Craniofacial development requires precise spatiotemporal regulation of multiple signaling pathways that crosstalk to coordinate the growth and patterning of the skull with surrounding tissues. Recent insights into these signaling pathways and previously uncharacterized progenitor cell populations have refined our understanding of skull patterning, bone mineralization and tissue homeostasis. Here, we touch upon classical studies and recent advances with an emphasis on developmental and signaling mechanisms that regulate the osteoblast lineage for the calvaria, which forms the roof of the skull. We highlight studies that illustrate the roles of osteoprogenitor cells and cranial suture-derived stem cells for proper calvarial growth and homeostasis. We also discuss genes and signaling pathways that control suture patency and highlight how perturbing the molecular regulation of these pathways leads to craniosynostosis. Finally, we discuss the recently discovered tissue and signaling interactions that integrate skull and cerebrovascular development, and the potential implications for both cerebrospinal fluid hydrodynamics and brain waste clearance in craniosynostosis.

Skull shapes have fascinated humans for millennia. Starting in ancient cultures and lasting until the 20th century, skull shaping, known as artificial cranial deformation, was commonly used for rites of passage, instilling communal unity, and to symbolize beauty and social status (Alfonso-Durrruty et al., 2015). Among the earliest written records of artificial cranial deformation are by Hippocrates in ‘On air, waters, and places’ (∼400 BCE). He describes a tribe known as the ‘Macrocephali’ that practiced artificial head lengthening by tightly wrapping newborn heads with cloth. By applying pressure during infancy and early childhood when the skull is malleable, mechanical deformation modulates physiological bone growth and permanently alters skull morphology. Similarly, the shape of the modern human skull, which is more dome-shaped compared to ancient hominids, evolved in response to internal forces and the expansion of the frontal and temporal lobes (Lieberman et al., 2002). These cultural practices and evolutionary adaptations exemplify the remarkable ability of the skull to grow and adapt in response to environmental and genetic factors, raising many questions about the signaling pathways and cell types that are involved.

The skull is a complex collection of flat bones composed of two main anatomical divisions. One aspect, the viscerocranium, forms the facial skeleton and is derived from the neural crest (Jin et al., 2016). The other division, the neurocranium, surrounds the brain (Kaucka and Adameyko, 2019). The neurocranium is further subdivided into the chondrocranium and calvarium, which form the skull base and skull vault, respectively. The calvarium includes the frontal and parietal bones, the squamous portion of the temporal bone, and the interparietal portion of the occipital bone (Holmes, 2012; Carter and Anslow, 2009; Ferguson et al., 2018) (Fig. 1). At the junction of these bones lie sutures, strips of fibrous tissue that act as synarthrotic joints. Sutures permit skull compliance, assist with shock absorption and are major signaling centers that regulate the shape of calvarial bones and their respective boundaries (Alaqeel et al., 2006; Siismets and Hatch, 2020; Opperman et al., 2000). Four major sutures separate the calvarial bones. The metopic/frontal suture separates the frontal bones, the coronal suture separates the frontal and parietal bones, and the sagittal suture lies on the dorsal midline between the parietal bones. Finally, the lambdoid suture lies between the parietal and occipital bones, which corresponds to the parietal-interparietal boundary in rodents (Fig. 1) (Li et al., 2021).

Fig. 1.

Comparative human and mouse skull anatomy. The major calvarial bones and intervening sutures are labeled. CS, coronal suture; Fb, frontal bone; FS, frontal suture; Ipb, interparietal bone; LS, lambdoid suture; MS, metopic suture; Nb, nasal bone; Ob, occipital bone; Pb, parietal bone; SS, sagittal suture; SqS, squamosal suture.

Fig. 1.

Comparative human and mouse skull anatomy. The major calvarial bones and intervening sutures are labeled. CS, coronal suture; Fb, frontal bone; FS, frontal suture; Ipb, interparietal bone; LS, lambdoid suture; MS, metopic suture; Nb, nasal bone; Ob, occipital bone; Pb, parietal bone; SS, sagittal suture; SqS, squamosal suture.

Some anatomical and embryological differences between human and animal calvaria exist. For example, cranial sutures typically fuse and ossify in adult humans, whereas most sutures remain patent in rodents and zebrafish (Quarto and Longaker, 2005). Also, comparative anatomical studies suggest the zebrafish ‘coronal suture’ is the anatomical equivalent of the lambdoidal suture in mammals (Teng et al., 2019). Despite these differences, the molecular mechanisms and major developmental processes in skull morphogenesis remain remarkably consistent between vertebrates (Cubbage and Mabee, 1996; Kanther et al., 2019; Topczewska et al., 2016; Ferguson and Atit, 2019). Thus, vertebrate models allow for the characterization of cellular interactions and signaling processes that control skull morphogenesis, and how these are affected in human craniofacial disorders. Here, primarily focusing on human studies and mouse models, we review key studies and recent insights that have shaped our understanding of the cellular and molecular regulation of calvarial patterning in normal development and disease states. We also discuss recently described populations of sutural stem cells that mediate postnatal skull growth and injury repair, as well as cutting-edge regenerative medicine approaches that may transform how we treat skull malformations. We conclude by describing how skull development is intertwined with that of the meningeal vasculature and then subsequently explain how craniosynostosis affects these processes, with a focus on the potential implications for brain health and function.

Development and differentiation of calvarial progenitors

Skull development requires the regulated commitment, differentiation and maturation of osteoblasts from undifferentiated mesenchymal precursor cells (Fig. 2A). Starting around mouse mid-gestation, between embryonic (E) days E8.5 and E9.5, neural crest cells from the mid-hindbrain domain along with the paraxial mesoderm migrate into the supraorbital ridge, forming a pool of calvarial progenitors in the supraorbital mesenchyme (SOM) (Ferguson and Atit, 2019; Jiang et al., 2002). This process is dependent upon canonical Wnt/β-catenin signaling and downstream factors, such as Apc (Hasegawa et al., 2002) and Twist1, a transcription factor required for cell migration and osteoprogenitor cell (OPC) specification from mesenchymal precursors (Tischfield et al., 2017; Goodnough et al., 2012). As cells migrate into the supraorbital ridge, they spatially segregate and form a distinct neural crest-mesoderm boundary (NeuCMesB) (Morriss-Kay and Wilkie, 2005; Holmes, 2012). Anterior to the eye, the neural crest comprises the rostral domain and gives rise to the frontal bone (Doro et al., 2019; Jiang et al., 2002). The caudal domain is composed of mesoderm-derived cells that form the parietal bone (Yoshida et al., 2008) (Fig. 2B).

Fig. 2.

Osteoblast differentiation in the supraorbital mesenchyme. (A) Molecular regulation of the osteoblast lineage. Mesenchymal stem cells (MSCs) give rise to osteoprogenitor cells (OPCs) and then to committed preosteoblasts, which differentiate into postmitotic osteoblasts. (B) Following migration into the supraorbital ridge, neural crest (NeuC, green)-derived and mesoderm (Mes, yellow)-derived OPCs form and condense in the supraorbital mesenchyme (SOM) between ∼E10.5 and E12.5. (C,D) Morphogen signaling drives differentiation of the osteoblast lineage. Wnt signaling (C) from the surface ectoderm is necessary for the specification of OPCs from uncommitted MSCs. β-Catenin activates Twist1 in OPCs, which prevents chondrogenesis by inhibiting Sox9 activity, while also maintaining OPC responsiveness to autocrine Wnt signaling. BMP signaling (D) activates Runx2, which drives Osx1 expression in committed preosteoblasts. Once Twist1 levels decline (arrowhead), the brake on Runx2 function is relieved (dashed lines), allowing upregulation of Osx1 expression. Osx1 is required for differentiation towards mature osteoblasts.

Fig. 2.

Osteoblast differentiation in the supraorbital mesenchyme. (A) Molecular regulation of the osteoblast lineage. Mesenchymal stem cells (MSCs) give rise to osteoprogenitor cells (OPCs) and then to committed preosteoblasts, which differentiate into postmitotic osteoblasts. (B) Following migration into the supraorbital ridge, neural crest (NeuC, green)-derived and mesoderm (Mes, yellow)-derived OPCs form and condense in the supraorbital mesenchyme (SOM) between ∼E10.5 and E12.5. (C,D) Morphogen signaling drives differentiation of the osteoblast lineage. Wnt signaling (C) from the surface ectoderm is necessary for the specification of OPCs from uncommitted MSCs. β-Catenin activates Twist1 in OPCs, which prevents chondrogenesis by inhibiting Sox9 activity, while also maintaining OPC responsiveness to autocrine Wnt signaling. BMP signaling (D) activates Runx2, which drives Osx1 expression in committed preosteoblasts. Once Twist1 levels decline (arrowhead), the brake on Runx2 function is relieved (dashed lines), allowing upregulation of Osx1 expression. Osx1 is required for differentiation towards mature osteoblasts.

After migration, from approximately E10.5 to E12.5, mesenchymal cells within the SOM undergo a bone lineage program to become specified into OPCs (Ferguson and Atit, 2019) (Fig. 2A). This process is dependent upon expression of Msx1 and Msx2, genes encoding homeobox transcription factors that control both OPC proliferation and differentiation (Liu et al., 1999; Satokata et al., 2000). The expression of these transcription factors requires fibroblast growth factors (FGFs) (Kim et al., 1998; Ignelzi et al., 2003) and the secretion of bone morphogenic proteins (BMPs) from the surrounding calvarial mesenchyme and meninges (Bonilla-Claudio et al., 2012; Rice et al., 2003). In mice, Msx1 and Msx2 are co-expressed in the frontal bone anlage and exhibit redundant functions during this period, whereas the parietal bone anlage solely expresses Msx2 at E12.5 (Han et al., 2007), as some molecular differences exist given their cell ontology.

The specification of osteogenic mesenchyme is dependent on paracrine Wnt signaling from the adjacent surface ectoderm. In mice, calvarial OPCs must be exposed to a gradient of Wnt10 and Wnt7b in order to sustain their osteogenic identity (Goodnough et al., 2014). Twist1 is required to transduce Wnt signaling and maintain Wnt responsiveness in the SOM (Goodnough et al., 2016). Subsequently, OPCs produce Wnt3a, Wnt5a, Wnt11 and Wnt16 in an autocrine manner to specify committed preosteoblasts and prevent cartilage formation through canonical Wnt and extracellular signal-regulated kinase (ERK) 1/2 signaling (Goodnough et al., 2014; Ibarra et al., 2021) (Fig. 2C).

As OPCs undergo specification, they express Twist1, along with Dlx5, low levels of alkaline phosphatase (ALP, encoded by Alpl) and Runx2, which is a master regulator and one of the earliest known markers for OPC specification (Komori et al., 1997; Otto et al., 1997; Ishii et al., 2003; Robledo et al., 2002). Regulating Runx2 activity is important to prevent premature differentiation into postmitotic osteoblasts. Twist1 is crucial for this process as it heterodimerizes with Runx2, inhibiting its ability to bind to DNA (Bialek et al., 2004). Once Twist1 expression declines, the brake on Runx2 function is relieved, permitting the differentiation of Osx1 (Sp7)/Runx2+ preosteoblasts towards more mature bone sialoprotein (Bsp or Ibsp)- and osteocalcin (Bglap)-expressing osteoblasts (Bialek et al., 2004) (Fig. 2A,D). Thus, Twist1 is necessary for specifying OPCs from mesenchymal progenitors and also controls the timing of osteoblast differentiation, ensuring the pool of proliferating OPCs is not prematurely depleted.

Once OPCs begin to upregulate Osx1 expression, which is activated by Runx2, they transition into a more mature preosteoblast state (Nishio et al., 2006). In both mice and zebrafish, Osx1 is essential for early osteoblast lineage commitment and the eventual differentiation of preosteoblasts into mature bone matrix-producing osteoblasts (Nakashima et al., 2002; Kague et al., 2016; Xu et al., 2018) (Fig. 2D). Loss of Osx1 in zebrafish causes excessive proliferation of OPCs and impedes their differentiation, leading to the formation of ectopic bones that take on the appearance of Wormian bones (Kague et al., 2016). In the absence of mouse Osx1, Runx2+ preosteoblasts begin to express chondrocyte-specific markers, such as Sox9, suggesting that Osx1 represses the differentiation of cartilage (Nakashima et al., 2002).

Supraorbital mesenchyme condensation

By E12.5 in mice, OPCs aggregate and form condensation centers in the SOM (Hall and Miyake, 2000) (Fig. 2B), which subsequently expand to form the flat bones. The initiation of condensation is influenced by the timing of OPC differentiation and maintenance of the progenitor pool. To facilitate the growth and maturation of the skull vault, OPCs must self-renew and gradually differentiate to ensure the progenitor pool is not prematurely depleted. Maintaining this balance requires the activities of multiple signaling molecules, including BMPs, transforming growth factor β (TGFβ) family proteins, Hedgehog (Hh) proteins, retinoic acid (RA) and Notch, which crosstalk to pattern the skull (Bonilla-Claudio et al., 2012; Li et al., 2015; Wurdak et al., 2005; Dong et al., 2010; Jeong et al., 2004; Pakvasa et al., 2021). Together, these signals set the equilibrium between cell proliferation and differentiation. For example, studies in mice have shown that BMP signaling is required for proper Dlx5, Runx2 and Osx1 expression, and failure to upregulate these transcription factors impedes osteoblast differentiation and calvarial growth (Nakashima et al., 2002; Miyama et al., 1999; Phimphilai et al., 2006; Baek et al., 2014) (Fig. 2D). Conversely, although the Notch mediator RBP-Jκ (Rbpj) and downstream effector Hey1 promote the proliferation of murine preosteoblasts via upregulation of Osx1 expression, they also suppress Runx2 activity to prevent premature differentiation into osteocalcin-expressing osteoblasts (Engin et al., 2008). Thus, there is likely crosstalk between Notch and BMP signaling (Pakvasa et al., 2021).

Dysregulation of these signaling pathways can impede mouse calvarial development by inducing premature OPC differentiation, thereby depleting the pools of OPCs and preventing SOM condensation (Singh et al., 2013; Maeno et al., 2011; Sun et al., 2013). For example, although murine BMP signaling is necessary to initiate the expression of Msx1, Msx2 and Runx2 (Rice et al., 2000, 2005; Beederman et al., 2013), preosteoblasts require Foxc1 to maintain the appropriate level of BMP responsiveness (Rice et al., 2003) (Fig. 2A). When Foxc1 is lost in mice, OPCs become overly sensitive to BMP signaling and upregulate Msx2 expression (Rice et al., 2003). Consequently, OPCs prematurely differentiate, which depletes the progenitor pool and inhibits condensation in SOM (Sun et al., 2013).

Apical expansion of calvarial progenitors

As mesenchymal condensations form, they expand bilaterally from the base of the skull to the apex of the head such that by ∼E15, the opposite edges of the paired frontal and parietal bones rudiments are juxtaposed (Fig. 3A). During expansion, growth factors and morphogens, including Hh and autocrine BMP signaling, control the proliferation of preosteoblasts at the leading edge, called the osteogenic front (OF) (Kim et al., 1998; Lenton et al., 2011; Veistinen et al., 2017; Jiang et al., 2019). Proliferating preosteoblasts express Osx1 and high levels of Runx2 (Ishii et al., 2015; Yoshida et al., 2008). They are responsive to FGF signaling, which is required for their proliferation and differentiation towards osteoblasts. Fgfr2 is required for proliferation at the leading edge (Iseki et al., 1999), whereas Fgfr1 controls the differentiation of the more mature osteoid-laying osteoblasts behind the OF (Johnson et al., 2000; Iseki et al., 1999) (Fig. 2A, Fig. 3B). Bordering these groups lies a population of cells that transiently express FGFR3 and are transitioning to the mature osteoblast state (Iseki et al., 1999). Activating mutations in FGFR3 affect calvarial development in mice and humans (Di Rocco et al., 2014; Nah et al., 2012), although less is known about the developmental functions of FGFR3 versus those of FGFR2. In zebrafish, the OF expresses Fgfr3 in a manner similar to that of mice (Ledwon et al., 2018) and FGFR3 is necessary for calvarial growth, osteoblast differentiation and suture formation (Dambroise et al., 2020).

Fig. 3.

Apical expansion and growth of the calvarium. (A) Between E12.5 and E15.5, the frontal bone (Fb) and parietal bone (Pb) primordia form as condensations in the supraorbital mesenchyme expand towards the apex of the head. (B) Apical expansion is influenced by the proliferation of preosteoblasts at the osteogenic front. Proliferating preosteoblasts require FGF2 signaling, which also serves to upregulate FGFR2 expression via Twist1. As cells differentiate, one daughter cell continues to divide at the leading edge, whereas the other exits the cell cycle to differentiate into a FGFR1+ osteoblast. Note that a small portion of the frontal bone contains coronal suture-derived osteoblasts, as shown in this example, whereas the coronal suture makes a major contribution to the parietal bone. Also, the periosteum is thought to be a source of osteoblasts. MSC, mesenchymal stem cell.

Fig. 3.

