Skip to main content
eLife logoLink to eLife
. 2025 Oct 28;14:RP103592. doi: 10.7554/eLife.103592

Gene regulatory dynamics during craniofacial development in a carnivorous marsupial

Laura E Cook 1,‡,, Charles Y Feigin 2, John D Hills 1, Davide M Vespasiani 3, Andrew J Pask 1,, Irene Gallego Romero 4,5,6,†,
Editors: Detlef Weigel7, Detlef Weigel8
PMCID: PMC12563553  PMID: 41146606

Abstract

Marsupials and placental mammals exhibit significant differences in reproductive and life history strategies. Marsupials are born highly underdeveloped after an extremely short period of gestation, leading to prioritized development of structures critical for post-birth survival in the pouch. Critically, they must undergo accelerated development of the orofacial region compared to placentals. Previously, we described the accelerated development of the orofacial region in the carnivorous Australian marsupial, the fat-tailed dunnart Sminthopsis crassicaudata, that has one of the shortest gestations of any mammal. By combining genome comparisons of the mouse and dunnart with functional data for the enhancer-associated chromatin modifications, H3K4me3 and H3K27ac, we investigated divergence of craniofacial regulatory landscapes between these species. This is the first description of genome-wide face regulatory elements in a marsupial, with 60,626 putative enhancers and 12,295 putative promoters described. We also generated craniofacial RNA-seq data for the dunnart to investigate expression dynamics of genes near predicted active regulatory elements. While genes involved in regulating facial development were largely conserved in mouse and dunnart, the regulatory landscape varied significantly. Additionally, a subset of dunnart-specific enhancers was associated with genes highly expressed only in dunnart relating to cranial neural crest proliferation, embryonic myogenesis, and epidermis development. Comparative RNA-seq analyses of facial tissue revealed dunnart-specific expression of genes involved in the development of the mechanosensory system. Accelerated development of the dunnart sensory system likely relates to the sensory cues received by the nasal–oral region during the postnatal journey to the pouch. Together, these data suggest that accelerated face development in the dunnart may be driven by dunnart-specific enhancer activity. Our study highlights the power of marsupial–placental comparative genomics for understanding the role of enhancers in driving temporal shifts in development.

Research organism: Other

eLife digest

Marsupials are a distinctive group of mammals best known for their defining trait: a pouch. Unlike monotremes, which lay eggs, and placental mammals, in which young develop fully in the womb, marsupials give birth to highly premature offspring. These young complete most of their development within the mother’s pouch.

Because of this strategy, marsupial newborns possess traits critical for immediate survival outside the womb. These include well-developed forelimbs for climbing into the pouch and a mature oral region for attaching to the teat and suckling.

Evolution has shaped the development of the head to match the diverse environments vertebrates inhabit. Marsupials diverged from placental mammals over 160 million years ago, and their requirement for early limb and oral development makes them a powerful system for investigating the genetic mechanisms underlying development.

The fat-tailed dunnart (Sminthopsis crassicaudata), often called a marsupial mouse, is emerging as an important model organism. With its short gestation period and extremely undeveloped state at birth, the dunnart provides an excellent comparison to the laboratory mouse – a well-established placental model – for studying evolutionary and developmental differences.

Cook et al. investigated the genes and regulatory elements driving early orofacial development in the dunnart, comparing their findings with existing craniofacial expression and epigenomic data in mice.

Their results showed that although the genes involved in craniofacial development are highly conserved between the two species, the regulatory elements controlling those genes differ markedly. In dunnarts, regulatory elements linked to skin, nervous system, and muscle development were highly active, whereas in mice they were inactive or only weakly expressed.

An important feature of marsupial development is the postnatal retention of the periderm (a transient outer cell layer in embryonic skin). This may support gas exchange in prematurely born young. In addition, elevated activity of genes regulating sensory and muscle development suggests early maturation of mechanosensory and olfactory systems, which are essential for the newborn’s journey to the pouch. Notably, pharyngeal and facial muscles develop before the skeletal system, contrasting with the developmental sequence in placental mammals.

In summary, marsupials exhibit a unique reproductive strategy, in which young are born in an extremely underdeveloped state and survival depends on accelerated development of specific traits after birth. By comparing gene regulation in marsupials and placentals, researchers gain insights into how evolution shaped developmental pathways. Expanding genomic resources for the dunnart, such as genome editing tools and transgenic models, will further enhance its role as a powerful comparative system for evolutionary developmental biology.

Introduction

The vertebrate head is a highly complex region of the body that plays a key role in an organism’s ecology by centralizing numerous structures involved in diet, sensory perception, and behavior (Francis-West and Crespo-Enriquez, 2016). Consequently, evolution has modified craniofacial development across lineages, producing a wide array of head morphologies concomitant with the diverse niches that vertebrates occupy. Craniofacial diversity among the major mammalian lineages, in particular, has long been of great interest, due to the striking differences in their developmental ontology.

Placental mammals are characterized by a long gestation with a high maternal investment during pregnancy, and a considerable degree of both orofacial and neurocranial development occurring in the embryo in utero. As a result, placental young experience little functional constraint during early developmental stages. By contrast, marsupials have a short gestation and give birth to highly altricial young that must crawl to the teat, typically located within the maternal pouch, where they complete the remainder of their development ex utero. This unique reproductive method is thought to have imposed strong pressures on the evolution and development of the limbs and head. In particular, marsupials show accelerated development of the nasal cavity, tongue, oral bones, and musculature relative to the development of the posterior end of the body (Nunn and Smith, 1998; Smith, 2006; Smith, 2001a; Cook et al., 2021), and generally when compared to placental embryonic development (Nunn and Smith, 1998; Smith, 2006; Smith, 1997). Additionally, aspects of the peripheral nervous system appear to be accelerated in the face. We and others have observed large medial nasal swellings early in marsupial development that are innervated and proposed to be necessary for the sensory needs of newborn young in their journey to the pouch (Veitch et al., 2000; Jones and Munger, 1985; Waite et al., 1994). Comparative morphometric studies have provided a wealth of evidence that this stark difference in craniofacial development has imposed different regimes of constraint on marsupial and placental mammals (Nunn and Smith, 1998; Smith, 2006; Cook et al., 2021; Smith, 1997; Spiekman and Werneburg, 2017; Koyabu et al., 2014; Goswami et al., 2016) with marsupials in particular showing significantly less interspecies variation in orofacial structures and nasal morphology than placental mammals (Goswami et al., 2016; Sears, 2004; Hüppi et al., 2018). In spite of these observations, the molecular mechanisms that underlie these differences in early craniofacial development between marsupial and placental mammals remain poorly understood.

Cis-acting regulatory regions have been proposed to play a significant role in morphological divergence in the face, with a number of well-described enhancers that fine-tune face shape in mammals (Attanasio et al., 2013). There is also some evidence of a role for regulatory regions in craniofacial heterochrony in marsupials. One recent study found a marsupial-specific region within a Sox9 enhancer that drives early and broad expression in pre-migratory neural crest cell domains contributing to early migration of cranial neural crest cells relative to the mouse (Wakamatsu and Suzuki, 2019; Wakamatsu et al., 2014). However, no study has thus far attempted to compare the overall regulatory landscape between marsupials and placentals at developmentally comparable stages. Such surveys have the potential to provide functional insights into the loci controlling craniofacial heterochrony in mammals and consequently the causative evolutionary changes in the genome that have driven the divergent ontogenies of marsupials and placentals.

In recent years, the fat-tailed dunnart (Sminthopsis crassicaudata, hereafter referred to as the dunnart) has emerged as a tractable marsupial model species (Cook et al., 2021; Suárez et al., 2017). Dunnarts are born after 13.5 days of gestation and craniofacial heterochrony in line with what has been reported in other marsupials is readily observable (Nunn and Smith, 1998; Smith, 2006; Cook et al., 2021; Smith, 1997; Spiekman and Werneburg, 2017; Koyabu et al., 2014; Goswami et al., 2016), making this species an excellent system for comparative studies with placental models. The dunnart provides an ideal model in which to study the regulatory landscape during craniofacial development in marsupials, as they rely entirely on the advanced chondrocranium with bony elements of the skeleton not present until approximately 24 hr after birth (Cook et al., 2021). To investigate a potential role for regulatory elements in this heterochrony, we used chromatin immunoprecipitation (ChIP)-sequencing and RNA-sequencing on craniofacial tissue (fronto-nasal, mandibular, and maxillary prominences) collected from newborn dunnart pouch young. We performed a detailed characterization of chromatin marks during early craniofacial development and then comparative analyses with the placental laboratory mouse. Our work provides valuable insights into genomic regions associated with regulatory elements regulating craniofacial development in marsupials and their potential role in craniofacial heterochrony.

Results

Defining craniofacial putative enhancer and promoter regions in the dunnart

After validating the ability of our antibodies to enrich for dunnart chromatin marks (see Appendix), ChIP-seq libraries were sequenced to average depth of 57 million reads and mapped to a de novo assembly of the dunnart genome generated for this study (see Methods). Peak calling with MACS2 (q < 0.05) identified 80,989 regions reproducibly enriched for H3K4me3 and 121,281 regions reproducibly enriched for H3K27ac in dunnart facial prominence tissue. As this is the first epigenomic profiling for this species, we performed extensive data quality control to ensure the robustness of the data. Similar to previous studies (Villar et al., 2015; Heintzman et al., 2009; Zhu et al., 2013; Cain et al., 2011; Santos-Rosa et al., 2002), we found that H3K4me3 often (62% of all H3K4me3 peaks) co-occupied the genome with H3K27ac (Figure 1A) while 50% of H3K27ac-enriched regions were only associated with this mark (Figure 1A). Active enhancers are generally enriched for H3K27ac (Villar et al., 2015; Creyghton et al., 2010) while sites of transcription initiation (active promoters) can be identified as being marked by both H3K27ac and H3K4me3 (Cain et al., 2011; Santos-Rosa et al., 2002).

Figure 1. Analysis workflow and quality control of H3K4me3 and H3K27ac peaks in the fat-tailed dunnart (Sminthopsis crassicaudata).

Figure 1.

(A) Drawing of a D0 dunnart pouch young with dissected orofacial tissue shown in gray. Short-read alignment and peak calling workflow and numbers of reproducible peaks identified for H3K27ac (orange), H3K4me3 (blue) for craniofacial tissue. (B) Log10 distance to the nearest TSS for putative enhancer (orange) and promoters (blue). (C) Log10 of peak intensity and peak length are represented as boxplots and violin plots for putative enhancers (orange) and promoters (blue). Peak intensities correspond to average fold enrichment values over total input DNA across biological replicates. (D) Dunnart pouch young on the day of birth. Scale bar = 1 mm. (E) Adult female dunnart carrying four young. Statistical significance (Wilcoxon, FDR-adjusted, p < 0.00001) compared to cluster 1 promoter-associated peaks is denoted by ****.

We initially defined promoters as those marked by only H3K4me3 or with >50% of reciprocal peak length for H3K27ac and H3K4me3 peaks, and enhancers as those marked only by H3K27ac, identifying 66,802 promoters and 60,626 enhancers. Enhancers were located on average 77 kb from TSS, while promoters were located on average 106 kb from the nearest TSS, despite there being a greater number of peaks located <1 kb from the TSS (1008 enhancers vs. 9023 promoters; Figure 1B). This was an unexpected finding as a large fraction (0.41) of promoters were located >3 kb from an annotated TSS (see Appendix). H3K4me3 activity at enhancers is well established (Koch and Andrau, 2011; Pekowska et al., 2011); however, compared to H3K4me3 activity at promoters, H3K4me3 levels at enhancers are low (Calo and Wysocka, 2013). This is in line with our observations that H3K4me3 levels at enhancers were nearly seven times lower than those observed at promoter regions (see Appendix for details). Distance from TSS is frequently used to filter putative promoters from other elements, therefore, we grouped peaks characterized as promoters based on their distance to the nearest TSS (see Methods), resulting in 12,295 high-confidence promoters for all of the following analyses (see Appendix).

Candidate dunnart regulatory elements are associated with highly expressed genes involved in muscle, skin, and bone development

Next, we asked what biological processes are associated with active regulatory regions in the dunnart face. To accomplish this, we linked peaks to genes in order to associate functional annotations of coding genes with the candidate regulatory elements that likely regulate their expression. To make use of resources available in model organisms such as GO databases, we converted all dunnart gene IDs to mouse orthologous genes for downstream applications. This reduced the dataset to 35,677 putative enhancers and 8589 promoters near (within 1 Mbp) genes with a one-to-one ortholog in mouse (Supplementary file 1a).

We found that gene annotations for both promoters and enhancers were enriched for 23% of the same GO terms, including cellular processes (protein localization to plasma membrane, protein localization to cell periphery, regulation of cell morphogenesis, positive regulation of cell migration) and development (axon development, camera-type eye development, muscle tissue development, striated muscle development). By contrast, 44% of GO terms were uniquely enriched among genes assigned to promoters and were related to mRNA processing, transcription, mRNA stability, cell cycle, and mRNA degradation (Figure 2A, Supplementary file 1b) and uniquely enriched GO terms for genes assigned to candidate enhancers corresponded to processes indicative of early embryonic development (Figure 2A, Supplementary file 1c).

Figure 2. Predicted functional enrichment for dunnart peaks.

(A) 304 significantly enriched GO terms clustered based on similarity of the terms. The function of the terms in each group is summarized by word clouds of the keywords. Rows marked by P were driven by genes linked to putative promoters, rows marked by E were driven by genes linked to putative enhancers. (B) Enriched TF motifs for transcription factor families (HOMER). PWM logos for preferred binding motifs of TFs are shown. The letter size indicates the probability of a TF binding given the nucleotide composition.

Figure 2.

Figure 2—figure supplement 1. Homer motif enrichment for dunnart promoters and enhancers for the top 20 enriched TF families.

Figure 2—figure supplement 1.

Log10 FDR adjusted p-value scores are shown as a heatmap in enhancers and promoters. Individual TFs are labeled on the right-hand side of the heatmap. Distinct colors represent TF families.

Terms related to facial skeleton development were enriched among genes assigned to putative dunnart enhancers, including bone cell development, muscle cell development, secondary palate development, roof of mouth development, and mesenchyme development, consistent with dunnart craniofacial morphology (Cook et al., 2021). Enhancers active near important palate genes (Won et al., 2023; Zarate et al., 2024) such as SHH, SATB2, MEF2C, SNAI2, and IRF6 in the dunnart at birth may highlight potential regulatory mechanisms driving early palatal closure. In addition, terms related to the development of the circulatory system, including regulation of vasculature development, circulatory system process, and blood circulation were enriched among genes linked to predicted enhancer regions (e.g., ACE, PDGFB, GATA4, GATA6, VEGFA). This is consistent with observations that show the oral region of newborn dunnarts is highly vascularized, with blood vessels visible through their translucent skin at birth (Cook et al., 2021).

To gain further insight into dunnart gene regulation at this developmental stage, we scanned putative enhancers and promoters for 440 known Homer vertebrate motifs and tested for enriched TFs (Heinz et al., 2010). Enhancers were significantly enriched for 170 TFs relative to a background set of random GC- and length-matched sequences (FDR-corrected, p < 0.01), including those with known roles in differentiation of cranial neural crest cells (TWIST, HOXA2), skeletal morphogenesis (DLX5, CREB5, HOXA2), bone development (ATF3, RUNX), cranial nerve development (ATOH1), and/or facial mesenchyme development (LHX2, FOXP1, MAFB; Figure 2B, Figure 2—figure supplement 1). Consistent with the GO enrichment, TFBS in promoter sequences were dominated by transcriptional initiation regulatory sequences, with significant enrichment for 13 TFs (FDR-corrected, p < 0.01) including RFX3, RFX2, NRF, NRF1, GRY, ZBTB33, RONIN, JUND, and GFX (Figure 2B, Figure 2—figure supplement 1).

Next, we assessed predicted target gene expression by performing bulk RNA-seq in dunnart face tissue collected on the day of birth. There were 12,153 genes reproducibly expressed at a level >1 TPM across three biological replicates, with the majority of genes expressed (67%; 8158/12,153) associated with an active enhancer and/or promoter peak. Most enhancers (78% of all enhancers, Supplementary file 1d) or promoters (87% of all promoters, Supplementary file 1e) were linked to a reproducibly expressed gene in the dunnart (Figure 3A). Additionally, the majority of reproducibly expressed genes near a regulatory peak (61%) were associated with both an enhancer and promoter region (Figure 3B), highlighting the correspondence in active regulatory elements and expressed genes at this time point in the dunnart face. Genes with a medium-to-high expression level at this stage (>10 TPM) and associated with at least one putative enhancer and promoter were enriched for biological processes including ‘in utero embryonic development’, ‘skeletal system development’, ‘muscle tissue development’, ‘skin development’, ‘vasculature development’, and ‘sensory organ development’ (Figure 3C, Supplementary file 1f). Enrichment for the term ‘in utero embryonic development’ is indicative of the altricial nature of the dunnart neonate. In a previous study, we showed that in the dunnart, ossification begins post-birth (day of birth young corresponding approximately to embryonic day (E) 12.5 in mouse) and that the dunnart neonate instead likely relies on the well-developed cranial muscles and an extremely large chondrocranium for structural head and feeding support during early pouch life (Cook et al., 2021; Clark and Smith, 1993). Consistent with this, we observed high expression (>20 TPM) of the key head myogenesis genes (Bentzinger et al., 2012; Buckingham, 2017; Lin et al., 2006), MYOD1, MYF6, MEF2C, PAX3, MYL1, and MYOG, essential genes regulating chondrogenesis (Lefebvre et al., 2019; Yip et al., 2019), SOX9, COL2A1, and FGFR1 and genes that act upstream of osteoblast differentiation (Qi et al., 2003; Huang et al., 2007), MSX1, MSX2, CEBPA/G, ALPL, DLX3, DLX5, FGFR1, and FGFR2 (Supplementary file 1d, e).

