Development is regulated by coordinated changes in gene expression. Control of these changes in expression is largely governed by the binding of transcription factors to specific regulatory elements. However, the packaging of DNA into chromatin prevents the binding of many transcription factors. Pioneer factors overcome this barrier owing to unique properties that enable them to bind closed chromatin, promote accessibility and, in so doing, mediate binding of additional factors that activate gene expression. Because of these properties, pioneer factors act at the top of gene-regulatory networks and drive developmental transitions. Despite the ability to bind target motifs in closed chromatin, pioneer factors have cell type-specific chromatin occupancy and activity. Thus, developmental context clearly shapes pioneer-factor function. Here, we discuss this reciprocal interplay between pioneer factors and development: how pioneer factors control changes in cell fate and how cellular environment influences pioneer-factor binding and activity.

During development, the single-celled zygote divides and differentiates, eventually giving rise to all the cells of an adult organism. The cells of the adult are specialized and phenotypically distinct, but all share a common genome formed at fertilization. Therefore, it is the differential interpretation of the shared genome that determines cell fate. Transcription factors orchestrate this process by recognizing and binding to specific DNA motifs to drive gene expression. However, the packaging of DNA into chromatin is a barrier to transcription-factor binding. The nucleosome is the basic repeating unit of chromatin and is composed of approximately 147 base pairs of DNA wrapped twice around an octamer of histones, comprising two each of H2A, H2B, H3 and H4. H1 linker histones stabilize nucleosomes and help establish higher-order chromatin structure, including arrays of nucleosomes (Fyodorov et al., 2018; Kornberg and Thomas, 1974; Li et al., 2007). Nucleosome compaction enables the approximately two meters of DNA to fit in the eukaryotic nucleus. This compact chromatin structure prevents many transcription factors from accessing their binding motifs. Instead, transcription factors often bind their motifs within open and accessible chromatin. By contrast, a specialized class of transcription factors, termed pioneer factors, bind closed, nucleosomal DNA to promote regions of open chromatin that can be subsequently bound by other transcription factors. Because of the ability to define open chromatin, pioneer factors act at the top of gene-regulatory networks to define cell fate.

Despite their ability to target closed, nucleosome-occupied binding sites, pioneer factors do not bind all genomic instances of their target motif. Instead, they have cell type-specific patterns of binding and activity in promoting chromatin accessibility. Emerging evidence suggests that pioneer-factor occupancy and activity is regulated by the cell type in which it is expressed. Below, we explore the reciprocal interplay between pioneer factors and development. We initially provide a perspective on the role of pioneer factors in driving developmental transitions. We then discuss the mechanism by which pioneer factors bind chromatin and generate accessible cis-regulatory regions. Finally, we focus on how developmental context can influence the ability of pioneer factors to access the genome and promote gene expression.

Because of their ability to promote widespread changes in gene expression, pioneer factors drive developmental transitions, and their dysregulation results in disease.

Cellular differentiation and tissue specification

Pioneer factors promote developmental transitions by defining the cis-regulatory regions of the gene-regulatory networks that drive specific cell fates. The first pioneer factor to be identified was Foxa1, discovered in mice based on its ability to bind and open the albumin enhancer, necessary for liver specification, prior to transcriptional activation of this gene (Gualdi et al., 1996). During endoderm differentiation, this initial opening of chromatin by Foxa1 facilitates the subsequent binding of tissue-specific transcription factors needed for lineage specification (Cirillo et al., 2002; Iwafuchi-Doi et al., 2016; Wang et al., 2015). Foregut endoderm-specific knockout of Foxa1 and Foxa2 results in the failure to initiate hepatic differentiation, highlighting the essential role of this family of pioneer factors in promoting the gene expression program required for normal development of the liver (Lee et al., 2005). In humans, the related factor FOXA2 is similarly required for pancreatic differentiation from pluripotent stem cells (Lee et al., 2019). Since the initial characterization of the pioneering functions of Foxa1, the vital roles of many different pioneer factors in promoting developmental transitions have been identified across a range of species (Balsalobre and Drouin, 2022; Barral and Zaret, 2024; Horisawa and Suzuki, 2023; Larson et al., 2021; Mayran and Drouin, 2018).

The importance of pioneer factors in development is evident from the conservation of their pioneering function across species. For example, in Caenorhabditis elegans, the FoxA ortholog PHA-4 is a pioneer factor required for specifying foregut fate. Like the mammalian FoxA family members, it binds nucleosomes in vitro and promotes chromatin accessibility and gene expression in vivo (Hsu et al., 2015). The Grainy head (GRH) family of transcription factors are a set of conserved pioneer factors defined by their similarity to the Grainy head protein that was originally discovered in the fruit fly (Drosophila melanogaster) (Bray et al., 1989; Dynlacht et al., 1989; Nüsslein-Volhard et al., 1984; Venkatesan et al., 2003). In Drosophila, Grh promotes chromatin accessibility and is required for expression of genes necessary for defining epithelial cell fate (Jacobs et al., 2018; Nevil et al., 2020; Wang and Samakovlis, 2012). The role of this pioneer factor in epithelial cell-fate specification is conserved in worms, frogs and mammals (Gasperoni et al., 2022). Mammals have three Grainy head-like (GRHL) proteins (Reese et al., 2019; Wilanowski et al., 2002). Expression of GRHL2 in naïve embryonic stem cells (ESCs) is required to maintain expression of epithelial genes as ESCs transition to formative epiblast-like cells (EpiLCs) (Chen et al., 2018). In the absence of GRHL2, EpiLCs fail to adopt an epithelial fate and instead transition to a mesenchymal state (Chen et al., 2018). These few examples highlight the conserved role of pioneer factors in promoting cell-fate specification in organisms ranging from nematodes to humans.

To enhance cell-fate transitions, pioneer factors not only activate specific gene-expression programs, but also shut off gene-expression networks for alternative cell fates. During mammalian pituitary development, Pax7 defines melanotrope identity by stimulating expression of melanotrope-specific genes and repressing genes that drive corticotrope fate (Budry et al., 2012). PU.1 (also known as SPI1) is a pioneer factor that plays a role in immune cell differentiation (Li et al., 2020) and functions through both activating and repressing gene expression (Li et al., 2020; Ungerbäck et al., 2018). This repression is thought to be indirect through the sequestration of factors away from non-PU.1 targets (Hosokawa et al., 2018). As hair follicle stem cells differentiate from multipotent embryonic progenitors, SOX9 opens enhancers of genes defining hair follicle stem cells. This results in the simultaneous silencing of epidermal enhancers through recruitment of co-factors away from these alternative-lineage enhancers (Yang et al., 2023). Pioneer factors can also directly facilitate repression. For example, all three FOXA proteins are important for silencing expression of alternative lineages during foregut differentiation. This is likely mediated by their recruitment of chromatin factors, including the enzyme PRDM1, which promotes a chromatin state that represses gene expression (Matsui et al., 2024). The pioneer factor OCT4 (also known as POU5F1) also interacts with a related chromatin modifying enzyme, PRDM14, suggesting this role in transcriptional repression may be shared amongst a distinct class of pioneer factors (Matsui et al., 2024). Thus, pioneer factors promote robust cell-fate changes by directly initiating chromatin accessibility at enhancers to drive gene expression and by repressing expression of genes that specify alternative cell fates.

Reprogramming in embryos and in culture

The capacity of pioneer factors to prompt large-scale changes in gene expression is required for cellular reprogramming both in the early embryo and in cell culture. After fertilization, the genome of the unified gametes is reprogrammed to generate the totipotent state of the early embryo. This conserved developmental transition relies on maternally deposited mRNAs and proteins, as the zygotic genome is transcriptionally quiescent. Gradual activation of zygotic transcription is coordinated with the degradation of these maternal products during the maternal-to-zygotic transition (MZT) (Schulz and Harrison, 2019; Vastenhouw et al., 2019). In all organisms studied to date, pioneer factors drive the efficient reprogramming of the zygotic genome during these early developmental stages. The Drosophila protein Zelda was the first major activator of the zygotic genome to be discovered (Liang et al., 2008). Zelda is required broadly for restructuring chromatin accessibility and influences the 3D structure of the genome (Hug et al., 2017; Schulz and Harrison, 2019; Schulz et al., 2015; Sun et al., 2015; Brennan et al., 2023). Zelda is not the only pioneer factor essential for the progressive activation of zygotic transcription in Drosophila. Additional pioneer factors, including GAGA factor (GAF; also known as Trl) and CLAMP, function with Zelda to restructure and activate the early embryonic genome (Duan et al., 2021; Gaskill et al., 2021). Whereas Zelda is preferentially required for the earliest wave of zygotic gene expression, GAF is required slightly later in development (Gaskill et al., 2021). Following the MZT, embryonic patterning is established by another pioneer-like factor, Odd paired (Opa) (Koromila et al., 2020; Soluri et al., 2020). These studies highlight the importance of sequential pioneer-factor activity in promoting the highly conserved reprogramming that occurs following fertilization.

Since the discovery of Zelda in flies, the essential role of pioneer factors in early embryonic reprogramming has been established in numerous other organisms. In zebrafish (Danio rerio), the pioneer factors Pou5f3, Sox19b and Nanog cooperatively regulate zygotic genome activation (Lee et al., 2013; Leichsenring et al., 2013; Miao et al., 2022; Veil et al., 2019). Similarly, in frogs (Xenopus laevis), Pou5f3 (also known as Oct4), Sox3 and Foxh1 remodel early embryonic chromatin to regulate activation of the zygotic genome (Blitz and Cho, 2021; Charney et al., 2017; Gentsch et al., 2019; Paraiso et al., 2019). The conserved pioneer factors DUX4 and Dux, found in humans and mice, respectively, promote chromatin accessibility and gene expression during the earliest stages of embryo development (De Iaco et al., 2017; Hendrickson et al., 2017; Whiddon et al., 2017). However, some Dux mutant embryos are viable, indicating that Dux is important but not essential for mouse embryogenesis (Chen and Zhang, 2019; De Iaco et al., 2017). Like flies, zebrafish and frogs, mammalian embryonic reprogramming involves multiple pioneer factors. The PRD-like homeobox domain transcription factor family (Obox1-Obox8) and orphan nuclear receptor Nr5a2 are key activators of zygotic transcription in mice (Gassler et al., 2022; Ji et al., 2023). Although neither Dux nor Obox4 is individually essential for early mouse development, they are redundantly required for viability and zygotic genome activation (Guo et al., 2024 preprint). The parallel requirement for the cooperative action of pioneer factors in early embryonic reprogramming in many species highlights the importance of understanding how these proteins function together to robustly drive this conserved developmental transition.