Apical expansion and growth of the calvarium. (A) Between E12.5 and E15.5, the frontal bone (Fb) and parietal bone (Pb) primordia form as condensations in the supraorbital mesenchyme expand towards the apex of the head. (B) Apical expansion is influenced by the proliferation of preosteoblasts at the osteogenic front. Proliferating preosteoblasts require FGF2 signaling, which also serves to upregulate FGFR2 expression via Twist1. As cells differentiate, one daughter cell continues to divide at the leading edge, whereas the other exits the cell cycle to differentiate into a FGFR1+ osteoblast. Note that a small portion of the frontal bone contains coronal suture-derived osteoblasts, as shown in this example, whereas the coronal suture makes a major contribution to the parietal bone. Also, the periosteum is thought to be a source of osteoblasts. MSC, mesenchymal stem cell.

Cell proliferation drives calvarial growth as mesenchymal condensations expand. Studies in mice suggest that the rate of expansion is partially determined by the number of OPCs in the condensed SOM, and subsequently by the numbers of proliferating preosteoblasts within the OF (Lana-Elola et al., 2007). Preosteoblasts proliferate through asymmetric cell division. In response to FGF2 signaling, one daughter cell continues to divide at the leading edge, whereas the other remains behind, exiting the cell cycle to differentiate into a more mature FGFR1+ osteoblast (Lana-Elola et al., 2007; Iseki et al., 1999). Additionally, FGF2 upregulates the expression of Twist1 in murine OPCs (Rice et al., 2000). In a positive feedback loop, Twist1 upregulates Fgfr2 expression to prime the sensitivity of OPCs to FGF2 (Johnson et al., 2000) (Fig. 3B). Interestingly, the protocadherin proteins Fat4 and Dchs1 also influence the proliferation of OPCs directly, as well as indirectly by controlling the relative balance of Yap (Yap1) and Taz (Tafazzin) (Crespo-Enriquez et al., 2019). In the absence of Fat4-Dchs1 signaling, levels of Yap become aberrantly high in OPCs, leading to excess formation of Yap-Runx2 complexes. This alters Runx2 activity, which expands the osteoprogenitor pool and increases proliferation at both the frontal and parietal bone OFs, while delaying the differentiation of Osx1+ preosteoblasts (Crespo-Enriquez et al., 2019).

Despite these data, the various mechanisms that control expansion, including cell proliferation and migration, are still somewhat unclear. Notably, lineage tracing in the SOM using an inducible En1-CreER driver found some labeled progenitors beyond the OF (Tran et al., 2010; Deckelbaum et al., 2012). These findings are supported by lipophilic tracer experiments tracking the migration of SOM progenitors (Yoshida et al., 2008). Additionally, although administration of nocodazole inhibited proliferation at the parietal bone OF by 44%, the reduction of parietal bone size was only 19% (Lana-Elola et al., 2007), suggesting that other unidentified mechanisms contribute to calvarial growth.

Canonical Wnt signaling

Calvarial OPCs are technically osteochondroprogenitor cells, possessing the ability to differentiate into both osteoblasts and chondrocytes in mice (Hill et al., 2005; Nakashima et al., 2002). However, calvarial OPCs directly differentiate into osteoblasts through intramembranous ossification, a process distinct from endochondral ossification of the long bones, which requires a cartilage intermediate. Canonical Wnt/β-catenin signaling in mice is required to repress chondrocyte differentiation (Hill et al., 2005; Day et al., 2005) by activating Twist1, which suppresses Sox9 function by binding to the 3′ UTR of the Sox9 gene (Goodnough et al., 2012) (Fig. 2C). Sox9 is a master regulator for chondrocyte differentiation and an inhibitor of Runx2 (Mori-Akiyama et al., 2003; Hill et al., 2005). Ablating β-catenin in murine OPCs using En1-Cre results in loss of Twist1 expression, upregulation of Sox9 activity and the conversion of flat bones to cartilage (Goodnough et al., 2012). Notably, balancing the levels of β-catenin activity is also essential for intramembranous ossification. For example, β-catenin upregulates Runx2 and Osx1 expression in committed preosteoblasts, and abolishing its activity perturbs osteoblast differentiation and skull development (Rodda and McMahon, 2006). Conversely, expressing a constitutively activated form of β-catenin inhibits osteoblast differentiation in mice by prolonging Twist1 expression, and ablating Twist1 expression restores normal skull development (Goodnough et al., 2012).

FGF signaling

Intramembranous ossification is also sensitive to the levels of FGF signaling because activating mutations in FGFR1/2 can antagonize canonical Wnt/β-catenin signaling (Ambrosetti et al., 2008). Sox2, a downstream target of FGF signaling, is expressed by preosteoblasts in the OF and can inhibit β-catenin activity to downregulate target genes (Mansukhani et al., 2005). Forced Sox2 expression inhibits the differentiation of osteoblasts and causes parietal bones to become thin and immature (Holmes et al., 2011). Additionally, an excess of certain FGFs, such as FGF8 and FGF9, induces endochondral ossification and severe calvarial dysplasia (Schmidt et al., 2018; Govindarajan and Overbeek, 2006). Ectopic FGF9-driven endochondral ossification is partially rescued by reducing the activity of the Wnt inhibitor Axin2 (Schmidt et al., 2018). Although the exact mechanisms behind ectopic calvarial endochondral ossification are not well understood, excess platelet-derived growth factor α (PDGFRα, encoded by Pdgfra) signaling may drive this process. Humans with Apert syndrome and activating mutations in FGFR2 have increased levels of PDGFRα and epidermal growth factor (EGF) signaling in the coronal suture (Miraoui et al., 2010). Furthermore, recent work in Apert mice has revealed that ectopic cartilage forms in place of coronal sutures and constitutive activation of PDGFRα signaling can induce suture closure through endochondral ossification (Peskett et al., 2017; He and Soriano, 2017).

Retinoic acid

Intramembranous ossification also requires protection from excessive exposure to all-trans RA, a vitamin A-derived metabolite that controls cell fate decisions (Morriss-Kay, 1993). As the meninges begin to differentiate and baso-apically expand at E12.5, skull progenitors are proximal to cells expressing Rdh10/Raldh2 to produce RA in the future arachnoid membrane (Siegenthaler et al., 2009). Exogenously applied RA inhibits calvarial development and induces the formation of ectopic cartilage (Jiang et al., 2002). In addition, humans, zebrafish and mice with inactivating mutations in the Cyp26b1 gene, which encodes a RA-degrading enzyme, exhibit skull dysplasia and abnormal calvarial development (Spoorendonk et al., 2008; Maclean et al., 2009; Laue et al., 2011). Taken together, these findings suggest that OPCs may fail to become committed preosteoblasts when exposed to elevated levels of RA. In support, in vitro studies have shown that RA antagonizes Wnt signaling and inhibits osteoblast differentiation (Lind et al., 2017; Roa et al., 2019).

Contributions from periosteum

Periosteal stem cells (PSCs) have recently been shown to be a source of calvarial osteoblasts in normal development (Debnath et al., 2018) (Fig. 3B). The periosteum has traditionally been viewed to be a source of osteoblasts in the context of fractures and other bone injuries (Utvåg et al., 1996; van Gastel et al., 2012), and the molecular identities of periosteal cells contributing to bone healing are unclear. However, cathepsin K (Ctsk)-positive PSCs have been identified in the calvarial periosteum and in cranial sutures of mice, and directly form bone through intramembranous ossification (Debnath et al., 2018). When Osx1 is deleted via Ctsk-Cre, these cells fail to differentiate into osteoblasts, leading to hypoplastic and hypomineralized calvarial bones, and enlarged cranial sutures (Debnath et al., 2018).

The dura, the outermost of the three membranes of the brain and spinal cord, produces morphogens that regulate calvarial growth. Starting at E9.5 in mice, the cranial mesenchyme begins to express Foxc1 (Inman et al., 2013), which is required for the growth and expansion of the dura and the calvarium (Mishra et al., 2016; Vivatbutsiri et al., 2008). In parallel with the apical expansion of skull progenitors, the dura differentiates in a baso-apical manner through a process dependent on Foxc1 and Wnt signaling (Zarbalis et al., 2007; DiNuoscio and Atit, 2019). The dura secretes growth factors, such as FGF2, BMP4 and TGFβ1 (Tgfb1), all of which contribute to osteogenesis (Machida et al., 2014; Mehrara et al., 1999). TGFβ signaling acts upstream of FGF signaling and induces FGFR2 expression in calvarial OPCs (Sasaki et al., 2006). When canonical (SMAD-dependent) TGFβ signaling is lost, calvarial OPCs shift to non-canonical (SMAD-independent) signaling and fail to properly develop (Ito et al., 2003; Iwata et al., 2012). Shifting the balance back to canonical TGFβ signaling can recover calvarial development (Ho et al., 2015). In postnatal mice, FGF2 is enriched in the dura and promotes osteogenesis (Li et al., 2007, 2010; Ogle et al., 2004). Upon FGF2 stimulation, dural fibroblasts proliferate, secrete TGFβ1/TGFβ3 (Tgfb3) and produce even more FGF2 to induce osteogenesis (Spector et al., 2002).

The dura also regulates suture patency during both embryonic and postnatal development. In vitro studies with E19 rat calvaria showed that coronal suture patency is dependent on dural-derived heparin-binding factors (Opperman et al., 1996). When the dura is removed from fetal rat calvaria, coronal sutures fuse (Opperman et al., 1993, 1995; Opperman, 2000; Ogle et al., 2004). Additional experiments have induced coronal suture fusion by removing TGFβ3 activity, whereas fusion can be prevented by removing TGFβ2 (Tgfb2) activity (Opperman et al., 2000). In a mouse model of intrauterine constraint, elevated levels of TGFβ1 and TGFβ3 correlate with increased incidence of suture closure (Kirschner et al., 2002). Additionally, the application of exogenous TGFβ2 in E15.5 mouse calvarial explants induces sagittal suture fusion via activation of ERK-MAPK signaling (Lee et al., 2006). In postnatal rodents, TGFβ and FGF signaling from the dura is regionally specified and the relative levels dictate suture fusion. For example, the coronal, lambdoid and sagittal sutures remain patent throughout adulthood, but the posterior frontal suture fuses in rodents by the third week (Hermann et al., 2013; Most et al., 1998). In comparison with the sagittal suture, the dura associated with the posterior suture shows higher levels of TGFβ1-3 and FGF2 expression (Roth et al., 1997; Most et al., 1998; Greenwald et al., 2000). Co-culture experiments have shown that the dura derived from the posterior frontal suture has a greater capacity to induce osteogenic gene expression versus that of sagittal suture dura (Warren et al., 2003b). Furthermore, autologous transplantation experiments in which the posterior frontal and sagittal sutures were surgically resected and swapped caused sagittal suture fusion, whereas the posterior frontal suture remained patent (Levine et al., 1998).

Coronal suture progenitor cell populations

Coordinated growth of the calvarial anlagen is necessary to form and position cranial sutures, which develop between the neighboring OFs of the paired frontal and parietal bones (Teng et al., 2019; Richtsmeier and Flaherty, 2013). The coronal suture serves as a model for understanding sutural development and patency, and the cell-autonomous and non-cell-autonomous mechanisms that regulate these processes. The coronal suture develops from a subpopulation of mesoderm-derived Engrailed1 (En1)-positive cells, which appears in the SOM by E11 in mice (Deckelbaum et al., 2012). These precursors proliferate and apically migrate between E11.5 and E13.5 to form the coronal suture, which gives rise to osteoblasts that make major and minor contributions to the parietal and frontal bones, respectively (Deckelbaum et al., 2012). In the sutural mesenchyme, En1 is necessary to maintain Twist1 and repress Msx2 expression and FGFR2 signaling to prevent sutural progenitor cells from aberrantly adopting an osteogenic phenotype (Deckelbaum et al., 2012). The murine sutural mesenchyme also contains a population of Prx1 (Prrx1)/Sca1 (Ly6a) double-positive stem cells that influence intramembranous ossification (Takarada et al., 2015). Recently, single-cell sequencing experiments have identified previously unreported populations of sutural progenitor cells and markers for specific subpopulations at the OF (Box 1). For intramembranous ossification to occur, these cells upregulate Runx2 expression and become recruited into the OF (Takarada et al., 2015).

Box 1. Suture progenitors and markers of the osteogenic front

Hhip+ progenitors have been specifically identified in the coronal suture and help control patterning (Holmes et al., 2021). RNA sequencing of the sutural mesenchyme during late gestation revealed a previously unreported population of Erg+/Pthlh+/Six2+ skull progenitors (Farmer et al., 2021). Interestingly, although both populations give rise to osteoprogenitor cells (OPCs) that contribute to the frontal and parietal bones, Hhip+ cells preferentially populate the parietal bones, whereas Erg+/Pthlh+/Six2+ cells tend to produce frontal bone OPCs. The same study found a subpopulation of dural cells with chondrogenic potential and a subpopulation of ligament-like cells that connects the frontal and parietal bones in both fetal and adult stages (Farmer et al., 2021). Importantly, this study identified previously unappreciated subpopulations of cells in the ectocranial mesenchyme, which might mediate the formation of the neural crest-mesoderm boundary. Single-cell RNA sequencing has also revealed four new subpopulations in the sutural mesenchyme and osteogenic front in the murine posterior frontal suture, which is distinct because it fuses postnatally (Holmes et al., 2020).

Regulation of cell-lineage and cell-fate boundaries

En1 is necessary to maintain the coronal suture by establishing a NeuCMesB (Fig. 4A), thereby preventing the invasion of neural crest cells into the sutural mesenchyme and mesoderm (Deckelbaum et al., 2012; Merrill et al., 2006). Neural crest-derived cells secrete pro-osteogenic growth factors (i.e. FGFs) and/or have the propensity to differentiate into bone, which can cause fusion. The NeuCMesB is also enforced by Twist1 and Eph-ephrin signaling. Through genetic interactions with Twist1, ephrin A2 (Efna2)- and ephrin A4 (Efna4)-expressing cells in the non-osteogenic mesenchyme lining the frontal bone maintain the separation of the cranial neural crest and mesoderm via repulsive interactions with EphA4+ cells in the sutural mesenchyme (Ting et al., 2009; Merrill et al., 2006). Twist1 also maintains the NeuCMesB by regulating jagged 1 (Jag1) expression. In turn, Jag1 maintains the mesenchymal architecture of the suture and inhibits osteogenesis by attenuating the levels of SMAD, ERK, Notch and Wnt signaling (Yen et al., 2010).

Fig. 4.

Suture morphology and pathogenesis of craniosynostosis. (A) The frontal bone (Fb), which is predominately derived from the neural crest (NeuC), and the coronal suture, which is derived from the mesoderm (Mes), form a natural neural crest-mesoderm boundary (NeuCMesB) in the skull. Note that a small portion of the frontal bone contains osteoblasts derived from the coronal suture, as shown in this example, so the boundary is unidirectional. Within the suture, a cell-fate border also exists that delineates uncommitted mesenchymal stem cells (MSCs) from cells undergoing specification into OPCs via exposure to morphogen signaling. Examples of various mechanisms postulated to cause suture fusion are shown. (B) Craniosynostosis. (Ba) Twist1 maintains the boundary between the NeuC and the Mes in the coronal suture by mediating repulsive EphA-ephrin A signaling. When these repulsive interactions are lost, the NeuC invades the OF into the sutural mesenchyme, causing ossification at the midline. (Bb) Twist1 haploinsufficiency can also cause premature differentiation of MSCs into OPCs at the sutural midline, depleting the pool of MSCs and leading to ossification. (Bc) Gain-of-function mutations in FGFR2 increase the proliferation of preosteoblasts at the osteogenic OF, causing the OF to encroach on the midline sutural mesenchyme. CS, coronal suture; Pb, parietal bone; OB, osteoblast; OF, osteogenic front.

Fig. 4.

Suture morphology and pathogenesis of craniosynostosis. (A) The frontal bone (Fb), which is predominately derived from the neural crest (NeuC), and the coronal suture, which is derived from the mesoderm (Mes), form a natural neural crest-mesoderm boundary (NeuCMesB) in the skull. Note that a small portion of the frontal bone contains osteoblasts derived from the coronal suture, as shown in this example, so the boundary is unidirectional. Within the suture, a cell-fate border also exists that delineates uncommitted mesenchymal stem cells (MSCs) from cells undergoing specification into OPCs via exposure to morphogen signaling. Examples of various mechanisms postulated to cause suture fusion are shown. (B) Craniosynostosis. (Ba) Twist1 maintains the boundary between the NeuC and the Mes in the coronal suture by mediating repulsive EphA-ephrin A signaling. When these repulsive interactions are lost, the NeuC invades the OF into the sutural mesenchyme, causing ossification at the midline. (Bb) Twist1 haploinsufficiency can also cause premature differentiation of MSCs into OPCs at the sutural midline, depleting the pool of MSCs and leading to ossification. (Bc) Gain-of-function mutations in FGFR2 increase the proliferation of preosteoblasts at the osteogenic OF, causing the OF to encroach on the midline sutural mesenchyme. CS, coronal suture; Pb, parietal bone; OB, osteoblast; OF, osteogenic front.

