Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Phylogenomics and Divergence Dating of Fungus-Farming Ants (Hymenoptera: Formicidae) of the Genera Sericomyrmex and Apterostigma

  • Ana Ješovnik ,

    schultzt@si.edu (TRS); jesovnika@si.edu (AJ)

    Affiliations Entomology Department, National Museum of Natural History, Smithsonian Institution, Washington, District of Columbia, United States of America, Maryland Center for Systematic Entomology, Department of Entomology, University of Maryland, College Park, Maryland, United States of America

  • Vanessa L. González,

    Affiliation Global Genome Initiative, National Museum of Natural History, Smithsonian Institution, Washington, District of Columbia, United States of America

  • Ted R. Schultz

    schultzt@si.edu (TRS); jesovnika@si.edu (AJ)

    Affiliation Entomology Department, National Museum of Natural History, Smithsonian Institution, Washington, District of Columbia, United States of America

Abstract

Fungus-farming ("attine") ants are model systems for studies of symbiosis, coevolution, and advanced eusociality. A New World clade of nearly 300 species in 15 genera, all attine ants cultivate fungal symbionts for food. In order to better understand the evolution of ant agriculture, we sequenced, assembled, and analyzed transcriptomes of four different attine ant species in two genera: three species in the higher-attine genus Sericomyrmex and a single lower-attine ant species, Apterostigma megacephala, representing the first genomic data for either genus. These data were combined with published genomes of nine other ant species and the honey bee Apis mellifera for phylogenomic and divergence-dating analyses. The resulting phylogeny confirms relationships inferred in previous studies of fungus-farming ants. Divergence-dating analyses recovered slightly older dates than most prior analyses, estimating that attine ants originated 53.6–66.7 million of years ago, and recovered a very long branch subtending a very recent, rapid radiation of the genus Sericomyrmex. This result is further confirmed by a separate analysis of the three Sericomyrmex species, which reveals that 92.71% of orthologs have 99% - 100% pairwise-identical nucleotide sequences. We searched the transcriptomes for genes of interest, most importantly argininosuccinate synthase and argininosuccinate lyase, which are functional in other ants but which are known to have been lost in seven previously studied attine ant species. Loss of the ability to produce the amino acid arginine has been hypothesized to contribute to the obligate dependence of attine ants upon their cultivated fungi, but the point in fungus-farming ant evolution at which these losses occurred has remained unknown. We did not find these genes in any of the sequenced transcriptomes. Although expected for Sericomyrmex species, the absence of arginine anabolic genes in the lower-attine ant Apterostigma megacephala strongly suggests that the loss coincided with the origin of attine ants.

Introduction

Fungus-farming (hereafter “attine”) ants are a monophyletic group in which all known species grow fungus for food. The most conspicuous attine ants, the leaf-cutting genera Atta and Acromyrmex, are dominant herbivores in Neotropical ecosystems [1] and have become model organisms for studies of symbiosis, higher eusociality, and coevolution [2,3]. The attine ants have been divided into two informal groups: the lower and the higher attine ants, the former paraphyletic with respect to the latter. Higher attine ants include, in addition to the leaf-cutters, the non-leaf-cutting genera Trachymyrmex and Sericomyrmex. All of the higher attine ants grow highly derived, obligately symbiotic fungal cultivars, in contrast to the lower attine ants, which grow facultatively symbiotic cultivars capable of living outside of the symbiosis [2].

The higher attine ant genus Sericomyrmex contains 19 nominal species and 3 nominal subspecies and has a broad Neotropical distribution [4,5]. Sericomyrmex ants are commonly collected in leaf-litter samples in biodiversity studies in South and Central America, but are very hard to identify to the species level. Species are morphologically very similar and within-nest variation is substantial, confounding easy recognition of species boundaries [6]. In addition, preliminary multiple-gene studies have shown surprisingly low molecular divergence (Ješovnik, unpublished), which makes this genus a compelling group in which to study speciation. The phylogenetic position of Sericomyrmex within the higher attine ant clade, and its similarities with Atta leaf-cutter ants [7], makes understanding the evolutionary history and biology of Sericomyrmex species important for reconstructing the origin and evolution of higher attine agriculture and for explaining the ecological success of the leaf-cutting genera Atta and Acromyrmex.

The lower attine ant genus Apterostigma, with 45 described species [8], is remarkable for its symbiotic plasticity. Like all other genera of lower attine ants, one clade of Apterostigma species grows fungi in the tribe Leucocoprineae, whereas, unlike any other lower or higher attine ant, another clade of Apterostigma species cultivates coral fungi in the distantly related family Pterulaceae [2]. Most remarkably, a recent study revealed the only known case of a lower attine ant cultivating a higher attine fungus: A. megacephala grows Leucoagaricus gongylophorus, the most highly derived and recently evolved leucocoprineaceous fungal species, an obligate symbiont otherwise known to be cultivated only by leaf-cutting ants [9].

Attine ant genomic studies [1012] have significantly advanced our understanding of the evolution of fungus-farming in ants. The goal of this study was to sequence, de novo assemble, and characterize transcriptomes for species in the genera Sericomyrmex and Apterostigma, in order to better understand the evolution of fungus-farming behavior and species boundaries within the genus Sericomyrmex. Here we report the first genomic data for both genera. For Sericomyrmex we sequenced the transcriptomes of three different morphospecies. These three morphospecies were chosen to be the most morphologically and molecularly diverged of the samples assembled for a taxonomic revision of the genus that were properly preserved for RNA extraction. We combined the data produced in this study with published ant and honey bee genomes [1016] in order to confirm the phylogenetic position of the genus Sericomyrmex and to infer divergence dates for Sericomyrmex. The transcriptome of A. megacephala is the first genomic data generated for any species in the Paleoattini, one of the two sister clades produced by the basal-most divergence in the fungus-farming ant phylogeny. The other sister clade, the Neoattini, includes the higher attines and all previously sequenced species, so genetic data for a paleoattine species significantly improves our ability to date the loss of arginine biosynthesis.

