Academia.eduAcademia.edu

TFAP2 paralogs regulate midfacial development in part through a conservedALXgenetic pathway

ABSTRACTCranial neural crest development is governed by positional gene regulatory networks (GRNs). Fine-tuning of the GRN components underly facial shape variation, yet how those in the midface are connected and activated remain poorly understood. Here, we show that concerted inactivation ofTfap2aandTfap2bin the murine neural crest even during the late migratory phase results in a midfacial cleft and skeletal abnormalities. Bulk and single-cell RNA-seq profiling reveal that loss of bothTfap2members dysregulated numerous midface GRN components involved in midface fusion, patterning, and differentiation. Notably,Alx1/3/4(Alx) transcript levels are reduced, while ChIP-seq analyses suggest TFAP2 directly and positively regulatesAlxgene expression.TFAP2andALXco-expression in midfacial neural crest cells of both mouse and zebrafish further implies conservation of this regulatory axis across vertebrates. Consistent with this notion,tfap2amutant zebrafish present abnormalalx3expression pat...

bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. TFAP2 paralogs regulate midfacial development in part through a conserved ALX genetic pathway. 2 Timothy T Nguyen1,2,3,4†, Jennyfer M Mitchell5†, Michaela D Kiel1,2,3, Kenneth L Jones6‡, Trevor J Williams5,6,7, James T Nichols5, and Eric Van Otterloo1,2,3,4,8* 4 1 6 City, IA, 52242, USA 2 8 12 Department of Periodontics, College of Dentistry & Dental Clinics, University of Iowa, Iowa City, IA, 52242, USA 3 10 Iowa Institute for Oral Health Research, College of Dentistry & Dental Clinics, University of Iowa, Iowa Department of Anatomy and Cell Biology, Carver College of Medicine, University of Iowa, Iowa City, IA, 52242, USA 4 Interdisciplinary Graduate Program in Genetics, University of Iowa, Iowa City, IA, 52242, USA 5 Department of Craniofacial Biology, University of Colorado Anschutz Medical Campus, Aurora, CO 80045, USA 14 6 Department of Pediatrics, University of Colorado Anschutz Medical Campus, Children's Hospital Colorado, Aurora, CO 80045, USA 16 7 Department of Cell and Developmental Biology, University of Colorado Anschutz Medical Campus, Aurora, CO, 80045, USA 18 20 8 Craniofacial Anomalies Research Center, University of Iowa, Iowa City, IA, 52242, USA † Equal contributions ‡ Current address: Department of Cell Biology, University of Oklahoma Health Sciences Center, Oklahoma City, OK 73104, USA 22 * Corresponding author: eric-vanotterloo@uiowa.edu Keywords: TFAP2, BOFS, Char syndrome, ALX, Frontonasal Dysplasia, neural crest 24 Summary Statement: Mouse and zebrafish genetic analyses and next-generation sequencing profiling identify a critical, conserved TFAP2-ALX transcriptional axis within the midface developmental gene 26 regulatory network. Competing Interest Statement: The authors declare no conflict of interest. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 28 ABSTRACT: Cranial neural crest development is governed by positional gene regulatory networks (GRNs). Fine- 30 tuning of the GRN components underly facial shape variation, yet how those in the midface are connected and activated remain poorly understood. Here, we show that concerted inactivation of Tfap2a and Tfap2b 32 in the murine neural crest even during the late migratory phase results in a midfacial cleft and skeletal abnormalities. Bulk and single-cell RNA-seq profiling reveal that loss of both Tfap2 members dysregulated 34 numerous midface GRN components involved in midface fusion, patterning, and differentiation. Notably, Alx1/3/4 (Alx) transcript levels are reduced, while ChIP-seq analyses suggest TFAP2 directly and positively 36 regulates Alx gene expression. TFAP2 and ALX co-expression in midfacial neural crest cells of both mouse and zebrafish further implies conservation of this regulatory axis across vertebrates. Consistent with this 38 notion, tfap2a mutant zebrafish present abnormal alx3 expression patterns, and the two genes display a genetic interaction in this species. Together, these data demonstrate a critical role for TFAP2 in regulating 40 vertebrate midfacial development in part through ALX transcription factor gene expression. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. INTRODUCTION: 42 The evolution of paired jaws in parallel with the development of neural crest cells (NCCs) has led to considerable breadth of facial shapes, enabling gnathostome vertebrates to thrive and exploit varied 44 ecological niches (Brandon et al., 2022; Martik and Bronner, 2021; Square et al., 2017; York and McCauley, 2020). The ability for cranial NCCs (CNCCs) to give rise to skeletal elements important for 46 mastication and housing numerous sensory organs is driven by coordination of epigenetic priming and transcriptional responses to local signaling inputs during embryonic development (Clouthier et al., 2010; 48 Kessler et al., 2023; Santagati and Rijli, 2003; Selleri and Rijli, 2023). Fine-tuning of these positional gene regulatory networks (GRNs) underscores facial shape variation; likewise, disruption of such important 50 processes can lead to severe craniofacial birth defects including orofacial clefting. In humans, this phenotype usually presents as a lateral or bilateral cleft of the upper lip and primary palate which presents 52 in ~1 in 700 births (Mai et al., 2019; Tolarova and Cervenka, 1998). However, clefting can also occur at the midline of the medial- and upper-face region (i.e., midface) (Tessier, 1976), generating a bifid nose 54 that is often part of the spectrum of pathologies such as Frontonasal Dysplasia (FND) (MIMs: 136760, 613451, 613456) (Vargel et al., 2021). Genes encoding the homeobox-containing ALX transcription factors 56 (ALX1/3/4) are the prominent underpinning of FND when mutated and have also been linked with midfacial development in mouse and zebrafish (Beverdam et al., 2001; Iyyanar et al., 2022; Lakhwani et al., 2010; 58 McGonnell et al., 2011; Mitchell et al., 2021; Qu et al., 1999; Yoon et al., 2022). Although ALX genes are clearly fundamental components of the GRNs regulating midfacial development, how they connect with 60 other midface genes into regulatory nodes remains uncertain. Of further importance, midface-specific modules are postulated to drive facial shape variation in humans (Claes et al., 2018; Feng et al., 2021; 62 Naqvi et al., 2022; Xiong et al., 2019). Therefore, uncovering and connecting these genes in the midface GRN remains central to understanding facial development, morphology, and pathology. 64 Encoded members of the TFAP2 gene family have previously been linked with facial development, particularly as transcription factors operating in GRNs across multiple NCC lineages in several vertebrate 66 species (de Croze et al., 2011; Dooley et al., 2019; Hovland et al., 2022; Kenny et al., 2022; Li and Cornell, 2007; Rothstein and Simoes-Costa, 2020; Schmidt et al., 2011; Seberg et al., 2017; Van Otterloo et al., bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 68 2010; Van Otterloo et al., 2012). TFAP2 paralogs share highly conserved amino acid sequences forming their DNA-binding and dimerization domains; moreover, their overlapping DNA-binding motifs (5’- 70 GCCNNNGGC-3’) and capacity to form homo- or heterodimers suggest functional redundancy (Eckert et al., 2005; Williams et al., 1988; Williams and Tjian, 1991). In particular, TFAP2A and TFAP2B have been 72 shown to regulate craniofacial development in numerous model organisms (Barrallo-Gimeno et al., 2004; Knight et al., 2003; Martino et al., 2016; Schorle et al., 1996; Zhang et al., 1996) and are also mutated in 74 human syndromes with facial pathologies. Dominant mutations in TFAP2A and TFAP2B cause BranchioOculo-Facial syndrome (BOFS) (MIM: 113620) (Milunsky et al., 2008) and Char syndrome (MIM: 169100) 76 (Satoda et al., 2000), respectively, conditions which include midface phenotypes such as hypertelorism, broad nasal tip, as well as an abnormal philtrum and nasal bridge. TFAP2 has also been implicated as a 78 central determinant to facial evolution (Prescott et al., 2015) and, thus, facial morphology appears particularly sensitive to the total functional levels of TFAP2 (Laugsch et al., 2019; Naqvi et al., 2022). 80 Supporting this notion, biochemical analyses suggest the BOFS and Char syndrome mutations interfere with the function of the remaining wild-type product or other TFAP2 paralogs (Li et al., 2013; Satoda et al., 82 2000). Moreover, simultaneous targeting of TFAP2A and TFAP2B in mice and zebrafish results in exacerbated craniofacial defects (Knight et al., 2005; Van Otterloo et al., 2018; Van Otterloo et al., 2022; 84 Woodruff et al., 2021). Mechanistically, TFAP2 orchestrates branchial arch (BA) patterning through modulating homeobox transcription factor gene expression (Knight et al., 2004; Knight et al., 2003; Van 86 Otterloo et al., 2018), thus raising the possibility that similar mechanisms operate in the midface. Here, we coupled mouse and zebrafish genetics with next-generation sequencing to extend our 88 analysis of TFAP2A and TFAP2B function in midfacial development and gene regulation. Our findings show that similar midfacial clefting phenotypes are observed when these two genes are lost either early or 90 late in mouse NCC development, suggesting that the critical aspects of their function reside in patterning and differentiation rather than migration. Using genome-wide molecular analysis in mouse coupled with 92 genetic approaches in zebrafish, we further link these TFAP2 genes with ALX family expression and function, placing them firmly in a GRN critical for midface development and evolution. 94 bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. RESULTS: 96 TFAP2A and TFAP2B cooperatively function in CNCCs during midface development. In our previous efforts to illuminate the combined role of TFAP2A and TFAP2B in CNCCs, we 98 identified that simultaneous loss of these genes in mice caused major defects in the development of both midface and BA1-derived jaw structures (Fig. S1A-F) (Van Otterloo et al., 2018). Subsequent profiling 100 focused on the disruption of GRNs patterning BA1, while mechanisms responsible for the midfacial cleft were not pursued in detail. These earlier studies also deployed a combination of null and floxed conditional 102 alleles such that lack of Tfap2 in NCCs was accompanied by their heterozygosity in the remaining embryo (Fig. 1A). Since the concerted action of TFAP2A and TFAP2B serves critical functions in the ectoderm 104 during facial development (Van Otterloo et al., 2022; Woodruff et al., 2021), we wished to ascertain if the midfacial clefting was caused solely by loss of the two genes in NCCs. 106 To address this, we utilized only floxed conditional alleles in combination with NCC-specific CRE recombinase transgenes (Fig. 1B). Moving forward, embryos derived from the prior breeding scheme (Fig. 108 1A) are denoted as the null/conditional model (Fig. 1C, red), while embryos from the updated paradigm (Fig. 1B) are termed as the conditional/conditional model (Fig. 1C, blue). First using Wnt1:CRE (Danielian 110 et al., 1998) in the updated paradigm, we conditionally deleted various allelic combinations of Tfap2 genes in pre-migratory NCCs and examined gross craniofacial development (Fig. 1C-F). CRE-negative control 112 mice at embryonic day 18.5 (E18.5) displayed appropriate midface formation that included laterally positioned vibrissae and an elongated snout (Fig. 1D). Littermates with NCC-specific loss of Tfap2a 114 (Tfap2aNCKO, also Tfap2b-heterozygous) or Tfap2b (Tfap2bNCKO, also Tfap2a-heterozygous) displayed mild and largely overlapping midfacial anomalies such as an indented snout (Fig. 1E, F, gold arrowhead) with 116 misplaced vibrissae on the dorsum (Fig. 1E’, F’, circled). By contrast, complete loss of Tfap2a and Tfap2b in NCCs (Tfap2NCKO) resulted in a fully penetrant midfacial cleft that was accompanied by a mildly 118 shortened snout (Fig. 1G; white arrowhead). The misplaced vibrissae present by loss of a single paralog were not observed in Tfap2NCKO embryos, but we surmised this was confounded by the facial cleft. Earlier bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 120 timepoints (E11.5, E12.5, E15.5) revealed similar phenotypic trends, with the cleft becoming evident at E11.5 (Fig. S2) during midface fusion. 122 Jaw defects were also apparent in the various mutants by gross examination or following skeletal analysis (Fig. S3), resembling those observed in our prior studies (Van Otterloo et al., 2018). Intriguingly, 124 the conditional/conditional embryos did not display a midline mandibular cleft, in contrast to previous findings in their Tfap2Δ/NCKO counterparts (Fig. S1). Therefore, the current analysis indicates that the upper 126 midface cleft and significant BA1 patterning defects can be attributed solely to loss of these two genes in NCCs, whereas mandibular clefting is uncovered by the reduced allelic dosage of Tfap2 elsewhere— 128 presumably the ectoderm. In summary, anomalies like the midfacial cleft suggest that proper facial development relies on 130 sufficient Tfap2 dosage specifically in NCCs. As observed in other NCC stages and derivatives (Rothstein and Simoes-Costa, 2020; Schmidt et al., 2011; Seberg et al., 2017; Van Otterloo et al., 2018), these 132 findings raise the hypothesis that TFAP2A and TFAP2B cooperatively function in NCC GRNs directing midfacial development. 134 Fusion of the frontonasal prominences require TFAP2 function in post-migratory CNCCs. 136 TFAP2 paralogs are postulated to be critical pioneer transcription factors residing throughout NCC development, marking genes poised to be activated once CNCCs have taken residency in the facial 138 prominences (Fernandez Garcia et al., 2019; Rada-Iglesias et al., 2012; Rothstein and Simoes-Costa, 2020). We next asked whether the midfacial cleft resulted from the lack of TFAP2 activity throughout NCC 140 development or after their migration. Previously assessed by β-galactosidase staining with the r26r-lacZ reporter (Soriano, 1999), E9.0 and E10.0 Tfap2Δ/NCKO embryos revealed no distinct changes in CNCC 142 migration into the facial complex (Van Otterloo et al., 2018), which comprise of the midface-associated frontonasal prominence (FNP) alongside the paired maxillary prominences (MxP) a mandibular 144 prominences (MdP). Extending this analysis in conditional/conditional animals, we observed that the distribution and intensity of β-galactosidase staining in the E10.5 midface was similar between Tfap2NCKO, bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 146 Tfap2aNCKO, Tfap2bNCKO, and control littermates heterozygous for Tfap2 in NCCs (Tfap2HET) (Fig. S4A), with the caveat that the developing cleft between the FNP alters overall morphology in Tfap2NCKO embryos 148 (Fig. S4A, arrowheads). Performing sectional immunofluorescence on E11.5 wild-type midface tissue, we found TFAP2A 150 and TFAP2B protein expression in CNCCs occupying the medial and lateral domains of the FNP (Fig. 2A; Fig. S4B). Additional comparisons of published E11.5 transcriptomes showed persistent Tfap2a and 152 Tfap2b transcript levels in FNP CNCCs in contrast to those in the MdP (Fig. 2B) (Hooper et al., 2020), Therefore, taking advantage of the Sox10:CRE transgene (Matsuoka et al., 2005) to inactivate Tfap2 when 154 CNCCs have begun reaching the FNP (~E9.0) (Hari et al., 2012; Jacques-Fricke et al., 2012), we next sought to determine the critical periods by which TFAP2 is needed for FNP fusion. Micro-computed 156 tomography (µCT) analysis at E12.5 showed that, indeed, similar to Wnt1:CRE animals (Fig. 2C), Sox10:CRE-Tfap2NCKO embryos presented a midface cleft not observed in controls (Fig. 2D). Quantification 158 of the linear distance between medial FNP tips were statistically significant between controls and their respective mutant backgrounds (Fig. 2E), with Sox10:CRE-Tfap2NCKO more mildly affected at this stage, 160 while ventral µCT sections revealed that both mutants shared abnormal nasal pit morphology (Fig. 2F, G). Lastly, examining the gross morphology of E18.5 Sox10:CRE embryos of the various Tfap2 allelic 162 combinations presented similar phenotypic trends as their Wnt1:CRE counterparts (Fig. S5). Thus, while TFAP2 persists throughout CNCC developmental progression, these data provide 164 evidence that the midface cleft does not arise from impaired TFAP2 activity during migration. Rather, retention of TFAP2A and TFAP2B expression levels in post-migratory CNCCs, alongside the Sox10:CRE 166 studies, refine the temporal requirement for both paralogs in midface fusion. 