Cell fate determination is a necessary and tightly regulated process for producing different cell types and structures during development. Cranial neural crest cells (CNCCs) are unique to vertebrate embryos and emerge from the neural plate borders into multiple cell lineages that differentiate into bone, cartilage, neurons and glial cells. We have previously reported that Irf6 genetically interacts with Twist1 during CNCC-derived tissue formation. Here, we have investigated the mechanistic role of Twist1 and Irf6 at early stages of craniofacial development. Our data indicate that TWIST1 is expressed in endocytic vesicles at the apical surface and interacts with β/δ-catenins during neural tube closure, and Irf6 is involved in defining neural fold borders by restricting AP2α expression. Twist1 suppresses Irf6 and other epithelial genes in CNCCs during the epithelial-to-mesenchymal transition (EMT) process and cell migration. Conversely, a loss of Twist1 leads to a sustained expression of epithelial and cell adhesion markers in migratory CNCCs. Disruption of TWIST1 phosphorylation in vivo leads to epidermal blebbing, edema, neural tube defects and CNCC-derived structural abnormalities. Altogether, this study describes a previously uncharacterized function of mammalian Twist1 and Irf6 in the neural tube and CNCCs, and provides new target genes for Twist1 that are involved in cytoskeletal remodeling.

Craniofacial morphogenesis requires precise interaction between multiple genes and gene environments that are essential for cell fate induction, specification and differentiation (Chai and Maxson, 2006; Nomura and Li, 1998; Padmanabhan and Ahmed, 1997). Disruption of the precisely organized morphogenesis leads to craniofacial disorders, which are the second most common congenital birth defects in humans (Hoyert et al., 2006; Joshi et al., 2014; Murray et al., 2007; Parker et al., 2010). During embryogenesis, signaling molecules from the notochord and mesoderm induce the formation of neuroectoderm. The junctions between non-neural and neural ectoderm are called the neural plate borders, which form distinct domains through inductive interactions with the adjacent cells and the underlying mesoderm (Groves and LaBonne, 2014; Litingtung and Chiang, 2000; Millonig et al., 2000). The CNCCs emerge from the neural fold borders of the midbrain and hindbrain regions, and migrate as unipotent and multipotent mesenchymal cells towards the frontonasal processes (Achilleos and Trainor, 2012; Zhang et al., 2021). The CNCCs give rise to the bone, cartilage, neuron and glial cells of the orofacial skeletal and peripheral nervous systems (Groves and LaBonne, 2014; He et al., 2014; Le Douarin et al., 2004; Noden and Trainor, 2005).

Twist1 encodes a transcription factor that belongs to the basic helix-loop-helix (bHLH) B-family. It was first identified in Drosophila melanogaster where its deletion resulted in disruption of mesodermal specification and ventral furrow formation (Lu et al., 2011; Nüsslein-Volhard et al., 1984; Simpson, 1983). In vertebrates, Twist1 is expressed first at the gastrulation stage in the primitive streak and then in the adjacent mesodermal cells (Füchtbauer, 1995; Stoetzel et al., 1995). At the neurulation stage, mammalian Twist1 is detected in the paraxial and lateral mesoderm, and the migratory CNCCs (Bildsoe et al., 2016; Füchtbauer, 1995; Ota et al., 2004; Soo et al., 2002). Twist1 expression is maintained in the CNCC-derived mesenchyme of the frontonasal and pharyngeal processes (Bildsoe et al., 2009; Füchtbauer, 1995). In humans, mutations in TWIST1 cause Saethre-Chotzen syndrome, which is characterized by craniosynostosis and cleft palate, Sweeney-Cox syndrome and Robinow-Sorauf syndrome (Kunz et al., 1999; Takenouchi et al., 2018; Seto et al., 2007). In mice, Twist1 is crucial for neural tube closure and for CNCC-derived craniofacial bone and cartilage formation (Chen et al., 2007; Bildsoe et al., 2013). TWIST1 has been shown to promote cell survival and proliferation of migratory CNCCs during craniofacial development (Bildsoe et al., 2013). Although both processes involve gaining potency and motility, the in vivo molecular function of TWIST1 as a master regulator of epithelial-to-mesenchymal transition (EMT) in CNCCs was thus far only described in Xenopus (Lander et al., 2013).

TWIST1 is a phosphoprotein with multiple residues that are phosphorylated by several kinases according to numerous cancer studies (Lu et al., 2011; Vichalkovski et al., 2010; Xue and Hemmings, 2012). TWIST1 phosphorylation is crucial for regulating its homo- and heterodimerization with other factors to control multiple cellular activities (Firulli and Conway, 2008; Hong et al., 2011; Lu et al., 2011; Vichalkovski et al., 2010). TWIST1 interacts with other proteins, such as TCF3 and chromatin remodeling factors, to guide CNCC migration during craniofacial development (Firulli et al., 2005; Fan et al., 2021). Misregulation of TWIST1 dimerization with HAND2 is caused by mutations of phosphoresidues in the PKA consensus domain or in the bHLH domain, which have been reported to be associated with Saethre-Chotzen syndrome (Firulli et al., 2005; El Ghouzzi et al., 1997). Mutations within the TWIST-box domain are also associated with the isolated form of craniosynostosis (Seto et al., 2007). TWIST1 is also expressed in cardiomyocytes and is repressed in patients with dilated cardiomyopathy. Knockdown of Twist1 in primary rat cardiomyocytes that are stimulated using phenylephrine/isoprenaline causes significant reduction in cell size (Baumgarten et al., 2013). Overexpression of Twist1T125D/S127D phosphomimetic alleles (equivalent to T121/S123 in human) in cardiomyocytes resulted in disrupted cell remodeling and heart formation of transient transgenic mouse embryos (Lu et al., 2011).

Our previous work demonstrated that the compound heterozygous mice (Irf6+/−; Twist1+/−) have craniofacial abnormalities, i.e. mandibular agnathia, fused maxilla, cleft palate and holoprosencephaly (Fakhouri et al., 2017). Unlike other IRF family members, interferon regulatory factor 6 (Irf6) is a broadly expressed transcription factor in the ectoderm and oral epithelium during embryogenesis. In mice, Irf6 promotes cell differentiation of proliferative epithelial cells, but its function in the neural tube and neural crest cell development is still unknown (Ingraham et al., 2006; Kousa et al., 2019; Richardson et al., 2006). We have recently shown that overexpression of Irf6 in ectoderm by K14-enhancer leads to exencephaly and failure of peritoneal skin development. Furthermore, the compound heterozygous mice for Irf6 and Tfap2a do not show neural tube defects compared to single Tfap2a heterozygous mice (Kousa et al., 2019).

In the current study, we sought to characterize the spatiotemporal expression of Twist1 and Irf6 during neural tube formation and their function in CNCCs. We generated multiple stable Twist1 phospho-incompetent mouse lines to identify the importance of TWIST1 phosphorylation in regulating neural tube and CNCC-derived craniofacial structures. We investigated the molecular role of Irf6 in regulating the integrity of the neural tube and its relation to AP2α expression. Our data show that TWIST1 protein and mRNA are detected at the apical side of dorsal neuroectodermal cells within expression locations similar to β-catenin and tight junction proteins. Further investigation suggests that TWIST1 is expressed in endocytic compartments at the apical surface and partially overlaps with endocytic markers LRP2 and RAB11b. Meanwhile, cytosolic TWIST1 interacts with β/δ-catenin during neural tube formation. We show that Twist1 is crucial to promote cell fate transition of pre-EMT neuroectodermal cells, and loss of Twist1 in neuroectoderm causes ectopic expression of IRF6 and E-cadherin in delaminated CNCCs. We demonstrate that three highly conserved phospho-residues in TWIST1 are crucial for its in vivo function in the neural tube and in CNCC-derived structures. Finally, our data shows that TWIST1 regulates Specc1l expression during CNCC development to possibly control cytoskeletal remodeling in CNCCs.

Spatiotemporal expression of TWIST1 and IRF6 during neural tube development

To determine the role of Twist1 and Irf6 in neural tube formation, we investigated their spatiotemporal expression in wild-type embryos. We performed immunofluorescent staining and in situ hybridization in embryos from E8.5 to E9.5. At E8.5, IRF6 was highly expressed in the basal and apical sides of neural plate, with low expression seen in the middle and weak IRF6 expression detected in some cells beneath the neural plate (Fig. 1A,A′). TWIST1 was unexpectedly detected in vesicle-like structures at the apical surface of neural plate (Fig. 1A,A′, white arrow; Fig. S1A-B). Nuclear TWIST1 was weakly expressed in a few cells at the neural plate borders (Fig. 1B,B′, white arrow). Robust nuclear expression of TWIST1 was observed in the mesodermal cells adjacent to the neural plate (Fig. 1A,B, green arrow). At E9.0, most dorsal and ventral cells in neural folds expressed IRF6, while TWIST1 was mostly expressed at the apical side and dorsal edges of neural folds. TWIST1 expression overlapped with IRF6 at the dorsal edges of the neural folds (Fig. 1E,E′, white arrow; Fig. S1C). To understand the role of TWIST1 apical expression, we investigated whether its expression was associated with adherens and tight junction proteins. TWIST1 apical expression in neuroectodermal cells had similar expression patterns to β-catenin (Fig. 1C,C′; Fig. S1C,D) and claudin 1 at E8.5 (Fig. 1D,D′) and E9.0 (Fig. 1F-G′; Fig. S1E). SOX9 was used as a marker of pre-EMT and migratory CNCCs (Fig. 1B-D, red arrow). TWIST1 apical expression was also observed at the dorsal edges of the neural folds immediately before fusion (Fig. 1H-H′, arrowhead). We tested the specificity of the anti-TWIST1 antibodies in Twist1-null embryos. Histological staining showed the neural tube closure defects in Twist1-null compared with wild-type embryos at E10.5 (Fig. S2A-B′). No signal for TWIST1 protein was detected in the neural tube or mesenchymal cells of a Twist1-null embryo compared with a wild type (Fig. S2C-F′). Notably, Twist1 null embryos had complete neural tube closure defects (Fig. 1J, white arrow) compared with wild-type littermates at E11.5 (Fig. 1I, white arrow). In Twist1 null embryos, the neural plate invaginated and lateral neural folds elevated towards the dorsal midline but the lateral folds failed to close (Fig. 1J, white arrow). To validate the immunofluorescence data, whole-mount in situ hybridization for Twist1 mRNA showed weak expression at the dorsal edges of the neural plate of E8.5 embryos (Fig. 1K-K′; Figs S3 and S4) and of neural folds of E9.5 embryos (Fig. 1L-M′; Fig. S5). Based on a previous publication reporting the expression of LRP2 and RAB11b, a marker of endocytic cellular membrane vesicles, at the apical surface (Kowalczyk et al., 2021), we investigated whether TWIST1 was expressed in endocytic vesicles of RAB11b compartments. The dual immunofluorescent staining data shows that TWIST1 was expressed in endocytic vesicles at the apical surface of neural folds (Fig. 1N-O′). TWIST1 expression partially overlapped with LRP2 at the apical side of neural tube (Fig. 1N,N′, inlet in N′; Fig. S1A,A′) and with RAB11b in endocytic compartments (Fig. 1O,O′, inlet in N′; Fig. S1B,B′). Rab11b mRNA was significantly reduced in hindbrain tissues of Twist1 conditional knockout (CKO), while no change was detected for Lrp2 expression (Fig. S6C,D).