Materials and Methods

Sample preparation and sequencing

No animal ethics approvals are required to conduct research on ants. Research in Brazil was covered by Brazilian Council of Research and Scientific Development permit Processo CNPq 001884/2012-3 and Instituto Chico Mendes de Conservação da Biodiversidade (ICMBio) collecting permit 14789–6. Research in Peru was covered by Ministerio de Agricultura Instituto Nacional de Recursos Naturales INRENA Autorización No. 034-2004-INRENA-IFFS-DCB; Modificación a la Autorización No. 034-2004-INRENA-IFFS-DCB; Ministerio de Agricultura Instituto Nacional de Recursos Naturales INRENA Autorización No. 12 C/C-2004-INRENA-IANP; Carta No. 553-2014-MINAGRI-DGFFS/DGEFFS; Ministerio de Agricultura Instituto Nacional de Recursos Naturales INRENA Autorización No. 088-2005-INRENA-IFFS-DCB; CARTA No. 0217-2012-SERNANP-JRNTAMB; Resolución del Jefe de la Reserva Nacional Tambopata No. 020-2012-SERNANP-JEF; Permiso para Fauna y Flora Silvestre No. 004905-AG-INRENA; Permiso para Fauna y Flora Silvestre No. 006771-AG-INRENA; and Permiso para Fauna y Flora Silvestre No. 009154-AG-DGFFS.

Specimens of three Sericomyrmex species were collected from live colonies in the field and preserved in RNAlater. Specimens of A. megacephala were collected from a live laboratory nest and also preserved in RNAlater. Voucher specimens for each of the species sequenced are deposited in the Department of Entomology, National Museum of Natural History, Washington, DC, USA (Table 1), as recorded in the NMNH K-EMu database (http://collections.nmnh.si.edu/search/ento/). In order to ensure sufficient quantities of RNA, we included ten workers per sample for each Sericomyrmex species and five workers for the larger species, A. megacephala, crushing them with sterilized wooden sticks to enable the RNAlater to rapidly penetrate the integument. All specimens were stored at -20°C until RNA extraction. We extracted RNA using the Promega SV Total RNA Isolation System, following the standard protocol. The Institute for Bioscience and Biotechnology Research (IBBR) at the University of Maryland prepared libraries using the standard Illumina TruSeq RNA Sample Preparation protocol (all libraries were normalized and redundant rRNA was removed), and performed 100bp pair-end sequencing on Illumina HiSeq.

Data cleaning and assembly

We concatenated the raw data and performed a quality check with FastQC [17] before and after trimming (Fig 1). Raw reads were cleaned and trimmed with Trimmomatic [18] and the resultant cleaned reads were assembled with Trinity [19,20] using the default parameters. Basic assembly statistics (number of transcripts, average transcript length, and N50) were obtained from Trinity. We processed the resulting assemblies in TransDecoder [21] to identify candidate coding regions within the transcripts. Predicted peptides were then filtered with CD-HIT-EST (98% global similarity) [22] to remove redundant sequences. After removal of duplicates, peptides were further filtered in order to select only one peptide per putative transcript by choosing the longest ORF (Open Reading Frame) per Trinity subcomponent with a custom-made Python script [23]. This step ensured that we had removed variation in the coding regions of our assemblies due to alternative splicing, closely related paralogs, and allelic diversity. Peptide sequences for the seven sampled ant genomes were filtered at 98% similarity in CD-HIT-EST. To calculate coverage we multiplied the number of raw reads with the expected length (100 bp), and divided that by total number of bases in the assembly [24].

Orthology Assignment and Alignment

We identified orthologs using OMA stand-alone v.0.99w software [25,26] with default settings on 100 cores on the Smithsonian Lattice high-performance computing cluster (Linux-based with AMD processors). We constructed an amino acid supermatrix by concatenating the set of OMA groups containing all taxa (1,317). Each putative ortholog group (from now on “gene”) was aligned individually using MAFFT [27]. Aligned genes were then trimmed with Gblocks [28] to cull regions of dubious alignment.

We separately ran an orthology assignment in OMA for a data subset containing only the Sericomyrmex transcriptomes. For this we used TransDecoder to identify the nucleotide sequences for the candidate protein-coding regions and performed the same filtering as described above with the larger dataset on the resultant nucleotide sequence data. Individual gene alignments were trimmed with Gblocks [28] and analyzed in Geneious v.8 [29].

Phylogenetic analyses

Maximum-likelihood inference was conducted with PhyML-PCMA [30,31]. We selected 10 PCs (principal components) in the PhyML-PCMA analyses and used empirical amino-acid frequencies. PhyML-PCMA estimates a model through the use of a principal component (PC) analysis; in this case using 10 PCs. Bootstraps were calculated in PhyML-PCMA for 100 replicates. Concomitantly, tree searches were conducted in PhyloBayes MPI 1.4e [32] using the site-heterogeneous CAT + GTR model of evolution. Three independent chains were run for 1295–1445 cycles, and the initial cycles discarded as burn-in were determined for each analyses using the “tracecomp” executable, with convergence assessed using the maximum bipartition discrepancies across chains (maxdiff < 0.1).

Divergence dating

Divergence dates were estimated using the Bayesian relaxed molecular clock approach as implemented in PhyloBayes v.3.3f [32] applying an auto-correlated model of clock relaxation [33,34]. Three calibration constraints (S1 Table), based on a recent study of myrmicine ants [35], were used with soft bounds [36] under a birth-death prior in PhyloBayes. PhyloBayes was run for 32,664 cycles, sampling posterior rates and dates every 5 cycles. The initial 5000 cycles were discarded as burn-in. We excluded the outgroup taxon Apis mellifera from final analyses because we considered it to be an inappropriate outgroup for the purposes of divergence dating. All other taxa are members of a single subfamily, the Myrmicinae, which occupies a highly derived position within the ant family Formicidae, whereas A. mellifera, the honey bee, is a highly derived member of a highly derived family within the superfamily Apoidea. The most recent common ancestor of the lineages containing the Apoidea and the Formicidae dates to at least 140 mya, whereas the ancestor of the Myrmicinae, the focus of our dating analyses, dates to ~99 mya [35,37,38]

Annotation and Ortholog Hit Ratio

We used BLASTx [39] (cut-off E-value 1e-5) to compare the unfiltered, assembled transcripts (Trinity output) against the non-redundant (nr) protein database of NCBI. Resulting xml files were used as input for CLC Workbench (CLC Inc., Aarhus, Denmark). We performed functional annotation and mapping to GO terms, using Blast2GO with default settings, in order to summarize functional categories of the genes, annotate our data set, and determine the quality of our transcriptomes. We also used the resultant xml output to run Orthology Hit Ratio calculations following previously described methods and scripts [40]. This analysis estimates the degree to which a transcriptome is fully sequenced and assembled by comparing the length of the contigs that recovered BLAST hits with the length of their top BLAST hits [41,42].