168 Appropriate midfacial bone and cartilage formation depends on Tfap2a and Tfap2b gene dosage. We next sought to leverage our transgenic models to better understand the developmental timing 170 of TFAP2 during midfacial bone and cartilage formation. Accordingly, we compared E18.5 skeletal preparations of CRE-negative control, Tfap2aNCKO, Tfap2bNCKO, and Tfap2NCKO backgrounds from both bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 172 conditional/conditional paradigms. Detailing the midfacial skeleton first in Sox10:CRE animals, controls showed appropriate ossification of frontal and nasal bones meeting at the midline (Fig. 3A). Underneath 174 the cranium lies the primary palate formed by the premaxilla bones, while the anterior portions of the vomer bones meet at the midline (Fig. 3A’). The nasal capsules are joined, connected to a compact cartilaginous 176 template for the nasal (i.e., ethmoid) labyrinth and nasal septum that sits adjacent to the anterior presphenoid (Fig. 3A’). Tfap2aNCKO and Tfap2bNCKO backgrounds exhibited small ectopic cartilage islands 178 forming on the calvaria, mildly increased gaps between the frontal bones (Fig. 3B, C). Comparing Tfap2aNCKO and Tfap2bNCKO mutants, the former exhibited gaps in the nasal bones and mild split in the 180 vomer bones (Fig. 3B, B’, white arrowhead, C, C’). By contrast, Tfap2NCKO embryos lacked nasal bones, exposing the underlying split nasal capsules (Fig. 3D, asterisk). Large cartilaginous ectopias were also 182 found in between the hypoplastic frontal bones that, in some individuals, appeared to stem from the nasal septum (Fig. 3D, gold arrowheads; Fig. S6A, A’). Interestingly, in one individual these ectopias were 184 replaced by severe frontal bone fractures (Fig. S6A”). Underneath the calvaria, Tfap2NCKO mutants presented clefting of the palatal processes of the premaxilla, increased separation between the split vomer 186 bones, less compacted nasal labyrinths, a thickened nasal septum, and malformed presphenoid bone (Fig. 3D’; Fig. S6B-B”, dashed lines). Although jaw and cranial base defects were not closely examined, it is 188 worth mentioning that major BA1 defects were in line with those observed in both Wnt1:CRE animals (Fig. S7) (Van Otterloo et al., 2018). Intriguingly, only one Tfap2NCKO mutant presented fusion of the paired jaws 190 (i.e., syngnathia) (Fig. S7E), a defect with higher prevalence in Wnt1:CRE-Tfap2NCKO and Tfap2Δ/NCKO embryos (Fig. S3) (Van Otterloo et al., 2018). Overall, the more common phenotype was fusion between 192 maxillary and jugal bones (Fig. S7F). Next examining the midface elements in Wnt1:CRE animals, both Tfap2aNCKO and Tfap2bNCKO 194 single mutants presented largely overlapping phenotypes as their Sox10:CRE equivalents, though with slight differences (Fig. 3E-G, E’-G’). For example, compared to controls (Fig. E, E’), both single conditional 196 mutants share gapped nasal bones and mishappened cartilage scaffolds connecting the frontal bone and anterior cranial base (pila pre/postoptica) (Fig. 3F, G, F’, G’). Strikingly, Wnt1:CRE-Tfap2NCKO embryos 198 presented nearly identical phenotypes like those observed with their Sox10:CRE counterparts, including a bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. complete lack of nasal bones, smaller frontal bones, cranial ectopias, and overt separation between the 200 premaxillary palate and vomer bones (Fig. 3H, H’). Moreover, additional preparation of E15.5 and E18.5 Tfap2Δ/NCKO skeletons displayed similar midfacial hypoplasia and ectopia phenotypes (Fig. S8). It is unclear 202 whether subtle phenotypic distinctions between Sox10:CRE and Wnt1:CRE mice are confounded by differences in CRE spatial expression patterns (Debbache et al., 2018). Nevertheless, overlaps in major 204 defects between animal models further emphasize critical dependence on Tfap2a and Tfap2b gene dosage during post-migratory CNCC events such as skeletal differentiation. 206 Combined loss of Tfap2a and Tfap2b dysregulates a post-migratory midfacial neural crest GRN 208 that includes Frontonasal Dysplasia-related Alx1/3/4 genes. Our mouse genetic analyses motivated the hypothesis that TFAP2A and TFAP2B acted upstream 210 at key post-migratory GRN modules associated with midface mesenchyme patterning and differentiation. Herein, we undertook transcriptomic profiling of controls and mutants during critical timepoints for midface 212 fusion: RNA-seq of E10.5 bulk FNP/MxP tissue and scRNA-seq of E11.5 CNCCs, respectively. With respect to the first approach, comparison between CRE-negative control and Tfap2Δ/NCKO mutants 214 uncovered 149 genes to be dysregulated between groups (86 down-regulated, 63 up-regulated) (Fig. 4A; Table S1). Ontology, enrichment, and pathway analysis via the Enrichr pipeline (Kuleshov et al., 2016) 216 revealed an overrepresentation of midfacial pathology categories associated with down-regulated genes (Fig. 4B). Similar to Tfap2Δ/NCKO profiles in the lower jaw (Van Otterloo et al., 2018), many of the terms 218 contained homeobox transcription factor genes including calvaria development-related Msx1 (Han et al., 2007; Ishii et al., 2005; Roybal et al., 2010) as well as Pax3 and Pax7 (Gu et al., 2022; Zalc et al., 2015). 220 Notably, terms like midline defect of the nose (p = 2.00e-05, adj. p < 0.05) as well as Frontonasal Dysplasia (6.33e-06, adj. p < 0.05) contained Alx1, Alx3, Alx4 (Fig. 4A, B), whose combinatorial loss lead to a midface 222 cleft strikingly similar to those in Tfap2NCKO and Tfap2Δ/NCKO (Fig. S2; Table S1) (Beverdam et al., 2001; Gu et al., 2022; Iyyanar et al., 2022; Lakhwani et al., 2010; Qu et al., 1999). Also reduced were Crabp1 and 224 Aldh1b1, encoding regulators of a retinoic acid pathway known to regulate midface fusion (Fig. 4A; Table bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. S1) (Gao et al., 2021; Kennedy and Dickinson, 2012; Lohnes et al., 1994; Williams and Bohnsack, 2019; 226 Wu et al., 2022). Unexpectedly, we noted increased expression of genes encoding collagens (e.g., Col14a1, Col3a1), periostin (Postn), osteoglycin (Ogn), and Fibronectin Leucine Rich Transmembrane 228 Protein 2 (Flrt2) (Fig. 4A; Table S1), components involved in skeletal formation (Bhatt et al., 2013; Dash and Trainor, 2020) and suggest premature activation of this process. Thus, by E10.5, transcriptional 230 changes are already emerging such as those involved in mammalian midfacial anomalies. Secondarily conducting scRNA-seq of CNCCs sorted from the heads of E11.5 Tfap2HET and 232 Tfap2NCKO littermates (Fig. 4C), we next examined the changing GRN architecture amidst midface fusion and the extent to which transcriptomic identities were compromised. Integration and Uniform Manifold 234 Approximation and Projection (UMAP) plotting of combined control and mutant genotypes by Seurat (Hao et al., 2021) generated 16 clusters that divided into three major cell populations (Fig. 4C). Concordant with 236 Wnt1:CRE cell labelling and other scRNA-seq studies (Jacques-Fricke et al., 2012; Soldatov et al., 2019), these cells clustered into groups for glia (e.g., Sox10, Erbb3, Plp1, Isl1; cluster 9), neurons and 238 components of the brain (e.g., Neurog2, Stmn2, Sox2, Otx2; clusters 4, 5, 7, 8, 10-13, 15), and the mesenchyme (e.g., Prrx1/2, Twist1, Fli1; clusters 0-3, 6) (Fig. 4C, Fig. S9). While total cell numbers 240 between samples were not drastically altered, mutant quantities for glial and mesenchymal populations were reduced (~7-8% each) while neuronal populations increased (~12%) compared to controls (Fig. 242 S10A). Gene expression-based cell cycle scoring did not detect noticeable changes in cell cycle progression of these populations (Fig. S10B), indicating cellular profiles between genotypes are not due 244 to altered proliferation states. While it is possible that TFAP2 participates in biasing CNCCs to a particular cell fate decision, we reasoned their loss alone is not sufficient to drive drastic fate switches akin to 246 previously reported factors (Cox et al., 2012; Fan et al., 2021; Scerbo and Monsoro-Burq, 2020; SimoesCosta and Bronner, 2016; Soldatov et al., 2019). Supporting this notion is no obvious expansion in 248 neurons, previously tested by whole-mount anti-Neurofilament immunostaining (Van Otterloo et al., 2018). To extend our bulk RNA-seq analyses, we implemented two scRNA-seq workflows in parallel for 250 the mesenchyme group (Fig. 4C): a “pseudobulk” approach for differential gene expression analyses and MAGIC imputation (van Dijk et al., 2018) to annotate major positional identities. The former identified 383 bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 252 dysregulated genes (124 down-regulated, 259 up-regulated; log fold-change threshold = 0.1, adj. p < 0.05), that contained overlaps with RNA profiles in the E10.5 midface (Table S1). For example, downregulated 254 genes yielded several Enrichr terms associated with transcription factor gene regulation and craniofacial features associated with midface molecules (e.g., Msx1, Pax3, Pax7, Alx1, Alx4) (Fig. S11A, Table S1). 256 Although Alx3 did not appear in our pseudobulk, we suspect this is partly due to it being the lowest expressing paralog in the whole mesenchyme alongside the limited sequencing depth. Intriguingly, the 258 upregulated GRN modules yielded members of the Fox transcription factor family (e.g., Foxd1, Foxf2, Foxp1, Foxp2), downstream targets of a Hedgehog signaling pathway often associated with midline facial 260 disorders when precociously activated (Fig. S11B, Table S1) (Brugmann et al., 2010; Jeong et al., 2004; Xu et al., 2022; Xu et al., 2018). Further, we continued to observe several extracellular matrix terms 262 resulting from the multiplicity of elevated collagen genes (e.g., Col1a1, Col2a1, Col3a1, Col4a1, Col6a1, Col6a2) (Fig. S11B, Table S1). 264 Confirming the positional specificity of the disrupted GRNs, we mapped expression levels of select genes on the computationally re-clustered mesenchyme UMAP, which comprises the presumptive FNP 266 (FNP-1/2, clusters 1, 2; Tfap2a/b+, Alx1/3/4+, Six2+), MxP (cluster 0; Six1+, Lhx6/8+), MdP (cluster 0, 4; Dlx1/2/3/4/5/6+, Hand2+), and branchial arch 2 (BA2, cluster 3; Hoxa2+) (Fig. 4D; Fig. S12; Fig. S13). 268 Consistent with previously reported lower face hypoplasia phenotypes (Van Otterloo et al., 2018), mutant cells were cumulatively underrepresented across MxP, MdP, and BA2 clusters (Fig. S14A, A’), 270 accompanying the decreased expression of various Dlx, Hand2, and Gsc transcripts (Table S1; Fig. S14B). Lastly, of the two FNP clusters, mutant cells increased in FNP-1 (58.4% mutant/41.6% control) while 272 decreasing in FNP-2 (26.8% mutant/73.2% control) (Fig. S14A, A’). These changes were coupled with reduced levels of Alx1/3/4 (Fig. 4E), Msx1, and Pax3/7, as well as upregulation of collagen and Fox genes 274 (Fig. S14C). Orthogonal approaches like whole mount in situ hybridization (Fig. 4F) and real-time PCR (Fig. 4G) analyses validated that, compared to their respective controls, both Tfap2NCKO and Tfap2Δ/NCKO 276 mutants exhibited reduced midfacial Alx gene expression. In total, profiling the loss of Tfap2 in CNCCs by transcriptomic and target gene expression profiling 278 suggests these transcription factors play crucial roles in regulating genes in the various positional bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. programs. Importantly, these data support the hypothesis that TFAP2 paralogs function upstream of critical 280 GRN nodes involved in midface fusion and skeletal formation, which include the Frontonasal Dysplasiaassociated Alx1/3/4 genes. 282 TFAP2 plays a direct role in the midface GRN in part by occupying Alx1/3/4 regulatory elements. 284 Our genome wide transcriptomic analyses identified important genes that were dysregulated by loss of Tfap2 in the midface CNCCs, but whether these were direct or indirect targets of these two 286 transcription factors remained unclear. Therefore, we conducted ChIP-seq in E11.5 whole facial prominence tissue using an antibody that recognizes TFAP2A/B/C (Fig. 5A, data not shown). We identified 288 13,778 enriched regions (i.e., TFAP2 ‘peaks’) relative to input (Fig. 5B, Table S2), while antibody specificity was reflected by the TFAP2 consensus binding motif being the most significantly enriched (p = 1.00e-154) 290 (Fig. 5C). Further analysis of TFAP2 binding coordinates revealed approximately 35% of them being in proximal promoters, 23% intronic, 16% intergenic, 9% in the 5’ UTR, while remaining peaks were found in 292 additional regions of the genome (Fig. 5C). Since we used whole facial prominence tissue for these studies, these peaks presumably reflect unique or shared targets within the FNP, MxP, MdP, as well as those 294 present in other tissues such as ectoderm and neuronal tissue. To focus on genes relevant to the midface, we cross-referenced these peak data with gene expression data for the E11.5 FNP (Hooper et al., 2020). 296 Of all genes with detectable midface expression (e.g., E11.5 FNP mesenchyme FPKM > 0, N = 15,842), the genes associated with at least one TFAP2 peak (N = 8,754) were expressed in the midface at a 298 significantly higher level than midface genes (N = 7,088) without a TFAP2 associated peak (p = 2.60e256) (Fig. 5D, Table S2). Further, despite the unique features of different datasets (e.g., E10.5 vs E11.5, 300 NCC-only vs whole face, etc.), ~66% (76 of 115) of the genes dysregulated in our overlapping bulk and scRNA-seq dataset, including both upper- and lower-face specific, were associated with a TFAP2 peak 302 (Table S2). We next assessed chromatin states at these ~13,800 genomic regions with the prediction that a 304 subset of these coordinates would be found within FNP expressed, CNCC-associated, accessible, and bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. ‘active’ regions of the genome. Using k-means clustering along with previously published histone ChIP306 seq and chromatin accessibility datasets generated from wild-type E10.5 FNP CNCCs (Minoux et al., 2017), 5 major clusters (clusters 3, 4, 5, 6, 7; cluster 1 and 2 had < 6 peaks) were identified (Fig. 5E, Table 308 S2). Of these five clusters, over half of TFAP2 bound regions (N = 7,995; 58% of all peaks) clustered into 2 major groups (clusters 5, 6) (Fig. 5E, Table S2). Both clusters contained substantial levels of the active 310 H3K4me2 and H3K27ac, accessible chromatin, and marked reduction in the repressive H3K27me3 mark (Fig. 5E). In contrast to clusters 5 and 6, clusters 3 and 4 displayed substantially reduced levels for open 312 chromatin and active H3K27ac accompanied by expanded H3K27me3 signal (i.e., repressed regions in E10.5 FNP CNCCs) (Fig. 5E, Table S2). Consistent with these profiles, genes associated with clusters 5 314 and 6 were highly expressed in the FNP, relative to genes without a TFAP2 peak, or genes associated with a peak(s) found in other clusters (Fig. 5F, Table S2). Pathway analysis for cluster 5 (1,658 peaks) and 316 cluster 6 (6,336 peaks) revealed terms associated with head and midfacial pathology (Fig. 5G). Notably, genes associated with cluster 3 and 4 were significantly enriched for skin (cluster 4), neuronal (clusters 3, 318 4), and cardiac (cluster 3) terms (Fig. S15), potentially reflecting TFAP2 binding in non-NCC tissues in which TFAP2 is expressed (e.g., ectoderm). Finally, cluster 7 displayed relatively little FNP associated 320 histone and chromatin signatures (Fig. 5E, Table S2) and displayed the fewest associated terms (data not shown). 