To further determine the significance of TWIST1 apical expression, we performed co-immunoprecipitation (co-IP) for TWIST1 to identify the protein interactors at E8.5-E9.0. The in vivo cellular fractionation co-IP blots show that the β-catenin and δ-catenin in the cytosolic fraction were pulled down with TWIST1 immunoprecipitation using anti-TWIST1 antibodies (Fig. 2A,B). In contrast, nuclear TWIST1 weakly interacted with β-catenin (Fig. 2A). We also performed dual immunostaining for TWIST1 and β-catenin using rabbit anti-β-catenin antibodies. β-Catenin was detected at the apical-lateral side of all the dorsal neuroectodermal cells (Fig. 2C,C′; Fig. S1F,G). Likewise, TWIST1 and δ-catenin were detected at the apical side of neural folds in a similar pattern of consecutive sections (Fig. 1D,D′; Fig. S1H-I′). Fig. 2E shows neural folds and a dotted-line square marks the areas indicated in Fig. 1F,G. β-Catenin apical expression detected in wild type was diffused in the Twist1cko/− embryo and became mostly cytosolic when detected using mouse anti-β-catenin antibodies (Fig. 2F-G′). To further determine the impact of Twist1 loss in neural tube formation, Hematoxylin and Eosin staining was performed in wild-type and Twist1cko/− embryonic tissues. Normal neural tube formation was observed in wild type (Fig. 2H,J), while Twist1cko/− embryos had multiple dorsolateral bend points and expansion of the neural tube (Fig. 2I,K). Histological staining of older wild-type embryos showed a normal development of the cephalic flexure of the midbrain and hindbrain at E12.5 (Fig. 2L,L′; Fig. S6F-F″), while Twist1cko/− showed abnormal patterning of the cephalic flexure of midbrain and hindbrain at E12.5 (Fig. 2M,M′, black arrows; Fig. S6H-H″).

Tracing CNCC migration in wild type and Twist1cko/− using ROSA26Tm1; Wnt1-Cre and Wnt1-Cre2 mouse lines

To determine the role of Twist1 in CNCCs, we used Wnt1-Cre and Wnt1-Cre2 to conditionally knockout (CKO) Twist1 in pre-EMT CNCCs. For cell tracing, we used R26Tm1 reporter gene and X-gal whole-mount staining. At E9.5, the blue staining of migratory CNCCs was detected in the cephalic flexure of the midbrain, frontonasal process and pharyngeal arches of wild-type embryos (Fig. 3A), and similar expression pattern was detected in wild type at E11.5 but with stronger staining in the rhombic lips (Fig. 3C). At E15.5, migratory CNCCs were observed in the otic pinna and orofacial regions (Fig. 3E). In mutant Twist1cko/− embryos, weaker blue staining was observed in the frontonasal and pharyngeal processes, while stronger staining was detected in the mid- and hindbrain regions at E9.5, E11.5 and E15.5, suggesting disruption of migratory CNCCs towards frontal processes (Fig. 3B,D,F). We noticed that the first pharyngeal arch of Twist1cko/− embryos was smaller at E11.5, and the embryos displayed severe brain abnormalities at E15.5 (Fig. 3D,F). The recently generated Wnt1-Cre2 mouse line has been used to validate the CNCC-specific deletion by the Wnt1-Cre line that has a midbrain developmental abnormality (Lewis et al., 2013). The Wnt1-Cre2 line showed a similar staining pattern of migratory CNCCs to Wnt1-Cre (Fig. 3G,I,K) and the craniofacial phenotype associated with Twist1cko/− (Fig. 3H,J,L). At E12.5 and E14.5, a dorsal image from the whole-mount staining of wild-type embryos showed blue staining in the hindbrain and in the trunk neural tube (Fig. S6E,E′; Fig. 3M, black arrow). Similarly, blue staining was observed in Twist1cko/− embryos, showing expanded staining in the midbrain and hindbrain compared with the wild type (Fig. S6G,G′; Fig. 3N, black arrow). The embryos with compound alleles of Twist1cko/− and Irf6+/− presented with severe hemorrhaging in the frontonasal process and craniofacial abnormalities compared with wild type (Fig. 3O,P; Fig. S6I-K). Stereomicroscope images of a wild-type embryo showed normal craniofacial development at E17.5 (Fig. 3Q), while a Twist1cko/− embryo showed severe exencephaly and abnormal frontonasal processes (Fig. 3R). Twist1cko/− embryos lacked cranial, frontonasal and maxillary bones, and had severe mandibular hypoplasia, as depicted by the skeletal staining (Fig. 3T), compared with wild-type littermate (Fig. 3S). Histological staining of a wild-type embryo showed normal development of the neural tube apical and basal layers of the hindbrain and midbrain at E10.5, respectively (Fig. 3U,W), while Twist1cko/− showed an expansion of the neural tube and partially detached cells at the basal side, and abnormal patterning of the cephalic flexure of the midbrain (Fig. 3V,X).

CNCC delamination and migration in WT and Twist1cko/− neural tube explants

We further investigated why a loss of Twist1 in pre-EMT neuroectodermal cells led to severely abnormal brain and craniofacial structures, even though a significant number of CNCCs migrated towards the frontonasal and pharyngeal processes. We also asked why the mutant migratory CNCCs did not form craniofacial bone and cartilage (Fig. 3Jj). To answer these questions, we developed a neural tube explant system using a dual reporter transgenic mouse line (ROSA26tm4.ACTB-tdTomato-EGFP), which allows us to capture cell morphogenesis and migration of somatic and CNCCs. The illustrative diagram depicts the procedure for dissecting the neural folds from embryos at E8.5 (Fig. 4A). We used this system to uncover the function of Twist1 in pre-EMT CNCCs, wherein lack of its expression in the neuroectodermal and pre-EMT neural crest cells can explain the phenotype observed in Twist1cko/−. The cells expressing the Wnt1-Cre transgene excise the membrane-associated mTomato reporter gene, allowing instead the expression of GFP. In wild-type explants, individual migratory CNCCs moved away from the dorsolateral sides of the neural folds (Fig. 4B,B′) and the migratory mesenchymal cells formed long protrusions (inset in Fig. 4B′). In the Twist1cko/− explant, a considerable amount of Twist1-deficient neuroectodermal cells remained within the neural folds and expanded ventrally (Fig. 4C), and the majority of the detached CNCCs did not transition into mesenchymal cells but instead maintained their cell-cell adhesion and epithelial morphology (Fig. 4C′, inset in C′). Quantitative measurements of the total number of migratory CNCCs were analyzed for each genotype. We detected approximately a 50% reduction in the total number of cells in mutant explants (Fig. 4D), and the average distance of all migratory CNCCs was reduced by 68% (Fig. 4E). We also measured the mean speed of individual CNCCs in wild type and Twist1cko/−. To determine the mean speed of individual CNCCs, we tracked the CNCCs as they moved away from the neural tube in wild type and Twist1cko/− (Fig. 4F,F′,G,G′). Micrographs show individual CNCC migration paths during time-lapse imaging (Fig. 4F′,G′, white lines). Wild-type CNCCs mean speed increased by 50% in comparison with Twist1cko/− (Fig. 4H).

Neural tube phenotype in Irf6 null embryos

IRF6 is expressed in the non-neural and neural ectoderm during neural tube formation. Its expression is relatively stronger in neural ectoderm than non-neural ectoderm (Fig. 1E,E′; Fig. S1C-E). We hypothesized that a complete loss of Irf6 affects neural tube formation, leading to craniofacial abnormalities. At E10.5, we noticed that the cellular organization of the edges of neural tube and apical membrane were disorganized in Irf6 null compared with wild-type littermates (Fig. 5A-B″, black arrow). Abnormal expansions from the brain cortex were also detected in Irf6-null embryos compared with wild type (Fig. 5C-D′). We checked the expression of SOX9, which is expressed in the neuroepithelial progenitors of glia (Farrell et al., 2011; Vogel and Wegner, 2021), and F-actin in wild-type and Irf6-null embryos. SOX9-positive cells were detected in the dorsal half of the neural tube (Fig. 5E,E′, white arrow). However, we found that the organization of SOX9-positive cells was altered and large intercellular gaps among SOX9 positive cells were detected in Irf6-null neural tube (Fig. 5F,F′, white arrows). A continuous expression of F-actin was noticed in the apical cells of the dorsal edge of wild-type neural tube (Fig. 5E″), while the apical membrane was discontinuous in Irf6-null embryos (Fig. 5F″). We also investigated the expression of non-neural ectodermal marker, AP2α, to uncover whether a loss of Irf6 alters its expression at the neural tube junction. We noticed that the number of AP2α-positive cells at the neural tube dorsal edges increased in Irf6-null embryos compared with wild-type littermates in four different biological replicates of each genotype (Fig. 5G,H). We quantified the number of AP2α-positive cells in the dorsal edges and observed a significant increase in Irf6 null compared with wild type (Fig. 5I).

