., [63, 64, 69]]. We are familiar with only one study in which five novel genes were inferred to originate from non-coding sequences in D. melanogaster [70]. However, unlike Bru-3 exon 8, four of these five genes had paralogous copies in the genome and thus their origins were associated with duplication-translocation.

The pre-existing location of sequences forming both SSs around exon 8 could play a critical role in the creation of this exon. In contrast, the deletion of one SS could instantly make the existing exon invisible to splicing machinery and thus would result in complete exon loss from mRNA and protein. The removal of translational constraint makes an exon sequence prone to mutational decay, akin to a pseudogene. We found one example of such 'turning off' of an existent exon in Bru-3. The deletion of the 3' SS flanking exon 6 along the lineage to the melanogaster group abolished its expression in this group, though this exon was included in mature mRNAs in D. pseudoobscura, D. persimilis, D. virilis and D. mojavensis and appeared functional in the other analyzed Drosophila species (Figures 2A and 3). Notably, the focal deletion of the 3' SS of exon 6 also took place between two large introns 5 and 6, which are 8463 bp and 1188 pb, respectively in D. pseudoobscura. Thus, this mutation also happened inside a large intron and led to the disappearance of exon 6 in the melanogaster group.

If large introns are conducive for exon gain and loss (i.e., exon turnover), we also expect to see more splicing mistakes in large introns compared to short introns. We identified one 'nonsense' exon 4a that originated from the largest intron of Bru-3. This exon was included into multiple ASTs in D. pseudoobscura, D. persimilis and D. melanogaster, and introduced multiple PTCs to mRNA (Table 1). Exon 4a is surrounded by conserved SSs in a proper orientation. The 3' SS located upstream of exon 4a can be a strong SS since it has a nearly 50 nucleotide-long pyrimidine stretch. This exon does not share major common characteristics of internal ASEs. First, its size is unusually large for internal ASEs in Drosophila. The average exon length in D. melanogaster is 150 nucleotides and fewer than 5% of exons have larger sizes than 454 nucleotides Figure 35 in [18]. Second, the total number of nucleotides in exon 4a is not divisible by three unlike in the majority of ASEs, especially in minor-form ASEs [e.g., [52, 53, 71]]. Both of these predicted features of ASEs result from the selection pressure to preserve an ORF (e.g., exons 3, 5, 6, 8 and 12 in Table 1). In addition, the sequence of exon 4a looks completely unrelated between D. pseudoobscura and D. melanogaster, while both species share the location of the SSs flanking this exon. These SSs are conserved between distant species because they likely function in recursive splicing of a 36 kb-long intron [17]. We think that the neighboring location of both SSs may bias toward the erroneous inclusion of the encompassed intronic sequence (i.e., 'nonsense' exon 4a) into Bru-3 mRNA.

Large introns have more opportinities to mutate than short introns due to the higher number of potentially mutable sites. In addition, the preferential location of large introns in genome areas with low recombination [72, 73] can further increase the spread of new mutations within them compared to short introns. We know from our work on recombination variation in D. pseudoobscura that Bru-3 resides in an extremely cold recombination spot [74]. Notably, the deletion-biased mutation ratio (relative to insertions) known for Drosophila [75, 76], which is expected to be especially strong in the areas of low recombination [77, 78], has not led to the shortening of large introns in Bru-3 and other genes. At the same time, the presence of large introns in genes likely increases the energetic cost of their transcription, and thus highly expressed genes in humans, Caenorhabditis and Drosophila tend to have small introns [e.g., [60, 79, 80]]. Therefore, large introns either have some selective advantage or they can be more easily tolerated in some genes. Comeron and Kreitman [73] suggested that the large introns could decrease crossover interference between exons in the areas with low recombination [but see also, [72, 81]], and thus they could be advantageous. Our data suggest that large introns can have yet another beneficial function. Large introns can contribute to the expansion of transcript diversity encoded by a single gene via both AS and de novo creation [82] or deletion of exons. Failures of proper splicing of large introns generate both exon skipping and intron retentions (i.e., via recursive splicing) in transcripts, which are screened out by nonsense-mediated mRNA decay (NMD) pathway [e.g., [83–85]]. The tolerance of suboptimal splicing efficiency and accuracy via NMD in turn may allow natural selection to 'sculpt' retained intronic sequences, like exon 4a in Bru-3, into a new translated exons, like exon 8 in Bru-3. Many species-specific ASEs in mammals were likely created de novo from purely intronic sequence [e.g., [15, 63, 64, 86]] potentially by the similar mechanism.

Conclusion

Drosophila Bru-3 has many exons that are separated by introns of extremely diverse lengths. The in-detail analysis of the ASTs of Bru-3 has inferred much greater diversity of ASTs that has been previously unknown. Our analysis of inclusion frequency of Bru-3 exons supported a positive correlation between exon skipping and surrounding intron length. The majority of cassette exons were located in the gene region with the longest introns, between exons 2 and 9. Conversely, the exons separated by short introns, i.e., less than 250 nucleotides, were always found splicing consecutively. In addition, cases of evolutionary gains and losses of exons were also mapped to long introns. Notably, we established that exon 8 was created in situ from intronic sequence positioned between cryptic SSs, which conserved among twelve Drosophila species. The presented findings support the role of gene structure in promoting and/or constraining gene's AS. However, a correlation between the intron length and AS does not imply the causality. Both gene structure and AS can be shaped by the interplay of selection, recombination and gene-gene interaction. Further study will determine whether the AS of Bru-3 is a general or special case.

