Next Article in Journal
Host-Specific Diversity of Culturable Bacteria in the Gut Systems of Fungus-Growing Termites and Their Potential Functions towards Lignocellulose Bioconversion
Previous Article in Journal
The Dynamic Distribution of Wolbachia and Rickettsia in AsiaII1 Bemisia tabaci
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Evolution of piggyBac Transposons in Apoidea

College of Animal Science and Technology, Yangzhou University, Yangzhou 225009, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Insects 2023, 14(4), 402; https://doi.org/10.3390/insects14040402
Submission received: 11 March 2023 / Revised: 15 April 2023 / Accepted: 19 April 2023 / Published: 21 April 2023
(This article belongs to the Section Insect Societies and Sociality)

Abstract

:

Simple Summary

Bees are one of the mega-diverse class of insects in Hymenoptera. As predators and the main crops pollinator, bees play an essential role in improving crop yield and the quality whiles providing substantial economic importance. Transposable elements (TEs) exist widely in nature and they make up a significant portion of prokaryotic and eukaryotic genomes. piggyBac (PB) transposons, belonging to DNA transposons, have been identified in plants, fungi, and animals. The advantages of the modified PB transposon system or high transposable efficiency and broad host ranges have led to their widespread application in transgenesis, mutagenesis, and gene therapy. At present, there is no systematic report on PB transposon in bees. Therefore, this study focused on annotating the transposons of the PB superfamily in bees to reveal the distribution, diversity, structural characteristics, evolutionary pattern and activity of PB transposons in the available genomes of bees.

Abstract

In this study, we investigated the presence of piggyBac (PB) transposons in 44 bee genomes from the Apoidea order, which is a superfamily within the Hymenoptera, which includes a large number of bee species crucial for pollination. We annotated the PB transposons in these 44 bee genomes and examined their evolution profiles, including structural characteristics, distribution, diversity, activity, and abundance. The mined PB transposons were divided into three clades, with uneven distribution in each genus of PB transposons in Apoidea. The complete PB transposons we discovered are around 2.23–3.52 kb in length and encode transposases of approximately 580 aa, with terminal inverted repeats (TIRs) of about 14 bp and 4 bp (TTAA) target-site duplications. Long TIRs (200 bp, 201 bp, and 493 bp) were also detected in some species of bees. The DDD domains of the three transposon types were more conserved, while the other protein domains were less conserved. Generally, most PB transposons showed low abundance in the genomes of Apoidea. Divergent evolution dynamics of PB were observed in the genomes of Apoidea. PB transposons in some identified species were relatively young, whiles others were older and with some either active or inactive. In addition, multiple invasions of PB were also detected in some genomes of Apoidea. Our findings highlight the contribution of PB transposons to genomic variation in these species and suggest their potential as candidates for future gene transfer tools.

1. Introduction

The Apoidea superfamily, which originated from the early-to-middle Cretaceous period, is one of the main categories of Hymenoptera Apocrita [1]. Apoidea has various morphological characteristics as a large group of holometabolous insects [2,3,4]. Bees are known for their short generation cycle, strong reproduction, parthenogenesis, and captive-bred. These insects have been extensively studied and can be used as model animals for studying the evolution of sociality, intra- and interspecific communication, and physiological adaptations [2,5,6,7,8,9,10]. Bees exhibit three different ways of life: sociality, solitary, and parasitism, with the majority being solitary [11,12]. Most bees are important pollinators of angiosperms and play crucial roles in ecosystem functions [3,4]. Apoidea insects directly impact 6% of the 128 most important crops consumed worldwide [13]. The Apoidea superfamily comprises seven widely recognized families: Megachilidae, Apidae, Andrenidae, Colletidae, Halictidae, Melittidae, and Stenotritidae [1,12,14].
Transposable elements (TEs) are DNA sequences that can be independently copied or broken from their original site and inserted into another site, playing a regulatory role in the evolution and gene regulation of all organisms. They ‘jump’ from one position in the genome to another through a series of processes, such as cutting and reintegrating, also called the jumping gene. The TE generally moves the TE sequence itself and is not dependent on the sequence relationship between the donor and recipient. TEs have been detected in various organisms, including bacteria, fungi, and insects [15], and they account for 15–47% of insect genomes, 35–69% of mammalian genomes, and about 90% of some plant genomes [16,17,18,19,20,21,22,23,24,25]. TEs also play important roles in shaping genome evolution and transcriptome regulation [15,26,27,28,29].
Tes are divided into two categories based on their transposable mechanisms. Class I or retrotransposons rely on RNA intermediates and transpose via a ‘copy-and-paste’ mechanism, whereas Class II or DNA transposons use DNA intermediates and usually transpose by ‘cut-and-paste’ mechanisms [30,31,32]. Several diverse DNA transposon superfamilies, such as piggyBac, Tc1/mariner, pogo, hAT, Helitron, and PIF-Harbinger, have been reported, and their evolutionary profiles have been well-defined for some groups, such as Tc1/mariner and pogo DNA transposons [12].
The active piggyBac (PB) transposon was initially discovered in the Lepidopteran cabbage looper moth (Trichoplusia ni), and subsequent genome sequencing and comparison across multiple species revealed its high diversity and widespread distribution [33,34]. PB elements have been identified in various eukaryote kingdoms, including fungi, plants, and animals (insects, crustaceans, amphibians, fishes, and mammals) [35,36]. The active PB transposon from T. ni has been developed into an efficient gene transfer vector [35] with potential applications in human gene therapy [37,38,39,40]. The PB transposon from T. ni is 2472 bp in length and contains TTAA target sites, with 14 bp terminal inverted repeats (TIR) and 19 bp sub-terminal asymmetric inverted repeats (STIR) located at 3 and 31 bp from the 5′ and 3′ TIRs, respectively [33,36,41,42]. It has an open reading frame (ORF) in the middle, which is 1782 bp long, encoding a protein of 594 amino acids with a molecular weight of 64 kDa [36,43]. PB transposases belong to DD [E/D]-transposases [44] and are currently defined using five domains, namely the N-terminal domain (NTD), the Dimerization and DNA-binding domain 1 (DDBD1), the catalytic domain (also named DDE/DDD) and the DDBD2 and C-terminal cysteine-rich domain (CRD) [45]. In addition, there is little obvious sequence similarity between the PB superfamily and other transposon superfamilies [35,44]. The majority of mammalian genomes, such as the human, mouse, rat, and dog genomes, contain decayed piggyBac transposons [21,36,46,47,48]. However, two families of piggyBac-like elements were found in Myotis lucifugus, and display recent activity, forming a strong cluster with the sea squirt, Ciona intestinalis [49]. Another study identified PB elements in Microcebus murinus, divided them into three categories, and found that those three types of PB transposons were relatively young [50]. In addition, 32 piggyBac-like elements in the red flour beetle Tribolium castaneumre are divided into 14 diverse groups [51]. Moreover, at least five domesticated piggyBac (PGBD) have been found in the human genome [36]. These findings highlight the wide distribution and high diversity of PB transposons, although their internal classification at the family level remains largely unknown.
Currently, there is limited information on DNA transposons in bees, with only pogo and mariner, belonging to the ITm group (IS630-Tc1-Mariner), being well-defined in bees [12,52]. In this paper, we investigated the genomes of 44 species of Apoidea and annotated the PB transposons in each species to determine their structural characteristics, overall distribution, and phylogenetic location and classification, and conducted evolutionary dynamics analysis. Our data shed light on the evolutionary pattern of PB transposons in Apoidea and enhanced our understanding of their contribution to the evolution of the Apoidea genome.