Expression pattern and level of epithelial genes and SOX9 in neural tube and migratory mesenchymal cells

Based on the CNCC-derived tissue phenotypes observed in Twist1cko/−, we investigated the expression pattern and level of SOX9 and epithelial genes involved in cell adhesion and differentiation to determine why a loss of Twist1 in delaminated CNCCs maintained their epithelial morphology. In wild-type embryos, IRF6 was mainly detected in the neural tube, whereas TWIST1 was highly expressed in migratory mesenchymal cells that were detached from the neural tube and migrated towards frontonasal and pharyngeal prominences at E9.5-E10.5 (Fig. 6A,A′; Fig. S7A-C). Ectopic IRF6 expression in partially detached mesenchymal cells was detected in a cluster of cells along the neural tube in Twist1cko/− embryos (Fig. 6B-B′, white arrow). No expression of E-cadherin (E-CAD) was present in mesenchymal cells in wild type (Fig. 6C,C′). However, staining for E-CAD in Twist1cko/− embryos showed that the partially detached mesenchymal cells express ectopic E-CAD (Fig. 6D,D′). SOX9 expression was used to mark migratory CNCCs along the neural tube and to distinguish CNCCs from mesodermal cells. In a wild-type embryo, an abundant number of migratory CNCCs showed co-expression of SOX9 and TWIST1 (Fig. 6E,E′ white arrows), whereas only a few TWIST1-positive CNCCs can be seen in the Twist1cko/− embryo (Fig. 6F-F′, white arrow). We also looked at the expression of tight junction proteins occludin and ZO1. Occludin is expressed at high levels at the dorsal junction of neural tube edges of wild-type embryos (Fig. 6G,G′), whereas almost no expression of was detected at the unfused junction of the dorsal edges of Twist1cko/− neural tube (Fig. 6H,H′). In neural tube explants, an interspaced and broken signal of ZO1 was seen among the migratory mesenchymal cells in wild type (Fig. 6I,I′), whereas a continuous junctional signal of ZO1 was observed among detached mesenchymal cells from the neural tube of Twist1cko/− (Fig. 6J,J′). We also looked at the expression of a mesenchymal marker, vimentin, and observed a disrupted expression pattern in the mesenchymal cells in Twist1cko/− embryos and neural tube explants compared with wild type (Fig. S7D-G′). As expected, Twist1 expression was significantly reduced in Twist1cko/− embryos, and a slight increase was seen in Irf6 mRNA in Twist1cko/− (Fig. 6K). mRNA encoding E-Cad (Cdh1) N-Cad (Cdh2) and β-catenin (Ctnnb1) were also measured in WT and Twist1cko/− (Fig. 6L). E-Cad expression was significantly increased in Twist1cko/− (Fig. 6L), which is consistent with the ectopic expression of E-CAD in partially detached mesenchymal cells (Fig. 6D,D′).

Quantitative levels of kinases, signaling molecules, and epithelial and mesenchymal marker genes

The expression of genes encoding kinases [Akt1, Akt2, members of the Pi3k family, Erk1 (Mapk3), members of the Ck2 family and Src]; epithelial markers [E-Cad, N-Cad and occludin (Ocln)]; neural and non-neural ectoderm markers, and CNCC specification drivers [Snail2 (Snai2), Msx1, Tfap2a, Hand2, Sox2, Sox10 and Hoxd10]; signaling molecules [Yap (Yap1), Wnt1, Wnt3, Jag1 and Mtor]; and actomyosin remodeling regulators (β-catenin, Rhoa, Rhoc and Arhgap29) was measured in dissected tissues of the hindbrain and first pharyngeal arch (Fig. S8). The mRNA expression of Akt2, Snai2, Msx1, Hand2, Sox10, Yap, β-catenin, Rhoa and Arhgap29 was significantly reduced in Twist1cko/− embryos at E10.5 (Fig. S8A,E,I,G,K). Protein analysis by western blot also showed decreased levels of AKT1, AKT2, AKT3, SNAI2 and WNT3 in Twist1cko/− embryos (Fig. S8B,F,H). Further details are explained in the legend of Fig. S8.

TWIST1 nuclear expression and phosphorylation in vivo and in the O9-1 cell line

TWIST1 has six phospho-residues that are highly conserved in vertebrates (Fig. S9A). Our data show that some subpopulations in the dorsal edges of neural folds and in the pre-EMT CNCCs express TWIST1 mostly in the cytosol (Fig. S9B-C′, arrows; Fig. S1A′,D,D′), and it becomes mainly nuclear in detached and migratory CNCCs at E9.5 and E10.5 (Fig. 6A,A′, arrow; Fig. S9D,D′; Fig. S7C). Our in vivo data show that phosphorylated TWIST1 protein can be detected in mesenchymal cells using monoclonal antibodies raised against TWIST1 phospho-S68, and polyclonal antibodies raised against TWIST1 phospho-S42 and phospho-S123 residues at E10.5 (Fig. S9E). TWIST1 unphosphorylated and phosphorylated forms were also detected in the O9-1 cell line (Fig. S9F). Further details are provided in the legend of Fig. S9.

Importance of TWIST1 phosphorylation in craniofacial tissues

To determine the importance of TWIST1 post-translational modifications in the neural tube and CNCC-derived tissues, we generated four phospho-incompetent mouse lines, including Twist1S18;20A, Twist1S42A, Twist1S68A and Twist1T121A; S123A using CRISPR/CAS9 technology. The founders of Twist1S42A mice did not give any progeny, and all ten Twist1T121A; S123A founders died before reaching 2 months of maturity. We confirmed the genomic changes in three founders of Twist1S18I;20A/+ and in four founders of Twist1S68A/+. The sequence of F6 generations confirmed the changes of serine 18 (S18) to isoleucine and serine 20 (S20) to alanine, and in the second mouse line, serine 68 (S68) to alanine (Fig. 7A,B).

The phenotypic characterization showed that the two phosphorylation sites, S18/20 and S68, are crucial for TWIST1 activity in craniofacial development. The two phospho-incompetent mouse lines have epidermal blebbing, severe edema along the neural tube and significant neural tube defects in Twist1S18I;20A/S18I;20A compared with wild-type littermates (Fig. 7C-F′). The skeletal staining exhibited a remarkable loss of maxillary and mandibular bones in Twist1S18I;20A/S18I;20A and a reduction in skull mineralization in Twist1S68A/S68A, respectively (Fig. 7G′,H′), compared with wild type (Fig. 7G,H). The histological staining of Twist1S68A/S68A head structures at E15 showed subepidermal blebbing, separation of fibroblast layers from the meninges and a detachment of the meninges from the surface of the brain cortex (Fig. 7J,J′,L,L) compared with wild-type littermates (Fig. 7I,I′,K,K′). We also noticed lymphocyte infiltration in Twist1S18I;20A/S18I;20A brain tissues and around the frontonasal processes compared with wild-type embryos (Fig. 7M,N). To determine the types of the infiltrated cells, immunofluorescence staining for the lymphocyte marker CD73 and B cell marker CD206 was used on tissues sections at E13.5. No signal was detected in the frontonasal sections of wild type (Fig. 7O,Q); however, enrichment of both markers was noticed in the frontonasal tissues of Twist1S18I;20A/S18I;20A embryos (Fig. 7P,R). The genotype and spectrum of craniofacial phenotype in the homozygous phospho-incompetent embryos are further described in Fig. S10.

Interaction between Twist1 and Specc1l in CNCCs

Many of the tested cytoskeletal remodeling regulators were significantly reduced at the mRNA level in Twist1cko/− embryos. We focused on Specc1l because of the phenotypic overlap in craniofacial disorders in humans and mice, and the importance of this gene in regulating the actomyosin cytoskeleton. We tested the hypothesis that TWIST1 regulates Specc1l expression in CNCCs and that both factors interact with adherens junction proteins. Specc1l has 17 exons plus 5′ and 3′ UTR (Fig. 8A). We mapped two putative regulatory elements in intron 1 and 2 based on epigenetic signatures (Fig. S11). We performed ChIP-PCR to determine whether TWIST1 in vivo binds to E1 and E2 putative elements. The results showed enrichment of TWIST1 protein at the putative Specc1l element E1 for both embryonic time points E9.5 and E10.5. The bound TWIST1 to E2 putative enhancer was increased at E10.5 as indicated by the PCR band intensity but not at E9.5 or E11.5 (Fig. 8B). mRNA expression of Specc1l was significantly reduced in tissues extracted from the hindbrain and first pharyngeal arch of Twist1cko/− compared with wild type (Fig. 8C). We matched the phenotype in Twist1cko/− and phospho-incompetent embryos with Specc1l loss-of-function mutant mouse embryos to compare the phenotypic overlap in craniofacial tissues and CNCCs. Twist1cko/− and phospho-incompetent mutant embryos showed brain hemorrhage, subepidermal blebbing and severe edema along the neural tube (Fig. 8D,E). Similarly, the Specc1lcGT/DC510 compound mutant alleles (described by Hall et al., 2020) show brain hemorrhage, subepidermal blebbing and severe edema along the neural tube compared with wild-type littermate embryos (Fig. 8F). Loss of Twist1 in CNCCs led to disruption in the EMT process and the persistence of cell-cell adhesion in migratory mesenchymal cells (Fig. 8H), and the in vivo accumulation of adherens junction proteins β-catenin and E-CAD in partially detached cells (Fig. 8J,L; Fig. S12B-B‴ and D-D‴) compared with wild type (Fig. 8G,I,K; Fig. S12A-A‴,C-C‴). Similarly, loss of Specc1l function caused the accumulation of membrane-associated β-catenin in migratory CNCCs (Fig. 8N-N″) compared with wild-type CNCCs (Fig. 8M-M″). We tested whether SPECC1L protein interacts with adherens junction proteins using U2OS cell lysate due to high expression in this cell line. SPECC1L and E-CAD were detected in the input of total protein lysate. E-CAD protein was pulled down with SPECC1L in a co-IP assay in a lysate immunoprecipitated by anti-SPECC1L antibodies, and no band was detected in the negative control (Fig. 8O).

Common variants near TWIST1 impact human facial shape

The different mouse models of Twist1 loss of expression and function indicate the crucial role of this gene in regulating the development of CNCC-derived structures. To determine the relevance of these animal model findings to humans, we revisited a recently published GWAS meta-analysis of 8246 individuals of recent European ancestry, which identified 203 genome-wide significant signals associated with normal-range facial shape, as previously described by White et al. (2021). When moving from the minor (A) to major (T) allele of the leading SNP, the major shape changes include a less protrusive and wider nose, a shorter and less protrusive upper lip, and a more prominent chin and forehead. These changes are visible in the morphs and in a heatmap that shows the same effect using the normal displacement at each point comprising the facial surface (∼8000 points), with blue representing inward depression and red representing outward protrusion (Fig. 9A). A signal at 7p21.1 near TWIST1 was significantly associated facial shape and the lead SNP (rs212672) is located ∼355 kb downstream of TWIST1 (Fig. 9B). When the face was partitioned into global-to-local anatomical segments, the strongest association with the lead SNP was observed for the whole face (P=1.31E−20) and for CNCC-derived facial regions involving the nose and upper lip (see rosette diagram in Fig. 9C).