Figure 3. Genes linked to craniofacial enhancers and promoters in the dunnart are reproducibly expressed and involved in embryonic vasculature, muscle, skin, and sensory system development.

Figure 3.

(A) The majority of nearest genes assigned to candidate enhancers and promoters were reproducibly expressed in dunnart face tissue, and (B) reproducibly expressed genes in dunnart were associated with both a promoter and enhancer region. (C) Biological processes enriched for genes medium to highly expressed (>10 TPM) and linked to both a promoter and enhancer region (FDR-corrected, p < 0.01).

Comparative analyses of regulatory elements in mouse and dunnart reveal conserved and dunnart-specific enhancers during craniofacial development

Previously, we defined the postnatal craniofacial development of the dunnart and characterized the developmental differences between dunnart and mouse (Cook et al., 2021). Marsupials, including dunnarts, have accelerated development of the cranial bones, musculature, and peripheral nerves when compared to placental mammals such as the mouse (Nunn and Smith, 1998; Smith, 2006; Smith, 2001a; Cook et al., 2021; Smith, 1997; Veitch et al., 2000; Jones and Munger, 1985; Waite et al., 1994). Having now characterized the chromatin landscape of the dunnart’s craniofacial region, we next compared it to that of a placental mammal. From our morphological study of dunnart craniofacial bone development, we estimated that craniofacial development in the newborn dunnart is approximately equivalent to E12.5 in mouse (Cook et al., 2021). However, we compared the dunnart to all available stages in the mouse to also provide insights into the regulation of additional craniofacial features (e.g., muscle, cranial nerves, sensory system, and skin). To do this, we took advantage of publicly available craniofacial ChIP-seq data for H3K4me3 and H3K27ac generated by the mouse ENCODE consortium (He et al., 2020) spanning multiple developmental time points (E10.5–E15.5).

After applying consistent peak calling and filtering parameters to the dunnart ChIP-seq data, we found ~17K enhancers and ~10K promoters per stage in the mouse (Figure 4A, see Appendix for details). The features of mouse predicted enhancer and promoters (percentage CpG, GC content, distance from TSS, peak length, and peak intensity) were consistent with the observations in the dunnart (Figure 4A, Figure 4—figure supplement 1). After building dunnart-mm10 liftover chains (see Methods and Appendix for details), we compared mouse and dunnart regulatory elements. The alignability (conserved sequence) for dunnart enhancers to the mouse genome was ~13% for 100 bp regions (Supplementary file 1v). Between 0.74% and 6.77% of enhancer regions out of all alignable enhancers were present in both mouse and dunnart (Supplementary file 1g). In contrast, between 45% and 57% of alignable promoter regions were present in mouse and dunnart (Supplementary file 1g). Although this is a small fraction of the total peaks (~8% for promoters and ~0.5% of enhancers; Supplementary file 1g), it suggests that, consistent with the literature (Villar et al., 2015), promoter regions are more stable over large evolutionary distances and that shifts in developmental timing of craniofacial marsupials may be more likely to be driven by recently evolved enhancer regions in marsupials.

Figure 4. Analysis workflow and features of H3K4me3 and H3K27ac enhancers and promoters in dunnart and mouse.

(A) Alignment filtering and peak calling workflow and (B) number of reproducible predicted regulatory elements identified in the dunnart and mouse embryonic stages. Log10 of distance to the nearest TSS for putative (C) enhancers and (D) promoters.

Figure 4.

Figure 4—figure supplement 1. Features of predicted promoters and enhancers in the dunnart and mouse.

Figure 4—figure supplement 1.

(A) Log10 peak intensity (measure of enrichment) for enhancers (orange) and promoters (blue) prior to filtering. (B, C) Log10 distance to the nearest TSS for enhancers (orange) and promoters (blue). After clustering and filtering for high-confidence promoters, we observed consistent patterns for features including (D) CpG and GC content and (E) Log10 peak intensity and Log10 peak length.

To further contextualize the regulatory patterns in the dunnart compared to mouse, we retrieved mouse gene expression data from embryonic facial prominence tissue collected from E10.5 to E15.5 in the mouse ENCODE database (He et al., 2020), and incorporated this into our comparative analyses. We considered only one-to-one orthologous genes in mouse and dunnart. Predicted dunnart enhancers with sequence conservation in mouse and matching peak activity at any of the five mouse embryonic stages (461 regions) were located near genes reproducibly expressed (>1 TPM, 89%) in both species (Supplementary file 1h). Genes in this group were enriched for core developmental and signaling processes: BMP signaling, cartilage development, ossification, skeletal development, and chondrocyte differentiation (Supplementary file 1i). Taking the predicted dunnart enhancers alignable to the mouse but without a matching peak, we looked at nearby genes and compared expression between mouse and dunnart. For the 4311 dunnart-specific enhancers, we found that 2310 (54%) were linked to genes expressed >1 TPM in all stages and species, suggesting that these genes in mouse could be regulated by a different set of regulatory regions or could be accounted for by the reduced enrichment in the mouse ENCODE ChIP-seq face datasets for H3K27ac (Figure 4—figure supplement 1). We found a smaller subset (179 regions, 114 unique genes) where the nearest genes were highly expressed (>10 TPM) only in the dunnart with low to no activity in the mouse at any of the embryonic stages (E10.5–E15.5; Supplementary file 1j). This included genes involved in cranial neural crest proliferation and migration (INKA2, TFAP2E, OVOL2, GPR161), keratinocyte proliferation (PLAU, HOXA1), embryonic myogenesis (PDF4), development of the ectodermal placodes and sensory systems (CNGA2, ELF5, EDN1, HOXA1, ATOH1, NPHP4, CFD, WNT2B) and vomeronasal sensory neuron development (TFAP2E; Supplementary file 1j).

Dunnart-specific expression of genes involved in the development of the mechanosensory system

Given the large evolutionary distance between the mouse and dunnart and low recovery of alignable regions, we performed a comparison between species at the gene level, by comparing genes assigned to putative enhancers and promoters between the dunnart and mouse. The number of genes that intersect can provide an idea of the similarities in genes and pathways regulated across a larger subset of the total regulatory dataset. The largest intersection size in genes with putative promoters was between the six mouse embryonic stages (1910 genes; 21.2%) and between the dunnart all six embryonic mouse stages (1908 genes, 21.2%; Figure 5A). Overlap between enhancers was more restricted, with 4483 predicted target genes (56%) being unique to the dunnart at D0 (Figure 5A). The top enriched terms for biological processes were largely shared across dunnart and mouse, with the exception of one GO term, ‘sensory system development’ (Figure 5—figure supplement 1). We further investigated this by incorporating gene expression data for mouse and dunnart for genes near putative enhancers. Genes highly expressed in dunnart but lowly or not expressed in mouse (537 total; see Supplementary file 1k) were related to three main developmental processes, ‘epidermis/skin development and keratinization’, ‘sensory system development’, and ‘muscle development and contraction’ (Figure 5C, D; Supplementary file 1l). The majority (70/114) of genes associated with sequence conserved dunnart-specific enhancers (see Supplementary file 1j) overlapped with the list of genes reported here.

Figure 5. Genes near enhancers highly expressed only in dunnart are involved in the development of the skin, muscle, and mechanosensory systems.

Gene set intersections across mouse (E10.5–E15.5) and dunnart (P0) for (A) genes near promoters and (B) genes near enhancers. (C) Gene ontology term enrichment for top 500 highly expressed dunnart genes. (D) Top 50 highly expressed genes (TPM) in dunnart compared with mouse embryonic stages. Scale bar in E. (E) Expression levels (TPM) for keratin genes across dunnart and mouse.

Figure 5.

Figure 5—figure supplement 1. Top GO enriched terms in the dunnart and mouse embryonic stages for (A) genes near enhancers and, (B) genes near high-confidence promoters.

Figure 5—figure supplement 1.

Figure 5—figure supplement 2. Heatmaps showing gene expression values (TPM) for common developmental genes across dunnart and mouse.

Figure 5—figure supplement 2.

Figure 5—figure supplement 3. Volcano plot for both up- and downregulated differentially expressed genes from mouse (any stage) and dunnart.

Figure 5—figure supplement 3.

Stringent log2 fold change and p-value cutoffs were used given the comparisons between datasets from two different species.

Genes critical for development of keratinocytes and the establishment of a skin barrier were highly expressed in dunnart facial tissue with lower expression or no transcripts expressed across the mouse embryonic stages including IGFBP2, SFN, AQP3, HOPX, KRT17, KRT7, KRT8, and KRT78 (Figure 5D, Supplementary file 1l). Keratin genes are also critical for the development of the mammalian mechanosensory system (Xiao et al., 2014; Doucet et al., 2013). Krt17-expressing epidermal keratinocytes are necessary to establish innervation patterns during development, and Krt8 and Atoh1 expression is required for the specification of the Merkel cells (touch sensory cells) (Doucet et al., 2013). KRT17, ATOH1, and KRT8 are all very highly expressed in dunnart compared to the low expression in mouse facial tissue. Other genes in the keratin family such as KRT35, KRT79, KRT4, and KRT80 did not vary greatly between mouse and dunnart (Figure 5E). Additionally, we observed high expression of genes involved in development of the olfactory system including CCK, OTX1, and ISLR (expressed in olfactory epithelium) (Liu and Liu, 2018; Simeone, 1998; Santoro and Jakob, 2018; Haering et al., 2015), and Ybx3 (Haering et al., 2015) (expressed in the nasal epithelium; Figure 5D, Supplementary file 1l). Muscle contraction genes such as TNNI2 (Robinson et al., 2007) and TNNT3 (Wei and Jin, 2016; Sung et al., 2003) and genes involved in skeletal muscle development (Jung and Ko, 2010; Jin et al., 2014; Feng and Jin, 2016; Aboalola and Han, 2017) (CAR3, ATP2A2, IGFBP6, and TRIM72) were upregulated in dunnart craniofacial tissue (Figure 5D, Supplementary file 1l). The majority of conserved toolkit genes involved in embryonic development had consistent expression across mouse and dunnart (Figure 5—figure supplement 2). To explore the relationship between genes expressed in the dunnart face and temporal gene expression dynamics during mouse development, we categorized mouse gene expression into five distinct temporal patterns (Figure 6—figure supplement 1). Each of these groups appeared to reflect biological processes occurring during development (Figure 6—figure supplement 1). Although dunnart facial development more closely resembles approximately E12.5 (Cook et al., 2021) in the mouse, when compared to the temporal gene expression dynamics during mouse craniofacial, dunnart expressed genes were associated with two distinct clusters: a set of genes upregulated specifically at E15.5 in mouse (cluster 2: OR = 1.30, CI = 1.15–1.46; Figure 6A; Figure 6—figure supplement 1) and a set of genes upregulated at E14.5 (cluster 3: OR = 1.99; CI = 1.78–2.24; Figure 6B; Figure 6—figure supplement 1). Cluster 2 genes were enriched for functions related to sensory perception, skin development, and keratinization (Figure 6C), while cluster 3 genes were enriched for muscle development (Figure 6D). These results align with our observations in dunnart enhancer and gene expression data, suggesting that shifts in the developmental timing of the skin, muscle, and sensory perception may play a role in marsupial early life history.

Figure 6. Dunnart-expressed genes are associated with two gene clusters with distinct temporal expression patterns in the mouse.

(A, B) Genes in clusters 2 and 3 plotted with their z-scaled temporal expression (logCPM). Color-coding represents membership value (degree to which data points of a gene belong to the cluster). Gene ontology enrichment for biological processes enriched in (C) cluster 2 and (D) cluster 3 (FDR-corrected p < 0.01).

Figure 6.

Figure 6—figure supplement 1. Temporal gene expression dynamics.

Figure 6—figure supplement 1.

(A) Z-scaled temporal expression (logCPM) plotted across embryonic time points for five clusters. Membership values are used to indicate to what degree a data point belongs to each cluster. (B) Top enriched biological processes for each cluster (FDR-corrected p < 0.01, background gene set is all differentially expressed genes used for clustering). (C) Odds ratio estimates with 95% confidence interval for genes expressed in dunnart in each cluster. (D) Number of genes in each cluster for mouse and dunnart.

Discussion

Marsupials display advanced development of the orofacial region relative to development of the central nervous system when compared to placental mammals (Smith, 2006; Smith, 1997). Specifically, the facial skeleton and muscular tissues begin development early relative to the neurocranium differentiation (Smith, 2006; Smith, 1997; Clark and Smith, 1993; Smith, 1999; Smith, 2001b; Smith, 2003; Sánchez-Villagra and Forasiepi, 2017). Although development of the central nervous system is protracted in marsupials compared to placentals, marsupials have well-developed peripheral motor nerves and sensory nerves (e.g., the trigeminal) at birth (Karlen and Krubitzer, 2007; Smith and Keyte, 2020). Despite increasing descriptions of craniofacial enhancers in model species (Attanasio et al., 2013; Brinkley et al., 2016; Samuels et al., 2020; Prescott et al., 2015; Rajderkar et al., 2024), the genetic drivers of variation in craniofacial development between mammalian species are largely unexplored. We examined the similarities and differences in genomic regions marked by H3K4me3 and H3K27ac between the dunnart and mouse during early craniofacial development and incorporated comparisons of gene expression between species.

Given the large evolutionary distance (~170 million years) between the mouse and dunnart (Bininda-Emonds et al., 2007; Luo et al., 2011) and high turnover of non-coding DNA sequences (Genereux et al., 2020; Andrews et al., 2023), regulatory elements in the dunnart were largely not sequence conserved in the mouse (~8% for promoters and ~0.5% of enhancers). Despite this, regions with conserved activity between the two species were predominantly near genes consistently expressed in both and were enriched for core craniofacial developmental and signaling processes. Additionally, 54% of dunnart enhancers aligned to the mouse genome but lacking active marks in the mouse were associated with genes expressed throughout all mouse developmental stages, suggesting these genes might be regulated by species-specific enhancers in mice. Moreover, despite differences in experimental methods, when taking a gene-level approach, there was significant overlap in genes near regulatory elements in both species, with high concordance in enriched biological process terms. The majority of craniofacial developmental genes had highly consistent gene expression levels across the mouse and dunnart, highlighting the substantial conservation of gene regulatory networks driving facial development. This supports the notion that conserved genes and signaling pathways are crucial for mammalian facial development (Roosenboom et al., 2016).

Although there was generally a high level of concordance between the dunnart and mouse, we discovered dunnart-specific regulatory elements near highly expressed genes in the dunnart that were lowly or not expressed in the mouse. These genes were related to three main developmental processes, ‘epidermis/skin development and keratinization’, ‘sensory system development’, and ‘muscle development and contraction’. In marsupials, pharyngeal muscles are well differentiated before birth preceding development of both the central nervous system and the skeletal system (Smith, 1994). Additionally, in the first 2 days immediately after birth, marsupial pouch young undergo considerable differentiation of the facial muscles (Smith, 1994). This pattern varies significantly from rodents where the events of skeletal and muscular development generally occur simultaneously (Smith, 1994). Consistent with this, we found that dunnart-expressed genes were associated with a cluster of mouse genes highly expressed at E14.5, a key time point for myogenesis in the mouse (Stockdale, 1992).

Our data also revealed a less evident developmental constraint experienced by marsupials, with dunnart-specific enhancer and gene expression data showing an enrichment of genes related to skin development. Newborn dunnarts exhibit high expression of KRT17 in the face and other genes essential for periderm establishment and maintenance (Hammond et al., 2019; Liu et al., 2016; Noiret et al., 2016; Kousa et al., 2017; Carroll, 2024; Vaziri Sani et al., 2010). Unlike placental mammals, marsupials retain the periderm post-birth (Smith and Keyte, 2020; Lillegraven, 1975; Pralomkarn et al., 1990), which may relate to neonatal transcutaneous gas exchange observed in marsupials in the first few days after birth (Simpson et al., 2011; Ferner, 2018). In placental mammals such as mouse and human, the periderm only exists in utero and disappears late in gestation with the eruption of hair and the development of a thick keratinized stratum corneum (Hammond et al., 2019). The persistence of the periderm aligns with the observation that dunnarts rely heavily on gas exchange through the skin and only begin lung respiration around 5 days postpartum, coinciding with the disappearance of the periderm (Simpson et al., 2011; Krause et al., 1978).