2. Materials and Methods

2.1. piggyBac Mining in Apoidea

All reference PB transposase sequences [36] were used to search the genomes of Apoidea in the WGS (whole-genome shotgun contigs) database of NCBI (https://www.ncbi.nlm.nih.gov, accessed on 22 November 2022) using TBLASTN (v. 2.12.0) [53] with the value of 1e-100. The best hits of TBLASTN [53] were extended to 2000 bp flanks, which were then BLASTN against the genome. The hits were downloaded and aligned to determine the boundaries of these transposons in each genome. The boundaries of the transposons (TIRs and TSDs) were manually checked, combined with the help of FastPCR (https://primerdigital.com/fastpcr.html, accessed on 25 November 2022) [54] and the Bioedit program (v.7.2.0) [55]. The representative sequences or consensus sequences (derived for PB > 10 copies in genome) of transposons in each genome were used to determine the genomic copies (coverage > 40% and similarity > 80% of queries) by BLAST.

2.2. Sequence Analysis and Phylogenetic Analysis

The transposases of the mined transposons were predicted by Genescan (https://www.bing.com/search?q=genescan, accessed on 29 November 2022), and the sequence identities were calculated using the BioEdit tool [55]. The sequence identity matrix was drawn with the HeatMap in the TBtools (v. 1.0987663) (https://github.com/CJ-Chen/TBtools/releases, accessed on 30 January 2023) [56]. The PB transposases were aligned using MAFFT [57] and manually edited by Bioedit [52]. The structure predicted by the transposase was presented by the tool Illustrator for Biological Sequences (IBS v. 1.0.3) [58]. The online website Welogo v.3.7.12 (http://weblogo.threeplusone.com/create.cgi, accessed on 16 December 2022) was used to create logo images of TIRs.
Overall, 13 PB transposase reference sequences collected from the previous report [36] and 10 IS1380 transposases downloaded from ISfinder (https://www-is.biotoul.fr/scripts/search-db.php, accessed on 23 January 2023) [59], which are distantly related transposases from prokaryotes [60], combined with the mined PB transposases in Apoidea, together submitted for phylogenetic analysis. A multiple sequence alignment was carried out by using the G-INS-I method of MAFFT (v7.310) [57]; then, the maximum likelihood method was used to construct the phylogenetic tree using IQ-TREE (v. 1.6.1), with an ultrafast bootstrap value of 1000. According to the Bayesian information criterion, the most suitable amino acid substitution model for these data was the VT+F+G4 model that was optimally matched according to BIC selected by ModelFinder embedded with the IQ-TREE program [61,62].

2.3. Evolutionary Dynamics Analysis

The PB transposon in each genome was annotated using the RepeatMasker program (http://www.repeatmasker.org/RMDownload.html, accessed on 20 December 2022). Then, the Kimura (K) divergence was calculated using the calcDivergenceFromAlign.pl package from RepeatMasker [63]. The Kimura (K) divergence reflects the activity of transposons on the relative time scale of each genome [64]. It is generally believed that young transposons have lower Kimura (K) divergence [65].

3. Results

3.1. Distribution of piggyBac Transposons in Apoidea

PB transposons showed an uneven distribution across the genus of Apoidea, as summarized in Figure 1 and Supplementary Table S1. Out of the 127 species examined, we detected a total of 44 PB elements in 44 species (34.65%) belonging to 27 genera within eight families of Apoidea. The majority of PB transposons were found in Apidae, with 26 species (38.81%) of Apidae containing PB transposons, while other families had fewer detections, such as three species (13.04%) in Halictidae, four species (66.67%) in Crabronidae, four species (57.14%) in Andrenidae, and four species (50%) in Megachilidae. PB transposons were only detected in one species in each of the following three families of bees, Colletidae, Melittidae, and Ampulicidae (Figure 1 and Supplementary Table S1).
Among the 27 genera, PB transposons were found in 12 species (36.36%) of Bombus, four species (57.14%) of Andrena, two species (66.67%) of Ceratina, two species (66.67%) of Megachile, and two species (40%) of Tetragonula. In some lineages, PB transposons were detected in only one species (Figure 1 and Supplementary Table S1).

3.2. Low Abundance of piggyBac Transposons in Apoidea

The copy number of PB in each genome was investigated by BLAST as described in the methods. Our data revealed significant variation in the copy numbers of PB transposons among different species, ranging from 1 (e.g., Ceratina calcata, Seladonia tumulorum) to 335 (Eufriesea mexicana) in the 44 species harboring PB elements (Table 1 and Supplementary Table S1). Generally, PB transposons were found in low copy numbers in most host genomes of Apoidea, with 28 host genomes containing less than 10 copies of PB. Only four species have more than 50 copies of PB in their genome (e.g., Bombus soroeensis, Bombus terrestris, Andrena haemorrhoa, and E. mexicana). Among these, very high copy numbers were identified in the genomes of A. haemorrhoa (236) and E. mexicana (335). Additionally, most PB transposons existed as truncated copies, with only 20 species harboring intact PB transposons (encoding transposase longer than 500 amino acids and detectable two-end TIRs), which accounted for 45.45% of the species containing PB copies. Although some species had high full copy numbers of PB, the intact copies were very low or not detectable. For example, B. terrestris contained 86 copies of PB, but only two of them were intact copies, while A. haemorrhoa, B. soroeensis, and E. mexicana had totals of 236, 58 and 335 in copy numbers, respectively, but intact copies were not detected in these genomes. In general, PB transposons were not significantly amplified in bees, and most species displayed a low abundance of PB (Table 1 and Supplementary Table S1).

3.3. Structural Organization of piggyBac Transposons in Apoidea

The total length of PB transposons ranges from 1.37 kb to 3.52 kb, and they contain a single ORF encoding for the transposase, flanked by TIRs and TSDs (Figure 2). Some species, such as Bombus bifarius, Ceratina calcarata, Frieseomelitta varia, and Andrena dorsata, have longer PB transposons (>3 kb) (Table 2 and Supplementary Table S1). Among them, A. dorsata has the longest transposon length (3518 bp) and encodes the longest transposase (681 aa). Most PB transposons carry short TIRs (<25 bp); however, few PB transposons harbor long TIRs, such as PB in B. bifarius (TIR: 201 bp), B. terrestris (TIR: 200 bp) and Exoneura robusta (TIR: 489 bp). The conserved and consistent TSD (TTAA) was observed in most PB transposons (Table 2 and Supplementary Table S1). The sequence logo of TIRs revealed two conserved motifs (CACTA and TACCG) in the TIRs, with the CACTA motif being the most highly conserved at the TIR end across these PB elements (Figure 3). High similarity between the left and right TIRs of each element was observed, displaying 100% sequence identity in 21 species (Supplementary Table S1).

3.4. Transposase Domain Organization

The evolutionary tree of the excavated PB transposases (≥100 aa) was constructed using the IQ-tree program [61], with 24 reference sequences [36], and the IS1380 transposases set as an outgroup. The phylogenetic tree revealed three distinct clades (A, B, and C) of PB transposases in Apoidea with robust bootstrap supports (≥78%), as summarized in Figure 4 and Supplementary Figure S1.
PB transposases are known to contain five well-defined domains: the N-terminal domain (NTD), the Dimerization and DNA-binding domain 1 (DDBD1), the catalytic domain (also named DDE/DDD), DDBD2, and the C-terminal cysteine-rich domain (CRD) [45]. Sequence analysis showed that the catalytic domains are highly conserved (Figure 5) with high sequence identities within and between the clades (ranging from 40% to 76% and 26% to 37%, respectively) (Figure 6c), while other domains, including DDBD1, DDBD2, CRD, and NTD, are less conserved (Supplementary Figure S2A–D) and show low sequence identities within and between clades (Figure 6a,b,d,e), particularly the NTDs, which display very low sequence identities between clades (≤ 10%, Figure 6a) and are poorly conserved across clades (Figure 6a). The three key catalytic residues (DDD), crucial for catalyzing the transposition reaction [43,60,66], are highly conserved in the three clade transposases. The insertion motif, identified between the second and third key catalytic D residues [45,67], is conserved across the three clades, with slight variations observed for the C terminal of the insertion motifs of the clade C transposases (Figure 5). Additionally, CRD, required for PB activity and thought to be the driver of TIR binding [45], exhibits seven cysteines with regular spacing that are highly conserved in the CRD of the three clade transposases (Supplementary Figure S2D). Residues of P131-N152, E175-K190, and V207-D228 in DDBD1 are also conserved among the three clade transposases [68] (Supplementary Figure S2B). The N-terminal of DDBD2 contains a highly conserved tryptophan (W) known to play a central role in the transposase activity and DNA hairpin formation and resolution [43,44] (Supplementary Figure S2C). The DDBD and catalytic domain, with its insertion motif, are known to collaborate to synapse TIRs and direct the scissile phosphates to the active sites comprising D268, D346, and D447 [45].