Gene search

Based on previous studies of attine ant genomes and ant transcriptomes [1012,43] we considered the following genes or gene families of particular interest: genes in the arginine metabolism pathway (arginase, nitric oxide synthase, argininosuccinate synthase, and argininosuccinate lyase), detoxification genes (cytochrome P450 monooxygenase), hexamerins (hex 70a, hex 70b, hex 70c and hex 110), desaturase, RYamide, and chitinases. The nucleotide or protein sequence for each of these genes was downloaded from NCBI GenBank or, in the case of the desaturase gene, obtained from the supplemental material of a previous study [43]. We used sequences from the closely related species Atta cephalotes and Acromyrmex echinatior for searches for arginase, nitric oxide synthase, and cytochrome P450. For chitinase searches we assembled protein and nucleotide sequences that were used in similar analyses in a study of the Atta cephalotes genome [10], for a total of ten different chitinase-like proteins from Acromyrmex echinatior, Nasonia vitripennis, and A. mellifera (S4 Table). For argininosuccinate synthase and argininosuccinate lyase, which are lost in leaf-cutter ants, we used copies from the closely related myrmicine ants Solenopsis invicta, Pogonomyrmex barbatus, and Wasmannia auropunctata. For hexamerin searches we used Apis mellifera sequences in order to repeat the methods of previous studies [10] and because some of the attine ant hexamerin sequences were unavailable. For searches for ryamide, which is absent in ants, we used a Drosphila melanogaster sequence. Full sequences of all genes used in our analyses, with GenBank accession numbers, can be found in S1 Text. We used BLAST+ standalone [44] to manually create a BLAST database for each of the transcriptomes separately (using unfiltered, assembled transcripts) and searched the created databases for each of the query sequences with the cut-off E-value 1e-5. Returned hits were then BLASTed against NCBI to confirm that they had returned the same protein.

Results and Discussion

We sequenced the transcriptomes of the lower attine ant Apterostigma megacephala and three morphospecies of the higher attine ant genus Sericomyrmex (Fig 1, Table 1). The total number of raw sequences was 329.6 Mb, varying from 72 to more than 91 million reads per sample. After trimming and cleaning, between 92.2% and 93.8% of the raw reads were used for assembly. The number of assembled contigs varied from 71,391 for S. cf. luederwaldti to 95,242 for A. megacephala. The N50, contig length, and other statistics for each of the sequenced taxa can be found in Table 2. The number of raw reads and assembled contigs were highest for A. megacephala, which can be attributed to the high quality of preservation of the extracted RNA for this sample.

thumbnail
Table 2. Transcriptome sequencing, assembly, and analysis statistics.

Sources of statistics: 1: FastQC, 2: Trimmomatic, 3: Trinitystats.pl, 4: Geneious, 5: Transdecoder, 6: Blast2GO, 7: calculated from other statistics, 8: Ortholog Hit Ratio.

https://doi.org/10.1371/journal.pone.0151059.t002

Coverage for the transcriptomes varied from 53.8 to 169.7 reads per base (Table 2). Ortholog Hit Ratios (OHR) were calculated in order to estimate transcriptome assembly quality and completeness [42]. An OHR value close to 1 suggests complete transcript assembly, with a value of 1 being an identical match. On average, our transcriptomes had high OHR values, with 47.7 to 53.5% of total contigs having an OHR greater than 0.8, and 63.7 to 69.5% of contigs having an OHR higher than 0.5 (Table 2, S1 Fig). A number of contigs recovered values over 1, which is not uncommon [41], and represents possible sequence expansions. Additionally, we calculated annotation metrics to estimate transcriptome quality (e.g., number of contigs that recovers at least one BLAST hit). Based on these metrics (Figs 2 and 3, Table 2, S2 and S3 Tables), all four transcriptomes are of high quality with values that are either within or slightly higher than those reported in similar studies [24,40,43].

thumbnail
Fig 2. Gene Ontology (GO) distributions by level, by species.

https://doi.org/10.1371/journal.pone.0151059.g002

thumbnail
Fig 3. Functional annotation.

Percentages of BLAST-hits, annotated sequences, mapped sequences, and no-BLAST hits for each species.

https://doi.org/10.1371/journal.pone.0151059.g003

Orthology assignment

We used OMA standalone software [26] to place predicted genes into orthologous groups. OMA’s advantage over standard bidirectional best-hit approaches is that OMA’s algorithm uses evolutionary distances instead of scores, considers distance inference uncertainty, includes many-to-many orthologous relations, and accounts for differential gene losses [26,45]. The number of orthologs obtained for the complete dataset (ten ingroup and three outgroup taxa) was 1,317 and the number of phylogenetically informative genes (n>4) was 11,447. The final concatenated matrix contains a total of 649,095 amino-acid sites present in all taxa.

Functional annotation

The percentage of reads that had at least one BLAST hit varied between 50% and 55.8% of total transcripts (Table 2), which is slightly higher than in similar transcriptome studies [40,43]. Final number of annotated sequences varied between 21,555 and 26,808 sequences. Our data were annotated with a wide range of functional categories represented on all levels of the Gene Ontology database and were comparable to other similar studies, with no biases toward any particular category (Fig 2, Table 2). For all four transcriptomes the top BLAST hit species was recovered as Acromyrmex echinatior, which is a higher-attine leaf-cutting species, a very close relative of Sericomyrmex, and a close relative of Apterostigma. This was also the species with the largest number of total BLAST hits for the three Sericomyrmex species. For A. megacephala the top BLAST hit species was the non fungus-farming ant Cerapachys biroi.