In vitro reprogramming of differentiated cells into induced pluripotent stem cells (iPSCs) through the expression of a cocktail of transcription factors shares many features with the much more efficient in vivo reprogramming that occurs in the early embryo. The initial cocktail of transcription factors used to generate iPSCs from mammalian fibroblasts was Oct4, Sox2, Klf4 and Myc (OSKM) (Takahashi and Yamanaka, 2006). Although at the time the underlying mechanisms by which these factors drove reprogramming was unclear, it has since become evident that the unique ability of these factors to reprogram relies on their pioneering abilities. Oct4, Sox2 and Klf4 (OSK) are pioneer factors and are sufficient to drive reprogramming independently of Myc (Nakagawa et al., 2008; Soufi et al., 2012, 2015). Nonetheless, Myc enhances OSK binding (Soufi et al., 2012). As in early embryonic reprogramming, OSK function cooperatively in the generation of iPSCs by binding and opening enhancers that drive the pluripotency gene network as well as silencing the somatic program by promoting the decommissioning of somatic enhancers (Chronis et al., 2017; Li et al., 2017). Nucleosome binding is correlated with the reprogramming ability of transcription factors (Fernandez Garcia et al., 2019), suggesting that pioneering activity is a driver of efficient reprogramming. Together, these examples highlight the importance of collaborative interactions among multiple pioneer factors in activating a pluripotency network and erasing the features of a prior cellular state.

Transdifferentiation

Pioneering activity is required for transdifferentiation, the process by which one mature somatic cell is transformed into another somatic cell type without transitioning through an intermediate, less-differentiated cell type. This is exemplified by the conversion of mouse or human fibroblasts into functional dopaminergic neurons through a cocktail of transcription factors that primarily depends on the pioneer factor Ascl1/ASCL1 (Marro and Yang, 2014; Vierbuchen et al., 2010; Wapinski et al., 2013). ASCL1 is a proneural basic helix-loop-helix (bHLH) transcription factor that induces neural differentiation in vivo (Bertrand et al., 2002). When exogenously expressed in fibroblasts, ASCL1 occupies most of the same sites it normally occupies in neurons and promotes chromatin accessibility. Binding to these sites persists as cells undergo transdifferentiation into neurons (Wapinski et al., 2013). FoxA proteins are also important in transdifferentiation. When expressed in mouse fibroblasts along with either Hnf1α and Gata4 or Hnf4α, FoxA proteins direct reprogramming into hepatocyte-like cells. In this context, all three members of the FoxA family function as pioneer factors (Huang et al., 2011; Sekiya and Suzuki, 2011). With the expression of other co-factors, Foxa2 can direct transdifferentiation of fibroblasts into dopaminergic neurons (Pfisterer et al., 2011).

Pioneer factors maintain developmental trajectories through mitosis

During mitosis, the nuclear envelope breaks down, the genome becomes highly condensed, transcription ceases, the 3D genome organization is largely disrupted, and many transcription factors dissociate from the chromosomes (Antonin and Neumann, 2016; Gottesfeld, 1997; Naumova et al., 2013). Nonetheless, the gene-regulatory networks that maintain cell-fate specification must be re-established following division. How this process is achieved is not well understood. Even though most transcription factors do not remain associated with the mitotic chromosomes, a small subset of factors is retained. Some of these factors are required for proper reactivation of gene expression upon mitotic exit, a mechanism termed ‘mitotic bookmarking’. Pioneer factors comprise a large portion of those factors that are retained on mitotic chromosomes (Bellec et al., 2022; Caravaca et al., 2013; Chervova et al., 2024; Deluz et al., 2016; Festuccia et al., 2016, 2019; Liu et al., 2017; Nevil et al., 2020; Price et al., 2023; Silvério-Alves et al., 2023; Teves et al., 2016), suggesting that the capacity to bind closed chromatin may support retention on the compacted mitotic chromatin. However, not all pioneer factors are mitotically retained, for example both Zelda (in flies) and Ascl1 (in mice) are evicted from chromatin during mitosis (Dufourt et al., 2018; Soares et al., 2021). Additionally, some non-pioneer factors are retained on mitotic chromatin and retention is not correlated with pioneering activity (Raccaud et al., 2019). Thus, the features that define a pioneer factor are separable from those that enable mitotic retention.

Mitotic bookmarking by pioneer factors may promote rapid gene activation and the preservation of cellular identity following division. This is a specific challenge during early embryogenesis owing to the rapid division cycles. During the MZT in Drosophila, GAF is mitotically retained and required for the rapid activation of a reporter gene and transcriptional memory across mitosis (Bellec et al., 2022). By contrast, Zelda, which is not mitotically retained, is not necessary for maintaining this transcriptional memory through mitosis (Dufourt et al., 2018). In mouse ESCs, the pioneer factors Esrrb and Nr5a2 both coat mitotic chromosomes and, together, facilitate reactivation of gene expression following mitosis (Chervova et al., 2024). A challenge to these studies is the difficulty in specifically depleting proteins only during mitosis. Most studies have relied on generating proteins tagged with a domain of cyclin B1 that drives mitotic degradation during the metaphase-anaphase transition (Kadauke et al., 2012). This system has been used to demonstrate that mitotic expression of the reprogramming factors Oct4 and Sox2 is required for maintaining pluripotency and mouse ESC differentiation (Deluz et al., 2016; Liu et al., 2017). Although most of these studies have been performed in culture, mice homozygous for Gata2 fused with a mitotic degron phenocopied the Gata2 knockout, suggesting that mitotic expression of Gata2 is necessary for function (Silvério-Alves et al., 2023). However, it is not clear how rapidly protein expression is re-established following mitotic degradation, making it challenging to disentangle a role during mitosis from that during early interphase. Future, more acute methods of disrupting the mitotic localization of pioneer factors will be essential for determining the functional significance of their retention on mitotic chromatin.

Pioneer factor dysregulation promotes disease

Disturbances to the finely tuned processes governing cell-fate decisions culminate in developmental disorders and oncogenic transformations. As a result of the power of pioneer factors to reshape cell fate, their misexpression can promote disease states. Pioneer factor expression levels are indicators of poor prognosis for a growing list of cancers (Chiou et al., 2010; Jozwik and Carroll, 2012; Lemma et al., 2022; Matsuoka et al., 2012; Saigusa et al., 2009; Sholl et al., 2010; Sunkel and Stanton, 2021).

Cancer stem cells (CSCs) are drivers of new tumors in many cancers (Reya et al., 2001). Like ESCs, CSCs are capable of self-renewal, and this capacity is thought to promote tumor formation (Yu et al., 2012; Reya et al., 2001). The shared ability to self-renew suggests that common regulators may exist between CSCs and ESCs (Reya et al., 2001). Indeed, the core reprogramming factors OCT4 and NANOG have been identified as CSC markers (Chiou et al., 2008), and their overexpression leads to tumor formation in mice and is associated with aggressive nasopharyngeal carcinoma and metastasis in humans (Liu et al., 2024; Luo et al., 2013; Ohnishi et al., 2014; Wei et al., 2014). The pioneer factor DUX4 is commonly expressed in metastatic tumors, and its expression is correlated with poor survival rates (Pineda and Bradley, 2024). This suggests a causal relationship between pioneer factor-mediated reprogramming and tumor pathogenicity. These and numerous additional examples demonstrate how overexpression of a pioneer factor can drive a stem-cell-like fate, which, in the wrong context, leads to unwanted proliferation and cancer.

Pioneer factors also facilitate cancer progression through their ability to promote distinct cellular fates. The GRHL family of transcription factors suppress the epithelial-to-mesenchymal transition, a cell-fate transition that promotes metastasis (Cieply et al., 2012, 2013). Based on this, GRHL factors were thought to function as tumor suppressors. However, GRHL family members function as both tumor suppressors and oncogenes depending on the cell type, highlighting the complex nature of cancer progression (Reese et al., 2019). This may be due to the diversity of co-factors expressed in the different cell types from which these cancers arise. For example, in estrogen receptor (ER)-positive breast cancer cell lines, GRHL2 influences chromatin binding of phosphorylated ER, likely through the promotion of chromatin accessibility (Reese et al., 2022). In androgen receptor-expressing prostate cell lines, GRHL2 instead promotes binding and expression of this hormone receptor (Paltoglou et al., 2017). FOXA1 pioneer factors also influence hormone-receptor binding and may even function together with GRHL2 in some cell types (Carroll et al., 2005; Cocce et al., 2019; Jozwik et al., 2016; Seachrist et al., 2021; Wang and Yin, 2015). Through their capacity to promote chromatin accessibility, pioneer factors allow unique cell type-specific factors to bind DNA and activate gene expression. The role of pioneer factors in disease is further exemplified by the misexpression of DUX4 in myoblasts, which results in the re-activation of genes and retrotransposons typically expressed during early embryonic development (Mocciaro et al., 2021; Young et al., 2013). Thus, the dysregulation of pioneer factors can lead to disease by promoting improper cell fates.

Translocations can generate chimeric pioneer factors with neomorphic functions that promote the disease state. For example, a translocation resulting in the expression of a chimera between PAX3 and FOXO1 is associated with rhabdomyosarcoma. The PAX3-FOXO1 chimera has features of pioneer factors, such as recognition of motifs within nucleosomes and activation of gene-regulatory networks involved in tumor development (Galili et al., 1993; Sunkel and Stanton, 2021). Another example of a chimeric oncoprotein is EWS-FLI1, which acts as an oncogenic pioneer factor to initiate the Ewing sarcoma gene-regulatory program in mesenchymal stem cells and is required for the growth and survival of these cancer cells (Riggi et al., 2008, 2014).