3.5. Evolution Dynamics of piggyBac in the Apoidea Genomes

To thoroughly investigate the evolutionary dynamics of the PB family in the genomes of Apoidea, we used RepeatMasker [63] to calculate the Kimura divergence of PB transposons in the PB-detected genomes. Divergent evolution dynamics of PB were observed in these species (Figure 7 and Supplementary Figure S3). Most copies of PB transposons in the genomes of fifteen species (A. haemorrhoa, Cerceris rybyensis, Ectemnius continuus, Exoneurella tridentata, Holcopasites calliopsidis, Lasioglossum pauxillum, Macropis europaea, Seladonia tumulorum, A. dorsata, F. varia, Megachile rotundata, Osmia lignaria, Stelis phaeoptera, Dufourea novaeangliae, and Nysson spinosus) represent K divergences of less than 2%, and seem to have invaded very recently. Furthermore, intact copies were detected in some species (A. dorsata, E. continuus, E. tridentata, F. varia, M. europaea, M. rotundata, L. pauxillum, S. tumulorum, H. calliopsidis, and N. spinosus) (Figure 7). These data indicated that PB transposons are young invaders in these bee genomes and may still have transposition activity.
On the other hand, PB copies in some species (B. bifarius, Bombus hypnorum, Bombus haemorrhoidalis, C. calcarata, Ctenoplectra terminalis, E. robusta, Habropoda laboriosa, and Mimumesa dahlbomi) show high K divergences (K > 15), suggesting that they may be old invaders in these lineages (Supplementary Figure S3) and may have lost their jumping activities. Additionally, multiple amplification waves were observed in Bombus balteatus, Bombus huntii, B. bifarius, Bombus ignitus, C. calcarata, C. terminalis, Ampulex compressa, Andrena minutula, B. haemorrhoidalis, E. robusta, H. laboriosa, Nomada fabriciana, B. terrestris, and S. phaeoptera, suggesting that these species might have experienced multiple invasions of PB transposon (Supplementary Figure S3).

4. Discussion

4.1. Distribution, Diversity and Copy Number of piggyBac in Apoidea

PB transposons are widely distributed in various vertebrates (Actinopterygii, Primate, and Rodentia) and invertebrates (Nematoda, Mollusca, Arthropods, Cnidaria, Stramenopiles, and Platyhelminthes) [17,35,36,69,70]. In our study, we found that PB transposons were detected in approximately one-third of bee species (34.65%, 44/127), and they were unevenly distributed across families and genera of bees. Notably, due to the high species diversity of Bombus, the number of PB transposons detected in Bombus was significantly higher than in other genera. Moreover, the transposons in Bombus were relatively conservative, with almost all PB transposons found in Bombus belonging to Clade B, suggesting a possible vertical propagation of PB transposons in Bombus. According to the available data, PB transposons have great diversity in arthropod species, and besides the PB family, Tc1/mariner, Busters and other transposons are found mostly in arthropods [71,72,73]. Here, we found that PB transposons also showed high diversity in Apoidea. Three distinct clades of PB transposases (A–C) were identified in Apoidea, with low sequence similarity among them. Megachile, Bombus, Ceratina, and Andrena all harbored two types of PB transposons. When determining the distribution, we considered PB transposons to be present in the genome when sequence similarity reached 80% and query coverage reached 40%. However, this approach may have led to the exclusion of short, truncated elements derived from PB, potentially underestimating the taxonomic distribution of PB transposons in Apoidea. In addition, due to the continuous update of genome sequencing data, new families are constantly appearing in the PB superfamily, meaning the number of PB transposon categories in Apoidea will be more than the currently identified, and their diversity may be higher than we predicted.
After invading the host genome, transposons undergo amplification, diversification, inactivation, and elimination. The presence of intact copies may indicate that transposons are in the amplification stage or have recently undergone amplification and may be active [74,75]. However, although PB transposons have invaded a large number of bee genomes, the total copy number is very low in most species, indicating that most PB transposons do not show significant amplification in bees. Bombus, which transitioned from solitary bees to social bees [12], has more copies than other genera, although the intact copies were relatively rare, suggesting that the transposon content may be somewhat related to the living environment. Most short copies or truncated copies of PB, due to lack of functional domains of transposases or TIRs, lost transposition activity. Other transposon families such as ZB and SB, which have been identified from animal genomes and belong to the Tc1/mariner superfamilies, exhibit high intact copy numbers (ranging from 10 to over 80 copies) in some genomes, with over 70% of species containing intact copies [76]. In contrast, intact copies of PB transposons are rare in Apoidea, with only 20 out of 44 species showing intact copies, and the copy numbers of intact copies are low in each detected species, with only H. calliopsidis having more than 5 intact copies in its genome, suggesting potential activity.

4.2. Structure Organization of piggyBac

The structure of most complete PB transposons in bees was found to be similar to the original sequence found in T. ni, consisting of five domains that have been previously defined [45]. TIRs play a crucial role in transposase recognition and target site cleavage [36,77,78]. Previous studies have shown that mutations in the first two pairs of bases in the TIRs can result in defects in the excision process [36,79]. However, our data showed that a large part of the identified PB transposons had CAC instead of CCC/GGG for their first three TIR bases. In seven bees (A. haemorrhoa, B. balteatus, Bombus confusus, B. haemorrhoidalis, B. huntii, B. terrestris, and C. calcarata), the TSDs of the PB transposons were not TTAA, which suggests possible mutations. Our data, along with previous studies, revealed that PB transposases may contain a highly divergent DDD domain, analogous to the widely distributed DDE domain found in many DNA transposases and retroviral integrases, which consists of two highly conserved aspartic acids and one glutamic acid [35]. The DDD domains of the three clades of transposases were highly conserved and are known to form an RNase H-like fold [43,60,66], with involvement in the nicking, hairpin resolution, and target joining of PB transposons [43,44]. Although the PB transposon in M. dahlbomi can encode protein of 596 aa with consistent 16 bp TIR at both ends, the second conserved residue in the catalytic domain of DDD mutated from “D” to “N”. Similarly, PB transposons in A. compressa encode a protein of 552 aa and have 14 bp TIR at both ends; the third conserved remnant in the catalytic domain of DDD mutated from “D” to “H”. These results suggest that PB transposons in these two species may have lost their transposable activity, which is supported by the late insertion age of transposons in M. dahlbomi, as determined by evolutionary dynamics analysis. In contrast, the DDBD1 and DDBD2 domains among the three types of PB transposons are not as conserved as DDD. DDBD1 and DDBD2 are special all-α-helix domains, knitting the protein together and interacting with TIR [45]. While the CRD of PB transposases are not conserved, they contain several highly conserved cysteines, with conserved distances between them. The CRD in the PB transposases do not match many other known C-rich domains [35]. NTD is the least conserved among all the domains, similar to other transposases, and may be involved in binding TIRs [35]. NTD residues show high variability, except for a more conserved methionine at the N-terminal and tryptophan at the C-terminal. Multiple sequence alignment indicates that Clade B transposons are more conserved than the other two clades.