Methods

Drosophila lines and their rearing

Two isolates of D. pseudoobscura - Mather 17 and Flagstaff 1993 [described in [87, 88]], D. melanogaster (Oregon-R), D. virilis (San Diego Drosophila Stock Center accession [SDDSC] 15010-1051.87) and D. persimilis Mount St. Helena 1993 (MSH93) were used for analysis of the AST diversity of Bru-3. To study exon gains and losses during Bru-3 evolution in Drosophila, we examined additional lines - D. simulans (SDDSC 14021-0251.195), D. erecta (SDDSC 14021-0224.01), D. ananassae (SDDSC 14024-0371.13), D. mojavensis (SDDSC 15081-1352-26), D. willistoni (SDDSC 14030-0811.24), and species from the obscura group: D. pseudoobscura bogotana El Recreo line [89], D. miranda (SDDSC 14011-0101.08), D. affinis (SDDSC 14012-0141.04), D. obscura (SDDSC 14011-0151.01), D. subobscura (SDDSC 14011-0131.09), and D. bifasciata (SDDSC 14012-0181.02). All flies were maintained under a constant regime of temperature (20°C) and humidity (80%) in 12-h dark-light cycle. A standard mixture of agar, dextrose and yeast was used to rear flies.

RNA extraction

Total RNA was extracted from 20-30 male and female flies using TRI reagent® and the DNA-free™ kit (Ambion). Only freshly collected flies were used for RNA extraction, tissue was ground in TRI reagent®, and proteins and lipids were removed using a chloroform wash. Total RNA was then precipitated and cleaned with isopropanol and ethanol, respectively. Co-precipitated DNA was digested with DNase I, and the remaining RNA was again precipitated and cleaned.

RT-PCR of ORF of Bru-3 mRNA

D. pseudoobscura, D. persimilis, D. melanogaster and D. virilis cDNA pools were prepared from RNA extractions using the SuperScript™ III first-strand Synthesis System for RT-PCR (Invitrogen). Reverse transcription was performed with Oligo(dT)20 at 50° for 50 min following the manufacturer's protocol. To amplify the putatively translated sequence of Bru-3, we designed primers specific to the 5' and 3'-ends of the ORF (i.e., Bru3Dpse(2)20F and Bru3-3UTR, Additional file 9), which are highly conserved among the twelve Drosophila species [37]. PCR was conducted in a 30 μl volume with 4 μl PCR buffer, 6 μl Q solution (Qiagen), 4.0-4.5 mM total MgCl2, 0.5 μM of each primer, and 3-4 units Taq DNA polymerase (New England BioLabs). A touchdown temperature profile was used for PCR: 10 cycles at an annealing temperature of 61° for 40 sec, 10 cycles at 60.5° for 37 sec, and finally 20 cycles at 60° for 35 sec. The denaturing and elongation temperatures were the same across all cycles: 94° for 40 sec and 72° for 2 min, respectively. 15 μl of PCR reaction was used to visualize an amplicon on a 1.5% TBE agarose gel. The remaining PCR reaction was used for cloning. We used the StrataClone™ PCR cloning vector pSC-A (Stratagene) following the manufacturer's protocol. Inserts were amplified from the transformed colonies via PCR in 30 μl, and PCR halves were arrayed on an agarose gel. The amplicons of different sizes from the same cDNA pool were preferentially sequenced to assess the diversity of ASTs expressed in adult flies. The remaining 15 μl of a PCR reaction were cleaned by applying 1.8 μl of a mixture (i.e., 1 U to 4 U) of Exonuclease I (Exo I) and Shrimp Alkaline Phosphatase (SAP, Fermentas) and incubating at 37° for 35 min. The cleaned amplicons were sequenced in both directions with BigDye v3.0 (Applied Biosystems) and run on an Applied Biosystem 3700 automated DNA sequencer.

Alignment of ASTs of Bru-3

Independent clones of Bru-3 transcripts amplified from the same cDNA pool of D. pseudoobscura had different lengths, and thus were ASTs of the same gene. To identify a number of exons and positions of exon junctions in D. pseudoobscura mRNA, a few longest sequences of Bru-3 were BLASTed [BLASTN, [90]] against the genomic sequence of D. pseudoobscura [r2.0.1, [91]] using Drosophila Species Genomes BLAST [DroSpeGe, [92]92]. We verified that the 5' (donor) and 3' (acceptor) SSs flanking introns [i.e., GT and AG, [38]] were directly adjacent to each inferred exon junction in the genomic sequence. Then, D. pseudoobscura sequences were manually aligned in Se-Al v2.0 a11 [93] following the inferred positions of exon junctions in Bru-3 mRNA. Sequenced transcripts of D. persimilis, D. melanogaster and D. virilis, as well as two D. melanogaster transcripts available at FlyBase.org were added manually to the entire alignment of D. pseudoobscura transcripts.

Inference of Bru-3 ORF from genomic sequence assemblies of Drosophila species

To infer the translated sequences of Bru-3 in twelve sequenced Drosophila species [37], the sequences of individual exons of D. pseudoobscura Bru-3 were BLASTNed [90] against the genomic sequence assemblies of Drosophila species (D. melanogaster r4.3.0, D. simulans, D. sechellia, D. yakuba, D. erecta, D. ananassae, D. pseudoobscura r2.0.1, D. persimilis, D. willistoni, D. mojavensis, D. virilis, and D. grimshawi) at DroSpeGe http://insects.eugenes.org/species/blast/. Initially, a high statistical significance threshold [i.e., 0.01 matches were expected to be found by chance according to the model of [94]] was used for BLASTNing. If no matching sequence was found in the first round, a lower significance threshold was used for the next search. We verified that each exon junction in mRNA had the 5' and 3' SS directly adjacent to it in the genomic sequences. The exonic sequences inferred for each Drosophila species were concatenated and manually aligned to our D. pseudoobscura sequence of Bru-3 using Se-Al v2.0a11 [93].