The unique ability of pioneer factors to access the genome and promote widespread changes in gene expression allows them to drive broad developmental transitions, but dysregulation of this activity results in disease. Thus, determining the mechanisms by which these proteins function will have important implications for understanding normal developmental processes and advancing targeted therapeutics.

Pioneer factors elicit changes in gene expression and cell identity by defining cell type-specific cis-regulatory regions. Promoters typically maintain accessibility across cell types so, although some pioneer factors in flies and mammals are associated with broadly accessible promoters, this binding is unlikely to drive developmental transitions because accessibility is shared across tissue types (Biggin and Tjian, 1988; Fuda et al., 2015; Lee et al., 2007; Li et al., 1994). However, a set of promoters have been identified in Drosophila that gain paused RNA polymerase II and chromatin accessibility over embryonic development, and this is driven by the pioneer factor Lola-I (Gaertner et al., 2012; Ramalingam et al., 2023). In contrast to promoters, accessibility at enhancer regions is regulated. Open enhancers are indicative of active gene expression, and pioneer factors largely elicit changes in gene expression by promoting accessibility at distally located enhancer elements (Spitz and Furlong, 2012).

Although pioneer factors share the ability to bind and open closed chromatin, the mechanisms by which such factors engage nucleosomal DNA and drive accessibility are varied (Zaret, 2020; Stoeber et al., 2024). Beyond creating accessible regions, pioneer factors can also influence chromatin by recruiting enzymes that chemically modify the histone tails or DNA and by regulating the 3D structure of the genome (Fig. 1).

Fig. 1.

Pioneer factors influence chromatin through diverse mechanisms. (A) Pioneer factors (PF, orange) bind nucleosomes and promote local DNA unwinding (indicated by orange lines). Binding may be stabilized by sequence-specific co-factors. Subsequent chromatin opening can occur through multiple mechanisms: pioneer-factor binding alone (top); recruitment of ATP-dependent chromatin remodelers (green) (middle); interaction with enzymes that catalyze post-translational modifications to the histones or DNA (purple) (bottom). (B) Many pioneer factors localize to sub-nuclear microenvironments that can mediate interactions between distal regions of chromatin (orange). Regions outside the DNA-binding domains of pioneer factors promote self-association, which may drive the formation of these microenvironments.

Fig. 1.

Pioneer factors influence chromatin through diverse mechanisms. (A) Pioneer factors (PF, orange) bind nucleosomes and promote local DNA unwinding (indicated by orange lines). Binding may be stabilized by sequence-specific co-factors. Subsequent chromatin opening can occur through multiple mechanisms: pioneer-factor binding alone (top); recruitment of ATP-dependent chromatin remodelers (green) (middle); interaction with enzymes that catalyze post-translational modifications to the histones or DNA (purple) (bottom). (B) Many pioneer factors localize to sub-nuclear microenvironments that can mediate interactions between distal regions of chromatin (orange). Regions outside the DNA-binding domains of pioneer factors promote self-association, which may drive the formation of these microenvironments.

Close modal

Mechanisms of chromatin binding

To promote changes in chromatin accessibility, pioneer factors must first access and bind to nucleosomal DNA. Although this is a defining feature, pioneer factors have a wide diversity of DNA-binding domains (DBDs) and access the DNA through distinct mechanisms (Barral and Zaret, 2024). Structural studies have suggested multiple mechanisms, including binding partial motifs, that enable pioneer factors to engage DNA even when it is wrapped around histones (Fernandez Garcia et al., 2019; Ozden et al., 2023; Soufi et al., 2015). The location of a pioneer factor binding motif on the nucleosome might also regulate pioneer-factor binding. High-throughput in vitro studies of hundreds of human transcription factors (413 DBD and 46 full-length proteins) identified a diversity of ways that pioneer factors bind nucleosomes (Zhu et al., 2018). Many factors prefer to bind to the entry/exit sites of nucleosomes, but others bind periodically to the solvent-exposed side of DNA or to the dyad axis (center of the nucleosomal DNA) (Zhu et al., 2018). For example, both mammalian and Drosophila pioneer factors, Nr5a2 and Lola-I, respectively, bind preferentially to motifs located at the edge of the nucleosome (Gassler et al., 2022; Ramalingam et al., 2023). Binding at the entry/exit site may be facilitated by transient exposure of DNA as it relaxes from the nucleosome (Isaac et al., 2016; Poirier et al., 2008). Although extensive work has characterized binding in vitro, how nucleosome positioning affects pioneer-factor binding in vivo remains less clear. Analysis of the binding preferences of Grh and Zelda in Drosophila cell culture using MNase (micrococcal nuclease digestion with deep sequencing) and ChIP-seq (chromatin immunoprecipitation followed by sequencing) data failed to identify a preferred position for their binding motif within the nucleosome (Gibson et al., 2024). Indeed, nucleosomes at silent regions of chromatin are not well-positioned, suggesting that the exact positioning of a binding motif on a nucleosome may not strongly regulate pioneer-factor binding within a cell (Stergachis et al., 2020; Abdulhay et al., 2020).

Some pioneer factors rely on engagement with histones for nucleosome binding. For example, Foxa1 interacts with the core histones through a C-terminal domain, and this domain is required for chromatin accessibility in the mouse embryo (Cirillo et al., 2002; Iwafuchi et al., 2020). Based on its crystal structure, the yeast pioneer factor Cbf1 is speculated to interact with the H2A tail (Donovan et al., 2023), and the pluripotency factor OCT4 interacts directly with the histone H3 tail through its DBD (Echigoya et al., 2020; MacCarthy et al., 2022). These examples highlight how the structure and composition of nucleosomes can affect pioneer factor–nucleosome interactions.

Co-factor interactions may also contribute to stable nucleosome binding, and nucleosome engagement can be regulated by the cooperative activities of multiple factors. In mouse ESCs, the nucleosome binding activity of Parp1 stabilizes occupancy of the pioneer factor Sox2, possibly by specifically promoting Sox2 binding to nucleosomes in which the DNA motif is positioned sub-optimally for Sox2 engagement (Liu and Kraus, 2017). In addition to proteins that generally stabilize interactions with chromatin, sequence-specific factors can also regulate pioneer-factor occupancy. OCT4 binding to the nucleosome edge is enhanced by SOX2 co-binding (Michael et al., 2020), and binding of ASCL1 to nucleosomes is promoted by heterodimerization with E12α (Fernandez Garcia et al., 2019). Similarly, the transcription factor AP-1 may facilitate binding of different pioneer factors, including Foxa1 and GATA3, in distinct cell types (Bi et al., 2020; Xu et al., 2023 preprint). The expression of multiple pioneer factors can also regulate occupancy. The pioneer factors Nanog, Pou5f3 and Sox19b are required for reprogramming of the zygotic genome in the zebrafish embryo and can bind either independently or cooperatively depending on the genomic locus (Miao et al., 2022). Thus, co-factors can direct pioneer-factor binding and are important for stable occupancy of pioneer factors at some loci.

Like all transcription factors, pioneer factors rely on their DBD for interactions with the genome. Indeed, the DBDs of many factors, including Foxa1, Nr5a2, Cbf1 and Zelda, are sufficient to bind mono-nucleosomes in vitro, demonstrating that these diverse binding domains can engage DNA when wrapped around histones (Cirillo, 1998; Donovan et al., 2023; Fernandez Garcia et al., 2019; Gassler et al., 2022; McDaniel et al., 2019). Nonetheless, these domains alone are often insufficient to stably occupy chromatin and additional protein regions are required for pioneering activity. For example, regions outside of the zinc-finger DBD of Zelda are required to facilitate robust binding to mono-nucleosomes in vitro (Fernandez Garcia et al., 2019; McDaniel et al., 2019). The in vivo chromatin environment is more complex than the nucleosomes used in vitro, and regions outside the DBD of pioneer factors are required for stable occupancy. For example, the DBD of Drosophila pioneer factors Grh and Zelda are not able to bind regions of closed chromatin that are pioneered by the full-length proteins (Gibson et al., 2024). Similarly, single-molecule assays of pioneer factors FOXA1 and SOX2 in human fibroblasts demonstrated that their DBDs alone are less efficient than the full-length proteins at scanning silent chromatin (Lerner et al., 2023). Overall, although pioneer factors share the ability to bind closed chromatin, they vary in the mechanisms by which they engage the genome.

Mechanisms of chromatin opening

Depending on how pioneer factors bind chromatin, they can promote accessibility either directly or indirectly. Pioneer factors can directly influence local chromatin opening by binding to the nucleosome (Cirillo et al., 2002; Donovan et al., 2019; Guan et al., 2023; Michael et al., 2020; Mivelaz et al., 2020; Sinha et al., 2023; Zhu et al., 2018). The DBD of FoxA family members are winged helix domains that are structurally similar to linker histones (Dai et al., 2021), allowing them to displace linker histone H1 and promote accessibility independently of additional factors both in vitro and in the mouse liver (Cirillo et al., 2002; Iwafuchi-Doi et al., 2016; Taube et al., 2010). In vitro binding of OCT4, SOX2 and SOX11 can destabilize the interactions of DNA with the core histones and more broadly shape inter-nucleosome interactions by repositioning histone tails (Dodonova et al., 2020; Guan et al., 2023; Michael et al., 2020). Regions outside the DBD may be important for this intrinsic ability to open chromatin, as is the case for PU.1 and OCT4 in vitro (Frederick et al., 2023; Sinha et al., 2023). Although the mechanisms of chromatin opening, like chromatin binding, are diverse, many pioneer factors share the intrinsic ability to directly open chromatin.