4.3. Evolution Dynamics of piggyBac in Apoidea

Generally, differential evolutionary dynamics were observed for PB elements in Apoidea, with most of the excavated PB transposons representing old copies indicated by high K divergence. However, PB transposons in several species (such as A. dorsata, E. continuus, E. tridentata, F. varia, M. europaea, M. rotundata, L. pauxillum, S. tumulorum, H. calliopsidis, and N. spinosus) showed very low K divergences, indicating that these transposons are recent invaders and might still be active, contributing to ongoing genome renewal. In addition, multiple waves of PB amplifications were detected for some species, suggesting repeat invasions of PB transposons. The copy number of each transposon appears to be related to its insertion age and evolutionary mode. For instance, PB transposons in B. hypnorum and B. bifarius had relatively old insertion times, and the copies of PB transposons in these genomes tend to be degenerated due to the accumulation of mutations. In contrast, PB transposons in H. calliopsidis display very low K divergences, indicating young insertion ages, and more intact copies were detected.

5. Conclusions

Our study provides the first comprehensive information on the distribution of PB transposons in 44 species of Apoidea and has analyzed their activity, phylogeny, and structural characteristics. In general, PB transposons are widely distributed in Apoidea and can be divided into three classes. Despite low overall sequence similarity among the three clades, each class of transposons exhibits highly conserved DDD domains. Furthermore, we detected low copy numbers of PB transposons in most Apoidea genomes. Interestingly, we observed divergent evolution dynamics of PB transposons in the genomes of Apoidea, with intact copies being rare in these 44 species.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/insects14040402/s1, Figure S1: Complete phylogenetic tree of piggyBac (PB) elements in Apoidea identified in this study together with 24 reference sequences. The scale bar of the phylogenetic tree represents 0.5. The sequences of IS1380 were used as the outgroup. Red, blue, and yellow circles, respectively, represented three types of PB transposons in Apoidea. (A–C). (Apo-Laspau: Lasioglossum pauxillum, Apo-Andful: Andrena fulva, Apo-Colgig: Colletes gigas, Apo-Cerryb: Cerceris rybyensis, Apo-Megrot: Megachile rotundata, Apo-Frivar: Frieseomelitta varia, Apo-Holcal: Holcopasites calliopsidis, Apo-Exotri: Exoneurella tridentata, Apo-Cercal: Ceratina calcarata, Apo-Seltum: Seladonia tumulorum, Apo-Anddor: Andrena dorsata, Apo-Ectcon: Ectemnius continuus, Apo-Mimdah: Mimumesa dahlbomi, Apo-Cteter: Ctenoplectra terminalis, Apo-Maceur: Macropis europaea, Apo-Andmin: Andrena minutula, Apo-Nysspi: Nysson spinosus, Apo-Hablab: Habropoda laboriosa, Apo-Bomter: Bombus terrestris, Apo-Bomvos: Bombus vosnesenskii, Apo-Megwil: Megachile willughbiella, Apo-Nomfab: Nomada fabriciana, Apo-Bomhae: Bombus haemorrhoidalis, Apo-Ampcom: Ampulex compressa, Apo-Dufnov: Dufourea novaeangliae, Apo-Stepha: Stelis phaeoptera, Apo-Eugdil: Euglossa dilemma, Apo-Bomtur: Bombus turneri, Apo-Bomhyp: Bombus hypnorum, Apo-Bomsor: Bombus soroeensis, Apo-Bombal: Bombus balteatus, Apo- Bomcon: Bombus confusus, Apo-Bomign: Bombus ignitus, Apo-Bombif: Bombus bifarius, Apo-Ceraus: Ceratina australensis). Figure S2: Multiple sequence alignment of different domains of three classes. Different conserved residues are colored differently: (A) Alignment of NTD of complete piggyBac (PB) transposases (≥500 aa). The black boxes are the more conservative areas; (B) Alignment of DDBD1 of complete PB transposases. The numbers above represent the position of the amino acid; (C) Alignment of DDBD2 of complete PB transposases; (D) Alignment of CRD of complete PB transposases. (Apo-Laspau: Lasioglossum pauxillum, Apo-Colgig: Colletes gigas, Apo-Frivar: Frieseomelitta varia, Apo-Megrot: Megachile rotundata, Apo-Cercal: Ceratina calcarata, Apo-Seltum: Seladonia tumulorum, Apo-Anddor: Andrena dorsata, Apo-Holcal: Holcopasites calliopsidis, Apo-Exotri: Exoneurella tridentata, Apo-Ectcon: Ectemnius continuus, Apo-Mimdah: Mimumesa dahlbomi, Apo-Cteter: Ctenoplectra terminalis, Apo-Maceur: Macropis europaea, Apo-Andmin: Andrena minutula, Apo-Nysspi: Nysson spinosus, Apo-Hablab: Habropoda laboriosa, Apo-Bomter: Bombus terrestris, Apo-Megwil: Megachile willughbiella, Apo-Bomhae: Bombus haemorrhoidalis, Apo-Ampcom: Ampulex compressa). Figure S3: Evolutionary dynamics of piggyBac (PB) in Apoidea. The y-axis represents the coverage (%) of each PB transposon in the genome, and the x-axis indicates the K divergence estimate (%), while the numbers in blankets represent the intact copy number and total copy number of PB transposon in each genome. The figure only shows species with high K values. At the top of each chart are species names. Table S1: Information about piggyBac (PB) elements in Apoidea.

Author Contributions

Conceptualization, X.L., C.S., B.G., Z.L. and T.J.; methodology, X.L., Z.G., Y.W., S.S. and C.S.; writing—original draft preparation, X.L., Z.G., F.W., E.A. and C.S.; writing—review and editing, X.L., E.A., Z.L., T.J. and C.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded with grants from the National Natural Science Foundation of China (32271508 and 31671313) and the High-end Talent Support Program of Yangzhou University to Chengyi Song.

