Skip to main content
Advertisement
  • Loading metrics

Genome evolution and divergence in cis-regulatory architecture is associated with condition-responsive development in horned dung beetles

Abstract

Phenotypic plasticity is thought to be an important driver of diversification and adaptation to environmental variation, yet the genomic mechanisms mediating plastic trait development and evolution remain poorly understood. The Scarabaeinae, or true dung beetles, are a species-rich clade of insects recognized for their highly diversified nutrition-responsive development including that of cephalic horns—evolutionarily novel, secondary sexual weapons that exhibit remarkable intra- and interspecific variation. Here, we investigate the evolutionary basis for horns as well as other key dung beetle traits via comparative genomic and developmental assays. We begin by presenting chromosome-level genome assemblies of three dung beetle species in the tribe Onthophagini (> 2500 extant species) including Onthophagus taurus, O. sagittarius, and Digitonthophagus gazella. Comparing these assemblies to those of seven other species across the order Coleoptera identifies evolutionary changes in coding sequence associated with metabolic regulation of plasticity and metamorphosis. We then contrast chromatin accessibility in developing head horn tissues of high- and low-nutrition O. taurus males and females and identify distinct cis-regulatory architectures underlying nutrition- compared to sex-responsive development, including a large proportion of recently evolved regulatory elements sensitive to horn morph determination. Binding motifs of known and new candidate transcription factors are enriched in these nutrition-responsive open chromatin regions. Our work highlights the importance of chromatin state regulation in mediating the development and evolution of plastic traits, demonstrates gene networks are highly evolvable transducers of environmental and genetic signals, and provides new reference-quality genomes for three species that will bolster future developmental, ecological, and evolutionary studies of this insect group.

Author summary

Phenotypic plasticity is the ability of a single genotype to produce multiple phenotypes in response to environmental variation and thus represents an important evolutionary mechanism for organismal adaptation and diversification. In this study, we investigate the genomic basis for phenotypic plasticity in Onthophagini horned dung beetles, a highly diverse (~2500 species) tribe of beetles renowned for their developmental plasticity. We assemble and annotate chromosome-level genome assemblies for three Onthophagini species and identify examples of gene family and coding sequence evolution potentially associated with nutrition-responsiveness in this insect group. When then compare chromatin conformation underlying head horn development and specification in Onthophagus taurus, a species with dramatic nutrition- and sex-based differences in horn shape. We find chromatin accessibility likely plays a critical role in specifying nutritional and sexual horn dimorphisms in this species. Further, we show nutrition- and sex-responsive horn development are controlled by largely distinct, rather than shared, regulatory architectures, and the acquisition of lineage-specific regulatory elements may have played an outsized role in the evolution of nutrition-responsive development of this trait. Our results highlight the significance of chromatin accessibility and regulatory element activity in the regulation of plastic phenotypes and the highly-evolvable nature of developmental gene networks.

Introduction

Phenotypic plasticity is the capacity of a single genotype to produce multiple phenotypes in response to environmental variation and constitutes a ubiquitous property of multicellular life [1]. Plasticity is thought to be an important driver of adaptation, allowing organisms to maintain high fitness in the face of environmental variability, as well as of diversification via evolutionary changes in the genetic architectures underlying plastic trait formation [2].

The ecological and evolutionary significance of phenotypic plasticity has received much attention, and diverse genes and signal transduction pathways have been identified as important mediators of plastic development across biological systems [3]. In addition to coding sequence, epigenetic modifications such as histone marking are predicted to provide important mechanisms of plastic gene expression regulation (for reviews, see refs [45]). Further, recent quantitative trait locus (QTL) analysis combined with genome editing by CRISPR-Cas9 has begun to establish first causal connections between several cis-regulatory elements and the plastic development of nematode feeding structures [6]. Yet despite these advances, the genomic basis underlying developmental plasticity and its evolution, and in particular the role of the non-coding genome and chromatin architecture in regulating conditional responses in trait formation, remain largely unknown.

One group of animals that exhibit an extreme degree of phenotypic plasticity are the true dung beetles (Scarabaeinae), a hyper-diverse clade (>6000 extant species) [7] found on every continent except Antarctica. The extraordinary evolutionary success of this group is attributable at least in part to their ability to exploit an abundant resource inaccessible to most other insects–dung. For nearly every species, the acquisition and utilization of dung is essential to each aspect of their life history. This includes not only consuming dung as a food source (coprophagy), but also as a resource for larval food provisioning and nest construction, thereby enabling a single offspring to complete development from egg to adult within the confines of an underground brood ball. One key adaptation aiding in this strategy is a highly diversified degree of nutrition-responsive (plastic) development. In the case of dung beetles, nutrition-responsive development is a flexible developmental response to variable and limited larval food quality and quantity, resulting in a wide range of adult body sizes, which in turn has fueled the evolution of alternative, body size-dependent morphological, physiological, and behavioral phenotypes [8]. Accordingly, phenotypic plasticity is predicted to be an evolutionary driver for many dung beetle adaptations. Furthermore, due to their diversity, abundance, pronounced environment-sensitive development, and unique feeding and reproductive traits, dung beetles have thus long served as important models for behavioral (e.g. status dependent selection and sperm competition models [9,10]), developmental (e.g. mechanisms of plasticity [11,12]), evolutionary (e.g. the origins of evolutionary novelties [13]), and ecological studies (e.g. meta-population theory [14], nutrient recycling, soil aeration [15,16]). However, despite the significance of dung beetles in both basic and applied science, a reference-quality genomic resource for any member of this insect group has so far been lacking and thus limited studies examining the genetic basis for dung beetle traits and adaptations.

Among the most conspicuous morphological trait of dung beetles are head horns—novel, highly diversified secondary sexual weapons used in reproductive competition [17]. Horns vary tremendously in shape, size, and number across and within species, mediate widespread sexual dimorphisms, and exhibit a high degree of nutrition-responsive development among conspecific males [18]. Most commonly, horn development is limited to, and often exaggerated, in males while females are hornless, though in some species both sexes may develop horns or–on rare occasions–horn development may be sex-reversed and more pronounced in females [19]. Here, nutrition responsiveness of horn formation provides an important axis of diversification, with horn size-body size scaling relationships ranging from isometric to positively allometric (exaggerated) to polyphenic. In polyphenic species, larval nutrition channels male development toward one of two alternate, and discretely different ontogenetic outcomes: fully horned major males (which as adults engage in aggressive combat to secure matings) or smaller-sized, nearly fully hornless minor males (which engage in sneaking tactics and sperm competition) [17,19]. Intriguingly, head horns lack homology to any other appendage or body part [20], and as such qualify as an evolutionary novelty even by the strictest of definitions [21], yet gains, losses, and modifications to horn structure are common among even closely related species [22]. Thus, beetle horns exhibit a high degree of evolutionary lability and represent a powerful natural system for understanding how complex traits originate and diversify [12,13,20,22].

In this study, we begin by assembling and annotating chromosome-level genome assemblies for three Onthophagine dung beetle species: Onthophagus taurus, Onthophagus sagittarius, and Digitonthophagus gazella (Fig 1A). The Onthophagini is a highly speciose (~ 2500 extant species) tribe of dung beetles found worldwide and includes one of the most species-rich genera on Earth, Onthophagus [23]. Through a series of whole genome alignments and comparative genomic assays including other published beetle genomes, we identify sequence gains and changes putatively associated with adaptations unique to dung beetle ecology and evolution.

thumbnail
Fig 1. Chromosome-level, reference genome assemblies for three dung beetle species.