Pioneer factors also indirectly mediate chromatin accessibility via interactions with ATP-dependent chromatin remodelers, enzymes that can restructure chromatin by altering the interaction between histones and DNA (Fig. 1A) (Ahmad et al., 2024; Clapier et al., 2017). Even pioneer factors that can initially open chromatin independently of other factors may require chromatin remodelers to fully remodel the genome. For example, in vitro, PU.1 alone opens chromatin within an array of multiple nucleosomes, but recruitment of the chromatin remodeler SWI/SNF promotes further chromatin opening, which may lead to more stable, widespread opening of chromatin (Frederick et al., 2023; Minderjahn et al., 2020). Oct4 recruits the central catalytic subunit of SWI/SNF chromatin remodeling complex, Brg1 (Smarca4), to shape the chromatin accessibility at target sites in mouse ESCs (Huang et al., 2021; King and Klose, 2017). GATA3 similarly depends on BRG1 to create accessible chromatin regions during the mesenchymal-to-epithelial transition in human breast cancer cells (Takaku et al., 2016). Numerous other pioneer factors have been shown to interact with members of the SWI/SNF complex (Gouhier et al., 2024; Judd et al., 2021; Mivelaz et al., 2020; Păun et al., 2023; Wolf et al., 2023), suggesting that many pioneer factors may function through SWI/SNF to stably open chromatin and promote expression of gene-regulatory networks. A single pioneer factor can also synergize with multiple ATP-dependent chromatin remodelers to open chromatin and drive gene expression. For example, GAF interacts with remodelers from both the SWI/SNF family and ISWI families (Lomaev et al., 2017; Nakayama et al., 2012); its interactions with PBAP (SWI/SNF) promote chromatin accessibility and RNA polymerase II recruitment, whereas its interactions with NURF (ISWI) ensure the efficient release of elongating RNA polymerase II (Judd et al., 2021).

Pioneer factors alter post-translational modifications to DNA and chromatin

Pioneer factors recruit histone-modifying enzymes to elicit changes in the chromatin structure (Fig. 1A). Chromatin structure can be modified by multiple post-translational modifications of histone tails. These histone modifications act as molecular marks associated with the active and inactive cis-regulatory regions that are crucial for proper development (Dong and Weng, 2013; Kouzarides, 2007). The effects of histone modifications are widespread, including functioning as binding platforms for specific proteins, promoting formation of an open chromatin structure, and establishing enhancer–promoter communication.

Active enhancers are accessible regions marked by acetylation on lysine 27 of histone 3 (H3K27ac) (Wang et al., 2008). Recruitment of histone acetyltransferases, such as p300 (Ep300) and/or CBP (CREBBP), which are general transcription co-activators, is responsible for H3K27ac at nucleosomes flanking active enhancers (Calo and Wysocka, 2013). Deposition of histone acetylation marks, including H3K27ac, correlates with zygotic genome activation (Li et al., 2014; Vastenhouw et al., 2019). Indeed, the pioneer factors Nanog, Pou5f3 and Sox19b are necessary for occupancy of p300 at target genes in zebrafish embryos. p300 is recruited to the genome through these pioneer factors and is required for zygotic genome activation (Chan et al., 2019; Miao et al., 2022). NANOG interacts with p300 in mammalian ESCs (Fang et al., 2014), and recruitment of p300/CBP is required for DUX4-mediated transcriptional activation in human myoblasts (Choi et al., 2016), suggesting that there may be a general requirement for interactions with this histone acetyltransferase. In Drosophila, the dependency of enhancer-associated histone 3 lysine 18 acetylation (H3K18ac) on the pioneer factor Zelda suggests that Zelda may also recruit a histone acetyltransferase (Li et al., 2014). Despite these conserved interactions, in zebrafish, most Nanog, Pou5f3 and Sox19b target genes remain open upon inhibition of p300, suggesting that p300-mediated acetylation may not be required to promote chromatin accessibility during zygotic genome activation (Miao et al., 2022). Thus, pioneer factors that drive zygotic genome activation promote acetylation but the relationship between this chromatin mark, pioneer activity, and gene expression remains incompletely understood.

Pioneer factors also recruit chromatin-modifying enzymes to elicit changes in the histone methylation state. FOXA1 recruits the histone methyltransferase MLL3 (KMT2C) to promote methylation of lysine 4 on histone 3 (H3K4me1), which in turn increases activation of ER-responsive genes in breast cancer cells (Jozwik et al., 2016). In mouse pituitary cells, Pax7 first recruits a histone demethylase, KDM1A, to remove the repressive H3K9me2 mark followed by recruitment of the MLL complex to deposit the activating H3K4me1/3 mark (Gouhier et al., 2024; Kawabe et al., 2012; McKinnell et al., 2008).

Chromatin structure is also modified through methylation of DNA. Pioneer factors regulate DNA methylation actively by recruiting demethylase enzymes, or passively promote the loss of methylation as the cell cycle progresses. In mouse muscle stem cells, DNA demethylation is mediated by Pax7 and is a crucial event for gene activation (Carrió et al., 2016). FOXA1 exhibits physical interactions with the demethylases TET1 and TET2 in vitro and in vivo, and its knockdown in breast cancer cells leads to heightened methylation within FOXA1-bound regions, suggesting that FOXA1 has a role in inducing site-specific demethylation that is TET1/2 dependent (Lemma et al., 2022). Klf4 also promotes DNA demethylation through interactions with Tet2 in mouse ESCs (Sardina et al., 2018). By contrast, FOXA2 and SOX2 binding leads to a loss of DNA methylation that is replication dependent, suggesting a passive loss of DNA methylation upon pioneer-factor binding, without the active recruitment of DNA demethylases (Donaghey et al., 2018; Vanzan et al., 2021). In summary, pioneer factors regulate different developmental gene networks by modulating histone and DNA modifications in unique ways.

Pioneer factors influence 3D chromatin structure

To package chromatin in the nucleus, it is structurally organized in 3D space. This packaging is organized at multiple scales from compartments of active (A) and inactive (B) chromatin to topologically associated domains (TADs) to loops (Theis and Harrison, 2023). Chromatin structure can govern long- and short-range DNA interactions, allowing for the proper regulation of gene expression and cell identity. TAD boundaries predominantly function to prevent unwanted interactions, whereas loop boundaries, which are governed by tethering elements, reflect enhancer–promoter interactions (Batut et al., 2022). These boundaries form sequentially during embryonic development, first at the level of chromatin loops and then further into TADs (Theis and Harrison, 2023). Pioneer-factor occupancy has been implicated in shaping this 3D genome organization. During the establishment of 3D chromatin structure in the early Drosophila embryo, binding sites for the pioneer factors Zelda and GAF are located at loop and TAD boundaries (Fig. 1B) (Batut et al., 2022; Hug et al., 2017; Levo et al., 2022; Ogiyama et al., 2018). Zelda is required for the initial formation of loops (Espinola et al., 2021) and later for defining individual TAD boundaries (Hug et al., 2017). GAF is similarly required for looping; depletion of the GAGA motif to which it binds results in loss of chromatin loops but does not seem to affect TAD formation (Gaskill et al., 2023; Li et al., 2023b; Ogiyama et al., 2018). The role of GAF in promoting 3D chromatin structure depends, in part, on the ability of GAF to interact with itself (Li et al., 2023b). The capacity of pioneer factors to multimerize, either through stable interactions or weak multivalent interactions, may promote the formation of sub-nuclear microenvironments that can shape 3D structure (Hayward-Lara et al., 2024). During somatic cell reprogramming of mouse embryonic fibroblasts, Oct4 facilitates TAD reorganization, and this depends on the ability of Oct4 to phase separate (Wang et al., 2021). In addition to their roles in forming TADs and loops, pioneer factors may influence whether regions of the genome are localized to the A or B compartments. For example, Pax7-mediated chromatin accessibility in mouse pituitary cells is correlated with a switch at pioneered loci from the inactive B compartment to the active A compartment through a process that requires cell division (Gouhier et al., 2024). Additional pioneer factors, such as Foxa2, AP-1 and KLF4, have been suggested to mediate changes to 3D chromatin structure (Di Giammartino et al., 2019; Hao et al., 2024; Phanstiel et al., 2017; Wolf et al., 2023).

Pioneer factors both directly, through binding nucleosomal DNA, and indirectly, through the recruitment of co-factors and chromatin-modifying enzymes, reshape the chromatin landscape at multiple levels. These modifications to the chromatin landscape result in robust changes to the gene-expression program and cellular state, driving normal developmental transitions and, when co-opted, leading to disease states. In the following section, we discuss how developmental state, in turn, feeds back on pioneer-factor activity.

Pioneer factors bind silenced regions of the genome, yet only a subset of their defined sequence-recognition motifs is bound in vivo, and these occupied binding sites vary between different cell types (Buecker et al., 2014; Cernilogar et al., 2019; Chronis et al., 2017; Donaghey et al., 2018; Gibson et al., 2024; Larson et al., 2021; Lupien et al., 2008; Mayran et al., 2018; Soufi et al., 2012). Both the binding and activity of pioneer factors can be influenced by developmental context (Gaskill et al., 2021; Gouhier et al., 2024; Maresca et al., 2023; Nevil et al., 2020; Schulz et al., 2015; Li et al., 2023a). The following examples highlight the role of cell fate in shaping pioneer factor function.

Chromatin structure regulates pioneer-factor binding and activity

Distinct cell types possess unique chromatin architecture that often reflects the gene-expression profile of the cell. Promoters are generally accessible, whereas enhancers have cell type-specific accessibility that is correlated with gene expression (Andersson and Sandelin, 2020; Heinz et al., 2015; Rada-Iglesias et al., 2011). As discussed above, active enhancers are marked with H3K27ac. Silent enhancers are marked by post-translational modifications that are associated with repressed chromatin, such as H3K27me3. Although pioneer factors influence both chromatin accessibility and chromatin modifications, these same chromatin features also regulate pioneer-factor binding and activity.