Data Availability Statement

All data needed to evaluate the conclusions in this paper are present either in the main text or in the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Danforth, B.N.; Sipes, S.; Fang, J.; Brady, S.G. The history of early bee diversification based on five genes plus morphology. Proc. Natl. Acad. Sci. USA 2006, 103, 15118–15123. [Google Scholar] [CrossRef]
  2. Sann, M.; Niehuis, O.; Peters, R.S.; Mayer, C.; Kozlov, A.; Podsiadlowski, L.; Bank, S.; Meusemann, K.; Misof, B.; Bleidorn, C.; et al. Phylogenomic analysis of Apoidea sheds new light on the sister group of bees. BMC Evol. Biol. 2018, 18, 71. [Google Scholar] [CrossRef] [PubMed]
  3. Michener, C.D. The Professional Development of an Entomologist. Annu. Rev. Entomol. 2007, 52, 1–15. [Google Scholar] [CrossRef] [PubMed]
  4. Zheng, B.-Y.; Cao, L.-J.; Tang, P.; van Achterberg, K.; Hoffmann, A.A.; Chen, H.-Y.; Chen, X.-X.; Wei, S.-J. Gene arrangement and sequence of mitochondrial genomes yield insights into the phylogeny and evolution of bees and sphecid wasps (Hymenoptera: Apoidea). Mol. Phylogenet. Evol. 2018, 124, 1–9. [Google Scholar] [CrossRef] [PubMed]
  5. Ohl, M.; Engel, M.S. Die Fossilgeschichte der Bienen und Ihrer Nächsten Verwandten (Hymenoptera: Apoidea). Denisia 2007, 20, 687–700. [Google Scholar]
  6. Debevec, A.H.; Cardinal, S.; Danforth, B.N. Identifying the sister group to the bees: A molecular phylogeny of Aculeata with an emphasis on the superfamily Apoidea. Zool. Scr. 2012, 41, 527–535. [Google Scholar] [CrossRef]
  7. Cardinal, S.; Danforth, B. The Antiquity and Evolutionary History of Social Behavior in Bees. PLoS ONE 2011, 6, e21086. [Google Scholar] [CrossRef] [PubMed]
  8. Cardinal, S.; Danforth, B. Bees diversified in the age of eudicots. Proc. R. Soc. B Biol. Sci. 2013, 280, 20122686. [Google Scholar] [CrossRef]
  9. Hedtke, S.M.; Patiny, S.; Danforth, B.N. The bee tree of life: A supermatrix approach to apoid phylogeny and biogeography. BMC Evol. Biol. 2013, 13, 138. [Google Scholar] [CrossRef]
  10. Danforth, B.N.; Fang, J.; Sipes, S. Analysis of family-level relationships in bees (Hymenoptera: Apiformes) using 28S and two previously unexplored nuclear genes: CAD and RNA polymerase II. Mol. Phylogenet. Evol. 2006, 39, 358–372. [Google Scholar] [CrossRef] [PubMed]
  11. Danforth, B.N.; Cardinal, S.; Praz, C.; Almeida, E.A.; Michez, D. The Impact of Molecular Data on Our Understanding of Bee Phylogeny and Evolution. Annu. Rev. Entomol. 2013, 58, 57–78. [Google Scholar] [CrossRef] [PubMed]
  12. Liu, Y.; Zong, W.; Diaby, M.; Lin, Z.; Wang, S.; Gao, B.; Ji, T.; Song, C. Diversity and Evolution of pogo and Tc1/mariner Transposons in the Apoidea Genomes. Biology 2021, 10, 940. [Google Scholar] [CrossRef] [PubMed]
  13. Lorenzo-Felipe, I.; Blanco, C.A.; Corona, M. Impact of Apoidea (Hymenoptera) on the World’s Food Production and Diets. Ann. Entomol. Soc. Am. 2020, 113, 407–424. [Google Scholar] [CrossRef]
  14. Hepburn, H. The Bees of the World. Afr. Zool. 2001, 36, 117. [Google Scholar] [CrossRef]
  15. Sang, Y.; Gao, B.; Diaby, M.; Zong, W.; Chen, C.; Shen, D.; Wang, S.; Wang, Y.; Ivics, Z.; Song, C. Incomer, a DD36E family of Tc1/mariner transposons newly discovered in animals. Mob. DNA 2019, 10, 45. [Google Scholar] [CrossRef]
  16. Han, M.-J.; Xu, H.-E.; Zhang, H.-H.; Feschotte, C.; Zhang, Z. Spy: A New Group of Eukaryotic DNA Transposons without Target Site Duplications. Genome Biol. Evol. 2014, 6, 1748–1757. [Google Scholar] [CrossRef]
  17. Holt, R.A.; Subramanian, G.M.; Halpern, A.; Sutton, G.G.; Charlab, R.; Nusskern, D.R.; Wincker, P.; Clark, A.G.; Ribeiro, J.C.; Wides, R.; et al. The Genome Sequence of the Malaria Mosquito Anopheles gambiae. Science 2002, 298, 129–149. [Google Scholar] [CrossRef]
  18. Kapitonov, V.V.; Jurka, J. Molecular paleontology of transposable elements in the Drosophila melanogaster genome. Proc. Natl. Acad. Sci. USA 2003, 100, 6569–6574. [Google Scholar] [CrossRef] [PubMed]
  19. Nene, V.; Wortman, J.R.; Lawson, D.; Haas, B.; Kodira, C.; Tu, Z.; Loftus, B.; Xi, Z.; Megy, K.; Grabherr, M.; et al. Genome Sequence of Aedes aegypti, a Major Arbovirus Vector. Science 2007, 316, 1718–1723. [Google Scholar] [CrossRef]
  20. Xu, H.-E.; Zhang, H.-H.; Xia, T.; Han, M.-J.; Shen, Y.-H.; Zhang, Z. BmTEdb: A collective database of transposable elements in the silkworm genome. Database 2013, 2013, bat055. [Google Scholar] [CrossRef]
  21. Lander, E.S.; Linton, L.M.; Birren, B.; Nusbaum, C.; Zody, M.C.; Baldwin, J.; Devon, K.; Dewar, K.; Doyle, M.; Fitzhugh, W.; et al. Initial sequencing and analysis of the human genome. Nature 2001, 409, 860–921. [Google Scholar] [CrossRef] [PubMed]
  22. Chinwalla, A.T.; Cook, L.L.; Delehaunty, K.D.; Fewell, G.A.; Fulton, L.A.; Fulton, R.S.; Graves, T.A.; Hillier, L.W.; Mardis, E.R.; McPherson, J.D.; et al. Initial Sequencing and Comparative Analysis of the Mouse Genome. Nature 2002, 420, 520–562. [Google Scholar] [CrossRef]
  23. De Koning, A.P.J.; Gu, W.; Castoe, T.A.; Batzer, M.A.; Pollock, D.D. Repetitive Elements May Comprise Over Two-Thirds of the Human Genome. PLoS Genet. 2011, 7, e1002384. [Google Scholar] [CrossRef] [PubMed]
  24. Feschotte, C.E.D.; Jiang, N.; Wessler, S.R. Plant transposable elements: Where genetics meets genomics. Nat. Rev. Genet. 2002, 3, 329–341. [Google Scholar] [CrossRef]
  25. Kidwell, M.G. Transposable elements and the evolution of genome size in eukaryotes. Genetica 2002, 115, 49–63. [Google Scholar] [CrossRef]
  26. Li, Y.; Li, C.; Xia, J.; Jin, Y. Domestication of Transposable Elements into MicroRNA Genes in Plants. PLoS ONE 2011, 6, e19212. [Google Scholar] [CrossRef] [PubMed]
  27. Rebollo, R.; Romanish, M.T.; Mager, D.L. Transposable Elements: An Abundant and Natural Source of Regulatory Sequences for Host Genes. Annu. Rev. Genet. 2012, 46, 21–42. [Google Scholar] [CrossRef]
  28. Almeida, L.; Silva, I.T.; Silva, W.A.; Castro, J.P.; Riggs, P.K.; Carareto, C.M.; Amaral, M.E.J. The contribution of transposable elements to Bos taurus gene structure. Gene 2007, 390, 180–189. [Google Scholar] [CrossRef] [PubMed]
  29. Shapiro, J.A. Mobile DNA and evolution in the 21st century. Mob. DNA 2010, 1, 4. [Google Scholar] [CrossRef]
  30. Wicker, T.; Sabot, F.; Hua-Van, A.; Bennetzen, J.L.; Capy, P.; Chalhoub, B.; Flavell, A.; Leroy, P.; Morgante, M.; Panaud, O.; et al. A unified classification system for eukaryotic transposable elements. Nat. Rev. Genet. 2007, 8, 973–982. [Google Scholar] [CrossRef] [PubMed]
  31. Bourque, G.; Burns, K.H.; Gehring, M.; Gorbunova, V.; Seluanov, A.; Hammell, M.; Imbeault, M.; Izsvák, Z.; Levin, H.L.; Macfarlan, T.S.; et al. Ten things you should know about transposable elements. Genome Biol. 2018, 19, 199. [Google Scholar] [CrossRef]
  32. Kojima, K.K. Structural and sequence diversity of eukaryotic transposable elements. Genes Genet. Syst. 2019, 94, 233–252. [Google Scholar] [CrossRef] [PubMed]