(A) Images of adult male and female dung beetle individuals representing the focal species of this study: Onthophagus taurus, Onthophagus sagittarius, Digitonthophagus gazella. Scale bar: 10 mm. (B) Intra- and interspecific head horn morphologies of these dung beetle species, including a striking nutrition-dependent male polyphenism in O. taurus and sex-reversed posterior horn development in O. sagittarius. (C) Link plot illustrating synteny relationships between genes of chromosome-length scaffolds of these three dung beetle species and the more distantly related model beetle species, Tribolium castaneum. Genome composition and summary statistics for each dung beetle species are provided on the right.

https://doi.org/10.1371/journal.pgen.1011165.g001

We then utilize these three reference genomes to investigate the role of the non-coding genome in the development and evolution of sex- and nutrition-dependent horn formation across our three focal species. We chose these three species because collectively they embody a remarkable diversity in head horn development reflective of much of the diversity contained within the clade, including: a highly exaggerated sexual dimorphism and male polyphenism in O. taurus; a rare example of sex-reversed horn development in O. sagittarius females paralleled by a secondary loss of horn polyphenism in males; and a modest sexual dimorphism and male polyphenism in the more distantly related D. gazella thought to reflect ancestral character states (Fig 1B). Specifically, we apply genome-wide chromatin accessibility assays to investigate the cis-regulatory basis for sex- and nutrition-biased horn formation in O. taurus and provide some of the first evolutionary insights into the chromatin architecture underlying nutrition-dependent morph determination and developmental plasticity broadly. Our approach and results open new ways of understanding how gene networks may evolve to generate novel structures and regulate environment-responsive development.

Results and discussion

Assembly of reference dung beetle genome assemblies

Chromosome-level scaffolds were constructed for each beetle species using a two-step assembly strategy: PacBio HiFi long read contig assembly followed by HiC proximity ligation sequencing. Importantly, each species’ assembly was generated and annotated using identical sequencing and bioinformatic pipelines (see Methods and Materials for details), minimizing technical bias during cross-species comparisons. The O. taurus assembly is 290.0 Mb in length and is composed of 53.3% repetitive elements (Fig 1C). Interestingly, the closely related O. sagittarius genome assembly is nearly twice as large (553.3 Mb), which is almost entirely attributable to expansion of non-coding, primarily repetitive (69.1% of assembly), sequence. The D. gazella assembly, on the other hand, revealed only a modest increase in genome size (327.3 Mb) relative to O. taurus. High contiguity and BUSCO single copy scores (97.1–98.9%) alongside low BUSCO duplication scores (0.4–1.0%) suggest accurate reference genome assemblies were achieved for each species (Fig 1C). The O. taurus, O. sagittarius, and D. gazella genomes have an estimated 18,389, 19,766, and 18,889 number of gene models, respectively, which is in the range of other published beetle genome assemblies compared in this study (Fig 2A) (12,873–23,987). At a genome-wide scale, gene composition is also comparable in terms of orthogroup membership (a set of orthologous genes with shared ancestry across species, see ref. [24]), including both the proportion of species-specific orthogroups (Fig 2B) and number of genes per orthogroup (Fig 2C). Therefore, we next compared individual orthogroup size and nucleotide sequence across ten beetle species to detect putatively adaptive changes to dung beetle coding sequence.

thumbnail
Fig 2. Comparative genomics across the Coleoptera identify dung beetle-specific evolutionary changes in gene content.

(A) Phylogenetic relationships of ten beetle species included in comparative genomic analyses. We focus on five phylogenetic positions (colored dots) for identifying putative changes in gene content associated with dung beetle-specific traits: the species tips of 1) O. taurus, 2) O. sagittarius, and 3) D. gazella, 4) the ancestral node of O. taurus and O. sagittarius, and 5) the ancestral node of all three dung beetles species analyzed here. (B) Proportion of each dung beetles’ gene set belonging to different orthogroup classifications. “Conserved” denotes orthogroups that include at least one gene from each of the ten beetle species included in the study. “Other” denotes orthogroups that have representation from some, but not all, of the ten beetle species. Onthophagine-ancestor, Onthophagus-specific, and lineage specific orthogroups are those with genes exclusively in 1) the three dung beetle species, 2) O. taurus and O. sagittarius, and 3) only one of the dung beetle species, respectively. (C) Average number of genes identified in each dung beetle species’ genome per orthogroup type. (D) Rapidly evolving putative gene families (orthogroups) at each phylogenetic position of interest potentially associated with dung beetle adaptions. This includes gene families with significant expansions or contractions relative to the rest of the tree, which could occur in multiple lineages (stacked bars on left) or at a single phylogenetic position (single bars on right). Circle size denotes relative size of gene family and each bar on top corresponds to one circle within the plot. Orange inset lists orthogroups with some of the largest expansions at the estimated Onthophagine node, which we predict includes evolutionary changes to gene family size most representative of dung beetle-specific adaptations. (E) Number of genes with evidence of episodic diversifying (positive) selection at each phylogenetic position of interest but no evidence of selection at any non-dung beetle node. Colored bars denote genes with evidence of positive selection detected exclusively at that phylogenetic position, whereas gray bars denote genes with evidence of positive selection within multiple dung beetle lineages.

https://doi.org/10.1371/journal.pgen.1011165.g002

Coding sequence evolution associated with dung beetle-specific adaptations

In total, we identified 19,407 orthogroups across the three Onthophagini species reported here and seven other coleopterans for which chromosome-scale resolution genomes (or close to) where available (Fig 2A). These seven non-dung beetle species represent a diverse set of feeding ecologies (e.g. predation: Photinus pyralis (firefly); phytophagy: Leptinotarsa decemlineata (Colorado potato beetle); wood feeding: Anoplophora glabripennis (Asian long-horned beetle)) and a range of phylogenetic distances relative to the Scarabaeinae. From these orthogroups, we identified evolutionary changes in gene content in dung beetles relative to species in other beetle families in two ways: 1) examination of rapidly evolving gene families within multi-copy orthogroups (14,414 orthogroups) using CAFE [25] and 2) testing for evidence of episodic diversifying selection among single copy orthologs (2,948 orthogroups tested) with BUSTED [26]. While gene family size changes and non-synonymous mutations are not necessarily indicative of adaptation, these analyses provide predictions of candidate genes and pathways for further assessment. We highlight evolutionary changes at five phylogenetic positions (Fig 2A): each beetle species tip, the ancestral node of O. taurus and O. sagittarius, and the ancestral node of all three dung beetle species examined here (Ot-Os-Dg node).

For each dung beetle species, we identified ~200–300 rapidly evolving (mostly expanded) orthogroups as well 110 inferred gains and losses at the Ot-Os-Dg node (Fig 2D and S2 Data). Here, we call rapidly evolving orthogroups as those with significant gains or losses at a dung beetle position but no significant changes (in the same direction) at any non-dung beetle point in the phylogeny. While rapidly evolving gene families detected at the tips of the phylogeny may represent species-specific copy number changes or adaptations, we focused first on changes inferred at the Ot-Os-Dg node (orange) which we predict may most closely represent changes in gene family size associated with the Onthophagine ancestor and dung beetle biology. Among these 110 rapidly evolving orthogroups at the Ot-Os-Dg node are gene families involved in metabolism, metamorphosis regulation, and odorant receptor genes (Fig 2D inset). Specifically, among the largest gene family expansions are genes involved in juvenile hormone (JH) production (JH acid O-methyltransferase and JH esterase), in line with earlier work suggesting this hormone may be involved in regulating male dimorphisms in O. taurus [27] and the broadly described importance of this pathway in insect development and metamorphosis [28]. We also identified orthogroup expansions of genes involved in lipid (lipase 3; phospholipase A1-like) and carbohydrate (facilitated trehalose transporters) metabolism, which may be associated with diversification of nutrition-responsive development in dung beetles. Lastly, among the largest gene families within metazoans are those encoding odorant receptor (OR) proteins [29], the diversification of which is predicted to have aided in the evolution of insect terrestriality given their critical function in scent detection [30]. Expansion of the OR 9a family was inferred at the dung beetle ancestral node as well as separate expansions of OR 85d and OR 67c families in O. taurus and O. sagittarius, respectively, which may be associated with the localization and utilization of dung beetles’ unique food resources. However, like all gene family evolution analyses, associations of orthogroup expansions with dung beetle adaptations are speculative, and functional assessment of these genes and pathways will be necessary to validate the predictions mentioned above.