Uncovering the underlying mechanism of Twist1 and Irf6 function in cell fate regulation is crucial to identify their regulatory pathways and associated genes that can control tissue development and differentiation. Our work tested the hypothesis that Twist1 is involved in neural tube formation and EMT of CNCCs, while Irf6 helps define neural tube dorsal edges and structural integrity. Our immunofluorescence staining data showed that TWIST1 expression has a similar pattern to β/δ-catenin and tight junction proteins at the apical side of neural plate and neural folds. This cellular overlap is validated by the dual immunofluorescence of TWIST1 and β-catenin. Furthermore, the in situ hybridization results showed that Twist1 mRNA is also weakly expressed in the dorsal edges of neural plate and apical side of neural folds. Our findings are consistent with a recent study indicating that a few cells of the neural folds express Twist1 mRNA when detected using single cell RNA-seq technology (Soldatov et al., 2019), although the scRNA-seq data relies on computational analysis to determine the spatial expression and the relative anatomical distribution of this population of cells (Soldatov et al., 2019). Additionally, Twist1 expression was detected in round cells residing in the neuroectoderm of the forebrain during the emergence of CNCCs at E8.5 (Füchtbauer, 1995).

To provide valuable insight on TWIST1 apical expression in the neural folds, we explored whether TWIST1 is co-expressed with LRP2 and RAB11b in endocytic vesicle compartments. We based our investigation on a previous report on LRP2 and RAB11b co-expression in endocytic compartments at the apical surface of neural folds (Kowalczyk et al., 2021). We found that TWIST1 is expressed in endocytic vesicles at the apical surface of neural folds, and its expression partially overlaps with RAB11b and LRP2. Recycling endosomes are important for apical constriction, and loss of Lrp2 leads to neural tube closure defects in mouse embryos (Kowalczyk et al., 2021). Similarly, loss of Twist1, β-catenin and δ-catenin in neuroectodermal cells leads to neural tube closure defects (Kowalczyk et al., 2021; Brault et al., 2001; Pieters et al., 2020). The cellular fractionation co-IP data shows that the cytosolic TWIST1 interacts with β/δ-catenins during neural tube closure, and Twist1 CKO using Wnt1-Cre disrupts the apical expression of β-catenin, which mostly becomes cytosolic. Notably, δ-catenin CKO by Wnt1-Cre also causes a focal loss of N-cadherin at the apical side of neural folds (Pieters et al., 2020). However, we cannot definitively disentangle the cytosolic versus nuclear contribution of Twist1 because it is a dynamic and continuous process, and the disruption of the neural tube formation might impact the delamination of the CNCCs. Therefore, our data suggests that the interaction of cytosolic TWIST1 with β/δ-catenins, and the expression of TWIST1 in endocytic vesicles at the apical surface might facilitate the apical constriction of neuroectodermal cells, whereas lack of Twist1 in neuroectodermal cells leads to an altered cell shape with multiple ectopic dorsolateral hinge points at the apical surface of the neural tube. Our findings are consistent with a recent publication showing that TWIST1 interacts with tight junction protein 1 (TJP1) along the cell membrane of two different cancer cells (Liu et al., 2022). Liu et al. also demonstrated that TWIST1 expression overlaps with TJP1 in vesicles at cellular membrane (Liu et al., 2022).

Twist1-null mice exhibited degeneration of the apical neuroectodermal cells, and increased cell death in migratory CNCCs (Bildsoe et al., 2009; Chen and Behringer, 1995; Soo et al., 2002). Although the adjacent cephalic mesenchyme activity contributes to neural fold elevation, it does not explain why the neural folds in Twist1-null embryos do elevate but fail to bend and close completely. Similarly, in Twist1cko/− in CNCCs, the neural folds bend as if to join but stop prematurely, leading to improper closure of the neural tube. Although the phenotypes caused by a lack of Twist1 in CNCCs and mesoderm can be explained by an indirect role of Twist1 in regulating neural tube formation (Bildsoe et al., 2013), Twist1 expression in the neural folds provides a complementary mechanism for explaining the neural tube closure defects in Twist1-null embryos, the multiple ectopic bending points and the expansion in the neural tube of Twist1cko/−. Although a complete loss of Twist1 did not disrupt the anterior-posterior patterning of the neural tube at E9.5, there was expansion in a few components of Shh and Fgf pathways (Soo et al., 2002). A direct role in the expansion of these components has not been previously reported because Twist1 expression was not described in the neural folds (Soo et al., 2002). Twist1 CKO in mesodermal cells disrupts neural fold elevation (Bildsoe et al., 2013), and the rescue experiment with wild-type mesenchymal cells in Twist1-null chimera embryos indicated that cephalic mesenchymal cells are involved in neural fold elevation (Chen and Behringer, 1995). For the function of IRF6 in neural tube, we have shown that overexpression of Irf6 in the basal layer of non-neural ectoderm suppresses Tfap2a in the ectoderm and causes neural tube abnormalities and exencephaly (Kousa et al., 2019). Our current findings show that AP2α expression was expanded into the neural tube in Irf6-null embryos, which could explain the disruption of neural tube edge integrity. In addition, the glial progenitor cells within the neural tube were disorganized in Irf6-null embryos with more noticeable gaps between the progenitor cells.

The Wnt1 ligand is a crucial regulator in the identification of neural plate border domains and in early induction of CNCCs (Parr et al., 1993). The Wnt1-Cre transgene drives the expression of Cre throughout the neural plate at E8.5 and later stages during neural tube development (Chang et al., 2015). Hence, Wnt1-Cre was ideal to delete Twist1 at early stages, if present, during early neural plate formation. Previous studies have suggested that Twist1 CKO in neuroectoderm by Wnt1-Cre does not disrupt CNCC migration towards the frontonasal and pharyngeal processes. However, the X-gal staining in Twist1-deficient CNCCs was reduced in frontonasal and pharyngeal arches, as Twist1 was considered a survival factor in migratory CNCCs (Chen et al., 2007, 2017; Zhang et al., 2012). Apoptosis in Twist1-deficient CNCCs might be one of the disrupted cellular functions behind the loss of craniofacial bone and cartilage. Yet no report has indicated whether mammalian Twist1 is involved in the events before CNCC delamination or whether the migratory Twist1-deficient CNCCs are normal mesenchymal cells. The data in this study from neural tube explants of embryos shows that Twist1 regulates EMT in CNCCs and that a loss of Twist1 in neuroectoderm leads to disruption of cell fate transition and delamination. Our findings also show that TWIST1 suppresses the expression of Irf6 and other epithelial factors during delamination. Irf6 is a terminal differentiation factor of proliferative epithelial cells in the ectoderm and its expression prevents EMT (Fakhouri et al., 2017; Ingraham et al., 2006; Thompson et al., 2019). In agreement with its role in EMT, Twist1 overexpression in the trunk neural tube converts the pre-EMT neuroectodermal cells into cells that resemble CNCCs (Soldatov et al., 2019). Recent studies on the EMT show that a complete conversion from an epithelial to a mesenchymal state is not necessary for cancer cells to migrate and multiple metastable states during cell transition have been described in cancer studies (Cousin, 2017; Nieto et al., 2016; Thiery et al., 2009). However, we did not observed a collective cell migration or epithelial-like cells in wild-type explants or in vivo histological staining. Lineage analysis may be the most appropriate way to detect the effect of Twist1 loss on target genes. Unfortunately, the lack of such analysis limits our interpretation of the impact of Twist1 on target gene expression in migratory CNCCs.

Determining the in vivo role of TWIST1 post-translational modification is vital in delineating its role in neural tube and CNCCs. TWIST1 is a phospho-protein and multiple phospho-residues are necessary for its stability, nuclear translocation and transcriptional regulation, as per previous cancer studies (Bourguignon et al., 2010; Zhao et al., 2017; Xu et al., 2017). Twist1 phosphorylation increases tumor cell motility in squamous cell carcinoma of the head and neck (Alexander et al., 2006; Su et al., 2011). In addition, overexpression of Twist1 phosphomimetic T125D; S127D led to limb abnormalities in transient transgenic mouse embryos (Firulli et al., 2007). We suggest that post-translational phosphorylation of TWIST1 is crucial for regulating its cellular activity in CNCCs. The Twist1S18;20A/S18;20A and Twist1S68A/S68A mutants embryos showed several craniofacial abnormalities, including epidermal blebbing, severe edema, meningeal detachment and neural tube defects. The skeletal staining of these phospho-incompetent mutant embryos showed bone loss and reduced skull mineralization of the frontonasal and maxillary bone. Yet the in vivo molecular function of the two phospho-sites needs to be delineated, including its importance in protein stability, nuclear translocation, the EMT process, cell survival and differentiation.

The data generated on the differentially expressed genes in Twist1cko/− embryos demonstrated that many factors involved in intercellular adhesion proteins and cytoskeletal remodeling are significantly altered. Previous studies have shown that mutations in adhesion protein and cytoskeletal regulator genes cause neural tube closure defects and improper delamination of CNCCs, leading to craniofacial disorders (Pieters et al., 2020; Wilson et al., 2016). One of the genes contributing to this delamination defect in mice is Specc1l. Mutations in TWIST1 and SPECC1L lead to craniosynostosis and orofacial clefting in humans (Bai et al., 2021; Bhoj et al., 2019; Goering et al., 2021; Hall et al., 2020). Specc1l is a gene that encodes a novel coiled-coil domain-containing protein, which stabilizes microtubules and actin of the cytoskeleton (Bhoj et al., 2019; Gfrerer et al., 2014). SPECC1L colocalizes with both cellular actin and tubulin (Saadi et al., 2011; Wilson et al., 2016). Thus, without sufficient SPECC1L, actin-cytoskeleton reorganization and cell adhesion are significantly impacted (Saadi et al., 2011). The overlap in cellular function and embryonic phenotype between Twist1 and Specc1l mutant mouse lines suggests that both genes are involved in a similar regulatory pathway to control cytoskeleton reorganization during cell delamination and migration. This study shows that TWIST1 directly binds to putative regulatory elements in the intronic regions of Specc1l and that loss of Twist1 leads to a significant reduction of Specc1l in hindbrain and first pharyngeal arch tissues, suggesting that TWIST1 acts as an activator for Specc1l.

In conclusion, our findings emphasize the direct role of Twist1 and Irf6 in neural tube development, and the conservation of TWIST1 function in regulating the EMT process in CNCCs by controlling the expression of epithelial genes and cytoskeletal regulators. Specc1l and Twist1 play a similar role in the delamination of CNCCs by remodeling the cell-cell adhesion and cytoskeletal reorganization. Thus, previous reports and the current study provide a reasonable explanation for the overlapping phenotypes in mice and humans. In addition, as our results highlight, reanalysis of the GWAS data in humans shows that variants near TWIST1 impact normal-range facial shape, including CNCC-derived structures of the midface (Claes et al., 2018; White et al., 2021). Notably, the most significant SNP near TWIST1 overlaps with epigenetic enhancer signatures in which the elements with the epigenetic marks were tested for enhancer activities to recapitulate TWIST1 endogenous expression (Hirsch et al., 2018). Understanding the transcriptional regulation of TWIST1 and its function in controlling the associated regulatory pathway involved in cell fate determination could lead to the identification of the missing heritability in families that are affected and at high risk of craniofacial birth defects.