  33. Cary, L.C.; Goebel, M.; Corsaro, B.G.; Wang, H.-G.; Rosen, E.; Fraser, M.J. Transposon mutagenesis of baculoviruses: Analysis of Trichoplusia ni transposon IFP2 insertions within the FP-locus of nuclear polyhedrosis viruses. Virology 1989, 172, 156–169. [Google Scholar] [CrossRef] [PubMed]
  34. Fraser, M.J.; Smith, G.E.; Summers, M.D. Acquisition of Host Cell DNA Sequences by Baculoviruses: Relationship between Host DNA Insertions and FP Mutants of Autographa californica and Galleria mellonella Nuclear Polyhedrosis Viruses. J. Virol. 1983, 47, 287–300. [Google Scholar] [CrossRef] [PubMed]
  35. Sarkar, A.; Sim, C.; Hong, Y.S.; Hogan, J.R.; Fraser, M.J.; Robertson, H.M.; Collins, F.H. Molecular evolutionary analysis of the widespread piggyBac transposon family and related “domesticated” sequences. Mol. Genet. Genom. 2003, 270, 173–180. [Google Scholar] [CrossRef]
  36. Bouallègue, M.; Rouault, J.-D.; Hua-Van, A.; Makni, M.; Capy, P. Molecular evolution of piggyBac superfamily: From selfishness to domestication. Genome Biol. Evol. 2017, 9, 323–339. [Google Scholar] [CrossRef]
  37. Ding, S.; Wu, X.; Li, G.; Han, M.; Zhuang, Y.; Xu, T. Efficient Transposition of the piggyBac (PB) Transposon in Mammalian Cells and Mice. Cell 2005, 122, 473–483. [Google Scholar] [CrossRef]
  38. Yusa, K.; Zhou, L.; Li, M.A.; Bradley, A.; Craig, N.L. A hyperactive piggyBac transposase for mammalian applications. Proc. Natl. Acad. Sci. USA 2011, 108, 1531–1536. [Google Scholar] [CrossRef] [PubMed]
  39. Manuri, P.V.R.; Wilson, M.H.; Maiti, S.N.; Mi, T.; Singh, H.; Olivares, S.; Dawson, M.J.; Huls, H.; Lee, D.A.; Rao, P.H.; et al. piggyBac Transposon/Transposase System to Generate CD19-Specific T Cells for the Treatment of B-Lineage Malignancies. Hum. Gene Ther. 2010, 21, 427–437. [Google Scholar] [CrossRef]
  40. Kolacsek, O.; Wachtl, G.; Fóthi, Á.; Schamberger, A.; Sándor, S.; Pergel, E.; Varga, N.; Raskó, T.; Izsvák, Z.; Apáti, Á.; et al. Functional indications for transposase domestications–Characterization of the human piggyBac transposase derived (PGBD) activities. Gene 2022, 834, 146609. [Google Scholar] [CrossRef]
  41. Fraser, M.J.; Clszczon, T.; Elick, T.; Bauser, C. Precise excision of TTAA-specific lepidopteran transposons piggyBac (IFP2) and tagalong (TFP3) from the baculovirus genome in cell lines from two species of Lepidoptera. Insect Mol. Biol. 1996, 5, 141–151. [Google Scholar] [CrossRef]
  42. Lobo, N.; Li, X.; Fraser, M.J. Transposition of the piggyBac element in embryos of Drosophila melanogaster, Aedes aegypti and Trichoplusiani. Mol. Gen. Genet. 1999, 261, 803–810. [Google Scholar] [CrossRef]
  43. Yusa, K. piggyBac Transposon. Microbiol. Spectr. 2015, 3, 875–892. [Google Scholar] [CrossRef] [PubMed]
  44. Mitra, R.; Fain-Thornton, J.; Craig, N.L. piggyBac can bypass DNA synthesis during cut and paste transposition. EMBO J. 2008, 27, 1097–1109. [Google Scholar] [CrossRef]
  45. Chen, Q.; Luo, W.; Veach, R.A.; Hickman, A.B.; Wilson, M.H.; Dyda, F. Structural basis of seamless excision and specific targeting by piggyBac transposase. Nat. Commun. 2020, 11, 3446. [Google Scholar] [CrossRef]
  46. Gibbs, R.A.; Weinstock, G.M.; Metzker, M.L.; Muzny, D.M.; Sodergren, E.J.; Scherer, S.; Scott, G.; Steffen, D.; Worley, K.C.; Burch, P.E.; et al. Genome sequence of the Brown Norway rat yields insights into mammalian evolution. Nature 2004, 428, 493–521. [Google Scholar] [CrossRef] [PubMed]
  47. Lindblad-Toh, K.; Wade, C.M.; Mikkelsen, T.S.; Karlsson, E.K.; Jaffe, D.B.; Kamal, M.; Clamp, M.; Chang, J.L.; Kulbokas, E.J., III; Zody, M.C.; et al. Genome sequence, comparative analysis and haplotype structure of the domestic dog. Nature 2005, 438, 803–819. [Google Scholar] [CrossRef]
  48. Pace, J.K.; Feschotte, C. The evolutionary history of human DNA transposons: Evidence for intense activity in the primate lineage. Genome Res. 2007, 17, 422–432. [Google Scholar] [CrossRef]
  49. Ray, D.A.; Feschotte, C.; Pagan, H.J.; Smith, J.D.; Pritham, E.J.; Arensburger, P.; Atkinson, P.W.; Craig, N.L. Multiple waves of recent DNA transposon activity in the bat, Myotis lucifugus. Genome Res. 2008, 18, 717–728. [Google Scholar] [CrossRef] [PubMed]
  50. Pagan, H.J.T.; Smith, J.D.; Hubley, R.M.; Ray, D.A. PiggyBac-ing on a Primate Genome: Novel Elements, Recent Activity and Horizontal Transfer. Genome Biol. Evol. 2010, 2, 293–303. [Google Scholar] [CrossRef] [PubMed]
  51. Wang, J.; Du, Y.; Wang, S.; Brown, S.J.; Park, Y. Large diversity of the piggyBac-like elements in the genome of Tribolium castaneum. Insect Biochem. Mol. Biol. 2008, 38, 490–498. [Google Scholar] [CrossRef] [PubMed]
  52. Ebert, P.R.; Hileman, J.P.; Nguyen, H.T. Primary sequence, copy number, and distribution of mariner transposons in the honey bee. Insect Mol. Biol. 1995, 4, 69–78. [Google Scholar] [CrossRef]
  53. Altschul, S.F.; Madden, T.L.; Schäffer, A.A.; Zhang, J.; Zhang, Z.; Miller, W.; Lipman, D.J. Gapped BLAST and PSI-BLAST: A new generation of protein database search programs. Nucleic Acids Res. 1997, 25, 3389–3402. [Google Scholar] [CrossRef] [PubMed]
  54. Kalendar, R.; Lee, D.; Schulman, A. FastPCR Software for PCR Primer and Probe Design and Repeat Search. Genes Genomes Genom. 2009, 3, 1–14. [Google Scholar]
  55. Yang, P.; Craig, P.A.; Goodsell, D.; Bourne, P.E. BioEditor—Simplifying macromolecular structure annotation. Bioinformatics 2003, 19, 897–898. [Google Scholar] [CrossRef] [PubMed]
  56. Chen, C.J.; Chen, H.; Zhang, Y.; Thomas, H.R.; Frank, M.H.; He, Y.H.; Xia, R. TBtools: An Integrative Toolkit Developed for Interactive Analyses of Big Biological Data. Mol. Plant 2020, 13, 1194–1202. [Google Scholar] [CrossRef]
  57. Yamada, K.D.; Tomii, K.; Katoh, K. Application of the MAFFT sequence alignment program to large data—Reexamination of the usefulness of chained guide trees. Bioinformatics 2016, 32, 3246–3251. [Google Scholar] [CrossRef] [PubMed]
  58. Liu, W.; Xie, Y.; Ma, J.; Luo, X.; Nie, P.; Zuo, Z.; Lahrmann, U.; Zhao, Q.; Zheng, Y.; Zhao, Y.; et al. IBS: An illustrator for the presentation and visualization of biological sequences. Bioinformatics 2015, 31, 3359–3361. [Google Scholar] [CrossRef] [PubMed]
  59. Siguier, P.; Perochon, J.; Lestrade, L.; Mahillon, J.; Chandler, M. ISfinder: The reference centre for bacterial insertion sequences. Nucleic Acids Res. 2006, 34, D32–D36. [Google Scholar] [CrossRef] [PubMed]
  60. Hickman, A.B.; Chandler, M.; Dyda, F. Integrating prokaryotes and eukaryotes: DNA transposases in light of structure. Crit. Rev. Biochem. Mol. Biol. 2010, 45, 50–69. [Google Scholar] [CrossRef]
  61. Nguyen, L.-T.; Schmidt, H.A.; Von Haeseler, A.; Minh, B.Q. IQ-TREE: A Fast and Effective Stochastic Algorithm for Estimating Maximum-Likelihood Phylogenies. Mol. Biol. Evol. 2015, 32, 268–274. [Google Scholar] [CrossRef]