Phylogeny and dating

Results from phylogenetic analyses of this dataset (Fig 4) are congruent with existing phylogenies, including a monophyletic Sericomyrmex clade that is the sister of Trachymyrmex zeteki, and with A. megacephala as the sister taxon to all other attine ant taxa included in this study. Considering the large number of characters in this dataset, the branches subtending the three Sericomyrmex species are reconstructed as remarkably short in both analyses. This result indicates a very low genetic divergence between Sericomyrmex species. To investigate this further we ran a separate orthology search including only the three Sericomyrmex species, which revealed 4,217 orthologous genes present in all three samples, varying in length from 303 to 10,530 bp. Of those 4,217 loci, 92.71% (3,910 sequences) are 99% to 100% pairwise identical, as defined by Geneious [29]. Of the remaining orthologs, 290 are between 90% and 99% pairwise identical. Combining these results, 99.59% of all orthologs shared by these three Sericomyrmex morphospecies are at least 90% pairwise identical. This result is unexpected given that we selected the most molecularly and morphologically divergent Sericomyrmex samples available for RNA extraction. Based on morphology and multiple gene sequences, of the samples chosen for this study S. cf. luederwaldti and S. cf. mayri were considered a priori to be more similar, so the high similarity of their transcriptome sequences is less surprising. However, S. cf. parvulus is morphologically distinctly different from the other two species and is considered to be a member of one of the most basally diverging lineages of Sericomyrmex species (Ješovnik, unpublished). Even though phylogenetic analyses of the transcriptome data agreed with prior analyses regarding these relationships, i.e., that S. cf. luederwaldti and S. cf. mayri are more closely related to each other than either is to S. cf. parvulus, the degree of divergence separating S. cf. parvulus from those two species is much lower than expected based on divergences separating similarly related taxa, e.g., some leaf-cutter species. In general, the transcriptomes indicate that even morphologically divergent Sericomyrmex species are separated by remarkably low genetic distances, which means that accurately recovering species boundaries may require more variable data from non-coding regions. Importantly, this pattern, also supported by our divergence-dating analyses (see next paragraph) indicates that Sericomyrmex has rapidly radiated into multiple, morphologically distinguishable species that occupy a broad geographic distribution [4,46] with only a small degree of accompanying genetic divergence, in contrast to most other comparable ant groups, including its sister taxon, the Trachymyrmex iheringi species group. This conclusion must obviously be investigated further, given that transcriptomes from only three Sericomyrmex species were included in this study. It will be particularly important to test whether this rapid radiation may have been driven by coevolution with a specialized clade of higher attine fungi, as has been suggested for Atta [2,47], especially since recent research suggests the possibility of high symbiont fidelity between Sericomyrmex species and a single species of fungal cultivar [48]. Another potentially important factor in this radiation could be major changes in genetic architecture such as chromosome duplications and rearrangements, because data from two species indicate that, compared to most other attine ants, Sericomyrmex has an unusually high number of chromosomes and an unusually large genome [49,50].

thumbnail
Fig 4. Time-calibrated phylogeny of attine ants.

Highlighted taxa are novel transcriptomes obtained for this study. This topology was recovered in both ML and Bayesian analysis. Asterisks (*) at nodes indicate bootstrap values of 100 and posterior probabilities of 1.0. Green error bars at nodes indicate minimum and maximum age estimates; the time scale at the bottom is in million of years.

https://doi.org/10.1371/journal.pone.0151059.g004

The divergence dating analysis, which included ant species only (the outgroup A. mellifera was excluded in the dating analysis), recovered the stem-group age (i.e., the earliest possible origin) of fungus-farming ants as 66.7 million years ago (mya) and the crown-group age (i.e., the latest possible origin) as 53.6 mya. The crown-group age estimate is consistent with previous estimates of 50–60 mya [2,9,35,50,51], whereas the stem-group age estimate is slightly younger than a stem-group estimate from a recent study (74.6 mya) of a large genomic data set [12]. However, these results are in some cases imprecise and are not directly comparable because neither included taxa from the closest non-fungus-farming relatives (the sister group) of the attine ants [35] and only this study included representatives of both clades resulting from the basal-most divergence at the origin of the fungus-farming ants. Our age estimate (33.3–53.6 mya) of this divergence between the Paleoattini (including the genera Myrmicocrypta, Apterostigma, and Mycocepurus) and the Neoattini (including all other attine ant genera) should be interpreted with caution because our taxon sampling is arguably inadequate for the task. Our taxon sampling for the higher attine ants is better and represents the first genomic data set that includes Sericomyrmex. We estimate the crown-group age of the higher attine ants to be 29.2 (31.3–27.4) mya; the crown-group age of the Sericomyrmex+Trachymyrmex zeteki clade (i.e., the Trachymyrmex iheringi group) as 22.6 (25.3–20.4) mya; and the crown-group age of the leaf-cutting ants (genera Atta and Acromyrmex) to be 17.9 (20.4–15.6) mya (Table 3). For comparison, Schultz and Brady [2] recovered ~16 mya, ~10.5, and ~8 mya mya for those same nodes, respectively, and Nygaard and colleagues estimate ~17 mya for the leaf-cutter origin [12]. Our dating analyses, based on transcriptome data, are most likely overestimating divergence times due to the very small taxon sample. This possible overestimation notwithstanding, the crown-group age of Sericomyrmex is reconstructed as very recent, estimated at 4.8 mya for the age of the ancestor of all extant species. For comparison, the stem- and crown-group ages of the clade containing the attine host/parasite species pair Mycocepurus goeldii and M. castrator are estimated to be 2.04 and 3.31 million years, respectively [52], and crown-group divergences between sister species pairs in the Cyphomyrmex wheeleri group range from 5.3 to 7.0 mya [3]. The branch separating Sericomyrmex from the most recent common ancestor it shares with the Trachymyrmex iheringi group (represented here by T. zeteki) is very long (Fig 4), which might be explained by taxon undersampling, especially if there are more extant Sericomyrmex species that could split that long branch if they had been included. We consider this unlikely, however, because the three sampled Sericomyrmex species were chosen to represent the full range of diversity in the genus both morphologically and molecularly and because other phylogenomic studies currently in preparation, employing much denser taxon sampling, corroborate that the genus Sericomyrmex is the sister group of the entire T. iheringi clade, i.e., that the branch subtending Sericomyrmex is unoccupied by other Trachymyrmex species.

thumbnail
Table 3. Crown-group and stem-group age estimates.

In units of million years ago, with standard error given in parentheses. An asterisk (*) indicates a posterior probability of 1.

https://doi.org/10.1371/journal.pone.0151059.t003

Gene family searches

In order to further investigate the sequenced transcriptomes, we chose several genes of interest based on previous ant genome and ant transcriptome studies [10,11,43] and performed separate searches for those genes. In general, the similarity of the results of our gene BLAST searches to those of other attine ant genomic studies serves as a confirmation of the completeness of our transcriptomes.

Arginine metabolic pathway genes.