Next, contrasting orthogroup membership gains and losses between O. taurus and O. sagittarius offers a way of identifying rapidly evolving gene families potentially associated with the exaggeration (O. taurus) and secondary loss (O. sagittarius) of nutritional plasticity. We identified 42 orthogroups with significant gene number expansions in O. taurus and corresponding losses in O. sagittarius as well as 30 significantly expanded orthogroups in O. sagittarius that are reduced in O. taurus (S2 Data). Among the most dynamic of these orthogroups is one that includes targets of rapamycin (TOR), for which we observed 9 gene gains in O. taurus and 3 losses in O. sagittarius, paralleling the evolutionary exaggeration and secondary loss of male polyphenism observed in the two species, respectively. TOR proteins are a well characterized group of kinases that play a central role in eukaryotic cell growth and nutrition sensing as a part of critical developmental signaling pathways [31]. Given the correlative nature of this result and the fact that TOR copy number can vary widely among lineages, future comparative functional analyses of this pathway during dung beetle larval development will have to determine what role, if any, TOR protein signaling may play in nutrition-responsive development within and among species.

Lastly, we estimated evidence of gene-wide episodic diversifying (positive) selection in the coding sequence of 2,948 single copy orthogroups at the same five phylogenetic positions as above (Fig 2A) to identify changes in protein sequence possibly associated with dung beetle-specific adaptations. While we found 220–260 genes with evidence of positive selection on each position tested, we did not detect an enrichment of lineage-specific positive selection on any of the five lineages tested relative to other lineages (Fig 2E and S3 Data). However, within each of the dung beetle species reported here, we found evidence of positive selection in numerous developmental transcription factors including six homeobox proteins; for comparison, we detected evidence for positive selection in four homeobox proteins in the other seven non-dung beetle species combined. Taken together, these results document a remarkable number of non-synonymous substitutions in developmental regulatory genes of dung beetles, some of which may play adaptive roles in these species’ development.

Sex and nutrition-responsive regulatory elements underlie intraspecific diversity in beetle horns

Beetle head horns are novel structures, i.e. they lack obvious homology to other body parts, and are considered hotspots for evolutionary diversification due to the extraordinary morphological variation found both within and across species [18]. Previous work has identified critical transcription factors and developmental pathways that play a role in beetle horn development including Doublesex [11], Hedgehog signaling [32], and the insulin transduction pathway [33,34], among many others [35,36]. However, the role of the cis-regulatory elements (CREs) in mediating horn growth and diversification as well as transducing sex- and nutrition-responsive signals into alternative developmental phenotypes via gene network interactions remains poorly understood. To address this void, we measured chromatin accessibility using ATAC-seq (Assay for Transposase-Accessible Chromatin) in the epithelial cells that compose the dorsal, posterior head and thus the location in which large males develop prominent horns, small males develop horn rudiments, and females develop a conspicuous ridge (Fig 1B). We focused on individuals who had just completed the larval to pupal transition and replicated our approach in both males and females reared in high and low nutrition conditions to identify putative regulatory elements involved in sex- and morph-dependent horn morphogenesis.

It total, we identified 68,038 open chromatin regions (OCRs) in dorsal epithelial cells of O. taurus pupae, a portion of which we predict will contain CREs involved in developmental regulation of this body region. At a genome-wide scale, the chromatin landscape of high and low nutrition females is nearly indistinguishable and only 25 OCRs (0.037%) are significantly differentially accessible between nutritional conditions in this sex (S4 Data). This result may not be surprising given, like most dung beetle species, O. taurus females do not develop head horns or any other nutrition-responsive dorsal head phenotype. Accordingly, we grouped high and low nutrition female samples into a single female sample group (seven biological replicates) for downstream accessibility comparisons.

In stark contrast, we found 401 differentially accessible OCRs in primordial horn tissue of males reared in high vs. low nutritional conditions, including 250 OCRs more accessible in major males and 151 OCRs more accessible in minor males (Fig 3A: top). We categorize this set of OCRs as putative nutrition-responsive regulatory sites. Interestingly, over half (52.3%) of the 151 OCRs more accessible in minor males are located on the X chromosome (Scaffold 8), suggesting many loci responsible for regulating plastic horn growth are concentrated on this chromosome. Of note, one of the genes with the highest number of nearby nutrition-responsive OCRs is doublesex (Fig 3B: top; 4 OCRs), which previous work has demonstrated to be highly differentially expressed between male morphs [12] and to play an essential role in regulating both nutrition and sex-responsive horn development in O. taurus [11,37]. This concordance lends confidence to the accessibility data presented here and suggests regulation of doublesex expression, and by extension horn plasticity, may be associated with altered binding efficacy of doublesex regulatory factors to nearby CREs. More broadly, genes with nearby male nutrition-responsive OCRs more frequently exhibit differential mRNA expression between primordial horn tissue of male morphs than genes lacking nutrition-responsive OCRs (Fig 4D) (RNA-seq data from ref. [12]), though this association is not supported at a genome-wide scale (chi-square test: p = 0.15). This weak global correlation between OCR accessibility and transcriptomic divergence has been reported in other systems [38,39] and suggests alternative accessibility of OCRs is consequential for differential gene expression regulating plastic horn growth at some, but certainly not all, loci.

thumbnail
Fig 3. Nutrition and sex-responsive open chromatin regions (OCRs) underlie intraspecific head horn variation in O. taurus.

(A) Manhattan plots of OCRs identified in primordial head horn epithelial cells, arranged by differential accessibility comparison: top- major (horned) males vs minor (hornless) males; middle- major (horned) males vs females; bottom- minor (hornless) males vs females. Gray points indicate non-differentially accessible OCRs for each comparison. Colored points indicate OCRs with significantly greater accessibility for each comparison. The red and dotted lines denote a false discovery rate (FDR) of 10% and 5%, respectively. The number of more accessible OCRs in each comparison are also provided at a 10% and 5% FDR (colored and gray numbers, respectively). Line plots on the right show the percentage of all differentially accessible OCRs that are located on each scaffold at the 10% significance threshold. Note the abundance of differentially accessible OCRs on the X chromosome (Scaffold 8) in minor male and female comparisons with major males. (B) Examples of genes with multiple male nutrition-responsive OCRs (top: doublesex) or sex-responsive OCRs (bottom: foxp1) upstream of their start codon (highlighted in gray) potentially involved differential gene expression via altered binding capacity of regulatory factors. Beetle head illustrations courtesy of Erica M. Nadolski.

https://doi.org/10.1371/journal.pgen.1011165.g003

thumbnail
Fig 4. Regulatory architectures underlying head horn shape are largely distinct.

(A) Venn diagram of shared and uniquely differentially accessible head horn OCRs identified in each pairwise comparison of O. taurus individuals. OCRs are detected in each tissue type. (B) Examples of binding motifs of developmental transcription factors significantly enriched in differentially accessible OCRs identified in both major (horned) male comparisons (yellow) or sex-based comparisons (purple). Shared major male OCRs (yellow) are predicted to contain regulatory elements especially important for exaggerated, nutrition-responsive horn development in this species. “% in background” refers to the percentage of OCRs containing a binding motif that are not differentially accessible in major males or between sexes. (C) Proportion of primordial head horn OCRs identified in O. taurus with orthology to loci in nine other beetle species included in this study. The “% shared” notes the percentage of the 68,038 OCRs identified on O. taurus that were also present in the genomes of one of the other beetle species. 3.1% of all OCRs were identified in every beetle species’ genome (conserved set) whereas 23.3% of all OCRs were only detected in the O. taurus genome. (D) Proportion of genes that are differentially expressed between primordial head horn tissues of O. taurus male morphs that do or do not possess at least one nearby (nutrition-responsive) differentially accessible OCR. (E) Proportion of head horn OCRs found only in the O. taurus genome with differential accessibility according to male nutrition (left) or sex (right). Chi-square test results are presented in (D) and (E).

https://doi.org/10.1371/journal.pgen.1011165.g004

The number of male nutrition-responsive OCRs (401) is comparable to the number of differentially accessible OCRs between females and minor males (318) (Fig 3A: bottom). However, these differences are dwarfed by the 2,163 OCRs differentially accessible between females and major males (Fig 3A: middle; see Fig 3B: bottom for example OCRs nearby foxp1). One possible explanation for this observation is a compounding effect of two biological factors shaping the development of this body region: 1) sex and 2) exaggerated nutrition-responsive growth. Like in the comparison of male morphs, a large proportion of OCRs more accessible in females relative to major males are located on the X chromosome (55.7%), but surprisingly this sex chromosome enrichment is absent in the minor male and female comparison (Fig 3A). Thus, somewhat contradictory to expectations, genes and regulatory elements on the X chromosome may play an outsized role in male horn polyphenism development, whereas sex-based differences in the same trait may originate from loci more uniformly distributed across the genome.