322 Given the significant down-regulation of all three Alx genes in Tfap2NCKO and Tfap2Δ/NCKO mutants (Fig. 4) and the midfacial cleft associated with their loss (Beverdam et al., 2001; Iyyanar et al., 2022; 324 Lakhwani et al., 2010; Qu et al., 1999), we next asked whether the Alx loci were directly bound by TFAP2. Assessment of peak files revealed that all 3 Alx loci had enriched regions of TFAP2 binding (Fig. 5H); 326 moreover, these regions were associated with cluster 5, concordant with active and open genomic regions at the site of TFAP2 binding (Fig. 5H). Consistent with the more limited expression of Alx1, Alx3, and Alx4 328 in jaw domains at E11.5, these genomic regions were associated with reduced H3K4me2 and H3K27ac and increased H3K27me3 in CNCCs occupying the MxP and MdP (Fig. 5H, and not shown) (Minoux et 330 al., 2017). Thus, transcription factor binding analysis suggests that TFAP2 paralogs occupy genomic bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. regions harboring active histone enhancer marks and associated with midfacial patterning and 332 differentiation genes, namely Alx genes. 334 Zebrafish tfap2a regulates alx3 expression and genetically interacts with alx3 during midfacial development. 336 While different vertebrates use different TFAP2 paralogs during CNCC development, a member commonly used is TFAP2A (Eckert et al., 2005; Meulemans and Bronner-Fraser, 2002). Like humans and 338 mice, the zebrafish tfap2a paralog is a critical determinant for formation of the craniofacial structures (Neuhauss et al., 1996; Schilling et al., 1996). Mining our published CNCC scRNA-seq data (Mitchell et 340 al., 2021; Stenzel et al., 2022), we observed tfap2a gene expression predominantly in the frontonasal cluster giving rise to the neurocranium (i.e., midface skeleton) (Fig. S16). Further co-expression analyses 342 indicated that tfap2a-positive cells also express alx3 (Fig. S16), suggesting that TFAP2A regulation of ALX gene expression is a conserved transcriptional axis in vertebrates. To test this, we examined alx3 344 expression in wild-types and tfap2a mutants (Fig. 6A-D”) (Knight et al., 2003; Schilling et al., 1996). We analyzed zebrafish embryos using in situ hybridization at 48 hours post-fertilization, a stage resembling 346 the mouse E10.5 stage when changes in Alx expression were identified in mouse Tfap2aΔ/NCKO FNP tissue (Fig. 4G). In wild-type embryos, alx3 expression is enriched in fli1a:EGFP-labelled CNCCs on the 348 developing roof of the mouth (Fig. 6A-A”, C-C”), as well as those ventral and medial to the nasal placodes (Fig. 6A-A”). In tfap2a mutants, alx3 expression patterns were altered, becoming reduced around the 350 periphery of the nasal placodes, and bifurcated into two laterally expanded domains (Fig. 6B, D, E). Similar to prior zebrafish BA studies (Barrallo-Gimeno et al., 2004) and our mouse study (Fig. S4A), CNCC number 352 and positioning appeared to be largely intact in the frontonasal region in tfap2a mutants (Fig. 6A’-D’, A”D”). These data strongly suggest that the transcriptional circuit identified in post-migratory CNCCs is 354 conserved between mice and zebrafish. Further testing the conservation hypothesis with the tfap2a and alx3 mutant alleles (Mitchell et al., 356 2021), we sought to determine if the regulatory axis could be read out as genetic interactions (Fig. 6F-J). bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Several previous genetic studies indicate that viscerocranium structures derived from BA2-7 appear the 358 most sensitive to tfap2a mutation, while the BA1 skeleton is less affected (Knight et al., 2003; Neuhauss et al., 1996; Schilling et al., 1996; Wang et al., 2011). In the anterior neurocranium, tfap2a mutants 360 produced incompletely penetrant ethmoid plate loss with variable expressivity, ranging from complete absence to small discontinuations (Barrallo-Gimeno et al., 2004). In mutants with a divided ethmoid plate, 362 the parasphenoid bone extended anteriorly, sometimes interfacing with the abnormal cartilage gap (Fig. 6G’) (Schilling et al., 1996). In this study, two incompletely penetrant tfap2a mutant-associated skeletal 364 phenotypes were observed that were not previously described. Specifically, tfap2a mutant larvae exhibited partial penetrance for cartilaginous ectopias proximal to the eyes (Fig. 6G, red arrowhead) as well as 366 ectopic cartilage rods projecting dorsally from the ethmoid plate (Fig. 6J). As alx3 gene dosage was reduced in tfap2a mutants, the penetrance for all three phenotypes increased. We also discovered a new 368 phenotype, never seen in single mutants, when copies of alx3 were removed from tfap2a mutants (Fig. 6I, J). The parasphenoid is expanded in animals mutant for both tfap2a and alx3. It is interesting to note that 370 a selection of these skeletal abnormalities parallel those observed in our mouse models. Specifically, the midfacial clefting and ectopic cartilage (Fig. 6I, I’) bear a striking resemblance to those found in Tfap2NCKO 372 and Tfap2Δ/NCKO mutant embryos, suggesting that midline facial defects in both TFAP2 mutant animal models are ALX3-dependent. 374 Performing the reciprocal genetic interaction experiment, we documented how alx3 mutantassociated phenotypes were affected upon removing wild-type copies of tfap2a. alx3 homozygous mutants 376 present cellular morphology changes in the anterior ethmoid plate, an anteriorly truncated parasphenoid, and ectopic midline cartilage (Fig. 6J, middle table, boxed in purple) (Mitchell et al., 2021). In alx3 mutant 378 homozygotes with reduced tfap2a dosage, the ethmoid plate and parasphenoid were severely impacted (Fig. 6J, middle table), precluding us from scoring tfap2a-dependent changes. Therefore, we examined 380 changes in alx3-associated phenotypes in alx3 heterozygotes when wild-type copies of tfap2a were removed. Concordant with the first phenotypic scoring, we found that the removal of tfap2a copies from 382 alx3 heterozygotes increased penetrance of the alx3 mutant phenotypes (Fig. 6J, bottom table). These findings demonstrate that tfap2a and alx3 genetically interact during zebrafish midfacial development. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 384 Taken together with our observations in mice, these results raise a model that the transcription factors encoded by TFAP2 and ALX genes function in a conserved, midface developmental pathway. 386 DISCUSSION: 388 The role of TFAP2A and TFAP2B in cranial neural crest cells during midface development. Positional gene regulatory networks endow CNCCs their ability to form unique facial structures 390 during embryogenesis. Despite the clear existence of GRN components directing midfacial development, shape, and pathology, our understanding of how they connect into transcription circuits is poorly 392 understood. Our previous studies had established an essential role for NCC-specific Tfap2a and Tfap2b in facial development, and in particular, jaw patterning (Van Otterloo et al., 2018). While there are severe 394 defects in both lower-face and midface regions, mechanisms underlying the latter were not investigated. In the current study, we show that TFAP2A and TFAP2B plays an essential role in CNCCs during midfacial 396 development in part through activation of a conserved ALX genetic pathway. Deploying mouse models, we found that TFAP2-dependent midface fusion is largely driven by their regulatory control of post- 398 migratory CNCCs. The combination of gene expression and transcription factor binding profiling suggest that TFAP2 paralogs act upstream of several GRN nodes critical for midface identity, midface fusion, and 400 skeletal formation—notably through ALX transcription factor gene expression. Consistent with the notion that these regulatory mechanisms are conserved across jawed vertebrates, we found that zebrafish tfap2a 402 is required for proper alx3 expression patterns and that both genes interact in the midface skeleton in this species. Given the continuum of midface anomalies associated with disrupted TFAP2A (BOFS), TFAP2B 404 (Char Syndrome), and ALX (FND) function in humans, these findings link features of previously disparate developmental disorders into a shared genetic hierarchy. 406 TFAP2 and ALX paralogs in vertebrate midface morphology and evolution. 408 Loss of TFAP2 genes in CNCCs indicates a central role for the encoded proteins in defining positional neural crest GRNs, in part, by modulating homeobox transcription factor pathways. While TFAP2 bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 410 is necessary for driving BA1 transcriptional programs—including DLX paralogs (Knight et al., 2003; Van Otterloo et al., 2018)—this study identified these factors execute a similar function in the midface. However, 412 in this tissue, TFAP2 drives a unique repertoire of transcription factor genes, including Pax3, Pax7, Msx1, and of particular interest, Alx1, Alx3, and Alx4. Past studies using induced pluripotent stem-cell derived 414 primate NCCs suggested TFAP2 and ALX proteins likely underlie species-specific facial variation (Prescott et al., 2015; Rada-Iglesias et al., 2012). More recent computational analyses of these datasets support the 416 idea that TFAP2 and ALX paralogs are linked to human facial morphology (Feng et al., 2021). Intraspecies comparisons further strengthen the role of both TFAP2 [stickleback, (Erickson et al., 2018)] and ALX 418 [Darwin’s finches, (Lamichhaney et al., 2015)] in facial evolution. Our in vivo ChIP-seq data suggests that the Alx loci (along with other lower-face and midface genes) are direct targets of TFAP2 binding. These 420 findings are consistent with the study by Prescott et al., which postulated that human TFAP2A directly regulated genes hallmarking the various positional identities, including the ALX family (Prescott et al., 422 2015). Although additional studies are required for further detailing the relationship between TFAP2 and ALX factors, our in vivo analyses in mice and zebrafish recontextualize these in vitro studies, suggesting 424 midface evolution is a direct output of gene regulation enacted by both transcription factor families. 426 TFAP2 and ALX transcription factor pathways facilitating midface fusion. How might a TFAP2-driven ALX node in the midface GRN ultimately influence CNCC behaviors 428 directing midface fusion? A combination of studies has provided some insight into the potential roles of ALX in this process. Most notably, ALX is a known regulator of skeletal differentiation across species. For 430 example, in sea urchin, ALX directly targets biomineralization and extracellular matrix genes (Khor and Ettensohn, 2020). In line with these observations, our previous cellular and skeletal analyses in zebrafish 432 implicate Alx3 as a critical determinant of midfacial chondrocyte differentiation timing (Mitchell et al., 2021). Similarly, in this study, we observed a significant increase in the gene expression of multiple collagens in 434 Tfap2NCKO mouse embryos (Fig. S11; Tables S1, S2)—whose deposition acts as a major readout of bone and cartilage formation (Dash and Trainor, 2020). Conceivably, aberrant activation of a differentiation GRN 436 mediated via loss of TFAP2 could preclude appropriate midface fusion in an ALX-dependent manner. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. There is precedent for ALX regulating additional aspects of post-migratory developmental 438 programs. In the case of cell survival, Alx1, Alx3, and Alx3/4 mutant mice exhibited elevated apoptosis in FNP CNCCs (Beverdam et al., 2001; Lakhwani et al., 2010; Zhao et al., 1996) and its periocular 440 mesenchyme lineage (Iyyanar et al., 2022), while FND3 (ALX1) patient-derived NCC lines showed an increased propensity to cell death (Pini et al., 2020). Consistent with a role in patterning, Alx1 and Alx1/4 442 mutant mice exhibited a reduction of Pax7 and an expansion of BA1-specific Lhx6 and Lhx8 into the FNP (Iyyanar et al., 2022). Interestingly, while a clear reduction in Pax7 (and Pax3) expression was noted in 444 Tfap2Δ/NCKO and Tfap2NCKO embryos, our expression profiling did not detect a concomitant expansion of Lhx6 and Lhx8. The degree to which a reduction of Alx expression in Tfap2 mutants contributes to the 446 post-migratory CNCC phenotypes observed remains elusive and will require follow-up studies. 448 Potential TFAP2 interplay with other midface positional factors. Our suite of animal and molecular studies suggests the TFAP2-ALX transcriptional circuit as a 450 major node operating in the midfacial CNCCs. Yet, also evident was the disruption of additional GRN components, such as signaling pathways and other transcription factors. For example, loss of Tfap2 452 resulted in decreased levels of various retinoic acid-related genes (Fig. 4A; Table S1, S2). This is of interest because loss of genes encoding components of the retinoic acid signaling pathway also results in a midline 454 cleft (Gao et al., 2021; Kennedy and Dickinson, 2012; Lohnes et al., 1994; Wu et al., 2022), while the morphogen itself can modulate TFAP2A gene expression and activity in vitro (Luscher et al., 1989). Thus, 456 TFAP2 may function in a regulatory feedback loop with this signaling cascade. Another pathway that is critical for midface development and that appeared to be disrupted upon loss of Tfap2 included Sonic 458 hedgehog signaling. While our gene expression profiling did not reveal significant changes in major Hedgehog signaling pathway molecules (e.g., Shh, Ptch1/2, Gli1/2/3, Smo), loss of Tfap2 resulted in an 460 upregulation of members in the Fox transcription factor family in the FNP. FOX factors normally act as downstream targets for this signaling input and functions during BA1 CNCC patterning and cartilage 462 differentiation (Jeong et al., 2004; Xu et al., 2022; Xu et al., 2018; Xu et al., 2021). Further, hyperactivation of Hedgehog activity is strongly linked with failure of the FNP to fuse at the midline (Brugmann et al., 2010; bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 464 Wu et al., 2022). Thus, it is possible that the TFAP2 midface pathway—and potentially ALX—plays a role in safeguarding FNP fusion in part by restricting Fox gene expression. 466 TFAP2 may also interact in parallel with other midfacial transcription factors. One such example includes SIX2, whose gene expression coincides with Tfap2 in the FNP (Fig. S12) (Minoux et al., 2017) 468 and is largely unaffected by loss of Tfap2 in NCCs. Loss of human SIX2 causes FND pathology (Hufnagel et al., 2016), while mice harboring a similar locus deletion produced a nasal cleft (Fogelgren et al., 2008). 470 Mouse genetic analyses suggested SIX paralogs contribute to Alx gene expression (Liu et al., 2019), thereby hinting at a transcriptional battery running in parallel with TFAP2 in the FNP. Such interplay at 472 midfacial GRN nodes requires further investigation, as it is unclear whether the SIX2 midface pathway is dependent or independent of TFAP2 activity. Given the collective contributions of both signaling pathways 474 and transcription factors to the midface GRN, addressing these remaining knowledge gaps will further advance our understanding of how TFAP2 contributes to facial shape variation, vertebrate facial evolution, 476 and the pathogenesis of midfacial disorders. 