Exon turnover in Bru-3

Two putatively novel exons of Bru-3, exons 6 and 8 (Figure 1A), were identified in D. pseudoobscura and D. persimilis. These exons were absent in the two transcripts of Bru-3 described in D. melanogaster (Flybase.org). We failed to find homologous exons in the melanogaster group species using both BLASTNing and TBLASTNing. To confirm that the identified exons were absent from a genomic sequence and their disappearance is not a genome assembly mistake, we amplified and sequenced the homologous regions encompassing exons 6 and 8 in D. melanogaster, D. simulans, D. erecta, D. ananassae, D. pseudoobscura, D. persimilis, D. mojavensis, D. willistoni and D. virilis using primers to the conserved sequences flanking these homologous regions (Additional file 9). The PCR and sequencing of amplicons were conducted as above. The obtained sequences were aligned with the corresponding genomic regions downloaded from DroSpeGe using the web-based T-Coffee [95] under default settings. The preliminary alignments were manually edited in Se-Al.

The above assay tested for mistakes in genome assemblies only at the homologous positions. However, exons 6 and 8 could be translocated into non-homologous locations and still participate in the formation of Bru-3 mRNA [e.g., via trans-splicing [96]]. Therefore, we also screened the cDNA pools of D. melanogaster, D. erecta, D. ananassae, D. mojavensis, D. willistoni, D. virilis and, as a positive control, D. pseudoobscura for the presence of sequences similar to exons 6 and 8 using PCR. This screen assumes that the sequences of exons 6 and 8 in the tested species remained conserved to those in D. pseudoobscura. The primers bound to exons 6 and 9 were used to screen cDNA for exon 6. For exon 8, we used two pairs of primers: the 1st pair was specific for exons 7 and 8; and the 2nd bound to exons 8 and 10 (Additional file 9). PCR was performed as above, but Q solution (Qiagen) was excluded from a PCR cocktail. Amplicons were sequenced in both directions to verify their specificity. Finally, to thoroughly analyze the origin of exon 8 within the obscura group, both tests were also performed for the species representing its four major subgroups: the obscura, subobscura, pseudoobscura and affinis subgroups [39, 40].

Screening cDNA pools for specific exon combination

Only transcript 6, which was identified in D. pseudoobscura, had exons 3, 4, and 5 spliced contiguously (Table 1). We failed to sample the ASTs with this particular exon combination from any other tested species. However, one of the two known transcripts of D. melanogaster has this exon combination (transcript 2 in Table 1). To test for the presence of this exon combination in other Drosophila species, we screened cDNA pools using PCR with primers bound to exons 3 and 5 (Additional file 9). We also screened for one potential exon combination - contiguously spliced exons 5, 6, and 7 - that has not been found in any tested species (Table 1). Amplicons were checked on a 2% TBA agarose gel, purified using Exo I and SAP (Fermentas) digestion and sequenced in both directions to verify their specificity.

Splicing constrains and molecular evolution

Splicing constraints (i.e., alternative or constitutive splicing) can affect the evolutionary potential of a gene region. Constitutively spliced regions are expected to experience stronger functional constraints than alternatively spliced regions of the same gene [e.g., [43]]. To test this prediction, we compared substitution patterns between two RRMs in Bru-3. Although, two regions of Bru-3 are homologous to the same RRM domain, and thus share the common ancestry, they are likely under different selection constrains. The C-terminal RRM of Bru-3 was constitutively spliced, whereas the N-terminal RRM was alternatively spliced in mRNA (Table 1). We used a conserved domain database [i.e., CDD at NCBI web page, [97]] to precisely localize two RRMs in Bru-3 and analyzed substitution patterns separately within both domains. The number of sites with substitutions, total number of substitutions, nucleotide diversity (¶) and its standard deviation (sd¶) were estimated in both domains among twelve Drosophila species [37] using DnaSP 4.20.2 [98]. In this context, we use ¶ as a measure of sequence diversity among the sampled species rather than within a species. We also calculated the ratios of synonymous/non-synonymous substitutions over the total number of synonymous/non-synonymous sites (KS and KA) for each exon between D. pseudoobscura and D. melanogaster.

5'- and 3'-end UTRs of Bru-3 in D. pseudoobscura and D. persimilis

To identify the transcribed but untranslated sequence of Bru-3, we used the FirstChoice® RNA ligase mediated-rapid amplification of cDNA ends (RLM-RACE) kit (Ambion) following the manufacturer's directions. All RACE reactions were performed on the total RNA extracted from adult females of the D. pseudoobscura Flagstaff 1993 strain. Bru-3Dmel-58F and PsBru3Pout are gene-specific primers used for 5'-end and 3'-end RACE, respectively (Additional file 9). The obtained sequence allowed us to design primers specific to both untranslated regions. Bru3Dmel(UTR)+760F and Bru3Dros(1)55R were used to amplify the 5' UTR in both lines of D. pseudoobscura, D. persimilis and D. melanogaster. To amplify the 3' UTR from both lines of D. pseudoobscura and D. persimilis, we used PCR with Bru3end-1F and Bru3UTR-2R primers (Additional file 9). PCR was conducted as above with Q solution (Qiagen). All amplicons were digested with Exo I and SAP (Fermentas) and sequenced directly, without intermediate cloning.

Additional files

DNA sequences are deposited in GenBank [GenBank: GQ497346-GQ497445].

References

  1. Sharp PA: Split genes and RNA splicing. Cell. 1994, 77 (6): 805-815. 10.1016/0092-8674(94)90130-9.

    Article  CAS  PubMed  Google Scholar 

  2. Black DL: Protein diversity from alternative splicing: a challenge for bioinformatics and post-genome biology. Cell. 2000, 103 (3): 367-370. 10.1016/S0092-8674(00)00128-8.

    Article  CAS  PubMed  Google Scholar 

  3. Kan Z, Rouchka EC, Gish WR, States DJ: Gene structure prediction and alternative splicing analysis using genomically aligned ESTs. Genome Res. 2001, 11 (5): 889-900. 10.1101/gr.155001.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  4. Graveley BR: Alternative splicing: increasing diversity in the proteomic world. Trends Genet. 2001, 17 (2): 100-107. 10.1016/S0168-9525(00)02176-4.