Many pioneer factors prefer to bind naïve chromatin that is largely devoid of histone marks and are excluded from silent, condensed heterochromatic regions (Gibson et al., 2024; Soufi et al., 2012). H3K9me3 is enriched at constitutive heterochromatin and limits binding of OSK in human fibroblasts (McCarthy et al., 2023; Soufi et al., 2012); OSK binding can be promoted by knocking down a histone methyltransferase that deposits H3K9me3, and this, in turn, increases reprogramming efficiency (Kim et al., 2021; Onder et al., 2012). H3K9me3 also impedes mouse embryonic reprogramming following nuclear transfer, suggesting that it is a barrier to the pioneer factor-mediated events of early embryonic development (Matoba et al., 2014). H3K9 methylation may be a general barrier preventing pioneer factors from accessing the genome. High levels of H3K9me3 limit binding of Pax7 during mouse pituitary specification (Mayran et al., 2018). Similarly, human FOXA1 binding primarily occurs in regions depleted of H3K9me2, and FOXA2 binding is enriched at regions depleted of H3K9me3 (Donaghey et al., 2018; Lupien et al., 2008). Together, these data suggest that repressive chromatin modifications restrict pioneer-factor occupancy.

Whereas constitutive heterochromatin is shared across cell types, facultative heterochromatin, characterized by H3K27me3, is cell type specific. Unlike H3K9 methylation, H3K27me3-marked facultative heterochromatin does not appear to limit binding of either Zelda or Grainy head in Drosophila cell lines. Removal of H3K27me3, through inhibition of the PRC2 enzyme that deposits it, does not result in increased occupancy of either pioneer factor at regions previously marked by H3K27me3 (Gibson et al., 2024). However, there are cell type-specific features that may influence pioneer binding. During asymmetric neural stem cell divisions in Drosophila, Zelda-mediated reprogramming is progressively limited (Larson et al., 2021). Knockdown of proteins that recruit histone deacetylases enhance Zelda-mediated reprogramming, suggesting that histone acetylation state regulates Zelda pioneering activity (Larson et al., 2021). These examples indicate that chromatin context can restrict pioneer-factor occupancy in a cell type-specific manner.

In addition to histone modifications, other features of chromatin may impact pioneer factor binding. H1 linker histones function to compact the genome and may, like H3K9me2/3, limit pioneer factor binding. In vitro, OCT4 binding to nucleosomes is impeded by the addition of H1, and in cell culture H1-dependent chromatin compaction restricts Pax7 recruitment (Echigoya et al., 2020; Gouhier et al., 2024). The nuclear lamina lines the inner nuclear membrane, provides structure to the nucleus and is associated with condensed heterochromatin (Shevelyov, 2023). Association of loci with the nuclear lamina may similarly limit pioneer-factor binding. During foregut development in C. elegans, PHA-4 binding to a reporter gene is increased when the nuclear lamina protein EMR-1 is knocked down (Fakhouri et al., 2010). Similarly, Foxa2/FOXA2 binding is enhanced in multiple mammalian models upon disruption of the nuclear lamina (Whitton et al., 2018; Wei et al., 2022).

The ability of chromatin context to promote pioneer-factor binding is less well established than the restrictive contexts discussed above. FOXA1 binding to chromatin is correlated with marks of active modifications (Lupien et al., 2008; Wang et al., 2015), but there is little data to support an instrumental role for these marks in directly shaping the binding. Conflicting structural studies about the role of H3K27ac in regulating OCT4 nucleosome binding also further limit our understanding of whether active marks promote chromatin occupancy of pioneer factors (Lian et al., 2023 preprint; Sinha et al., 2023). By contrast to the unclear role of histone modifications in promoting pioneer-factor binding, nucleosomes promote binding. Both in the early embryo and upon exogenous expression, pioneer factors in multiple species preferentially bind and open chromatin regions enriched for nucleosomes (Ballaré et al., 2013; Gibson et al., 2024; Miao et al., 2022; Veil et al., 2019). These data suggest that pioneer factors may preferentially bind motifs wrapped around histones or that interactions with the histone proteins themselves promote pioneer-factor binding. Together, it is evident that, although pioneer factors can access their binding sites within closed chromatin, the chromatin structure itself can influence the ability of these factors to stably bind and promote accessibility (Fig. 2A).

Fig. 2.

Cell type-specific features shape pioneer-factor activity. (A) Post-translational modifications to histones can either promote or inhibit pioneer-factor binding. In cell type 1, activating modifications (green) promote pioneer factor (PF, orange) binding to loci that are not occupied by the factor in the naïve, unmodified chromatin shown in cell type 2. In cell type 3, repressive modifications (pink) occlude pioneer-factor binding to regions that are bound in the unmodified chromatin of cell type 2. (B) As cells differentiate, they express a distinct set of protein co-factors (red, blue, green). In the red cell, the red co-factor directs pioneer-factor binding to a locus that is distinct from that bound by the pioneer factor when the blue co-factor is expressed. In cells in which no co-factor is expressed (beige), the pioneer-factor binds but is unable to promote chromatin accessibility. In the green cell type, the presence of the green co-factor allows for the pioneer factor to indirectly modify the surrounding chromatin. (C) Differential levels of pioneer-factor expression result in distinct regions of binding and chromatin accessibility. Whereas some loci require a high concentration of pioneer factor expression for chromatin accessibility (top), others can be opened and promote gene expression at low concentrations (bottom). (D) Different protein isoforms, generated by splicing, can result in distinct affinities for chromatin and influence pioneering activity. When exon 2 (dark blue) is present, the pioneer factor binds more stably to chromatin, leading to induced accessibility. When exon 2 is not included, the pioneer factor fails to stably bind or open chromatin. (E) Cell type-specific post-translational modifications to the pioneer factor can regulate binding and activity by either inhibiting (green) or promoting (purple) interactions with chromatin and other factors. Promoting interactions with other factors can lead to a diversity of effects on chromatin, in this case modification to the histone tails.

Fig. 2.

Cell type-specific features shape pioneer-factor activity. (A) Post-translational modifications to histones can either promote or inhibit pioneer-factor binding. In cell type 1, activating modifications (green) promote pioneer factor (PF, orange) binding to loci that are not occupied by the factor in the naïve, unmodified chromatin shown in cell type 2. In cell type 3, repressive modifications (pink) occlude pioneer-factor binding to regions that are bound in the unmodified chromatin of cell type 2. (B) As cells differentiate, they express a distinct set of protein co-factors (red, blue, green). In the red cell, the red co-factor directs pioneer-factor binding to a locus that is distinct from that bound by the pioneer factor when the blue co-factor is expressed. In cells in which no co-factor is expressed (beige), the pioneer-factor binds but is unable to promote chromatin accessibility. In the green cell type, the presence of the green co-factor allows for the pioneer factor to indirectly modify the surrounding chromatin. (C) Differential levels of pioneer-factor expression result in distinct regions of binding and chromatin accessibility. Whereas some loci require a high concentration of pioneer factor expression for chromatin accessibility (top), others can be opened and promote gene expression at low concentrations (bottom). (D) Different protein isoforms, generated by splicing, can result in distinct affinities for chromatin and influence pioneering activity. When exon 2 (dark blue) is present, the pioneer factor binds more stably to chromatin, leading to induced accessibility. When exon 2 is not included, the pioneer factor fails to stably bind or open chromatin. (E) Cell type-specific post-translational modifications to the pioneer factor can regulate binding and activity by either inhibiting (green) or promoting (purple) interactions with chromatin and other factors. Promoting interactions with other factors can lead to a diversity of effects on chromatin, in this case modification to the histone tails.

Close modal

Tissue-specific co-factor availability influences pioneering activity

Because co-factors influence pioneer-factor occupancy, developmentally regulated expression of these proteins impacts the ability of pioneer factors to stably bind and open the genome (Fig. 2B). Collaborative interactions between OSK and stage-specific transcription factors govern pluripotency-enhancer selection to reprogram somatic cells to pluripotency (Chronis et al., 2017). Consistent with this, Oct4 occupies different regions in the genome when it is expressed alone or with other reprogramming factors, suggesting that Oct4 cell type-specific binding is strongly influenced or redirected by additional co-factors (Chronis et al., 2017; Donaghey et al., 2018). This is similar to what is observed for the orthologs of these factors in zebrafish embryos (Miao et al., 2022). As stem cells transition between naïve and primed pluripotency, there is widespread relocalization of Oct4, which is regulated, in part, by expression of the transcription factor Otx2 in the primed pluripotent cells (Buecker et al., 2014). OCT4 and SOX2 cooperate in human neural crest specification but bind to regions distinct from those occupied by these factors in ESCs. This distinct binding is driven in part by binding with the neural crest transcription factor TFAP2 and highlights how the same pioneer factors can be co-opted by tissue-specific co-factors to drive distinct gene-expression programs (Hovland et al., 2022). Similarly, the capacity of SOX2 to promote the pluripotent fate versus the neuronal fate is regulated by the ability of neuronally expressed co-factors PAX6 and ATRX to promote localization of SOX2 to neuronal-specific regulatory regions (Bunina et al., 2020; Zhang et al., 2019). Thus, developmentally regulated expression of additional factors influences pioneer-factor occupancy. This contrasts with the relatively limited evidence that histone marks actively promote pioneer-factor binding. Indeed, upon exogenous expression of Zelda and Grh there is no clear chromatin profile that correlates with cell type-specific binding events, but motif analysis shows an enrichment for motifs bound by factors that are expressed in a tissue-specific manner. This suggests that expression of these transcription factors may modulate the binding of Zelda and Grh in a tissue-specific manner (Gibson et al., 2024). Similar data for FOXA2 suggest that coordinated involvement of multiple factors may be a general feature to help specify pioneer-factor binding (Meers et al., 2019).

Co-factors can also affect pioneer-factor activity. Despite the ability of pioneer factors to bind nucleosomal DNA and promote local opening, most pioneer factors only promote chromatin accessibility at a subset of bound regions (Gibson et al., 2024; Soufi et al., 2012; Veil et al., 2019). Thus, in many cases binding is separable from pioneer activity, and developmental state can regulate the capacity of pioneer factors to promote chromatin accessibility. For example, the co-factor GATA4 is co-expressed with FOXA2 in human endoderm and co-expression of this co-factor in a human fibroblast cell line promotes FOXA2 occupancy at endoderm-specific binding sites. However, this binding does not lead to accessibility, suggesting that additional factors must be recruited to promote accessibility (Donaghey et al., 2018). In Drosophila, Grh binding remains stable across developmental stages, but the gene expression pattern and chromatin accessibility driven by Grh varies through development (Nevil et al., 2017, 2020). This suggests that pioneering activity of Grh, rather than its binding, is regulated by tissue-specific features, such as co-factors. In pituitary stem cells, Pax7 can bind chromatin but requires the co-factor Tpit to promote accessibility (Mayran et al., 2018, 2019). Thus, during certain developmental stages, chromatin accessibility may be established through cooperation of pioneer factors with other non-pioneer factors.