The recurrence of genes related to the development of the sensory systems in our comparative analyses spotlights a highly unique aspect of early pouch life in marsupials. Newborn marsupial young require highly developed sensory systems (mechanosensory and olfactory systems) to respond to the cues that guide them to the teat inside the mother’s pouch, with the nasal-oral region making regular contact with the mother’s belly during the crawl to the pouch (Waite et al., 1994; Desmarais et al., 2016; Gemmell and Nelson, 1989). The snout epidermis of newborn marsupials has been shown to be innervated with the presence of mature Merkel cells with connected nerve terminals (Veitch et al., 2000; Jones and Munger, 1985; Desmarais et al., 2016). We observe high expression of key genes involved in the development of Merkel cells and sensory placodes uniquely in the dunnart face. In the mouse, Merkel cell development does not begin until later at approximately E16.5 (Jenkins et al., 2019). Upregulation of Merkel cell genes in dunnart facial tissue could be a result of prioritization of sensory development required for life outside the womb in a highly underdeveloped state.

In summary, our work suggests that craniofacial developmental constraints in the dunnart may be driven by dunnart-specific gene regulation. Future studies will apply single-cell multi-modal technologies to dunnart tissues to investigate the specific regulatory differences between CNS and orofacial development.

Methods

Tissue collection

All animal procedures, including laboratory breeding, were conducted in accordance with the current Australian Code for the Care and Use of Animals for Scientific Purposes (NHMRC, 2013) and were approved by The University of Melbourne Animal Ethics Committee (AEC: 1513686.2) and with the appropriate Wildlife Permit (number 10008652) from the Department of Environment, Land, Water and Planning. Animals were housed in a breeding colony in the School of BioSciences, The University of Melbourne. For details on animal husbandry and collection of dunnart pouch young, refer to Cook et al., 2021. Details of pouch young used in this study are presented in Supplementary file 1m. After removal of pouch young from the teat, young were killed by decapitation and craniofacial tissue dissected using insulin needles (Becton Dickinson). The tongue, neural tube, and eye primordia were removed to limit tissue collected to the facial prominences. Specifically, the mandibular, maxillary, and fronto-nasal prominences were collected and snap frozen in liquid nitrogen. All pouch young were determined to be <24 hr old, but to account for variability in time since birth, we scored pouch young based on head shape. Immediately after birth, the dunnart has a flat neurocranium that by approximately 1 day after birth has begun to round (see Cook et al., 2021 for more details). We combined craniofacial tissue from 50 pouch young into two replicates, ensuring that sex, head shape, and parentage were accounted for (Supplementary file 1m).

Immunofluorescence

We assessed the reactivity of the ChIP antibodies in the dunnart using immunostaining on dunnart head sections. Frontal head sections (7 µm thick) on superfrost slides (Platinum Pro, Grale) were deparaffinized and rehydrated according to standard methods (Bancroft and Gamble, 2008), followed by antigen retrieval using pH 8.8 unmasking solution (Vector) at 99°C for 30 min. We then incubated the sections with either rabbit anti-H3K4me3 primary antibody (1:500 dilution; Abcam ab8580) or rabbit anti-H3K27ac primary antibody (1:500, Abcam ab4729). Sections were then incubated with Alexa Fluor 555 donkey anti-rabbit antibody (1:500 dilution; Abcam in 10% horse serum in PBS with 0.1% Triton X-100; Sigma). All sections were counterstained with 300 nM DAPI to visualize cell nuclei. We observed no staining in the negative controls (no primary antibody). Images were captured using fluorescence microscopy (BX51 Microscope and DP70 Camera; Olympus) and processed in ImageJ v 2.0.0 (Schneider et al., 2012).

ChIP and sequencing

ChIP was performed using the MAGnify Chromatin Immunoprecipitation System (Thermo Fisher, 492024) according to the manufacturer’s instructions. Briefly, frozen dunnart tissue samples (see Supplementary file 1m) were diced quickly with two razor blades in cold dPBS (Gibco) followed by crosslinking in 1% formaldehyde solution (Sigma) for 10 min. We then added 0.125 M glycine and incubated for 5 min at room temperature to neutralize the formaldehyde. Chromatin was fragmented to 300 bp average size by sonication on a Covaris S2 using the following parameters: [duty cycle = 5%, intensity = 2, cycles per burst = 200, cycle time = 60 s, cycles = 10, temperature = 4°C, power mode = frequency sweeping, degassing mode = continuous]. In each ChIP experiment, we used sheared chromatin from each replicate for immunoprecipitation with antibodies against H3K4me3 (Abcam ab8580) and H3K27ac (Abcam ab4729). An input control was included for each replicate. DNA was purified according to kit instructions using a DynaMag-PCR Magnet (Thermo Fisher).

We assessed the success of the immunoprecipitation in dunnart craniofacial tissues by performing qPCR for primers designed to amplify genomic regions expected to be occupied by H3K4me3 and/or H3K27ac based on mouse and human enhancers active in facial regions of E11.5 mice in VISTA enhancer (Visel et al., 2007) and GeneHancer (Fishilevich et al., 2017; Supplementary file 1n). We also designed primers that amplify regions we predicted would be unoccupied by these histone modifications (enhancers active in heart tissue of E11.5 mice in VISTA enhancer browser; Visel et al., 2007; Supplementary file 1n). qPCR using SYBR Green Supermix (Bio-Rad Laboratories) was performed in triplicate on a QuantStudio 5 System (Thermo Fisher) as per the manufacturer’s instructions (primers listed in Supplementary file 1n). The cycling conditions were [one cycle of 95°C for 15 s, followed by 40 cycles of 95°C for 15 s, 57°C for 30 s, and 72°C for 30 s]. A dissociation curve was also generated for each primer pair. No-template controls were included in triplicate on each plate as a negative control. Regions expected to be enriched in the test sample were quantified by expressing the test sample as a fold change relative to a control sample (no antibody control).

Illumina sequencing libraries were prepared from ChIP-enriched DNA by GENEWIZ (Suzhou, China). Libraries were constructed following the manufacturer’s protocol (NEBNext UltraTMII DNA Library Prep Kit for Illumina). For each sample, a minimum of 10 ng of ChIP product was used and libraries were multiplexed and sequenced on an Illumina HiSeq 4000 instrument according to the manufacturer’s instructions (Illumina, San Diego, CA, USA). Sequencing was carried out using a 2 × 150 paired-end configuration to an average depth of 57 million read pairs per sample.

Genome assembly and annotation

In order to generate a genome for the dunnart, tissue was collected and four sequencing libraries were prepared following existing methods (Sambrook et al., 1989). Four libraries were generated to improve molecular complexity and genome representation of input DNA. These libraries were then sequenced using the following technologies: Illumina X Ten 2 × 150 bp, PacBio Sequel I CLR, ONT PromethION, and ONT GridION (Supplementary file 1o).

Quality trimming and residual adaptor removal from dunnart Illumina libraries (Library 1) was performed using trimmomatic v0.38 (Bolger et al., 2014) with options [PE SLIDINGWINDOW 5:20, MINLEN 75, AVGQUAL 30]. Contigs were assembled using 200 Gb of PacBio CLR subreads (Library 2) using Flye v2.7 (Kolmogorov et al., 2019) with options [–iterations 4 –trestle –pacbio-raw –genome-size 3.0g]. Removal of redundant contigs and two rounds of short- and long-read scaffolding were performed using Redundans 0.14a (Pryszcz and Gabaldón, 2016) with options [--nogapclosing --limit 1.0]. Inputs for redundancy scaffolding were short-insert paired-end reads (Library 1) and 6.5 gigabases of Oxford Nanotechnology reads corresponding to two libraries (Library 3 and Library 4). Scaffolds then underwent two rounds of polishing to improve base quality using Pilon v1.23 (Walker et al., 2014) with Illumina Library 1 as input and with options [--vcf --diploid --chunksize 10000000 --fix snps,indexls,gaps --minqual 15]. The resulting genome assembly had a total size of 2.84 Gb and an N50 length of 23 megabases (Mb). We used BUSCO v5.2.2 to assess genome completeness [v3.0.2, -l mammalia_odb10 -m genome]. BUSCO gene recovery was 89.9% for complete orthologs in the mammalia_odb10 lineage dataset which includes 9226 BUSCOs. Together, these metrics indicate that the assembly is of comparable completeness and contiguity to other recently published marsupial genomes (Feigin et al., 2018; Feigin et al., 2022; Renfree et al., 2011; Murchison et al., 2012; Mikkelsen et al., 2007; Peel et al., 2021; Brandies et al., 2020) and therefore represents an excellent resource for downstream functional genomic experiments. The resulting de novo assembly with a resulting scaffold N50 of 23 Mb and a total size of 2.84 Gb makes it comparable to other marsupial genomes (Feigin et al., 2018; Renfree et al., 2011; Murchison et al., 2012; Mikkelsen et al., 2007). The G+C content of the de novo contigs in the dunnart (~36.25%) was similar to other marsupial species Tasmanian devil: 36.4% (Murchison et al., 2012), thylacine: ~36% (Feigin et al., 2018), wallaby: 34–38.8% (Renfree et al., 2011), opossum: 37.8% (Mikkelsen et al., 2007), woylie: 38.6% (Peel et al., 2021), and the brown antechinus: 36.2% (Brandies et al., 2020).

Gene annotations from the high-quality genome assembly of the Tasmanian devil (Sarcophilus harrisii, GCF_902635505.1 – mSarHar1.11), which has a divergence time with the dunnart of approximately 40 million years, were accessed from NCBI and then lifted over to dunnart scaffolds using the program liftoff v1.0 (Shumate and Salzberg, 2021) with option [--d 4].

Whole-genome alignment

To compute pairwise genome alignments between the mouse and dunnart, we used the mouse mm10 assembly as the reference. We first built pairwise alignments using Lastz and axtChain to generate co-linear alignment chains (Kent et al., 2003), using the previously described Lastz parameters for vertebrates, [K = 2400, L = 3000, Y = 3400, H = 200] with the HoxD55 scoring matrix (Sharma and Hiller, 2017). After building chains, patchChain (Sharma and Hiller, 2017) was applied to extract all the unaligned loci and create local alignment jobs for each one. The new local alignments were combined with the original local alignments for an improved set of chains. We then applied chainCleaner (Suarez et al., 2017) with the parameters [-LRfoldThreshold = 2.5 -doPairs -LRfoldThresholdPairs = 10 -maxPairDistance = 10,000 -maxSuspectScore = 100,000 -minBrokenChainScore = 75,000] to improve the specificity of the alignment. After generating an improved set of chains, we applied chainPreNet, chainNet, and ChainSubset to filter, produce the alignment nets and create a single chain file using only the chains that appear in the alignment nets (Kent et al., 2003). Alignment nets are a hierarchical collection of the chains that attempt to capture orthologous alignments (Kent et al., 2003). Chain fragments were joined using chainStitchId and dunnart to mouse chains generated using chainSwap (Kent et al., 2003). For quality control, maf files were generated using netToAxt and axtToMaf (Kent et al., 2003). Block counts, block lengths, and pairwise divergence in the alignments were assessed using MafFilter (Dutheil et al., 2014).

ChIP sequencing data analysis

First, we assessed the raw sequencing read quality using FastQC v0.11.9 (Andrews, 2010). Raw data were processed by adapter trimming and low-quality read removal using Cutadapt v1.9.1 (Martin, 2011) [-q 20a AGATCGGAAGAGCACACGTCTGAACTCCAGTCA -A AGATCGGAAGAGCGTCGTGTAGGGAAAGAGTGT --max-n 0.10 -m 75]. ChIP sequencing statistics for raw reads and trimmed reads are described in Supplementary file 1p. Sequencing reads were aligned to the dunnart genome with Bowtie2 v.2.3.5.1 (Langmead and Salzberg, 2012) [-q -X 2000 --very-sensitive]. Unfiltered aligned reads from ChIP-seq experiments performed using mouse embryonic facial prominence for E10.5, E11.5, E12.5, E13.5, E14.5, and E15.5 were downloaded from https://www.encode.org/ (He et al., 2020) (accession details described in Supplementary file 1q). For both dunnart and mouse-aligned reads, low-quality and unpaired reads were removed using Samtools v.1.9 (Li et al., 2009) [-q 30 -f 2] and duplicate reads removed by the MarkDuplicates tool from picard v.2.23.1 (https://broadinstitute.github.io/picard/). Mapping statistics and library complexity for dunnart and mouse reads are described in Supplementary file 1r and s, respectively. Effective genome size for the dunnart was calculated using the [unique-kmers.py] script from khmer v.2.0 (Crusoe et al., 2015).

Peaks were called on the dunnart-aligned reads using MACS2 v.2.1.1 (Zhang et al., 2008) [-f BAMPE] and the mouse aligned reads using MACS2 v.2.1.1 with default parameters for mm10, using total DNA input as control and retaining all statistically significant enrichment regions (FDR-corrected p < 0.01). Reproducible consensus peaks for biological replicates within a species were determined using the ENCODE3 [overlap_peaks.py] script (He et al., 2020). Enriched regions were considered reproducible when they overlapped in two biological replicates within a species by a minimum of 50% of their length using bedtools intersect v2.29.2 (Quinlan and Hall, 2010). Peak-calling statistics for dunnart and mouse are described in Supplementary file 1t and u, respectively. Similar to Villar et al., 2015, we overlapped H3K4me3 and H3K27ac reproducible peaks to determine promoter-associated peaks (marked by only H3K4me3 or both H3K4me3 and H3K27ac) and enhancer-associated peaks (marked only by H3K27ac). H3K4me3 and H3K27ac reproducible peaks were overlapped to determine genomic regions enriched for H3K4me3, H3K27ac, or both marks using bedtools intersect v2.29.2 (Quinlan and Hall, 2010). Double-marked H3K4me3 and H3K27ac elements were defined as regions reproducibly marked by H3K4me3 and H3K27ac and overlapping by a minimum 50% of their reciprocal length and were merged with bedtools v2.29.2 (Quinlan and Hall, 2010). Mouse and dunnart peaks called on aligned reads are deposited at 10.7554/eLife.103592.1.

Gene ontology enrichment analyses

For the dunnart peaks, gene annotations lifted over from the Tasmanian devil annotation were associated with ChIP-seq peaks using the default settings for the annotatePeak function in ChIPseeker v1.26.2 (Yu et al., 2015). As there is no equivalent gene ontology database for dunnart, we converted the Tasmanian devil RefSeq IDs to Ensembl v103 IDs using biomaRt v2.46.3 (Durinck et al., 2009; Durinck et al., 2005), and then converted these to mouse Ensembl v103. In this way, we were able to use the mouse Ensembl gene ontology annotations for the dunnart gene domains. We were able to assign Devil Ensembl IDs to 74% of genes with peaks, and mouse IDs to 95% of genes with a devil Ensembl ID. For calculating enrichment in GO in the dunnart, the list of Tasmanian devil genes with an orthologous Ensembl gene in the mouse was used as the background list. All gene ontology analyses were performed using clusterProfiler v4.1.4 (Wu et al., 2021), with gene ontology from the org.Mm.eg.db v3.12.0 database (Carlson, 2020), setting an FDR-corrected p-value threshold of 0.01 for statistical significance.

Short sequence motifs enriched in dunnart peaks were identified with Homer v4.11.1 (Heinz et al., 2010) using [findMotifsGenome.pl]. In this case, random GC- and length-match sequences for all promoters and enhancers were used as the background set to test for enrichment compared to random expectation. Enriched motifs were clustered into Homer motif families (Heinz et al., 2010).

Mouse and dunnart peak comparison

Dunnart and mouse peaks called from normalized input reads were filtered to 100-bp regions centered on the peak summit. Dunnart peak coordinates were lifted over using liftOver (Kent et al., 2003) to the mouse (mm10) genome and then back to the dunnart genome. Alignable peaks were kept if after reciprocal liftOver they had the same nearest gene call (Supplementary file 1v). Alignable peaks were then intersected with enhancer and promoters at each stage in the mouse with bedtools intersect v.2.30.0 (Quinlan and Hall, 2010) to assess peaks with conserved activity (Supplementary file 1w).

RNA-sequencing and analyses

The reads for three replicates were generated from facial prominences of P0 dunnart pouch young. Each replicate is a pool of two pouch young. Libraries were prepared using the Illumina stranded mRNA kit and were sequenced to a depth ≥50 M read pairs (i.e., 25 M F + 25 M R) in 2 × 100 bp format. The raw reads were first trimmed using Trimmomatic v0.39 (Bolger et al., 2014) [ILLUMINACLIP:2:30:10 SLIDINGWINDOW:5:15 MINLEN:50 AVGQUAL:20]. Reads were mapped with hisat2 v2.21 (Zhang et al., 2021) [--fr --no-mixed --no-discordant] and alignments were filtered using samtools view (Li et al., 2009) [-f 1 -F 2316]. Finally, the count table was generated using featureCounts from the Subread package v2.0.1 (Liao et al., 2013) [-s 2 -p -t exon --minOverlap 10]. Mouse gene count tables were acquired from ENCODE (see Supplementary file 1q). Mouse and dunnart gene expression data were normalized for library size with edgeR v.4.0.16 (Robinson et al., 2010) and lowly expressed genes filtered (>2 cpm) for at least two out of three replicates.