Twist1 also maintains suture patency through its ability to heterodimerize with other basic helix-loop-helix (bHLH) proteins. In mice and human cell lines, Twist1 can signal by forming homodimers (T/T) or heterodimers with bHLH E proteins (T/E) (Fan et al., 2020; Mikheeva et al., 2019). T/T homodimers stimulate osteogenesis by upregulating the levels of Fgfr2 and periostin (Postn) and are preferentially localized to the murine OF, whereas T/E heterodimers are enriched in the mid-sutural mesenchyme (Connerney et al., 2006). Thus, Twist1 helps maintain a cell-fate border within the sutural mesenchyme – distinct from the NeuCMesB – that separates the OF from the mid-sutural mesenchyme (Fig. 4A). In mice, dimerization is influenced by the relative abundance of Twist1 and the bHLH E proteins, but also by Ids, which are bHLH proteins that preferentially bind to E proteins and disrupt the formation of T/E heterodimers by outcompeting Twist1 (Connerney et al., 2008). Thus, at sites of Id enrichment, such as the OF, T/T homodimers predominate and promote the proliferation of preosteoblasts by upregulating Fgfr2 expression (Connerney et al., 2008). By contrast, because Id expression is low in the murine sutural mesenchyme, T/E dimers form and repress responsiveness to pro-osteogenic signaling (Connerney et al., 2006). Thus, controlling the balance of Twist1, bHLH E proteins and Ids is crucial for maintaining suture patency and growth at the OF.

Twist1 expression in the sutural mesenchyme is controlled by BCL11B (Kyrylkova et al., 2016), a zinc finger protein that is enriched in the cranial sutures throughout mid-to-late calvarial development (Holmes et al., 2015). BCL11B is switched on at E14.5 in the coronal suture, whereas its expression is initiated at E16.5 in the sagittal, interfrontal and lambdoid sutures (Holmes et al., 2015). Similar to Twist1 haploinsufficient animals discussed below, Bcl11b knockout mice develop craniosynostosis and show upregulated expression of Runx2 and Fgfr2 in the sutural mesenchyme, and also increased cell proliferation at OFs (Kyrylkova et al., 2016; Rice et al., 2000; Bialek et al., 2004).

Postnatal sutures contain heterogeneous populations of multipotent mesenchymal stem cells (MSCs) (Doro et al., 2017). Sutural MSCs have overlapping, but also distinct, molecular signatures and show various capacities to differentiate into bone, cartilage and dura. The ontological relationships between sutural MSCs are unclear and it is unknown whether residing in spatially distinct locations within the sutural mesenchyme is important for the determination of cell fate. Cell ablation studies and findings from craniosynostosis models, however, have shown that sutural MSCs are required for postnatal growth of the calvarium, injury repair and suture patency (Wilk et al., 2017; Zhao et al., 2015). Thus, sutural MSCs promote skull expansion while the brain grows in adolescents, and later help maintain bone homeostasis and turnover in adults.

Sutural MSCs are characterized by unique or overlapping expression of Gli1, Axin2, Prx1 and CD51 (Itgav)/CD200 (Maruyama et al., 2016; Zhao et al., 2015; Di Pietro et al., 2020; Wilk et al., 2017; Menon et al., 2021). Several populations demonstrate clonal expansion, long-term self-renewal and the ability to differentiate (Maruyama et al., 2016; Wilk et al., 2017; Zhao et al., 2015). Most sutural MSCs express Gli1, whereas subsets also co-express Axin2 and Prx1 (Zhao et al., 2015; Di Pietro et al., 2020; Wilk et al., 2017). A landmark study in 2015 showed that Gli1-lacZ mice exhibit lacZ expression in the periosteum, dura and sutural mesenchyme of newborns (Zhao et al., 2015). Gli1-lacZ expression gradually becomes restricted to the sutural mesenchyme, such that by 1 month of age, lacZ-expressing cells are mainly detected at the midline, with expression absent in Runx2+/Osx1+ preosteoblasts at the OF (Zhao et al., 2015). This pattern holds consistent for most adult sutures. Interestingly, Gli1-lacZ cells are absent in the posterior frontal suture, the only suture that fuses in rodents. Lineage labeling has further revealed that Gli1+ sutural MSCs give rise to the sutural mesenchyme, as well as osteoblasts, the periosteal dura and osteoclasts, with the entire skull and surrounding tissue labeled after several months. Following injury and in response to IHH signaling, Gli1+ cells can mobilize to the region of repair and differentiate into the periosteum, dura and osteocytes (Zhao et al., 2015). Thus, Gli1+ sutural MSCs represent an essential pool of multipotent MSCs responsible for calvarial growth and turnover.

In mice, a subset of Gli1+ sutural MSCs also expresses Prx1 and differentiates into both chondrocytes and osteoblasts in vitro (Ouyang et al., 2014; Wilk et al., 2017). During mouse embryogenesis, Prx1 is expressed in the osteogenic mesenchyme and is required for the development of both the cranial neural crest-derived and paraxial mesoderm-derived calvarial bones as well as the intervening sutures (Ouyang et al., 2014; Wilk et al., 2017). Ablating these cells in the mouse postnatal sutural mesenchyme, however, does not stunt calvarial growth, but rather the ability of the skull to repair itself following injury. Furthermore, these cells are absent in the calvarial periosteum and dura and are distinct from Osx1-expressing preosteoblasts, and their abundance declines with age (Wilk et al., 2017). Prx1+/Gli1+ MSCs, therefore, appear to constitute a distinct lineage that can differentiate into osteoblasts under specific contexts (i.e. injury) and environmental conditions (i.e. exposure to Wnt signaling).

Similar to Gli1+ MSCs, Axin2+ MSCs are found at the midline of all sutures except the posterior frontal suture in mice (Maruyama et al., 2016). Axin2+ MSCs have much in common with Gli1+ MSCs, although they are a more restricted subset, making up ∼15% of mesenchymal cells. Axin2 expression is localized to the sutural midline by the first postnatal week and is maintained in adults, matching the spatiotemporal profile of Gli1+ MSCs (Maruyama et al., 2016; Zhao et al., 2015). Lineage labeling has shown that Axin2+ MSCs can differentiate into osteoblasts that are incorporated throughout the calvarium, as well as osteocytes (Maruyama et al., 2016). Following injury, Axin2-lacZ cells surround the repair site, and their descendants express markers for both preosteoblasts (Osx1) and osteoclasts (Sost), suggesting their involvement in bone homeostasis and repair. Transplantation of purified Axin2-lacZ-expressing MSCs into the site of injury hastens the healing process and results in bone engraftment (Maruyama et al., 2016). Thus, Axin2+ MSCs appear to be closely related to Gli1+ MSCs and constitute a long-lasting source of sutural MSCs that build and maintain the calvarium.

Premature fusion of the cranial sutures is called craniosynostosis, a disorder that affects one in ∼2000-2500 live births (Lajeunie et al., 1995; Boulet et al., 2008; Miller et al., 2017). Craniosynostosis can be idiopathic or occur as part of a defined genetic syndrome. Although all sutures have the potential to fuse prematurely, the sagittal suture is the most commonly affected in idiopathic craniosynostosis (Stanton et al., 2022). Close to 80 genes have been implicated in various forms of craniosynostosis (Twigg and Wilkie, 2015; Goos and Mathijssen, 2019). Although clinical manifestations of craniosynostosis are most commonly identified after birth, many – but not all – of the pathophysiological mechanisms seem to affect processes that regulate early suture patterning. As such, genes involved in craniosynostosis tend to be enriched in signaling pathways that control the proliferation and differentiation of OPCs, as well as the maintenance of the NeuCMesB (Stanton et al., 2022).

Twist1 signaling

Saethre–Chotzen syndrome, caused by heterozygous missense and loss-of-function mutations (haploinsufficiency) in Twist1, is the second leading cause of syndromic craniosynostosis and primarily affects the coronal suture (el Ghouzzi et al., 1997). Findings suggest a dual paradigm wherein the coronal suture first assumes a partially osteogenic character, affecting the cell-fate border in the mesoderm, before allowing the invasion of neural crest cells into the mid-sutural mesoderm, obliterating the NeuCMesB (Ishii et al., 2015) (Fig. 4A,Ba). In Twist1+/− mice, the coronal suture displays ectopic pro-osteogenic gene expression, including Fgfr2, Alpl and Msx2 (Merrill et al., 2006; Rice et al., 2000). As early as E12.5, an aberrant osteogenic character is evident, as detected by Notch2 expression (Yen et al., 2010). Twist1 haploinsufficiency is postulated to shift the balance between T/E and T/T dimers, causing T/T dimers to form in the mid-sutural mesenchyme, thereby promoting pro-osteogenic gene expression and cell fate and depleting uncommitted MSCs (Firulli et al., 2005; Connerney et al., 2006, 2008) (Fig. 4Bb). Inhibiting the expression of Fgfr2 in the sutural mesenchyme can rescue craniosynostosis, suggesting that this counteracts pro-osteogenic gene induction. Furthermore, it is suggested that missense mutations in Twist1 may affect its ability to repress Runx2 activity, promoting premature osteoblast differentiation and suture fusion (Bialek et al., 2004; Yoshida et al., 2005).

Several Twist1-associated genes and downstream effectors have also been implicated in craniosynostosis, pointing to the loss of the NeuCMesB as another underlying mechanism (Ting et al., 2009; Kamath et al., 2002; Sharma et al., 2013). Mutations in EFNA4 cause sagittal synostosis in humans (Merrill et al., 2006) and mutation of the gene encoding the binding partner of EFNA4, Epha4, causes craniosynostosis in mice (Ting et al., 2009). Epha4 acts downstream from Twist1 to prevent migratory neural crest-derived OPCs from invading the sutural mesenchyme, which is observed in Twist1+/− mice, and Twist1+/−;Epha4+/− compound heterozygotes display more severe craniosynostosis (Ting et al., 2009). Furthermore, it is suggested that Twist1 regulates the Notch pathway, functionally linking their respective roles in NeuCMesB formation (Yen et al., 2010). Mutations in Jag1, a constituent of the Notch pathway, cause Alagille syndrome, which can be associated with craniosynostosis (Oda et al., 1997; Kamath et al., 2002). Jag1 is expressed in the embryonic coronal suture and its inactivation leads to fusion (Yen et al., 2010).

Twist1 forms heterodimers with the bHLH E protein Tcf12. Mutations in TCF12 cause coronal craniosynostosis in humans, closely resembling Saethre–Chotzen syndrome (Fan et al., 2020; Sharma et al., 2013). Twist1+/−;Tcf12+/− compound heterozygotes display severe craniosynostosis, suggesting that these proteins physically interact to maintain the cell-fate boundary within the coronal suture (Sharma et al., 2013; Ting et al., 2022). In both mice and zebrafish, Twist1-Tcf12 heterodimers control OPC proliferation at the OF (Teng et al., 2018). When heterodimers are lost, progenitor pools at the OF proliferate excessively, whereas sutural MSCs become depleted (Teng et al., 2018) (Fig. 4Bb,Bc). Interestingly, these mice also show neural crest- and mesoderm-cell mixing within the suture, suggesting that the balance of these dimers also maintains the NeuCMesB.

FGF signaling

Gain-of-function mutations in the FGF family of receptors are among the most extensively studied, causing several forms of syndromic craniosynostosis including Apert (FGFR2), Crouzon (FGFR2), Muenke (FGFR3) and Pfeiffer syndromes (FGFR1 and FGFR2) (Sawh-Martinez and Steinbacher, 2019). Two different amino acid substitutions in Fgfr2 (S252W and P253R), which increase affinity of the mutant receptor for FGF ligands (Anderson, 1998; Chen et al., 2003; Wang et al., 2005; Yin et al., 2008), account for 99% of the cases of Apert syndrome, characterized by coronal suture fusion (Park et al., 1995; Wilkie et al., 1995). FGFR2 activates ERK1/2 signaling at the OF, controlling the proliferation and differentiation of preosteoblasts (Neben et al., 2014). In mice expressing constitutively active Fgfr2S252W, cell proliferation is increased and even extends beyond the OF to encroach the mid-sutural mesenchyme (Wang et al., 2005) (Fig. 4Bc). Using RNA interference to silence the mutant Fgfr2S252W transcript reduces ERK1/2 activity and restores patency (Shukla et al., 2007). Notably, limiting activation of FGFR2S252W to the mesoderm is sufficient to induce craniosynostosis, as neural crest-specific activation has no effect. Thus, it is proposed that coronal synostosis in Apert syndrome is not caused by failure to maintain the NeuCMesB (as cell mixing is not observed), but rather by the cell-fate border within the suture that separates the mid-sutural mesenchyme from OPCs closer to the OF (Holmes and Basilico, 2012).

BMP signaling

BMP signaling and control over osteoblast differentiation is also implicated in craniosynostosis. Loss of Bmp2/4 in the mouse cranial mesoderm affects the NeuCMesB and causes narrowing of the coronal suture (Tischfield et al., 2017). Conversely, constitutive activation of BMPR1A and downstream effectors (SMADs) in the neural crest promotes premature osteoblast differentiation and suture fusion, which is partially rescued by BMP receptor kinase inhibitors (Komatsu et al., 2013). BMP9, a potent inducer of osteogenesis, is also involved as it is more highly expressed in fused versus patent sutures (Luther et al., 2011; Song et al., 2020). Fused sutures are more responsive to BMP9 signaling, and forced expression increases osteogenic gene expression and matrix mineralization in sutural MSCs (Song et al., 2020). In addition, noggin (Nog), a BMP antagonist that is repressed by FGF2 signaling, is expressed in the dura of patent sutures. Gain-of-function mutations in FGFR2 receptors may attenuate noggin signaling, leading to upregulation of BMP activity and osteogenesis of the sutural mesenchyme (Warren et al., 2003a). Interestingly, sensory nerves, which innervate the periosteum of the coronal, sagittal and lambdoid sutures (Kosaras et al., 2009), may also play a role in maintaining suture patency. These nerves secrete BMP inhibitors, such as Fstl1, preventing excess BMP/TGFβ signaling and depletion of sutural MSCs (Tower et al., 2021).

Hedgehog signaling

Finally, Hh signaling and activation of the Gli family of transcription factors have been implicated in craniosynostosis. Humans with Greig cephalopolysyndactyly syndrome have disrupting mutations in Gli3 and some show synostosis of the lambdoid suture (Vortkamp et al., 1991; Johnston et al., 2005). Gli3-null mice exhibit lambdoid suture fusion, ectopic expression of Runx2, and imbalances between OPC proliferation and differentiation (Rice et al., 2010). Furthermore, Carpenter syndrome, an autosomal-recessive form of syndromic craniosynostosis marked by multi-suture fusion, is associated with mutations in Rab23 (Jenkins et al., 2007). A recent study showed that Rab23 is required to maintain embryonic suture patency (Hasan et al., 2020). In mice, Rab23 normally antagonizes Hh signaling through the regulation of the Smoothened (Smo) and Gli proteins, which modulate the expression of Hh target genes (Eggenschwiler et al., 2001, 2006). By repressing Gli1 and FGF signaling, Rab23 maintains normal levels of OPC proliferation and prevents excessive preosteoblast formation in sutures (Hasan et al., 2020).

Mechanical force is implicated in craniosynostosis

Skulls are subjected to mechanical forces that are transmitted to the sutures. These forces alter gene expression, affect the equilibrium between differentiation and proliferation, and influence the maintenance of sutural MSCs. For example, tensional forces in rodent calvaria can upregulate canonical TGFβ signaling and maintain patency of the posterior frontal suture (Tholpady et al., 2007). Furthermore, compressive forces can induce aberrant fusion of the sagittal suture (Oppenheimer et al., 2009, 2012; Heller et al., 2007). During embryonic development, fetal constraint can alter the expression of morphogens, such as IHH, at the OF and induce craniosynostosis (Jacob et al., 2007). Although little is known about how mechanical forces influence sutural gene expression, studies suggest that cilia might play a role in sutural mechanotransduction. Primary cilia are important for maintaining the expression of osteogenic genes in human MSCs (Hoey et al., 2012; Chen et al., 2016), and ciliopathies, such as Sensenbrenner syndrome, have been associated with craniosynostosis (Tiberio et al., 2021). Interestingly, polycystin-1 (PC1, encoded by Pkd1), a mechanotransduction protein in primary cilia, is overexpressed in patients with craniosynostosis, and PC1 induces Runx2 expression through ERK signaling (Katsianou et al., 2021).

Involvement of sutural progenitor cells in craniosynostosis

Sutural MSCs also maintain patency. Ablating postnatal Gli1+ MSCs with diphtheria toxin causes craniosynostosis and a dramatic reduction in skull growth, suggesting loss of Gli1+ MSCs may contribute to suture fusion (Zhao et al., 2015). Indeed, these cells are depleted in sutures undergoing fusion in Twist1+/− mice and are undetectable in fused sutures (Zhao et al., 2015). Similar findings have been obtained with CD51+/CD200+ MSCs, which are present in sutures of postnatal rodents, with the exception of the posterior frontal suture. In Twist1+/− mice, these cells are depleted in coronal sutures undergoing fusion (Menon et al., 2021). Furthermore, by inducing sagittal suture fusion via small molecule inhibition of TGFβ signaling, a significant decrease of CD51+/CD200+ MSCs is observed. By contrast, activating Wnt signaling via exogenous Wnt3a application in the posterior frontal suture, which normally has low levels of Wnt signaling and is devoid of CD51+/CD200+ MSCs, stimulates proliferation and increases the number of CD51+/CD200+ cells, leading to suture patency (Menon et al., 2021). This may explain why the posterior frontal suture, which normally fuses by endochondral ossification several weeks after birth, is naturally devoid of Gli1+ and CD51+/200+ MSCs (Menon et al., 2021). Thus, both cell ablation studies and findings in craniosynostosis models show that MSCs are crucial for suture patency. Moreover, sutural stem cell transplantation can restore patency in craniosynostosis mouse models (Box 2).