The first genomic studies of the higher-attine leaf-cutting ant genera Atta and Acromyrmex demonstrated losses of two genes in the arginine anabolic pathway, whereas all other ants for which data are available, including species in the same subfamily (Myrmicinae) as the attine ants, have functional copies of those genes [10,11]. It was suggested that the leaf-cutting ants, or even all higher attine ants, had become obligately dependent on their fungi for arginine. Such a metabolic division of labor is perhaps not surprising in the highly derived higher attine ants, given that their fungi are also obligate symbionts, and similar examples are known from other mutually obligate symbioses [5355]. However, a more recent study, which included one lower attine ant species that is relatively closely related to the higher attines, indicates that those two genes, argininosuccinate synthase and argininosuccinate lyase, were lost prior to the origin of the higher attines [12], begging the question of at what point in attine evolution the ants became obligately dependent on their fungal cultivars for the amino acid arginine.

Our BLAST searches for argininosuccinate synthase and argininosuccinate lyase in the three Sericomyrmex species did not retrieve any hits, which, because Sericomyrmex species are higher attines, was to be expected. However, BLAST searches also failed to find these genes in A. megacephala. A. megacephala is a member of the Paleoattini, one of two sister clades formed by the basal-most divergence in the fungus-farming ant phylogeny. The other sister clade, the Neoattini, contains Sericomyrmex and all previously studied attine ant species, so the absence of these two genes in a paleoattine indicates that the dependence of attine ants on their fungal cultivars for arginine arose in the common ancestor of all fungus-farming ants. This result is unlikely to be due to inadequate transcriptome data for two reasons. First, our Ortholog Hit Ratios are high, and, second, searches for two other enzymes in the arginine metabolic pathway, the catabolic genes arginase and nitric oxide synthase, produced hits in all of our transcriptomes (Table 4). Functional copies of these two genes are likewise present in the genomes of Atta and Acromyrmex [10,11].

thumbnail
Table 4. Gene searches.

Number of gene copies and isoforms per species.

https://doi.org/10.1371/journal.pone.0151059.t004

Hexamerins.

Hexamerins are another protein family that has been associated with the specialized diet of attine ants [10]. Four copies of hexamerins are commonly found in insects, hex 70a, hex 70b, hex 70c, and hex 110, serving as storage proteins during development as well as in the adult stage [56]. In the genome study of the leaf-cutting ant Atta cephalotes, the gene hex 70c was not found [10]. BLAST searches with Apis mellifera hexamerins recovered two to four different hexamerins per ant transcriptome (Table 4), but unfortunately we were not able to distinguish between the different copies. In the transcriptome of S. cf. luederwaldti our search recovered only two copies.

RYamide.

We found no gene sequences for the RYamide protein in the Sericomyrmex and Apterostigma transcriptomes. RYamide proteins are recently discovered and have been found in all insects for which genomes are available [57] except for ants. The functional roles of RYamides are poorly known, but it has been suggested that in ants the absence of RYamide is connected with the differentiation between reproductive and non-reproductive castes [11].

Desaturases

The desaturase gene family plays an important role in the synthesis of cuticular hydrocarbons (CHC), which are one of the key elements of nestmate recognition in social insects. Using the desat1 gene of the ant Formica exsecta [43] as a query sequence, we found five copies of desaturase in A. megacephala and eight copies in each of the three Sericomyrmex species, all with various numbers of isoforms (Table 4). When the found copies were BLASTed back to the NCBI protein database, the highest similarity was found with delta(Δ)11 desaturase genes of other ants. This finding is comparable to that of a similar study of the Atta cephalotes genome, in which six out of eleven identified desaturase-like genes matched Δ11 desaturase [10].

Cytochrome P450.

Cytochrome P450 (CytP450) is a large protein family, members of which are found in enzymatic pathways central to the metabolism of toxic compounds as well as to development and reproduction [58]. We found 31 copies of CytP450 in A. megacephala, and between 38 and 42 copies in the three Sericomyrmex morphospecies (Table 4). These numbers are surprisingly small. The 54 copies of CytP450 identified in Atta cephalotes is regarded as a reduced number in comparison to other ants (136 in Camponotus floridanus and 107 in Harpegnathos saltator) [59]. Like fungus-farming ants, both of the two other insects with known low numbers of CytP450, the honey bee Apis mellifera (with 62 copies) and the body louse Pediculus humanus (with 39 copies) [60,61] have specialized diets and it has been suggested that insects with predictable diets may have a reduced need to metabolize toxins.

Chitinases.

Chitinases, enzymes with chitinolytic activity, play important roles in digestion and moulting in insects. The number of chitinases in Hymenoptera is considered reduced in comparison to that in Drosophila melanogaster [10]. In studies of attine ants, however, chitinases have been shown to have experienced positive selection, presumably because of their importance in the digestion of the chitinous cell walls of their fungal cultivars [12]. We found 10 chitinase-like genes, in different copy numbers, in all three Sericomyrmex species. In A. megacephala we found 9 of them, with CG8460, chitinase-like protein 1, absent. Interestingly, chitinase gene K06A9.b, which is found in A. mellifera and A. cephalotes but is not known in D. melanogaster, was found in all of our transcriptomes. It was beyond the scope of this paper to test for positive selection in the chitinase genes, but we found 13–23 copies of Chitinase 3, for which positive selection has been detected in other attine ants [12]. Details on copy numbers for each of the chitinases for each of the species can be found in S4 Table.

Our transcriptomes were constructed from worker-caste ants only, and some of the identified genes of interest, including the arginine metabolic pathway and CytP450 genes, are expressed at higher levels in the larval stage [10,58]. Their absence in our assemblies could therefore be attributed to insufficient sequencing depth due to low levels of expression. This seems unlikely, however, because, as noted above, our Ortholog Hit Ratios are high and we detected the other genes in the arginine metabolic pathway, arginase and nitric oxide synthase. Rather, our results suggest that the reduction in CytP450 genes and the losses of argininosuccinate synthase and argininosuccinate lyase likely occurred at the origin of fungus-farming ants. An alternative hypothesis, at least for the arginine metabolic pathway genes, is that they were lost twice, once in A. megacephala and once in the ancestor of the higher attine ants. This appears initially plausible because A. megacephala is the only lower attine known to cultivate a higher-attine fungal cultivar and it is clear that A. megacephala secondarily acquired its cultivar during the course of its evolution [9]. Hence, if arginine dependence is specifically associated with higher attine fungal cultivars, then it could have occurred in parallel in the higher attine ants and in A. megacephala. On closer inspection, we consider this hypothesis unlikely because the two arginine anabolic genes are also absent in the lower attine ant Cyphomyrmex costatus, which cultivates a lower-attine fungus. Clearly, genomic data from other species of paleoattine and basally diverging neoattine lineages are needed to better understand the history of genomic evolution in the fungus-farming ants.