    Article  CAS  PubMed  Google Scholar 

  5. Adams MD, Celniker SE, Holt RA, Evans CA, Gocayne JD, Amanatides PG, Scherer SE, Li PW, Hoskins RA, Galle RF, et al: The genome sequence of Drosophila melanogaster. Science. 2000, 287 (5461): 2185-2195. 10.1126/science.287.5461.2185.

    Article  PubMed  Google Scholar 

  6. Misra S, Crosby M, Mungall C, Matthews B, Campbell K, Hradecky P, Huang Y, Kaminker J, Millburn G, Prochnik S, et al: Annotation of the Drosophila melanogaster euchromatic genome: a systematic review. Genome Biol. 2002, 3 (12): 0081-0083. 10.1186/gb-2002-3-12-research0083.

    Article  Google Scholar 

  7. Schuler GD: Pieces of the puzzle: expressed sequence tags and the catalog of human genes. J Mol Med. 1997, 75 (10): 694-698. 10.1007/s001090050155.

    Article  CAS  PubMed  Google Scholar 

  8. Clamp M, Fry B, Kamal M, Xie X, Cuff J, Lin MF, Kellis M, Lindblad-Toh K, Lander ES: From the cover: distinguishing protein-coding and noncoding genes in the human genome. Proc Natl Acad Sci USA. 2007, 104 (49): 19428-19433. 10.1073/pnas.0709013104.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  9. Budagyan B, Loraine A: Gene length and alternative transcription in fruit fly. Computational Systems Bioinformatics Conference, 2004: 2004, Stanford, CA, USA: Proceedings. 2004, IEEE, 515-516.

    Google Scholar 

  10. Ast G: How did alternative splicing evolve?. Nat Rev Genet. 2004, 5 (10): 773-782. 10.1038/nrg1451.

    Article  CAS  PubMed  Google Scholar 

  11. McGuire AM, Pearson MD, Neafsey DE, Galagan JE: Cross-kingdom patterns of alternative splicing and splice recognition. Genome Biol. 2008, 9: R50-10.1186/gb-2008-9-3-r50.

    Article  PubMed Central  PubMed  Google Scholar 

  12. Fox-Walsh KL, Dou Y, Lam BJ, Hung S-P, Baldi PF, Hertel KJ: The architecture of pre-mRNAs affects mechanisms of splice-site pairing. Proc Natl Acad Sci USA. 2005, 102 (45): 16176-16181. 10.1073/pnas.0508489102.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  13. Graveley BR: Mutually exclusive splicing of the insect Dscam pre-mRNA directed by competing intronic RNA secondary structures. Cell. 2005, 123 (1): 65-73. 10.1016/j.cell.2005.07.028.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  14. Kim E, Magen A, Ast G: Different levels of alternative splicing among eukaryotes. Nucl Acids Res. 2007, 35 (1): 125-131. 10.1093/nar/gkl924.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  15. Roy M, Kim N, Xing Y, Lee C: The effect of intron length on exon creation ratios during the evolution of mammalian genomes. RNA. 2008, 14: 1-13. 10.1261/rna.1024908.

    Article  Google Scholar 

  16. Hatton AR, Subramaniam V, Lopez AJ: Generation of alternative Ultrabithorax isoforms and stepwise removal of a large intron by resplicing at exon-exon junctions. Mol Cell. 1998, 2 (6): 787-796. 10.1016/S1097-2765(00)80293-2.

    Article  CAS  PubMed  Google Scholar 

  17. Burnette JM, Miyamoto-Sato E, Schaub MA, Conklin J, Lopez AJ: Subdivision of large introns in Drosophila by recursive splicing at nonexonic elements. Genetics. 2005, 170 (2): 661-674. 10.1534/genetics.104.039701.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  18. International Human Genome Sequencing Consortium: Initial sequencing and analysis of the human genome. Nature. 2001, 409 (6822): 860-921. 10.1038/35057062.

    Article  Google Scholar 

  19. Wilkinson MF, Shyu A-B: Multifunctional regulatory proteins that control gene expression in both the nucleus and the cytoplasm. BioEssays. 2001, 23 (9): 775-787. 10.1002/bies.1113.

    Article  CAS  PubMed  Google Scholar 

  20. Delaunay J, Mee GL, Ezzeddine N, Labesse G, Terzian C, Capri M, Ait-Ahmed O: The Drosophila Bruno paralogue Bru-3 specifically binds the EDEN translational repression element. Nucl Acids Res. 2004, 32 (10): 3070-3082. 10.1093/nar/gkh627.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  21. Kenan DJ, Query CC, Keene JD: RNA recognition: towards identifying determinants of specificity. Trends Biol Sci. 1991, 16: 214-220. 10.1016/0968-0004(91)90088-D.

    Article  CAS  Google Scholar 

  22. Good PJ, Chen Q, Warner SJ, Herring DC: A family of human RNA-binding proteins related to the Drosophila Bruno translational regulator. J Biol Chem. 2000, 275 (37): 28583-28592. 10.1074/jbc.M003083200.

    Article  CAS  PubMed  Google Scholar 

  23. Philips AV, Timchenko LT, Cooper TA: Disruption of splicing regulated by a CUG-binding protein in myotonic dystrophy. Science. 1998, 280 (5364): 737-741. 10.1126/science.280.5364.737.

    Article  CAS  PubMed  Google Scholar 

  24. Ladd AN, Charlet-B N, Cooper TA: The CELF family of RNA binding proteins Is implicated in cell-specific and developmentally regulated aternative splicing. Mol Cell Evol. 2001, 21 (4): 1285-1296.

    CAS  Google Scholar 

  25. Monroe D: Genomic Clues to DNA Treasure Sometimes Lead Nowhere. Science. 2009, 325 (5937): 142-143. 10.1126/science.325_142.