Box 2. Regenerative medicine for craniosynostosis

Craniosynostosis is often treated by removing the affected suture(s), followed by cranial vault remodeling, an invasive procedure causing significant blood loss (Stanton et al., 2022). Without surgical intervention, intracranial pressure (ICP) may rise and impede brain development, causing cognitive deficits. Regenerative medicine approaches via mesenchymal stem cell (MSC) transplantation into affected sutures may transform how we treat craniosynostosis. Recently, Gli1+ MSCs have been surgically implanted within an osteotomy mimicking the coronal suture in Twist1+/− mice (Yu et al., 2021). One month after surgery, Gli1+ MSCs populated the center of the bony gap, whereas some populated the osteogenic front, periosteum and dura. Six months later, the implanted cells and their progeny were detected in these tissues further away from the implantation site, with the proportion increasing by 1 year (Yu et al., 2021). Importantly, endogenous Gli1+ progenitor cells from the periosteal dura were recruited and were necessary to sustain the regenerated suture. Overall, this procedure restored skull growth in Twist1+/− mice to levels that approximated those of controls, normalized ICP, and rescued brain development and cognitive deficits (Yu et al., 2021). Transplantation of CD51+/CD200+ MSCs has also been reported to rescue coronal synostosis in Twist1+/− mice (Menon et al., 2021). Following surgical implantation of CD51+/CD200+ MSCs into an incision site along with exogenous Wnt3a protein, they recruited neighboring cells, similar to Gli1+ MSCs, and created a favorable microenvironment that was sufficient to restore the sutural mesenchyme (Menon et al., 2021).

Dural venous sinuses are large superficial cerebral veins located in the periosteal dura attached to the skull. They develop in proximity to the coronal suture and parietal bone anlagen (Tischfield et al., 2017) (Fig. 5A). As the coronal suture develops and the parietal bone expands, the dural venous sinuses are actively growing and remodeling. In mice, this process requires paracrine BMP4 signaling from preosteoblasts and dura. Inactivating Twist1 in the neural crest and the cranial mesoderm via Sm22a-Cre shifts the NeuCMesB boundary caudally and perturbs the development of the coronal suture and parietal bone by E12.75. Concurrently, the dural venous sinuses fail to properly grow and remodel, leading to vessel regression and/or hypoplasia by E14.5 (Tischfield et al., 2017) (Fig. 5B). In addition, BMP4 expression is reduced in the sutural mesenchyme and the OF. Conditional inactivation of Bmp2/Bmp4 via Sm22a-Cre also shifts the NeuCMesB boundary caudally and affects the development of the coronal suture and parietal bone, causing similar effects on the growth and remodeling of the dural venous sinuses. Moreover, inactivating BMPR2 in venous endothelial cells results in hypoplastic vessels (Tischfield et al., 2017). Thus, the skull and dura crosstalk with endothelial cells to regulate venous angiogenesis, suggesting that the growth and patterning of the calvarium and underlying blood vessels co-evolved in mammals.

Fig. 5.

Integrated development of the skull and dural venous sinuses. (A) Growth and remodeling of the dural venous sinuses requires paracrine BMP signaling from preosteoblasts and the dura. This process is dependent on Twist1 for proper generation of OPCs and expansion of the developing dura. (B) The transverse sinus develops from the primitive head vein and the associated anterior and middle capillary plexi, which are present at E11.75 in mouse (left). By E12.75 (middle), the middle plexus grows into the anterior plexus, which is located in the dura immediately adjacent to the coronal suture (CS) and the parietal bone. This forms the transverse sinus and a new drainage route, causing the primitive head vein to regress as it is no longer the primary vessel draining blood from the brain. Knocking out Twist1 activity in the dura, sutural and supraorbital mesenchyme (top right panel) stunts parietal bone development and the formation of the coronal suture. The loss of preosteoblasts and hypoplastic dura diminishes paracrine BMP signaling, leading to delayed remodeling and poorly developed transverse sinuses, one of which regresses, leading to unilateral absence of the vessel. By E13.75 (B′), the sigmoid sinus remodels in response to temporal bone (Tb) development. Both the sigmoid and transverse sinuses often fail to properly form in craniosynostosis. Fb, frontal bone; Pb, parietal bone.

Fig. 5.

Integrated development of the skull and dural venous sinuses. (A) Growth and remodeling of the dural venous sinuses requires paracrine BMP signaling from preosteoblasts and the dura. This process is dependent on Twist1 for proper generation of OPCs and expansion of the developing dura. (B) The transverse sinus develops from the primitive head vein and the associated anterior and middle capillary plexi, which are present at E11.75 in mouse (left). By E12.75 (middle), the middle plexus grows into the anterior plexus, which is located in the dura immediately adjacent to the coronal suture (CS) and the parietal bone. This forms the transverse sinus and a new drainage route, causing the primitive head vein to regress as it is no longer the primary vessel draining blood from the brain. Knocking out Twist1 activity in the dura, sutural and supraorbital mesenchyme (top right panel) stunts parietal bone development and the formation of the coronal suture. The loss of preosteoblasts and hypoplastic dura diminishes paracrine BMP signaling, leading to delayed remodeling and poorly developed transverse sinuses, one of which regresses, leading to unilateral absence of the vessel. By E13.75 (B′), the sigmoid sinus remodels in response to temporal bone (Tb) development. Both the sigmoid and transverse sinuses often fail to properly form in craniosynostosis. Fb, frontal bone; Pb, parietal bone.

Recent insights into pathways that govern the maintenance, proliferation and cell-fate decisions of the osteoblast lineage have advanced our understanding of normal skull development. In particular, single-cell RNA-sequencing has identified previously unreported subpopulations of osteoblasts, sutural stem cells and dural cells (Holmes et al., 2021, 2020; Farmer et al., 2021). Understanding how these cells function and interact to regulate calvarial development and sutural homeostasis may enhance our ability to treat craniofacial disorders. Notably, a recent human genome-wide association study identified different loci that influence skull shape and are shared between humans and mice (Roosenboom et al., 2018). Expanding on these types of analyses will be beneficial, as variations in skull patterning may also affect vascular development, central nervous system fluid balance and/or waste clearance.

The effects of craniosynostosis and elevated intracranial pressure (ICP) on cerebrospinal fluid (CSF) balance and brain waste clearance also need to be investigated. Elevated ICP impedes CSF influx along perivascular channels in the brain (i.e. the glymphatic system), which is necessary to clear waste and transport it to dural lymphatic vessels (Iliff et al., 2012; Bolte et al., 2020). Affecting this process is associated with cognitive deficits and the buildup of tau and β-amyloid plaques in mice (Da Mesquita et al., 2018; Iliff et al., 2014). Importantly, cognitive deficits are present in craniosynostosis but, in most cases, the underlying causes are unknown (Brooks et al., 2018). Thus, it will be important to test whether brain waste clearance is impeded by craniosynostosis and whether animal models have combined deficits in the glymphatic and lymphatic systems (Box 3). Although craniosynostosis is currently treated through surgical interventions (Kyutoku and Inagaki, 2017), surgeries are not always effective, especially in syndromic craniosynostosis. In addition to sutural stem cell transplantation that can restore suture patency and normalize ICP in mouse models (Box 2), recent work has shown that applying cyclical mechanical loading to the frontal bone, which generates tensile force on the coronal suture, contributes to the normalization of skull growth in Crouzon mice (Moazen et al., 2022). As opening the skull vault can impair glymphatic function (Plog et al., 2020; Hablitz and Nedergaard, 2021), these minimally invasive methods may also reduce the risk of harmful side effects.

Box 3. Neurovascular findings in craniosynostosis

Meningeal (dural) lymphatic vessels are patterned along the dural venous sinuses (Louveau et al., 2015; Aspelund et al., 2015). These vessels are required for neuroimmune surveillance and, in conjunction with the glymphatic system, the removal of waste and neurotoxins from the brain (Da Mesquita et al., 2018; Louveau et al., 2018). Dorsal lymphatic ‘hotspots’, highly branched regions specialized for immune cell and waste uptake, are located at the sinus confluence and along the transverse sinuses (Louveau et al., 2018). In addition to hypoplasia/loss of the transverse sinuses (purple vessels) in Twist1 craniosynostosis models, recent findings show that dorsal lymphatic networks (green vessels) and hotspot regions are hypoplastic and less complex as well (double arrows) (Ang et al., 2022). Failure of lymphatic networks to properly grow and expand was attributed to hypoplastic dura and/or loss of the venous sinuses, as venous smooth muscle is postulated to produce vascular endothelial growth factor C (VEGFC), which controls lymphatic growth (Antila et al., 2017; Mukherjee and Dixon, 2021; Lund et al., 2012). Loss/hypoplasia of the dural venous sinuses and jugular vein stenosis are also commonly associated with syndromic craniosynostosis (Tischfield et al., 2017; Copeland et al., 2018). Venous sinus anomalies can affect fluid balance in the head and are thought to contribute to elevated intracranial pressure (ICP), which, if not properly managed, can cause cognitive deficits and vision loss (Stanton et al., 2022; Taylor et al., 2001; Hayward and Gonsalez, 2005; Renier et al., 1982). Notably, dural lymphatic vessels are sensitive to elevated ICP. In traumatic brain injury models, these vessels undergo pathological changes that hinder waste clearance (Bolte et al., 2020). Thus, it remains possible that elevated ICP also contributes to lymphatic deficits in craniosynostosis mice and potentially in humans.

The figures were adapted from the following BioRender templates: ‘skull (lateral view)', ‘skull (superior view)’, ‘mouse skull (with mandible)’, ‘mouse skull (dorsal)’, ‘mouse brain (dorsal, with veins)’, ‘mouse embryo (E11.5)’, ‘mouse embryo (E12.5)’, ‘mouse embryo (E15.5)’, ‘mesenchymal stromal cell', ‘osteoprogenitor cell' and ‘osteoblast', by BioRender.com (2022). Retrieved from https://app.biorender.com/biorender-templates.

Funding

Funding was provided by a Busch Biomedical Grant (to M.A.T.) and the Robert Wood Johnson Foundation (74260 to M.A.T.). Partial funding was provided by a National Institutes of Health/National Institute of Dental and Craniofacial Research grant (R01DE030480 to R.R.R.). Deposited in PMC for release after 12 months.