  62. Kalyaanamoorthy, S.; Minh, B.Q.; Wong, T.K.F.; Von Haeseler, A.; Jermiin, L.S. ModelFinder: Fast model selection for accurate phylogenetic estimates. Nat. Methods 2017, 14, 587–589. [Google Scholar] [CrossRef] [PubMed]
  63. Tarailo-Graovac, M.; Chen, N. Using RepeatMasker to Identify Repetitive Elements in Genomic Sequences. Curr. Protoc. Bioinform. 2009, 25, 4–10. [Google Scholar] [CrossRef] [PubMed]
  64. Kimura, M. A simple method for estimating evolutionary rates of base substitutions through comparative studies of nucleotide sequences. J. Mol. Evol. 1980, 16, 111–120. [Google Scholar] [CrossRef] [PubMed]
  65. Dunemann, S.M.; Wasmuth, J.D. Horizontal transfer of a retrotransposon between parasitic nematodes and the common shrew. Mob. DNA 2019, 10, 24. [Google Scholar] [CrossRef] [PubMed]
  66. Yuan, Y.-W.; Wessler, S.R. The catalytic domain of all eukaryotic cut-and-paste transposase superfamilies. Proc. Natl. Acad. Sci. USA 2011, 108, 7884–7889. [Google Scholar] [CrossRef]
  67. Ghanim, G.E.; Kellogg, E.H.; Nogales, E.; Rio, D.C. Structure of a P element transposase–DNA complex reveals unusual DNA structures and GTP-DNA contacts. Nat. Struct. Mol. Biol. 2019, 26, 1013–1022. [Google Scholar] [CrossRef] [PubMed]
  68. Keith, J.H.; Schaeper, C.A.; Fraser, T.S.; Fraser, M.J. Mutational analysis of highly conserved aspartate residues essential to the catalytic core of the piggyBac transposase. BMC Mol. Biol. 2008, 9, 73. [Google Scholar] [CrossRef]
  69. Horn, C.; Offen, N.; Nystedt, S.; Häcker, U.; Wimmer, E.A. piggyBac-Based Insertional Mutagenesis and Enhancer Detection as a Tool for Functional Insect Genomics. Genetics 2003, 163, 647–661. [Google Scholar] [CrossRef]
  70. Aparicio, S.; Chapman, J.; Stupka, E.; Putnam, N.; Chia, J.-M.; Dehal, P.; Christoffels, A.; Rash, S.; Hoon, S.; Smit, A.; et al. Whole-Genome Shotgun Assembly and Analysis of the Genome of Fugu rubripes. Science 2002, 297, 1301–1310. [Google Scholar] [CrossRef] [PubMed]
  71. Gao, B.; Zong, W.; Miskey, C.; Ullah, N.; Diaby, M.; Chen, C.; Wang, X.; Ivics, Z.; Song, C. Intruder (DD38E), a recently evolved sibling family of DD34E/Tc1 transposons in animals. Mob. DNA 2020, 11, 32. [Google Scholar] [CrossRef] [PubMed]
  72. Shen, D.; Gao, B.; Miskey, C.; Chen, C.; Sang, Y.; Zong, W.; Wang, S.; Wang, Y.; Wang, X.; Ivics, Z.; et al. Multiple Invasions of Visitor, a DD41D Family of Tc1/mariner Transposons, throughout the Evolution of Vertebrates. Genome Biol. Evol. 2020, 12, 1060–1073. [Google Scholar] [CrossRef] [PubMed]
  73. Guan, Z.; Shi, S.; Diaby, M.; Danley, P.; Ullah, N.; Puzakov, M.; Gao, B.; Song, C. Horizontal transfer of Buster transposons across multiple phyla and classes of animals. Mol. Phylogenet. Evol. 2022, 173, 107506. [Google Scholar] [CrossRef]
  74. Schaack, S.; Gilbert, C.; Feschotte, C. Promiscuous DNA: Horizontal transfer of transposable elements and why it matters for eukaryotic evolution. Trends Ecol. Evol. 2010, 25, 537–546. [Google Scholar] [CrossRef]
  75. Shi, S.; Puzakov, M.; Guan, Z.; Xiang, K.; Diaby, M.; Wang, Y.; Wang, S.; Song, C.; Gao, B. Prokaryotic and Eukaryotic Horizontal Transfer of Sailor (DD82E), a New Superfamily of IS630-Tc1-Mariner DNA Transposons. Biology 2021, 10, 1005. [Google Scholar] [CrossRef]
  76. Jia, W.; Asare, E.; Liu, T.; Zhang, P.; Wang, Y.; Wang, S.; Shen, D.; Miskey, C.; Gao, B.; Ivics, Z.; et al. Horizontal Transfer and Evolutionary Profiles of Two Tc1/DD34E Transposons (ZB and SB) in Vertebrates. Genes 2022, 13, 2239. [Google Scholar] [CrossRef] [PubMed]
  77. Elick, T.A.; Lobo, N.; Fraser, M.J. Analysis of the cis-acting DNA elements required for piggyBac transposable element excision. Mol. Genet. Genom. 1997, 255, 605–610. [Google Scholar] [CrossRef]
  78. Li, X.; Lobo, N.; Bauser, C.A.; Fraser, M.J. The minimum internal and external sequence requirements for transposition of the eukaryotic transformation vector piggyBac. Mol. Genet. Genom. 2001, 266, 190–198. [Google Scholar] [CrossRef] [PubMed]
  79. Haniford, D.; Kleckner, N. Tn 10 transposition in vivo: Temporal separation of cleavages at the two transposon ends and roles of terminal basepairs subsequent to interaction of ends. EMBO J. 1994, 13, 3401–3411. [Google Scholar] [CrossRef]
Figure 1. Distribution of piggyBac (PB) transposons in Apoidea among different genera. The tree on the left is the phylogenetic relationship between all genera in which transposons were detected, and different colors represent different families. The numbers in parentheses are the number of species in the families and genera used to examine, and “n” represents the number of species with PB transposons detected in each genus.
Figure 1. Distribution of piggyBac (PB) transposons in Apoidea among different genera. The tree on the left is the phylogenetic relationship between all genera in which transposons were detected, and different colors represent different families. The numbers in parentheses are the number of species in the families and genera used to examine, and “n” represents the number of species with PB transposons detected in each genus.
Insects 14 00402 g001
Figure 2. Structural organization of complete and representative piggyBac (PB) elements in Apoidea. (1) The bright blue part represents the terminal inverted repeats (TIRs) at both ends of the PB transposons; (2) The purple part represents the ORF; (3) The yellow part in the middle represents the catalytic domains; (4) The corresponding length of each section is marked below.
Figure 2. Structural organization of complete and representative piggyBac (PB) elements in Apoidea. (1) The bright blue part represents the terminal inverted repeats (TIRs) at both ends of the PB transposons; (2) The purple part represents the ORF; (3) The yellow part in the middle represents the catalytic domains; (4) The corresponding length of each section is marked below.
Insects 14 00402 g002
Figure 3. The logo figure of piggyBac (PB) Transposons’ TIR (n = 40). The online website Welogo v.3.7.12 (http://weblogo.threeplusone.com/create.cgi, accessed on 16 December 2022) was used to create logo images of the TIR sequences. The value 2 on the y axis stands for maximum possible frequency.
Figure 3. The logo figure of piggyBac (PB) Transposons’ TIR (n = 40). The online website Welogo v.3.7.12 (http://weblogo.threeplusone.com/create.cgi, accessed on 16 December 2022) was used to create logo images of the TIR sequences. The value 2 on the y axis stands for maximum possible frequency.
Insects 14 00402 g003
Figure 4. Phylogenetic tree of piggyBac (PB) elements in Apoidea identified in this study together with 24 reference sequences. The sequences of IS1380 were used as the outgroup. Red, blue, and green, respectively, represented three types of PB transposons (A–C). The complete evolutionary tree is in Figure S1.