For mouse gene expression data, the resulting gene list was tested for temporally, differentially expressed genes with the [DBanalysis] function in the TCseq package (Wu and Gu, 2020), which implements edgeR to fit read counts to a negative binomial generalized linear model. Differentially expressed genes with an absolute log2 fold change >2 compared to the starting time point (E10.5) were determined. The optimal division of clusters was determined using the Calinski criterion implemented with the [cascadeKM] function in the vegan v. package (Oksanen, 2020) with the parameters [inf.gr = 1, sup.g = 10 iter = 1000, criterion = ‘calinski’]. Time-clustering for five clusters was performed using the [timeclust] function with the parameters: [algo = ‘cm’, k = 5, standardize = TRUE] which performs unsupervised soft clustering of gene expression patterns into five clusters with similar z-scaled temporal patterns. Orthologous genes reproducibly expressed >1 TPM in the dunnart were compared to the list of genes for each cluster using Fisher’s exact test followed by p-value corrections for multiple testing with the Holm method. Gene ontology enrichment was performed using [enrichGO] as part of the clusterProfiler v4.10.1 (Wu et al., 2021) with the 9933 genes present across the clusters as background and a false discovery rate of 1%.

Acknowledgements

St Vincent’s Institute acknowledges the infrastructure support it receives from the National Health and Medical Research Council Independent Research Institutes Infrastructure Support Program and from the Victorian Government through its Operational Infrastructure Support Program. The authors thank Eva Suric, Tania Long, and Darren Cipolla for their technical contributions and help with animal husbandry.

Appendix 1

Validation of ChIP-seq in dunnart craniofacial tissue

As the H3K4me3 and H3K27ac antibodies used in this work had not been previously used in marsupials, we first tested their reactivity in the dunnart using immunofluorescence and observed strong positive staining for both antibodies (Appendix 1—figure 1). We validated the levels of enrichment in the ChIP libraries using qPCR and primers for sequences in the dunnart that were orthologous to the craniofacial enhancers mm428, mm387, mm423, hs466 and GH07J005561 and an enhancer active in the heart as a negative control, hs222 (Supplementary file 1n). All primer sets were enriched as a percentage of the input control, and as expected, no enrichment was observed in the rabbit and mouse mock IgG negative control samples (Appendix 1—figure 1). Although the dunnart ortholog of the human heart enhancer (hs222) was also enriched in both ChIP samples, this enhancer has only been tested at E11.5 in mouse embryos (Visel et al., 2007) and may have alternative roles at other stages in development or in different species.

Having validated the ability of our antibodies to enrich for dunnart chromatin marks, ChIP-seq libraries were sequenced to average depth of 57 million reads. Reads were mapped to a de novo assembly of the dunnart genome generated for this study (Supplementary file 1o), which we annotated using the publicly available high-quality genome assembly of the Tasmanian devil (mSarHar1.11; see Methods for additional details). After mapping and filtering, we retained an average of 43 million mapped reads per library (Supplementary file 1p). All samples had an NRF of between 0.76 and 0.85 (Supplementary file 1r; He et al., 2020). Read coverage was highly correlated for replicates within each mark (Pearson’s r > 0.99) and also between marks (Pearson’s r > 0.88; Appendix 1—figure 2). Read coverage for replicate input controls was also correlated, but as expected, due to enrichment for specific genomic regions in IP samples, there was no correlation in read coverage observed between the input controls and IP samples (Pearson’s r range = 0.03–0.08, Appendix 1—figure 2). Similarly, alignment BAM files for H3K4me3 and H3K27ac IP replicates show strong and specific enrichment indicated by a prominent and steep rise of the cumulative sum toward the highest rank, whereas input control alignment BAM files are roughly diagonal, highlighting even coverage across the genomic windows (Appendix 1—figure 2). Fraction of reads in peaks was approximately 60% for consensus H3K27ac and H3K4me3 peaks (Supplementary file 1t).

Investigating promoters located >3 kb from an annotated TSS

To investigate the large number of putative promoters located >3 kb from an annotated TSS, we first looked at the relationship between the number of peaks per gene and the distance to the next closest gene. After annotating peaks with the nearest gene call using ChIPseeker (Yu et al., 2015), 72% of genes had at least 1 predicted promoter or enhancer, with 81 genes having more than 50 promoters (Appendix 1—figure 3). The number of promoters per gene was weakly positively correlated with distance to the next closest gene (Pearson’s r = 0.36, p < 2.2 × 10–16), suggesting that this observation might be partially due to the presence of unannotated transcripts between genes (Appendix 1—figure 3). Enhancers can also be associated with H3K4me3 (Wu and Gu, 2020). However, enrichment levels (peak intensity) tend to be lower than that of H3K27ac (Kim et al., 2010). Hence, peaks characterized as promoters that were located more than 3 kb from a known TSS may instead represent active enhancer regions (Kim et al., 2010). To assess this, we compared mean peak intensity in H3K4me3 peaks located greater than 3 kb from the nearest TSS to H3K4me3 peaks located within 3 kb of the TSS. We found that H3K4me3 peaks located closer to the TSS had a stronger peak signal (mean = 46.10) than distal H3K4me3 peaks (mean = 6.95; Wilcoxon FDR-adjusted p < 2.2 × 10–16). This suggests that although some distal promoter peaks may be due to missingness in the annotation, the majority likely represent peaks associated with enhancer regions.

Defining high-confidence promoter regions

This analysis identified three clusters: upstream distal peaks (cluster 3, 31,285 peaks), downstream distal peaks (cluster 2, 22,863 peaks), and peaks centered on the TSS (cluster 1, 12,295 peaks; Appendix 1—figure 4). Cluster 1 peaks had a higher intensity (Appendix 1—figure 4) and length (Appendix 1—figure 4) when compared to clusters 2 and 3 (Wilcoxon FDR-adjusted p < 2.2 × 10–16) and compared to all predicted enhancers and promoters (Wilcoxon FDR-adjusted p < 2.2 × 10–16). Furthermore, peaks in cluster 1 (those closest to annotated TSS) had a higher GC and CpG island content than clusters 2 and 3 (Wilcoxon FDR-adjusted p < 2.2 × 10–16) and higher than enhancers (Wilcoxon FDR-adjusted p < 2.2 × 10–16), consistent with known features of mammalian promoters (Appendix 1—figure 4). H3K4me3 ChIP-seq peak intensity has previously been correlated with transcriptional activity (Barski et al., 2007) and a higher peak intensity has also been observed in promoters in mammalian liver tissue (Villar et al., 2015). We thus used cluster 1 (12,295 peaks) to define a set of high-confidence promoters for all of the following analyses.

Investigating differences in the number of mouse and dunnart peaks called

After applying consistent peak calling and filtering parameters to the dunnart ChIP-seq data, we found that the number of peaks was fairly consistent across mouse embryonic stages (~20–30,000 total peaks; Figure 4A). This number was significantly lower than the total number of peaks observed in the dunnart (~150,000 peaks; Figure 4A). We investigated this further and found that the strength and specificity of enrichment differed between the mouse and dunnart datasets. Dunnart alignment BAM files for H3K4me3 and H3K27ac immunoprecipitation replicates show strong and specific enrichment (Appendix 1—figure 2); however, mouse alignment BAM files show a weaker enrichment with the immunoprecipitation samples for H3K27ac being closer to the input (Appendix 1—figure 5). This was also reflected in the similarity between mouse IPs and input controls based on read coverage, with correlation coefficients higher (H3K4me3, Pearson’s r range = 0.17–0.32, Appendix 1—figure 6; H3K27ac, Pearson’s r range = 0.39–0.64, Appendix 1—figure 6) than the corresponding dunnart values (Appendix 1—figure 2). Therefore, this is likely a technical confounder that may be responsible for the lower numbers of enriched regions called in the mouse, as peaks with lower enrichment signal might not be identified by the peak caller. This was consistent with the distribution of peak enrichment values in the dunnart and mouse for H3K4me3, with a lower mean peak enrichment value in the dunnart compared to all mouse stages (Kruskal–Wallis FDR-corrected p = 6.7 × 10–12; Appendix 1—figure 2). However, this was not the case for peak enrichment values for H3K27ac peaks (Appendix 1—figure 2), which generally have lower peak enrichment values than observed in H3K4me3 (Villar et al., 2015; Kim et al., 2010)

Whole-genome alignment between mouse and dunnart

In order to compare activity in orthologous mouse and dunnart peaks, liftOver chains are needed to find the corresponding coordinates in the alternative genome. After building a dunnart-mm10 liftOver chain, we examined the quality of the WGA between mouse and dunnart. Approximately a quarter of both genomes was recovered in the WGA, with 26% of the dunnart genome and 28% of the mouse genome present with an average mismatch between the dunnart and mouse of 37.2% (Appendix 1—figure 7). Previous WGAs between placentals and marsupials (opossum vs. human, tammar wallaby vs. human) have recovered approximately 6–8% of the genomes with LastZ (Howe et al., 2021). The higher recovery in our WGA may be due to the use of new genome alignment tools like chainCleaner and patchChain. Given the higher conservation of coding regions than non-coding regions, exon coverage is a useful metric for assessing the quality of the alignment. 92% of mouse exon sequence and 69% of dunnart exon sequence was present in the WGA (Appendix 1—figure 7). Coverage of the exons recovered after liftOver was 82.7% and 57.8% in the mouse and dunnart, respectively (Appendix 1—figure 7). The lower exon coverage in the dunnart could be due to the missingness in the dunnart annotation. The majority of the alignment blocks fell within the 100 bp to 1000 bp size range with a total of 578.3 Mb of DNA sequence alignment out of a total of 801 Mb in the alignment within this range (Appendix 1—figure 7). This is very similar to the existing opossum-human WGA where 83% of alignment blocks fall within the size range of 100–1000 bp (Howe et al., 2021).

We then used the liftOver chains to find corresponding coordinates for dunnart peaks in the mouse genome. Given the majority of the alignment blocks for the mouse and dunnart WGA fell between 100 and 1000 bp in length (Appendix 1—figure 7), we first assessed the impact of peak length (centered on the peak summit, which represents the point with the highest signal in the peak region) on interspecies recovery of peaks. We defined dunnart peak regions as alignable if they could be reciprocally lifted over to the mouse genome and then back to the dunnart genome. Unsurprisingly, for both enhancers and promoters, the number of alignable peaks recovered decreased with increasing peak length (Supplementary file 1v and w). Promoters had the highest number of regions alignable (17%) using 50 bp peak summit widths, which decreased to 1.8% with a 500 bp peak summit length (Supplementary file 1w). Similarly, 19% of enhancers were alignable using a 50-bp peak summit length, and this decreased to 2% with 500 bp peak summit length (Supplementary file 1v). To assess the accuracy of the reciprocal liftOver, we next investigated whether alignable peaks in the mouse and dunnart mapped to the same nearest Ensembl gene before and after liftOver. Accuracy was high for both the promoters and enhancers (from 94% to 100%; Supplementary file 1v, w).

Appendix 1—figure 1. Validation of antibodies in dunnart craniofacial tissue with immunofluorescence and qPCR.

Appendix 1—figure 1.

Localization of (A) H3K27ac (pink, scale bar = 250 µm) and (B) H3K4me3 (pink, scale bar = 80 µm), in D0 dunnart head sections with nuclei stained with DAPI (blue) (C) enhancer regions expected to be enriched in the IP samples presented as the percentage of the input control sample as measured by qPCR.

Appendix 1—figure 2. deepTools quality control plots for dunnart subsampled aligned BAM files.

Appendix 1—figure 2.

(A). Overall similarity between BAM files based on read coverage within genomic regions with Pearson correlation coefficients plotted for H3K27ac, H3K4me3, and input control. (B) Fingerprint plot showing a profile of cumulative read coverages for each BAM file. All reads overlapping a window (bin) of the specified length are counted, sorted, and plotted for H3K27ac, H3K4me3, and input control.

Appendix 1—figure 3. Number of peaks per gene for enhancer- and promoter-associated peaks.

Appendix 1—figure 3.

(A) Log10 number of peaks per gene. Genes with greater than 50 peaks are noted. (B) Scatter plot with distance to the closest gene on the x-axis and number of peaks per gene on the y-axis. There is a weak but significant correlation between the number of peaks per gene and the distance to the next closest gene in both enhancers and promoters.

Appendix 1—figure 4. Subsetting high-confidence promoter-associated peaks with k-means clustering.

Appendix 1—figure 4.

(A) Barplot for the number of promoter-associated peaks per cluster and histogram showing the distribution of peak distance from the nearest TSS in each cluster. (B) Genomic annotations for promoter-associated peaks in each cluster. (C) GC content, (D) Log10 peak length, (E) CpG content, and (F) Log10 of clustered promoter-associated peaks, enhancer-associated peaks, and unclustered promoter-associated peaks.

Appendix 1—figure 5. deepTools fingerprint plot for mouse subsampled aligned BAM files.

Appendix 1—figure 5.

Cumulative read coverages for each BAM file. All reads overlapping a window (bin) of the specified length are counted, sorted, and plotted for (A) H3K27ac and (B) H3K4me3.

Appendix 1—figure 6. deepTools correlation plots for mouse subsampled aligned BAM files.

Appendix 1—figure 6.

Overall similarity between BAM files based on read coverage within genomic regions with Pearson correlation coefficients plotted for (A) H3K27ac and (B) H3K4me3.

Appendix 1—figure 7. Genome alignment between the fat-tailed dunnart Sminthopsis crassicaudata and mouse (Mus musculus).

Appendix 1—figure 7.

(A) Genome alignment workflow. (B) Genome coverage and exon coverage (bp) between mouse and dunnart. Coverage of exons recovered after LiftOver. (C) Block size distribution for the dunnart including number of blocks and total size of the blocks.

Funding Statement

The funders had no role in study design, data collection, and interpretation, or the decision to submit the work for publication.

Contributor Information

Laura E Cook, Email: lecook@lbl.gov.

Irene Gallego Romero, Email: irene.gallego@svi.edu.au.

Detlef Weigel, Max Planck Institute for Biology Tübingen, Germany.

Detlef Weigel, Max Planck Institute for Biology Tübingen, Germany.

Funding Information

This paper was supported by the following grants:

  • Australian Government Research Training Program Scholarship to Laura E Cook.

  • Australian Research Council DP160103683 to Andrew J Pask.

Additional information

Competing interests

No competing interests declared.

is funded in part by Colossal Biosciences. However, Colossal Biosciences had no direct influence on this project design, data collection, analysis, interpretation, or conclusions presented in this paper.

Author contributions

Conceptualization, Data curation, Formal analysis, Investigation, Visualization, Methodology, Writing – original draft, Writing – review and editing.

Formal analysis, Investigation, Writing – review and editing.

Investigation.

Formal analysis.

Conceptualization, Resources, Supervision, Funding acquisition, Writing – review and editing.

Resources, Supervision, Funding acquisition, Writing – original draft, Writing – review and editing.

Ethics

All animal procedures, including laboratory breeding, were conducted in accordance with the current Australian Code for the Care and Use of Animals for Scientific Purposes and were approved by The University of Melbourne Animal Ethics Committee (AEC: 1513686.2) and with the appropriate Wildlife Permit (number 10008652) from the Department of Environment, Land, Water and Planning.

Additional files

Supplementary file 1. All supplementary tables.

(a) Annotated dunnart peak nearest gene ID conversions. (b) Enriched gene ontology terms for dunnart genes near promoters. (c) Enriched gene ontology terms for dunnart genes near enhancers. (d) Dunnart enhancers and nearest gene expression (TPM). (e) Dunnart promoters and nearest gene expression (TPM). (f) Enriched GO terms for dunnart enhancers and promoters where nearest gene expression ≥10 TPM (FDR <1%). (g) Summary of mouse–dunnart conserved peak activity for a. high-confidence putative promoters and b. putative enhancers. (h) Enhancers with conserved activity in mouse and dunnart with incorporated stage and species gene expression values (TPM). (i) Gene ontology terms for biological processes enriched in genes expressed in mouse and dunnart (>1 TPM) near conserved enhancers. (j) Dunnart-specific enhancers where the nearest genes were highly expressed (>10 TPM) only in the dunnart with low to no activity in the mouse at any of the embryonic stages (E10.5–E15.5). (k) Highly expressed genes (TPM) in dunnart where expression is lower or absent in mouse. (l) Biological processes enrichment for highly expressed genes in dunnart where expression is lower or absent in mouse. (m) Dunnart facial prominence tissue samples. (n) qPCR primer sequences for ChIP enrichment validation. (o) Details of libraries used in genome assembly. (p) DNA fragments isolated after pull down were sequenced by GENEWIZ. Paired-end sequencing at an average depth of 57 million reads. a. Raw reads sequencing statistics, b. Read statistics after CutAdapt filtering. (q) ENCODE mouse embryonic facial prominence ChIP-seq and gene expression quantification file accession numbers and details used in study. (r) Read alignment quality metrics for H3K4me3 and H3K27ac dunnart ChIP-seq samples. (s) Read alignment quality metrics for mouse ENCODE ChIP-seq data. (t) Peak calling quality metrics for H3K4me3 and H3K27ac dunnart ChIP-seq samples. (u) Peak calling quality metrics for mouse ChIP-seq peaks after peak calling with MACS2. (v) Summary of alignable enhancer-associated peaks in the mouse and dunnart. (w) Summary of alignable high-confidence promoter-associated peaks in mouse and dunnart.

elife-103592-supp1.xlsx (3.3MB, xlsx)
MDAR checklist

Data availability

The data generated in this study have been deposited in NCBI’s Gene Expression Omnibus (Barrett et al., 2013) and are available through GEO Series accession number GSE188990. Accession IDs for previously published datasets from the ENCODE consortium (He et al., 2020) are given in Supplementary file 1. Processed data and an IGV session for browsing are available in a figshare repository. All analyses described were carried out using custom bash, Python3, and R v4.1.0 scripts and are available at https://github.com/lecook/chipseq-cross-species (copy archived at Cook, 2025).