    Article  CAS  PubMed  Google Scholar 

  26. Tamura K, Subramanian S, Kumar S: Temporal patterns of fruit fly (Drosophila) evolution revealed by mutation clocks. Mol Biol Evol. 2004, 21 (1): 36-44. 10.1093/molbev/msg236.

    Article  CAS  PubMed  Google Scholar 

  27. Glazov EA, Pheasant M, McGraw EA, Bejerano G, Mattick JS: Ultraconserved elements in insect genomes: a highly conserved intronic sequence implicated in the control of homothorax mRNA splicing. Genome Res. 2005, 15 (6): 800-808. 10.1101/gr.3545105.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  28. Adams MD, Kelley JM, Gocayne JD, Dubnick M, Polymeropoulos MH, Xiao H, Merril CR, Wu A, Olde B, Moreno RF, et al: Complementary DNA sequencing: expressed sequence tags and human genome project. Science. 1991, 252 (5013): 1651-1656. 10.1126/science.2047873.

    Article  CAS  PubMed  Google Scholar 

  29. Johnson JM, Castle J, Garrett-Engele P, Kan Z, Loerch PM, Armour CD, Santos R, Schadt EE, Stoughton R, Shoemaker DD: Genome-wide survey of human alternative pre-mRNA splicing with exon junction microarrays. Science. 2003, 302 (5653): 2141-2144. 10.1126/science.1090100.

    Article  CAS  PubMed  Google Scholar 

  30. Mortazavi A, Williams BA, McCue K, Schaeffer L, Wold B: Mapping and quantifying mammalian transcriptomes by RNA-Seq. Nature Methods. 2008, 5: 621-628. 10.1038/nmeth.1226.

    Article  CAS  PubMed  Google Scholar 

  31. Modrek B, Lee C: A genomic view of alternative splicing. Nat Genet. 2002, 30: 13-19. 10.1038/ng0102-13.

    Article  CAS  PubMed  Google Scholar 

  32. Itoh H, Washio T, Tomita M: Computational comparative analyses of alternative splicing regulation using full-length cDNA of various eukaryotes. RNA. 2004, 10 (7): 1005-1018. 10.1261/rna.5221604.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  33. Pan Q, Shai O, Misquitta C, Zhang W, Saltzman AL, Mohammad N, Babak T, Siu H, Hughes TR, Morris QD, et al: Revealing global regulatory features of mammalian alternative splicing using a quantitative microarray platform. Mol Cell. 2004, 16 (6): 929-941. 10.1016/j.molcel.2004.12.004.

    Article  CAS  PubMed  Google Scholar 

  34. Lee C, Roy M: Analysis of alternative splicing with microarrays: successes and challenges. Genome Biol. 2004, 5 (7): 231-10.1186/gb-2004-5-7-231.

    Article  PubMed Central  PubMed  Google Scholar 

  35. Graveley BR: Molecular biology: power sequencing. Nature. 2008, 453 (7199): 1197-1198. 10.1038/4531197b.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  36. Sorek R, Shamir R, Ast G: How prevalent is functional alternative splicing in the human genome?. Trends Genet. 2004, 20 (2): 68-71. 10.1016/j.tig.2003.12.004.

    Article  CAS  PubMed  Google Scholar 

  37. Clark AG, Eisen MB, Smith DR, Bergman CM, Oliver B, Markow TA, Kaufman TC, Kellis M, Gelbart WM: Evolution of genes and genomes on the Drosophila phylogeny. Nature. 2007, 450 (7167): 203-218. 10.1038/nature06341.

    Article  PubMed  Google Scholar 

  38. Patel AA, Steitz JA: Splicing double: insights from the second spliceosome. Nat Rev Mol Cell Biol. 2003, 4: 960-970. 10.1038/nrm1259.

    Article  CAS  PubMed  Google Scholar 

  39. O'Grady PM: Reevaluation of phylogeny in the Drosophila obscura species group based on combined analysis of nucleotide sequences. Mol Phylog Evol. 1999, 12 (2): 124-139. 10.1006/mpev.1998.0598.

    Article  Google Scholar 

  40. Prud'homme B, Gompel N, Rokas A, Kassner VA, Williams TM, Yeh S-D, True JR, Carroll SB: Repeated morphological evolution through cis-regulatory changes in a pleiotropic gene. Nature. 2006, 440 (7087): 1050-1053. 10.1038/nature04597.

    Article  PubMed  Google Scholar 

  41. Guo M, Lo PC, Mount SM: Species-specific signals for the splicing of a short Drosophila intron in vitro. Mol Cell Evol. 1993, 13 (2): 1104-1118.

    CAS  Google Scholar 

  42. Talerico M, Berget SM: Intron definition in splicing of small Drosophila introns. Mol Cell Evol. 1994, 14 (5): 3434-3445.

    CAS  Google Scholar 

  43. Xing Y, Lee C: Alternative splicing and RNA selection pressure -- evolutionary consequences for eukaryotic genomes. Nat Rev Genet. 2006, 7 (7): 499-509. 10.1038/nrg1896.

    Article  CAS  PubMed  Google Scholar 

  44. Stolc V, Gauhar Z, Mason C, Halasz G, van Batenburg MF, Rifkin SA, Hua S, Herreman T, Tongprasit W, Barbano PE, et al: A Gene expression map for the euchromatic genome of Drosophila melanogaster. Science. 2004, 306 (5696): 655-660. 10.1126/science.1101312.

    Article  CAS  PubMed  Google Scholar 

  45. Royce TE, Rozowsky JS, Bertone P, Samanta M, Stolc V, Weissman S, Snyder M, Gerstein M: Issues in the analysis of oligonucleotide tiling microarrays for transcript mapping. Trends Genet. 2005, 21 (8): 466-475. 10.1016/j.tig.2005.06.007.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  46. Kopelman NM, Lancet D, Yanai I: Alternative splicing and gene duplication are inversely correlated evolutionary mechanisms. Nat Genet. 2005, 37 (6): 588-589. 10.1038/ng1575.