Figure 4. Phylogenetic tree of piggyBac (PB) elements in Apoidea identified in this study together with 24 reference sequences. The sequences of IS1380 were used as the outgroup. Red, blue, and green, respectively, represented three types of PB transposons (A–C). The complete evolutionary tree is in Figure S1.
Insects 14 00402 g004
Figure 5. Multiple sequence alignments of the DDD domains of complete piggyBac (PB) transposases (≥ 500 aa) of three classes. The black frame represents the insertion motif of the DDD domain, and different conserved residues are colored differently. (Apo-Laspau: Lasioglossum pauxillum, Apo-Colgig: Colletes gigas, Apo-Frivar: Frieseomelitta varia, Apo-Megrot: Megachile rotundata, Apo-Cercal: Ceratina calcarata, Apo-Seltum: Seladonia tumulorum, Apo-Anddor: Andrena dorsata, Apo-Holcal: Holcopasites calliopsidis, Apo-Exotri: Exoneurella tridentata, Apo-Ectcon: Ectemnius continuus, Apo-Mimdah: Mimumesa dahlbomi, Apo-Cteter: Ctenoplectra terminalis, Apo-Maceur: Macropis europaea, Apo-Andmin: Andrena minutula, Apo-Nysspi: Nysson spinosus, Apo-Hablab: Habropoda laboriosa, Apo-Bomter: Bombus terrestris, Apo-Megwil: Megachile willughbiella, Apo-Bomhae: Bombus haemorrhoidalis, Apo-Ampcom: Ampulex compressa). Multiple sequence alignments of other domains are in Supplementary Figure S2.
Figure 5. Multiple sequence alignments of the DDD domains of complete piggyBac (PB) transposases (≥ 500 aa) of three classes. The black frame represents the insertion motif of the DDD domain, and different conserved residues are colored differently. (Apo-Laspau: Lasioglossum pauxillum, Apo-Colgig: Colletes gigas, Apo-Frivar: Frieseomelitta varia, Apo-Megrot: Megachile rotundata, Apo-Cercal: Ceratina calcarata, Apo-Seltum: Seladonia tumulorum, Apo-Anddor: Andrena dorsata, Apo-Holcal: Holcopasites calliopsidis, Apo-Exotri: Exoneurella tridentata, Apo-Ectcon: Ectemnius continuus, Apo-Mimdah: Mimumesa dahlbomi, Apo-Cteter: Ctenoplectra terminalis, Apo-Maceur: Macropis europaea, Apo-Andmin: Andrena minutula, Apo-Nysspi: Nysson spinosus, Apo-Hablab: Habropoda laboriosa, Apo-Bomter: Bombus terrestris, Apo-Megwil: Megachile willughbiella, Apo-Bomhae: Bombus haemorrhoidalis, Apo-Ampcom: Ampulex compressa). Multiple sequence alignments of other domains are in Supplementary Figure S2.
Insects 14 00402 g005
Figure 6. Sequence identities between A–C clades. The numbers in the heatmap are the percentage of the average values of the sequences’ identities of the two types of transposons in the corresponding row and column, and “n” represents the number of each type of transposon sequence (ae). The average values of sequence identities were measured by pairwise comparison of sequences of NTD (a), DDBD1 (b), DDD (c), DDBD2 (d), and CRD (e) of complete PB transposases (≥500 aa).
Figure 6. Sequence identities between A–C clades. The numbers in the heatmap are the percentage of the average values of the sequences’ identities of the two types of transposons in the corresponding row and column, and “n” represents the number of each type of transposon sequence (ae). The average values of sequence identities were measured by pairwise comparison of sequences of NTD (a), DDBD1 (b), DDD (c), DDBD2 (d), and CRD (e) of complete PB transposases (≥500 aa).
Insects 14 00402 g006
Figure 7. Evolutionary dynamics of piggyBac (PB) in Apoidea. The y-axis represents each PB transposon’s coverage (%) in the genome, and the x-axis indicates the K divergence estimate (%). In contrast, the numbers in brackets represent the intact copy number and total copy number of PB transposon in each genome. The figure only shows species with low K values. At the top of each chart are species names. More information is displayed in Figure S3.
Figure 7. Evolutionary dynamics of piggyBac (PB) in Apoidea. The y-axis represents each PB transposon’s coverage (%) in the genome, and the x-axis indicates the K divergence estimate (%). In contrast, the numbers in brackets represent the intact copy number and total copy number of PB transposon in each genome. The figure only shows species with low K values. At the top of each chart are species names. More information is displayed in Figure S3.
Insects 14 00402 g007
Table 1. Full and intact copy numbers of piggyBac (PB) transposons in genomes of Apoidea.
Table 1. Full and intact copy numbers of piggyBac (PB) transposons in genomes of Apoidea.
Distribution aAmpulicidaeAndrenidaeApidaeColletidaeCrabronidaeHalictidaeMegachilidaeMelittidae
Full Copy Number32–2361–33544–101–81–86
Average b3.0061.2530.004.007.005.004.756.00
Number of Species with Full Transposons142614341
Intact Copy Number10–30–810–40–10–34
Average C1.001.250.621.002.000.671.254.00
Number of Species with Intact Transposons12813221
a Distribution: eight families of Apoidea in which PB transposons were detected. b Average full copy (transposons flanked by detectable TSDs and TIRs) number of transposons in genomes of Apoidea. C Average intact copy (transposons flanked by detectable TSDs and TIRs and encoded ≥500 aa transposases) number of transposons in genomes of Apoidea.
Table 2. Structure organization of piggyBac (PB) transposons.
Table 2. Structure organization of piggyBac (PB) transposons.
Distribution aSpecies NumberLength of Full Transposons (bp) bNumber of
Species with Full
Transposons
Length of
Intact
Transposon (bp) C
Number of Species with Intact
Transposons
Transposase Length (aa)TIR Length (bp)TSD
Ampulicidae1238012380155214TTAA
Andrenidae42519–351842519–3518283–6815–23TTAA
Apidae26368–3518261670–3413824–6127–493TTAA
Colletidae1283812838158916TTAA
Crabronidae42334–271242334–27123449–60113–18TTAA
Halictidae32326–259632326–25962477–58013–17TTAA
Megachilidae4900–286842292–28682206–58115–20TTAA
Melittidae1264612646156713TTAA
a Distribution: eight families of Apoidea in which PB transposons were detected. b Full transposon: transposons flanked by detectable TSDs (target site duplication) and TIRs. C Intact transposon: transposons flanked by detectable TSDs and TIRs and encoded ≥500 aa transposases.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, X.; Guan, Z.; Wang, F.; Wang, Y.; Asare, E.; Shi, S.; Lin, Z.; Ji, T.; Gao, B.; Song, C. Evolution of piggyBac Transposons in Apoidea. Insects 2023, 14, 402. https://doi.org/10.3390/insects14040402

AMA Style

Li X, Guan Z, Wang F, Wang Y, Asare E, Shi S, Lin Z, Ji T, Gao B, Song C. Evolution of piggyBac Transposons in Apoidea. Insects. 2023; 14(4):402. https://doi.org/10.3390/insects14040402

Chicago/Turabian Style

Li, Xueyuan, Zhongxia Guan, Feng Wang, Yali Wang, Emmanuel Asare, Shasha Shi, Zheguang Lin, Ting Ji, Bo Gao, and Chengyi Song. 2023. "Evolution of piggyBac Transposons in Apoidea" Insects 14, no. 4: 402. https://doi.org/10.3390/insects14040402

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

Article Metrics

Back to TopTop