The following datasets were generated:

Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. dunnart genome. figshare.

Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. Gene regulatory dynamics during craniofacial development in a carnivorous marsupial. NCBI Gene Expression Omnibus. GSE188990

Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. dunnart and mouse peak characterisation. figshare.

Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. tc-seq processed data. figshare.

Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. liftover chains. figshare.

References

  1. Aboalola D, Han VKM. Insulin-like growth factor binding protein-6 alters skeletal muscle differentiation of human mesenchymal stem cells. Stem Cells International. 2017;2017:2348485. doi: 10.1155/2017/2348485. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Andrews S. FastQC: a quality control tool for high throughput sequence data [online] v3Babraham Bioinformatics. 2010 https://www.bioinformatics.babraham.ac.uk/projects/fastqc/
  3. Andrews G, Fan K, Pratt HE, Phalke N, Karlsson EK, Lindblad-Toh K, Gazal S, Moore JE, Weng Z, Zoonomia Consortium§ Mammalian evolution of human cis-regulatory elements and transcription factor binding sites. Science. 2023;380:eabn7930. doi: 10.1126/science.abn7930. [DOI] [PubMed] [Google Scholar]
  4. Attanasio C, Nord AS, Zhu Y, Blow MJ, Li Z, Liberton DK, Morrison H, Plajzer-Frick I, Holt A, Hosseini R, Phouanenavong S, Akiyama JA, Shoukry M, Afzal V, Rubin EM, FitzPatrick DR, Ren B, Hallgrímsson B, Pennacchio LA, Visel A. Fine tuning of craniofacial morphology by distant-acting enhancers. Science. 2013;342:1241006. doi: 10.1126/science.1241006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Bancroft JD, Gamble M. Theory and Practice of Histological Techniques. Elsevier Health Sciences; 2008. [Google Scholar]
  6. Barrett T, Wilhite SE, Ledoux P, Evangelista C, Kim IF, Tomashevsky M, Marshall KA, Phillippy KH, Sherman PM, Holko M, Yefanov A, Lee H, Zhang N, Robertson CL, Serova N, Davis S, Soboleva A. NCBI GEO: archive for functional genomics data sets—update. Nucleic Acids Research. 2013;41:D991–D995. doi: 10.1093/nar/gks1193. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Barski A, Cuddapah S, Cui K, Roh TY, Schones DE, Wang Z, Wei G, Chepelev I, Zhao K. High-resolution profiling of histone methylations in the human genome. Cell. 2007;129:823–837. doi: 10.1016/j.cell.2007.05.009. [DOI] [PubMed] [Google Scholar]
  8. Bentzinger CF, Wang YX, Rudnicki MA. Building muscle: molecular regulation of myogenesis. Cold Spring Harbor Perspectives in Biology. 2012;4:a008342. doi: 10.1101/cshperspect.a008342. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Bininda-Emonds ORP, Cardillo M, Jones KE, MacPhee RDE, Beck RMD, Grenyer R, Price SA, Vos RA, Gittleman JL, Purvis A. The delayed rise of present-day mammals. Nature. 2007;446:507–512. doi: 10.1038/nature05634. [DOI] [PubMed] [Google Scholar]
  10. Bolger AM, Lohse M, Usadel B. Trimmomatic: a flexible trimmer for Illumina sequence data. Bioinformatics. 2014;30:2114–2120. doi: 10.1093/bioinformatics/btu170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Brandies PA, Tang S, Johnson RSP, Hogg CJ, Belov K. The first Antechinus reference genome provides a resource for investigating the genetic basis of semelparity and age-related neuropathologies. GigaByte. 2020;2020:gigabyte7. doi: 10.46471/gigabyte.7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Brinkley JF, Fisher S, Harris MP, Holmes G, Hooper JE, Jabs EW, Jones KL, Kesselman C, Klein OD, Maas RL, Marazita ML, Selleri L, Spritz RA, van Bakel H, Visel A, Williams TJ, Wysocka J, Chai Y, the FaceBase Consortium The FaceBase Consortium: a comprehensive resource for craniofacial researchers. Development. 2016;143:2677–2688. doi: 10.1242/dev.135434. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Buckingham M. Gene regulatory networks and cell lineages that underlie the formation of skeletal muscle. PNAS. 2017;114:5830–5837. doi: 10.1073/pnas.1610605114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Cain CE, Blekhman R, Marioni JC, Gilad Y. Gene expression differences among primates are associated with changes in a histone epigenetic modification. Genetics. 2011;187:1225–1234. doi: 10.1534/genetics.110.126177. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Calo E, Wysocka J. Modification of enhancer chromatin: what, how, and why? Molecular Cell. 2013;49:825–837. doi: 10.1016/j.molcel.2013.01.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Carlson M. Org.mm.eg.db: genome wide annotation for mouse. R Package version 3.12.0Bioconductor. 2020 https://bioconductor.org/packages/org.Mm.eg.db/
  17. Carroll SH. Neural crest and periderm-specific requirements of irf6 during neural tube and craniofacial development. bioRxiv. 2024 doi: 10.1101/2024.06.11.598425. [DOI] [PMC free article] [PubMed]
  18. Clark CT, Smith KK. Cranial osteogenesis in Monodelphis domestica (Didelphidae) and Macropus eugenii (Macropodidae) Journal of Morphology. 1993;215:119–149. doi: 10.1002/jmor.1052150203. [DOI] [PubMed] [Google Scholar]
  19. Cook LE, Newton AH, Hipsley CA, Pask AJ. Postnatal development in a marsupial model, the fat-tailed dunnart (Sminthopsis crassicaudata; Dasyuromorphia: Dasyuridae) Communications Biology. 2021;4:1028. doi: 10.1038/s42003-021-02506-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Cook LE. Chipseq-cross-species. swh:1:rev:74138447485798537e187a5c531e5d1dc1e85c11Software Heritage. 2025 https://archive.softwareheritage.org/swh:1:dir:a8d371f866b38955d3bbb3e2997cc4a9903ef938;origin=https://github.com/lecook/chipseq-cross-species;visit=swh:1:snp:ffdcfcb4c42639233df7d8e2f38bdc40fe7859f2;anchor=swh:1:rev:74138447485798537e187a5c531e5d1dc1e85c11
  21. Creyghton MP, Cheng AW, Welstead GG, Kooistra T, Carey BW, Steine EJ, Hanna J, Lodato MA, Frampton GM, Sharp PA, Boyer LA, Young RA, Jaenisch R. Histone H3K27ac separates active from poised enhancers and predicts developmental state. PNAS. 2010;107:21931–21936. doi: 10.1073/pnas.1016071107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Crusoe MR, Alameldin HF, Awad S, Boucher E, Caldwell A, Cartwright R, Charbonneau A, Constantinides B, Edvenson G, Fay S, Fenton J, Fenzl T, Fish J, Garcia-Gutierrez L, Garland P, Gluck J, González I, Guermond S, Guo J, Gupta A, Herr JR, Howe A, Hyer A, Härpfer A, Irber L, Kidd R, Lin D, Lippi J, Mansour T, McA’Nulty P, McDonald E, Mizzi J, Murray KD, Nahum JR, Nanlohy K, Nederbragt AJ, Ortiz-Zuazaga H, Ory J, Pell J, Pepe-Ranney C, Russ ZN, Schwarz E, Scott C, Seaman J, Sievert S, Simpson J, Skennerton CT, Spencer J, Srinivasan R, Standage D, Stapleton JA, Steinman SR, Stein J, Taylor B, Trimble W, Wiencko HL, Wright M, Wyss B, Zhang Q, Zyme E, Brown CT. The khmer software package: enabling efficient nucleotide sequence analysis. F1000Research. 2015;4:900. doi: 10.12688/f1000research.6924.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Desmarais MJ, Beauregard F, Cabana T, Pflieger JF. Facial mechanosensory influence on forelimb movement in newborn opossums, monodelphis domestica. PLOS ONE. 2016;11:e0148352. doi: 10.1371/journal.pone.0148352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Doucet YS, Woo SH, Ruiz ME, Owens DM. The touch dome defines an epidermal niche specialized for mechanosensory signaling. Cell Reports. 2013;3:1759–1765. doi: 10.1016/j.celrep.2013.04.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Durinck S, Moreau Y, Kasprzyk A, Davis S, De Moor B, Brazma A, Huber W. BioMart and Bioconductor: a powerful link between biological databases and microarray data analysis. Bioinformatics. 2005;21:3439–3440. doi: 10.1093/bioinformatics/bti525. [DOI] [PubMed] [Google Scholar]
  26. Durinck S, Spellman PT, Birney E, Huber W. Mapping identifiers for the integration of genomic datasets with the R/Bioconductor package biomaRt. Nature Protocols. 2009;4:1184–1191. doi: 10.1038/nprot.2009.97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Dutheil JY, Gaillard S, Stukenbrock EH. MafFilter: a highly flexible and extensible multiple genome alignment files processor. BMC Genomics. 2014;15:53. doi: 10.1186/1471-2164-15-53. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Feigin CY, Newton AH, Doronina L, Schmitz J, Hipsley CA, Mitchell KJ, Gower G, Llamas B, Soubrier J, Heider TN, Menzies BR, Cooper A, O’Neill RJ, Pask AJ. Genome of the Tasmanian tiger provides insights into the evolution and demography of an extinct marsupial carnivore. Nature Ecology & Evolution. 2018;2:182–192. doi: 10.1038/s41559-017-0417-y. [DOI] [PubMed] [Google Scholar]
  29. Feigin C, Frankenberg S, Pask A. A chromosome-scale hybrid genome assembly of the extinct tasmanian tiger (Thylacinus cynocephalus) Genome Biology and Evolution. 2022;14:evac048. doi: 10.1093/gbe/evac048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Feng HZ, Jin JP. Carbonic anhydrase III is expressed in mouse skeletal muscles independent of fiber type-specific myofilament protein isoforms and plays a role in fatigue resistance. Frontiers in Physiology. 2016;7:597. doi: 10.3389/fphys.2016.00597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Ferner K. Skin structure in newborn marsupials with focus on cutaneous gas exchange. Journal of Anatomy. 2018;233:311–327. doi: 10.1111/joa.12843. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Fishilevich S, Nudel R, Rappaport N, Hadar R, Plaschkes I, Iny Stein T, Rosen N, Kohn A, Twik M, Safran M, Lancet D, Cohen D. GeneHancer: genome-wide integration of enhancers and target genes in GeneCards. Database. 2017;2017:bax028. doi: 10.1093/database/bax028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Francis-West P, Crespo-Enriquez I. Vertebrate Embryo: Craniofacial Development. John Wiley & Sons; 2016. [DOI] [Google Scholar]
  34. Gemmell RT, Nelson J. Ultrastructure of the olfactory system of three newborn marsupial species. The Anatomical Record. 1989;225:203–208. doi: 10.1002/ar.1092250305. [DOI] [PubMed] [Google Scholar]
  35. Genereux DP, Serres A, Armstrong J, Johnson J, Marinescu VD, Murén E, Juan D, Bejerano G, Casewell NR, Chemnick LG, Damas J, Di Palma F, Diekhans M, Fiddes IT, Garber M, Gladyshev VN, Goodman L, Haerty W, Houck ML, Hubley R, Kivioja T, Koepfli K-P, Kuderna LFK, Lander ES, Meadows JRS, Murphy WJ, Nash W, Noh HJ, Nweeia M, Pfenning AR, Pollard KS, Ray DA, Shapiro B, Smit AFA, Springer MS, Steiner CC, Swofford R, Taipale J, Teeling EC, Turner-Maier J, Alfoldi J, Birren B, Ryder OA, Lewin HA, Paten B, Marques-Bonet T, Lindblad-Toh K, Karlsson EK, Zoonomia Consortium A comparative genomics multitool for scientific discovery and conservation. Nature. 2020;587:240–245. doi: 10.1038/s41586-020-2876-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Goswami A, Randau M, Polly PD, Weisbecker V, Bennett CV, Hautier L, Sánchez-Villagra MR. Do developmental constraints and high integration limit the evolution of the marsupial oral apparatus? Integrative and Comparative Biology. 2016;56:404–415. doi: 10.1093/icb/icw039. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Haering C, Kanageswaran N, Bouvain P, Scholz P, Altmüller J, Becker C, Gisselmann G, Wäring-Bischof J, Hatt H. Ion transporter NKCC1, modulator of neurogenesis in murine olfactory neurons. The Journal of Biological Chemistry. 2015;290:9767–9779. doi: 10.1074/jbc.M115.640656. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Hammond NL, Dixon J, Dixon MJP. Periderm: Life-cycle and function during orofacial and epidermal development. Seminars in Cell & Developmental Biology. 2019;91:75–83. doi: 10.1016/j.semcdb.2017.08.021. [DOI] [PubMed] [Google Scholar]
  39. He Y, Hariharan M, Gorkin DU, Dickel DE, Luo C, Castanon RG, Nery JR, Lee AY, Zhao Y, Huang H, Williams BA, Trout D, Amrhein H, Fang R, Chen H, Li B, Visel A, Pennacchio LA, Ren B, Ecker JR. Spatiotemporal DNA methylome dynamics of the developing mouse fetus. Nature. 2020;583:752–759. doi: 10.1038/s41586-020-2119-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Heintzman ND, Hon GC, Hawkins RD, Kheradpour P, Stark A, Harp LF, Ye Z, Lee LK, Stuart RK, Ching CW, Ching KA, Antosiewicz-Bourget JE, Liu H, Zhang X, Green RD, Lobanenkov VV, Stewart R, Thomson JA, Crawford GE, Kellis M, Ren B. Histone modifications at human enhancers reflect global cell-type-specific gene expression. Nature. 2009;459:108–112. doi: 10.1038/nature07829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Heinz S, Benner C, Spann N, Bertolino E, Lin YC, Laslo P, Cheng JX, Murre C, Singh H, Glass CK. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Molecular Cell. 2010;38:576–589. doi: 10.1016/j.molcel.2010.05.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Howe KL, Achuthan P, Allen J, Allen J, Alvarez-Jarreta J, Amode MR, Armean IM, Azov AG, Bennett R, Bhai J, Billis K, Boddu S, Charkhchi M, Cummins C, Da Rin Fioretto L, Davidson C, Dodiya K, El Houdaigui B, Fatima R, Gall A, Garcia Giron C, Grego T, Guijarro-Clarke C, Haggerty L, Hemrom A, Hourlier T, Izuogu OG, Juettemann T, Kaikala V, Kay M, Lavidas I, Le T, Lemos D, Gonzalez Martinez J, Marugán JC, Maurel T, McMahon AC, Mohanan S, Moore B, Muffato M, Oheh DN, Paraschas D, Parker A, Parton A, Prosovetskaia I, Sakthivel MP, Salam AIA, Schmitt BM, Schuilenburg H, Sheppard D, Steed E, Szpak M, Szuba M, Taylor K, Thormann A, Threadgold G, Walts B, Winterbottom A, Chakiachvili M, Chaubal A, De Silva N, Flint B, Frankish A, Hunt SE, IIsley GR, Langridge N, Loveland JE, Martin FJ, Mudge JM, Morales J, Perry E, Ruffier M, Tate J, Thybert D, Trevanion SJ, Cunningham F, Yates AD, Zerbino DR, Flicek P. Ensembl 2021. Nucleic Acids Research. 2021;49:D884–D891. doi: 10.1093/nar/gkaa942. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Huang W, Yang S, Shao J, Li YP. Signaling and transcriptional regulation in osteoblast commitment and differentiation. Frontiers in Bioscience. 2007;12:3068–3092. doi: 10.2741/2296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Hüppi E, Sánchez-Villagra MR, Tzika AC, Werneburg I. Ontogeny and phylogeny of the mammalian chondrocranium: the cupula nasi anterior and associated structures of the anterior head region. Zoological Letters. 2018;4:29. doi: 10.1186/s40851-018-0112-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Jenkins BA, Fontecilla NM, Lu CP, Fuchs E, Lumpkin EA. The cellular basis of mechanosensory Merkel-cell innervation during development. eLife. 2019;8:e42633. doi: 10.7554/eLife.42633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Jin S, Kim J, Willert T, Klein-Rodewald T, Garcia-Dominguez M, Mosqueira M, Fink R, Esposito I, Hofbauer LC, Charnay P, Kieslinger M. Ebf factors and MyoD cooperate to regulate muscle relaxation via Atp2a1. Nature Communications. 2014;5:3793. doi: 10.1038/ncomms4793. [DOI] [PubMed] [Google Scholar]
  47. Jones TE, Munger BL. Early differentiation of the afferent nervous system in glabrous snout skin of the opossum (Monodelphis domesticus) Somatosensory Research. 1985;3:169–184. doi: 10.3109/07367228509144582. [DOI] [PubMed] [Google Scholar]
  48. Jung SY, Ko YG. TRIM72, a novel negative feedback regulator of myogenesis, is transcriptionally activated by the synergism of MyoD (or myogenin) and MEF2. Biochemical and Biophysical Research Communications. 2010;396:238–245. doi: 10.1016/j.bbrc.2010.04.072. [DOI] [PubMed] [Google Scholar]
  49. Karlen SJ, Krubitzer L. The functional and anatomical organization of marsupial neocortex: evidence for parallel evolution across mammals. Progress in Neurobiology. 2007;82:122–141. doi: 10.1016/j.pneurobio.2007.03.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Kent WJ, Baertsch R, Hinrichs A, Miller W, Haussler D. Evolution’s cauldron: Duplication, deletion, and rearrangement in the mouse and human genomes. PNAS. 2003;100:11484–11489. doi: 10.1073/pnas.1932072100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Kim TK, Hemberg M, Gray JM, Costa AM, Bear DM, Wu J, Harmin DA, Laptewicz M, Barbara-Haley K, Kuersten S, Markenscoff-Papadimitriou E, Kuhl D, Bito H, Worley PF, Kreiman G, Greenberg ME. Widespread transcription at neuronal activity-regulated enhancers. Nature. 2010;465:182–187. doi: 10.1038/nature09033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Koch F, Andrau JC. Initiating RNA polymerase II and TIPs as hallmarks of enhancer activity and tissue-specificity. Transcription. 2011;2:263–268. doi: 10.4161/trns.2.6.18747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Kolmogorov M, Yuan J, Lin Y, Pevzner PA. Assembly of long, error-prone reads using repeat graphs. Nature Biotechnology. 2019;37:540–546. doi: 10.1038/s41587-019-0072-8. [DOI] [PubMed] [Google Scholar]
  54. Kousa YA, Roushangar R, Patel N, Walter A, Marangoni P, Krumlauf R, Klein OD, Schutte BC. IRF6 and SPRY4 signaling interact in periderm development. Journal of Dental Research. 2017;96:1306–1313. doi: 10.1177/0022034517719870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Koyabu D, Werneburg I, Morimoto N, Zollikofer CPE, Forasiepi AM, Endo H, Kimura J, Ohdachi SD, Truong Son N, Sánchez-Villagra MR. Mammalian skull heterochrony reveals modular evolution and a link between cranial development and brain size. Nature Communications. 2014;5:3625. doi: 10.1038/ncomms4625. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Krause WJ, Cutts JH, Leeson CR. Postnatal development of the epidermis in a marsupial, Didelphis virginiana. Journal of Anatomy. 1978;125:85–99. [PMC free article] [PubMed] [Google Scholar]
  57. Langmead B, Salzberg SL. Fast gapped-read alignment with Bowtie 2. Nature Methods. 2012;9:357–359. doi: 10.1038/nmeth.1923. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Lefebvre V, Angelozzi M, Haseeb A. SOX9 in cartilage development and disease. Current Opinion in Cell Biology. 2019;61:39–47. doi: 10.1016/j.ceb.2019.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Li H, Handsaker B, Wysoker A, Fennell T, Ruan J, Homer N, Marth G, Abecasis G, Durbin R, 1000 Genome Project Data Processing Subgroup The Sequence Alignment/Map format and SAMtools. Bioinformatics. 2009;25:2078–2079. doi: 10.1093/bioinformatics/btp352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Liao Y, Smyth GK, Shi W. The Subread aligner: fast, accurate and scalable read mapping by seed-and-vote. Nucleic Acids Research. 2013;41:e108. doi: 10.1093/nar/gkt214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Lillegraven JA. Biological considerations of the marsupial-placental dichotomy. Evolution; International Journal of Organic Evolution. 1975;29:707–722. doi: 10.1111/j.1558-5646.1975.tb00865.x. [DOI] [PubMed] [Google Scholar]