Conclusions

Results from our phylogenetic and dating analyses suggest that the genus Sericomyrmex has undergone a very recent, rapid diversification, reflected by short branch lengths and recent divergence dates. Most surprisingly, the overall genetic similarity between the three Sericomyrmex morphospecies is unexpectedly high. It is our hope that these results will inspire further investigation into the genetic mechanisms underlying rapid radiation with little genetic change in Sericomyrmex. In addition to this phenomenon and its implications for speciation, recent research on social parasites and agro-predators makes this genus an exciting model system for studying behavioral ecology, coevolution, and chemical ecology [6264]. The transcriptomes sequenced in this study could provide the foundation for future research in gene discovery, phylogenomics, population genomics, and conservation genetic studies [65,66]. In our analyses of gene families (arginine metabolic pathway, Cytochrome P450, hexamerins, RYamide, desaturase, and chitinase) we failed to find two arginine anabolic enzymes in the paleoattine ant A. megacephala, suggesting that the loss of these enzymes, confirmed previously only in the clade Neoattini, was likely lost at the origin of the fungus-farming ants. We recovered a surprisingly small number of CytP450 genes, which are associated with detoxification and unspecialized diets. Additional research is required to confirm these gene losses, since our data do not include larval transcripts, and therefore cannot be interpreted as conclusive. Finally, we hope the transcriptome of A. megacephala will prove a useful tool for future research on the underlying genetics that makes this species the only lower attine ant known to cultivate a higher attine fungus [9].

Supporting Information

S1 Fig. Ortholog Hit Ratio graphs.

Ortholog Hit Ratio values for each of the species sequenced.

https://doi.org/10.1371/journal.pone.0151059.s001

(PDF)

S1 Table. PhyloBayes dating analysis calibrations.

Calibration priors used in PhyloBayes divergence dating analysis. Unit: Million of years ago.

https://doi.org/10.1371/journal.pone.0151059.s002

(XLSX)

S2 Table. GO distributions by level.

GO distributions by level, by species, with GO-ID numbers, the output of BLASTtoGO Annotation analyses.

https://doi.org/10.1371/journal.pone.0151059.s003

(XLSX)

S3 Table. Enzyme code distribution.

Number of sequences annotated with different Enzyme code per species, the output of BLASTtoGO Annotation analyses.

https://doi.org/10.1371/journal.pone.0151059.s004

(XLSX)

S4 Table. Chitinase genes.

Number of chitinase genes copies in transcriptomes, for each gene searched, for each species.

https://doi.org/10.1371/journal.pone.0151059.s005

(XLSX)

S1 Text. Gene searches.

The text file with the DNA sequences used in gene searches, with GenBank accession numbers included.

https://doi.org/10.1371/journal.pone.0151059.s006

(TXT)

Acknowledgments

We would like to thank Matt Kweskin, Ben Ruben, Rebecca Dikow, Bastian Bentlage, and Ehsan Kayal for help with bioinformatic analyses. Heraldo Vasconcelos, F. Drummond Martins, and E. Esteves helped with permits and logistics of fieldwork in Brazil; and Carlos Salazar Paz, Antonio Morizaki Taura, Gustavo Suarez de Freitas, Ernesto John Flórez Leiva, Alex Cruz, Katheryn Sarmiento Canales, Ella Karina Ramírez Cuadros, and Frank Azorsa helped with permits for Peru. Jeffrey Sosa-Calvo collected colonies of Apterostigma megacephala. Sanne Nygaard, Koos Boomsma, and Guojie Zhang shared their (at the time, unpublished) data and helped with advice. Kim Mitter helped with RNA extraction, and Michael Lloyd with figure preparation. Nick Silverson, Michael Lloyd, Nicole Gerardo, and two anonymous reviewers made valuable comments that greatly improved the manuscript.

Author Contributions

  1. Conceived and designed the experiments: AJ TRS.
  2. Performed the experiments: AJ.
  3. Analyzed the data: AJ VLG.
  4. Contributed reagents/materials/analysis tools: TRS.
  5. Wrote the paper: AJ VLG TRS.
  6. Field work: AJ TRS.