Alaqeel
,
S.
,
Hinton
,
R.
and
Opperman
,
L.
(
2006
).
Cellular response to force application at craniofacial sutures
.
Orthod. Craniofac. Res.
9
,
111
-
122
.
Alfonso-Durrruty
,
M. P.
,
Giles
,
B. T.
,
Misarti
,
N.
,
San Roman
,
M.
and
Morello
,
F.
(
2015
).
Antiquity and geographic distribution of cranial modification among the prehistoric groups of Fuego-Patagonia, Chile: Cranial Modification in Fuego-Patagonia
.
Am. J. Phys. Anthropol.
158
,
607
-
623
.
Ambrosetti
,
D.
,
Holmes
,
G.
,
Mansukhani
,
A.
and
Basilico
,
C.
(
2008
).
Fibroblast growth factor signaling uses multiple mechanisms to inhibit wnt-induced transcription in osteoblasts
.
Mol. Cell. Biol.
28
,
4759
-
4771
.
Anderson
,
J.
(
1998
).
Apert syndrome mutations in fibroblast growth factor receptor 2 exhibit increased affinity for FGF ligand
.
Hum. Mol. Genet.
7
,
1475
-
1483
.
Ang
,
P. S.
,
Matrongolo
,
M. J.
and
Tischfield
,
M. A.
(
2022
).
The growth and expansion of meningeal lymphatic networks are affected in craniosynostosis
.
Development
149
,
dev200065
.
Antila
,
S.
,
Karaman
,
S.
,
Nurmi
,
H.
,
Airavaara
,
M.
,
Voutilainen
,
M. H.
,
Mathivet
,
T.
,
Chilov
,
D.
,
Li
,
Z.
,
Koppinen
,
T.
,
Park
,
J.-H.
et al. 
(
2017
).
Development and plasticity of meningeal lymphatic vessels
.
J. Exp. Med.
214
,
3645
-
3667
.
Aspelund
,
A.
,
Antila
,
S.
,
Proulx
,
S. T.
,
Karlsen
,
T. V.
,
Karaman
,
S.
,
Detmar
,
M.
,
Wiig
,
H.
and
Alitalo
,
K.
(
2015
).
A dural lymphatic vascular system that drains brain interstitial fluid and macromolecules
.
J. Exp. Med.
212
,
991
-
999
.
Baek
,
J.-E.
,
Choi
,
J.-Y.
and
Kim
,
J.-E.
(
2014
).
Skeletal analysis and differential gene expression in Runx2/Osterix double heterozygous embryos
.
Biochem. Biophys. Res. Commun.
451
,
442
-
448
.
Beederman
,
M.
,
Lamplot
,
J. D.
,
Nan
,
G.
,
Wang
,
J.
,
Liu
,
X.
,
Yin
,
L.
,
Li
,
R.
,
Shui
,
W.
,
Zhang
,
H.
,
Kim
,
S. H.
et al. 
(
2013
).
BMP signaling in mesenchymal stem cell differentiation and bone formation
.
JBiSE
06
,
32
-
52
.
Bialek
,
P.
,
Kern
,
B.
,
Yang
,
X.
,
Schrock
,
M.
,
Sosic
,
D.
,
Hong
,
N.
,
Wu
,
H.
,
Yu
,
K.
,
Ornitz
,
D. M.
,
Olson
,
E. N.
et al. 
(
2004
).
A twist code determines the onset of osteoblast differentiation
.
Dev. Cell
6
,
423
-
435
.
Bolte
,
A. C.
,
Dutta
,
A. B.
,
Hurt
,
M. E.
,
Smirnov
,
I.
,
Kovacs
,
M. A.
,
McKee
,
C. A.
,
Ennerfelt
,
H. E.
,
Shapiro
,
D.
,
Nguyen
,
B. H.
,
Frost
,
E. L.
et al. 
(
2020
).
Meningeal lymphatic dysfunction exacerbates traumatic brain injury pathogenesis
.
Nat. Commun.
11
,
4524
.
Bonilla-Claudio
,
M.
,
Wang
,
J.
,
Bai
,
Y.
,
Klysik
,
E.
,
Selever
,
J.
and
Martin
,
J. F.
(
2012
).
Bmp signaling regulates a dose-dependent transcriptional program to control facial skeletal development
.
Development
139
,
709
-
719
.
Boulet
,
S. L.
,
Rasmussen
,
S. A.
and
Honein
,
M. A.
(
2008
).
A population-based study of craniosynostosis in metropolitan Atlanta, 1989–2003
.
Am. J. Med. Genet.
146A
,
984
-
991
.
Brooks
,
E. D.
,
Beckett
,
J. S.
,
Yang
,
J.
,
Timberlake
,
A. T.
,
Sun
,
A. H.
,
Chuang
,
C.
and
Persing
,
J. A.
(
2018
).
The etiology of neuronal development in craniosynostosis: a working hypothesis
.
J. Craniofac. Surg.
29
,
49
-
55
.
Carter
,
R. M. S.
and
Anslow
,
P.
(
2009
).
Imaging of the Calvarium
.
Semin. Ultrasound CT MRI
30
,
465
-
491
.
Chen
,
L.
,
Li
,
D.
,
Li
,
C.
,
Engel
,
A.
and
Deng
,
C.-X.
(
2003
).
A Ser250Trp substitution in mouse fibroblast growth factor receptor 2 (Fgfr2) results in craniosynostosis
.
Bone
33
,
169
-
178
.
Chen
,
J. C.
,
Hoey
,
D. A.
,
Chua
,
M.
,
Bellon
,
R.
and
Jacobs
,
C. R.
(
2016
).
Mechanical signals promote osteogenic fate through a primary cilia–mediated mechanism
.
FASEB J.
30
,
1504
-
1511
.
Connerney
,
J.
,
Andreeva
,
V.
,
Leshem
,
Y.
,
Muentener
,
C.
,
Mercado
,
M. A.
and
Spicer
,
D. B.
(
2006
).
Twist1 dimer selection regulates cranial suture patterning and fusion
.
Dev. Dyn.
235
,
1334
-
1346
.
Connerney
,
J.
,
Andreeva
,
V.
,
Leshem
,
Y.
,
Mercado
,
M. A.
,
Dowell
,
K.
,
Yang
,
X.
,
Lindner
,
V.
,
Friesel
,
R. E.
and
Spicer
,
D. B.
(
2008
).
Twist1 homodimers enhance FGF responsiveness of the cranial sutures and promote suture closure
.
Dev. Biol.
318
,
323
-
334
.
Copeland
,
A. E.
,
Hoffman
,
C. E.
,
Tsitouras
,
V.
,
Jeevan
,
D. S.
,
Ho
,
E. S.
,
Drake
,
J. M.
and
Forrest
,
C. R.
(
2018
).
Clinical significance of venous anomalies in syndromic craniosynostosis
.
Plast. Reconstr. Surg. Glob. Open
6
,
e1613
.
Crespo-Enriquez
,
I.
,
Hodgson
,
T.
,
Zakaria
,
S.
,
Cadoni
,
E.
,
Shah
,
M.
,
Allen
,
S.
,
Al-Khishali
,
A.
,
Mao
,
Y.
,
Yiu
,
A.
,
Petzold
,
J.
et al. 
(
2019
).
Dchs1-Fat4 regulation of osteogenic differentiation in mouse
.
Development
146
,
dev176776
.
Cubbage
,
C. C.
and
Mabee
,
P. M.
(
1996
).
Development of the cranium and paired fins in the zebrafish Danio rerio (Ostariophysi, Cyprinidae)
.
J. Morphol.
229
,
121
-
160
.
Dambroise
,
E.
,
Ktorza
,
I.
,
Brombin
,
A.
,
Abdessalem
,
G.
,
Edouard
,
J.
,
Luka
,
M.
,
Fiedler
,
I.
,
Binder
,
O.
,
Pelle
,
O.
,
Patton
,
E. E.
et al. 
(
2020
).
Fgfr3 is a positive regulator of osteoblast expansion and differentiation during zebrafish skull vault development
.
J. Bone Miner. Res.
35
,
1782
-
1797
.
Da Mesquita
,
S.
,
Louveau
,
A.
,
Vaccari
,
A.
,
Smirnov
,
I.
,
Cornelison
,
R. C.
,
Kingsmore
,
K. M.
,
Contarino
,
C.
,
Onengut-Gumuscu
,
S.
,
Farber
,
E.
,
Raper
,
D.
et al. 
(
2018
).
Functional aspects of meningeal lymphatics in ageing and Alzheimer's disease
.
Nature
560
,
185
-
191
.
Day
,
T. F.
,
Guo
,
X.
,
Garrett-Beal
,
L.
and
Yang
,
Y.
(
2005
).
Wnt/β-catenin signaling in mesenchymal progenitors controls osteoblast and chondrocyte differentiation during vertebrate skeletogenesis
.
Dev. Cell
8
,
739
-
750
.
Debnath
,
S.
,
Yallowitz
,
A. R.
,
McCormick
,
J.
,
Lalani
,
S.
,
Zhang
,
T.
,
Xu
,
R.
,
Li
,
N.
,
Liu
,
Y.
,
Yang
,
Y. S.
,
Eiseman
,
M.
et al. 
(
2018
).
Discovery of a periosteal stem cell mediating intramembranous bone formation
.
Nature
562
,
133
-
139
.
Deckelbaum
,
R. A.
,
Holmes
,
G.
,
Zhao
,
Z.
,
Tong
,
C.
,
Basilico
,
C.
and
Loomis
,
C. A.
(
2012
).
Regulation of cranial morphogenesis and cell fate at the neural crest-mesoderm boundary by engrailed 1
.
Development
139
,
1346
-
1358
.
DiNuoscio
,
G.
and
Atit
,
R. P.
(
2019
).
Wnt/β–catenin signaling in the mouse embryonic cranial mesenchyme is required to sustain the emerging differentiated meningeal layers
.
Genesis
57
,
e23279
.
Di Pietro
,
L.
,
Barba
,
M.
,
Prampolini
,
C.
,
Ceccariglia
,
S.
,
Frassanito
,
P.
,
Vita
,
A.
,
Guadagni
,
E.
,
Bonvissuto
,
D.
,
Massimi
,
L.
,
Tamburrini
,
G.
et al. 
(
2020
).
GLI1 and AXIN2 are distinctive markers of human calvarial mesenchymal stromal cells in nonsyndromic craniosynostosis
.
IJMS
21
,
4356
.
Di Rocco
,
F.
,
Biosse Duplan
,
M.
,
Heuzé
,
Y.
,
Kaci
,
N.
,
Komla-Ebri
,
D.
,
Munnich
,
A.
,
Mugniery
,
E.
,
Benoist-Lasselin
,
C.
and
Legeai-Mallet
,
L.
(
2014
).
FGFR3 mutation causes abnormal membranous ossification in achondroplasia
.
Hum. Mol. Genet.
23
,
2914
-
2925
.
Dong
,
Y.
,
Jesse
,
A. M.
,
Kohn
,
A.
,
Gunnell
,
L. M.
,
Honjo
,
T.
,
Zuscik
,
M. J.
,
O'Keefe
,
R. J.
and
Hilton
,
M. J.
(
2010
).
RBPjκ-dependent Notch signaling regulates mesenchymal progenitor cell proliferation and differentiation during skeletal development
.
Development
137
,
1461
-
1471
.
Doro
,
D. H.
,
Grigoriadis
,
A. E.
and
Liu
,
K. J.
(
2017
).
Calvarial suture-derived stem cells and their contribution to cranial bone repair
.
Front. Physiol.
8
,
956
.
Doro
,
D.
,
Liu
,
A.
,
Grigoriadis
,
A. E.
and
Liu
,
K. J.
(
2019
).
The osteogenic potential of the neural crest lineage may contribute to craniosynostosis
.
Mol. Syndromol.
10
,
48
-
57
.
Eggenschwiler
,
J. T.
,
Espinoza
,
E.
and
Anderson
,
K. V.
(
2001
).
Rab23 is an essential negative regulator of the mouse Sonic hedgehog signalling pathway
.
Nature
412
,
194
-
198
.
Eggenschwiler
,
J. T.
,
Bulgakov
,
O. V.
,
Qin
,
J.
,
Li
,
T.
and
Anderson
,
K. V.
(
2006
).
Mouse Rab23 regulates Hedgehog signaling from Smoothened to Gli proteins
.
Dev. Biol.
290
,
1
-
12
.
el Ghouzzi
,
V. E.
,
Merrer
,
M. L.
,
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
.
Engin
,
F.
,
Yao
,
Z.
,
Yang
,
T.
,
Zhou
,
G.
,
Bertin
,
T.
,
Jiang
,
M. M.
,
Chen
,
Y.
,
Wang
,
L.
,
Zheng
,
H.
,
Sutton
,
R. E.
et al. 
(
2008
).
Dimorphic effects of Notch signaling in bone homeostasis
.
Nat. Med.
14
,
299
-
305
.
Fan
,
X.
,
Waardenberg
,
A. J.
,
Demuth
,
M.
,
Osteil
,
P.
,
Sun
,
J. Q. J.
,
Loebel
,
D. A. F.
,
Graham
,
M.
,
Tam
,
P. P. L.
and
Fossat
,
N.
(
2020
).
TWIST1 homodimers and heterodimers orchestrate lineage-specific differentiation
.
Mol. Cell. Biol.
40
,
e00663-19
.
Farmer
,
D. T.
,
Mlcochova
,
H.
,
Zhou
,
Y.
,
Koelling
,
N.
,
Wang
,
G.
,
Ashley
,
N.
,
Bugacov
,
H.
,
Chen
,
H.-J.
,
Parvez
,
R.
,
Tseng
,
K.-C.
et al. 
(
2021
).
The developing mouse coronal suture at single-cell resolution
.
Nat. Commun.
12
,
4797
.
Ferguson
,
J. W.
and
Atit
,
R. P.
(
2019
).
A tale of two cities: The genetic mechanisms governing calvarial bone development
.
Genesis
57
,
e23248
.
Ferguson
,
J.
,
Devarajan
,
M.
,
DiNuoscio
,
G.
,
Saiakhova
,
A.
,
Liu
,
C.-F.
,
Lefebvre
,
V.
,
Scacheri
,
P. C.
and
Atit
,
R. P.
(
2018
).
PRC2 Is Dispensable in Vivo for β-Catenin-Mediated Repression of Chondrogenesis in the Mouse Embryonic Cranial Mesenchyme
.
G3 (Bethesda)
8
,
491
-
503
.
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
.
Goodnough
,
L. H.
,
Chang
,
A. T.
,
Treloar
,
C.
,
Yang
,
J.
,
Scacheri
,
P. C.
and
Atit
,
R. P.
(
2012
).
Twist1 mediates repression of chondrogenesis by β-catenin to promote cranial bone progenitor specification
.
Development
139
,
4428
-
4438
.
Goodnough
,
L. H.
,
DiNuoscio
,
G. J.
,
Ferguson
,
J. W.
,
Williams
,
T.
,
Lang
,
R. A.
and
Atit
,
R. P.
(
2014
).
Distinct requirements for cranial ectoderm and mesenchyme-derived Wnts in specification and differentiation of osteoblast and dermal progenitors
.
PLoS Genet.
10
,
e1004152
.
Goodnough
,
L. H.
,
Dinuoscio
,
G. J.
and
Atit
,
R. P.
(
2016
).
Twist1 contributes to cranial bone initiation and dermal condensation by maintaining wnt signaling responsiveness: Twist1 Contributions to Cranial Bone and Dermis
.
Dev. Dyn.
245
,
144
-
156
.
Goos
,
J. A. C.
and
Mathijssen
,
I. M. J.
(
2019
).
Genetic causes of craniosynostosis: an update
.
Mol. Syndromol.
10
,
6
-
23
.
Govindarajan
,
V.
and
Overbeek
,
P. A.
(
2006
).
FGF9 can induce endochondral ossification in cranial mesenchyme
.
BMC Dev. Biol.
6
,
7
.
Greenwald
,
J. A.
,
Mehrara
,
B. J.
,
Spector
,
J. A.
,
Warren
,
S. M.
,
Crisera
,
F. E.
,
Fagenholz
,
P. J.
,
Bouletreau
,
P. J.
and
Longaker
,
M. T.
(
2000
).
Regional differentiation of cranial suture-associated dura mater in vivo and in vitro: implications for suture fusion and patency
.
J. Bone Miner. Res.
15
,
2413
-
2430
.
Hablitz
,
L. M.
and
Nedergaard
,
M.
(
2021
).
The glymphatic system: a novel component of fundamental neurobiology
.
J. Neurosci.
41
,
7698
-
7711
.
Hall
,
B. K.
and
Miyake
,
T.
(
2000
).
All for one and one for all: condensations and the initiation of skeletal development
.
BioEssays
22
,
138
-
147
.
Han
,
J.
,
Ishii
,
M.
,
Bringas
,
P.
,
Maas
,
R. L.
,
Maxson
,
R. E.
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
.
Hasan
,
M. R.
,
Takatalo
,
M.
,
Ma
,
H.
,
Rice
,
R.
,
Mustonen
,
T.
and
Rice
,
D. P.
(
2020
).
RAB23 coordinates early osteogenesis by repressing FGF10-pERK1/2 and GLI1
.
eLife
9
,
e55829
.
Hasegawa
,
S.
,
Sato
,
T.
,
Akazawa
,
H.
,
Okada
,
H.
,
Maeno
,
A.
,
Ito
,
M.
,
Sugitani
,
Y.
,
Shibata
,
H.
,
Miyazaki
,
J.
,
Katsuki
,
M.
et al. 
(
2002
).
Apoptosis in neural crest cells by functional loss of APC tumor suppressor gene
.
Proc. Natl. Acad. Sci. U.S.A.
99
,
297
-
302
.
Hayward
,
R.
and
Gonsalez
,
S.
(
2005
).
How low can you go? Intracranial pressure, cerebral perfusion pressure, and respiratory obstruction in children with complex craniosynostosis
.
J. Neurosurg.
102
,
16
-
22
.
He
,
F.
and
Soriano
,
P.
(
2017
).
Deregulated PDGFRα signaling alters coronal suture morphogenesis and leads to craniosynostosis through endochondral ossification
.
Development
144
,
4026
-
4036
.
Heller
,
J. B.
,
Gabbay
,
J. S.
,
Wasson
,
K.
,
Mitchell
,
S.
,
Heller
,
M. M.
,
Zuk
,
P.
and
Bradley
,
J. P.
(
2007
).
Cranial suture response to stress: expression patterns of Noggin and Runx2
.
Plast. Reconstr. Surg.
119
,
2037
-
2045
.
Hermann
,
C. D.
,
Richards
,
M. A.
,
Chang
,
R.
,
Olivares-Navarrete
,
R.
,
Williams
,
J. K.
,
Guldberg
,
R. E.