    Article  CAS  PubMed  Google Scholar 

  47. Modrek B, Lee CJ: Alternative splicing in the human, mouse and rat genomes is associated with an increased frequency of exon creation and/or loss. Nat Genet. 2003, 34 (2): 177-180. 10.1038/ng1159.

    Article  CAS  PubMed  Google Scholar 

  48. Xing Y, Lee C: Evidence of functional selection pressure for alternative splicing events that accelerate evolution of protein subsequences. Proc Natl Acad Sci USA. 2005, 102 (38): 13526-13531. 10.1073/pnas.0501213102.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  49. Baek D, Green P: Sequence conservation, relative isoform frequencies, and nonsense-mediated decay in evolutionarily conserved alternative splicing. Proc Natl Acad Sci USA. 2005, 102 (36): 12813-12818. 10.1073/pnas.0506139102.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  50. Chen F-C, Wang S-S, Chen C-J, Li W-H, Chuang T-J: Alternatively and constitutively spliced exons are subject to different evolutionary forces. Mol Biol Evol. 2006, 23 (3): 675-682. 10.1093/molbev/msj081.

    Article  PubMed  Google Scholar 

  51. Sorek R, Ast G: Intronic sequences flanking alternatively spliced exons are conserved between human and mouse. Genome Res. 2003, 13 (7): 1631-1637. 10.1101/gr.1208803.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  52. Sorek R, Shemesh R, Cohen Y, Basechess O, Ast G, Shamir R: A non-EST-based method for exon-skipping prediction. Genome Res. 2004, 14 (8): 1617-1623. 10.1101/gr.2572604.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  53. Resch A, Xing Y, Alekseyenko A, Modrek B, Lee C: Evidence for a subpopulation of conserved alternative splicing events under selection pressure for protein reading frame preservation. Nucl Acids Res. 2004, 32 (4): 1261-1269. 10.1093/nar/gkh284.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  54. Berget SM: Exon Recognition in Vertebrate Splicing. J Biol Chem. 1995, 270 (6): 2411-2414.

    Article  CAS  PubMed  Google Scholar 

  55. Thummel CS: Mechanisms of transcriptional timing in Drosophila. Science. 1992, 255 (5040): 39-40. 10.1126/science.1553530.

    Article  CAS  PubMed  Google Scholar 

  56. Swinburne IA, Silver PA: Intron delays and transcriptional timing during development. Dev Cell. 2008, 14 (3): 324-330. 10.1016/j.devcel.2008.02.002.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  57. Neugebauer KM: On the importance of being co-transcriptional. J Cell Sci. 2002, 115 (20): 3865-3871. 10.1242/jcs.00073.

    Article  CAS  PubMed  Google Scholar 

  58. Proudfoot NJ: Dawdling polymerases allow introns time to splice. Nature Struct Biol. 2003, 10 (11): 876-878. 10.1038/nsb1103-876.

    Article  CAS  PubMed  Google Scholar 

  59. Haddrill P, Charlesworth B, Halligan D, Andolfatto P: Patterns of intron sequence evolution in Drosophila are dependent upon length and GC content. Genome Biol. 2005, 6 (8): R67-10.1186/gb-2005-6-8-r67.

    Article  PubMed Central  PubMed  Google Scholar 

  60. Marais G, Nouvellet P, Keightley PD, Charlesworth B: Intron size and exon evolution in Drosophila. Genetics. 2005, 170 (1): 481-485. 10.1534/genetics.104.037333.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  61. Sorek R: The birth of new exons: mechanisms and evolutionary consequences. RNA. 2007, 13 (10): 1603-1608. 10.1261/rna.682507.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  62. Lev-Maor G, Sorek R, Shomron N, Ast G: The birth of an alternatively spliced exon: 3' splice-site selection in Alu exons. Science. 2003, 300 (5623): 1288-1291. 10.1126/science.1082588.

    Article  CAS  PubMed  Google Scholar 

  63. Wang W, Zheng H, Yang S, Yu H, Li J, Jiang H, Su J, Yang L, Zhang J, McDermott J, et al: Origin and evolution of new exons in rodents. Genome Res. 2005, 15 (9): 1258-1264. 10.1101/gr.3929705.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  64. Alekseyenko AV, Kim N, Lee CJ: Global analysis of exon creation versus loss and the role of alternative splicing in 17 vertebrate genomes. RNA. 2007, 13 (5): 661-670. 10.1261/rna.325107.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  65. Nurminsky DI, Nurminskaya MV, Aguiar DD, Hartl DL: Selective sweep of a newly evolved sperm-specific gene in Drosophila. Nature. 1998, 396 (6711): 572-575. 10.1038/25126.

    Article  CAS  PubMed  Google Scholar 

  66. Chen S-T, Cheng H-C, Barbash DA, Yang H-P: Evolution of hydra, a recently evolved testis-expressed gene with nine alternative first exons in Drosophila melanogaster. PLoS Genetics. 2007, 3 (7): e107-10.1371/journal.pgen.0030107.

    Article  PubMed Central  PubMed  Google Scholar 

  67. Betran E, Wang W, Jin L, Long M: Evolution of the Phosphoglycerate mutase processed gene in human and chimpanzee revealing the origin of a new primate gene. Mol Biol Evol. 2002, 19 (5): 654-663.

    Article  CAS  PubMed  Google Scholar 

  68. Sorek R, Lev-Maor G, Reznik M, Dagan T, Belinky F, Graur D, Ast G: Minimal conditions for exonization of intronic sequences: 5' splice site formation in Alu exons. Mol Cell. 2004, 14 (2): 221-231. 10.1016/S1097-2765(04)00181-9.