  62. Lin CY, Yung RF, Lee HC, Chen WT, Chen YH, Tsai HJ. Myogenic regulatory factors Myf5 and Myod function distinctly during craniofacial myogenesis of zebrafish. Developmental Biology. 2006;299:594–608. doi: 10.1016/j.ydbio.2006.08.042. [DOI] [PubMed] [Google Scholar]
  63. Liu H, Leslie EJ, Jia Z, Smith T, Eshete M, Butali A, Dunnwald M, Murray J, Cornell RA. Irf6 directly regulates Klf17 in zebrafish periderm and KLF4 in murine oral epithelium, and dominant-negative KLF4 variants are present in patients with cleft lip and palate. Human Molecular Genetics. 2016;25:766–776. doi: 10.1093/hmg/ddv614. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Liu X, Liu S. Cholecystokinin selectively activates short axon cells to enhance inhibition of olfactory bulb output neurons. The Journal of Physiology. 2018;596:2185–2207. doi: 10.1113/JP275511. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Luo ZX, Yuan CX, Meng QJ, Ji Q. A Jurassic eutherian mammal and divergence of marsupials and placentals. Nature. 2011;476:442–445. doi: 10.1038/nature10291. [DOI] [PubMed] [Google Scholar]
  66. Martin M. Cutadapt removes adapter sequences from high-throughput sequencing reads. EMBnet.Journal. 2011;17:10. doi: 10.14806/ej.17.1.200. [DOI] [Google Scholar]
  67. Mikkelsen TS, Wakefield MJ, Aken B, Amemiya CT, Chang JL, Duke S, Garber M, Gentles AJ, Goodstadt L, Heger A, Jurka J, Kamal M, Mauceli E, Searle SMJ, Sharpe T, Baker ML, Batzer MA, Benos PV, Belov K, Clamp M, Cook A, Cuff J, Das R, Davidow L, Deakin JE, Fazzari MJ, Glass JL, Grabherr M, Greally JM, Gu W, Hore TA, Huttley GA, Kleber M, Jirtle RL, Koina E, Lee JT, Mahony S, Marra MA, Miller RD, Nicholls RD, Oda M, Papenfuss AT, Parra ZE, Pollock DD, Ray DA, Schein JE, Speed TP, Thompson K, VandeBerg JL, Wade CM, Walker JA, Waters PD, Webber C, Weidman JR, Xie X, Zody MC, Graves JAM, Ponting CP, Breen M, Samollow PB, Lander ES, Lindblad-Toh K, Broad Institute Genome Sequencing Platform. Broad Institute Whole Genome Assembly Team Genome of the marsupial Monodelphis domestica reveals innovation in non-coding sequences. Nature. 2007;447:167–177. doi: 10.1038/nature05805. [DOI] [PubMed] [Google Scholar]
  68. Murchison EP, Schulz-Trieglaff OB, Ning Z, Alexandrov LB, Bauer MJ, Fu B, Hims M, Ding Z, Ivakhno S, Stewart C, Ng BL, Wong W, Aken B, White S, Alsop A, Becq J, Bignell GR, Cheetham RK, Cheng W, Connor TR, Cox AJ, Feng ZP, Gu Y, Grocock RJ, Harris SR, Khrebtukova I, Kingsbury Z, Kowarsky M, Kreiss A, Luo S, Marshall J, McBride DJ, Murray L, Pearse AM, Raine K, Rasolonjatovo I, Shaw R, Tedder P, Tregidgo C, Vilella AJ, Wedge DC, Woods GM, Gormley N, Humphray S, Schroth G, Smith G, Hall K, Searle SMJ, Carter NP, Papenfuss AT, Futreal PA, Campbell PJ, Yang F, Bentley DR, Evers DJ, Stratton MR. Genome sequencing and analysis of the Tasmanian devil and its transmissible cancer. Cell. 2012;148:780–791. doi: 10.1016/j.cell.2011.11.065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. NHMRC . Australian Code for the Care and Use of Animals for Scientific Purposes. National Health And Medical Council; 2013. [Google Scholar]
  70. Noiret M, Mottier S, Angrand G, Gautier-Courteille C, Lerivray H, Viet J, Paillard L, Mereau A, Hardy S, Audic Y. Ptbp1 and Exosc9 knockdowns trigger skin stability defects through different pathways. Developmental Biology. 2016;409:489–501. doi: 10.1016/j.ydbio.2015.11.002. [DOI] [PubMed] [Google Scholar]
  71. Nunn CL, Smith KK. Statistical analyses of developmental sequences: the craniofacial region in marsupial and placental mammals. The American Naturalist. 1998;152:82–101. doi: 10.1086/286151. [DOI] [PubMed] [Google Scholar]
  72. Oksanen J. vegan: Community Ecology Package. 2.7-1CRAN. 2020 https://cran.r-project.org/package=vegan
  73. Peel E, Silver L, Brandies P, Hogg CJ, Belov K. A reference genome for the critically endangered woylie, Bettongia penicillata ogilbyi. GigaByte. 2021;2021:gigabyte35. doi: 10.46471/gigabyte.35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Pekowska A, Benoukraf T, Zacarias-Cabeza J, Belhocine M, Koch F, Holota H, Imbert J. H3K4 tri-methylation provides an epigenetic signature of active enhancers: Epigenetic signature of active enhancers. The EMBO Journal. 2011;30:4198–4210. doi: 10.1038/emboj.2011.295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Pralomkarn T, Nelson J, Gemmell RT. Postnatal development of the skin of the marsupial native cat Dasyurus hallucatus. Journal of Morphology. 1990;205:233–242. doi: 10.1002/jmor.1052050210. [DOI] [PubMed] [Google Scholar]
  76. Prescott SL, Srinivasan R, Marchetto MC, Grishina I, Narvaiza I, Selleri L, Gage FH, Swigut T, Wysocka J. Enhancer divergence and cis-regulatory evolution in the human and chimp neural crest. Cell. 2015;163:68–83. doi: 10.1016/j.cell.2015.08.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Pryszcz LP, Gabaldón T. Redundans: an assembly pipeline for highly heterozygous genomes. Nucleic Acids Research. 2016;44:e113. doi: 10.1093/nar/gkw294. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Qi H, Aguiar DJ, Williams SM, La Pean A, Pan W, Verfaillie CM. Identification of genes responsible for osteoblast differentiation from human mesodermal progenitor cells. PNAS. 2003;100:3305–3310. doi: 10.1073/pnas.0532693100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Quinlan AR, Hall IM. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics. 2010;26:841–842. doi: 10.1093/bioinformatics/btq033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Rajderkar SS, Paraiso K, Amaral ML, Kosicki M, Cook LE, Darbellay F, Spurrell CH, Osterwalder M, Zhu Y, Wu H, Afzal SY, Blow MJ, Kelman G, Barozzi I, Fukuda-Yuzawa Y, Akiyama JA, Afzal V, Tran S, Plajzer-Frick I, Novak CS, Kato M, Hunter RD, von Maydell K, Wang A, Lin L, Preissl S, Lisgo S, Ren B, Dickel DE, Pennacchio LA, Visel A. Dynamic enhancer landscapes in human craniofacial development. Nature Communications. 2024;15:2030. doi: 10.1038/s41467-024-46396-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Renfree MB, Papenfuss AT, Deakin JE, Lindsay J, Heider T, Belov K, Rens W, Waters PD, Pharo EA, Shaw G, Wong ESW, Lefèvre CM, Nicholas KR, Kuroki Y, Wakefield MJ, Zenger KR, Wang C, Ferguson-Smith M, Nicholas FW, Hickford D, Yu H, Short KR, Siddle HV, Frankenberg SR, Chew KY, Menzies BR, Stringer JM, Suzuki S, Hore TA, Delbridge ML, Patel HR, Mohammadi A, Schneider NY, Hu Y, O’Hara W, Al Nadaf S, Wu C, Feng Z-P, Cocks BG, Wang J, Flicek P, Searle SMJ, Fairley S, Beal K, Herrero J, Carone DM, Suzuki Y, Sugano S, Toyoda A, Sakaki Y, Kondo S, Nishida Y, Tatsumoto S, Mandiou I, Hsu A, McColl KA, Lansdell B, Weinstock G, Kuczek E, McGrath A, Wilson P, Men A, Hazar-Rethinam M, Hall A, Davis J, Wood D, Williams S, Sundaravadanam Y, Muzny DM, Jhangiani SN, Lewis LR, Morgan MB, Okwuonu GO, Ruiz SJ, Santibanez J, Nazareth L, Cree A, Fowler G, Kovar CL, Dinh HH, Joshi V, Jing C, Lara F, Thornton R, Chen L, Deng J, Liu Y, Shen JY, Song X-Z, Edson J, Troon C, Thomas D, Stephens A, Yapa L, Levchenko T, Gibbs RA, Cooper DW, Speed TP, Fujiyama A, Graves JAM, O’Neill RJ, Pask AJ, Forrest SM, Worley KC. Genome sequence of an Australian kangaroo, Macropus eugenii, provides insight into the evolution of mammalian reproduction and development. Genome Biology. 2011;12:R81. doi: 10.1186/gb-2011-12-8-r81. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Robinson P, Lipscomb S, Preston LC, Altin E, Watkins H, Ashley CC, Redwood CS. Mutations in fast skeletal troponin I, troponin T, and beta-tropomyosin that cause distal arthrogryposis all increase contractile function. FASEB Journal. 2007;21:896–905. doi: 10.1096/fj.06-6899com. [DOI] [PubMed] [Google Scholar]
  83. Robinson MD, McCarthy DJ, Smyth GK. edgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26:139–140. doi: 10.1093/bioinformatics/btp616. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Roosenboom J, Hens G, Mattern BC, Shriver MD, Claes P. Exploring the underlying genetics of craniofacial morphology through various sources of knowledge. BioMed Research International. 2016;2016:3054578. doi: 10.1155/2016/3054578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Sambrook J, Fritsch EF, Maniatis T. Molecular Cloning: A Laboratory Manual. Cold Spring Harbor Laboratory Press; 1989. [Google Scholar]
  86. Samuels BD, Aho R, Brinkley JF, Bugacov A, Feingold E, Fisher S, Gonzalez-Reiche AS, Hacia JG, Hallgrimsson B, Hansen K, Harris MP, Ho TV, Holmes G, Hooper JE, Jabs EW, Jones KL, Kesselman C, Klein OD, Leslie EJ, Li H, Liao EC, Long H, Lu N, Maas RL, Marazita ML, Mohammed J, Prescott S, Schuler R, Selleri L, Spritz RA, Swigut T, van Bakel H, Visel A, Welsh I, Williams C, Williams TJ, Wysocka J, Yuan Y, Chai Y. FaceBase 3: analytical tools and FAIR resources for craniofacial and dental research. Development. 2020;147:dev191213. doi: 10.1242/dev.191213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Sánchez-Villagra MR, Forasiepi AM. On the development of the chondrocranium and the histological anatomy of the head in perinatal stages of marsupial mammals. Zoological Letters. 2017;3:1. doi: 10.1186/s40851-017-0062-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Santoro SW, Jakob S. Gene expression profiling of the olfactory tissues of sex-separated and sex-combined female and male mice. Scientific Data. 2018;5:180260. doi: 10.1038/sdata.2018.260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Santos-Rosa H, Schneider R, Bannister AJ, Sherriff J, Bernstein BE, Emre NCT, Schreiber SL, Mellor J, Kouzarides T. Active genes are tri-methylated at K4 of histone H3. Nature. 2002;419:407–411. doi: 10.1038/nature01080. [DOI] [PubMed] [Google Scholar]
  90. Schneider CA, Rasband WS, Eliceiri KW. NIH Image to ImageJ: 25 years of image analysis. Nature Methods. 2012;9:671–675. doi: 10.1038/nmeth.2089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Sears KE. Constraints on the morphological evolution of marsupial shoulder girdles. Evolution; International Journal of Organic Evolution. 2004;58:2353–2370. doi: 10.1111/j.0014-3820.2004.tb01609.x. [DOI] [PubMed] [Google Scholar]
  92. Sharma V, Hiller M. Increased alignment sensitivity improves the usage of genome alignments for comparative gene annotation. Nucleic Acids Research. 2017;45:8369–8377. doi: 10.1093/nar/gkx554. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Shumate A, Salzberg SLL. Liftoff: accurate mapping of gene annotations. Bioinformatics. 2021;37:1639–1643. doi: 10.1093/bioinformatics/btaa1016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Simeone A. Otx1 and Otx2 in the development and evolution of the mammalian brain. The EMBO Journal. 1998;17:6790–6798. doi: 10.1093/emboj/17.23.6790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Simpson SJ, Flecknoe SJ, Clugston RD, Greer JJ, Hooper SB, Frappell PB. Structural and functional development of the respiratory system in a newborn marsupial with cutaneous gas exchange. Physiological and Biochemical Zoology. 2011;84:634–649. doi: 10.1086/662557. [DOI] [PubMed] [Google Scholar]
  96. Smith KK. Development of craniofacial musculature in Monodelphis domestica (Marsupialia, Didelphidae) Journal of Morphology. 1994;222:149–173. doi: 10.1002/jmor.1052220204. [DOI] [PubMed] [Google Scholar]
  97. Smith KK. Comparative patterns of craniofacial development in eutherian and metatherian mammals. Evolution; International Journal of Organic Evolution. 1997;51:1663–1678. doi: 10.1111/j.1558-5646.1997.tb01489.x. [DOI] [PubMed] [Google Scholar]
  98. Smith KK. Early Cranial Development in Marsupial Mammals: The Origins of Heterochrony. American Zoologist; 1999. [Google Scholar]
  99. Smith KK. The evolution of mammalian development. Bulletin of the Museum of Comparative Zoology. 2001a;156:119–135. [Google Scholar]
  100. Smith KK. Early development of the neural plate, neural crest and facial region of marsupials. Journal of Anatomy. 2001b;199:121–131. doi: 10.1046/j.1469-7580.2001.19910121.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Smith KK. Time’s arrow: heterochrony and the evolution of development. The International Journal of Developmental Biology. 2003;47:613–621. [PubMed] [Google Scholar]
  102. Smith KK. Craniofacial development in marsupial mammals: developmental origins of evolutionary change. Developmental Dynamics. 2006;235:1181–1193. doi: 10.1002/dvdy.20676. [DOI] [PubMed] [Google Scholar]
  103. Smith KK, Keyte AL. Adaptations of the marsupial newborn: birth as an extreme environment. The Anatomical Record. 2020;303:235–249. doi: 10.1002/ar.24049. [DOI] [PubMed] [Google Scholar]
  104. Spiekman SNF, Werneburg I. Patterns in the bony skull development of marsupials: high variation in onset of ossification and conserved regions of bone contact. Scientific Reports. 2017;7:43197. doi: 10.1038/srep43197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Stockdale FE. Myogenic cell lineages. Developmental Biology. 1992;154:284–298. doi: 10.1016/0012-1606(92)90068-r. [DOI] [PubMed] [Google Scholar]
  106. Suarez HG, Langer BE, Ladde P, Hiller M. chainCleaner improves genome alignment specificity and sensitivity. Bioinformatics. 2017;33:1596–1603. doi: 10.1093/bioinformatics/btx024. [DOI] [PubMed] [Google Scholar]
  107. Suárez R, Paolino A, Kozulin P, Fenlon LR, Morcom LR, Englebright R, O’Hara PJ, Murray PJ, Richards LJ. Development of body, head and brain features in the Australian fat-tailed dunnart (Sminthopsis crassicaudata; Marsupialia: Dasyuridae); A postnatal model of forebrain formation. PLOS ONE. 2017;12:e0184450. doi: 10.1371/journal.pone.0184450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Sung SS, Brassington AME, Krakowiak PA, Carey JC, Jorde LB, Bamshad M. Mutations in TNNT3 cause multiple congenital contractures: a second locus for distal arthrogryposis type 2B. American Journal of Human Genetics. 2003;73:212–214. doi: 10.1086/376418. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Vaziri Sani F, Kaartinen V, El Shahawy M, Linde A, Gritli-Linde A. Developmental changes in cellular and extracellular structural macromolecules in the secondary palate and in the nasal cavity of the mouse. European Journal of Oral Sciences. 2010;118:221–236. doi: 10.1111/j.1600-0722.2010.00732.x. [DOI] [PubMed] [Google Scholar]
  110. Veitch CE, Nelson J, Gemmell RT. Birth in the brushtail possum, Trichosurus vulpecula (Marsupialia: Phalangeridae) Australian Journal of Zoology. 2000;48:691. doi: 10.1071/ZO00033. [DOI] [Google Scholar]
  111. Villar D, Berthelot C, Aldridge S, Rayner TF, Lukk M, Pignatelli M, Park TJ, Deaville R, Erichsen JT, Jasinska AJ, Turner JMA, Bertelsen MF, Murchison EP, Flicek P, Odom DT. Enhancer evolution across 20 mammalian species. Cell. 2015;160:554–566. doi: 10.1016/j.cell.2015.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Visel A, Minovitsky S, Dubchak I, Pennacchio LA. VISTA Enhancer Browser--a database of tissue-specific human enhancers. Nucleic Acids Research. 2007;35:D88–D92. doi: 10.1093/nar/gkl822. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Waite PM, Marotte LR, Leamey CA. Timecourse of development of the wallaby trigeminal pathway: I. periphery to brainstem. The Journal of Comparative Neurology. 1994;350:75–95. doi: 10.1002/cne.903500106. [DOI] [PubMed] [Google Scholar]
  114. Wakamatsu Y, Nomura T, Osumi N, Suzuki K. Comparative gene expression analyses reveal heterochrony for Sox9 expression in the cranial neural crest during marsupial development. Evolution & Development. 2014;16:197–206. doi: 10.1111/ede.12083. [DOI] [PubMed] [Google Scholar]
  115. Wakamatsu Y, Suzuki K. Sequence alteration in the enhancer contributes to the heterochronic Sox9 expression in marsupial cranial neural crest. Developmental Biology. 2019;456:31–39. doi: 10.1016/j.ydbio.2019.08.010. [DOI] [PubMed] [Google Scholar]
  116. Walker BJ, Abeel T, Shea T, Priest M, Abouelliel A, Sakthikumar S, Cuomo CA, Zeng Q, Wortman J, Young SK, Earl AM. Pilon: an integrated tool for comprehensive microbial variant detection and genome assembly improvement. PLOS ONE. 2014;9:e112963. doi: 10.1371/journal.pone.0112963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Wei B, Jin JP. TNNT1, TNNT2, and TNNT3: Isoform genes, regulation, and structure-function relationships. Gene. 2016;582:1–13. doi: 10.1016/j.gene.2016.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Won HJ, Kim JW, Won HS, Shin JO. Gene regulatory networks and signaling pathways in palatogenesis and cleft palate: a comprehensive review. Cells. 2023;12:1954. doi: 10.3390/cells12151954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Wu M, Gu L. TCseq: time course sequencing data analysis. 3.21Bioconductor. 2020 https://bioconductor.org/packages/release/bioc/html/TCseq.html
  120. Wu T, Hu E, Xu S, Chen M, Guo P, Dai Z, Feng T, Zhou L, Tang W, Zhan L, Fu X, Liu S, Bo X, Yu G. clusterProfiler 4.0: A universal enrichment tool for interpreting omics data. The Innovation. 2021;2:100141. doi: 10.1016/j.xinn.2021.100141. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Xiao Y, Williams JS, Brownell I. Merkel cells and touch domes: more than mechanosensory functions? Experimental Dermatology. 2014;23:692–695. doi: 10.1111/exd.12456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Yip RKH, Chan D, Cheah KSE. Mechanistic insights into skeletal development gained from genetic disorders. Current Topics in Developmental Biology. 2019;133:343–385. doi: 10.1016/bs.ctdb.2019.02.002. [DOI] [PubMed] [Google Scholar]
  123. Yu G, Wang LG, He QY. ChIPseeker: an R/Bioconductor package for ChIP peak annotation, comparison and visualization. Bioinformatics. 2015;31:2382–2383. doi: 10.1093/bioinformatics/btv145. [DOI] [PubMed] [Google Scholar]
  124. Zarate YA, Bosanko K, Derar N, Fish JL. Abnormalities in pharyngeal arch-derived structures in SATB2-associated syndrome. Clinical Genetics. 2024;106:209–213. doi: 10.1111/cge.14540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Zhang Y, Liu T, Meyer CA, Eeckhoute J, Johnson DS, Bernstein BE, Nusbaum C, Myers RM, Brown M, Li W, Liu XS. Model-based analysis of ChIP-Seq (MACS) Genome Biology. 2008;9:R137. doi: 10.1186/gb-2008-9-9-r137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Zhang Y, Park C, Bennett C, Thornton M, Kim D. Rapid and accurate alignment of nucleotide conversion sequencing reads with HISAT-3N. Genome Research. 2021;31:1290–1295. doi: 10.1101/gr.275193.120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Zhu J, Adli M, Zou JY, Verstappen G, Coyne M, Zhang X, Durham T, Miri M, Deshpande V, De Jager PL, Bennett DA, Houmard JA, Muoio DM, Onder TT, Camahort R, Cowan CA, Meissner A, Epstein CB, Shoresh N, Bernstein BE. Genome-wide chromatin state transitions associated with developmental and environmental cues. Cell. 2013;152:642–654. doi: 10.1016/j.cell.2012.12.033. [DOI] [PMC free article] [PubMed] [Google Scholar]