References

  1. 1. Wirth R, Herz H, Ryel RJ, Beyschlag W, Holldobler B. Herbivory of leaf-cutting ants: a case study on Atta colombica in the tropical rainforest of Panama. Ecol Stud. 2003;164:230.
  2. 2. Schultz TR, Brady SG. Major evolutionary transitions in ant agriculture. Proc Natl Acad Sci U S A. 2008;105(14):5435–40. pmid:18362345
  3. 3. Mehdiabadi NJ, Mueller UG, Brady SG, Himler AG, Schultz TR. Symbiont fidelity and the origin of species in fungus-growing ants. Nat Commun. 2012 May 15;3:840. pmid:22588302
  4. 4. Mayhe-Nunes Antonio, Jaffe K. On the biogeografy of Attini (Hymenoptera: Formicidae). Ecotropicos. 1998;11(1):45–54.
  5. 5. Bolton B. An online catalog of the ants of the world. 2014. Available: http://antcat.org. Accessed 3 October 2015.
  6. 6. Wheeler WM. A new guest-ant and other new Formicidae from Barro Colorado Island, Panama. Biol Bull. 1925;49(3):150–84.
  7. 7. Weber NA. A ten-year colony of Sericomyrmex urichi (Hymenoptera: Formicidae). Ann Entomol Soc Am. 1976;69(5):815–9.
  8. 8. Lattke J. Revision del genero Apterostigma Mayr (Hymenoptera: Formicidae). Arq Zool. 1997;34(5):121–221.
  9. 9. Schultz TR, Sosa-Calvo J, Brady SG, Lopes CT, Mueller UG, Bacci M, et al. The Most Relictual Fungus-Farming Ant Species Cultivates the Most Recently Evolved and Highly Domesticated Fungal Symbiont Species. Am Nat. 2015 May;185(5):693–703. pmid:25905511
  10. 10. Suen G, Teiling C, Li L, Holt C, Abouheif E, Bornberg-Bauer E, et al. The Genome Sequence of the Leaf-Cutter Ant Atta cephalotes Reveals Insights into Its Obligate Symbiotic Lifestyle. PLOS Genet. 2011;7(2):e1002007. pmid:21347285
  11. 11. Nygaard S, Zhang G, Schiøtt M, Li C, Wurm Y, Hu H, et al. The genome of the leaf-cutting ant Acromyrmex echinatior suggests key adaptations to advanced social life and fungus farming. Genome Res. 2011;21(8):1339–48. pmid:21719571
  12. 12. Nygaard S, Hu H, Li C, Schiøtt M, Chen Z, Yang Z, et al. Reciprocal genomic evolution in the ant-fungus agricultural symbiosis. *Nat Commun. 2016.
  13. 13. Smith CR, Smith CD, Robertson HM, Helmkampf M, Zimin A, Yandell M, et al. Draft genome of the red harvester ant Pogonomyrmex barbatus. Proc Natl Acad Sci U S A. 2011;108:5667–72. pmid:21282651
  14. 14. Wurm Y, Wang J, Riba-Grognuz O, Corona M, Nygaard S, Hunt BG, et al. The genome of the fire ant Solenopsis invicta. Proc Natl Acad Sci U S A. 2011;108:5679–84. pmid:21282665
  15. 15. Elsik CG, Worley KC, Bennett AK, Beye M, Camara F, Childers CP, et al. Finding the missing honey bee genes: lessons learned from a genome upgrade. BMC Genomics. 2014;15:86. pmid:24479613
  16. 16. Munoz-Torres MC, Reese JT, Childers CP, Bennett AK, Sundaram JP, Childs KL, et al. Hymenoptera Genome Database: integrated community resources for insect species of the order Hymenoptera. Nucleic Acids Res. 2011;39:658–62.
  17. 17. Andrews S. FastQC: A Quality Control tool for High Throughput Sequence Data. 2010. Available: http://www.bioinformatics.babraham.ac.uk/projects/fastqc
  18. 18. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;1–7.
  19. 19. Grabherr MG, Haas BJ, Yassour M, Levin JZ, Thompson DA, Amit I, et al. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat Biotechnol. 2011;29:644–52. pmid:21572440
  20. 20. Haas BJ, Papanicolaou A, Yassour M, Grabherr M, Philip D, Bowden J, et al. De novo transcript sequence reconstruction from RNA-Seq: reference generation and analysis with Trinity. Nat Protoc. 2013;8:1–43.
  21. 21. Haas BJ, Papanicolaou A, Yassour M, Grabherr M, Blood PD, Bowden J, et al. De novo transcript sequence reconstruction from RNA-seq using the Trinity platform for reference generation and analysis. Nat Protoc. 2013;8:1494–512. pmid:23845962
  22. 22. Fu L, Niu B, Zhu Z, Wu S, Li W. CD-HIT: Accelerated for clustering the next-generation sequencing data. Bioinformatics. 2012;28:3150–2. pmid:23060610
  23. 23. Python Software Foundation. Available: http://www.python.org
  24. 24. Riesgo A, Andrade SCS, Sharma PP, Novo M, Pérez-Porro AR, Vahtera V, et al. Comparative description of ten transcriptomes of newly sequenced invertebrates and efficiency estimation of genomic sampling in non-model taxa. Front Zool. 2012;9(1):33. pmid:23190771
  25. 25. Roth ACJ, Gonnet GH, Dessimoz C. Algorithm of OMA for large-scale orthology inference. BMC Bioinformatics. 2008;9:518. pmid:19055798
  26. 26. Altenhoff AM, Schneider A, Gonnet GH, Dessimoz C. OMA 2011: Orthology inference among 1000 complete genomes. Nucleic Acids Res. 2011;39.
  27. 27. Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol Biol Evol. 2013;30:772–80. pmid:23329690
  28. 28. Castresana J. Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Mol Biol Evol. 2000;17:540–52. pmid:10742046
  29. 29. Kearse M, Moir R, Wilson A, Stones-Havas S, Cheung M, Sturrock S, et al. Geneious Basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics. 2012;28(12):1647–9. pmid:22543367
  30. 30. Zoller S, Schneider A. Improving phylogenetic inference with a semiempirical amino acid substitution model. Mol Biol Evol. 2013;30(2):469–79. pmid:23002090
  31. 31. Guindon S, Lethiec F, Duroux P, Gascuel O. PHYML Online—A web server for fast maximum likelihood-based phylogenetic inference. Nucleic Acids Res. 2005;33.
  32. 32. Lartillot N, Rodrigue N, Stubbs D, Richer J. PhyloBayes MPI: phylogenetic reconstruction with infinite mixtures of profiles in a parallel environment. Syst Biol. 2013;62(4):611–5. pmid:23564032
  33. 33. Lepage T, Bryant D, Philippe H, Lartillot N. A general comparison of relaxed molecular clock models. Mol Biol Evol. 2007;24(12):2669–80. pmid:17890241
  34. 34. Rehm P, Borner J, Meusemann K, von Reumont BM, Simon S, Hadrys H, et al. Dating the arthropod tree based on large-scale transcriptome data. Mol Phylogenet Evol. 2011;61(3):880–7. pmid:21945788
  35. 35. Ward PS, Brady SG, Fisher BL, Schultz TR. The evolution of myrmicine ants: Phylogeny and biogeography of a hyperdiverse ant clade (Hymenoptera: Formicidae). Syst Entomol. 2015;40(1):61–81.