,
Vidakovic
,
B.
,
Schwartz
,
Z.
and
Boyan
,
B. D.
(
2013
).
Biphasic fusion of the murine posterior frontal suture
.
Plast. Reconstr. Surg.
131
,
727
-
740
.
Hill
,
T. P.
,
Später
,
D.
,
Taketo
,
M. M.
,
Birchmeier
,
W.
and
Hartmann
,
C.
(
2005
).
Canonical Wnt/β-catenin signaling prevents osteoblasts from differentiating into chondrocytes
.
Dev. Cell
8
,
727
-
738
.
Ho
,
T.-V.
,
Iwata
,
J.
,
Ho
,
H. A.
,
Grimes
,
W. C.
,
Park
,
S.
,
Sanchez-Lara
,
P. A.
and
Chai
,
Y.
(
2015
).
Integration of comprehensive 3D microCT and signaling analysis reveals differential regulatory mechanisms of craniofacial bone development
.
Dev. Biol.
400
,
180
-
190
.
Hoey
,
D. A.
,
Tormey
,
S.
,
Ramcharan
,
S.
,
O'Brien
,
F. J.
and
Jacobs
,
C. R.
(
2012
).
Primary cilia–mediated mechanotransduction in human mesenchymal stem cells
.
Stem Cells
30
,
2561
-
2570
.
Holmes
,
G.
(
2012
).
The role of vertebrate models in understanding craniosynostosis
.
Childs Nerv. Syst.
28
,
1471
-
1481
.
Holmes
,
G.
and
Basilico
,
C.
(
2012
).
Mesodermal expression of Fgfr2S252W is necessary and sufficient to induce craniosynostosis in a mouse model of Apert syndrome
.
Dev. Biol.
368
,
283
-
293
.
Holmes
,
G.
,
Bromage
,
T. G.
and
Basilico
,
C.
(
2011
).
The Sox2 high mobility group transcription factor inhibits mature osteoblast function in transgenic mice
.
Bone
49
,
653
-
661
.
Holmes
,
G.
,
van Bakel
,
H.
,
Zhou
,
X.
,
Losic
,
B.
and
Jabs
,
E. W.
(
2015
).
BCL11B expression in intramembranous osteogenesis during murine craniofacial suture development
.
Gene Expr. Patterns
17
,
16
-
25
.
Holmes
,
G.
,
Gonzalez-Reiche
,
A. S.
,
Lu
,
N.
,
Zhou
,
X.
,
Rivera
,
J.
,
Kriti
,
D.
,
Sebra
,
R.
,
Williams
,
A. A.
,
Donovan
,
M. J.
,
Potter
,
S. S.
et al. 
(
2020
).
Integrated transcriptome and network analysis reveals spatiotemporal dynamics of calvarial suturogenesis
.
Cell Reports
32
,
107871
.
Holmes
,
G.
,
Gonzalez-Reiche
,
A. S.
,
Saturne
,
M.
,
Motch Perrine
,
S. M.
,
Zhou
,
X.
,
Borges
,
A. C.
,
Shewale
,
B.
,
Richtsmeier
,
J. T.
,
Zhang
,
B.
,
van Bakel
,
H.
et al. 
(
2021
).
Single-cell analysis identifies a key role for Hhip in murine coronal suture development
.
Nat. Commun.
12
,
7132
.
Ibarra
,
B. A.
,
Machen
,
C.
and
Atit
,
R. P.
(
2021
).
Wnt-dependent activation of ERK mediates repression of chondrocyte fate during calvarial development
.
JDB
9
,
23
.
Ignelzi
,
M. A.
,
Wang
,
W.
and
Young
,
A. T.
(
2003
).
Fibroblast growth factors lead to increased Msx2 expression and fusion in calvarial sutures
.
J. Bone Miner. Res.
18
,
751
-
759
.
Iliff
,
J. J.
,
Wang
,
M.
,
Liao
,
Y.
,
Plog
,
B. A.
,
Peng
,
W.
,
Gundersen
,
G. A.
,
Benveniste
,
H.
,
Vates
,
G. E.
,
Deane
,
R.
,
Goldman
,
S. A.
et al. 
(
2012
).
A paravascular pathway facilitates CSF flow through the brain parenchyma and the clearance of interstitial solutes, including amyloid β
.
Sci. Transl. Med.
4
,
147ra111
.
Iliff
,
J. J.
,
Chen
,
M. J.
,
Plog
,
B. A.
,
Zeppenfeld
,
D. M.
,
Soltero
,
M.
,
Yang
,
L.
,
Singh
,
I.
,
Deane
,
R.
and
Nedergaard
,
M.
(
2014
).
Impairment of glymphatic pathway function promotes tau pathology after traumatic brain injury
.
J. Neurosci.
34
,
16180
-
16193
.
Inman
,
K. E.
,
Purcell
,
P.
,
Kume
,
T.
and
Trainor
,
P. A.
(
2013
).
Interaction between Foxc1 and Fgf8 during mammalian jaw patterning and in the pathogenesis of syngnathia
.
PLoS Genet.
9
,
e1003949
.
Iseki
,
S.
,
Wilkie
,
A. O. M.
and
Morriss-Kay
,
G. M.
(
1999
).
Fgfr1 and Fgfr2 have distinct differentiation- and proliferation- related roles in the developing skull vault
.
Development
126
,
5611
-
5620
.
Ishii
,
M.
,
Merrill
,
A. E.
,
Chan
,
Y.-S.
,
Gitelman
,
I.
,
Rice
,
D. P. C.
,
Sucov
,
H. M.
and
Maxson
,
R. E.
(
2003
).
Msx2 and Twist cooperatively control the development of the neural crest-derived skeletogenic mesenchyme of the murine skull vault
.
Development
130
,
6131
-
6142
.
Ishii
,
M.
,
Sun
,
J.
,
Ting
,
M.-C.
and
Maxson
,
R. E.
(
2015
).
The development of the calvarial bones and sutures and the pathophysiology of craniosynostosis
. In
Current Topics in Developmental Biology
, pp.
131
-
156
.
Elsevier
.
Ito
,
Y.
,
Yeo
,
J. Y.
,
Chytil
,
A.
,
Han
,
J.
,
Bringas
,
P.
,
Nakajima
,
A.
,
Shuler
,
C. F.
,
Moses
,
H. L.
and
Chai
,
Y.
(
2003
).
Conditional inactivation of Tgfbr2 in cranial neural crest causes cleft palate and calvaria defects
.
Development
130
,
5269
-
5280
.
Iwata
,
J.
,
Hacia
,
J. G.
,
Suzuki
,
A.
,
Sanchez-Lara
,
P. A.
,
Urata
,
M.
and
Chai
,
Y.
(
2012
).
Modulation of noncanonical TGF-β signaling prevents cleft palate in Tgfbr2 mutant mice
.
J. Clin. Invest.
122
,
873
-
885
.
Jacob
,
S.
,
Wu
,
C.
,
Freeman
,
T. A.
,
Koyama
,
E.
and
Kirschner
,
R. E.
(
2007
).
Expression of indian hedgehog, BMP-4 and noggin in craniosynostosis induced by fetal constraint
.
Ann. Plast. Surg.
58
,
215
-
221
.
Jenkins
,
D.
,
Seelow
,
D.
,
Jehee
,
F. S.
,
Perlyn
,
C. A.
,
Alonso
,
L. G.
,
Bueno
,
D. F.
,
Donnai
,
D.
,
Josifiova
,
D.
,
Mathijssen
,
I. M. J.
,
Morton
,
J. E. V.
et al. 
(
2007
).
RAB23 mutations in carpenter syndrome imply an unexpected role for hedgehog signaling in cranial-suture development and obesity
.
Am. J. Hum. Genet.
80
,
1162
-
1170
.
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
.
Jiang
,
X.
,
Iseki
,
S.
,
Maxson
,
R. E.
,
Sucov
,
H. M.
and
Morriss-Kay
,
G. M.
(
2002
).
Tissue origins and interactions in the mammalian skull vault
.
Dev. Biol.
241
,
106
-
116
.
Jiang
,
Y.
,
Zhang
,
S.
,
Mao
,
C.
,
Lai
,
Y.
,
Wu
,
D.
,
Zhao
,
H.
,
Liao
,
C.
and
Chen
,
W.
(
2019
).
Defining a critical period in calvarial development for Hedgehog pathway antagonist-induced frontal bone dysplasia in mice
.
Int. J. Oral. Sci.
11
,
3
.
Jin
,
S.-W.
,
Sim
,
K.-B.
and
Kim
,
S.-D.
(
2016
).
Development and growth of the normal cranial vault: an embryologic review
.
J. Korean Neurosurg. Soc.
59
,
192
.
Johnson
,
D.
,
Iseki
,
S.
,
Wilkie
,
A. O. M.
and
Morriss-Kay
,
G. M.
(
2000
).
Expression patterns of Twist and Fgfr1, -2 and -3 in the developing mouse coronal suture suggest a key role for Twist in suture initiation and biogenesis
.
Mech. Dev.
91
,
341
-
345
.
Johnston
,
J. J.
,
Olivos-Glander
,
I.
,
Killoran
,
C.
,
Elson
,
E.
,
Turner
,
J. T.
,
Peters
,
K. F.
,
Abbott
,
M. H.
,
Aughton
,
D. J.
,
Aylsworth
,
A. S.
,
Bamshad
,
M. J.
et al. 
(
2005
).
Molecular and clinical analyses of greig cephalopolysyndactyly and pallister-hall syndromes: robust phenotype prediction from the type and position of GLI3 mutations
.
Am. J. Hum. Genet.
76
,
609
-
622
.
Kague
,
E.
,
Roy
,
P.
,
Asselin
,
G.
,
Hu
,
G.
,
Simonet
,
J.
,
Stanley
,
A.
,
Albertson
,
C.
and
Fisher
,
S.
(
2016
).
Osterix/Sp7 limits cranial bone initiation sites and is required for formation of sutures
.
Dev. Biol.
413
,
160
-
172
.
Kamath
,
B. M.
,
Stolle
,
C.
,
Bason
,
L.
,
Colliton
,
R. P.
,
Piccoli
,
D. A.
,
Spinner
,
N. B.
and
Krantz
,
I. D.
(
2002
).
Craniosynostosis in Alagille syndrome
.
Am. J. Med. Genet.
112
,
176
-
180
.
Kanther
,
M.
,
Scalici
,
A.
,
Rashid
,
A.
,
Miao
,
K.
,
Van Deventer
,
E.
and
Fisher
,
S.
(
2019
).
Initiation and early growth of the skull vault in zebrafish
.
Mech. Dev.
160
,
103578
.
Katsianou
,
M.
,
Papavassiliou
,
K. A.
,
Zoi
,
I.
,
Gargalionis
,
A. N.
,
Panagopoulos
,
D.
,
Themistocleous
,
M. S.
,
Piperi
,
C.
,
Papavassiliou
,
A. G.
and
Basdra
,
E. K.
(
2021
).
Polycystin–1 modulates RUNX2 activation and osteocalcin gene expression via ERK signaling in a human craniosynostosis cell model
.
J. Cell. Mol. Med.
25
,
3216
-
3225
.
Kaucka
,
M.
and
Adameyko
,
I.
(
2019
).
Evolution and development of the cartilaginous skull: from a lancelet towards a human face
.
Semin. Cell Dev. Biol.
91
,
2
-
12
.
Kim
,
H.-J.
,
Rice
,
D. P. C.
,
Kettunen
,
P. J.
and
Thesleff
,
I.
(
1998
).
FGF-, BMP- and Shh-mediated signalling pathways in the regulation of cranial suture morphogenesis and calvarial bone development
.
Development
125
,
1241
-
1251
.
Kirschner
,
R. E.
,
Gannon
,
F. H.
,
Xu
,
J.
,
Wang
,
J.
,
Karmacharya
,
J.
,
Bartlett
,
S. P.
and
Whitaker
,
L. A.
(
2002
).
Craniosynostosis and altered patterns of fetal TGF-beta expression induced by intrauterine constraint
.
Plast. Reconstr. Surg.
109
,
2338
-
2346
.
Komatsu
,
Y.
,
Yu
,
P. B.
,
Kamiya
,
N.
,
Pan
,
H.
,
Fukuda
,
T.
,
Scott
,
G. J.
,
Ray
,
M. K.
,
Yamamura
,
K.
and
Mishina
,
Y.
(
2013
).
Augmentation of smad-dependent BMP signaling in neural crest cells causes craniosynostosis in mice: Bmp Signaling and Craniosynostosis
.
J. Bone Miner. Res.
28
,
1422
-
1433
.
Komori
,
T.
,
Yagi
,
H.
,
Nomura
,
S.
,
Yamaguchi
,
A.
,
Sasaki
,
K.
,
Deguchi
,
K.
,
Shimizu
,
Y.
,
Bronson
,
R. T.
,
Gao
,
Y.-H.
,
Inada
,
M.
et al. 
(
1997
).
Targeted disruption of Cbfa1 results in a complete lack of bone formation owing to maturational arrest of osteoblasts
.
Cell
89
,
755
-
764
.
Kosaras
,
B.
,
Jakubowski
,
M.
,
Kainz
,
V.
and
Burstein
,
R.
(
2009
).
Sensory innervation of the calvarial bones of the mouse
.
J. Comp. Neurol.
515
,
331
-
348
.
Kyrylkova
,
K.
,
Iwaniec
,
U. T.
,
Philbrick
,
K. A.
and
Leid
,
M.
(
2016
).
BCL11B regulates sutural patency in the mouse craniofacial skeleton
.
Dev. Biol.
415
,
251
-
260
.
Kyutoku
,
S.
and
Inagaki
,
T.
(
2017
).
Review of past reports and current concepts of surgical management for craniosynostosis
.
Neurol. Med. Chir. (Tokyo)
57
,
217
-
224
.
Lajeunie
,
E.
,
Merrer
,
M. L.
,
Bonaïti-Pellie
,
C.
,
Marchac
,
D.
and
Renier
,
D.
(
1995
).
Genetic study of nonsyndromic coronal craniosynostosis
.
Am. J. Med. Genet.
55
,
500
-
504
.
Lana-Elola
,
E.
,
Rice
,
R.
,
Grigoriadis
,
A. E.
and
Rice
,
D. P. C.
(
2007
).
Cell fate specification during calvarial bone and suture development
.
Dev. Biol.
311
,
335
-
346
.
Laue
,
K.
,
Pogoda
,
H.-M.
,
Daniel
,
P. B.
,
van Haeringen
,
A.
,
Alanay
,
Y.
,
von Ameln
,
S.
,
Rachwalski
,
M.
,
Morgan
,
T.
,
Gray
,
M. J.
,
Breuning
,
M. H.
et al. 
(
2011
).
Craniosynostosis and multiple skeletal anomalies in humans and zebrafish result from a defect in the localized degradation of retinoic acid
.
Am. J. Hum. Genet.
89
,
595
-
606
.
Ledwon
,
J. K.
,
Turin
,
S. Y.
,
Gosain
,
A. K.
and
Topczewska
,
J. M.
(
2018
).
The expression of fgfr3 in the zebrafish head
.
Gene Expression Patterns
29
,
32
-
38
.
Lee
,
S.-W.
,
Choi
,
K.-Y.
,
Cho
,
J.-Y.
,
Jung
,
S.-H.
,
Song
,
K.-B.
,
Park
,
E.-K.
,
Choi
,
J.-Y.
,
Shin
,
H.-I.
,
Kim
,
S.-Y.
,
Woo
,
K.-M.
et al. 
(
2006
).
TGF-β2 stimulates cranial suture closure through activation of the Erk-MAPK pathway
.
J. Cell. Biochem.
98
,
981
-
991
.
Lenton
,
K.
,
James
,
A. W.
,
Manu
,
A.
,
Brugmann
,
S. A.
,
Birker
,
D.
,
Nelson
,
E. R.
,
Leucht
,
P.
,
Helms
,
J. A.
and
Longaker
,
M. T.
(
2011
).
Indian hedgehog positively regulates calvarial ossification and modulates bone morphogenetic protein signaling
.
Genesis
49
,
784
-
796
.
Levine
,
J. P.
,
Bradley
,
J. P.
,
Roth
,
D. A.
,
McCarthy
,
J. G.
and
Longaker
,
M. T.
(
1998
).
Studies in cranial suture biology: regional dura mater determines overlying suture biology
.
Plast. Reconstr. Surg.
101
,
1441
-
1447
.
Li
,
S.
,
Quarto
,
N.
and
Longaker
,
M. T.
(
2007
).
Dura mater-derived FGF-2 mediates mitogenic signaling in calvarial osteoblasts
.
Am. J. Physiol. Cell Physiol.
293
,
C1834
-
C1842
.
Li
,
S.
,
Quarto
,
N.
and
Longaker
,
M. T.
(
2010
).
Activation of FGF signaling mediates proliferative and osteogenic differences between neural crest derived frontal and mesoderm parietal derived bone
.
PLoS ONE
5
,
e14033
.
Li
,
S.
,
Quarto
,
N.
,
Senarath-Yapa
,
K.
,
Grey
,
N.
,
Bai
,
X.
and
Longaker
,
M. T.
(
2015
).
Enhanced activation of canonical Wnt signaling confers mesoderm-derived parietal bone with similar osteogenic and skeletal healing capacity to neural crest-derived frontal bone
.
PLoS ONE
10
,
e0138059
.
Li
,
B.
,
Wang
,
Y.
,
Fan
,
Y.
,
Ouchi
,
T.
,
Zhao
,
Z.
and
Li
,
L.
(
2021
).
Cranial suture mesenchymal stem cells: insights and advances
.
Biomolecules
11
,
1129
.
Lieberman
,
D. E.
,
McBratney
,
B. M.
and
Krovitz
,
G.
(
2002
).
The evolution and development of cranial form in Homo sapiens
.
Proc. Natl. Acad. Sci. U.S.A.
99
,
1134
-
1139
.
Lind
,
T.
,
Öhman
,
C.
,
Calounova
,
G.
,
Rasmusson
,
A.
,
Andersson
,
G.
,
Pejler
,
G.
and
Melhus
,
H.
(
2017
).
Excessive dietary intake of vitamin A reduces skull bone thickness in mice
.
PLoS ONE
12
,
e0176217
.
Liu
,
Y.-H.
,
Tang
,
Z.
,
Kundu
,
R. K.
,
Wu
,
L.
,
Luo
,
W.
,
Zhu
,
D.
,
Sangiorgi
,
F.
,
Snead
,
M. L.
and
Maxson
,
R. E.
(
1999
).