Distinct gene regulatory architectures underlie evolution of horn morphology

While horn development is dependent on both sex and nutritional signals, the degree of overlap between these gene networks underlying the evolution of intraspecific horn variation is mostly unknown. Earlier work suggested that sexually dimorphic horn formation may have evolved from sexually monomorphic horn development via the acquisition of mechanisms inhibiting horn development in females [18]. This perspective is supported by the observation that down-regulation of female doublesex isoforms induced horn development in otherwise hornless females [11]. Yet knockdown of doublesex expression mediates intersex phenotypes in diverse organisms and traits, and thus may not have directly driven the evolution of sexually dimorphic horn development in beetles. Earlier work also suggested that hornless-ness in both females and low nutrition males may be due to shared regulatory mechanisms [40], however, subsequent work has failed to support this notion. For example, manipulations of both insulin- and hedgehog signaling induce the formation of large horns in low nutrition males without affecting female dorsal head development [32,34].

In this study, we find that differentially accessible OCRs between developing horn tissue of male morphs and females are largely distinct from one another (Fig 4A), suggesting nutritional plasticity of O. taurus horn development is regulated by gene interactions largely separate from those governing sexual dimorphism. For example, of 318 OCRs differentially accessible between minor males and females, only 25 (7.86%) are also differentially accessible between major and minor males. Numerically, the largest overlap in differentially accessible OCRs is found when major males are compared to both females and minor males (124 OCRs), however, the same comparison identifies 254 differentially accessible OCRs (63.3%) as unique to the major vs minor male comparison and a remarkable 1,992 (92.1%) as unique to the major male vs female comparison (Fig 4A).

Importantly, both nutrition and sex-responsive OCRs are enriched for binding motifs of diverse developmental transcription factors relative to non-differentially accessible OCRs (Fig 4B), suggesting condition-dependent chromatin configuration provides a powerful molecular mechanism by which distinct suites of regulatory factors may exert their functions to determine head horn phenotypes. Similarly, “non-enriched” transcription factors that are known to play a role in specifying beetle head horn development also have putative binding sites within nutrition and sex-responsive OCRs (for a well-characterized example, see S5 Data for list of Doublesex hits). However, for these transcription factors, the proportion of binding sites residing within sex- and/or nutrition-responsive OCRs is not significantly different than the proportion of non-differentially accessible (background) OCRs with putative binding sites. This result may not be surprising given many transcription factors regulate gene expression in a variety of developmental and homeostatic contexts, which would require binding to distinct genomic loci. More generally, our results provide further support for the notion that sex- and morph-specific development are underlain by largely distinct regulatory landscapes and are thus developmentally and evolutionarily decoupled. Future work aimed at identifying which upstream regulatory molecules (e.g. pioneer factors) generate condition-dependent differences in chromatin accessibility will be critical to provide a more mechanistic understanding of the evolution of head horn regulation in this species.

Recently evolved CREs play an outsized role in nutrition-responsive horn development

Most non-coding genomic elements such as cis-regulatory elements are generally subject to lower conservation rates relative to coding sequence across evolutionary time [41,42]. Therefore, we sought to understand what role, if any, recently evolved cis-regulatory elements play in O. taurus head horn development. Here, we are defining recently evolved (putative) regulatory elements as genomic regions (identified as larger OCRs from the ATAC-seq data) found in O. taurus but not detected in any other of the nine beetle genomes surveyed, which total 15,850 (23.3%) of all OCRs identified in this tissue type (Fig 4C) (see Methods and Materials for details on how orthologous regions were identified). By comparison, only 3.1% of all OCRs have an orthologous genomic region in each of the other 9 beetle species. We use the term “recently evolved” in lieu of “novel” when describing these genomic elements as we presently cannot distinguish between 1) truly novel regulatory sequences present only in O. taurus and 2) highly variable regulatory sequences for which synteny can no longer be assigned. Still, regardless of how these genomic elements originated, the relatively short evolutionary time examined (~5 my diverged from O. sagittarius) suggests these putative regulatory regions may be especially pertinent to horn development and evolution in this species.

When comparing chromatin accessibility between male morphs, we found 23.2% of non-differentially accessible OCRs have recently evolved in O. taurus (Fig 4E), which is nearly identical to the proportion of recently evolved elements among all 68,038 OCRs (23.3%, see above). Strikingly, the proportion of recently evolved regulatory elements increases to 32.9% among male nutrition-responsive OCRs in primordial horn epithelial tissue (Fig 4E). This enrichment of recently evolved regulatory elements lies in stark contrast to sex-responsive OCRs, for which only 13.0% of OCRs appear exclusively in the O. taurus genome (Fig 4E). Intriguingly, these recently evolved OCRs are highly enriched for repetitive elements like TEs (77.5% overlap at least one repetitive element) compared to conserved OCRs (26.7%), suggesting TE activity may have played a role in the origination of these putative regulatory regions (S1 Fig).

One explanation for this opposing relationship between nutrition- and sex-responsive OCRs may be attributable to the evolutionary history of the molecular mechanisms underlying these two sets of phenotypes. Sexually dimorphic horn development is common not just in the Onthophagini but the family Scarabaeidae and thus likely originated early in scarab evolution. If so, we would not predict recently evolved regulatory elements to play an especially important role in the sex-biased development of this trait given the core of its developmental regulatory architecture evolved deep in this lineage’s history. In contrast, nutritional plasticity, and in particular the extreme polyphenic development of O. taurus head horns, constitutes a more recently derived trait, whose diversification is more likely to be impacted by the acquisition of lineage-specific regulatory elements and the establishment of novel regulatory interactions these afford–a phenomenon that may be important for the evolution of novel traits in other systems as well.

Conclusions

Nutrition-responsive phenotypic plasticity is a ubiquitous property of living systems and important biological mechanism for adaptation and trait diversification [1,2]. In this study, we first used comparative genomics to identify coding and non-coding changes in dung beetles associated with metabolite and developmental regulation that are potentially underpinning the evolution of nutritionally plastic phenotypes in these species. These biological pathways are involved in lipid and carbohydrate metabolism, juvenile hormone production, and embryonic transcriptional regulation, which collectively are now motivating new hypotheses for how nutrition-responsive development may have evolved in this group of insects.

We then carried out an in-depth genomic and developmental analysis of one of these traits, head horns–secondary sexual weapons with an extreme degree of inter- and intraspecific variation. Chromatin configuration assays suggest alternative forms of intraspecific horn variation (sexual dimorphism vs. nutritional plasticity) are instructed by discrete regulatory architectures, a conclusion corroborated by two independent pieces of support: 1) largely distinct sex- and nutrition-responsive head horn OCRs; and 2) a disproportionate amount of OCRs that are nutrition-, but not sex-, responsive and only found in the O. taurus genome. In other words, accessibility of putative cis-regulatory elements covaries extensively with sex and, in males, nutritional state, of which the latter set is disproportionally composed of recently evolved regulatory elements. Future work comparing the overlap of sex- and nutrition-dependent changes in regulatory element activity of other body regions with varying degrees of sex and nutrition responsiveness will be critical to contextualizing these results and understanding how gene regulatory mechanisms evolve across disparate traits.