478 MATERIALS & METHODS: Animal husbandry and procedures 480 Mice: Experiments were all conducted in accordance with all applicable guidelines and regulations, following the “Guide for the Care and Use of Laboratory Animals of the National Institutes of Health”. The 482 animal protocols utilized were approved by the Institutional Animal Care and Use Committee of the University of Iowa (protocols #9012197 and #2032197) and University of Colorado (protocol #14). Noon 484 on the day of an observable copulatory plug was denoted as embryonic day 0.5 (E0.5). Additionally, all dissections were performed in 1X PBS (Fisher Scientific) treated with Diethyl Pyrocarbonate (Fisher 486 Scientific) (i.e., DEPC-PBS). Yolk sacs or tails clips were used to extract DNA for all PCR-genotyping as previously described, with the listed primers (Table S3, Tab 1) (Van Otterloo et al., 2018). During single- 488 cell RNA sequencing (scRNA-seq), we deployed a rapid lysis protocol using reagents from the SYBR Green Extract-N-Amp Tissue PCR Kit (Sigma) (Doan and Monuki, 2008). Yolk sacs were incubated in a bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 490 4:1 mixture of Extraction and Tissue Preparation Solutions first at room temperature (RT) for 10 min and subsequently at 95°C for 3 min. Neutralization Solution B was then added to the lysate, which was briefly 492 vortexed prior to immediate PCR-genotyping. Embryonic lethality was denoted for a subset of E18.5 embryos in the Wnt1:CRE model (see 494 ‘Mouse alleles and breeding schemes’, Table S4), a consequence of reduced norepinephrine and compromised survival of neural crest cell (NCC)-derived tissues in the sympathetic nervous system 496 (Schmidt et al., 2011). To prevent this lethality, pregnant dams received a water supplement to drink ad libitum (Hendershot et al., 2008; Lim et al., 2000). USP grade L-phenylephrine (β-adrenergic receptor 498 agonist), isoproterenol (α-adrenergic receptor agonist), and L-ascorbic acid (agonist preservative) were added to their drinking water from E8.5 to E18.5, which was replaced every 3-5 days. Like previous studies 500 (Abe et al., 2021), we note the supplementation did not alter craniofacial phenotypes in control or mutant embryos compared to those without supplementation. 502 Zebrafish: All work with zebrafish was approved by the University of Colorado Institutional Animal Care and Use Committee (protocol #00188). All zebrafish experiments were performed on the AB strain. 504 Animals were maintained, staged, and genotyped as previously described (Knight et al., 2003; Mitchell et al., 2021). 506 Mouse alleles and breeding schemes Wnt1:CRE (Danielian et al., 1998), Sox10:CRE (Matsuoka et al., 2005), Gt(ROSA)26Sortm1Sor/J 508 [r26r-lacZ; (Soriano, 1999)], and Gt(ROSA)26Sortm4(ACTB-tdTomato,-EGFP)Luo/J [mTmG; (Muzumdar et al., 2007)] mice were obtained from the Jackson Laboratory (Bar Harbor, ME). The Tfap2a and Tfap2b 510 alleles used in this study include Tfap2a-null (Zhang et al., 1996), Tfap2atm2Will/J [Tfap2a floxed conditional; (Brewer et al., 2004)], Tfap2b-null (Martino et al., 2016), and Tfap2btm2Will [Tfap2b floxed 512 conditional; (Van Otterloo et al., 2018)]. Females in this study were double homozygous for Tfap2a and Tfap2b floxed conditional alleles 514 (Fig. 1A-C). For lineage tracing and fluorescence-activated cell sorting, the dams were also homozygous for r26r-lacZ and mTmG alleles, respectively. Conversely, the sires were CRE hemizygous and bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 516 heterozygous for either Tfap2a/b-null alleles (used in the null/conditional model, Fig. 1A) or Tfap2a/b floxed conditional alleles (used in the conditional/conditional model, Fig. 1B) (Van Otterloo et al., 2018; Woodruff 518 et al., 2021). Using either NCC-specific Wnt1:CRE or Sox10:CRE transgene, we were able to acquire CRE-negative controls as well as littermates that harbored NCC-specific heterozygosity for Tfap2a and 520 Tfap2b (Tfap2HET), NCC-specific loss of Tfap2a (Tfap2aNCKO), NCC-specific loss of Tfap2b (Tfap2bNCKO), and NCC-specific loss of both Tfap2a and Tfap2b (Tfap2NCKO) (Fig. 1C). To distinguish between breeding 522 schemes, genotype nomenclature for the null/conditional model was denoted with the “Δ/NCKO” superscript (i.e., Tfap2aΔ/NCKO, Tfap2bΔ/NCKO, Tfap2Δ/NCKO) (Fig. 1C). 524 Zebrafish alleles The transgenic line Tg(fli1a:EGFP)y1 (Lawson and Weinstein, 2002) and the germline mutant lines 526 for tfap2a lowts213 and alx3co3003 have been previously reported (Knight et al., 2003; Mitchell et al., 2021). Scanning electron microscopy 528 Preparation of E11.5 CRE-negative control, Tfap2HET, and Tfap2NCKO littermate heads and imaging was performed as previously described (Van Otterloo et al., 2022). Briefly, samples were fixed in 4% 530 paraformaldehyde, and then rinsed in 1X PBS. Subsequently, samples were rinsed in ddH2O, air-dried for at least 3 days, mounted on an SEM stub, and sputter coated for 4.5 minutes at 5 mA using a 532 gold/palladium target in an Emitech K550X sputter coater. Scanning electron micrographs were acquired using a Hitachi (Tokyo, Japan) S4800 electron microscope operated in high-vacuum mode at 1.8kV. 534 Imaging was performed through the University of Iowa Central Microscopy Research Facility workflow. β-galactosidase staining 536 E10.5 littermates with Wnt1:CRE-mediated recombination of the r26r-lacZ reporter allele were subject to β-galactosidase staining as previously described (Van Otterloo et al., 2016). bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 538 Tissue section analyses Hematoxylin and Eosin staining: Histology staining analyses of E17.5 control and Tfap2Δ/NCKO 540 embryos was performed as previously described (Van Otterloo et al., 2018). Cryosectioning: Cryosectioning was performed as previously described (Van Otterloo et al., 2016), 542 with minor modifications. In brief, E11.5 wild-type CD-1 mice were fixed overnight in 4% PFA at 4°C. Fixed embryos were then transferred through increasing concentrations of Tissue Tek OCT compound (Sakura 544 Finetek USA, Inc.), embedded in 100% OCT, and stored at -80°C when not immediately processed. Embryonic heads were sectioned in a horizontal fashion (10 µm slices) using a Microm HM525 NX Cryostat 546 Microtome (ThermoFisher) housed within the University of Iowa Central Microscopy Research Facility, mounted onto positively charged glass slides, and were left in a closed container at RT overnight to further 548 adhere. We used the eMouseAtlas database (Armit et al., 2017) to reference the horizontal section angle. Immunofluorescence: Sectioned tissue was blocked for 1.5 hours in 5% milk or 3% BSA dissolved 550 in 0.1% Triton-X/PBS (PBSTx). Wild-type sections were used to examine protein expression of TFAP2A (3B5-supernatant, Santa Cruz Biotechnology; 1:25 dilution) and TFAP2B (2509S, Cell Signaling 552 Technology; 1:25 dilution). After washing with PBSTx, samples were incubated with Alexa Fluor 488/568 secondary antibodies (Invitrogen; 1:500 dilution) for 1.5 hours. We used DAPI to counterstain nuclei. Stains 554 were imaged via Zen software on a Zeiss 700 LSM confocal microscope. Images were exported as Zstacked, maximum intensity projection TIFF images. 556 Micro-computed tomography The University of Iowa Small Animal Imaging Core Facility provided scanning and software 558 services. E12.5 CRE-negative control and Tfap2NCKO littermates were prepared as previously described (Hsu et al., 2016). All fixing, washing, and incubation steps were performed at 4°C. In brief, embryos were 560 fixed in 4% PFA overnight and washed thrice and stored in 1X PBS until the day of scanning was determined. Three days prior to scanning, embryos were incubated in 1 N (v/v) iodine solution (Sigma). 562 Control and mutant embryos were then mounted above 1% agarose in the same screw-top conical tube, submerged in 1X PBS, and immediately imaged at 5- or 6-µm resolution. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 564 Sectional analysis and measurements of the µCT scans were performed on Dragonfly software (Object Research Systems). Embryo scans were oriented to a frontal plane to examine nasal 566 morphologies. Embryos were also oriented in a horizontal fashion, spanning from the nasal pits to the hindbrain to quantify the distance between the anterior tips of the mFNP. Control and mutant pairs were 568 imaged using the same parameters. A Student’s t-test was performed on GraphPad Prism using three independent control-mutant pairs to compare mFNP tip distances. A p-value less than 0.05 was considered 570 statistically significant. Concurrent Alizarin Red and Alcian Blue staining preparation 572 Concurrent Alizarin Red/Alcian Blue staining of E18.5 embryos and E15.5 Alcian Blue staining of null/conditional embryos were performed as previously described (Van Otterloo et al., 2016). Note that 574 Alcian blue staining for E18.5 nasal capsule elements are highly variable. Bulk RNA-sequencing 576 Tissue preparation, cDNA library generation, and sequencing: E10.5 FNP and MxP bulk tissue from three control and three Tfap2Δ/NCKO littermates were processed for RNA-seq as previously described 578 (Van Otterloo et al., 2016). After PCR-genotyping, samples stored in RNAlater (ThermoFisher) were processed for RNA (Norgen Biotek) and further purified with the RNAeasy kit (Qiagen). mRNA quality was 580 determined with DNA Analysis ScreenTape (Agilent Technologies), and cDNA libraries were generated using the Illumina TruSeq Stranded mRNA Sample Prep Kit (Illumina). Samples were sequenced on the 582 Illumina HiSeq2500 platform as 150-bp single-end reads. Library construction and sequencing was carried out by the Genomics and Microarray Core on the University of Colorado Anschutz Medical Campus. 584 RNA-seq dataset processing and analyses: Following sequencing, reads were demultiplexed, and FASTQ files processed using two distinct bioinformatic pipelines. Two pipelines were used to help reduce 586 false positives. In the first approach, FASTQ files were mapped to the mm10 genome using STAR Aligner (Dobin et al., 2013) with default settings. Gene expression was then quantified using StringTie (Pertea et 588 al., 2016; Pertea et al., 2015), with the following settings: -e (expression estimation mode), -G (mm10 reference annotation file), -A (gene abundances). Following quantification, count matrices were produced bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 590 using the prepDE.py script, available from StringTie, and used as input for DESeq2 (Love et al., 2014). Differential gene expression analysis was conducted in DESeq2 using the DESeqDataSetFromMatrix 592 function. For the second approach, FASTQ files were pseudo-aligned to the mm10 genome using the kallisto (Bray et al., 2016) quant function with the following settings: -b 50, --single, -l 100, -s 20. After 594 alignment, differential gene expression analysis was conducted using sleuth (Pimentel et al., 2017), along with the sleuth_prep function and the following settings: full_model = genotype, gene_model = TRUE, 596 read_bootstrap_tpm = TRUE, extra_bootstrap_summary = TRUE, transformation_function = function(x) log2(x + 0.05)). The resulting ‘sleuth object’ was then subjected to the sleuth_fit function, followed by the 598 sleuth_wt function to calculate differentially expressed genes between genotypes. The kallisto-sleuth pipeline output was used, along with ggplot2 (Wickham, 2016) and ggrepel, to generate and visualize 600 differentially expressed genes by volcano plot. Genes that both had, 1) a fold-change greater or less than +1.25 or -1.25, respectively, and an adjusted p-value less than 0.05 (pipeline 1), and 2) a q-value less than 602 0.1 (pipeline 2) (Table S1), were used for ontology, enrichment, and pathway analysis using the Enrichr platform (Chen et al., 2013; Kuleshov et al., 2016; Xie et al., 2021). 604 Reanalysis and visualization of published RNA-seq datasets: To directly compare the E11.5 mandibular and frontonasal prominence mesenchyme (e.g., Fig. 2) a procedure like ‘pipeline 2’ above 606 (kallisto and sleuth) was used. Specifically, triplicate bam files were downloaded from the Facebase.org repository (FB00000867) for both E11.5 mandibular prominence mesenchyme (Biosample 30H6; RID 608 33SA, 33SE, 33SJ) and E11.5 frontonasal prominence mesenchyme (Biosample 2Y6P; RID: 33TT, 33TY, 33V2). Next, bam files were converted to fastq files using the bam2fq function in samtools (Li et al., 2009), 610 Generated FASTQ files were then used along with kallisto and sleuth for differential gene expression analysis. Data was plotted using ggplot2 and ggrepel. 612 Mouse single-cell RNA-sequencing Single-cell dissociation and cell sorting: To isolate CNCCs, E11.5 embryonic heads were cut just 614 below BA2 and the otic vesicle by microdissection. Using Wnt1:CRE-mediated recombination of the mTmG reporter allele (Muzumdar et al., 2007), CRE-positive embryos were selected for further processing. Tissue bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 616 samples were then subject to a cold protease single-cell dissociation protocol (Adam et al., 2017), with modifications. Embryo heads were incubated in a protease cocktail comprised of Bacillus licheniformis 618 Subtilisin A (Creative Enzymes), Accumax, and Accutase (Fisher Scientific) suspended in DEPC-PBS as previously described (Sekiguchi and Hauser, 2019). The tissue was then carefully disrupted using a 620 disposable Eppendorf pellet pestle and placed on a rotator at 4°C for 30 min. During this step, samples were gently triturated every 10 minutes using a wide bore 1 mL pipette tip. The enzymatic reaction was 622 quenched using an equal volume of 10% heat-inactivated FBS (Fisher Scientific) in 1X phenol red-free DMEM (ThermoFisher). Dissociated cells were washed thrice in DEPC-PBS by centrifugation at 600 x g 624 at 4°C for 5 min and then transferred, through Bel Art SP Scienceware Flowmi 40-μm strainers (ThermoFisher), into chilled flow tubes precoated with DEPC-PBS. After GFP-positive sorting on the 626 University of Iowa Flow Cytometry Facility’s ARIA II system (Becton Dickinson), CNCCs were resuspended in DEPC-PBS containing 0.04% non-acetylated BSA (New England Biolabs). Cell viability was 628 determined to be >95% by Trypan Blue staining. During cell sorting, genotypes were confirmed by rapid genotyping, allowing isolation of GFP-positive cells from two Tfap2HET control and two sibling-matched 630 Tfap2NCKO mutant embryos. Cells from each “biological replicate” were combined to generate a single control and mutant sample. 632 Library preparation & sequencing: ~6,000 GFP-positive CNCCs from each condition were subject to library preparation by the University of Iowa Genomics Division on the 3’ expression scRNA-seq 10X 634 Chromium v3.1 pipeline. Libraries were sequenced on an Illumina NovaSeq 6000 platform as 100-bp paired-end reads. Over 20,000 reads per cell were acquired, resulting in ~1.5 x 108 reads per condition. 636 Quality control, integration, and UMAP visualization: Sequencing results were demultiplexed and converted to a FASTQ format using the Illumina bcl2fastq software. Reads were processed and aligned to 638 the mm10 reference genome assembly with the Cell Ranger Count function. Seurat v4.3.0 was used for quality control, filtering out cells with >10% mitochondrial counts, less than 200 features, and more than 640 7,500 features. We then integrated Tfap2HET and Tfap2NCKO conditions (Hao et al., 2021), where RNA counts were normalized and FindVariableFeatures was run with the following parameters: 642 selection.method = “vst”, nfeatures = 2,000. Cell clustering by Uniform Manifold Approximation and bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Projection (UMAP) plots was deployed by RunPCA and then RunUMAP with dims 1:30. The UMAP 644 visualizing all CNCCs was generated with FindClusters using a resolution of 0.40. Major CNCC derivatives were annotated based on published sectional immunofluorescence and scRNA-seq profiling (Soldatov et 646 al., 2019). Gene expression was plotted on UMAP plots by FeaturePlots. Gene expression mapping of CNCC lineage markers was done using the FeaturePlot function. 648 Cell cycle analysis: The CellCycleScoring function was performed as previously described, with minor modifications (Hao et al., 2021). The UMAP resolution was changed to 0.05 to segregate the major 650 CNCC lineages as individual clusters. Bioconductor BiomaRt v2.46.3 (Durinck et al., 2005) was used to convert a published list of cell cycle genes (Kowalczyk et al., 2015) from human to mouse gene 652 nomenclature. Cell cycle scores were then mapped directly onto the UMAP and graphed, as a percentage relative to each CNCC lineage, using GraphPad Prism. 654 Cellular distribution analysis: To examine genotype cellular distributions in each cluster, each condition was expressed as a percentage per cluster. These analyses were visualized in GraphPad Prism. 