  36. 36. Yang Z, Rannala B. Bayesian estimation of species divergence times under a molecular clock using multiple fossil calibrations with soft bounds. Mol Biol Evol. 2006;23:212–26. pmid:16177230
  37. 37. Brady SG, Schultz TR, Fisher BL, Ward PS. Evaluating alternative hypotheses for the early evolution and diversification of ants. Proc Natl Acad Sci U S A. 2006;103:18172–7. pmid:17079492
  38. 38. Johnson BR, Borowiec ML, Chiu JC, Lee EK, Atallah J, Ward PS. Phylogenomics Resolves Evolutionary Relationships among Ants, Bees, and Wasps. Curr Biol. 2013;23(20):1–5.
  39. 39. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. Basic local alignment search tool. J Mol Biol. 1990;215(3):403–10. pmid:2231712
  40. 40. Ewen-Campen B, Shaner N, Panfilio KA, Suzuki Y, Roth S, Extavour CG. The maternal and early embryonic transcriptome of the milkweed bug Oncopeltus fasciatus. BMC Genomics. BioMed Central Ltd; 2011;12(1):61.
  41. 41. O’Neil ST, Emrich SJ. Assessing De Novo transcriptome assembly metrics for consistency and utility. BMC Genomics. 2013;14:465. pmid:23837739
  42. 42. O’Neil ST, Dzurisin JDK, Carmichael RD, Lobo NF, Emrich SJ, Hellmann JJ. Population-level transcriptome sequencing of nonmodel organisms Erynnis propertius and Papilio zelicaon. BMC Genomics. 2010;11:310. pmid:20478048
  43. 43. Badouin H, Belkhir K, Gregson E, Galindo J, Sundström L, Martin SJ, et al. Transcriptome characterisation of the ant Formica exsecta with new insights into the evolution of desaturase genes in social hymenoptera. PLOS One. 2013;8(7).
  44. 44. Camacho C, Coulouris G, Avagyan V, Ma N, Papadopoulos J, Bealer K, et al. BLAST+: architecture and applications. BMC Bioinformatics. 2009;10(1):1–9.
  45. 45. Dalquen DA, Altenhoff AM, Gonnet GH, Dessimoz C. The Impact of Gene Duplication, Insertion, Deletion, Lateral Gene Transfer and Sequencing Error on Orthology Inference: A Simulation Study. PLOS One. 2013;8(2):e56925. pmid:23451112
  46. 46. Sánchez-Peña SR. Some Fungus-Growing Ants (Hymenoptera: Formicidae) from Northeastern Mexico. Florida Entomol. 2010;93:501–4.
  47. 47. Mueller UG, Scott JJ, Ishak HD, Cooper M, Rodrigues A. Monoculture of leafcutter ant gardens. PLOS One. 2010;5(9):1–7.
  48. 48. De Fine Licht HH, Boomsma JJ. Variable interaction specificity and symbiont performance in Panamanian Trachymyrmex and Sericomyrmex fungus-growing ants. BMC Evol Biol. 2014;14(1):244.
  49. 49. Tsutsui ND, Suarez AV, Spagna JC, Johnston JS. The evolution of genome size in ants. BMC Evol Biol. 2008;8(64):1–9.
  50. 50. Cardoso DC, das Graças Pompolo S, Cristiano MP, Tavares MG. The Role of Fusion in Ant Chromosome Evolution: Insights from Cytogenetic Analysis Using a Molecular Phylogenetic Approach in the Genus Mycetophylax. PLOS One. 2014;9(1):e87473. pmid:24489918
  51. 51. Rabeling C, Gonzales O, Schultz TR, Bacci M, Garcia MVB, Verhaagh M, et al. Cryptic sexual populations account for genetic diversity and ecological success in a widely distributed, asexual fungus-growing ant. Proc Natl Acad Sci U S A. 2011;108(30):12366–71. pmid:21768368
  52. 52. Rabeling C, Schultz TR, Pierce NE, Bacci M. A social parasite evolved reproductive isolation from its fungus-growing ant host in sympatry. Curr Biol. 2014;24(17):2047–52. pmid:25155509
  53. 53. Graf J, Ruby EG. Host-derived amino acids support the proliferation of symbiotic bacteria. Proc Natl Acad Sci. 1998;95(4):1818–22. pmid:9465100
  54. 54. Klasson L. Evolution of minimal-gene-sets in host-dependent bacteria. Trends Microbiol. 2004;12(1):37–43. pmid:14700550
  55. 55. Feldhaar H, Straka J, Krischke M, Berthold K, Stoll S, Mueller MJ, et al. Nutritional upgrading for omnivorous carpenter ants by the endosymbiont Blochmannia. BMC Biol. 2007;5(48).
  56. 56. Martins JR, Nunes FMF, Cristino AS, Simões ZLP, Bitondi MMG. The four hexamerin genes in the honey bee: structure, molecular evolution and function deduced from expression patterns in queens, workers and drones. BMC Mol Biol. 2010;11:23. pmid:20346164
  57. 57. Hauser F, Neupert S, Williamson M, Predel R, Tanaka Y, Grimmelikhuijzen CJP. Genomics and peptidomics of neuropeptides and protein hormones present in the parasitic wasp Nasonia vitripennis. J Proteome Res. 2010;9(10):5296–310. pmid:20695486
  58. 58. Feyereisen R. Insect P450 Enzymes. Annu Rev Entomol. 1999;44(1):507–33.
  59. 59. Bonasio R, Zhang G, Ye C, Mutti NS, Fang X, Qin N, et al. Genomic comparison of the ants Camponotus floridanus and Harpegnathos saltator. Science. 2010;329(5995):1068–71. pmid:20798317
  60. 60. Claudianos C, Ranson H, Johnson RM, Biswas S, Schuler MA, Berenbaum MR, et al. A deficit of detoxification enzymes: pesticide sensitivity and environmental response in the honeybee. Insect Mol Biol. 2006;15(5):615–36. pmid:17069637
  61. 61. Kirkness EF, Haas BJ, Sun W, Braig HR, Perotti MA, Clark JM, et al. Genome sequences of the human body louse and its primary endosymbiont provide insights into the permanent parasitic lifestyle. Proc Natl Acad Sci U S A. 2010;107(27):12168–73. pmid:20566863
  62. 62. Adams RMM, Jones TH, Jeter AW, De Fine Licht HH, Schultz TR, Nash DR. A comparative study of exocrine gland chemistry in Trachymyrmex and Sericomyrmex fungus-growing ants. Biochem Syst Ecol. 2012;40:91–7.
  63. 63. Adams RMM, Liberti J, Illum AA, Jones TH, Nash DR, Boomsma JJ. Chemically armed mercenary ants protect fungus-farming societies. Proc Natl Acad Sci U S A. 2013;110:15752–7. pmid:24019482
  64. 64. Bruner G, Wcislo WT, Fernández-Marín H. Prudent inquilines and proactive hosts: behavioral dynamics between an ant social parasite, Megalomyrmex symmetochus and its fungus-growing ant host, Sericomyrmex amabilis. Insectes Soc. 2013;61(1):83–8.
  65. 65. Peters RS, Meusemann K, Petersen M, Mayer C, Wilbrandt J, Ziesmann T, et al. The evolutionary history of holometabolous insects inferred from transcriptome-based phylogeny and comprehensive morphological data. BMC Evol Biol. 2014;14(1):52. pmid:24646345
  66. 66. Romiguier J, Lourenco J, Gayral P, Faivre N, Weinert LA, Ravel S, et al. Population genomics of eusocial insects: the costs of a vertebrate-like effective population size. J Evol Biol. 2014;27(3):593–603. pmid:26227898