Protein-intrinsic features control pioneer-factor binding and function

In addition to extrinsic features, pioneer-factor activity is also regulated by protein-intrinsic properties of the pioneer factors themselves, including expression level, isoform expression and post-translational modifications (Fig. 2C-E). These properties are, in turn, regulated by developmental context.

Concentration is a key factor influencing pioneering activity (Fig. 2C), and when overexpressed pioneer factors can be oncogenic. Upon ectopic induction, chromatin binding and accessibility is correlated with the expression levels of Grh and Zelda (Gibson et al., 2024). Zelda overexpression in neural stem cells similarly results in novel occupancy of closed regions, suggesting that higher expression levels promote binding to closed chromatin (Gibson et al., 2024). Levels of Nanog, Pou5f3 and Sox19b expression have a concentration-dependent effect, with protein levels correlating with chromatin accessibility (Miao et al., 2022). In models of Pax7-induced melanotrope production, chromatin opening is similarly correlated with levels of Pax7 binding (Gouhier et al., 2024). Although overall expression levels clearly influence pioneer activity, the local concentration of pioneer factors may also be regulated at specific loci. Imaging of numerous pioneer factors, including FOXA1, KLF4, SOX2, GAF, Nanog and Zelda, demonstrate that they are not uniformly distributed in the nucleus but instead are concentrated in localized microenvironments (Dufourt et al., 2018; Gaskill et al., 2021, 2023; Hayward-Lara et al., 2024; Ji et al., 2024; Kuznetsova et al., 2023; Mir et al., 2018; Morin et al., 2022; Nguyen et al., 2022; Raff et al., 1994; Sharma et al., 2021). These pioneer factor microenvironments are formed by various mechanisms, including the multivalent interactions mediated by intrinsically disordered regions that are found in many transcription factors (Boeynaems et al., 2018; Ferrie et al., 2022; Peng et al., 2020; Rippe, 2022). These localized regions of relatively concentrated factors may promote chromatin opening and gene expression (Boija et al., 2018; Chong et al., 2018; Tang et al., 2022). Nonetheless, the relationship between concentration and ability to activate gene expression is not simple. Although hub formation can promote transcription, when concentrated above a certain threshold, proteins may promote transcriptional repression (Chong et al., 2022; Gaskill et al., 2023). Many mechanistic studies of pioneer factors rely on overexpression in culture. These conditions are similar to those applied for iPSC induction, which requires the exogenous expression of multiple pioneer factors to initiate the necessary reprogramming events. It is increasingly clear that the levels of expression can influence pioneer factor activity, and thus it is essential for studies to be performed in the context of normal development.

Because domains outside of the DBD influence pioneer-factor binding and activity, isoform expression can regulate function (Fig. 2D). Protein structure can vary with developmental context owing to splicing or alternative transcription start sites. For example, the Drosophila pioneer factor Grh has multiple isoforms that are regulated by both processes. There are two characterized splice isoforms that both contain the conserved DBD but differ in the N-terminal region and are expressed in different Drosophila tissues (Uv et al., 1997; Almeida and Bray, 2005; Bray and Kafatos, 1991). In intestinal stem cells, one isoform is thought to prevent differentiation whereas ectopic expression of the other promotes differentiation (Dominado et al., 2022 preprint). Similarly, the mammalian GRHL genes are differentially spliced, suggesting a conserved regulatory mechanism (Lambourne et al., 2024 preprint; Wilanowski et al., 2002). In addition to splicing, there are multiple transcription start sites that result in proteins with N-terminal domains of varying lengths, and these isoforms have tissue-specific expression patterns. Given that regions outside of the DBD strongly influence pioneer-factor activity, it is evident that these developmentally regulated isoforms could significantly impact pioneering function.

Protein activity can be influenced by post-translational modifications, and these modifications can differ depending on cellular context. Many pioneer factors are post-translationally modified (Brumbaugh et al., 2012; Cho et al., 2018; Park et al., 2021; Spelat et al., 2012; Sutinen et al., 2014; Wei et al., 2007; Williams et al., 2020; Wolfrum et al., 2003). Cellular signaling often results in downstream changes in post-translational modifications. For example, phosphorylation regulates conserved functions of Grh in a context-dependent manner. Grh is essential in both epidermal barrier formation and wound healing (Sundararajan et al., 2020; Wang and Samakovlis, 2012). Although phosphorylation is dispensable for maintenance of the epidermal barrier, it is required for the activation of wound enhancers upon wounding (Kim and McGinnis, 2011).

Post-translational modifications can influence pioneer activity by regulating total and local concentrations. For example, Sox2 levels are controlled through competing modifications; methylation of Sox2 resulting in ubiquitylation and subsequent degradation, whereas phosphorylation prevents methylation and therefore stabilizes the protein (Fang et al., 2014). Acetylation of Sox2 affects local concentration by disrupting the nuclear localization signal and promoting nuclear export (Baltus et al., 2009; Williams et al., 2020). Nuclear import of SOX11 is controlled by phosphorylation (Balta et al., 2018; Williams et al., 2020). Thus, the local and total concentrations of pioneer factors are regulated by post-translational modifications, which are subject to developmental regulation.

Post-translational modifications also regulate the interaction of pioneer factors with both the DNA and co-factors (Fig. 2E). The affinity of GAF, Pax7 and HMG-14 (HMGN1) for chromatin are reduced by acetylation and phosphorylation, which can influence interaction with both the negatively charged DNA as well as the nucleosomes (Aran-Guiu et al., 2010; Bergel et al., 2000; Bonet et al., 2005; Sincennes et al., 2021). During asymmetric division of satellite stem cells in mice, Pax7 is differentially regulated in the two resulting daughter cell lineages by methylation on the N terminus. This methylation allows Pax7 to interact with the MLL complex to deposit H3K4me3 and induce expression of a gene marking the committed satellite fate (Kawabe et al., 2012). Protein activity is broadly regulated through development by a myriad of different mechanisms, and pioneer factors are no exception.

The above examples highlight the reciprocal relationship between pioneer factors and development. Ultimately, understanding how this special class of transcription factors engages with the genome to drive cell-state transitions will have fundamental implications for our understanding of both normal development and disease progression.

The capacity of pioneer factors to promote chromatin accessibility enables them to drive major developmental transitions. However, this same function promotes disease states when dysregulated. Thus, it is imperative to understand the mechanisms governing pioneer-factor activity. Although extensive work has been done to characterize pioneer-factor function, many of these studies rely on in vitro and cell-culture systems. These systems have been essential for understanding the mechanisms by which pioneer factors bind and open chromatin, but provide limited insight into the relationship between pioneer factors and development. The emerging evidence that pioneering function is regulated by developmental context makes it clear that in vivo studies provide a necessary complement to studies in culture and in a test tube. To understand how these powerful factors drive development and disease, it will be essential to uncover the reciprocal relationship between pioneer factor-driven developmental transitions and how cellular state influences pioneering activity.

We thank Audrey Marsh, Annemarie Branks and members of the Harrison lab for helpful discussions. We apologize to those whose research we did not discuss or cite owing to space limitations.

Funding

Research in the Harrison lab is supported by the National Institutes of Health (NIH) (R35 GM136298 and R01 NS111647). M.M.F. was supported by an NIH National Research Service Award (T32 GM00713). M.M.H. is a Romnes Faculty Fellow and Vilas Faculty Mid-Career Investigator at the University of Wisconsin-Madison. E.F.T.-Z. is supported by a fellowship from the University of Wisconsin-Madison Stem Cell and Regenerative Medicine Center. Deposited in PMC for release after 12 months.