656 For the initial UMAP containing all sorted cells, a resolution of 0.05 was used. For the MAGIC-clustered UMAP, a resolution of 0.10 was used (see “Facial prominence mesenchyme annotation”). 658 Facial prominence mesenchyme annotation: To better define the facial prominence subpopulations, we subset and re-clustered the “Mesenchyme” clusters using MAGIC v2.0.3 (van Dijk et 660 al., 2018) with RunUMAP dims 1:20. Major mesenchyme populations were then annotated based on known gene markers, which were plotted on the MAGIC-clustered UMAP with DefaultAssay = 662 “MAGIC_RNA”. A cluster resolution of 0.10 was chosen to best visualize each prominence population and the second branchial arch as individual clusters based on published in situ data, ChIP-seq, ATAC-seq, and 664 transcriptomic datasets (Gu et al., 2022; Hooper et al., 2020; Minoux et al., 2017). To visualize coexpression of two genes on the UMAP, the flag blend = “TRUE” was used in the FeaturePlots function. 666 Gene-set and gene expression analyses: The UMAP resolution was set to 0.05 to distinguish the major cell groupings, to which we then subset the “pseudobulked” mesenchyme group. We then used the 668 FindMarkers function with ident.1 = Tfap2HET and ident.2 = Tfap2NCKO on CNCCs, to generate a list of bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. differentially expressed genes (adjusted p-value < 0.05 and a log fold-change threshold of 0.10) (Table 670 S2, Tab 9). For gene-set enrichment, ontology, and pathway analysis, lists of either up- or downregulated genes, generated from the mesenchyme “peudobulk” approach, in the Tfap2NCKO condition were inputted 672 in the Enrichr pipeline. Control-versus-mutant expression analysis of select genes was visualized as a violin plot using the VlnPlot function on the MAGIC-clustered UMAP, with DefaultAssay = “RNA”. 674 Differential gene expression was validated for a select set of genes-of-interest by targeted approaches (see real-time PCR). Gene-set overlap with bulk RNA-seq datasets: First, to determine overlapping genes between the 676 E11.5 scRNA-seq and E10.5 bulk RNA-seq datasets, gene lists of differentially expressed genes were 678 compiled (Table S1, Tab 11). Datasets used included the following outputs: E11.5 scRNA-seq mesenchyme “pseudobulk” (Table S1, Tab 10), E10.5 bulk FNP/MxP (i.e., ‘upper-face’) pipeline 1 (Table 680 S1, Tab 4) and pipeline 2 (Table S1, Tab 7), and E10.5 bulk MdP (i.e., ‘lower-face’) pipeline 1 and pipeline 2 (Van Otterloo et al., 2018). Second, the VLOOKUP function in Excel was used to intersect all 4 E10.5 682 gene lists with the E11.5 pseudobulk gene list and generate a comprehensive overlap table (Table S1, Tab 2). 684 Real-time PCR For the null/conditional model, E10.5 bulk FNP tissue was acquired from CRE-negative controls 686 and Tfap2Δ/NCKO mutants by micro-dissection. For the conditional/conditional model, E11.5 MxP and FNP tissue from Tfap2HET and Tfap2NCKO embryos were microdissected, and CNCCs were collected by the 688 Wnt1:CRE-mTmG sorting strategy. For the latter, single-cell dissociation was conducted using 0.25% trypsin (ThermoFisher) for 15 minutes at 37°C. Samples were either stored in RNAlater (ThermoFisher) at 690 -20°C or immediately processed for RNA using the RNeasy Mini Kit as per manufacturer’s protocol (Qiagen). Real-time analyses were performed as previously described (Van Otterloo et al., 2018). After 692 PCR-genotyping, cDNA was prepared with SuperScript III First-Strand Synthesis Kit (Invitrogen/ThermoFisher), where equal amounts of RNA template were processed from each condition. 694 Real-time reactions were performed on a CFX Connect instrument (Bio-Rad) with either SYBR Green or bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Select Master Mixes (Applied Biosystems, ThermoFisher). Primers targeting Alx1, Alx3, and Alx4 696 transcripts were designed to target exons flanking intronic sequences. Relative mRNA expression levels were normalized to B2m or Actb transcripts. Real-time PCR primers are listed in Table S3, Tab 2. 698 Whole mount in situ hybridization Control and Tfap2Δ/NCKO embryos at E9.0 and E10.5 were subject to whole mount in situ 700 hybridization with Alx3 and Alx4 RNA probes as previously described (Van Otterloo et al., 2018). Primers used to generate the probes are listed in Table S3, Tab 3. 702 Chromatin Immunoprecipitation followed by sequencing Tissue preparation, library construction, and sequencing: Chromatin immunoprecipitation (ChIP) 704 was conducted, essentially as previously described (Van Otterloo et al., 2022), with the exception that full facial prominences were used, rather than just isolated ectoderm. Further modifications included the use 706 of an anti-TFAP2A antibody (5 µg, sc-184, Santa Cruz Biotechnology) and Dynabeads (ThermoFisher Scientific). Following precipitation and subsequent quality control analysis (ScreenTape analysis, Agilent), 708 samples (i.e., sc-184 IP and input DNA) were sequenced on the Illumina HiSeq 2500 platform as 150-bp single-end reads. Library construction and sequencing was carried out by the Genomics and Microarray 710 Core at the University of Colorado Anschutz Medical Campus. Peak calling and analyses: Following sequencing and demultiplexing, FASTQ files were mapped 712 to the mm10 genome using BWA (Li and Durbin, 2009), followed by sam to bam file conversion using Samtools (Li et al., 2009). To identify ‘enriched’ regions (i.e., regions with significantly more reads in the 714 TFAP2 IP versus input), MACS2 was utilized with the following settings: -f BAM, -g mm, --keep-dup 2 (Table S2, Tab 3). Motif and peak annotation were completed using the findMotifsGenome.pl and the 716 annotatePeaks.pl functions in HOMER (Heinz et al., 2010). Heatmap visualization of enriched regions was completed using the bamCoverage, computeMatrix and plotHeatmap function in deepTools (Ramirez et 718 al., 2016), with the following settings: -of ‘bigwig’, --ignoreDuplicates, -bs 25, --normalizeUsing RPKM (for bamCoverage); reference-point, -b 3000, -a 3000 –referencePoint center (for computeMatrix); -- 720 whatToShow ‘heatmap and colorbar’, --missingDataColor 1, --kmeans 7, --refPointLabel “peak” (for bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. plotHeatmap). For computeMatrix (Fig. 5E) the following scaled bigwig datasets were downloaded from 722 GEO (Barrett et al., 2013) and used (along with the MACS2 generated ‘TFAP2-bound’ bed file): GSM2371708, E10.5 FNP H3K4me2; GSM2371717, E10.5 FNP H3K27ac; GSM2371732, E10.5 FNP 724 ATAC-seq; GSM2371713, E10.5 FNP H3K27me3 (peak-cluster annotations in Table S2, Tab 3) (Minoux et al., 2017). Additional datasets used for analysis included: GSM2371710, E10.5 MdP H3K4me2; 726 GSM2371719, E10.5 MdP H3K27ac; GSM2371734, E10.5 MdP ATAC-seq; GSM2371715, E10.5 MdP H3K27me3. The Integrative Genomics Viewer (Robinson et al., 2011) was used for gene level visualization 728 of these datasets at Alx loci. Pathway analysis was conducted using GREAT (McLean et al., 2010) with the following setting: ‘Basal plus extension’, ‘Proximal 5 kb upstream, 1 kb downstream, plus Distal: up to 730 100 kb’. To intersect peak coordinates with dysregulated genes, a data matrix was generated (Table S2, 732 Tab 11) that included all genes in the GREAT database (mm10, n = 20,510), their TFAP2 peak association (yes/no) using the criteria outlined above, along with gene lists from the E10.5 bulk RNA-seq analysis and 734 the E11.5 mesenchyme “pseudobulk” analysis. The E10.5 bulk RNA-seq datasets included the ‘upperface’ pipeline 1, pipeline 2, and their overlap (this study) along with the ‘lower-face’ pipeline 1, pipeline 2, 736 and their overlap (Van Otterloo et al., 2018). Total genes with or without peaks (excluding those that did not have a 1:1 match between datasets) were summed and summarized (Table S2, Tab 2). To correlate 738 peak status with gene expression in the midface, genes with and without an associated TFAP2 peak were paired with previously calculated FPKM values from the E11.5 FNP mesenchyme (Hooper et al., 2020) 740 (Table S2, Tab 12). Data was plotted and visualized using ggplot2 and statistical significance calculated using ggpubr. 742 Zebrafish single-cell RNA sequencing analysis Expression mapping on UMAPs of published zebrafish CNCC scRNA-seq datasets were performed 744 as previously described (Stenzel et al., 2022). bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. in situ hybridization chain reaction v3.0 and fluorescence microscopy 746 We ordered oligos from Molecular Instruments targeting the same alx3 5’ UTR sequence previously used for in situ hybridization (Mitchell et al., 2021) and followed the manufacturer’s protocol for whole- 748 mount zebrafish embryos and larvae (Choi et al., 2016). High salt content in buffers used during the hybridization chain reaction (HCR) v3.0 protocol made genotyping difficult. To circumvent this, we used 750 Promega GoTaq Flexi (M8295) in “M” buffer (2mM MgCl2, 14mM Tris-HCl pH 8.4, 68.25 mM KCl, 0.0013% gelatin, 1.8 mg/mL BSA, 140 µM each dNTP). Embryos or larvae were then embedded in 0.2% agarose 752 and imaged with an Andor Dragonfly 301 spinning disk confocal system. Acquisition parameters and fluorescence adjustments were applied linearly and equally to all samples. 754 Zebrafish genetic interaction analyses Zebrafish larvae 6 days post-fertilization were subject to concurrent Alizarin Red/Alcian Blue 756 staining preparations as previously described (Walker and Kimmel, 2007). Nomarski imaging, genotypeblinded phenotype scoring, and statistical analysis were performed as previously described (Bailon- 758 Zambrano et al., 2022; Mitchell et al., 2021; Sucharov et al., 2019). In brief, larval skeletons were dissected and flat mounted for imaging on a Leica DMi8 inverted microscope equipped with a Leica DMC2900. For 760 penetrance scores, we performed the Fisher’s exact test using GraphPad. All sample sizes and p-values are included in Fig. 6F. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 762 FOOTNOTES: Acknowledgements: The authors would like to acknowledge the University of Iowa Small Animal Imaging 764 Core, Central Microscopy Research Facility, Flow Cytometry Facility, and Genomics Division for their technical support in µCT, tissue processing, and scRNA-seq experiments. These facilities are funded 766 through user fees and generous financial support from the Carver College of Medicine. To Dr. Michael Chimenti from the Bioinformatics Division, Professor Huojun Cao from the College of Dentistry Division of 768 Biostatistics and Computational Biology, Nate Mullin, and Yann Vanrobaeys for their insights into scRNAseq analyses. To Jamie Thompson for technical input on immunofluorescence experiments and additional 770 care for animals used in this study. Lastly, we are grateful to all Van Otterloo lab members and our colleagues in the Craniofacial Interest Group for their invaluable feedback, particularly Professors Colin 772 Kenny, Robert Cornell, Brad Amendt, Martine Dunnwald, and John Manak. Funding: Training support was provided by the National Institute for Dental and Craniofacial Research 774 (NIDCR) [T90DE023520 to TTN; F32DE029995 to JMM]. This research was funded by the University of Iowa Graduate and Professional Student Government Research Grant (to TTN), NIDCR [2R01DE12728 776 to TJW; R01DE029193 to JTN; R00DE026823 to EVO], alongside University of Iowa College of Dentistry start-up & seed grant funds (to EVO). 778 Author Contributions: Conceptualization: TTN, JMM, TJW, JTN, EVO 780 Methodology: TTN, JMM, TJW, JTN, EVO Software: TTN, JMM, KLJ, EVO 782 Validation: TTN, JMM, EVO Formal analysis: TTN, JMM, JTN, EVO 784 Investigation: TTN, JMM, MDK, EVO Resources: KLJ, TW, JTN, EVO 786 Data Curation: TTN, JMM, EVO Writing – original draft: TTN, JMM 788 Writing – review and editing: TTN, JMM, TJW, JTN, EVO Visualization: TTN, JMM, EVO 790 Supervision: TJW, JTN, EVO Project administration: TJW, JTN, EVO 792 Funding acquisition: TTN, TJW, JTN, EVO bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. FIGURE LEGENDS: 794 Fig. 1 – TFAP2A and TFAP2B cooperatively function in CNCCs during midfacial development. (A, B) Schematics depicting the mouse breeding scheme used for the null/conditional model (A), previously 796 described (Van Otterloo et al., 2018), and the conditional/conditional model (B) used in this study. Cartoons are adapted from BioRender. (C) Table summarizing a subset of the combination of alleles that can be 798 obtained using the breeding scheme highlighted in panels A or B, along with current and previous shorthand nomenclature. Note, null/conditional genotypes are denoted with a “Δ” superscript in the current 800 study. (D-G) Lateral (D-G) or frontal (D’-G’) views of conditional/conditional mouse embryos at embryonic day 18.5 (E18.5), with indicated genotypes. Gold arrowheads in panels E and F point to indented snouts. 802 Insets in panels D'-G’ include higher magnification images of the snout, with misplaced vibrissae outlined in a gold dashed circle. White arrowhead in G indicates the shortened snout and the asterisk marks the 804 short mandible. N = 5 per genotype. Fig. 2 – TFAP2A and TFAP2B function during post-migratory CNCCs to facilitate midface fusion. 806 (A) Immunofluorescent analysis of TFAP2A (left, green) and TFAP2B (right, red) on a horizontal tissue section through an E11.5 wild type embryonic midface. White dashed lines demarcate ectoderm- 808 mesenchyme boundaries. (B) Volcano plot displaying differentially expressed genes between E11.5 CNCCs occupying the mandibular prominence (MdP, violet) and those occupying the frontonasal 810 prominence (FNP, orange). (C, D) µCT horizontal sections of E12.5 control-Tfap2NCKO littermates in either the Wnt1:CRE (C) or Sox10:CRE (D) models. Yellow dashed lines outline the medial domains of the 812 frontonasal prominence (mFNP). (E) Quantification of the linear distance between the tips of the mFNP domains in control and Tfap2NCKO embryos in both mouse models. N = 3 per genotype. Student’s t-test, ** 814 p < 0.01, *** p < 0.001. (F, G) µCT frontal sections through the snout. Gold dashed lines highlight the midfacial cleft. Additional abbreviations: d, dorsal; ect, ectoderm; lFNP, lateral FNP; ne, nasal epithelium; 816 np, nasal pit; v, ventral. Fig. 3 – Formation of midfacial bone and cartilage structures depends on Tfap2a and Tfap2b gene 818 dosage in CNCCs. (A-H) Concurrent Alizarin Red/Alcian Blue stained skeletal preparations of E18.5 bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. embryos, with indicated genotypes, from Sox10:CRE (A-D) or Wnt1:CRE (E-H) based breeding schemes. 