One possible explanation for these findings may be linked to the idea that head horns lack obvious homology to other body parts in insects or non-insect hexapods. In contrast to legs, antennae, and mouthparts (which are all serial homologs of ventral appendages) and thoracic horns (which have recently been identified as wing serial homologs [13,20]), head horns appear to be unique elaborations of the dorsal head, and sufficiently individuated to have undergone remarkable diversification in shape, relative size, number, and in part, precise location. Lacking homology to other body parts may have promoted, permitted, or alternatively limited head horn development to the inclusion of novel regulatory interactions into gene networks due to the lack of pre-existing mechanisms that could be co-opted, and/or to minimize developmental constraint arising from pleiotropy. Comparing our findings on head horns to that of other diversified structures with established homology (such as ventral appendages or thoracic beetle horns) alongside other novel structures (such as butterfly wing spots or firefly lanterns) will be informative for understanding if the patterns of cis-regulatory evolution identified here are symptomatic for the early stages of morphological innovation in evolution.

Methods and materials

Genome sequencing and assembly

O. taurus individuals were collected from Chapel Hill, North Carolina, USA (35.936, -79.128), O. sagittarius were collected nearby Kilcoy, Queensland, AU (-26.951, 152.605), and D. gazella individuals were collected nearby Legonyane, South Africa. For each species, F0 individuals from these wild populations were crossed and offspring reared in artificial brood balls to the late pupal stage. Once this developmental stage was reached, a single male individual from each species were flash frozen in liquid nitrogen, stored at -80° C, and shipped overnight to Dovetail Genomics (Scotts Valley, CA, USA) on dry ice for library preparation and DNA sequencing.

A dual sequencing approach was implemented for de-novo genome assembly of each dung beetle species: 1) PacBio HiFi CCS long read sequencing for contig assembly and 2) Dovetail Omni-C proximity ligation sequencing (DNAse I digestion) for building contigs into chromosome-length scaffolds. PacBio coverage and read N50 lengths of (86.6X, 13.8 Kb), (30.8X, 12.2Kb), and (72.4X, 12.8 Kb) were achieved for O. taurus, O. sagittarius, and D. gazella, respectively, on a Sequel II platform. Omni-C libraries were sequenced on an Illumina HiSeqX platform to 30.8 million, 102.2 million, and 31.6 million 150 bp paired-end reads for each species, respectively. Contigs were constructed using the PacBio HiFi reads via hifiasm v. 0.13 [43] under default parameters. Omni-C read pairs were aligned to the contig assembly using bwa [44] and scaffolds were assembled using the HiRise software pipeline [45]. Genome summary statistics are provided in Fig 1C. BUSCO v5.4.2 was implemented to estimate the completeness and redundancy of gene content within each assembly by referencing the “insecta_odb10” lineage dataset.

Gene modelling and annotation

After the final version of each genome assembly was completed, we used RepeatModeler v2.0.1 [46] to create a library of repetitive elements for each species’ genome. This library was then used to mask repetitive elements within each genome assembly with RepeatMasker v.4.1.1 using the most sensitive pre-set parameters (-s), which maximizes detection of repetitive elements at the expense of computational time. Soft-masked genome assemblies were used as input for proceeding gene modeling and annotation analyses.

We implemented the BRAKER2 [47] pipeline to generate an ab-initio, preliminary set of gene and isoform models for each dung beetle species, aided by the inclusion of RNA-seq data from each species12 and the UniRef90 Ecdysozoan protein database (accessed Feb. 2022) [48]. The “-etpmode” pipeline includes a large suite of software including Augustus [49], Genmark-EP+ [50], Genemark-ET [51], DIAMOND [52], and SAMtools [53]. Gene models were assigned putative identification by BLAST-ing each model to the UniRef90 Ecdysozoan protein database and the Onthophagus taurus v. 2.0 protein models (e-value 1e-5).

Genome-wide synteny analysis was carried out using the gene-anchored MCScanX method [54] under default parameters. Pairwise alignments were performed in order of increasing phylogenetic distance starting with O. taurus, illustrated in Fig 1C: Ot-Os, Os-Dg, Dg-Tc following an all-to-all BLASTP similarity search carried out between each pair of species’ protein models (-max_target_seqs 5, -evalue 1e-10). Synteny results from these analyses were then visualized using SynVisio [55].

Comparative genomic analyses across ten Coleopteran species

Comparative genome analyses included the three newly assembled dung beetle genomes presented in this study alongside seven other high-quality beetle genomes representing a diverse set of families across the Coleoptera. These genomes include the common eastern firefly Photinus pyralis [56], emerald ash borer Agrilus planipennis, red flour beetle Tribolium castaneum [57], mountain pine beetle Dendroctonus ponderosae [58], Asian longhorned beetle Anoplophora glabripennis [59], Colorado potato beetle Leptinotarsa decemlineata [60], and the Japanese rhinoceros beetle Trypoxylus dichotomus [61]. For all analyses of gene content between species, only the longest isoform of each species’ gene was used. Protein models of these longest isoforms were input into OrthoFinder v.2.5.4 [24] to identify orthogroups across all ten beetle species, which served as the basis for analysis of rapidly evolving gene families and evidence of positive selection within coding sequence described below. Putative annotations for each orthogroup were assigned by BLAST-ing protein models (e-value < 1e-5) from each orthogroup to the NCBI non-redundant invertebrate peptide database (accessed November 2022) and filtering for significant hits to non-repeating protein descriptions (S1 Data).

Gene family size evolution analyses

Nucleotide alignments of single copy orthogroups identified above were input into IQ-Tree v.2.2.0 [62] to generate an initial beetle species tree using the best fit model “JTT+F+R8” (1000 bootstrap replicates) identified by ModelFinder [63], and a time-calibrated species tree was estimated with r8s [64]. While only 10 taxa were included in this analysis and species tree determination was not a primary objective of this study, each node within this tree had 100% bootstrap support. Furthermore, its topology and estimated branch lengths are congruent with previously published, more in-depth phylogenomic analyses of Coleopteran relationships [65], providing confidence for this tree’s inclusion into gene family evolution analyses. After filtering out exceedingly large orthogroups (5), 14,414 multi-copy orthogroups remained for rapid gene family size evolution analysis. We input these orthogroups and the species tree described above into CAFE v5 [25], while modelling for errors in gene family sizes associated with different assembly and annotation methods across the genomes included here [66]. Rapidly evolving gene families putatively associated with dung beetle specific traits (Fig 2D) are those orthogroups with significant expansions or contractions (p-value < 0.05) at a dung beetle species tip or ancestral node but no significant changes at any other non-dung beetle position (see S2 Data).

Tests for episodic diversifying selection

To test for evidence for episodic diversifying selection within beetle coding sequence, nucleotide sequences of 2,660 single copy orthogroups were aligned across species using MUSCLE v5.1 [67]. Nucleotide alignments were input into BUSTED v.4.0 [26] while accounting for synonymous rate variation in the model to test for gene-wide evidence of positive selection for every orthogroup at six phylogenetic (foreground) positions: 1–3) each dung beetle species tip, 4) ancestral node of O. taurus and O. sagittarius, 5) ancestral node of all three dung beetle species reported here, and 6) any non-dung beetle position. P-values were adjusted using the Bonferroni method for multiple comparisons correction in R. Genes with evidence of positive selection putatively associated with dung beetle adaptation were those with significant evidence of selection (adjusted p-value < 0.05) at a dung beetle position and no evidence of positive selection coding sequence of any non-dung beetle species (S3 Data).