eLife Assessment

Detlef Weigel 1

This important study of regulatory elements and gene expression in the craniofacial region of the fat-tailed dunnart shows that, compared to placental mammals, marsupial craniofacial tissue develops in a precocious manner, with enhancer regulatory elements as primary driver of this difference. The compelling data, including a new dunnart genome assembly, provide an invaluable reference for future mammalian evolution studies, especially once additional developmental time point for the fat-tailed dunnart become available.

Reviewer #1 (Public review):

Anonymous

Summary:

Compared to placental mammals, marsupials have a short gestation period and give birth to altricial young. To assist with the detection and response to cues that direct the neonate joeys to the mother's pouch, as well as latching onto the teat, marsupial craniofacial development at this stage is rapid and heterochronous relative to placentals. Cook et al. have presented an important study on the transcriptomic and epigenomic signature underlying this heterochronous development of craniofacial features across mammals, using the fat-tailed dunnart as a marsupial model.

Given the lack of a dunnart genome, the authors prepared long and short read sequence datasets to assemble and annotate a novel genome to allow for mapping of RNAseq and ChIPseq data against H3K4me3 and H3K27ac, which allowed for identification of putative promoter and enhancer sites in dunnart. They found that genes proximal to these regulatory loci were enriched for functions related to bone, skin, muscle and embryonic development, verifying the precocious state of newborn dunnart facial tissue. When compared with mouse, the authors found a much higher proportion of promoter regions aligned between species than for enhancer regions, and subsequent profiling identified regulatory elements conserved across species and are important for mammalian craniofacial development. In contrast, identification of dunnart-specific enhancers and patterns of RNA expression further confirm the precocious state of muscle development, as well as for sensory system development, in dunnart, suggesting that early formation of these features are critical for neonate marsupials.

Marsupials are emerging as an important model for studying mammalian development and evolution, and the authors have performed a novel and thorough analysis that helps to elucidate the regulatory profile underlying craniofacial heterochrony. Impressively, this study also includes the assembly of a new marsupial reference genome that will benefit many future studies of mammalian developmental biology.

Strengths:

The genome assembly method was thorough, using two different long-read methods (Pacbio and ONT) to generate the long reads for contig and scaffold construction, increasing the quality of the final assembled genome, which was effectively annotated and used for functional analysis of orthologous regulatory elements.

The birth of altricial young in marsupials is an important feature of their development that is distinct from placental mammals which are separated by about 160 million years of evolution. Very little is known, however, about the regulatory profile that contributes to the advanced craniofacial development required for joey survival. This is one of the few epigenomic studies performed in marsupials (of any organ) and the first performed in fat-tailed dunnart (also of any organ), which begins to address this lack of knowledge.

The study also provides evidence supporting the important role enhancer elements play in mammalian phenotypic evolution, relative to promoters.

Weaknesses:

Biological replicates of facial tissue were collected at a single developmental time point of the fat-tailed dunnart within the first postnatal day (P0), and analysed this in the context of similar mouse facial samples from the ENCODE consortium at six developmental time points, where previous work from the authors have shown that the younger mouse samples (E11.5-12.5) approximately corresponds to the dunnart developmental stage (Cook et al. 2021). However, it would be useful to have samples from at least one older dunnart time point, for example, at a developmental stage equivalent to mouse E15.5. This would provide additional insight into the extent of accelerated face development in dunnart relative to mouse, i.e. how long do the regulatory elements that are activated early in dunnart remain active for and does their function later influence other aspects of craniofacial development?

eLife. 2025 Oct 28;14:RP103592. doi: 10.7554/eLife.103592.3.sa2

Author response

Laura Emily Cook 1, Charles Y Feigin 2, John D Hills 3, Davide M Vespasiani 4, Andrew Pask 5, Irene Gallego Romero 6

The following is the authors’ response to the original reviews.

Reviewer #1 (Public review):

Summary:

Cook et al. have presented an important study on the transcriptomic and epigenomic signature underlying craniofacial development in marsupials. Given the lack of a dunnart genome, the authors also prepared long and short-read sequence datasets to assemble and annotate a novel genome to allow for the mapping of RNAseq and ChIPseq data against H3K4me3 and H3K27ac, which allowed for the identification of putative promoter and enhancer sites in dunnart. They found that genes proximal to these regulatory loci were enriched for functions related to bone, skin, muscle and embryonic development, highlighting the precocious state of newborn dunnart facial tissue. When compared with mouse, the authors found a much higher proportion of promoter regions aligned between species than for enhancer regions, and subsequent profiling identified regulatory elements conserved across species and are important for mammalian craniofacial development. In contrast, the identification of dunnart-specific enhancers and patterns of RNA expression further confirm the precocious state of muscle development, as well as for sensory system development, in dunnart suggesting that early formation of these features are critical for neonate marsupials likely to assist with detecting and responding to cues that direct the joeys to the mother's teat after birth. This is one of the few epigenomic studies performed in marsupials (of any organ) and the first performed in fat-tailed dunnart (also of any organ). Marsupials are emerging as an important model for studying mammalian development and evolution and the authors have performed a novel and thorough analysis, impressively including the assembly of a new marsupial reference genome that will benefit many future studies.

Strengths:

The study provides multiple pieces of evidence supporting the important role enhancer elements play in mammalian phenotypic evolution, namely the finding of a lower proportion of peaks present in both dunnart and mouse for enhancers than for promoters, and dunnart showing more genes uniquely associated with it's active enhancers than any other combination of mouse and dunnart samples, whereas this pattern was less pronounced than for promoter-associated genes. In addition, rigorous parameters were used for the cross-species analyses to identify the conserved regulatory elements and the dunnart-specific enhancers. For example, for the results presented in Figure 1, I agree that it is a little surprising that the average promoter-TSS distance is greater than that for enhancers, but that this could be related to the possible presence of unannotated transcripts between genes. The authors addressed this well by examining the distribution of promoter-TSS distances and using proximal promoters (cluster #1) as high confidence promoters for downstream analyses.

The genome assembly method was thorough, using two different long read methods (Pacbio and ONT) to generate the long reads for contig and scaffold construction, increasing the quality of the final assembled genome.

Weaknesses:

Biological replicates of facial tissue were collected at a single developmental time point of the fat-tailed dunnart within the first postnatal day (P0), and analysed this in the context of similar mouse facial samples from the ENCODE consortium at six developmental time points, where previous work from the authors have shown that the younger mouse samples (E11.5-12.5) approximately corresponds to the dunnart developmental stage (Cook et al. 2021). However, it would be useful to have samples from at least one older dunnart time point, for example, at a developmental stage equivalent to mouse E15.5. This would provide additional insight into the extent of accelerated face development in dunnart relative to mouse, i.e. how long do the regulatory elements that activated early in dunnart remain active for and does their function later influence other aspects of craniofacial development?

We thank the reviewer for their feedback and agree that the inclusion of multiple postnatal stages in the dunnart would give further valuable insights to the comparative analyses. Unfortunately, we were limited by the pouch young available and prioritized ensuring robust data at a single stage for this study. We hope to expand this work to more stages in future studies.

The authors refer to the development of the CNS being delayed in marsupials relative to placental mammals, however, evidence shows how development of the dunnart brain (whole brain or cortex) is protracted compared to mouse, by a factor of at least 2 times, rather than delayed per se (Workman et al. 2013; Paolino et al. 2023). In addition, there is evidence that cortical formation and cell birth may begin at approximately the same stage across species equivalent to the neonate period in dunnart (E10.5 in mouse), and that shortly after this at the stage equivalent to mouse E12.5, the dunnart cortex shows signs of advanced neurogenesis followed by a protracted phase of neuronal maturation (Paolino et al. 2023). Therefore, it is possible that marsupial CNS development appears delayed relative to mouse but instead begins at the same stage and then proceeds to develop on a different timing scale.

The comparison here is not directly between CNS development in placental and marsupials but CNS development relative to development of a subset of structures of the cranial skeleton and musculature (as first proposed by Kathleen Smith 1997). For example, Smith 1997 found that in eutherians, evagination of the telencephalon and appearance of the pigment in the eye occur before the ossification of the premaxilla, maxilla, and dentary. However, in marsupials, evagination of the telencephalon and appearance of the pigment in the eye occur concurrently with condensation of cartilage in the basicranium and the ossification of the premaxilla, maxilla, and dentary. Smith 1997 reports both a delay in the initiation of CNS development in marsupials relative to craniofacial ossification and a protraction of CNS development compared to placental mammals.

This also highlights the challenges of correlating different staging systems between placentals and marsupials as stages determined as equivalent can change depending on which developmental events are used. The protracted development of the CNS in marsupials (Smith 1997, Workman et al. 2013; Paolino et al. 2023) still supports the hypothesis that during the short gestation period in marsupials structures required for life outside the womb in an embryonic-like state, such as the orofacial region, are likely prioritized.

We have clarified this based on the reviewers feedback and added text referring to the protraction of marsupial CNS development to the Discussion section.

[New text]: Marsupials display advanced development of the orofacial region relative to development of the central nervous system when compared to placental mammals[3,6].

[New text]: Although development of the central nervous system is protracted in marsupials compared to placentals, marsupials have well-developed peripheral motor nerves and sensory nerves (eg. the trigeminal) at birth [5].

Reviewer #2 (Public review):

This study by Cook and colleagues utilizes genomic techniques to examine gene regulation in the craniofacial region of the fat-tailed dunnart at perinatal stages. Their goal is to understand how accelerated craniofacial development is achieved in marsupials compared to placental mammals.

The authors employ state-of-the-art genomic techniques, including ChIP-seq, transcriptomics, and high-quality genome assembly, to explore how accelerated craniofacial development is achieved in marsupials compared to placental mammals. This work addresses an important biological question and contributes a valuable dataset to the field of comparative developmental biology. The study represents a commendable effort to expand our understanding of marsupial development, a group often underrepresented in genomic studies.