820 Anterior (a) is to the left. Bone is stained magenta while cartilage is stained blue. However, note that cartilage staining towards the anterior end is variable that precluded thorough examination of the nasal 822 capsules. Top-down, dorsal, (A-H) or bottom-up, ventral, (A’-H’) views of the midfacial skeleton, with the maxillary bone removed. Gold dashed lines (A-H) outline the peripheral edges of the calvarial bones. White 824 dashed lines (B, F, G) outline nasal bone gaps. Gold arrowheads (D, F-H) point to cartilaginous ectopias sitting adjacent to the frontal bones. White arrowheads (B’, F’) point to a gap between the anterior portion 826 of the vomer bones. Asterisks (D, H) highlight missing nasal bones. N = 5 per genotype. Scale bar = 1 mm. Additional abbreviations: cs, coronal suture; fr, frontal bone; na, nasal bone; ns, nasal septum; nl, 828 nasal/ethmoid labyrinth; p, posterior; pl, palatine bone; pmx, premaxilla; ppp, palatal process of the palatine; pppm, palatal process of the premaxilla; ppro, pila preoptica; ppso, pila postoptica; pr, parietal 830 bone; ps, presphenoid; vm, vomer bone. Fig. 4 – Transcriptomic analyses reveal NCC-specific loss of Tfap2 compromises midface Alx1/3/4 832 gene expression. (A) Bulk RNA-seq volcano plot analysis of frontonasal and maxillary prominence tissue (FNP, MxP) isolated from E10.5 control and Tfap2Δ/NCKO littermates (cartoon workflow on the left). Down- 834 regulated genes are visualized in purple while up-regulated genes are gold. (B) Enrichr (Kuleshov et al., 2016) terms based on the downregulated gene-set from the bulk RNA-seq analysis. (C) Overview of the 836 scRNA-seq experiment performed on E11.5 Tfap2HET and Tfap2NCKO littermates. The top box displays the workflow of sorting GFP-positive head cranial neural crest cells (CNCCs) from branchial arch 2 (BA2) and 838 upward. Cartoon partly derived from BioRender. Displayed at the bottom is the resulting Uniform Manifold Approximation and Projection (UMAP) plot of the three major groupings and genes enriched in each. 840 Outlined is the mesenchyme population that is further subset and re-clustered (D) using MAGIC imputation (van Dijk et al., 2018). Dashed lines on the mesenchyme UMAP demarcate the identified positional identity 842 clusters, with their respective gene signatures, of the head CNCCs. These include clusters for the FNP (two generated), MxP, mandibular prominence (MdP), and BA2. (E) Violin plots of Alx1/3/4 gene 844 expression between conditions and divided based on clusters. Boxed are the FNP clusters. (F) Whole mount in situ hyberization for Alx3/4 between E10.5 controls and Tfap2Δ/NCKO mutants, viewed in a frontal bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 846 fashion. Scale bar = 500 µm. Dashed lines outline the FNP, while arrowheads point to the dysregulated Alx3/4 staining in the MdP that was previously characterized (Van Otterloo et al., 2018). (G) Real-time 848 PCR analysis for Alx1/3/4 gene expression in the conditional/conditional model (left) and null/conditional model (right), with indicated genotypes and timepoints. Note that in the conditional/conditional model used 850 sorted CNCCs from the FNP/MxP. Meanwhile, the null/conditional model examined expression profiles in bulk tissue isolated from individual medial and lateral domains of the FNP (mFNP, lFNP). 852 Fig. 5 – ChIP-seq profiling identifies TFAP2 paralogs directly bind Alx1/3/4 regulatory elements. (A) Schematic depicting workflow for anti-TFAP2 chromatin immunoprecipitation followed by sequencing 854 (ChIP-seq) of wild-type E11.5 C57BL/6 facial prominence tissues. (B) Density heatmap displaying readdepth at the 13,778 TFAP2 ‘peaks’ (y-axis) detected by anti-TFAP2 ChIP-seq, relative to non- 856 immunoprecipitated input. (C) A hollow pie chart summarizing the distribution of TFAP2 ‘peaks’ throughout the genome, relative to key features (e.g., promoters, introns, etc.). The TFAP2 motif, along with the 858 significance of its enrichment in all TFAP2 ‘peaks’, is displayed in the center. (D) Box-and-whisker graph plotting the FPKM values of 16,573 genes with frontonasal prominence (FNP) mesenchymal expression 860 at E11.5 (Hooper et al., 2020), with or without an associated TFAP2 peak (p-value = 2.60e-256). (E) Density heatmaps for anti-H3K4me2 (blue), anti-H3K27ac (green), ATAC-seq (forest green), and anti- 862 H3K27me3 (red) profiles of E10.5 FNP cranial neural crest cells (CNCCs) (Minoux et al., 2017) at the TFAP2-positive coordinates. These regions are divided by k-means clustering. Boxed are clusters 5 and 864 6, whose coordinates are correlated with high H3K4me2, high H3K27ac, high ATAC (i.e., chromatin accessibility), and low H3K27me3 signals. (F) Box-and-whisker graphs, as in panel D, but further 866 partitioning the ‘peak’ group based on k-means clusters from panel E. (G) GREAT (McLean et al., 2010) pathway analysis of genes associated with cluster 5 TFAP2 peaks. Significant “Human Phenotype” terms 868 are listed, with the most significant on the top. Navy shaded boxes are associated with craniofacial-specific terms. (H) IGV browser views of the Alx1, Alx3, and Alx4 loci overlayed with the epigenome signatures 870 (blue, H3K4me2; green, H3K27ac; forest green, ATAC; red, H3K27me3) of E10.5 CNCCs residing in the mandibular prominence (MdP, top 4 tracks) and FNP (tracks 5-8), as well as the cluster 5-assigned TFAP2 872 peaks (bottom, gray highlights). bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Fig. 6 – Gene expression and epistasis studies indicate zebrafish alx3 is genetically downstream 874 of tfap2a. (A-D) Frontal views of alx3 in situ hybridization (violet) and fli1a:EGFP-labelled cranial neural crest cells (green) in a wild-type control (A, C) and a tfap2a-/- sibling (B, D) at 48 hours post-fertilization 876 (hpf). Note, panels C and D are magnified views of the roof of the mouth, as boxed in panels A and B. (E) Schematic summarizing expression patterns observed, N = 5 per genotype. (F-I) Whole mount, lateral 878 images of Alizarin Red/Alcian Blue stained skeletal preparations of 6-day-old zebrafish heads, with indicated genotypes. The anterior is to the left. Red arrowhead. (F’-I’) Neurocrania were dissected and flat 880 mounted. Red arrowhead indicates ectopic cartilage near the eye socket. Black arrowhead marks dorsally projecting ectopic ethmoid plate cartilage. (J) Tables summarizing penetrance, samples, and statistics of 882 the epistasis experiments as follow: top, the removal of wild-type alx3 copies from tfap2a homozygous mutants; middle, the removal of wild-type tfap2a copies from alx3 homozygous mutants; bottom, the 884 removal of wild-type tfap2a copies from alx3 heterozygotes. Boxed values correspond to the genotypes shown in panels G, H, and I. For example, in yellow are tfap2a mutant phenotypes and their penetrance 886 (corresponding to G, G’); boxed in green those for when both alx3 copies are removed in tfap2a-mutant backgrounds (corresponding to I, I’); lastly, boxed in purple are those for alx3 mutants (H, H’). Asterisk 888 indicates significant difference compared with the far-left column. Carets indicate significant difference compared with middle column by Fisher’s exact test. Scale bars = 100 µm. Abbreviations: ep, ethmoid 890 plate; m, mouth; n, nares; ND, not determined; ps, parasphenoid. Fig. S1 – Histological analysis of midface phenotypes in Tfap2∆/NCKO mutants. (A-F) Coronal sections 892 of H&E stained anterior (A-C) and posterior (D-F) regions of the snout of E17.5 embryos, indicated by genotype. Note, the schematic in the top left highlights relative position of sections in A-C (‘Top’) and D-F 894 (‘Bottom’). Yellow lines in panels C and F demarcate the midfacial cleft. Scale bars = 500 µm. N = 3 per genotype. Abbreviations: i, incisor; ns, nasal septum; t, tongue; v, vibrissae. 896 Fig. S2 – TFAP2A and TFAP2B cooperative function in CNCCs during midface fusion. (A-C) Gross craniofacial morphology, as detected by scanning electron microscopy (A) or brightfield imaging (B, C) of 898 E11.5 (A), E12.5 (B), or E15.5 (C) embryonic midfaces, as indicated by genotype. Panel A is ventral views, whereas panels B and C are top-down views of the cranium and midface. Black dotted lines in the E12.5 bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 900 images highlight the frontonasal prominences. White arrowheads point to the midface cleft. N = 3 per genotype. Scale bar = 1 mm. Abbreviations: FNP, frontonasal prominence; lFNP, lateral FNP; mFNP, 902 medial FNP; np, nasal pit. Fig. S3 – Tfap2NCKO mutant skeletal phenotypes recapitulate those observed in Tfap2∆/NCKO mutants. 904 (A) An E18.5 Wnt1:CRE Tfap2NCKO skeletal preparation showing ‘maxilla-like’ medial projections on the mandible that meet at the midline (left) along with syngnathia of the left jaw viewed laterally in isolation 906 (right). (B) Isolated left mandible from E18.5 Tfap2NCKO embryos, with indicated genotype. These phenotypes match those previously described in Tfap2∆/NCKO embryos (Van Otterloo et al., 2018). Scale 908 bar = 1 mm. Abbreviations: agp, angular process; cdp, condylar process; crp, coronoid process; fpm, frontal process of the maxilla; jg, jugal bone; li, lower incisor; mc, Meckel’s cartilage; mx-l, maxilla-like; 910 ppmx, palatal process of the maxilla; zpm, zygomatic process of the maxilla. Fig. S4 – Lineage tracing and immunofluorescence analysis suggest a post-migratory role for 912 TFAP2A and TFAP2B during midfacial development. (A) Front view of E10.5 embryos, with indicated genotypes. Cranial neural crest cells are β-galactosidase-stained from Wnt1:CRE-mediated recombination 914 of the r26r-lacZ reporter allele. Dashed yellow lines mark the midline-proximal edges of the medial domains of the frontonasal prominences (mFNP). White arrowhead in Tfap2NCKO embryo points to the increased 916 gap between the mFNP. Scale bar = 500 µm. (B) Co-staining of nuclei (DAPI, blue), TFAP2A (gold), and TFAP2B (red) in midface tissue. The top-left image, derived from BioRender, indicates the plane of each 918 section that TFAP2 co-staining was performed. The bottom-left image, derived from the eMouse atlas (Armit et al., 2017), is a representative orientation of the tissue section. Boxed is the region examined in 920 the immunofluorescent images. The white arrowhead indicates absent TFAP2B protein expression in the nasal epithelium, in contrast to the top panel. White dashed lines indicate boundaries between 922 mesenchyme and epithelium. Additional abbreviations: d, dorsal; ect, ectoderm; lFNP, lateral domain of the FNP; MxP, maxillary prominence; ne, nasal epithelium; np, nasal pit; v, ventral. 924 Fig. S5 – A post-migratory CNCC role for TFAP2A and TFAP2B during midfacial development. Frontal view of E18.5 Sox10:CRE-Tfap2NCKO animals, with indicated genotypes. Scale bar = 1 mm. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 926 Fig. S6 – Defects in the midfacial skeleton of Sox10:CRE Tfap2NCKO mutants. (A-A”, B-B”) Three individual E18.5 Sox10:CRE-Tfap2NCKO skeletal preparations in top-down (A) or bottom-up (B) view. 928 Anterior is to the left. Yellow dashed lines in the top-down view outline the peripheral edges of the calvarial bones while those in the bottom-up view outline the presphenoid bone. Yellow arrowheads point to the 930 cartilaginous ectopias stemming from the nasal septum. Note, in contrast to the first two embryos, the maxillary bone has been removed from the third. Further, the third exhibits isolated Alizarin red stained 932 tissue (i.e., bony ‘islands’, individually outlined) on the calvaria instead of the cartilaginous ectopias. Scale bar = 1 mm. Abbreviations: bs, basisphenoid; cs, coronal suture; fr, frontal bone; ip, interpareital bone; nl, 934 nasal/ethmoid labyrinth; ns, nasal septum; pl, palatine; pmx, premaxilla; ppp, palatal process of the palatine; ppmx, palatal process fo the premaxilla; ppro, pila postoptica; ppso, pila postoptica; pr, pareital 936 bone; vm, vomer bone. Fig. S7 – Jaw phenotypes in Sox10:CRE-Tfap2NCKO mutants partially recapitulate those observed 938 in Wnt1:CRE-Tfap2NCKO and -Tfap2∆/NCKO mutants. (A-F) E18.5 maxillary and mandibular bones in isolation, with indicated genotypes. In panels A through D, the left-side elements are presented; panels E 940 and F show the right-side elements. In panel B, circled is the missing jugal bone and thicked zygomatic process of the maxilla. In panel E, the sygnanthia is circled. In panel F, fusion between the the jugal and 942 thickened zygomatic process of the maxilla is circled. Scale bar = 1 mm. Abbreviations: agp, angular process; cdp, condylar process; crp, coronoid process; fpm, frontal process of the maxilla; jg, jugal bone; 944 li, lower incisor; mc, Meckel’s cartilage; ppmx, palatal process of the maxilla; sq, squamosal bone; zpm, zygomatic process of the maxilla. 946 Fig. S8 – Wnt1:CRE-Tfap2∆/NCKO mutants present defects in the midfacial skeleton. (A) Oblique frontal view of Alcian Blue stained preparations for E15.5 Tfap2∆/NCKO mutant embryos, indicated by genotypes. 948 The midface is outlined by black dotted lines. The black arrowhead points to ectopic cartilage forming superior to the midface cleft. (A’) Bottom-up ventral view, anterior to the left, of chondrocrania, as in panel 950 A, with a focus on nasal cartilage elements. Black arrowheads point to structural aberrations which presumably would have formed the anterior nasal capsules and labyrinths. (B) Top-down view of E18.5 952 craniums, anterior up, from the indicated genotypes. Gold dashed lines outline the edge of the calvarial bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. bones at the midline. The asterisk marks the midface cleft and missing nasal bones. N = 3 per genotype. 954 Scale bar = 500 µm. Abbreviations: cs, coronal suture; e, eye; fr, frontal bone; ip, interparietal bone; ms, metopic suture; na, nasal bone; ns, nasal septum; pn, paries nasi; ppro, pila preoptica; ppso, pila 956 postoptica; pr, parietal bone; ss, sagittal suture. Fig. S9 – Gene expression-based scRNA-seq annotation of the major CNCC lineages. Gene 958 expression of select genes mapped onto the Uniform Manifold Approximation and Projection (UMAP) plot in Fig. 4C. These genes were selected based on published scRNA-seq datasets (Soldatov et al., 2019). 960 Note that cells from the brain are included in this dataset because Wnt1:CRE labels parts of the brain. Fig. S10 – Cellular profiling of the three major CNCC lineages in the developing head. (A) Tfap2HET 962 and Tfap2NCKO representation in the integrated dataset, divided by the three major groupings and expressed as a percentage. The dashed line marks the fifty-percent threshold. (B) Gene expression-based 964 cell cycle scoring for indiivdual Tfap2HET and Tfap2NCKO conditions. Cell cycle scores are mapped on the Uniform Manifold Approximation and Projection (UMAP) plots of individual conditions (left) and visualized 966 as percentages per CNCC lineage. Fig. S11 – Enrichment analyses of downregulated genes in the scRNA-seq mesenchyme group. 968 Terms from the Enrichr pipeline (Kuleshov et al., 2016), based on (A) downregulated or (B) upregulated genes identified through the mesenchyme psuedobulk analysis (Table S2, Tab 9). 970 Fig. S12 – Gene expression-based scRNA-seq annotation of the major CNCC mesenchymal populations. MAGIC-based (van Dijk et al., 2018) gene expression of select genes mapped onto the 972 Uniform Manifold Approximation and Projection (UMAP) plot in Fig. 4D. These genes were selected based on published transcriptomic datasets (Gu et al., 2022; Hooper et al., 2020), epigenome datasets (Minoux 974 et al., 2017), and published in situ data. Fig. S13 – Tfap2 and Alx paralog transcripts are enriched in the murine FNP CNCCs. Gene 976 expression of Tfap2a and Tfap2b (blue) overlaid with individual Alx paralogs (red) onto the Uniform Manifold Approximation and Projection (UMAP) plot. Color saturation correlates to gene expression levels, 978 while degree of co-expression is read out as a blending of the two colors. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Fig. S14 – Cellular and gene expression profiling in the CNCC-derived mesenchyme. (A) Cellular 980 distributions of Tfap2HET and Tfap2NCKO conditions as Uniform Manifold Approximation and Projection (UMAP) plots (A) or as a percentage in each cluster (A’). (B) Violin expression plots of branchial arch (BA)- 982 enriched genes previously identified (Van Otterloo et al., 2018). Boxed are clusters for the maxillary prominence (MxP), mandibular prominence (MdP), and BA2. (C) Violin expression plots for genes 984 dysregulated in the midface. Boxed are the frontonasal prominence (FNP) clusters. Fig. S15 – Additional GREAT pathway analysis on TFAP2 genomic binding. Terms enriched in cluster 986 3 and 4 peaks that contained positive TFAP2 ChIP-seq signal. Note, peaks in these clusters are significantly enriched for skin (cluster 4), neuronal (clusters 3, 4), and cardiac (cluster 3) terms. 988 Fig. S16 – tfap2a and alx3 transcripts are enriched in the zebrafish frontonasal CNCCs. Gene expression of tfap2a and alx3 overlaid together on a Uniform Manifold Approximation and Projection 990 (UMAP) plot of a published single-cell RNA-seq dataset generated from sorted zebrafish cranial neural crest cells (Stenzel et al., 2022). Color saturation correlates to gene expression levels, while degree of co- 992 expression is read out as a blending of the two colors. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. REFERENCES 994 996 998 1000 1002 1004 1006 1008 1010 1012 1014 1016 1018 1020 1022 1024 1026 1028 1030 1032 1034 1036 1038 1040 Abe, M., Cox, T. C., Firulli, A. B., Kanai, S. M., Dahlka, J., Lim, K. C., Engel, J. D. and Clouthier, D. E. (2021). GATA3 is essential for separating patterning domains during facial morphogenesis. Development 148. Adam, M., Potter, A. S. and Potter, S. S. (2017). Psychrophilic proteases dramatically reduce single-cell RNA-seq artifacts: a molecular atlas of kidney development. Development 144, 3625-3632. Armit, C., Richardson, L., Venkataraman, S., Graham, L., Burton, N., Hill, B., Yang, Y. and Baldock, R. A. (2017). eMouseAtlas: An atlas-based resource for understanding mammalian embryogenesis. Dev Biol 423, 1-11. Bailon-Zambrano, R., Sucharov, J., Mumme-Monheit, A., Murry, M., Stenzel, A., Pulvino, A. T., Mitchell, J. M., Colborn, K. L. and Nichols, J. T. (2022). Variable paralog expression underlies phenotype variation. Elife 11. Barrallo-Gimeno, A., Holzschuh, J., Driever, W. and Knapik, E. W. (2004). Neural crest survival and differentiation in zebrafish depends on mont blanc/tfap2a gene function. Development 131, 1463-1477. Barrett, T., Wilhite, S. E., Ledoux, P., Evangelista, C., Kim, I. F., Tomashevsky, M., Marshall, K. A., Phillippy, K. H., Sherman, P. M., Holko, M., et al. (2013). NCBI GEO: archive for functional genomics data sets--update. Nucleic Acids Res 41, D991-995. Beverdam, A., Brouwer, A., Reijnen, M., Korving, J. and Meijlink, F. (2001). Severe nasal clefting and abnormal embryonic apoptosis in Alx3/Alx4 double mutant mice. Development 128, 3975-3986. Bhatt, S., Diaz, R. and Trainor, P. A. (2013). Signals and switches in Mammalian neural crest cell differentiation. Cold Spring Harb Perspect Biol 5. Brandon, A. A., Almeida, D. and Powder, K. E. (2022). Neural crest cells as a source of microevolutionary variation. Semin Cell Dev Biol. Bray, N. L., Pimentel, H., Melsted, P. and Pachter, L. (2016). Near-optimal probabilistic RNA-seq quantification. Nat Biotechnol 34, 525-527. Brewer, S., Feng, W., Huang, J., Sullivan, S. and Williams, T. (2004). Wnt1-Cre-mediated deletion of AP-2alpha causes multiple neural crest-related defects. Dev Biol 267, 135-152. Brugmann, S. A., Allen, N. C., James, A. W., Mekonnen, Z., Madan, E. and Helms, J. A. (2010). A primary ciliadependent etiology for midline facial disorders. Hum Mol Genet 19, 1577-1592. Chen, E. Y., Tan, C. M., Kou, Y., Duan, Q., Wang, Z., Meirelles, G. V., Clark, N. R. and Ma'ayan, A. (2013). Enrichr: interactive and collaborative HTML5 gene list enrichment analysis tool. BMC Bioinformatics 14, 128. Choi, H. M., Calvert, C. R., Husain, N., Huss, D., Barsi, J. C., Deverman, B. E., Hunter, R. C., Kato, M., Lee, S. M., Abelin, A. C., et al. (2016). Mapping a multiplexed zoo of mRNA expression. Development 143, 3632-3637. Claes, P., Roosenboom, J., White, J. D., Swigut, T., Sero, D., Li, J., Lee, M. K., Zaidi, A., Mattern, B. C., Liebowitz, C., et al. (2018). Genome-wide mapping of global-to-local genetic effects on human facial shape. Nat Genet 50, 414-423. Clouthier, D. E., Garcia, E. and Schilling, T. F. (2010). Regulation of facial morphogenesis by endothelin signaling: insights from mice and fish. Am J Med Genet A 152A, 2962-2973. Cox, S. G., Kim, H., Garnett, A. T., Medeiros, D. M., An, W. and Crump, J. G. (2012). An essential role of variant histone H3.3 for ectomesenchyme potential of the cranial neural crest. PLoS Genet 8, e1002938. Danielian, P. S., Muccino, D., Rowitch, D. H., Michael, S. K. and McMahon, A. P. (1998). Modification of gene activity in mouse embryos in utero by a tamoxifen-inducible form of Cre recombinase. Curr Biol 8, 13231326. Dash, S. and Trainor, P. A. (2020). The development, patterning and evolution of neural crest cell differentiation into cartilage and bone. Bone 137, 115409. de Croze, N., Maczkowiak, F. and Monsoro-Burq, A. H. (2011). Reiterative AP2a activity controls sequential steps in the neural crest gene regulatory network. Proc Natl Acad Sci U S A 108, 155-160. Debbache, J., Parfejevs, V. and Sommer, L. (2018). Cre-driver lines used for genetic fate mapping of neural crest cells in the mouse: An overview. Genesis 56, e23105. Doan, L. and Monuki, E. S. (2008). Rapid genotyping of mouse tissue using Sigma's Extract-N-Amp Tissue PCR Kit. J Vis Exp. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 1042 1044 1046 1048 1050 1052 1054 1056 1058 1060 1062 1064 1066 1068 1070 1072 1074 1076 1078 1080 1082 1084 1086 1088 1090 Dobin, A., Davis, C. A., Schlesinger, F., Drenkow, J., Zaleski, C., Jha, S., Batut, P., Chaisson, M. and Gingeras, T. R. (2013). STAR: ultrafast universal RNA-seq aligner. Bioinformatics 29, 15-21. Dooley, C. M., Wali, N., Sealy, I. M., White, R. J., Stemple, D. L., Collins, J. E. and Busch-Nentwich, E. M. (2019). The gene regulatory basis of genetic compensation during neural crest induction. PLoS Genet 15, e1008213. Durinck, S., Moreau, Y., Kasprzyk, A., Davis, S., De Moor, B., Brazma, A. and Huber, W. (2005). BioMart and Bioconductor: a powerful link between biological databases and microarray data analysis. Bioinformatics 21, 3439-3440. Eckert, D., Buhl, S., Weber, S., Jager, R. and Schorle, H. (2005). The AP-2 family of transcription factors. Genome Biol 6, 246. Erickson, P. A., Baek, J., Hart, J. C., Cleves, P. A. and Miller, C. T. (2018). Genetic Dissection of a Supergene Implicates Tfap2a in Craniofacial Evolution of Threespine Sticklebacks. Genetics 209, 591-605. Fan, X., Masamsetti, V. P., Sun, J. Q., Engholm-Keller, K., Osteil, P., Studdert, J., Graham, M. E., Fossat, N. and Tam, P. P. (2021). TWIST1 and chromatin regulatory proteins interact to guide neural crest cell differentiation. Elife 10. Feng, Z., Duren, Z., Xiong, Z., Wang, S., Liu, F., Wong, W. H. and Wang, Y. (2021). hReg-CNCC reconstructs a regulatory network in human cranial neural crest cells and annotates variants in a developmental context. Commun Biol 4, 442. Fernandez Garcia, M., Moore, C. D., Schulz, K. N., Alberto, O., Donague, G., Harrison, M. M., Zhu, H. and Zaret, K. S. (2019). Structural Features of Transcription Factors Associating with Nucleosome Binding. Mol Cell 75, 921-932 e926. Fogelgren, B., Kuroyama, M. C., McBratney-Owen, B., Spence, A. A., Malahn, L. E., Anawati, M. K., Cabatbat, C., Alarcon, V. B., Marikawa, Y. and Lozanoff, S. (2008). Misexpression of Six2 is associated with heritable frontonasal dysplasia and renal hypoplasia in 3H1 Br mice. Dev Dyn 237, 1767-1779. Gao, T., Wright-Jin, E. C., Sengupta, R., Anderson, J. B. and Heuckeroth, R. O. (2021). Cell-autonomous retinoic acid receptor signaling has stage-specific effects on mouse enteric nervous system. JCI Insight 6. Gu, R., Zhang, S., Saha, S. K., Ji, Y., Reynolds, K., McMahon, M., Sun, B., Islam, M., Trainor, P. A., Chen, Y., et al. (2022). Single-cell transcriptomic signatures and gene regulatory networks modulated by Wls in mammalian midline facial formation and clefts. Development 149. Han, J., Ishii, M., Bringas, P., Jr., Maas, R. L., Maxson, R. E., Jr. and Chai, Y. (2007). Concerted action of Msx1 and Msx2 in regulating cranial neural crest cell differentiation during frontal bone development. Mech Dev 124, 729-745. Hao, Y., Hao, S., Andersen-Nissen, E., Mauck, W. M., 3rd, Zheng, S., Butler, A., Lee, M. J., Wilk, A. J., Darby, C., Zager, M., et al. (2021). Integrated analysis of multimodal single-cell data. Cell 184, 3573-3587 e3529. Hari, L., Miescher, I., Shakhova, O., Suter, U., Chin, L., Taketo, M., Richardson, W. D., Kessaris, N. and Sommer, L. (2012). Temporal control of neural crest lineage generation by Wnt/beta-catenin signaling. Development 139, 2107-2117. Heinz, S., Benner, C., Spann, N., Bertolino, E., Lin, Y. C., Laslo, P., Cheng, J. X., Murre, C., Singh, H. and Glass, C. K. (2010). Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol Cell 38, 576-589. Hendershot, T. J., Liu, H., Clouthier, D. E., Shepherd, I. T., Coppola, E., Studer, M., Firulli, A. B., Pittman, D. L. and Howard, M. J. (2008). Conditional deletion of Hand2 reveals critical functions in neurogenesis and cell typespecific gene expression for development of neural crest-derived noradrenergic sympathetic ganglion neurons. Dev Biol 319, 179-191. Hooper, J. E., Jones, K. L., Smith, F. J., Williams, T. and Li, H. (2020). An Alternative Splicing Program for Mouse Craniofacial Development. Front Physiol 11, 1099. Hovland, A. S., Bhattacharya, D., Azambuja, A. P., Pramio, D., Copeland, J., Rothstein, M. and Simoes-Costa, M. (2022). Pluripotency factors are repurposed to shape the epigenomic landscape of neural crest cells. Dev Cell 57, 2257-2272 e2255. Hsu, C. W., Wong, L., Rasmussen, T. L., Kalaga, S., McElwee, M. L., Keith, L. C., Bohat, R., Seavitt, J. R., Beaudet, A. L. and Dickinson, M. E. (2016). Three-dimensional microCT imaging of mouse development from early post-implantation to early postnatal stages. Dev Biol 419, 229-236. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 1092 1094 1096 1098 1100 1102 1104 1106 1108 1110 1112 1114 1116 1118 1120 1122 1124 1126 1128 1130 1132 1134 1136 1138 1140 1142 Hufnagel, R. B., Zimmerman, S. L., Krueger, L. A., Bender, P. L., Ahmed, Z. M. and Saal, H. M. (2016). A new frontonasal dysplasia syndrome associated with deletion of the SIX2 gene. Am J Med Genet A 170A, 487491. Ishii, M., Han, J., Yen, H. Y., Sucov, H. M., Chai, Y. and Maxson, R. E., Jr. (2005). Combined deficiencies of Msx1 and Msx2 cause impaired patterning and survival of the cranial neural crest. Development 132, 4937-4950. Iyyanar, P. P. R., Wu, Z., Lan, Y., Hu, Y. C. and Jiang, R. (2022). Alx1 Deficient Mice Recapitulate Craniofacial Phenotype and Reveal Developmental Basis of ALX1-Related Frontonasal Dysplasia. Front Cell Dev Biol 10, 777887. Jacques-Fricke, B. T., Roffers-Agarwal, J. and Gammill, L. S. (2012). DNA methyltransferase 3b is dispensable for mouse neural crest development. PLoS One 7, e47794. Jeong, J., Mao, J., Tenzen, T., Kottmann, A. H. and McMahon, A. P. (2004). Hedgehog signaling in the neural crest cells regulates the patterning and growth of facial primordia. Genes Dev 18, 937-951. Kennedy, A. E. and Dickinson, A. J. (2012). Median facial clefts in Xenopus laevis: roles of retinoic acid signaling and homeobox genes. Dev Biol 365, 229-240. Kenny, C., Dilshat, R., Seberg, H. E., Van Otterloo, E., Bonde, G., Helverson, A., Franke, C. M., Steingrimsson, E. and Cornell, R. A. (2022). TFAP2 paralogs facilitate chromatin access for MITF at pigmentation and cell proliferation genes. PLoS Genet 18, e1010207. Kessler, S., Minoux, M., Joshi, O., Ben Zouari, Y., Ducret, S., Ross, F., Vilain, N., Salvi, A., Wolff, J., Kohler, H., et al. (2023). A multiple super-enhancer region establishes inter-TAD interactions and controls Hoxa function in cranial neural crest. Nat Commun 14, 3242. Khor, J. M. and Ettensohn, C. A. (2020). Transcription Factors of the Alx Family: Evolutionarily Conserved Regulators of Deuterostome Skeletogenesis. Front Genet 11, 569314. Knight, R. D., Javidan, Y., Nelson, S., Zhang, T. and Schilling, T. (2004). Skeletal and pigment cell defects in the lockjaw mutant reveal multiple roles for zebrafish tfap2a in neural crest development. Dev Dyn 229, 87-98. Knight, R. D., Javidan, Y., Zhang, T., Nelson, S. and Schilling, T. F. (2005). AP2-dependent signals from the ectoderm regulate craniofacial development in the zebrafish embryo. Development 132, 3127-3138. Knight, R. D., Nair, S., Nelson, S. S., Afshar, A., Javidan, Y., Geisler, R., Rauch, G. J. and Schilling, T. F. (2003). lockjaw encodes a zebrafish tfap2a required for early neural crest development. Development 130, 5755-5768. Kowalczyk, M. S., Tirosh, I., Heckl, D., Rao, T. N., Dixit, A., Haas, B. J., Schneider, R. K., Wagers, A. J., Ebert, B. L. and Regev, A. (2015). Single-cell RNA-seq reveals changes in cell cycle and differentiation programs upon aging of hematopoietic stem cells. Genome Res 25, 1860-1872. Kuleshov, M. V., Jones, M. R., Rouillard, A. D., Fernandez, N. F., Duan, Q., Wang, Z., Koplev, S., Jenkins, S. L., Jagodnik, K. M., Lachmann, A., et al. (2016). Enrichr: a comprehensive gene set enrichment analysis web server 2016 update. Nucleic Acids Res 44, W90-97. Lakhwani, S., Garcia-Sanz, P. and Vallejo, M. (2010). Alx3-deficient mice exhibit folic acid-resistant craniofacial midline and neural tube closure defects. Dev Biol 344, 869-880. Lamichhaney, S., Berglund, J., Almen, M. S., Maqbool, K., Grabherr, M., Martinez-Barrio, A., Promerova, M., Rubin, C. J., Wang, C., Zamani, N., et al. (2015). Evolution of Darwin's finches and their beaks revealed by genome sequencing. Nature 518, 371-375. Laugsch, M., Bartusel, M., Rehimi, R., Alirzayeva, H., Karaolidou, A., Crispatzu, G., Zentis, P., Nikolic, M., Bleckwehl, T., Kolovos, P., et al. (2019). Modeling the Pathological Long-Range Regulatory Effects of Human Structural Variation with Patient-Specific hiPSCs. Cell Stem Cell 24, 736-752 e712. Lawson, N. D. and Weinstein, B. M. (2002). In vivo imaging of embryonic vascular development using transgenic zebrafish. Dev Biol 248, 307-318. Li, H. and Durbin, R. (2009). Fast and accurate short read alignment with Burrows-Wheeler transform. Bioinformatics 25, 1754-1760. Li, H., Handsaker, B., Wysoker, A., Fennell, T., Ruan, J., Homer, N., Marth, G., Abecasis, G., Durbin, R. and Genome Project Data Processing, S. (2009). The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078-2079. Li, H., Sheridan, R. and Williams, T. (2013). Analysis of TFAP2A mutations in Branchio-Oculo-Facial Syndrome indicates functional complexity within the AP-2alpha DNA-binding domain. Hum Mol Genet 22, 3195-3206. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 1144 1146 1148 1150 1152 1154 1156 1158 1160 1162 1164 1166 1168 1170 1172 1174 1176 1178 1180 1182 1184 1186 1188 1190 1192 Li, W. and Cornell, R. A. (2007). Redundant activities of Tfap2a and Tfap2c are required for neural crest induction and development of other non-neural ectoderm derivatives in zebrafish embryos. Dev Biol 304, 338-354. Lim, K. C., Lakshmanan, G., Crawford, S. E., Gu, Y., Grosveld, F. and Engel, J. D. (2000). Gata3 loss leads to embryonic lethality due to noradrenaline deficiency of the sympathetic nervous system. Nat Genet 25, 209212. Liu, Z., Li, C., Xu, J., Lan, Y., Liu, H., Li, X., Maire, P., Wang, X. and Jiang, R. (2019). Crucial and Overlapping Roles of Six1 and Six2 in Craniofacial Development. J Dent Res 98, 572-579. Lohnes, D., Mark, M., Mendelsohn, C., Dolle, P., Dierich, A., Gorry, P., Gansmuller, A. and Chambon, P. (1994). Function of the retinoic acid receptors (RARs) during development (I). Craniofacial and skeletal abnormalities in RAR double mutants. Development 120, 2723-2748. Love, M. I., Huber, W. and Anders, S. (2014). Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol 15, 550. Luscher, B., Mitchell, P. J., Williams, T. and Tjian, R. (1989). Regulation of transcription factor AP-2 by the morphogen retinoic acid and by second messengers. Genes Dev 3, 1507-1517. Mai, C. T., Isenburg, J. L., Canfield, M. A., Meyer, R. E., Correa, A., Alverson, C. J., Lupo, P. J., Riehle-Colarusso, T., Cho, S. J., Aggarwal, D., et al. (2019). National population-based estimates for major birth defects, 20102014. Birth Defects Res 111, 1420-1435. Martik, M. L. and Bronner, M. E. (2021). Riding the crest to get a head: neural crest evolution in vertebrates. Nat Rev Neurosci. Martino, V. B., Sabljic, T., Deschamps, P., Green, R. M., Akula, M., Peacock, E., Ball, A., Williams, T. and WestMays, J. A. (2016). Conditional deletion of AP-2beta in mouse cranial neural crest results in anterior segment dysgenesis and early-onset glaucoma. Dis Model Mech 9, 849-861. Matsuoka, T., Ahlberg, P. E., Kessaris, N., Iannarelli, P., Dennehy, U., Richardson, W. D., McMahon, A. P. and Koentges, G. (2005). Neural crest origins of the neck and shoulder. Nature 436, 347-355. McGonnell, I. M., Graham, A., Richardson, J., Fish, J. L., Depew, M. J., Dee, C. T., Holland, P. W. and Takahashi, T. (2011). Evolution of the Alx homeobox gene family: parallel retention and independent loss of the vertebrate Alx3 gene. Evol Dev 13, 343-351. McLean, C. Y., Bristor, D., Hiller, M., Clarke, S. L., Schaar, B. T., Lowe, C. B., Wenger, A. M. and Bejerano, G. (2010). GREAT improves functional interpretation of cis-regulatory regions. Nat Biotechnol 28, 495-501. Meulemans, D. and Bronner-Fraser, M. (2002). Amphioxus and lamprey AP-2 genes: implications for neural crest evolution and migration patterns. Development 129, 4953-4962. Milunsky, J. M., Maher, T. A., Zhao, G., Roberts, A. E., Stalker, H. J., Zori, R. T., Burch, M. N., Clemens, M., Mulliken, J. B., Smith, R., et al. (2008). TFAP2A mutations result in branchio-oculo-facial syndrome. Am J Hum Genet 82, 1171-1177. Minoux, M., Holwerda, S., Vitobello, A., Kitazawa, T., Kohler, H., Stadler, M. B. and Rijli, F. M. (2017). Gene bivalency at Polycomb domains regulates cranial neural crest positional identity. Science 355. Mitchell, J. M., Sucharov, J., Pulvino, A. T., Brooks, E. P., Gillen, A. E. and Nichols, J. T. (2021). The alx3 gene shapes the zebrafish neurocranium by regulating frontonasal neural crest cell differentiation timing. Development 148. Muzumdar, M. D., Tasic, B., Miyamichi, K., Li, L. and Luo, L. (2007). A global double-fluorescent Cre reporter mouse. Genesis 45, 593-605. Naqvi, S., Hoskens, H., Wilke, F., Weinberg, S. M., Shaffer, J. R., Walsh, S., Shriver, M. D., Wysocka, J. and Claes, P. (2022). Decoding the Human Face: Progress and Challenges in Understanding the Genetics of Craniofacial Morphology. Annu Rev Genomics Hum Genet 23, 383-412. Neuhauss, S. C., Solnica-Krezel, L., Schier, A. F., Zwartkruis, F., Stemple, D. L., Malicki, J., Abdelilah, S., Stainier, D. Y. and Driever, W. (1996). Mutations affecting craniofacial development in zebrafish. Development 123, 357-367. Pertea, M., Kim, D., Pertea, G. M., Leek, J. T. and Salzberg, S. L. (2016). Transcript-level expression analysis of RNAseq experiments with HISAT, StringTie and Ballgown. Nat Protoc 11, 1650-1667. Pertea, M., Pertea, G. M., Antonescu, C. M., Chang, T. C., Mendell, J. T. and Salzberg, S. L. (2015). StringTie enables improved reconstruction of a transcriptome from RNA-seq reads. Nat Biotechnol 33, 290-295. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 1194 1196 1198 1200 1202 1204 1206 1208 1210 1212 1214 1216 1218 1220 1222 1224 1226 1228 1230 1232 1234 1236 1238 1240 1242 1244 Pimentel, H., Bray, N. L., Puente, S., Melsted, P. and Pachter, L. (2017). Differential analysis of RNA-seq incorporating quantification uncertainty. Nat Methods 14, 687-690. Pini, J., Kueper, J., Hu, Y. D., Kawasaki, K., Yeung, P., Tsimbal, C., Yoon, B., Carmichael, N., Maas, R. L., Cotney, J., et al. (2020). ALX1-related frontonasal dysplasia results from defective neural crest cell development and migration. EMBO Mol Med 12, e12013. Prescott, S. L., Srinivasan, R., Marchetto, M. C., Grishina, I., Narvaiza, I., Selleri, L., Gage, F. H., Swigut, T. and Wysocka, J. (2015). Enhancer divergence and cis-regulatory evolution in the human and chimp neural crest. Cell 163, 68-83. Qu, S., Tucker, S. C., Zhao, Q., deCrombrugghe, B. and Wisdom, R. (1999). Physical and genetic interactions between Alx4 and Cart1. Development 126, 359-369. Rada-Iglesias, A., Bajpai, R., Prescott, S., Brugmann, S. A., Swigut, T. and Wysocka, J. (2012). Epigenomic annotation of enhancers predicts transcriptional regulators of human neural crest. Cell Stem Cell 11, 633648. Ramirez, F., Ryan, D. P., Gruning, B., Bhardwaj, V., Kilpert, F., Richter, A. S., Heyne, S., Dundar, F. and Manke, T. (2016). deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res 44, W160-165. Robinson, J. T., Thorvaldsdottir, H., Winckler, W., Guttman, M., Lander, E. S., Getz, G. and Mesirov, J. P. (2011). Integrative genomics viewer. Nat Biotechnol 29, 24-26. Rothstein, M. and Simoes-Costa, M. (2020). Heterodimerization of TFAP2 pioneer factors drives epigenomic remodeling during neural crest specification. Genome Res 30, 35-48. Roybal, P. G., Wu, N. L., Sun, J., Ting, M. C., Schafer, C. A. and Maxson, R. E. (2010). Inactivation of Msx1 and Msx2 in neural crest reveals an unexpected role in suppressing heterotopic bone formation in the head. Dev Biol 343, 28-39. Santagati, F. and Rijli, F. M. (2003). Cranial neural crest and the building of the vertebrate head. Nat Rev Neurosci 4, 806-818. Satoda, M., Zhao, F., Diaz, G. A., Burn, J., Goodship, J., Davidson, H. R., Pierpont, M. E. and Gelb, B. D. (2000). Mutations in TFAP2B cause Char syndrome, a familial form of patent ductus arteriosus. Nat Genet 25, 4246. Scerbo, P. and Monsoro-Burq, A. H. (2020). The vertebrate-specific VENTX/NANOG gene empowers neural crest with ectomesenchyme potential. Sci Adv 6, eaaz1469. Schilling, T. F., Piotrowski, T., Grandel, H., Brand, M., Heisenberg, C. P., Jiang, Y. J., Beuchle, D., Hammerschmidt, M., Kane, D. A., Mullins, M. C., et al. (1996). Jaw and branchial arch mutants in zebrafish I: branchial arches. Development 123, 329-344. Schmidt, M., Huber, L., Majdazari, A., Schutz, G., Williams, T. and Rohrer, H. (2011). The transcription factors AP2beta and AP-2alpha are required for survival of sympathetic progenitors and differentiated sympathetic neurons. Dev Biol 355, 89-100. Schorle, H., Meier, P., Buchert, M., Jaenisch, R. and Mitchell, P. J. (1996). Transcription factor AP-2 essential for cranial closure and craniofacial development. Nature 381, 235-238. Seberg, H. E., Van Otterloo, E., Loftus, S. K., Liu, H., Bonde, G., Sompallae, R., Gildea, D. E., Santana, J. F., Manak, J. R., Pavan, W. J., et al. (2017). TFAP2 paralogs regulate melanocyte differentiation in parallel with MITF. PLoS Genet 13, e1006636. Sekiguchi, R. and Hauser, B. (2019). Preparation of Cells from Embryonic Organs for Single-Cell RNA Sequencing. Curr Protoc Cell Biol 83, e86. Selleri, L. and Rijli, F. M. (2023). Shaping faces: genetic and epigenetic control of craniofacial morphogenesis. Nat Rev Genet. Simoes-Costa, M. and Bronner, M. E. (2016). Reprogramming of avian neural crest axial identity and cell fate. Science 352, 1570-1573. Soldatov, R., Kaucka, M., Kastriti, M. E., Petersen, J., Chontorotzea, T., Englmaier, L., Akkuratova, N., Yang, Y., Haring, M., Dyachuk, V., et al. (2019). Spatiotemporal structure of cell fate decisions in murine neural crest. Science 364. Soriano, P. (1999). Generalized lacZ expression with the ROSA26 Cre reporter strain. Nat Genet 21, 70-71. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 1246 1248 1250 1252 1254 1256 1258 1260 1262 1264 1266 1268 1270 1272 1274 1276 1278 1280 1282 1284 1286 1288 1290 1292 1294 Square, T., Jandzik, D., Romasek, M., Cerny, R. and Medeiros, D. M. (2017). The origin and diversification of the developmental mechanisms that pattern the vertebrate head skeleton. Dev Biol 427, 219-229. Stenzel, A., Mumme-Monheit, A., Sucharov, J., Walker, M., Mitchell, J. M., Appel, B. and Nichols, J. T. (2022). Distinct and redundant roles for zebrafish her genes during mineralization and craniofacial patterning. Front Endocrinol (Lausanne) 13, 1033843. Sucharov, J., Ray, K., Brooks, E. P. and Nichols, J. T. (2019). Selective breeding modifies mef2ca mutant incomplete penetrance by tuning the opposing Notch pathway. PLoS Genet 15, e1008507. Tessier, P. (1976). Anatomical classification facial, cranio-facial and latero-facial clefts. J Maxillofac Surg 4, 69-92. Tolarova, M. M. and Cervenka, J. (1998). Classification and birth prevalence of orofacial clefts. Am J Med Genet 75, 126-137. van Dijk, D., Sharma, R., Nainys, J., Yim, K., Kathail, P., Carr, A. J., Burdziak, C., Moon, K. R., Chaffer, C. L., Pattabiraman, D., et al. (2018). Recovering Gene Interactions from Single-Cell Data Using Data Diffusion. Cell 174, 716-729 e727. Van Otterloo, E., Feng, W., Jones, K. L., Hynes, N. E., Clouthier, D. E., Niswander, L. and Williams, T. (2016). MEMO1 drives cranial endochondral ossification and palatogenesis. Dev Biol 415, 278-295. Van Otterloo, E., Li, H., Jones, K. L. and Williams, T. (2018). AP-2α and AP-2β cooperatively orchestrate homeobox gene expression during branchial arch patterning. Development 145. Van Otterloo, E., Li, W., Bonde, G., Day, K. M., Hsu, M. Y. and Cornell, R. A. (2010). Differentiation of zebrafish melanophores depends on transcription factors AP2 alpha and AP2 epsilon. PLoS Genet 6, e1001122. Van Otterloo, E., Li, W., Garnett, A., Cattell, M., Medeiros, D. M. and Cornell, R. A. (2012). Novel Tfap2-mediated control of soxE expression facilitated the evolutionary emergence of the neural crest. Development 139, 720-730. Van Otterloo, E., Milanda, I., Pike, H., Thompson, J. A., Li, H., Jones, K. L. and Williams, T. (2022). AP-2alpha and AP-2beta cooperatively function in the craniofacial surface ectoderm to regulate chromatin and gene expression dynamics during facial development. Elife 11. Vargel, I., Canter, H. I., Kucukguven, A., Aydin, A. and Ozgur, F. (2021). ALX-Related Frontonasal Dysplasias: Clinical Characteristics and Surgical Management. Cleft Palate Craniofac J, 10556656211019621. Walker, M. B. and Kimmel, C. B. (2007). A two-color acid-free cartilage and bone stain for zebrafish larvae. Biotech Histochem 82, 23-28. Wang, W. D., Melville, D. B., Montero-Balaguer, M., Hatzopoulos, A. K. and Knapik, E. W. (2011). Tfap2a and Foxd3 regulate early steps in the development of the neural crest progenitor population. Dev Biol 360, 173-185. Wickham, H. (2016). ggplot2 : Elegant Graphics for Data Analysis. In Use R!,, pp. 1 online resource (XVI, 260 pages 232 illustrations, 140 illustrations in color. Cham: Springer International Publishing : Imprint: Springer,. Williams, A. L. and Bohnsack, B. L. (2019). What's retinoic acid got to do with it? Retinoic acid regulation of the neural crest in craniofacial and ocular development. Genesis 57, e23308. Williams, T., Admon, A., Luscher, B. and Tjian, R. (1988). Cloning and expression of AP-2, a cell-type-specific transcription factor that activates inducible enhancer elements. Genes Dev 2, 1557-1569. Williams, T. and Tjian, R. (1991). Characterization of a dimerization motif in AP-2 and its function in heterologous DNA-binding proteins. Science 251, 1067-1071. Woodruff, E. D., Gutierrez, G. C., Van Otterloo, E., Williams, T. and Cohn, M. J. (2021). Anomalous incisor morphology indicates tissue-specific roles for Tfap2a and Tfap2b in tooth development. Dev Biol 472, 6774. Wu, Y., Kurosaka, H., Wang, Q., Inubushi, T., Nakatsugawa, K., Kikuchi, M., Ohara, H., Tsujimoto, T., Natsuyama, S., Shida, Y., et al. (2022). Retinoic Acid Deficiency Underlies the Etiology of Midfacial Defects. J Dent Res 101, 686-694. Xie, Z., Bailey, A., Kuleshov, M. V., Clarke, D. J. B., Evangelista, J. E., Jenkins, S. L., Lachmann, A., Wojciechowicz, M. L., Kropiwnicki, E., Jagodnik, K. M., et al. (2021). Gene Set Knowledge Discovery with Enrichr. Curr Protoc 1, e90. Xiong, Z., Dankova, G., Howe, L. J., Lee, M. K., Hysi, P. G., de Jong, M. A., Zhu, G., Adhikari, K., Li, D., Li, Y., et al. (2019). Novel genetic loci affecting facial shape variation in humans. Elife 8. bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. 1296 1298 1300 1302 1304 1306 1308 1310 1312 Xu, J., Liu, H., Lan, Y. and Jiang, R. (2022). The transcription factors Foxf1 and Foxf2 integrate the SHH, HGF and TGFbeta signaling pathways to drive tongue organogenesis. Development 149. Xu, P., Balczerski, B., Ciozda, A., Louie, K., Oralova, V., Huysseune, A. and Crump, J. G. (2018). Fox proteins are modular competency factors for facial cartilage and tooth specification. Development 145. Xu, P., Yu, H. V., Tseng, K. C., Flath, M., Fabian, P., Segil, N. and Crump, J. G. (2021). Foxc1 establishes enhancer accessibility for craniofacial cartilage differentiation. Elife 10. Yoon, B., Yeung, P., Santistevan, N., Bluhm, L. E., Kawasaki, K., Kueper, J., Dubielzig, R., VanOudenhove, J., Cotney, J., Liao, E. C., et al. (2022). Zebrafish models of alx-linked frontonasal dysplasia reveal a role for Alx1 and Alx3 in the anterior segment and vasculature of the developing eye. Biol Open 11. York, J. R. and McCauley, D. W. (2020). The origin and evolution of vertebrate neural crest cells. Open Biol 10, 190285. Zalc, A., Rattenbach, R., Aurade, F., Cadot, B. and Relaix, F. (2015). Pax3 and Pax7 play essential safeguard functions against environmental stress-induced birth defects. Dev Cell 33, 56-66. Zhang, J., Hagopian-Donaldson, S., Serbedzija, G., Elsemore, J., Plehn-Dujowich, D., McMahon, A. P., Flavell, R. A. and Williams, T. (1996). Neural tube, skeletal and body wall defects in mice lacking transcription factor AP2. Nature 381, 238-241. Zhao, Q., Behringer, R. R. and de Crombrugghe, B. (1996). Prenatal folic acid treatment suppresses acrania and meroanencephaly in mice mutant for the Cart1 homeobox gene. Nat Genet 13, 275-283. A B Tfap2a +/flox ; Tfap2b +/flox ; NCC-CRE Tg/+ Tfap2a +/null ; Tfap2b +/null ; NCC-CRE Tg/+ X X bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. Tfap2a flox/flox ; Tfap2b flox/flox Tfap2a flox/flox ; Tfap2b flox/flox C Nomenclature in this Study Tfap2a CTRL Null/ Conditional any combination Tfap2Δ/HET + / flox + / flox Tfap2aΔ/NCKO null / flox + / flox Tfap2bΔ/NCKO + / flox null / flox Tfap2Δ/NCKO null / flox null / flox CTRL Conditional/ Conditional any combination Tfap2HET + / flox + / flox Tfap2aNCKO flox / flox + / flox Tfap2bNCKO + / flox flox / flox Tfap2NCKO flox / flox flox / flox D CTRL 1 mm E18.5 F Tfap2b Tfap2bNCKO NCC-CRE negative positive Previous Nomenclature (Van Otterloo et al., 2018) CTRL SCM-A SCM-B DCM negative positive not applicable D’ E Tfap2aNCKO E’ F’ G Tfap2NCKO G’ * TFAP2A DAPI TFAP2B DAPI ne ne np np mFNP mFNP d B -log10 (Q-value) A FNP-enriched genes Not enriched MdP-enriched genes bioRxiv preprint version posted June 16, 2023. The copyright holder for this preprint lFNP doi: https://doi.org/10.1101/2023.06.16.545376; thislFNP (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. ect ect E11.5 v 100 µm Wnt1:CRECTRL C Wnt1:CRETfap2NCKO log2FC D Sox10:CRECTRL Sox10:CRETfap2NCKO F Wnt1:CRECTRL Wnt1:CRETfap2NCKO d v lFNP mFNP 500 µm E np E12.5 np mFNP lFNP 500 µm G E12.5 Sox10:CRECTRL Sox10:CRETfap2NCKO CTRL A na e Tfap2aNCKO B fr fr pr Sox10:CRE p Tfap2NCKO D * na na fr pr cs a Tfap2bNCKO C pr cs pr fr pmx cs cs E18.5 B’ A’ ppso ppro vm pppm vm pppm ppp bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. ns vm pppm ppso? pppm ppp pl pmx ppp vm ps ns ns ns nl nl D’ ppso ppro ppp pmx pl pmx C’ ppso ppro pl nl pl pmx nl as a p CTRL E Tfap2aNCKO F na na fr Wnt1:CRE * pr fr cs Tfap2NCKO H H na pr fr pr cs a Tfap2bNCKO G cs pr pmx fr cs E18.5 p F’ E’ E’ ppso ppro vm G’ G’ ppso ppro ppp vm vm ns ns ps pl ppp vm ns pmx nl pppm ns pmx nl ppp pppm pppm ppso ppro ppp pppm pmx H’ ppso ppro pl pl pl nl pmx a p nl A FNP Up-regulated Down-regulated bulk RNA-seq log2FC DisGeNET 3x CTRL vs 3x Tfap2Δ/NCKO Human Phenotype Ontology MdP -log10(Q-value) MxP Jensen DISEASES WikiPathway 2021 Human B -log10(p-value) BA2: Dlx1/2/3/4/5/6 Hoxa2 Emx2 bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. GFP-positive CNCCs Glia: Sox10 Erbb3 Plp1 Isl1 10X Chromium single-cell RNA-seq UMAP_2 Tfap2HET + Tfap2NCKO Littermates MdP: Dlx1/2/3/4/5/6 Foxf1/2 Hand2 Mesenchyme: Prrx1/2 Twist1 Fli1 MxP MdP sort midface CNCCs Tfap2HET Tfap2NCKO Relative mRNA expression (vs. B2m housekeeping gene) FNP Alx3 Alx4 FNP-2: Alx1/3/4 Pax3/7 Tfap2HET Tfap2NCKO F Alx3 UMAP_1 G FNP-1: Alx1/3/4 Six2 UMAP_1 UMAP_2 Autonomic & Sensory Neurons + Brain: Neurog2 Stmn2 Sox2 Otx2 MxP: Lhx6/8 Six1 Alx1 Gene Expression Levels Cut just below BA2 & otic vesicle E D FACS using Wnt1:CRE-mTmG CTRL Tfap2aΔ/NCKO Tfap2Δ/NCKO CTRL Tfap2Δ/NCKO E10.5 E10.5 Relative mRNA expression (vs. Actb housekeeping gene) C Alx4 CTRL Tfap2Δ/NCKO E10.5 E10.5 A B C D bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license. F FNP MdP H G E bioRxiv preprint doi: https://doi.org/10.1101/2023.06.16.545376; this version posted June 16, 2023. The copyright holder for this preprint (which was not certified by peer review) is the author/funder, who has granted bioRxiv a license to display the preprint in perpetuity. It is made available under aCC-BY-NC-ND 4.0 International license.