ATAC-seq Sample and Data Preparation

Onthophagus taurus individuals were collected from Chapel Hill, North Carolina (same population as the individual used for genome sequencing) and kept in laboratory conditions as described in Moczek and Nijhout, 2002 at 24°C. Six female and three male adult individuals were bred over the course of one week, after which brood balls collected and 1st instar larvae moved to artificial brood balls made in 12-well plates (an established protocol, detailed in ref. [68]). To simulate high and low nutritional conditions, larvae were fed manure from either grass-fed cows (high nutrition) or hay-fed cows (low nutrition) over the course of larval development (method detailed in ref. [69]). Plates were checked every 12 hours for pupating individuals. Distinguishing high and low nutrition males of O. taurus pupae is straightforward, as the fully developed horns of large (major) and horn rudiments of small (minor) male morphs become externally visible and easily recognized upon pupation. To distinguish high and low nutrition females, only pupae of the lower and upper quartile of the normal pupal weight distribution were selected for ATAC-seq sampling, i.e. pupae > 140 mg for the high-nutrition set and < 110 mg for the low nutrition set. Male pupae of corresponding weight classes invariably metamorphose into alternate horned and hornless morphs, respectively [70]. In total, we collected 16 biological replicates for ATAC-sequencing: 7 females (including 4 high and 3 low nutrition), 4 high nutrition males, and 5 low nutrition males. One high nutrition male and one low nutrition female ATAC-seq libraries required greater than 20 PCR cycles to properly amplify (see details in next paragraph) and as a result, were not sequenced due to poor library construction and complexity.

Upon the onset of pupation, live dorsal head epithelial tissue was dissected from O. taurus individuals in molecular-grade 1X PBS (phosphate-buffered saline) in sterilized 3-well glass dissection plates as described in ref. [12]. After dissection, an estimated 50,000 cells (determined by counting DAPI stained cells in a hemocytometer) were immediately transferred to a 1.5 mL conical tube and processed for ATAC-sequencing [71] using the Omni-Seq protocol [72]. Sequencing libraries were generated by amplifying open chromatin fragments with PCR, having determined the optimal number of amplification cycles with qPCR, as described in ref [73]. Sample were sequenced on an Illumina NextSeq 550 instrument to a minimum depth of 21.8 million 50 bp paired-end reads per sample, including a median 34.8 million reads after trimming and filtering per sample (17.4 million proper pairs) with Trimmomatic v. 0.39 (parameters: leading:10 trailing:10 slidingwindow:4:15 minlen:25) [74].

Filtered, paired-end ATAC-seq reads were aligned to the O. taurus genome assembly described above with Bowtie v.2.2.5 [75]. Alignments were filtered for minimum mapping quality (MAPQ) score of 20 and PCR duplicates were removed with Picard Tools v.2.18.7 (broadinstitute.github.io/picard). After all filtering steps, a median of 14.2 million high quality alignments remained across all samples. ATAC-seq peaks (hereafter open chromatin regions: OCRs) were called with macs2 v.2.2.6 [76] using the parameters:—nomodel—keep-dup = auto—shift 100—extsize 200 -g 2.9e8 -f BAMPE -q 0.05. All OCR sets were merged into a consensus set of 83,925 OCRs with the bedtools v.2.30.0 merge command (d = 0) [77], and peak accessibility counts were calculated with the bedtools multicov command.

ATAC-seq Data analysis

Most statistical analyses of the ATAC-seq count data were carried out in R v.4.2.1. First, OCRs with extremely low accessibility counts were removed (3 counts-per-million required in at least 5 biological replicates), leaving a final set of 68,038 OCRs for downstream analysis (S4 Data). Differential accessibility analyses between sample groups (i.e. high [major] and low [minor] nutrition males, females) was calculated in DESeq2 v.1.36.0 [78] wherein significantly differentially accessible OCRs were called as having false-discovery rate support < 10%. OCRs were putatively assigned to a gene if it was located within 25 Kb (up- or downstream) of the gene’s translational start site. While the transcription start site would normally be a more appropriate genomic reference for assigning OCRs to nearby coding sequence, annotation of these features is currently not sufficient for every gene in the O. taurus genome, so the translation start site was instead chosen as a shared reference point among all gene models.

To identify orthologous OCRs across beetle species, a whole genome alignment was carried out on the soft repeat-masked genome assemblies of the ten beetle species analyzed in this study using Cactus v.2.3 [79]. The HAL file generated from this whole-genome alignment along with OCR coordinates were input into HALPER [80] with the parameters: -max_frac 4 -min_len 30 -protect_dist 1 to identify orthologous OCRs of O. taurus dorsal head epithelia in the genomes of each of the other nine beetle species analyzed. Motif enrichment analyses of OCR sets of interest were carried out in homer v.4.11 [81] with the parameters: -50,50 -mset insects -fdr 10 on a background set of the remaining (non-selected) OCRs identified in this study. Of note, validated transcription factor binding motifs are not available for O. taurus and enrichment analyses were searched against a database of Drosophila melanogaster. As a result, the closest match of some enriched motifs is to transcription factors not present in the O. taurus genome (e.g. bicoid). Gene and repeat annotations, comparative genomic results, and ATAC-seq result files have been deposited on Dryad [82].

Supporting information

S1 Data. Annotation of Orthogroup Membership Gene IDs.

https://doi.org/10.1371/journal.pgen.1011165.s001

(TXT)

S2 Data. Results of CAFE gene family evolution analyses.

https://doi.org/10.1371/journal.pgen.1011165.s002

(XLSX)

S3 Data. Results of HYPHY-BUSTED evidence of episodic diversifying selection analyses.

https://doi.org/10.1371/journal.pgen.1011165.s003

(TXT)

S4 Data. Results of ATAC-seq statistical analyses.

https://doi.org/10.1371/journal.pgen.1011165.s004

(XLSX)

S5 Data. List of top 10% doublsex binding motif hits to Onthophagus taurus genome.

https://doi.org/10.1371/journal.pgen.1011165.s005

(TXT)

S1 Fig. Comparison of lineage-specific and shared OCRs repetitive element content.

A) Percent of total OCRs containing at least one repetitive element. B) Number and class of repetitive element for the 100 most abundant repeats present in each OCR type.

https://doi.org/10.1371/journal.pgen.1011165.s006

(DOCX)

S1 Text. List of commands describing major computational steps involved in this study.

https://doi.org/10.1371/journal.pgen.1011165.s007

(TXT)

Acknowledgments

We thank Erica M. Nadolski for beetle head and horn illustrations.