The dunnart's unique biology, characterized by a short gestation and rapid craniofacial development, provides a powerful model for examining developmental timing and gene regulation. The authors successfully identified putative regulatory elements in dunnart facial tissue and linked them to genes involved in key developmental processes such as muscle, skin, bone, and blood formation. Comparative analyses between dunnart and mouse chromatin landscapes suggest intriguing differences in deployment of regulatory elements and gene expression patterns.

Strengths

(1) The authors employ a broad range of cutting-edge genomic tools to tackle a challenging model organism. The data generated - particularly ChIP-seq and RNA-seq from craniofacial tissue - are a valuable resource for the community, which can be employed for comparative studies. The use of multiple histone marks in the ChIP-seq experiments also adds to the utility of the datasets.

(2) Marsupial occupy an important phylogenetic position, but they remain an understudied group. By focusing on the dunnart, this study addresses a significant gap in our understanding of mammalian development and evolution. Obtaining enough biological specimens for these experiments studies was likely a big challenge that the authors were able to overcome.

(3) The comparison of enhancer landscapes and transcriptomes between dunnarts and can serve as the basis of subsequent studies that will examine the mechanisms of developmental timing shifts. The authors also carried out liftover analyses to identify orthologous enhancers and promoters in mice and dunnart.

Weaknesses and Recommendations

(1) The absence of genome browser tracks for ChIP-seq data makes it difficult to assess the quality of the datasets, including peak resolution and signal-to-noise ratios. Including browser tracks would significantly strengthen the paper by provide further support for adequate data quality.

We have put together an IGV session with the dunnart genome, annotation and ChIP-seq tracks. This is now available in the FigShare data repository (10.7554/eLife.103592.1).

(2) The first two figures of the paper heavily rely in gene orthology analysis, motif enrichment, etc, to describe the genomic data generated from the dunnart. The main point of these figures is to demonstrate that the authors are capturing the epigenetic signature of the craniofacial region, but this is not clearly supported in the results. The manuscript should directly state what these analyses aim to accomplish - and provide statistical tests that strengthen confidence on the quality of the datasets.

As this is the first epigenomic profiling for this species we performed extensive data quality control (See Supplementary Tables 2-3, 18, 20-23 and Supplementary Figures 1-3, 6-11). These figures and corresponding Supplementary Tables show the robustness of the data, including well-described metrics for assessing promoters and enhancers, GO terms relevant to craniofacial development and binding motifs for key developmental TF families.

We have emphasised this aspect of the work more strongly in the results section, particularly in [Defining craniofacial putative enhancer- and promoter regions in the dunnart].

(3) The observation that "promoters are located on average 106 kb from the nearest TSS" raises significant concerns about the quality of the ChIP-seq data and/or genome annotation. The results and supplemental information suggest a combination of factors, including unannotated transcripts and enhancer-associated H3K4me3 peaks - but this issue is not fully resolved in the manuscript. The authors should confirm that this is not caused by spurious peaks in the CHIP-seq analysis - and possibly improve genome annotation with the transcriptomic datasets presented in the study.

Spurious ChIP-seq peaks could be possible as there is no “blacklisted regions” database for the dunnart to filter on, however we used a no-IP control, a stringent FDR of 0.01 and peaks had to be reproducible in two biological replicates when calling peaks - all of which should reduce the likelihood of false positives.

H3K4me3 activity at enhancers is well-established, in particular when enhancer sequences are also bound by RNA Pol II (Koch and Andrau, 2011; Pekowska et al., 2011). However, compared to H3K4me3 activity at promoters, H3K4me3 levels at enhancers are low (Calo and Wysocka, 2013). This is in line with our observations that H3K4me3 levels at enhancers are much lower than observed at promoter regions (see Supplementary Note 2). We found that H3K4me3 peaks located closer to the TSS had a stronger peak signal (mean = 46.10) than distal H3K4me3 peaks (mean = 6.95; Wilcoxon FDR-adjusted p < 2.2 x 10-16). This suggests that although some distal promoter peaks may be due to missingness in the annotation, the majority likely represent peaks associated with enhancer regions. We have emphasized this finding more strongly in the results section:

[New text]: H3K4me3 activity at enhancers is well-established[25,26], however, compared to H3K4me3 activity at promoters, H3K4me3 levels at enhancers are low[27]. This is in line with our observations where H3K4me3 levels at distal enhancer peaks are nearly 7 times lower than those observed at promoter regions (see SupNote2).

(4) The comparison of gene regulation between a single dunnart stage (P1) and multiple mouse stages lacks proper benchmarking. Morphological and gene expression comparisons should be integrated to identify equivalent developmental stages. This "alignment" is essential for interpreting observed differences as true heterochrony rather than intrinsic regulatory differences.

Given the developmental differences between eutherian and marsupial mammals it is challenging to assign the dunnart a precise “equivalent” developmental stage to the mouse. From our morphological and developmental characterisation (see Cook et al. 2020 Nat Comms Bio) based on ossification patterns the dunnart orofacial region on the day of birth appears to be similar to that of an E12.5 mouse embryo (just prior to the observation of ossified craniofacial bones). However, when we compared both regulatory elements and expressed genes between the dunnart at this stage (P1) and 5 developmental stages in the mouse, there is no obvious equivalent stage. For example, when we simply compare genes linked to enhancer peaks, the group with the largest intersection between dunnart and any mouse stage are ~500 genes that are present in dunnart, and mouse stages E10.5, E12.5 - E15.5, Figure 5B. When we then compare genes expressed in the dunnart to temporal gene expression dynamics during mouse development we find that the largest overlap is with genes highly expressed at E14.5 or E15.5 in the mouse (Figure 6, Supplementary Figure 5). We have strengthened the rationale for the selected mouse stages in the comparative analyses section of the results.

(5) The low conservation of putative enhancers between mouse and dunnart (0.74-6.77%) is surprising given previous reports of higher tissue-specific enhancer conservation across mammals. The authors should address whether this low conservation reflects genuine biological divergence or methodological artifacts (e.g., peak-calling parameters or genome quality). Comparisons with published studies could contextualize these findings.

The reported range (0.74 - 6.77%) refers to the number regions called as an active enhancer peak in both species (conserved activity) divided by the total number of dunnart peaks alignable to the mouse genome, which we expect to be low given sequence turnover rates and the evolutionary distance separating dunnart and mice. The alignability (conserved sequence) for dunnart enhancers to the mouse genome was ~13% for 100bp regions and can be found in Supplementary Table 22, we have now clarified this in the main text.

[New Text]: After building dunnart-mm10 liftover chains (see Methods and SupNote5) we compared mouse and dunnart regulatory elements. The alignability (conserved sequence) for dunnart enhancers to the mouse genome was ~13% for 100bp regions (Supplementary Table 22).

The activity conservation range reported here is consistent with previously reported for marsupial-placental enhancer comparisons (Villar et al. 2015), where ~1% of conserved liver-specific human enhancers had conserved activity to opossum. Follow up studies in Berthelot et al 2018 also found that approximately 1% of human liver enhancers were conserved across the placental mammals included in the study.

(6) Focusing only on genes associated with shared enhancers excludes potentially relevant genes without clear regulatory conservation. A broader analysis incorporating all orthologous genes may reveal additional insights into craniofacial heterochrony.

We appreciate the reviewers comment, we understand that a broader analysis may provide some additional insights to this question however in this study our focus was understanding the enhancers driving craniofacial development in these species. We linked enhancers with gene expression data as additional evidence of regulatory programs involved in craniofacial development. The majority (~70%) of genes reproducibly expressed were linked to an active enhancer and/or promoter. This has now been highlighted in the result section.

[New Text]: There were 12,153 genes reproducibly expressed at a level > 1 TPM across three biological replicates, with the majority of genes 67% of genes expressed (67%; 8158/12153) associated with near an active enhancer and/or promoter peak.

In conclusion, this study provides an important dataset for understanding marsupial craniofacial development and highlights the potential of genomic approaches in non-traditional model organisms. However, methodological limitations, including incomplete genome annotation and lack of developmental benchmarking weaken the robustness and of the findings. Addressing these issues would significantly enhance the study's utility to the field and its ability to support the study's central conclusion that dunnart-specific enhancers drive accelerated craniofacial development.

Reviewer #1 (Recommendations for the authors):

Minor comments and corrections:

(1) ChIP-seq FRiP fractions were much higher in dunnart samples than in mouse. Is this related to any differences in sample preparation they are aware of in the ENCODE datasets of mouse, such as different anti-histone antibodies used (and therefore different efficiency of binding to the same histone markers across species)? The authors appear to have addressed something similar with respect to the much lower enriched peak number observed in the mouse sample relative to dunnart in Supp note 4. I suspect the "technical cofounder" they refer to there is affecting both the FRiP scores and the higher correlation coefficients between IP and input in mouse.

We chose the same antibodies used in the mouse craniofacial tissue ENCODE experiments however, the procedure is slightly different. We used the MAGnify Chromatin Immunoprecipitation System while in the ENCODE assays performed by Bing Ren’s group in 2012 was an in-house lab protocol for MicroChIP. Given that the samples for mouse and dunnart were not processed together, by the same researcher, with the same protocol there could be any number of technical cofounders impacting enrichment. A low FRiP score suggests low specificity as the majority of reads are in non-specific regions (low enrichment), consistent with the higher correlation between IP and input in mouse. The data quality also appears to vary between H3K27ac and H3K4me3 in the mouse (Supplementary Table 21), with H3K4me3 FRiP scores more similar to those observed in our dunnart experiments. This suggests a potential confounder specific to the mouse H3K27ac IP. QC metrics (FRiP, bam correlation) are consistent between H3K27ac and H3K4me3 IPs in our experiments (Supplementary Table 20).

(2) Some of the promoter peak numbers in Supp table 1 do not match the numbers in the main text.

We have corrected the incorrect number reported in the text for promoter peaks with orthologous genes (8590 -> 8597).

(3) In Supp tables 2 and 3, the number of GO terms similar across tables is 466, which is ~42% of total number of enriched GO terms. However the authors mention that only 23% of terms were the same between promoters and enhancers, and a value of 42% was applied to the proportion of terms uniquely enriched for terms associated with genes assigned to promoters only. Unless I'm reading these Supp tables incorrectly, is it possible the proportions were mixed up?

Thanks for catching this. The lists provided in Supplementary Table 2 were incorrect. The Supplementary Tables and in text description has been corrected to reflect this.

(4) Would be helpful to add a legend for the mouse samples in Supp Figure 10.

We have added the labels to the plot.

(5) In Supp note 5, regarding the percentage of alignable peaks recovered, the percentages mentioned for the 50bp and 500bp peak summit lengths for enhancers and promoters do not seem to match the values in Supp tables 22 and 23.

Thank you for catching this - we have corrected the Supplementary Tables and in text.

(6) Please provide additional information to explain how dunnart RNA expression was associated with the five temporal expression clusters found in the mouse data shown in Figure 6 given there is only one dunnart time point and so the species temporal pattern's could not be compared, i.e. how was the odds ratio calculated and was this applied iteratively for dunnart against each mouse age and within each temporal cluster?

The TCseq package takes the mouse expression data across all 6 stages and calls differentially expressed genes with an absolute log2 fold-change > 2 compared to the starting time-point (E10.5). The mouse gene expression patterns were clustered into 5 clusters that each show distinct temporal expression patterns (see Supplementary Figure 5D). The output from this is 5 lists where within each list are unique genes that share a temporal pattern. These lists of mouse genes were then each compared to the orthologous genes expressed in the dunnart using a Fishers Exact test with corrections for multiple testing using the Holm method. We have added additional details in the methods:

[New text]: Orthologous genes reproducibly expressed >1 TPM in the dunnart were compared to the list of genes for each cluster using Fisher’s Exact Test followed by p-value corrections for multiple testing with the Holm method.

(7) SupFile1 and SupFile2 - which supplementary note or figure are these referring to?

Apologies for this error. These items were meant to link to the FigShare repository where the supplementary files can be found. We have corrected this using the DOI for the repository.

Reviewer #2 (Recommendations for the authors):

(1) Authors should clarify that the mouse ENCODE data used for the comparisons was obtained from craniofacial tissue.

This has now been corrected to clarify that the mouse ENCODE data used was from craniofacial tissues. ENCODE mouse embryonic facial prominence ChIP-seq and gene expression quantification file accession numbers and details used in study can be found in Supplementary Table 17.

(2) Given the large differences in TPM for highly expressed genes shown in Figure 5, a MA or volcano plot would provide a more comprehensive view of global transcriptome differences between species.

We have added this plot as Supplementary Figure 13.

(3) It is unclear whether the enrichment analysis was performed for mouse genes, dunnart genes, or both.

In reference to Figure 5, Gene Ontology enrichment analysis was performed on the top 500 highly expressed genes in dunnart. Because there is not an ontology database for dunnart gene IDs, these top 500 dunnart gene IDs were converted to the orthologous gene ID in mouse before performing the enrichment analysis. We apologise for the lack of clarity and have added additional text in the results section to make this clearer. In addition, the relevant methods section now reads:

[New text]: As there is no equivalent gene ontology database for dunnart, we converted the Tasmanian devil RefSeq IDs to Ensembl v103 using biomaRt v2.46.3 and then converted these to mouse Ensembl v103 IDs. In this way we were able to use the mouse Ensembl Gene Ontology annotations for the dunnart gene domains. All gene ontology analyses were performed using clusterProfiler v4.1.4[117], with Gene Ontology from the org.Mm.eg.db v3.12.0 database[118], setting an FDR-corrected p-value threshold of 0.01 for statistical significance.

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Data Citations

    1. Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. dunnart genome. figshare. [DOI] [PMC free article] [PubMed]
    2. Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. Gene regulatory dynamics during craniofacial development in a carnivorous marsupial. NCBI Gene Expression Omnibus. GSE188990 [DOI] [PMC free article] [PubMed]
    3. Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. dunnart and mouse peak characterisation. figshare. [DOI]
    4. Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. tc-seq processed data. figshare. [DOI] [PMC free article] [PubMed]
    5. Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. liftover chains. figshare. [DOI] [PMC free article] [PubMed]

    Supplementary Materials

    Supplementary file 1. All supplementary tables.

    (a) Annotated dunnart peak nearest gene ID conversions. (b) Enriched gene ontology terms for dunnart genes near promoters. (c) Enriched gene ontology terms for dunnart genes near enhancers. (d) Dunnart enhancers and nearest gene expression (TPM). (e) Dunnart promoters and nearest gene expression (TPM). (f) Enriched GO terms for dunnart enhancers and promoters where nearest gene expression ≥10 TPM (FDR <1%). (g) Summary of mouse–dunnart conserved peak activity for a. high-confidence putative promoters and b. putative enhancers. (h) Enhancers with conserved activity in mouse and dunnart with incorporated stage and species gene expression values (TPM). (i) Gene ontology terms for biological processes enriched in genes expressed in mouse and dunnart (>1 TPM) near conserved enhancers. (j) Dunnart-specific enhancers where the nearest genes were highly expressed (>10 TPM) only in the dunnart with low to no activity in the mouse at any of the embryonic stages (E10.5–E15.5). (k) Highly expressed genes (TPM) in dunnart where expression is lower or absent in mouse. (l) Biological processes enrichment for highly expressed genes in dunnart where expression is lower or absent in mouse. (m) Dunnart facial prominence tissue samples. (n) qPCR primer sequences for ChIP enrichment validation. (o) Details of libraries used in genome assembly. (p) DNA fragments isolated after pull down were sequenced by GENEWIZ. Paired-end sequencing at an average depth of 57 million reads. a. Raw reads sequencing statistics, b. Read statistics after CutAdapt filtering. (q) ENCODE mouse embryonic facial prominence ChIP-seq and gene expression quantification file accession numbers and details used in study. (r) Read alignment quality metrics for H3K4me3 and H3K27ac dunnart ChIP-seq samples. (s) Read alignment quality metrics for mouse ENCODE ChIP-seq data. (t) Peak calling quality metrics for H3K4me3 and H3K27ac dunnart ChIP-seq samples. (u) Peak calling quality metrics for mouse ChIP-seq peaks after peak calling with MACS2. (v) Summary of alignable enhancer-associated peaks in the mouse and dunnart. (w) Summary of alignable high-confidence promoter-associated peaks in mouse and dunnart.

    elife-103592-supp1.xlsx (3.3MB, xlsx)
    MDAR checklist

    Data Availability Statement

    The data generated in this study have been deposited in NCBI’s Gene Expression Omnibus (Barrett et al., 2013) and are available through GEO Series accession number GSE188990. Accession IDs for previously published datasets from the ENCODE consortium (He et al., 2020) are given in Supplementary file 1. Processed data and an IGV session for browsing are available in a figshare repository. All analyses described were carried out using custom bash, Python3, and R v4.1.0 scripts and are available at https://github.com/lecook/chipseq-cross-species (copy archived at Cook, 2025).

    The following datasets were generated:

    Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. dunnart genome. figshare.

    Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. Gene regulatory dynamics during craniofacial development in a carnivorous marsupial. NCBI Gene Expression Omnibus. GSE188990

    Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. dunnart and mouse peak characterisation. figshare.

    Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. tc-seq processed data. figshare.

    Cook LE, Feigin CY, Hills JD, Vespasiani DM, Pask AJ, Gallego Romero I. 2023. liftover chains. figshare.


    Articles from eLife are provided here courtesy of eLife Sciences Publications, Ltd

    RESOURCES