Abdulhay
,
N. J.
,
McNally
,
C. P.
,
Hsieh
,
L. J.
,
Kasinathan
,
S.
,
Keith
,
A.
,
Estes
,
L. S.
,
Karimzadeh
,
M.
,
Underwood
,
J. G.
,
Goodarzi
,
H.
,
Narlikar
,
G. J.
et al.
(
2020
).
Massively multiplex single-molecule oligonucleosome footprinting
.
eLife
9
,
e59404
.
Ahmad
,
K.
,
Brahma
,
S.
and
Henikoff
,
S.
(
2024
).
Epigenetic pioneering by SWI/SNF family remodelers
.
Mol. Cell
84
,
194
-
201
.
Almeida
,
M. S.
and
Bray
,
S. J.
(
2005
).
Regulation of post-embryonic neuroblasts by Drosophila Grainyhead
.
Mech. Dev.
122
,
1282
-
1293
.
Andersson
,
R.
and
Sandelin
,
A.
(
2020
).
Determinants of enhancer and promoter activities of regulatory elements
.
Nat. Rev. Genet.
21
,
71
-
87
.
Antonin
,
W.
and
Neumann
,
H.
(
2016
).
Chromosome condensation and decondensation during mitosis
.
Curr. Opin. Cell Biol.
40
,
15
-
22
.
Aran-Guiu
,
X.
,
Ortiz-Lombardía
,
M.
,
Oliveira
,
E.
,
Bonet Costa
,
C.
,
Odena
,
M. A.
,
Bellido
,
D.
and
Bernués
,
J.
(
2010
).
Acetylation of GAGA factor modulates its interaction with DNA
.
Biochemistry
49
,
9140
-
9151
.
Ballaré
,
C.
,
Castellano
,
G.
,
Gaveglia
,
L.
,
Althammer
,
S.
,
González-Vallinas
,
J.
,
Eyras
,
E.
,
Le Dily
,
F.
,
Zaurin
,
R.
,
Soronellas
,
D.
,
Vicent
,
G. P.
et al.
(
2013
).
Nucleosome-driven transcription factor binding and gene regulation
.
Mol. Cell
49
,
67
-
79
.
Balsalobre
,
A.
and
Drouin
,
J.
(
2022
).
Pioneer factors as master regulators of the epigenome and cell fate
.
Nat. Rev. Mol. Cell Biol.
23
,
449
-
464
.
Balta
,
E.-A.
,
Wittmann
,
M.-T.
,
Jung
,
M.
,
Sock
,
E.
,
Haeberle
,
B. M.
,
Heim
,
B.
,
Von Zweydorf
,
F.
,
Heppt
,
J.
,
Von Wittgenstein
,
J.
,
Gloeckner
,
C. J.
et al.
(
2018
).
Phosphorylation modulates the subcellular localization of SOX11
.
Front. Mol. Neurosci.
11
,
211
.
Baltus
,
G. A.
,
Kowalski
,
M. P.
,
Zhai
,
H.
,
Tutter
,
A. V.
,
Quinn
,
D.
,
Wall
,
D.
and
Kadam
,
S.
(
2009
).
Acetylation of Sox2 induces its nuclear export in embryonic stem cells
.
Stem Cells
27
,
2175
-
2184
.
Barral
,
A.
and
Zaret
,
K. S.
(
2024
).
Pioneer factors: roles and their regulation in development
.
Trends Genet.
40
,
134
-
148
.
Batut
,
P. J.
,
Bing
,
X. Y.
,
Sisco
,
Z.
,
Raimundo
,
J.
,
Levo
,
M.
and
Levine
,
M. S.
(
2022
).
Genome organization controls transcriptional dynamics during development
.
Science
375
,
566
-
570
.
Bellec
,
M.
,
Dufourt
,
J.
,
Hunt
,
G.
,
Lenden-Hasse
,
H.
,
Trullo
,
A.
,
Zine El Aabidine
,
A.
,
Lamarque
,
M.
,
Gaskill
,
M. M.
,
Faure-Gautron
,
H.
,
Mannervik
,
M.
et al.
(
2022
).
The control of transcriptional memory by stable mitotic bookmarking
.
Nat. Commun.
13
,
1176
.
Bergel
,
M.
,
Herrera
,
J. E.
,
Thatcher
,
B. J.
,
Prymakowska-Bosak
,
M.
,
Vassilev
,
A.
,
Nakatani
,
Y.
,
Martin
,
B.
and
Bustin
,
M.
(
2000
).
Acetylation of novel sites in the nucleosomal binding domain of chromosomal protein HMG-14 by p300 alters its interaction with nucleosomes
.
J. Biol. Chem.
275
,
11514
-
11520
.
Bertrand
,
N.
,
Castro
,
D. S.
and
Guillemot
,
F.
(
2002
).
Proneural genes and the specification of neural cell types
.
Nat. Rev. Neurosci.
3
,
517
-
530
.
Bi
,
M.
,
Zhang
,
Z.
,
Jiang
,
Y.-Z.
,
Xue
,
P.
,
Wang
,
H.
,
Lai
,
Z.
,
Fu
,
X.
,
De Angelis
,
C.
,
Gong
,
Y.
,
Gao
,
Z.
et al.
(
2020
).
Enhancer reprogramming driven by high-order assemblies of transcription factors promotes phenotypic plasticity and breast cancer endocrine resistance
.
Nat. Cell Biol.
22
,
701
-
715
.
Biggin
,
M. D.
and
Tjian
,
R.
(
1988
).
Transcription factors that activate the Ultrabithorax promoter in developmentally staged extracts
.
Cell
53
,
699
-
711
.
Blitz
,
I. L.
and
Cho
,
K. W. Y.
(
2021
).
Control of zygotic genome activation in Xenopus
. In:
Current Topics in Developmental Biology
(ed.
S. Y.
Sokol
), pp.
167
-
204
.
Elsevier
.
Boeynaems
,
S.
,
Alberti
,
S.
,
Fawzi
,
N. L.
,
Mittag
,
T.
,
Polymenidou
,
M.
,
Rousseau
,
F.
,
Schymkowitz
,
J.
,
Shorter
,
J.
,
Wolozin
,
B.
,
Van Den Bosch
,
L.
et al.
(
2018
).
Protein phase separation: a new phase in cell biology
.
Trends Cell Biol.
28
,
420
-
435
.
Boija
,
A.
,
Klein
,
I. A.
,
Sabari
,
B. R.
,
Dall'Agnese
,
A.
,
Coffey
,
E. L.
,
Zamudio
,
A. V.
,
Li
,
C. H.
,
Shrinivas
,
K.
,
Manteiga
,
J. C.
,
Hannett
,
N. M.
et al.
(
2018
).
Transcription factors activate genes through the phase-separation capacity of their activation domains
.
Cell
175
,
1842
-
1855.e16
.
Bonet
,
C.
,
Fernández
,
I.
,
Aran
,
X.
,
Bernués
,
J.
,
Giralt
,
E.
and
Azorín
,
F.
(
2005
).
The GAGA protein of Drosophila is phosphorylated by CK2
.
J. Mol. Biol.
351
,
562
-
572
.
Bray
,
S. J.
and
Kafatos
,
F. C.
(
1991
).
Developmental function of Elf-1: an essential transcription factor during embryogenesis in Drosophila
.
Genes Dev.
5
,
1672
-
1683
.
Bray
,
S. J.
,
Burke
,
B.
,
Brown
,
N. H.
and
Hirsh
,
J.
(
1989
).
Embryonic expression pattern of a family of Drosophila proteins that interact with a central nervous system regulatory element
.
Genes Dev.
3
,
1130
-
1145
.
Brennan
,
K. J.
,
Weilert
,
M.
,
Krueger
,
S.
,
Pampari
,
A.
,
Liu
,
H.-Y.
,
Yang
,
A. W. H.
,
Morrison
,
J. A.
,
Hughes
,
T. R.
,
Rushlow
,
C. A.
,
Kundaje
,
A.
et al.
(
2023
).
Chromatin accessibility in the Drosophila embryo is determined by transcription factor pioneering and enhancer activation
.
Dev. Cell
58
,
1898
-
1916.e9
.
Brumbaugh
,
J.
,
Hou
,
Z.
,
Russell
,
J. D.
,
Howden
,
S. E.
,
Yu
,
P.
,
Ledvina
,
A. R.
,
Coon
,
J. J.
and
Thomson
,
J. A.
(
2012
).
Phosphorylation regulates human OCT4
.
Proc. Natl. Acad. Sci. USA
109
,
7162
-
7168
.
Budry
,
L.
,
Balsalobre
,
A.
,
Gauthier
,
Y.
,
Khetchoumian
,
K.
,
L'Honoré
,
A.
,
Vallette
,
S.
,
Brue
,
T.
,
Figarella-Branger
,
D.
,
Meij
,
B.
and
Drouin
,
J.
(
2012
).
The selector gene Pax7 dictates alternate pituitary cell fates through its pioneer action on chromatin remodeling
.
Genes Dev.
26
,
2299
-
2310
.
Buecker
,
C.
,
Srinivasan
,
R.
,
Wu
,
Z.
,
Calo
,
E.
,
Acampora
,
D.
,
Faial
,
T.
,
Simeone
,
A.
,
Tan
,
M.
,
Swigut
,
T.
and
Wysocka
,
J.
(
2014
).
Reorganization of enhancer patterns in transition from naive to primed pluripotency
.
Cell Stem Cell
14
,
838
-
853
.
Bunina
,
D.
,
Abazova
,
N.
,
Diaz
,
N.
,
Noh
,
K.-M.
,
Krijgsveld
,
J.
and
Zaugg
,
J. B.
(
2020
).
Genomic rewiring of SOX2 chromatin interaction network during differentiation of ESCs to postmitotic neurons
.
Cell Syst.
10
,
480
-
494.e8
.
Calo
,
E.
and
Wysocka
,
J.
(
2013
).
Modification of enhancer chromatin: what, how, and why?
Mol. Cell
49
,
825
-
837
.
Caravaca
,
J. M.
,
Donahue
,
G.
,
Becker
,
J. S.
,
He
,
X.
,
Vinson
,
C.
and
Zaret
,
K. S.
(
2013
).
Bookmarking by specific and nonspecific binding of FoxA1 pioneer factor to mitotic chromosomes
.
Genes Dev.
27
,
251
-
260
.
Carrió
,
E.
,
Magli
,
A.
,
Muñoz
,
M.
,
Peinado
,
M. A.
,
Perlingeiro
,
R.
and
Suelves
,
M.
(
2016
).
Muscle cell identity requires Pax7-mediated lineage-specific DNA demethylation
.
BMC Biol.
14
,
30
.
Carroll
,
J. S.
,
Liu
,
X. S.
,
Brodsky
,
A. S.
,
Li
,
W.
,
Meyer
,
C. A.
,
Szary
,
A. J.
,
Eeckhoute
,
J.
,
Shao
,
W.
,
Hestermann
,
E. V.
,
Geistlinger
,
T. R.
et al.
(
2005
).
Chromosome-wide mapping of estrogen receptor binding reveals long-range regulation requiring the forkhead protein FoxA1