References

  1. 1. West-Eberhard MJ. Developmental plasticity and evolution. 1st ed. Oxford: Oxford University Press; 2003.
  2. 2. Pfennig DW, Wund MA, Snell-Rood EC, Cruickshank T, Schlichting CD, Moczek AP. Phenotypic plasticity’s impacts on diversification and speciation. Trends in Ecology & Evolution. 2010; 25, 459–467. pmid:20557976
  3. 3. Pfennig DW. Phenotypic Plasticity & Evolution: Causes, Consequences, Controversies. 1st ed. Milton Park: Taylor & Francis; 2021.
  4. 4. Duncan EJ, Gluckman PD, Dearden PK. Epigenetics, plasticity, and evolution: How do we link epigenetic change to phenotype? J. Exp. Zool. 2014; 322B, 208–220. pmid:24719220
  5. 5. Schlichting CD, Wund MA. Phenotypic plasticity and epigenetic marking: an assessment of evidence for genetic accommodation. Evolution 2014; 68, 656–672. pmid:24410266
  6. 6. Dardiry M, Eberhard G, Witte H, Rödelsperger C, Lightfoot JW, Sommer RJ. Divergent combinations of cis-regulatory elements control the evolution of phenotypic plasticity. PLoS Biol. 2023; 21, e3002270.
  7. 7. Tarasov S, Génier F. Innovative bayesian and parsimony phylogeny of dung beetles (Coleoptera, Scarabaeidae, Scarabaeinae) enhanced by ontology-based partitioning of morphological characters. PLoS One. 2015; 10, e0116671. pmid:25781019
  8. 8. Moczek AP. Phenotypic plasticity and the origins of diversity: a case study on horned beetles. In: Whitman D, Ananthakrishnan TN, editors. Phenotypic plasticity in insects: mechanisms and consequences. Milton Park: Taylor & Francis; 2009. pp. 81–134.
  9. 9. Simmons LW, Emlen DJ, Tomkins JL. Sperm competition games between sneaks and guards: a comparative analysis using dimorphic male beetles. Evolution. 2007; 61, 2684–2692. pmid:17941836
  10. 10. Hunt J, Simmons LW. Status-dependent selection in the dimorphic beetle Onthophagus taurus. Proc Biol Sci. 2001; 268, 2409–2414.
  11. 11. Kijimoto T, Moczek AP, Andrews J. Diversification of doublesex function underlies morph-, sex-, and species-specific development of beetle horns. Proceedings of the National Academy of Sciences. 2012; 109, 20526–20531.
  12. 12. Casasa S, Zattara EE, Moczek AP. Nutrition-responsive gene expression and the developmental evolution of insect polyphenism. Nat Ecol Evol. 2020; 4, 970–978. pmid:32424280
  13. 13. Hu Y, Linz DM, Moczek AP. Beetle horns evolved from wing serial homologs. Science. 2019; 366, 1004–1007. pmid:31754001
  14. 14. Hanski I, Cambefort Y. 1st ed. Dung Beetle Ecology. Princeton: Princeton University Press; 2014.
  15. 15. Bornemissza GF. Could dung eating insects improve our pastures? Journal of the Australian Institute of Agricultural Science. 1960; 26, 54–56.
  16. 16. Mittal IC. Natural manuring and soil conditioning by dung beetles. Tropical Ecology. 1993; 34, 150–159.
  17. 17. Moczek AP, Emlen DJ. Male horn dimorphism in the scarab beetle, Onthophagus taurus: do alternative reproductive tactics favour alternative phenotypes? Animal Behaviour. 2000; 59, 459–466.
  18. 18. Emlen DJ, Hunt J, Simmons LW. Evolution of sexual dimorphism and male dimorphism in the expression of beetle horns: phylogenetic evidence for modularity, evolutionary lability, and constraint. The American Naturalist. 2005; 166, S42–S68. pmid:16224711
  19. 19. Davidson PL, Nadolski EM, Moczek AP. Gene regulatory networks underlying the development and evolution of plasticity in horned beetles. Current Opinion in Insect Science. 2023; 60, 101114. pmid:37709168
  20. 20. Linz DM, Moczek AP. Integrating evolutionarily novel horns within the deeply conserved insect head. BMC Biol. 2020; 18, 41. pmid:32312271
  21. 21. Muller GB, Wagner GP. Novelty in evolution: restructuring the concept. Annual Review of Ecology and Systematics. 1991; 22, 229–256.
  22. 22. Emlen DJ, Corley Lavine L, Ewen-Campen B. On the origin and evolutionary diversification of beetle horns. Proceedings of the National Academy of Sciences. 2007; 104, 8661–8668. pmid:17494751
  23. 23. Tarasov SI, Solodovnikov AY. Phylogenetic analyses reveal reliable morphological markers to classify mega-diversity in Onthophagini dung beetles (Coleoptera: Scarabaeidae: Scarabaeinae). Cladistics. 2011; 27, 490–528. pmid:34875798
  24. 24. Emms DM, Kelly S. OrthoFinder: phylogenetic orthology inference for comparative genomics. Genome Biol. 2019; 20, 238. pmid:31727128
  25. 25. Mendes FK, Vanderpool D, Fulton B, Hahn MW. CAFE 5 models variation in evolutionary rates among gene families. Bioinformatics. 2020; 36, 5516–5518.
  26. 26. Murrell B, Weaver S, Smith MD, Wertheim JO, Murrell S, Aylward A, et al. Gene-wide identification of episodic selection. Molecular Biology and Evolution. 2015; 32, 1365–1371. pmid:25701167
  27. 27. Emlen DJ, Nijhout HF. Hormonal control of male horn length dimorphism in the dung beetle Onthophagus taurus (Coleoptera: Scarabaeidae). Journal of Insect Physiology. 1999; 45, 45–53.
  28. 28. Riddiford LM. Juvenile hormone action: A 2007 perspective. Journal of Insect Physiology. 2008; 54, 895–901. pmid:18355835
  29. 29. Sánchez-Gracia A, Vieira FG, Rozas J. Molecular evolution of the major chemosensory gene families in insects. Heredity. 2009; 103, 208–216. pmid:19436326
  30. 30. Robertson HM, Warr CG, Carlson JR. Molecular evolution of the insect chemoreceptor gene superfamily in Drosophila melanogaster. Proceedings of the National Academy of Sciences. 2003; 100, 14537–14542.
  31. 31. Loewith R, Hall MN. Target of rapamycin (TOR) in nutrient signaling and growth control. Genetics. 2011; 189, 1177–1201. pmid:22174183
  32. 32. Kijimoto T, Moczek AP. Hedgehog signaling enables nutrition-responsive inhibition of an alternative morph in a polyphenic beetle. Proceedings of the National Academy of Sciences. 2016; 113, 5982–5987.
  33. 33. Snell-Rood EC, Moczek AP. Insulin signaling as a mechanism underlying developmental plasticity: the role of FOXO in a nutritional polyphenism. PLoS One. 2012; 7, e34857. pmid:22514679
  34. 34. Casasa S, Moczek AP. Insulin signalling’s role in mediating tissue-specific nutritional plasticity and robustness in the horn-polyphenic beetle Onthophagus taurus. Proceedings of the Royal Society B: Biological Sciences. 2018; 285, 20181631.
  35. 35. Kijimoto T, Pespeni M, Beckers O, Moczek AP. Beetle horns and horned beetles: emerging models in developmental evolution and ecology. WIREs Developmental Biology. 2013; 2, 405–418. pmid:23799584
  36. 36. Casasa S, Moczek AP. Evolution of, and via, developmental plasticity: insights through the study of scaling relationships. Integrative and Comparative Biology. 2019; 59, 1346–1355. pmid:31147701
  37. 37. Ledón-Rettig CC, Zattara EE, Moczek AP. Asymmetric interactions between doublesex and tissue- and sex-specific target genes mediate sexual dimorphism in beetles. Nat Commun. 2017; 8, 14593. pmid:28239147
  38. 38. Shibata Y, Sheffield NC, Fedrigo O, Babbitt CC, Wortham M, Tewari AK, et al. Extensive evolutionary changes in regulatory element activity during human origins are associated with altered gene expression and positive selection. PLoS Genetics. 2012; 8, e1002789. pmid:22761590
  39. 39. Davidson PL, Byrne M, Wray GA. Evolutionary changes in the chromatin landscape contribute to reorganization of a developmental gene network during rapid life history evolution in sea urchins. Molecular Biology and Evolution. 2022; 39, msac172. pmid:35946348
  40. 40. Emlen DJ, Szafran Q, Corley LS, Dworkin I. Insulin signaling and limb-patterning: candidate pathways for the origin and evolutionary diversification of beetle horns. Heredity. 2006; 97, 179–191. pmid:16850039
  41. 41. Lynch M, Bobay LM, Catania F, Gout JF, Rho M. The repatterning of eukaryotic genomes by random genetic drift. Annual Review of Genomics and Human Genetics. 2011; 12, 347–366. pmid:21756106
  42. 42. Villar D, Berthelot C, Aldridge S, Rayner TF, Lukk M, Pignatelli M, et al. Enhancer evolution across 20 mammalian species. Cell. 2015; 160, 554–566. pmid:25635462