Mouse strains

All the mice used in this study were generated using the C57BL/6J genetic background. Twist1 heterozygous mice were generated using EIIA-Cre mouse line using the B6;129S7-Twist1fl/fl [obtained from the Mutant Mouse Resource and Research Center (MMRRC) supported by the National Institutes of Health). The mouse lines 129S4.Cg-E2f1Tg(Wnt1-cre2)Sor/J (022137), BL6.129(Cg)-Gt(ROSA)26Sortm4(ACTB-tdTomato,−EGFP)Luo/J (007676) and BL6.129S4-Gt(ROSA)26Sortm1Sor/J (003474) were obtained from the Jackson Laboratory. The two different cell-tracing strategies R26tm4(ACTB-tdTomato,−EGFP) and R26tm1.lacZ were used to track in vivo CNCC formation and migration. Twist1S18;20A/+, Twist1S42A/+, Twist1S68A/+ and Twist1T121;S123A/+ phospho-incompetent founders were generated using CRISPR/Cas9 method at the Baylor College of Medicine. We backcrossed all phospho-mutant founders to C57BL/6J wild-type mice for six generations to avoid non-specific genomic alterations. Genomic DNA from the founders was sequenced to confirm the substitution of S18/20 and S68. Twist1S18I;20A/+ and Twist1S68A/+ heterozygous mice were crossed to obtain homozygous Twist1S18I;20A/S18I;20A and homozygous Twist1S68A/S68A embryos. These embryos were generated to determine the effects on the CNCC-derived craniofacial bone and cartilage, and the earliest time-point of detecting a craniofacial pathology. The animal work was approved by the Center for Laboratory Animal Medicine and Care committee at UT Health Houston under the Approved Animal protocol AWC-19-0045.

Mouse handling, embryo extraction, and genotyping

Embryos were extracted based on the presence of a copulation plug and the number of days after the last known delivery. Pregnant females were euthanized with CO2 followed by cervical dislocation. Embryos were genotyped for Twist1fl/fl, Twist1+/−, Twist1−/−, Twist1fl/−, Wnt1-Cre1, Wnt1-Cre2, Irf6−/+, Twist1S18I;20A/S18I;20A and Twist1S68A/S68A alleles via DNA extraction, allele-specific PCR and gel electrophoresis. The ratio of genotype-to-phenotype of Twist1 CKO and phosphor-incompetent mice was within the expected ratio for mouse.

Histological and immunofluorescent staining

Embryos of wild-type, Twist1−/−, Twist1cko/− and phospho-incompetent lines were collected for histological and immunofluorescence staining. Immunofluorescent staining was performed as previously described (Fakhouri et al., 2012). Briefly, mouse tissues were deparaffinized and rehydrated in a series of ethanol dilutions. The slides were boiled for 10 min in 10 mM sodium citrate buffer for antigen retrieval. Sections were blocked with anti-mouse IgG Fab fragment for 30 min, and then with 10% (v/v) normal goat serum and 1% BSA (v/v) in PBS for 1 h, then incubated overnight at 4°C with the following primary antibodies: mouse anti-AKT1/2/3 (1:150, Santa Cruz Biotechnology, sc-81434,), mouse anti-β-catenin (1:200, Santa Cruz Biotechnology, sc-7963), rabbit anti-β-catenin (1:150, Abcam, 16051), mouse anti-δ-catenin (1:150, BD Biosciences, 611537), rabbit anti-LRP2 (1:150, Abcam, ab76969), rabbit anti-RAB11b (1:150, Proteintech, 15903-I-AP), mouse anti-caludin 1 (1:200, Santa Cruz Biotechnology, sc-166338), mouse anti-N-cadherin (1:200, Santa Cruz Biotechnology, sc-59987), mouse anti-occludin (1:150, Santa Cruz Biotechnology, sc-133256), mouse anti-vimentin (1:150, Santa Cruz Biotechnology, sc-6260), mouse anti-E-cadherin (1:150, BD Biosciences, BD-610182), mouse anti-TWIST1 (1:200, Abcam, ab50887), SOX9 (1:200, Abcam, ab185966), mouse anti-ZO-1 (1:200, Santa Cruz Biotechnology, sc-33725), mouse anti-TFAP2α (1:50, DSHB, PCRP-TFAP2A-2C2), rabbit anti-CD73 (1:150, Abcam, ab133582), rabbit anti-CD206 (1:150, Abcam, ab64693), phalloidin (1:1000, Abcam, ab176757) and rabbit anti-IRF6 (1:500: Fakhouri et al., 2012). The secondary antibodies were goat anti-rabbit (1:150, A21429, Molecular Probes) and goat anti-mouse (1:150, A11029, Molecular Probes). We stained nuclei using DAPI (D3571, Invitrogen). An X-Cite Series 120Q laser and a CoolSnap HQ2 photometric camera (Andor Neo/Zyla) installed in a fluorescent microscope (Nikon Eclipse Ni) were used to capture images. For the immunofluorescence staining, sections from more than four embryos of each genotype, wild type and Twist1cko/−, were used for each embryonic time point tested.

Twist1 mRNA probe for in situ hybridization

The Twist1 probe for in situ hybridization was designed to recognize a 0.8 kb region in the 3′ UTR region of mouse Twist1, starting at 188 bp and ending at 1062 bp after the Twist1 coding sequence ends. This avoids any conserved regions homologous with other bHLH factors and encompasses the sequence of the shorter probe described by Wolf et al. (1991). Further details are provided in Fig. S3.

Whole-mount embryo β-gal staining

R26Tm1 transgenic mice were used to perform X-gal staining for cell tracing in whole-mount embryos. Whole-mount embryo staining at E9.5, E11.5 and E15.5 was performed as previously described (Metwalli et al., 2018). Stained wild-type and Twist1cko/− embryos were then visualized, analyzed and compared using a stereomicroscope and NIS Elements AR software. The X-gal staining was performed on more than three biological replicates of each genotype and embryonic stage tested.

Skeletal staining

Twist1 phospho-incompetent embryos at E16.5 were dissolved in 2% KOH until the skin dissolved, and then stained with Alcian Blue solution overnight for cartilage and Alizarin Red for bone, as previously described (Thompson et al., 2019). The embryos were then submerged in 10% glycerol and 1% KOH to remove excess stain and remaining tissues for 3-5 days.

Neural tube explant excision and CNCC culture

E8.5-E9.5 embryos from the Twist1fl/−;Irf6−/+;Wnt1-Cre2 crosses described previously were extracted from their amniotic sacs and maintained in an organ culture medium. Using microdissection instruments, E8.5-E9.5 embryos were removed from their amniotic sacs. Transverse cuts were made through the optic vesicle and first pharyngeal arch. After the removal of extraneous tissue, the mid and hindbrain regions were sectioned and cultured on collagen type I-coated petri dishes (10 µg/cm2) in alpha-MEM medium (10% fetal bovine serum, 100 U/ml penicillin G, 100 μg/ml streptomycin, and 1% non-essential and essential amino acids) and placed in a 5% CO2 incubator at 37°C overnight to allow for appropriate adherence.

Time-lapse image acquisition and analysis

Time-lapse imaging of explants was performed using the A1R MP Confocal and Multiphoton Microscope (Nikon Instruments). Using overlaid contrast of red (TRITC), green (FITC) and blue (Hoechst) channels, fluorescence color changes from tissue explants were analyzed. Green fluorescence indicated the GFP reporter found only in CNCCs, while red fluorescence indicated all other tissues that express RFP, with Hoechst serving as a nuclei counterstain. Neural tubes from more than three biological replicates of each genotype were imaged in an environmentally controlled chamber maintained at 37°C and 5% CO2, and the migration of CNCCs was captured using automated acquisition over a 21 h period. The net migration of cell bodies over a 10 h period was calculated using IMARIS software (Bitplane). Parameters included a spot diameter and connected components algorithm used for cell tracking. Cells also were tracked frame by frame using the manual-tracking spot feature of IMARIS. About 16 wild-type and 22 Twist1cko/− CNCCs were tracked. A two-tailed Mann–Whitney U-test was used with a significance level ****P<0.0001.

Western blot, protein phosphorylation and RT-qPCR

We performed immunoblots to measure the quantitative levels of TWIST1 (ab50887), SNAIL2 (sc-166476), AKT1 (sc-81434), E-CAD (sc-8426), N-CAD (sc-59987), β-catenin (sc-7963) and WNT3 (sc-74537). mRNA expression of Twist1 and Irf6 in the hindbrain and first pharyngeal arch was also measured in three pooled biological and four technical replicates at E9.5. We also measured the mRNA level of these genes in addition to Pi3k, Ck2, Erk1, vimentin, occludin, Msx1, Tfap2a, Hand2, Yap, Wnt1, Rhoa, Rhoa, Arhgap29, Sox2, Sox10, Hoxd10, Jag1, Src and Mtor. The levels of phosphorylated and unphosphorylated forms of TWIST1 at S42, S68 and S127 were detected via immunoblot as previous described (Fakhouri et al., 2017; Hong et al., 2011). We received aliquots of rabbit polyclonal antibodies for TWIST1 P-S42 and P-S127 from Dr Brian A. Hemmings (Friedrich Miescher Institute for Biomedical Research, Basel, Switzerland; Vichalkovski et al., 2010). We purchased the monoclonal TWIST1 phospho-S68 from Abcam (ab187008). Monoclonal antibodies for TWIST1 (1:200, Abcam, ab50887) were used for immunofluorescence, immunoblotting and co-IP. We used the same antibodies previously described for E-CAD and β-catenin for immunofluorescence, western and co-IP. For the digestion with phosphatase enzyme, protein phosphatase 2A1 bovine (PP2A; Millipore-Sigma, P6993) was used to digest 100 µl total protein extracted from O9-1 CNCCs. About 10 µl PP2A in 50 mM TRIS-HCl buffer containing 2-mercaptoethanol was used and the solution was incubated at 30°C for 30 min and then analyzed by western blot using TWIST1 monoclonal antibodies (Abcam, ab50887).

Sanger sequencing

Purified genomic DNA from mouse tail snips and purified PCR products were submitted for Sanger sequencing at Genewiz Company. We sequenced exon 1 of the Twist1-null and the two Twist1 phospho-incompetent mouse lines for S18/20 and S68 regions.