Msx2 gene dosage influences the number of proliferative osteogenic cells in growth centers of the developing murine skull: a possible mechanism for Msx2-mediated craniosynostosis in humans
.
Dev. Biol.
205
,
260
-
274
.
Louveau
,
A.
,
Smirnov
,
I.
,
Keyes
,
T. J.
,
Eccles
,
J. D.
,
Rouhani
,
S. J.
,
Peske
,
J. D.
,
Derecki
,
N. C.
,
Castle
,
D.
,
Mandell
,
J. W.
,
Lee
,
K. S.
et al. 
(
2015
).
Structural and functional features of central nervous system lymphatic vessels
.
Nature
523
,
337
-
341
.
Louveau
,
A.
,
Herz
,
J.
,
Alme
,
M. N.
,
Salvador
,
A. F.
,
Dong
,
M. Q.
,
Viar
,
K. E.
,
Herod
,
S. G.
,
Knopp
,
J.
,
Setliff
,
J. C.
,
Lupi
,
A. L.
et al. 
(
2018
).
CNS lymphatic drainage and neuroinflammation are regulated by meningeal lymphatic vasculature
.
Nat. Neurosci.
21
,
1380
-
1391
.
Lund
,
A. W.
,
Duraes
,
F. V.
,
Hirosue
,
S.
,
Raghavan
,
V. R.
,
Nembrini
,
C.
,
Thomas
,
S. N.
,
Issa
,
A.
,
Hugues
,
S.
and
Swartz
,
M. A.
(
2012
).
VEGF-C promotes immune tolerance in B16 melanomas and cross-presentation of tumor antigen by lymph node lymphatics
.
Cell Reports
1
,
191
-
199
.
Luther
,
G.
,
Wagner
,
E. R.
,
Zhu
,
G.
,
Kang
,
Q.
,
Luo
,
Q.
,
Lamplot
,
J.
,
Bi
,
Y.
,
Luo
,
X.
,
Luo
,
J.
,
Teven
,
C.
et al. 
(
2011
).
BMP-9 induced osteogenic differentiation of mesenchymal stem cells: molecular mechanism and therapeutic potential
.
Curr. Gene Ther.
11
,
229
-
240
.
Machida
,
A.
,
Okuhara
,
S.
,
Harada
,
K.
and
Iseki
,
S.
(
2014
).
Difference in apical and basal growth of the frontal bone primordium in Foxc1 ch/ch mice: Foxc1 function in skull bone growth
.
Congenital Anom.
54
,
172
-
177
.
Maclean
,
G.
,
Dollé
,
P.
and
Petkovich
,
M.
(
2009
).
Genetic disruption of CYP26B1 severely affects development of neural crest derived head structures, but does not compromise hindbrain patterning
.
Dev. Dyn.
238
,
732
-
745
.
Maeno
,
T.
,
Moriishi
,
T.
,
Yoshida
,
C. A.
,
Komori
,
H.
,
Kanatani
,
N.
,
Izumi
,
S.
,
Takaoka
,
K.
and
Komori
,
T.
(
2011
).
Early onset of Runx2 expression caused craniosynostosis, ectopic bone formation, and limb defects
.
Bone
49
,
673
-
682
.
Mansukhani
,
A.
,
Ambrosetti
,
D.
,
Holmes
,
G.
,
Cornivelli
,
L.
and
Basilico
,
C.
(
2005
).
Sox2 induction by FGF and FGFR2 activating mutations inhibits Wnt signaling and osteoblast differentiation
.
J. Cell Biol.
168
,
1065
-
1076
.
Maruyama
,
T.
,
Jeong
,
J.
,
Sheu
,
T.-J.
and
Hsu
,
W.
(
2016
).
Stem cells of the suture mesenchyme in craniofacial bone development, repair and regeneration
.
Nat. Commun.
7
,
10526
.
Mehrara
,
B. J.
,
Most
,
D.
,
Chang
,
J.
,
Bresnick
,
S.
,
Turk
,
A.
,
Schendel
,
S. A.
,
Gittes
,
G. K.
and
Longaker
,
M. T.
(
1999
).
Basic fibroblast growth factor and transforming growth factor beta-1 expression in the developing dura mater correlates with calvarial bone formation
.
Plast. Reconstr. Surg.
104
,
435
-
444
.
Menon
,
S.
,
Salhotra
,
A.
,
Shailendra
,
S.
,
Tevlin
,
R.
,
Ransom
,
R. C.
,
Januszyk
,
M.
,
Chan
,
C. K. F.
,
Behr
,
B.
,
Wan
,
D. C.
,
Longaker
,
M. T.
et al. 
(
2021
).
Skeletal stem and progenitor cells maintain cranial suture patency and prevent craniosynostosis
.
Nat. Commun.
12
,
4640
.
Merrill
,
A. E.
,
Bochukova
,
E. G.
,
Brugger
,
S. M.
,
Ishii
,
M.
,
Pilz
,
D. T.
,
Wall
,
S. A.
,
Lyons
,
K. M.
,
Wilkie
,
A. O. M.
and
Maxson
,
R. E.
(
2006
).
Cell mixing at a neural crest-mesoderm boundary and deficient ephrin-Eph signaling in the pathogenesis of craniosynostosis
.
Hum. Mol. Genet.
15
,
1319
-
1328
.
Mikheeva
,
S. A.
,
Camp
,
N. D.
,
Huang
,
L.
,
Jain
,
A.
,
Jung
,
S. Y.
,
Avci
,
N. G.
,
Tokita
,
M.
,
Wolf-Yadlin
,
A.
,
Zhang
,
J.
,
Tapscott
,
S. J.
et al. 
(
2019
).
TWIST1 heterodimerization with E12 requires coordinated protein phosphorylation to regulate periostin expression
.
Cancers
11
,
1392
.
Miller
,
K. A.
,
Twigg
,
S. R. F.
,
McGowan
,
S. J.
,
Phipps
,
J. M.
,
Fenwick
,
A. L.
,
Johnson
,
D.
,
Wall
,
S. A.
,
Noons
,
P.
,
Rees
,
K. E. M.
,
Tidey
,
E. A.
et al. 
(
2017
).
Diagnostic value of exome and whole genome sequencing in craniosynostosis
.
J. Med. Genet.
54
,
260
-
268
.
Miraoui
,
H.
,
Ringe
,
J.
,
Haupl
,
T.
and
Marie
,
P. J.
(
2010
).
Increased EFG- and PDGF -receptor signaling by mutant FGF-receptor 2 contributes to osteoblast dysfunction in Apert craniosynostosis
.
Hum. Mol. Genet.
19
,
1678
-
1689
.
Mishra
,
S.
,
Choe
,
Y.
,
Pleasure
,
S. J.
and
Siegenthaler
,
J. A.
(
2016
).
Cerebrovascular defects in Foxc1 mutants correlate with aberrant WNT and VEGF—A pathways downstream of retinoic acid from the meninges
.
Dev. Biol.
420
,
148
-
165
.
Miyama
,
K.
,
Yamada
,
G.
,
Yamamoto
,
T. S.
,
Takagi
,
C.
,
Miyado
,
K.
,
Sakai
,
M.
,
Ueno
,
N.
and
Shibuya
,
H.
(
1999
).
A BMP-inducible Gene, Dlx5, regulates osteoblast differentiation and mesoderm induction
.
Dev. Biol.
208
,
123
-
133
.
Moazen
,
M.
,
Hejazi
,
M.
,
Savery
,
D.
,
Jones
,
D.
,
Marghoub
,
A.
,
Alazmani
,
A.
and
Pauws
,
E.
(
2022
).
Mechanical loading of cranial joints minimizes the craniofacial phenotype in Crouzon syndrome
.
Sci. Rep.
12
,
9693
.
Mori-Akiyama
,
Y.
,
Akiyama
,
H.
,
Rowitch
,
D. H.
and
de Crombrugghe
,
B.
(
2003
).
Sox9 is required for determination of the chondrogenic cell lineage in the cranial neural crest
.
Proc. Natl. Acad. Sci. U.S.A.
100
,
9360
-
9365
.
Morriss-Kay
,
G.
(
1993
).
Retinoic acid and craniofacial development: Molecules and morphogenesis
.
BioEssays
15
,
9
-
15
.
Morriss-Kay
,
G. M.
and
Wilkie
,
A. O. M.
(
2005
).
Growth of the normal skull vault and its alteration in craniosynostosis: insights from human genetics and experimental studies: Growth of the normal skull vault and its alteration in craniosynostosis
.
J. Anat.
207
,
637
-
653
.
Most
,
D.
,
Levine
,
J. P.
,
Chang
,
J.
,
Sung
,
J.
,
McCarthy
,
J. G.
,
Schendel
,
S. A.
and
Longaker
,
M. T.
(
1998
).
Studies in cranial suture biology: up-regulation of transforming growth factor-beta1 and basic fibroblast growth factor mRNA correlates with posterior frontal cranial suture fusion in the rat
.
Plast. Reconstr. Surg.
101
,
1431
-
1440
.
Mukherjee
,
A.
and
Dixon
,
J. B.
(
2021
).
Mechanobiology of lymphatic vessels
. In
Vascular Mechanobiology in Physiology and Disease
(ed.
M.
Hecker
and
D. J.
Duncker
), pp.
191
-
232
.
Cham
:
Springer International Publishing
.
Nah
,
H.-D.
,
Koyama
,
E.
,
Agochukwu
,
N. B.
,
Bartlett
,
S. P.
and
Muenke
,
M.
(
2012
).
Phenotype profile of a genetic mouse model for Muenke syndrome
.
Childs Nerv. Syst.
28
,
1483
-
1493
.
Nakashima
,
K.
,
Zhou
,
X.
,
Kunkel
,
G.
,
Zhang
,
Z.
,
Deng
,
J. M.
,
Behringer
,
R. R.
and
de Crombrugghe
,
B.
(
2002
).
The novel zinc finger-containing transcription factor osterix is required for osteoblast differentiation and bone formation
.
Cell
108
,
17
-
29
.
Neben
,
C. L.
,
Idoni
,
B.
,
Salva
,
J. E.
,
Tuzon
,
C. T.
,
Rice
,
J. C.
,
Krakow
,
D.
and
Merrill
,
A. E.
(
2014
).
Bent bone dysplasia syndrome reveals nucleolar activity for FGFR2 in ribosomal DNA transcription
.
Hum. Mol. Genet.
23
,
5659
-
5671
.
Nishio
,
Y.
,
Dong
,
Y.
,
Paris
,
M.
,
O'Keefe
,
R. J.
,
Schwarz
,
E. M.
and
Drissi
,
H.
(
2006
).
Runx2-mediated regulation of the zinc finger Osterix/Sp7 gene
.
Gene
372
,
62
-
70
.
Oda
,
T.
,
Elkahloun
,
A. G.
,
Pike
,
B. L.
,
Okajima
,
K.
,
Krantz
,
I. D.
,
Genin
,
A.
,
Piccoli
,
D. A.
,
Meltzer
,
P. S.
,
Spinner
,
N. B.
,
Collins
,
F. S.
et al. 
(
1997
).
Mutations in the human Jagged1 gene are responsible for Alagille syndrome
.
Nat. Genet.
16
,
235
-
242
.
Ogle
,
R. C.
,
Tholpady
,
S. S.
,
McGlynn
,
K. A.
and
Ogle
,
R. A.
(
2004
).
Regulation of cranial suture morphogenesis
.
Cells Tissues Organs
176
,
54
-
66
.
Oppenheimer
,
A. J.
,
Rhee
,
S. T.
,
Goldstein
,
S. A.
and
Buchman
,
S. R.
(
2009
).
Force-induced craniosynostosis in the murine sagittal suture
.
Plast. Reconstr. Surg.
124
,
1840
-
1848
.
Oppenheimer
,
A. J.
,
Rhee
,
S. T.
,
Goldstein
,
S. A.
and
Buchman
,
S. R.
(
2012
).
Force-induced craniosynostosis via paracrine signaling in the murine sagittal suture
.
J. Craniofac. Surg.
23
,
573
-
577
.
Opperman
,
L. A.
(
2000
).
Cranial sutures as intramembranous bone growth sites
.
Dev. Dyn.
219
,
472
-
485
.
Opperman
,
L. A.
,
Sweeney
,
T. M.
,
Redmon
,
J.
,
Persing
,
J. A.
and
Ogle
,
R. C.
(
1993
).
Tissue interactions with underlying dura mater inhibit osseous obliteration of developing cranial sutures
.
Dev. Dyn.
198
,
312
-
322
.
Opperman
,
L. A.
,
Passarelli
,
R. W.
,
Morgan
,
E. P.
,
Reintjes
,
M.
and
Ogle
,
R. C.
(
1995
).
Cranial sutures require tissue interactions with dura mater to resist osseous obliteration in vitro
.
J. Bone Miner. Res.
10
,
1978
-
1987
.
Opperman
,
L. A.
,
Passarelli
,
R. W.
,
Nolen
,
A. A.
,
Gampper
,
T. J.
,
Lin
,
K. Y. K.
and
Ogle
,
R. C.
(
1996
).
Dura mater secretes soluble heparin-binding factors required for cranial suture morphogenesis
.
In Vitro Cell. Dev. Biol. Anim.
32
,
627
-
632
.
Opperman
,
L. A.
,
Adab
,
K.
and
Gakunga
,
P. T.
(
2000
).
Transforming growth factor–β2 and TGF–β3 regulate fetal rat cranial suture morphogenesis by regulating rates of cell proliferation and apoptosis
.
Dev. Dyn.
219
,
237
-
247
.
Otto
,
F.
,
Thornell
,
A. P.
,
Crompton
,
T.
,
Denzel
,
A.
,
Gilmour
,
K. C.
,
Rosewell
,
I. R.
,
Stamp
,
G. W. H.
,
Beddington
,
R. S. P.
,
Mundlos
,
S.
,
Olsen
,
B. R.
et al. 
(
1997
).
Cbfa1, a candidate gene for cleidocranial dysplasia syndrome, is essential for osteoblast differentiation and bone development
.
Cell
89
,
765
-
771
.
Ouyang
,
Z.
,
Chen
,
Z.
,
Ishikawa
,
M.
,
Yue
,
X.
,
Kawanami
,
A.
,
Leahy
,
P.
,
Greenfield
,
E. M.
and
Murakami
,
S.
(
2014
).
Prx1 and 3.2kb Col1a1 promoters target distinct bone cell populations in transgenic mice
.
Bone
58
,
136
-
145
.
Pakvasa
,
M.
,
Haravu
,
P.
,
Boachie-Mensah
,
M.
,
Jones
,
A.
,
Coalson
,
E.
,
Liao
,
J.
,
Zeng
,
Z.
,
Wu
,
D.
,
Qin
,
K.
,
Wu
,
X.
et al. 
(
2021
).
Notch signaling: Its essential roles in bone and craniofacial development
.
Genes Dis.
8
,
8
-
24
.
Park
,
W.-J.
,
Theda
,
C.
,
Maestri
,
N. E.
,
Meyers
,
G. A.
,
Fryburg
,
J. S.
,
Dufresne
,
C.
,
Cohen
,
M. M.
and
Jabs’
,
E. W.
(
1995
).
Analysis of phenotypic features and FGFR2 mutations in apert syndrome
.
Am. J. Hum. Genet.
57
,
321
-
328
.
Peskett
,
E.
,
Kumar
,
S.
,
Baird
,
W.
,
Jaiswal
,
J.
,
Li
,
M.
,
Patel
,
P.
,
Britto
,
J. A.
and
Pauws
,
E.
(
2017
).
Analysis of the Fgfr2C342Y mouse model shows condensation defects due to misregulation of Sox9 expression in prechondrocytic mesenchyme
.
Biol. Open
6
,
223
-
231
.
Phimphilai
,
M.
,
Zhao
,
Z.
,
Boules
,
H.
,
Roca
,
H.
and
Franceschi
,
R. T.
(
2006
).
BMP Signaling Is Required for RUNX2-Dependent Induction of the Osteoblast Phenotype
.
J. Bone Miner. Res.
21
,
637
-
646
.
Plog
,
B. A.
,
Lou
,
N.
,
Pierre
,
C. A.
,
Cove
,
A.
,
Kenney
,
H. M.
,
Hitomi
,
E.
,
Kang
,
H.
,
Iliff
,
J. J.
,
Zeppenfeld
,
D. M.
,
Nedergaard
,
M.
et al. 
(
2020
).
When the air hits your brain: decreased arterial pulsatility after craniectomy leading to impaired glymphatic flow
.
J. Neurosurg.
133
,
210
-
223
.
Quarto
,
N.
and
Longaker
,
M. T.
(
2005
).
The Zebrafish (Danio rerio): a model system for cranial suture patterning
.
Cells Tissues Organs
181
,
109
-
118
.
Renier
,
D.
,
Sainte-Rose
,
C.
,
Marchac
,
D.
and
Hirsch
,
J. F.
(
1982
).
Intracranial pressure in craniostenosis
.
J. Neurosurg.
57
,
370
-
377
.
Rice
,
D. P.
,
Aberg
,
T.
,
Chan
,
Y.
,
Tang
,
Z.
,
Kettunen
,
P. J.
,
Pakarinen
,
L.
,
Maxon
,
R. E.
and
Thesleff
,
I.
(
2000
).
Integration of FGF and Twist in calvarial bone and suture development
.
Development
127
,
1845
-
1855
.
Rice
,
D. P. C.
,
Connor
,
E. C.
,
Veltmaat
,
J. M.
,
Lana-Elola
,
E.
,
Veistinen
,
L.
,
Tanimoto
,
Y.
,
Bellusci
,
S.
and
Rice
,
R.
(
2010
).
Gli3Xt-J/Xt-J mice exhibit lambdoid suture craniosynostosis which results from altered osteoprogenitor proliferation and differentiation
.
Hum. Mol. Genet.
19
,
3457
-
3467
.
Rice
,
R.
,
Rice
,
D. P. C.
,
Olsen
,
B. R.
and
Thesleff
,
I.
(
2003
).
Progression of calvarial bone development requires Foxc1 regulation of Msx2 and Alx4
.
Dev. Biol.
262
,
75
-
87
.
Rice
,
R.
,
Rice
,
D. P. C.
and
Thesleff
,
I.
(
2005
).
Foxc1 integrates Fgf and Bmp signaling independently of twist or noggin during calvarial bone development
.
Dev. Dyn.
233
,
847
-
852
.
Richtsmeier
,
J. T.
and
Flaherty
,
K.
(
2013
).
Hand in glove: brain and skull in development and dysmorphogenesis
.
Acta Neuropathol.
125
,
469
-
489
.
Roa
,
L. A.
,
Bloemen
,
M.
,
Carels
,
C. E. L.
,
Wagener
,
F. A. D. T. G.
and
Von den Hoff
,
J. W.
(
2019
).
Retinoic acid disrupts osteogenesis in pre-osteoblasts by down-regulating WNT signaling