  43. 43. Cheng H, Concepcion GT, Feng X, Zhang H, Li H. Haplotype-resolved de novo assembly using phased assembly graphs with hifiasm. Nat Methods. 2021; 18, 170–175. pmid:33526886
  44. 44. Li H, Durbin R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics. 2009; 25, 1754–1760. pmid:19451168
  45. 45. Putnam NH, O’Connell BL, Stites JC, Rice BJ, Blanchette M, Calef R, et al. Chromosome-scale shotgun assembly using an in vitro method for long-range linkage. Genome Res. 2016; 26, 342–350. pmid:26848124
  46. 46. Flynn JM, Hubley R, Goubert C, Rosen J, Clark AG, Feschotte C, et al. RepeatModeler2 for automated genomic discovery of transposable element families. Proceedings of the National Academy of Sciences. 2022; 117, 9451–9457.
  47. 47. Brůna T, Hoff KJ, Lomsadze A, Stanke M, Borodovsky M. BRAKER2: automatic eukaryotic genome annotation with GeneMark-EP+ and AUGUSTUS supported by a protein database. NAR Genomics and Bioinformatics. 2021; 3, lqaa108. pmid:33575650
  48. 48. Suzek BE, Wang Y, Huang H, McGarvey PB, Wu CH, and the UniProt Consortium. UniRef clusters: a comprehensive and scalable alternative for improving sequence similarity searches. Bioinformatics. 2015; 31, 926–932. pmid:25398609
  49. 49. Stanke M, Keller O, Gunduz I, Hayes A, Waack S, Morgenstern B. AUGUSTUS: ab initio prediction of alternative transcripts. Nucleic Acids Research. 2006; 34, W435–W439. pmid:16845043
  50. 50. Brůna T, Lomsadze A, Borodovsky M. GeneMark-EP+: eukaryotic gene prediction with self-training in the space of genes and proteins. NAR Genomics and Bioinformatics. 2020; 2, lqaa026. pmid:32440658
  51. 51. Lomsadze A, Burns PD, Borodovsky M. Integration of mapped RNA-Seq reads into automatic training of eukaryotic gene finding algorithm. Nucleic Acids Research. 2014; 42, e119. pmid:24990371
  52. 52. Buchfink B, Xie C, Huson DH. Fast and sensitive protein alignment using DIAMOND. Nat Methods. 2015; 12, 59–60. pmid:25402007
  53. 53. Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, et al. The sequence alignment/map format and SAMtools. Bioinformatics. 2009; 25, 2078–2079. pmid:19505943
  54. 54. Wang Y, Tang H, DeBarry JD, Tan X, Li J, Wang X, et al. MCScanX: a toolkit for detection and evolutionary analysis of gene synteny and collinearity. Nucleic Acids Research. 2012; 40, e49. pmid:22217600
  55. 55. Bandi V, Gutwin C. Interactive exploration of genomic conservation. In Proceedings of the 46th Graphics Interface Conference on Proceedings of Graphics Interface 2020 (GI’20). Waterloo, Canada.
  56. 56. Fallon TR, Lower S.E., Chang CH, Bessho-Uehara M, Martin GJ, Bewick AJ, et al. Firefly genomes illuminate parallel origins of bioluminescence in beetles. eLife. 2018; 7, e36495. pmid:30324905
  57. 57. Herndon N, Shelton J, Gerischer L, Ioannidis P, Ninova M, Dönitz J, et al. Enhanced genome assembly and a new official gene set for Tribolium castaneum. BMC Genomics. 2020; 21, 47.
  58. 58. Keeling CI, Campbell EO, Batista PD, Shegelski VA, Trevoy SAL, Huber DPW, et al. Chromosome-level genome assembly reveals genomic architecture of northern range expansion in the mountain pine beetle, Dendroctonus ponderosae Hopkins (Coleoptera: Curculionidae). Molecular Ecology Resources. 2022; 22, 1149–1167.
  59. 59. McKenna DD, Scully ED, Pauchet Y, Hoover K, Kirsch R, Geib SM, et al. Genome of the Asian longhorned beetle (Anoplophora glabripennis), a globally significant invasive species, reveals key functional and evolutionary innovations at the beetle–plant interface. Genome Biology. 2016; 17, 227.
  60. 60. Schoville SD, Chen YH, Andersson MN, Benoit JB, Bhandari A, Bowsher JH, et al. A model species for agricultural pest genomics: the genome of the Colorado potato beetle, Leptinotarsa decemlineata (Coleoptera: Chrysomelidae). Sci Rep. 2018; 8, 1931.
  61. 61. Morita S, Shibata TF, Nishiyama T, Kobayashi Y, Yamaguchi K, Toga K, et al. The draft genome sequence of Japanese rhinoceros beetle Trypoxylus dichotomus. bioRxiv: 2022.01.10.475740 [Preprint]. 2022 [cited 2023 June 17]. Available from: https://www.biorxiv.org/content/10.1101/2022.01.10.475740v1.
  62. 62. Minh BQ, Schmidt HA, Chernomor O, Schrempf D, Woodhams MD, von Haeseler A, et al. IQ-TREE 2: new models and efficient methods for phylogenetic inference in the genomic era. Molecular Biology and Evolution. 2020; 37, 1530–1534. pmid:32011700
  63. 63. Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods 2017. 14, 587–589. pmid:28481363
  64. 64. Sanderson MJ. r8s: inferring absolute rates of molecular evolution and divergence times in the absence of a molecular clock. Bioinformatics. 2003; 19, 301–302. pmid:12538260
  65. 65. McKenna DD, Shin S, Ahrens D, Balke M, Beza-Beza C, Clarke DJ, et al. The evolution and genomic basis of beetle diversity. Proceedings of the National Academy of Sciences. 2019; 116, 24729–24737. pmid:31740605
  66. 66. Han MV, Thomas GWC, Lugo-Martinez J, Hahn MW. Estimating gene gain and loss rates in the presence of error in genome assembly and annotation using CAFE 3. Molecular Biology and Evolution. 2013; 30, 1987–1997. pmid:23709260
  67. 67. Edgar RC. MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Research. 2004; 32, 1792–1797. pmid:15034147
  68. 68. Shafiei M, Moczek AP, Nijhout HF. Food availability controls the onset of metamorphosis in the dung beetle Onthophagus taurus (Coleoptera: Scarabaeidae). Physiological Entomology. 2001; 26, 173–180.
  69. 69. Rohner PT, Moczek AP. Evolutionary and plastic variation in larval growth and digestion reveal the complex underpinnings of size and age at maturation in dung beetles. Ecology and Evolution. 2021; 11, 15098–15110. pmid:34765163
  70. 70. Moczek AP, Nijhout HF. A method for sexing final instar larvae of the genus Onthophagus Latreille (Coleoptera: Scarabaeidae). The Coleopterists Bulletin. 2002; 56, 279–284.
  71. 71. Buenrostro JD, Giresi PG, Zaba LC, Chang HY, Greenleaf WJ. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nat Methods. 2013; 10, 1213–1218. pmid:24097267
  72. 72. Corces MR, Trevino AE, Hamilton EG, Greenside PG, Sinnott-Armstrong NA, Vesuna S, et al. An improved ATAC-seq protocol reduces background and enables interrogation of frozen tissues. Nat Methods. 2017; 14, 959–962. pmid:28846090
  73. 73. Buenrostro JD, Wu B, Chang HY, Greenleaf WJ. ATAC-seq: a method for assaying chromatin accessibility genome-wide. Current Protocols in Molecular Biology. 2015; 109, 21.29.1–21.29.9. pmid:25559105
  74. 74. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014; 30, 2114–2120. pmid:24695404
  75. 75. Langmead B, Salzberg SL Fast gapped-read alignment with Bowtie 2. Nat Methods. 2012; 9, 357–359. pmid:22388286
  76. 76. Zhang Y, Liu T, Meyer CA, Eeckhoute J, Johnson DS, Bernstein BE, et al. Model-based analysis of ChIP-Seq (MACS). Genome Biology. 2008; 9, R137. pmid:18798982
  77. 77. Quinlan AR, Hall IM. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics. 2010; 26, 841–842. pmid:20110278
  78. 78. Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biology. 2014; 15, 550. pmid:25516281
  79. 79. Armstrong J, Hickey G, Diekhans M, Fiddes IT, Novak AM, Deran A, et al. Progressive Cactus is a multiple-genome aligner for the thousand-genome era. Nature. 2020; 587, 246–251. pmid:33177663
  80. 80. Zhang X, Kaplow IM, Wirthlin M, Park TY, Pfenning AR. HALPER facilitates the identification of regulatory element orthologs across species. Bioinformatics. 2020; 36, 4339–4340. pmid:32407523
  81. 81. Heinz S, Benner C, Spann N, Bertolino E, Lin YC, Laslo P, et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Molecular Cell. 2010; 38, 576–589.
  82. 82. Davidson PL, Moczek AP. Data from: Genome evolution and divergence in cis-regulatory architecture underlies condition-responsive development in horned dung beetles. Dryad Digital Repository. http://dx.doi.org/10.5061/dryad.wdbrv15t8