Co-IP assay for protein-protein interactions

We extracted total protein from mid and hindbrain tissues dissected from embryos at E8.5-E9.0. Tissues were ground with a plastic pestle on dry ice and the samples were frozen and thawed twice before centrifugation to remove undissolved materials. Purified total protein was incubated with Protein A/G conjugated to magnetic beads. We used monoclonal antibodies (Abs) for TWIST1 (1:200, Abcam, ab50887), E-CAD (1:150, Santa Cruz Biotechnology, sc-8426), δ-catenin (1:150, BD Biosciences, 611537) and β-catenin (1:150, Santa Cruz Biotechnology, sc-7963). We performed cellular fractionation for total protein extracted from E8.5 and E9.5 hindbrain tissues using a cytoplasmic buffer [50 mM Tris-HCl (pH 8), 137.5 mM NaCl, 0.05% Triton X-100, 10% glycerol, 5 mM EDTA and Protease Inhibitor Cocktail (Thermo Scientific)] and a nuclear buffer [50 mM Tris-HCl (pH 8), 137.5 mM NaCl, 0.1% Triton X-100, 0.5% SDS, 10% glycerol, 5 mM EDTA and Protease Inhibitor Cocktail (Thermo Scientific)]. The cytosolic and nuclear fractions were incubated individually with Protein A/G conjugated to magnetic beads for the co-IP assay.

ChIP-PCR for TWIST1 binding to Specc1l putative regulatory elements

Chromatin immunoprecipitation (ChIP) was performed in tissues dissected from hindbrain and first pharyngeal arch. We used TWIST1 monoclonal antibodies to detect the binding to two putative enhancer elements within intron 1 and 2 of Specc1l. We also pulled down fragmented chromatins with IgG antibodies as a negative control for non-specific binding. The ChIP-PCR was performed as previously described (Kousa et al., 2019).

Statistical analysis

The β-actin and ribosomal protein S16 (Rps16) were used for normalization of western blot and RT-qPCR data as internal controls, respectively. A two-tailed Student's t-test analysis was applied for the statistical analysis between the wild-type control and Twist1 CKO quantitative data. The difference in average mean value was considered statistically significant if P<0.05.

We thank April Zhang and Yazan Hasan for their excellent help with the immunofluorescent imaging and histological staining of the mouse embryonic tissues. Special thanks to Derrick Thomas for his help with the revised manuscript. We are very grateful to Dr Brian A. Hemmings at the Friedrich Miescher Institute for Biomedical Research, Basel, who sent us sufficient aliquots of rabbit polyclonal antibodies for TWIST1 P-S42 and P-S127, and to Drs Pierre McCrea at MD Anderson Cancer Research Institute, Noriaki Ono at UT Health School of Dentistry at Houston and David Sheikh-Hamad for providing us with aliquots of the δ-catenin, rabbit β-catenin and LRP2 antibodies, respectively. We thank Dr Brendan Lee and Ms Racel Cela from Baylor College of Medicine for their tremendous help maintaining the Twist1 phospho-incompetent mouse founders at the Baylor mouse facility. We further thank the Genetically Engineered Mouse Core at Baylor College of Medicine for generating the Twist1 phospho-incompetent mouse lines.

Author contributions

Conceptualization: J.W.B., W.D.F.; Methodology: J.W.B., S.J., R.A., V.K.X., K.M.H., L.N., M.S., J.P.G., P.N., S.L., H.H., I.S., M.C.F.-C., W.D.F.; Software: K.M.H.; Validation: J.W.B., S.J., R.A., M.S., J.P.G., P.N., M.C.F.-C., W.D.F.; Formal analysis: J.W.B., S.J., R.A., V.K.X., K.M.H., L.C., L.N., M.S., J.P.G., P.N., S.L., H.H., P.C., S.M.W., I.S., M.C.F.-C., W.D.F.; Investigation: J.W.B., S.J., V.K.X., L.N., S.L., W.D.F.; Resources: W.D.F.; Data curation: J.W.B., R.A., L.C., P.C., S.M.W., W.D.F.; Writing - original draft: J.W.B., S.J., R.A., V.K.X., W.D.F.; Writing - review & editing: J.W.B., R.A., L.C., L.N., H.H., P.C., S.M.W., I.S., M.C.F.-C., W.D.F.; Visualization: J.W.B., S.J., R.A., V.K.X., K.M.H., L.C., L.N., M.S., J.P.G., P.N., S.L., H.H., P.C., S.M.W., I.S., M.C.F.-C.; Supervision: I.S., M.C.F.-C., W.D.F.; Project administration: W.D.F.; Funding acquisition: W.D.F.

Funding

This study was funded by start-up funds from the University of Texas Health Science Center at Houston School of Dentistry, and the Rolanette and Berdon Lawrence Bone Disease Program of Texas. The facial GWAS work was supported by grants from the National Institute of Dental and Craniofacial Research (U01-DE020078, R01-DE016148 and R01-DE027023). I.S. was supported in part by the National Institute of Dental and Craniofacial Research (R01-DE026172). Deposited in PMC for release after 12 months.