    Article  CAS  PubMed  Google Scholar 

  69. Kondrashov FA, Koonin EV: Evolution of alternative splicing: deletions, insertions and origin of functional parts of proteins from intron sequences. Trends Genet. 2003, 19 (3): 115-119. 10.1016/S0168-9525(02)00029-X.

    Article  CAS  PubMed  Google Scholar 

  70. Levine MT, Jone CD, Kern AD, Lindfors HA, Begun DJ: Novel genes derived from noncoding DNA in Drosophila melanogaster are frequently X-linked and exhibit testis-biased expression. Proc Natl Acad Sci USA. 2006, 103 (26): 9935-9939. 10.1073/pnas.0509809103.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  71. Kim E, Goren A, Ast G: Alternative splicing: current perspectives. BioEssays. 2008, 30 (1): 38-47. 10.1002/bies.20692.

    Article  CAS  PubMed  Google Scholar 

  72. Carvalho AB, Clark AG: Genetic recombination: intron size and natural selection. Nature. 1999, 401 (6751): 344-344. 10.1038/43827.

    Article  CAS  PubMed  Google Scholar 

  73. Comeron JM, Kreitman M: The correlation between intron length and recombination in Drosophila: dynamic equilibrium between mutational and selective forces. Genetics. 2000, 156 (3): 1175-1190.

    PubMed Central  CAS  PubMed  Google Scholar 

  74. Fitzpatrick CL, Stevison LS, Noor MAF: Fine-scale crossover rate and interference along the XR-chromosome arm of Drosophila pseudoobscura. Drosophila Information Service. 2009,

    Google Scholar 

  75. Petrov DA, Lozovskaya ER, Hartl DL: High intrinsic rate of DNA loss in Drosophila. Nature. 1996, 384 (6607): 346-349. 10.1038/384346a0.

    Article  CAS  PubMed  Google Scholar 

  76. Petrov DA, Hartl DL: High rate of DNA loss in the Drosophila melanogaster and Drosophila virilis species groups. Mol Biol Evol. 1998, 15 (3): 293-302.

    Article  CAS  PubMed  Google Scholar 

  77. Hill WG, Robertson A: The effect of linkage on limits to artificial selection. Genet Res. 1966, 8: 269-294. 10.1017/S0016672300010156.

    Article  CAS  PubMed  Google Scholar 

  78. Felsenstein J: The evolutionary advantage of recombination. Genetics. 1974, 78 (2): 737-756.

    PubMed Central  CAS  PubMed  Google Scholar 

  79. Castillo-Davis CI, Mekhedov SL, Hartl DL, Koonin EV, Kondrashov FA: Selection for short introns in highly expressed genes. Nat Genet. 2002, 31 (4): 415-418.

    CAS  PubMed  Google Scholar 

  80. Vinogradov AE: Compactness of human housekeeping genes: selection for economy or genomic design?. Trends Genet. 2004, 20 (5): 248-253. 10.1016/j.tig.2004.03.006.

    Article  CAS  PubMed  Google Scholar 

  81. Lynch M: Intron evolution as a population-genetic process. Proc Natl Acad Sci USA. 2002, 92: 6118-6123. 10.1073/pnas.092595699.

    Article  Google Scholar 

  82. Babushok D, Ostertag E, Kazazian H: Current topics in genome evolution: molecular mechanisms of new gene formation. Cell Mol Life Sci. 2007, 64 (5): 542-554. 10.1007/s00018-006-6453-4.

    Article  CAS  PubMed  Google Scholar 

  83. Wagner E, Lykke-Andersen J: mRNA surveillance: the perfect persist. J Cell Sci. 2002, 115 (15): 3033-3038.

    CAS  PubMed  Google Scholar 

  84. Gatfield D, Unterholzner L, Ciccarelli FD, Bork P, Izaurralde E: Nonsense-mediated mRNA decay in Drosophila: at the intersection of the yeast and mammalian pathways. EMBO J. 2003, 22: 3960-3970. 10.1093/emboj/cdg371.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  85. Lareau LF, Inada M, Green RE, Wengrod JC, Brenner SE: Unproductive splicing of SR genes associated with highly conserved and ultraconserved DNA elements. Nature. 2007, 446 (7138): 926-929. 10.1038/nature05676.

    Article  CAS  PubMed  Google Scholar 

  86. Nekrutenko A: Identification of novel exons from rat-mouse comparisons. J Mol Evol. 2004, 59 (5): 703-708. 10.1007/s00239-004-2653-4.

    Article  CAS  PubMed  Google Scholar 

  87. Noor MA: Speciation driven by natural selection in Drosophila. Nature. 1995, 375: 674-675. 10.1038/375674a0.

    Article  CAS  PubMed  Google Scholar 

  88. Ortíz-Barrientos D, Counterman BA, Noor MAF: The genetics of speciation by reinforcement. PLoS Biol. 2004, 2 (12): 2256-2263. 10.1371/journal.pbio.0020416.

    Article  Google Scholar 

  89. Chang AS, Noor MAF: The genetics of hybrid male sterility between the allopatric species pair Drosophila persimilis and D. pseudoobscura bogotana: dominant sterility alleles in collinear autosomal regions. Genetics. 2007, 176 (1): 343-349. 10.1534/genetics.106.067314.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  90. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ: Basic local alignment search tool. J Mol Biol. 1990, 215 (3): 403-410.