.
Int. J. Biochem. Cell Biol.
116
,
105597
.
Robledo
,
R. F.
,
Rajan
,
L.
,
Li
,
X.
and
Lufkin
,
T.
(
2002
).
The Dlx5 and Dlx6 homeobox genes are essential for craniofacial, axial, and appendicular skeletal development
.
Genes Dev.
16
,
1089
-
1101
.
Rodda
,
S. J.
and
McMahon
,
A. P.
(
2006
).
Distinct roles for Hedgehog and canonical Wnt signaling in specification,differentiation and maintenance of osteoblast progenitors
.
Development
133
,
3231
-
3244
.
Roosenboom
,
J.
,
Lee
,
M. K.
,
Hecht
,
J. T.
,
Heike
,
C. L.
,
Wehby
,
G. L.
,
Christensen
,
K.
,
Feingold
,
E.
,
Marazita
,
M. L.
,
Maga
,
A. M.
,
Shaffer
,
J. R.
et al. 
(
2018
).
Mapping genetic variants for cranial vault shape in humans
.
PLoS ONE
13
,
e0196148
.
Roth
,
D. A.
,
Longaker
,
M. T.
,
McCarthy
,
J. G.
,
Rosen
,
D. M.
,
McMullen
,
H. F.
,
Levine
,
J. P.
,
Sung
,
J.
and
Gold
,
L. I.
(
1997
).
Studies in cranial suture biology: part I. increased immunoreactivity for TGF-β isoforms (β1, β2, and β3) during rat cranial suture fusion
.
J. Bone Miner. Res.
12
,
311
-
321
.
Sasaki
,
T.
,
Ito
,
Y.
,
Bringas
,
P.
,
Chou
,
S.
,
Urata
,
M. M.
,
Slavkin
,
H.
and
Chai
,
Y.
(
2006
).
TGFβ-mediated FGF signaling is crucial for regulating cranial neural crest cell proliferation during frontal bone development
.
Development
133
,
371
-
381
.
Satokata
,
I.
,
Ma
,
L.
,
Ohshima
,
H.
,
Bei
,
M.
,
Woo
,
I.
,
Nishizawa
,
K.
,
Maeda
,
T.
,
Takano
,
Y.
,
Uchiyama
,
M.
,
Heaney
,
S.
et al. 
(
2000
).
Msx2 deficiency in mice causes pleiotropic defects in bone growth and ectodermal organ formation
.
Nat. Genet.
24
,
391
-
395
.
Sawh-Martinez
,
R.
and
Steinbacher
,
D. M.
(
2019
).
Syndromic craniosynostosis
.
Clin. Plast. Surg.
46
,
141
-
155
.
Schmidt
,
L.
,
Taiyab
,
A.
,
Melvin
,
V. S.
,
Jones
,
K. L.
and
Williams
,
T.
(
2018
).
Increased FGF8 signaling promotes chondrogenic rather than osteogenic development in the embryonic skull
.
Dis. Model. Mech.
11
,
dmm031526
.
Sharma
,
V. P.
,
Fenwick
,
A. L.
,
Brockop
,
M. S.
,
McGowan
,
S. J.
,
Goos
,
J. A. C.
,
Hoogeboom
,
A. J. M.
,
Brady
,
A. F.
,
Jeelani
,
N. O.
,
Lynch
,
S. A.
et al. 
(
2013
).
Mutations in TCF12, encoding a basic helix-loop-helix partner of TWIST1, are a frequent cause of coronal craniosynostosis
.
Nat. Genet.
45
,
304
-
307
.
Shukla
,
V.
,
Coumoul
,
X.
,
Wang
,
R.-H.
,
Kim
,
H.-S.
and
Deng
,
C.-X.
(
2007
).
RNA interference and inhibition of MEK-ERK signaling prevent abnormal skeletal phenotypes in a mouse model of craniosynostosis
.
Nat. Genet.
39
,
1145
-
1150
.
Siegenthaler
,
J. A.
,
Ashique
,
A. M.
,
Zarbalis
,
K.
,
Patterson
,
K. P.
,
Hecht
,
J. H.
,
Kane
,
M. A.
,
Folias
,
A. E.
,
Choe
,
Y.
,
May
,
S. R.
,
Kume
,
T.
et al. 
(
2009
).
Retinoic acid from the meninges regulates cortical neuron generation
.
Cell
139
,
597
-
609
.
Siismets
,
E. M.
and
Hatch
,
N. E.
(
2020
).
Cranial neural crest cells and their role in the pathogenesis of craniofacial anomalies and coronal craniosynostosis
.
JDB
8
,
18
.
Singh
,
N.
,
Gupta
,
M.
,
Trivedi
,
C. M.
,
Singh
,
M. K.
,
Li
,
L.
and
Epstein
,
J. A.
(
2013
).
Murine craniofacial development requires Hdac3-mediated repression of Msx gene expression
.
Dev. Biol.
377
,
333
-
344
.
Song
,
D.
,
Huang
,
S.
,
Zhang
,
L.
,
Liu
,
W.
,
Huang
,
B.
,
Feng
,
Y.
,
Liu
,
B.
,
He
,
T.-C.
,
Huang
,
D.
and
Reid
,
R. R.
(
2020
).
Differential responsiveness to BMP9 between patent and fused suture progenitor cells from craniosynostosis patients
.
Plast. Reconstr. Surg.
145
,
552e
-
562e
.
Spector
,
J. A.
,
Greenwald
,
J. A.
,
Warren
,
S. M.
,
Bouletreau
,
P. L.
,
Detch
,
R. C.
,
Fegenholz
,
P. J.
,
Cicero
,
F. E.
and
Longaker
,
M. T.
(
2002
).
Dura mater biology: autocrine and paracrine effects of fibroblast growth factor 2
.
Plast. Reconstr. Surg.
109
,
645
-
654
.
Spoorendonk
,
K. M.
,
Peterson-Maduro
,
J.
,
Renn
,
J.
,
Trowe
,
T.
,
Kranenbarg
,
S.
,
Winkler
,
C.
and
Schulte-Merker
,
S.
(
2008
).
Retinoic acid and Cyp26b1 are critical regulators of osteogenesis in the axial skeleton
.
Development
135
,
3765
-
3774
.
Stanton
,
E.
,
Urata
,
M.
,
Chen
,
J.-F.
and
Chai
,
Y.
(
2022
).
The clinical manifestations, molecular mechanisms and treatment of craniosynostosis
.
Disease Models & Mechanisms
15
,
dmm049390
.
Sun
,
J.
,
Ishii
,
M.
,
Ting
,
M.-C.
and
Maxson
,
R.
(
2013
).
Foxc1 controls the growth of the murine frontal bone rudiment by direct regulation of a Bmp response threshold of Msx2
.
Development
140
,
1034
-
1044
.
Takarada
,
T.
,
Nakazato
,
R.
,
Tsuchikane
,
A.
,
Fujikawa
,
K.
,
Iezaki
,
T.
,
Yoneda
,
Y.
and
Hinoi
,
E.
(
2015
).
Genetic analysis of Runx2 function during intramembranous ossification
.
Development
143
,
211
-
218
.
Taylor
,
A.
,
Butt
,
W.
,
Rosenfeld
,
J.
,
Shann
,
F.
,
Ditchfield
,
M.
,
Lewis
,
E.
,
Klug
,
G.
,
Wallace
,
D.
,
Henning
,
R.
and
Tibballs
,
J.
(
2001
).
A randomized trial of very early decompressive craniectomy in children with traumatic brain injury and sustained intracranial hypertension
.
Childs Nerv. Syst.
17
,
154
-
162
.
Teng
,
C. S.
,
Ting
,
M.
,
Farmer
,
D. T.
,
Brockop
,
M.
,
Maxson
,
R. E.
and
Crump
,
J. G.
(
2018
).
Altered bone growth dynamics prefigure craniosynostosis in a zebrafish model of Saethre-Chotzen syndrome
.
eLife
7
,
e37024
.
Teng
,
C. S.
,
Cavin
,
L.
,
Maxson
,
R. E.
,
Sánchez-Villagra
,
M. R.
and
Crump
,
J. G.
(
2019
).
Resolving homology in the face of shifting germ layer origins: Lessons from a major skull vault boundary
.
eLife
8
,
e52814
.
Tholpady
,
S. S.
,
Freyman
,
T. F.
,
Chachra
,
D.
and
Ogle
,
R. C.
(
2007
).
Tensional forces influence gene expression and sutural state of rat calvariae In Vitro
.
Plast. Reconstr. Surg.
120
,
601
-
611
.
Tiberio
,
F.
,
Parolini
,
O.
and
Lattanzi
,
W.
(
2021
).
Ciliary signaling and mechanotransduction in the pathophysiology of Craniosynostosis
.
Genes
12
,
1073
.
Ting
,
M.-C.
,
Wu
,
N. L.
,
Roybal
,
P. G.
,
Sun
,
J.
,
Liu
,
L.
,
Yen
,
Y.
and
Maxson
,
R. E.
(
2009
).
EphA4 as an effector of Twist1 in the guidance of osteogenic precursor cells during calvarial bone growth and in craniosynostosis
.
Development
136
,
855
-
864
.
Ting
,
M.
,
Farmer
,
D. J. T.
,
Teng
,
C. S.
,
He
,
J.
,
Chai
,
Y.
,
Crump
,
J. G.
and
Maxson
,
R. E.
(
2022
).
Embryonic requirements for Tcf12 in the development of the mouse coronal suture
.
Development
149
,
dev199575
.
Tischfield
,
M. A.
,
Robson
,
C. D.
,
Gilette
,
N. M.
,
Chim
,
S. M.
,
Sofela
,
F. A.
,
DeLisle
,
M. M.
,
Gelber
,
A.
,
Barry
,
B. J.
,
MacKinnon
,
S.
,
Dagi
,
L. R.
et al. 
(
2017
).
Cerebral vein malformations result from loss of twist1 expression and BMP signaling from skull progenitor cells and dura
.
Dev. Cell
42
,
445
-
461.e5
.
Topczewska
,
J. M.
,
Shoela
,
R. A.
,
Tomaszewski
,
J. P.
,
Mirmira
,
R. B.
and
Gosain
,
A. K.
(
2016
).
The morphogenesis of cranial sutures in Zebrafish
.
PLoS ONE
11
,
e0165775
.
Tower
,
R. J.
,
Li
,
Z.
,
Cheng
,
Y.-H.
,
Wang
,
X.-W.
,
Rajbhandari
,
L.
,
Zhang
,
Q.
,
Negri
,
S.
,
Uytingco
,
C. R.
,
Venkatesan
,
A.
,
Zhou
,
F.-Q.
et al. 
(
2021
).
Spatial transcriptomics reveals a role for sensory nerves in preserving cranial suture patency through modulation of BMP/TGF-β signaling
.
Proc. Natl. Acad. Sci. U.S.A.
118
,
e2103087118
.
Tran
,
T. H.
,
Jarrell
,
A.
,
Zentner
,
G. E.
,
Welsh
,
A.
,
Brownell
,
I.
,
Scacheri
,
P. C.
and
Atit
,
R.
(
2010
).
Role of canonical Wnt signaling/β-catenin via Dermo1 in cranial dermal cell development
.
Development
137
,
3973
-
3984
.
Twigg
,
S. R. F.
and
Wilkie
,
A. O. M.
(
2015
).
A genetic-pathophysiological framework for craniosynostosis
.
Am. J. Hum. Genet.
97
,
359
-
377
.
Utvåg
,
S. E.
,
Grundnes
,
O.
and
Reikeraos
,
O.
(
1996
).
Effects of periosteal stripping on healing of segmental fractures in rats
.
J. Orthop. Trauma
4
,
279
-
284
.
van Gastel
,
N.
,
Torrekens
,
S.
,
Roberts
,
S. J.
,
Moermans
,
K.
,
Schrooten
,
J.
,
Carmeliet
,
P.
,
Luttun
,
A.
,
Luyten
,
F. P.
and
Carmeliet
,
G.
(
2012
).
Engineering vascularized bone: osteogenic and proangiogenic potential of murine periosteal cells
.
Stem Cells
30
,
2460
-
2471
.
Veistinen
,
L. K.
,
Mustonen
,
T.
,
Hasan
,
M. R.
,
Takatalo
,
M.
,
Kobayashi
,
Y.
,
Kesper
,
D. A.
,
Vortkamp
,
A.
and
Rice
,
D. P.
(
2017
).
Regulation of calvarial osteogenesis by concomitant de-repression of GLI3 and activation of IHH targets
.
Front. Physiol.
8
,
1036
.
Vivatbutsiri
,
P.
,
Ichinose
,
S.
,
Hytönen
,
M.
,
Sainio
,
K.
,
Eto
,
K.
and
Iseki
,
S.
(
2008
).
Impaired meningeal development in association with apical expansion of calvarial bone osteogenesis in the Foxc1 mutant
.
J. Anat.
212
,
603
-
611
.
Vortkamp
,
A.
,
Gessler
,
M.
and
Grzeschik
,
K.-H.
(
1991
).
GLI3 zinc-finger gene interrupted by translocations in Greig syndrome families
.
Nature
352
,
539
-
540
.
Wang
,
Y.
,
Xiao
,
R.
,
Yang
,
F.
,
Karim
,
B. O.
,
Iacovelli
,
A. J.
,
Cai
,
J.
,
Lerner
,
C. P.
,
Richtsmeier
,
J. T.
,
Leszl
,
J. M.
,
Hill
,
C. A.
et al. 
(
2005
).
Abnormalities in cartilage and bone development in the Apert syndrome FGFR2+/S252W mouse
.
Development
132
,
3537
-
3548
.
Warren
,
S. M.
,
Brunet
,
L. J.
,
Harland
,
R. M.
,
Economides
,
A. N.
and
Longaker
,
M. T.
(
2003a
).
The BMP antagonist noggin regulates cranial suture fusion
.
Nature
422
,
625
-
629
.
Warren
,
S. M.
,
Greenwald
,
J. A.
,
Nacamuli
,
R. P.
,
Fong
,
K. D.
,
Song
,
H. J. M.
,
Fang
,
T. D.
,
Mathy
,
J. A.
and
Longaker
,
M. T.
(
2003b
).
Regional dura mater differentially regulates osteoblast gene expression
.
J. Craniofac. Surg.
14
,
363
-
370
.
Wilk
,
K.
,
Yeh
,
S.-C. A.
,
Mortensen
,
L. J.
,
Ghaffarigarakani
,
S.
,
Lombardo
,
C. M.
,
Bassir
,
S. H.
,
Aldawood
,
Z. A.
,
Lin
,
C. P.
and
Intini
,
G.
(
2017
).
Postnatal calvarial skeletal stem cells expressing PRX1 reside exclusively in the calvarial sutures and are required for bone regeneration
.
Stem Cell Reports
8
,
933
-
946
.
Wilkie
,
A. O. M.
,
Slaney
,
S. F.
,
Oldridge
,
M.
,
Poole
,
M. D.
,
Ashworth
,
G. J.
,
Hockley
,
A. D.
,
Hayward
,
R. D.
,
David
,
D. J.
,
Pulleyn
,
L. J.
,
Rutland
,
P.
et al. 
(
1995
).
Apert syndrome results from localized mutations of FGFR2 and is allelic with Crouzon syndrome
.
Nat. Genet.
9
,
165
-
172
.
Wurdak
,
H.
,
Ittner
,
L. M.
,
Lang
,
K. S.
,
Leveen
,
P.
,
Suter
,
U.
,
Fischer
,
J. A.
,
Karlsson
,
S.
,
Born
,
W.
and
Sommer
,
L.
(
2005
).
Inactivation of TGFbeta signaling in neural crest stem cells leads to multiple defects reminiscent of DiGeorge syndrome
.
Genes Dev.
19
,
530
-
535
.
Xu
,
R.
,
Khan
,
S. K.
,
Zhou
,
T.
,
Gao
,
B.
,
Zhou
,
Y.
,
Zhou
,
X.
and
Yang
,
Y.
(
2018
).
Gαs signaling controls intramembranous ossification during cranial bone development by regulating both Hedgehog and Wnt/β-catenin signaling
.
Bone Res.
6
,
33
.
Yen
,
H.-Y.
,
Ting
,
M.-C.
and
Maxson
,
R. E.
(
2010
).
Jagged1 functions downstream of Twist1 in the specification of the coronal suture and the formation of a boundary between osteogenic and non-osteogenic cells
.
Dev. Biol.
347
,
258
-
270
.
Yin
,
L.
,
Du
,
X.
,
Li
,
C.
,
Xu
,
X.
,
Chen
,
Z.
,
Su
,
N.
,
Zhao
,
L.
,
Qi
,
H.
,
Li
,
F.
,
Xue
,
J.
et al. 
(
2008
).
A Pro253Arg mutation in fibroblast growth factor receptor 2 (Fgfr2) causes skeleton malformation mimicking human Apert syndrome by affecting both chondrogenesis and osteogenesis
.
Bone
42
,
631
-
643
.
Yoshida
,
T.
,
Phylactou
,
L. A.
,
Uney
,
J. B.
,
Ishikawa
,
I.
,
Eto
,
K.
and
Iseki
,
S.
(
2005
).
Twist is required for establishment of the mouse coronal suture
.
J. Anat.
206
,
437
-
444
.
Yoshida
,
T.
,
Vivatbutsiri
,
P.
,
Morriss-Kay
,
G.
,
Saga
,
Y.
and
Iseki
,
S.
(
2008
).
Cell lineage in mammalian craniofacial mesenchyme
.
Mech. Dev.
125
,
797
-
808
.
Yu
,
M.
,
Ma
,
L.
,
Yuan
,
Y.
,
Ye
,
X.
,
Montagne
,
A.
,
He
,
J.
,
Ho
,
T.-V.
,
Wu
,
Y.
,
Zhao
,
Z.
,
Sta Maria
,
N.
et al. 
(
2021
).
Cranial suture regeneration mitigates skull and neurocognitive defects in craniosynostosis
.
Cell
184
,
243
-
256.e18
.
Zarbalis
,
K.
,
Siegenthaler
,
J. A.
,
Choe
,
Y.
,
May
,
S. R.
,
Peterson
,
A. S.
and
Pleasure
,
S. J.
(
2007
).
Cortical dysplasia and skull defects in mice with a Foxc1 allele reveal the role of meningeal differentiation in regulating cortical development
.
Proc. Natl. Acad. Sci. U.S.A.
104
,
14002
-
14007
.
Zhao
,
H.
,
Feng
,
J.
,
Ho
,
T.-V.
,
Grimes
,
W.
,
Urata
,
M.
and
Chai
,
Y.
(
2015
).
The suture provides a niche for mesenchymal stem cells of craniofacial bones
.
Nat. Cell Biol.
17
,
386
-
396
.

Competing interests

The authors declare no competing or financial interests.