Achilleos
,
A.
and
Trainor
,
P.
(
2012
).
Neural crest stem cells: discovery, properties and potential for therapy
.
Cell Res.
22
,
288
-
304
.
Alexander
,
N. R.
,
Tran
,
N. L.
,
Rekapally
,
H.
,
Summers
,
C. E.
,
Glackin
,
C.
and
Heimark
,
R. L.
(
2006
).
N-cadherin gene expression in prostate carcinoma is modulated by integrin-dependent nuclear translocation of Twist1
.
Cancer Res.
66
,
3365
-
3369
.
Bai
,
S.
,
Geng
,
Y.
,
Duan
,
H.
,
Xu
,
L.
,
Yu
,
Z.
,
Yuan
,
J.
and
Wei
,
M.
(
2021
).
A novel p.Pro871Leu missense mutation in SPECC1L gene causing craniosynostosis in a patient
.
Orthod. Craniofac. Res.
24
,
480
-
485
.
Baumgarten
,
A.
,
Bang
,
C.
,
Tschirner
,
A.
,
Engelmann
,
A.
,
Adams
,
V.
,
von Haehling
,
S.
,
Doehner
,
W.
,
Pregla
,
R.
,
Anker
,
M. S.
,
Blecharz
,
K.
et al. 
(
2013
).
TWIST1 regulates the activity of ubiquitin proteasome system via the miR-199/214 cluster in human end-stage dilated cardiomyopathy
.
Int. J. Cardiol.
168
,
1447
-
1452
.
Bhoj
,
E. J.
,
Haye
,
D.
,
Toutain
,
A.
,
Bonneau
,
D.
,
Nielsen
,
I. K.
,
Lund
,
I. B.
,
Bogaard
,
P.
,
Leenskjold
,
S.
,
Karaer
,
K.
,
Wild
,
K. T.
et al. 
(
2019
).
Phenotypic spectrum associated with SPECC1L pathogenic variants: new families and critical review of the nosology of Teebi, Opitz GBBB, and Baraitser-Winter syndromes
.
Eur. J. Med. Genet.
62
,
103588
.
Bildsoe
,
H.
,
Loebel
,
D. A.
,
Jones
,
V. J.
,
Chen
,
Y. T.
,
Behringer
,
R. R.
and
Tam
,
P. P.
(
2009
).
Requirement for Twist1 in frontonasal and skull vault development in the mouse embryo
.
Dev. Biol.
331
,
176
-
188
.
Bildsoe
,
H.
,
Loebel
,
D. A.
,
Jones
,
V. J.
,
Hor
,
A. C.
,
Braithwaite
,
A. W.
,
Chen
,
Y. T.
,
Behringer
,
R. R.
and
Tam
,
P. P.
(
2013
).
The mesenchymal architecture of the cranial mesoderm of mouse embryos is disrupted by the loss of Twist1 function
.
Dev. Biol.
374
,
295
-
307
.
Bildsoe
,
H.
,
Fan
,
X.
,
Wilkie
,
E. E.
,
Ashoti
,
A.
,
Jones
,
V. J.
,
Power
,
M.
,
Qin
,
J.
,
Wang
,
J.
,
Tam
,
P. P. L.
and
Loebel
,
D. A. F.
(
2016
).
Transcriptional targets of TWIST1 in the cranial mesoderm regulate cell-matrix interactions and mesenchyme maintenance
.
Dev. Biol.
418
,
189
-
203
.
Bourguignon
,
L. Y.
,
Wong
,
G.
,
Earle
,
C.
,
Krueger
,
K.
and
Spevak
,
C. C.
(
2010
).
Hyaluronan-CD44 interaction promotes c-Src-mediated twist signaling, microRNA-10b expression, and RhoA/RhoC up-regulation, leading to Rho-kinase-associated cytoskeleton activation and breast tumor cell invasion
.
J. Biol. Chem.
285
,
36721
-
36735
.
Brault
,
V.
,
Moore
,
R.
,
Kutsch
,
S.
,
Ishibashi
,
M.
,
Rowitch
,
D. H.
,
McMahon
,
A. P.
,
Sommer
,
L.
,
Boussadia
,
O.
and
Kemler
,
R.
(
2001
).
Inactivation of the β-catenin gene by Wnt1-Cre-mediated deletion results in dramatic brain malformation and failure of craniofacial development
.
Development
128
,
1253
-
1264
.
Chai
,
Y.
and
Maxson
,
R. E.
Jr
. (
2006
).
Recent advances in craniofacial morphogenesis
.
Dev. Dyn.
235
,
2353
-
2375
.
Chang
,
Y.-W.
,
Su
,
Y.-J.
,
Hsiao
,
M.
,
Wei
,
K.-C.
,
Lin
,
W.-H.
,
Liang
,
C.-L.
,
Chen
,
S.-C.
and
Lee
,
J.-L.
(
2015
).
Diverse targets of β-catenin during the epithelial-mesenchymal transition define cancer stem cells and predict disease relapse
.
Cancer Res.
75
,
3398
-
3410
.
Chen
,
Z. F.
and
Behringer
,
R. R.
(
1995
).
Twist is required in head mesenchyme for cranial neural tube morphogenesis
.
Genes Dev.
9
,
686
-
699
.
Chen
,
Y. T.
,
Akinwunmi
,
P. O.
,
Deng
,
J. M.
,
Tam
,
O. H.
and
Behringer
,
R. R.
(
2007
).
Generation of a Twist1 conditional null allele in the mouse
.
Genesis
45
,
588
-
592
.
Chen
,
J.
,
Yuan
,
W.
,
Wu
,
L.
,
Tang
,
Q.
,
Xia
,
Q.
,
Ji
,
J.
,
Liu
,
Z.
,
Ma
,
Z.
,
Zhou
,
Z.
,
Cheng
,
Y.
et al. 
(
2017
).
PDGF-D promotes cell growth, aggressiveness, angiogenesis and EMT transformation of colorectal cancer by activation of Notch1/Twist1 pathway
.
Oncotarget
8
,
9961
-
9973
.
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
.
Cousin
,
H.
(
2017
).
Cadherins function during the collective cell migration of Xenopus Cranial Neural Crest cells: revisiting the role of E-cadherin
.
Mech. Dev.
148
,
79
-
88
.
El Ghouzzi
,
V.
,
Le Merrer
,
M.
,
Perrin-Schmitt
,
F.
,
Lajeunie
,
E.
,
Benit
,
P.
,
Renier
,
D.
,
Bourgeois
,
P.
,
Bolcato-Bellemin
,
A. L.
,
Munnich
,
A.
and
Bonaventure
,
J.
(
1997
).
Mutations of the TWIST gene in the Saethre-Chotzen syndrome
.
Nat. Genet.
15
,
42
-
46
.
Fakhouri
,
W. D.
,
Rhea
,
L.
,
Du
,
T.
,
Sweezer
,
E.
,
Morrison
,
H.
,
Fitzpatrick
,
D.
,
Yang
,
B.
,
Dunnwald
,
M.
and
Schutte
,
B. C.
(
2012
).
MCS9.7 enhancer activity is highly, but not completely, associated with expression of Irf6 and p63
.
Dev. Dyn.
241
,
340
-
349
.
Fakhouri
,
W. D.
,
Metwalli
,
K.
,
Naji
,
A.
,
Bakhiet
,
S.
,
Quispe-Salcedo
,
A.
,
Nitschke
,
L.
,
Kousa
,
Y. A.
and
Schutte
,
B. C.
(
2017
).
Intercellular genetic interaction between Irf6 and Twist1 during craniofacial development
.
Sci. Rep.
7
,
7129
.
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
,
e62873
.
Farrell
,
B. C.
,
Power
,
E. M.
and
Mc Dermott
,
K. W.
(
2011
).
Developmentally regulated expression of Sox9 and microRNAs 124,128 and 23 in neuroepithelial stem cells in the developing spinal cord
.
Int. J. Dev. Neurosci.
29
,
31
-
36
.
Firulli
,
A. B.
and
Conway
,
S. J.
(
2008
).
Phosphoregulation of Twist1 provides a mechanism of cell fate control
.
Curr. Med. Chem.
15
,
2641
-
2647
.
Firulli
,
B. A.
,
Krawchuk
,
D.
,
Centonze
,
V. E.
,
Vargesson
,
N.
,
Virshup
,
D. M.
,
Conway
,
S. J.
,
Cserjesi
,
P.
,
Laufer
,
E.
and
Firulli
,
A. B.
(
2005
).
Altered Twist1 and Hand2 dimerization is associated with Saethre-Chotzen syndrome and limb abnormalities
.
Nat. Genet.
37
,
373
-
381
.
Firulli
,
B. A.
,
Redick
,
B. A.
,
Conway
,
S. J.
and
Firulli
,
A. B.
(
2007
).
Mutations within helix I of Twist1 result in distinct limb defects and variation of DNA binding affinities
.
J. Biol. Chem.
282
,
27536
-
27546
.
Füchtbauer
,
E. M.
(
1995
).
Expression of M-twist during postimplantation development of the mouse
.
Dev. Dyn.
204
,
316
-
322
.
Gfrerer
,
L.
,
Shubinets
,
V.
,
Hoyos
,
T.
,
Kong
,
Y.
,
Nguyen
,
C.
,
Pietschmann
,
P.
,
Morton
,
C. C.
,
Maas
,
R. L.
and
Liao
,
E. C.
(
2014
).
Functional analysis of SPECC1L in craniofacial development and oblique facial cleft pathogenesis
.
Plast. Reconstr. Surg.
134
,
748
-
759
.
Goering
,
J. P.
,
Isai
,
D. G.
,
Hall
,
E. G.
,
Wilson
,
N. R.
,
Kosa
,
E.
,
Wenger
,
L. W.
,
Umar
,
Z.
,
Yousaf
,
A.
,
Czirok
,
A.
and
Saadi
,
I.
(
2021
).
SPECC1L-deficient primary mouse embryonic palatal mesenchyme cells show speed and directionality defects
.
Sci. Rep.
11
,
1452
.
Groves
,
A. K.
and
LaBonne
,
C.
(
2014
).
Setting appropriate boundaries: fate, patterning and competence at the neural plate border
.
Dev. Biol.
389
,
2
-
12
.
Hall
,
E. G.
,
Wenger
,
L. W.
,
Wilson
,
N. R.
,
Undurty-Akella
,
S. S.
,
Standley
,
J.
,
Augustine-Akpan
,
E. A.
,
Kousa
,
Y. A.
,
Acevedo
,
D. S.
,
Goering
,
J. P.
,
Pitstick
,
L.
et al. 
(
2020
).
SPECC1L regulates palate development downstream of IRF6
.
Hum. Mol. Genet.
,
29
,
845
-
858
.
He
,
F.
,
Hu
,
X.
,
Xiong
,
W.
,
Li
,
L.
,
Lin
,
L.
,
Shen
,
B.
,
Yang
,
L.
,
Gu
,
S.
,
Zhang
,
Y.
and
Chen
,
Y.
(
2014
).
Directed Bmp4 expression in neural crest cells generates a genetic model for the rare human bony syngnathia birth defect
.
Dev. Biol.
391
,
170
-
181
.
Hirsch
,
N.
,
Eshel
,
R.
,
Bar Yaacov
,
R.
,
Shahar
,
T.
,
Shmulevich
,
F.
,
Dahan
,
I.
,
Levaot
,
N.
,
Kaplan
,
T.
,
Lupiáñez
,
D. G.
and
Birnbaum
,
R. Y.
(
2018
).
Unraveling the transcriptional regulation of TWIST1 in limb development
.
PLoS Genet.
14
,
e1007738
.
Hong
,
J.
,
Zhou
,
J.
,
Fu
,
J.
,
He
,
T.
,
Qin
,
J.
,
Wang
,
L.
,
Liao
,
L.
and
Xu
,
J.
(
2011
).
Phosphorylation of serine 68 of Twist1 by MAPKs stabilizes Twist1 protein and promotes breast cancer cell invasiveness
.
Cancer Res.
71
,
3980
-
3990
.
Hoyert
,
D. L.
,
Mathews
,
T. J.
,
Menacker
,
F.
,
Strobino
,
D. M.
and
Guyer
,
B.
(
2006
).
Annual summary of vital statistics: 2004
.
Pediatrics
117
,
168
-
183
.
Ingraham
,
C. R.
,
Kinoshita
,
A.
,
Kondo
,
S.
,
Yang
,
B.
,
Sajan
,
S.
,
Trout
,
K. J.
,
Malik
,
M. I.
,
Dunnwald
,
M.
,
Goudy
,
S. L.
,
Lovett
,
M.
et al. 
(
2006
).
Abnormal skin, limb and craniofacial morphogenesis in mice deficient for interferon regulatory factor 6 (Irf6)
.
Nat. Genet.
38
,
1335
-
1340
.
Joshi
,
N.
,
Hamdan
,
A. M.
and
Fakhouri
,
W. D.
(
2014
).
Skeletal malocclusion: a developmental disorder with a life-long morbidity
.
J. Clin. Med. Res
6
,
399
-
408
.
Kousa
,
Y. A.
,
Zhu
,
H.
,
Fakhouri
,
W. D.
,
Lei
,
Y.
,
Kinoshita
,
A.
,
Roushangar
,
R. R.
,
Patel
,
N. K.
,
Agopian
,
A. J.
,
Yang
,
W.
,
Leslie
,
E. J.
et al. 
(
2019
).
The TFAP2A-IRF6-GRHL3 genetic pathway is conserved in neurulation
.
Hum. Mol. Genet
28
,
1726
-
1737
.
Kowalczyk
,
I.
,
Lee
,
C.
,
Schuster
,
E.
,
Hoeren
,
J.
,
Trivigno
,
V.
,
Riedel
,
L.
,
Görne
,
J.
,
Wallingford
,
J. B.
,
Hammes
,
A.
and
Feistel
,
K.
(
2021
).
Neural tube closure requires the endocytic receptor Lrp2 and its functional interaction with intracellular scaffolds