    Article  CAS  PubMed  Google Scholar 

  91. Richards S, Liu Y, Bettencourt BR, Hradecky P, Letovsky S, Nielsen R, Thornton K, Hubisz MJ, Chen R, Meisel RP, et al: Comparative genome sequencing of Drosophila pseudoobscura: chromosomal, gene, and cis-element evolution. Genome Res. 2005, 15 (1): 1-18. 10.1101/gr.3059305.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  92. Gilbert DG: DroSpeGe: rapid access database for new Drosophila species genomes. Nucl Acids Res. 2007, 35 (suppl_1): D480-485. 10.1093/nar/gkl997.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  93. Rambaut A: Se-Al v2.0a11. 2002, Department of Zoology, University of Oxford, Oxford, UK

    Google Scholar 

  94. Karlin S, Altschul SF: Methods for assessing the statistical significance of molecular sequence features by using general scoring schemes. Proc Natl Acad Sci USA. 1990, 87 (6): 2264-2268. 10.1073/pnas.87.6.2264.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  95. 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 

  96. Labrador M, Corces VG: Extensive Exon Reshuffling Over Evolutionary Time Coupled to Trans-Splicing in Drosophila. Genome Res. 2003, 13 (10): 2220-2228. 10.1101/gr.1440703.

    Article  PubMed Central  CAS  PubMed  Google Scholar 

  97. Marchler-Bauer A, Anderson JB, Cherukuri PF, DeWeese-Scott C, Geer LY, Gwadz M, He S, Hurwitz DI, Jackson JD, Ke Z, et al: CDD: a Conserved Domain Database for protein classification. Nucl Acids Res. 2005, 33 (suppl_1): D192-196.

    PubMed Central  CAS  PubMed  Google Scholar 

  98. Rozas J, Sanchez-DelBarrio JC, Messeguer X, Rozas R: DnaSP, DNA polymorphism analyses by the coalescent and other methods. Bioinformatics. 2003, 19 (18): 2496-2497. 10.1093/bioinformatics/btg359.

    Article  CAS  PubMed  Google Scholar 

  99. Swofford DL: PAUP*. Phylogenetic analysis using parsimony (*and other methods). 2000, Sunderland, MA, U.S.A.: Sinauer Associates, Inc, 4:

    Google Scholar 

Download references

Acknowledgements

We thank K. Dusinberre, our associate editor and two anonymous reviewers for comments on the manuscript and helpful discussion. D. pseudoobscura bogotana, D. melanogaster and D. mojavensis strains were contributed by A. Chang. N.P.K. was been supported in part by a postdoctoral fellowship in Evolutionary Genomics and Molecular Evolution from Duke University and the Duke Institute for Genome Sciences and Policy. M.A.F.N. is supported by NIH grant GM076051 and NSF grants 0509780 and 0715484.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Nikolai P Kandul.

Additional information

Authors' contributions

NPK conceived the study, performed the bench-work and analysis, and wrote the manuscript. MAFN supervised the analysis, wrote the script for intron analysis, and contributed to writing the manuscript. Both authors approved the final manuscript.

Electronic supplementary material

12863_2009_725_MOESM1_ESM.XLS

Additional file 1: The genomic position of each exon of Bru-3 in the twelve sequenced Drosophila species [37]at Drosophila Species Genomes BLAST [DroSpeGe, [92]]. The table lists positions of Bru-3 exons in genomes of the twelve sequenced Drosophila species. (XLS 45 KB)

12863_2009_725_MOESM2_ESM.PDF

Additional file 2: Diversity of Bru-3 ASTs identified in each of four Drosophila species. The table lists ASTs identified in each of four tested Drosophila species. (PDF 97 KB)

12863_2009_725_MOESM3_ESM.PDF

Additional file 3: cDNA screens for Bru-3 transcripts containing exon 6. The RT-PCR gel image shows that exons 5 and 6, and exons 5, 6 and 7 are included consecutively in mature mRNA of Bru-3 in both lines of D. pseudoobscura and D. persimilis. (PDF 187 KB)

12863_2009_725_MOESM4_ESM.PDF

Additional file 4: The alignment of the genomic sequences homologous by position to exon 6 in D. pseudoobscura and D. persimilis. The color-coded alignment of genomic sequences shows the local deletions/insertions and/or substitutions at the 5' SS flaking exon 6 likely led to the loss of coding potential and subsequent decay of exon 6 in the melanogaster lineage: D. anannassae, D. erecta, D. melanogaster and D. simulans. (PDF 94 KB)

12863_2009_725_MOESM5_ESM.PDF

Additional file 5: The alignment of the genomic sequences homologous by position to exon 8 in D. pseudoobscura and D. persimilis. The color-coded alignment of genomic sequences shows that, while all Drosophila species shared both 3' and 5' SS flanking exon 8, the region corresponding to exon 8 did not encode a polypeptide chain in any tested species except D. pseudoobscura and D. persimilis. (PDF 167 KB)

12863_2009_725_MOESM6_ESM.PDF

Additional file 6: The alignment of the genomic sequences encompassing exon 8 from species of the obscura group. The color-coded alignment of genomic sequences shows that exon 8 is conserved at the sequence level and can be translated in all tested species of the obscura group. (PDF 145 KB)

12863_2009_725_MOESM7_ESM.PDF

Additional file 7: The alignment of intron 9 for twelve Drosophila species [37]. The color-coded alignment of genomic sequences shows that the species of the melanogaster group, D. simulans, D. sechelia, D. erecta, D. yakuba, D. melanogaster and D. ananassae, evolved an alternative 3' SS of intron 9. (PDF 133 KB)

12863_2009_725_MOESM8_ESM.PDF

Additional file 8: Intron sizes of Bru-3 in the twelve sequenced Drosophila species [37]at Drosophila Species Genomes BLAST [DroSpeGe, [92]]. The table lists Bru-3 intron lengths for the twelve sequenced Drosophila species. (PDF 117 KB)

12863_2009_725_MOESM9_ESM.PDF

Additional file 9: Primer applications and sequences. The table lists the sequences of primers used in this study. (PDF 44 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

Kandul, N.P., Noor, M.A. Large introns in relation to alternative splicing and gene evolution: a case study of Drosophila bruno-3. BMC Genet 10, 67 (2009). https://doi.org/10.1186/1471-2156-10-67

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/1471-2156-10-67

Keywords