.
Development
148
,
dev195008
.
Kunz
,
J.
,
Hudler
,
M.
and
Fritz
,
B.
(
1999
).
Identification of a frameshift mutation in the gene TWIST in a family affected with Robinow-Sorauf syndrome
.
J. Med. Genet
36
,
650
-
652
.
Lander
,
R.
,
Nasr
,
T.
,
Ochoa
,
S. D.
,
Nordin
,
K.
,
Prasad
,
M. S.
and
Labonne
,
C.
(
2013
).
Interactions between Twist and other core epithelial-mesenchymal transition factors are controlled by GSK3-mediated phosphorylation
.
Nat. Commun.
4
,
1542
.
Le Douarin
,
N. M.
,
Creuzet
,
S.
,
Couly
,
G.
and
Dupin
,
E.
(
2004
).
Neural crest cell plasticity and its limits
.
Development
131
,
4637
-
4650
.
Lewis
,
A. E.
,
Vasudevan
,
H. N.
,
O'Neill
,
A. K.
,
Soriano
,
P.
and
Bush
,
J. O.
(
2013
).
The widely used Wnt1-Cre transgene causes developmental phenotypes by ectopic activation of Wnt signaling
.
Dev. Biol.
379
,
229
-
234
.
Litingtung
,
Y.
and
Chiang
,
C.
(
2000
).
Specification of ventral neuron types is mediated by an antagonistic interaction between Shh and Gli3
.
Nat. Neurosci.
3
,
979
-
985
.
Liu
,
X. Q.
,
Shao
,
X. R.
,
Liu
,
Y.
,
Dong
,
Z.-X.
,
Chan
,
S.-H.
,
Shi
,
Y.-Y.
,
Chen
,
S.-N.
,
Qi
,
L.
,
Zhong
,
Li.
,
Yu
,
Y.
et al. 
(
2022
).
Tight junction protein 1 promotes vasculature remodeling via regulating USP2/TWIST1 in bladder cancer
.
Oncogene
41
,
502
-
514
.
Lu
,
S.
,
Nie
,
J.
,
Luan
,
Q.
,
Feng
,
Q.
,
Xiao
,
Q.
,
Chang
,
Z.
,
Shan
,
C.
,
Hess
,
D.
,
Hemmings
,
B. A.
and
Yang
,
Z.
(
2011
).
Phosphorylation of the Twist1-family basic helix-loop-helix transcription factors is involved in pathological cardiac remodeling
.
PLoS One
6
,
e19251
.
Metwalli
,
K. A.
,
Do
,
M. A.
,
Nguyen
,
K.
,
Mallick
,
S.
,
Kin
,
K.
,
Farokhnia
,
N.
,
Jun
,
G.
and
Fakhouri
,
W. D.
(
2018
).
Interferon regulatory factor 6 is necessary for salivary glands and pancreas development
.
J. Dent. Res.
97
,
226
-
236
.
Millonig
,
J. H.
,
Millen
,
K. J.
and
Hatten
,
M. E.
(
2000
).
The mouse Dreher gene Lmx1a controls formation of the roof plate in the vertebrate CNS
.
Nature
403
,
764
-
769
.
Murray
,
S. A.
,
Oram
,
K. F.
and
Gridley
,
T.
(
2007
).
Multiple functions of Snail family genes during palate development in mice
.
Development
134
,
1789
-
1797
.
Nieto
,
M. A.
,
Huang
,
R. Y.-J.
,
Jackson
,
R. A.
and
Thiery
,
J. P.
(
2016
).
Emt: 2016
.
Cell
166
,
21
-
45
.
Noden
,
D. M.
and
Trainor
,
P. A.
(
2005
).
Relations and interactions between cranial mesoderm and neural crest populations
.
J. Anat.
207
,
575
-
601
.
Nomura
,
M.
and
Li
,
E.
(
1998
).
Smad2 role in mesoderm formation, left-right patterning and craniofacial development
.
Nature
393
,
786
-
790
.
Nüsslein-Volhard
,
C.
,
Wieschaus
,
E.
and
Kluding
,
H.
(
1984
).
Mutations affecting the pattern of the larval cuticle in Drosophila melanogaster. 1. Zygotic loci on the 2nd chromosome
.
Wilhelm Roux's Arch. Dev. Biol.
193
,
267
-
282
.
Ota
,
M. S.
,
Loebel
,
D. A.
,
O'Rourke
,
M. P.
,
Wong
,
N.
,
Tsoi
,
B.
and
Tam
,
P. P.
(
2004
).
Twist is required for patterning the cranial nerves and maintaining the viability of mesodermal cells
.
Dev. Dyn.
230
,
216
-
228
.
Padmanabhan
,
R.
and
Ahmed
,
I.
(
1997
).
Retinoic acid-induced asymmetric craniofacial growth and cleft palate in the TO mouse fetus
.
Reprod. Toxicol.
11
,
843
-
860
.
Parker
,
S. E.
,
Mai
,
C. T.
,
Canfield
,
M. A.
,
Rickard
,
R.
,
Wang
,
Y.
,
Meyer
,
R. E.
,
Anderson
,
P.
,
Mason
,
C. A.
,
Collins
,
J. S.
,
Kirby
,
R. S.
et al. 
(
2010
).
Updated National Birth Prevalence estimates for selected birth defects in the United States, 2004-2006
.
Birth Defects Res. A Clin. Mol. Teratol.
88
,
1008
-
1016
.
Parr
,
B. A.
,
Shea
,
M. J.
,
Vassileva
,
G.
and
McMahon
,
A. P.
(
1993
).
Mouse Wnt genes exhibit discrete domains of expression in the early embryonic CNS and limb buds
.
Development
119
,
247
-
261
.
Pieters
,
T.
,
Sanders
,
E.
,
Tian
,
H.
,
van Hengel
,
J.
and
van Roy
,
F.
(
2020
).
Neural defects caused by total and Wnt1-Cre mediated ablation of p120ctn in mice
.
BMC Dev. Biol.
20
,
17
.
Richardson
,
R. J.
,
Dixon
,
J.
,
Malhotra
,
S.
,
Hardman
,
M. J.
,
Knowles
,
L.
,
Boot-Handford
,
R. P.
,
Shore
,
P.
,
Whitmarsh and
,
A.
and
Dixon
,
M. J.
(
2006
).
Irf6 is a key determinant of the keratinocyte proliferation-differentiation switch
.
Nat. Genet.
38
,
1329
-
1334
.
Saadi
,
I.
,
Alkuraya
,
F. S.
,
Gisselbrecht
,
S. S.
,
Goessling
,
W.
,
Cavallesco
,
R.
,
Turbe-Doan
,
A.
,
Petrin
,
A. L.
,
Harris
,
J.
,
Siddiqui
,
U.
,
Grix
,
A. W.
et al. 
(
2011
).
Deficiency of the cytoskeletal protein SPECC1L leads to oblique facial clefting
.
Am. J. Hum. Genet.
89
,
44
-
55
.
Seto
,
M. L.
,
Hing
,
A. V.
,
Chang
,
J.
,
Hu
,
M.
,
Kapp-Simon
,
K. A.
,
Patel
,
P. K.
,
Burton
,
B. K.
,
Kane
,
A. A.
,
Smyth
,
M. D.
,
Hopper
,
R.
et al. 
(
2007
).
Isolated sagittal and coronal craniosynostosis associated with TWIST box mutations
.
Am. J. Med. Genet. A
143A
,
678
-
686
.
Simpson
,
P.
(
1983
).
Maternal–zygotic gene interactions during formation of the dorsoventral pattern in Drosophila embryos
.
Genetics
105
,
615
-
632
.
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
,
eaas9536
.
Soo
,
K.
,
O'Rourke
,
M. P.
,
Khoo
,
P. L.
,
Steiner
,
K. A.
,
Wong
,
N.
,
Behringer
,
R. R.
and
Tam
,
P. P.
(
2002
).
Twist function is required for the morphogenesis of the cephalic neural tube and the differentiation of the cranial neural crest cells in the mouse embryo
.
Dev. Biol.
247
,
251
-
270
.
Stoetzel
,
C.
,
Weber
,
B.
,
Bourgeois
,
P.
,
Bolcato-Bellemin
,
A. L.
and
Perrin-Schmitt
,
F.
(
1995
).
Dorso-ventral and rostro-caudal sequential expression of M-twist in the postimplantation murine embryo
.
Mech. Dev.
51
,
251
-
263
.
Su
,
Y.-W.
,
Xie
,
T.-X.
,
Sano
,
D.
and
Myers
,
J. N.
(
2011
).
IL-6 stabilizes Twist and enhances tumor cell motility in head and neck cancer cells through activation of casein kinase 2
.
PLoS One
6
,
e19412
.
Takenouchi
,
T.
,
Sakamoto
,
Y.
,
Sato
,
H.
,
Suzuki
,
H.
,
Uehara
,
T.
,
Ohsone
,
Y.
and
Kosaki
,
K.
(
2018
).
Ablepharon and craniosynostosis in a patient with a localized TWIST1 basic domain substitution
.
Am. J. Med. Genet. A
176
,
2777
-
2780
.
Thiery
,
J. P.
,
Acloque
,
H.
,
Huang
,
R. Y.
and
Nieto
,
M. A.
(
2009
).
Epithelial-mesenchymal transitions in development and disease
.
Cell
139
,
871
-
890
.
Thompson
,
J.
,
Mendoza
,
F.
,
Tan
,
E.
,
Bertol
,
J. W.
,
Gaggar
,
A. S.
,
Jun
,
G.
,
Biguetti
,
C.
and
Fakhouri
,
W. D.
(
2019
).
A cleft lip and palate gene, Irf6, is involved in osteoblast differentiation of craniofacial bone
.
Dev. Dyn.
248
,
221
-
232
.
Vichalkovski
,
A.
,
Gresko
,
E.
,
Hess
,
D.
,
Restuccia
,
D. F.
and
Hemmings
,
B. A.
(
2010
).
PKB/AKT phosphorylation of the transcription factor Twist-1 at Ser42 inhibits p53 activity in response to DNA damage
.
Oncogene
29
,
3554
-
3565
.
Vogel
,
J. K.
and
Wegner
,
M.
(
2021
).
Sox9 in the developing central nervous system: a jack of all trades?
Neural. Regen. Res.
16
,
676
-
677
.
White
,
J. D.
,
Indencleef
,
K.
,
Naqvi
,
S.
,
Eller
,
R. J.
,
Hoskens
,
H.
,
Roosenboom
,
J.
,
Lee
,
M. K.
,
Li
,
J.
,
Mohammed
,
J.
,
Richmond
,
S.
et al. 
(
2021
).
Insights into the genetic architecture of the human face
.
Nat. Genet.
53
,
45
-
53
.
Wilson
,
N. R.
,
Olm-Shipman
,
A. J.
,
Acevedo
,
D. S.
,
Palaniyandi
,
K.
,
Hall
,
E. G.
,
Kosa
,
E.
,
Stumpff
,
K. M.
,
Smith
,
G. J.
,
Pitstick
,
L.
,
Liao
,
E. C.
et al. 
(
2016
).
SPECC1L deficiency results in increased adherens junction stability and reduced cranial neural crest cell delamination
.
Sci. Rep.
6
.
17735
.
Wolf
,
C.
,
Thisse
,
C.
,
Stoetzel
,
C.
,
Thisse
,
B.
,
Gerlinger
,
P.
and
Perrin-Schmitt
,
F.
(
1991
).
The M-twist gene of Mus is expressed in subsets of mesodermal cells and is closely related to the Xenopus X-twi and the Drosophila twist genes
.
Dev. Biol.
143
,
363
-
373
.
Xu
,
Y.
,
Lee
,
D. K.
,
Feng
,
Z.
,
Xu
,
Y.
,
Bu
,
W.
,
Li
,
Y.
,
Liao
,
L.
and
Xu
,
J.
(
2017
).
Breast tumor cell-specific knockout of Twist1 inhibits cancer cell plasticity, dissemination, and lung metastasis in mice
.
Proc. Natl. Acad. Sci. USA
114
,
11494
-
11499
.
Xue
,
G.
and
Hemmings
,
B. A.
(
2012
).
Phosphorylation of basic helix-loop-helix transcription factor Twist in development and disease
.
Biochem. Soc. Trans.
40
,
90
-
93
.
Zhang
,
Y.
,
Blackwell
,
E. L.
,
McKnight
,
M. T.
,
Knutsen
,
G. R.
,
Vu
,
W. T.
and
Ruest
,
L. B.
(
2012
).
Specific inactivation of Twist1 in the mandibular arch neural crest cells affects the development of the ramus and reveals interactions with hand2
.
Dev. Dyn.
241
,
924
-
940
.
Zhang
,
A.
,
Aslam
,
H.
,
Sharma
,
N.
,
Warmflash
,
A.
and
Fakhouri
,
W. D.
(
2021
).
Conservation of epithelial-to-mesenchymal transition process in neural crest cells and metastatic cancer
.
Cells Tissues Organs
210
,
151
-
172
.
Zhao
,
Z.
,
Rahman
,
M. A.
,
Chen
,
Z. G.
and
Shin
,
D. M.
(
2017
).
Multiple biological functions of Twist1 in various cancers
.
Oncotarget
8
,
20380
-
20393
.

Competing interests

The authors declare no competing or financial interests.

Supplementary information