.
Cell
122
,
33
-
43
.
Cernilogar
,
F. M.
,
Hasenöder
,
S.
,
Wang
,
Z.
,
Scheibner
,
K.
,
Burtscher
,
I.
,
Sterr
,
M.
,
Smialowski
,
P.
,
Groh
,
S.
,
Evenroed
,
I. M.
,
Gilfillan
,
G. D.
et al.
(
2019
).
Pre-marked chromatin and transcription factor co-binding shape the pioneering activity of Foxa2
.
Nucleic Acids Res.
47
,
9069
-
9086
.
Chan
,
S. H.
,
Tang
,
Y.
,
Miao
,
L.
,
Darwich-Codore
,
H.
,
Vejnar
,
C. E.
,
Beaudoin
,
J.-D.
,
Musaev
,
D.
,
Fernandez
,
J. P.
,
Benitez
,
M. D. J.
,
Bazzini
,
A. A.
et al.
(
2019
).
Brd4 and P300 confer transcriptional competency during zygotic genome activation
.
Dev. Cell
49
,
867
-
881.e8
.
Charney
,
R. M.
,
Forouzmand
,
E.
,
Cho
,
J. S.
,
Cheung
,
J.
,
Paraiso
,
K. D.
,
Yasuoka
,
Y.
,
Takahashi
,
S.
,
Taira
,
M.
,
Blitz
,
I. L.
,
Xie
,
X.
et al.
(
2017
).
Foxh1 occupies cis-regulatory modules prior to dynamic transcription factor interactions controlling the mesendoderm gene program
.
Dev. Cell
40
,
595
-
607.e4
.
Chen
,
Z.
and
Zhang
,
Y.
(
2019
).
Loss of DUX causes minor defects in zygotic genome activation and is compatible with mouse development
.
Nat. Genet.
51
,
947
-
951
.
Chen
,
A. F.
,
Liu
,
A. J.
,
Krishnakumar
,
R.
,
Freimer
,
J. W.
,
DeVeale
,
B.
and
Blelloch
,
R.
(
2018
).
GRHL2-dependent enhancer switching maintains a pluripotent stem cell transcriptional subnetwork after exit from naive pluripotency
.
Cell Stem Cell
23
,
226
-
238.e4
.
Chervova
,
A.
,
Molliex
,
A.
,
Baymaz
,
H. I.
,
Papadopoulou
,
T.
,
Mueller
,
F.
,
Hercul
,
E.
,
Fournier
,
D.
,
Dubois
,
A.
,
Gaiani
,
N.
,
Beli
,
P.
et al.
(
2024
).
Mitotic bookmarking redundancy by nuclear receptors in pluripotent cells
.
Nat. Struct. Mol. Biol.
31
,
513
-
522
.
Chiou
,
S.-H.
,
Yu
,
C.-C.
,
Huang
,
C.-Y.
,
Lin
,
S.-C.
,
Liu
,
C.-J.
,
Tsai
,
T.-H.
,
Chou
,
S.-H.
,
Chien
,
C.-S.
,
Ku
,
H.-H.
and
Lo
,
J.-F.
(
2008
).
Positive correlations of Oct-4 and nanog in oral cancer stem-like cells and high-grade oral squamous cell carcinoma
.
Clin. Cancer Res.
14
,
4085
-
4095
.
Chiou
,
S.-H.
,
Wang
,
M.-L.
,
Chou
,
Y.-T.
,
Chen
,
C.-J.
,
Hong
,
C.-F.
,
Hsieh
,
W.-J.
,
Chang
,
H.-T.
,
Chen
,
Y.-S.
,
Lin
,
T.-W.
,
Hsu
,
H.-S.
et al.
(
2010
).
Coexpression of Oct4 and Nanog enhances malignancy in lung adenocarcinoma by inducing cancer stem cell–like properties and epithelial–mesenchymal transdifferentiation
.
Cancer Res.
70
,
10433
-
10444
.
Cho
,
Y.
,
Kang
,
H. G.
,
Kim
,
S.-J.
,
Lee
,
S.
,
Jee
,
S.
,
Ahn
,
S. G.
,
Kang
,
M. J.
,
Song
,
J. S.
,
Chung
,
J.-Y.
,
Yi
,
E. C.
et al.
(
2018
).
Post-translational modification of OCT4 in breast cancer tumorigenesis
.
Cell Death Differ.
25
,
1781
-
1795
.
Choi
,
S. H.
,
Gearhart
,
M. D.
,
Cui
,
Z.
,
Bosnakovski
,
D.
,
Kim
,
M.
,
Schennum
,
N.
and
Kyba
,
M.
(
2016
).
DUX4 recruits p300/CBP through its C-terminus and induces global H3K27 acetylation changes
.
Nucleic Acids Res.
44
,
5161
-
5173
.
Chong
,
S.
,
Dugast-Darzacq
,
C.
,
Liu
,
Z.
,
Dong
,
P.
,
Dailey
,
G. M.
,
Cattoglio
,
C.
,
Heckert
,
A.
,
Banala
,
S.
,
Lavis
,
L.
,
Darzacq
,
X.
et al.
(
2018
).
Imaging dynamic and selective low-complexity domain interactions that control gene transcription
.
Science
361
,
eaar2555
.
Chong
,
S.
,
Graham
,
T. G. W.
,
Dugast-Darzacq
,
C.
,
Dailey
,
G. M.
,
Darzacq
,
X.
and
Tjian
,
R.
(
2022
).
Tuning levels of low-complexity domain interactions to modulate endogenous oncogenic transcription
.
Mol. Cell
82
,
2084
-
2097.e5
.
Chronis
,
C.
,
Fiziev
,
P.
,
Papp
,
B.
,
Butz
,
S.
,
Bonora
,
G.
,
Sabri
,
S.
,
Ernst
,
J.
and
Plath
,
K.
(
2017
).
Cooperative binding of transcription factors orchestrates reprogramming
.
Cell
168
,
442
-
459.e20
.
Cieply
,
B.
,
Riley
,
P.
,
Pifer
,
P. M.
,
Widmeyer
,
J.
,
Addison
,
J. B.
,
Ivanov
,
A. V.
,
Denvir
,
J.
and
Frisch
,
S. M.
(
2012
).
Suppression of the epithelial-mesenchymal transition by Grainyhead-like-2
.
Cancer Res.
72
,
2440
-
2453
.
Cieply
,
B.
,
Farris
,
J.
,
Denvir
,
J.
,
Ford
,
H. L.
and
Frisch
,
S. M.
(
2013
).
Epithelial-mesenchymal transition and tumor suppression are controlled by a reciprocal feedback loop between ZEB1 and Grainyhead-like-2
.
Cancer Res.
73
,
6299
-
6309
.
Cirillo
,
L. A.
(
1998
).
Binding of the winged-helix transcription factor HNF3 to a linker histone site on the nucleosome
.
EMBO J.
17
,
244
-
254
.
Cirillo
,
L. A.
,
Lin
,
F. R.
,
Cuesta
,
I.
,
Friedman
,
D.
,
Jarnik
,
M.
and
Zaret
,
K. S.
(
2002
).
Opening of compacted chromatin by early developmental transcription factors HNF3 (FoxA) and GATA-4
.
Mol. Cell
9
,
279
-
289
.
Clapier
,
C. R.
,
Iwasa
,
J.
,
Cairns
,
B. R.
and
Peterson
,
C. L.
(
2017
).
Mechanisms of action and regulation of ATP-dependent chromatin-remodelling complexes
.
Nat. Rev. Mol. Cell Biol.
18
,
407
-
422
.
Cocce
,
K. J.
,
Jasper
,
J. S.
,
Desautels
,
T. K.
,
Everett
,
L.
,
Wardell
,
S.
,
Westerling
,
T.
,
Baldi
,
R.
,
Wright
,
T. M.
,
Tavares
,
K.
,
Yllanes
,
A.
et al.
(
2019
).
The lineage determining factor GRHL2 collaborates with FOXA1 to establish a targetable pathway in endocrine therapy-resistant breast cancer
.
Cell Rep.
29
,
889
-
903.e10
.
Dai
,
S.
,
Qu
,
L.
,
Li
,
J.
and
Chen
,
Y.
(
2021
).
Toward a mechanistic understanding of DNA binding by forkhead transcription factors and its perturbation by pathogenic mutations
.
Nucleic Acids Res.
49
,
10235
-
10249
.
De Iaco
,
A.
,
Planet
,
E.
,
Coluccio
,
A.
,
Verp
,
S.
,
Duc
,
J.
and
Trono
,
D.
(
2017
).
DUX-family transcription factors regulate zygotic genome activation in placental mammals
.
Nat. Genet.
49
,
941
-
945
.
Deluz
,
C.
,
Friman
,
E. T.
,
Strebinger
,
D.
,
Benke
,
A.
,
Raccaud
,
M.
,
Callegari
,
A.
,
Leleu
,
M.
,
Manley
,
S.
and
Suter
,
D. M.
(
2016
).
A role for mitotic bookmarking of SOX2 in pluripotency and differentiation
.
Genes Dev.
30
,
2538
-
2550
.
Di Giammartino
,
D. C.
,
Kloetgen
,
A.
,
Polyzos
,
A.
,
Liu
,
Y.
,
Kim
,
D.
,
Murphy
,
D.
,
Abuhashem
,
A.
,
Cavaliere
,
P.
,
Aronson
,
B.
,
Shah
,
V.
et al.
(
2019
).
KLF4 is involved in the organization and regulation of pluripotency-associated three-dimensional enhancer networks
.
Nat. Cell Biol.
21
,
1179
-
1190
.
Dodonova
,
S. O.
,
Zhu
,
F.
,
Dienemann
,
C.
,
Taipale
,
J.
and
Cramer
,
P.
(
2020
).
Nucleosome-bound SOX2 and SOX11 structures elucidate pioneer factor function
.
Nature
580
,
669
-
672
.
Dominado
,
N.
,
Casagranda
,
F.
,
Heaney
,
J.
,
Siddall
,
N. A.
,
Abud
,
H. E.
and
Hime
,
G. R.
(
2022
).
Alternate Grainy head isoforms regulate Drosophila midgut intestinal stem cell differentiation
.
bioRxiv
Donaghey
,
J.
,
Thakurela
,
S.
,
Charlton
,
J.
,
Chen
,
J. S.
,
Smith
,
Z. D.
,
Gu
,
H.
,
Pop
,
R.
,
Clement
,
K.
,
Stamenova
,
E. K.
,
Karnik
,
R.
et al.
(
2018
).
Genetic determinants and epigenetic effects of pioneer-factor occupancy
.
Nat. Genet.
50
,
250
-
258
.
Dong
,
X.
and
Weng
,
Z.
(
2013
).
The correlation between histone modifications and gene expression
.
Epigenomics
5
,
113
-
116
.
Donovan
,
B. T.
,
Chen
,
H.
,
Jipa
,
C.
,
Bai
,
L.
and
Poirier
,
M. G.
(
2019
).
Dissociation rate compensation mechanism for budding yeast pioneer transcription factors
.
eLife
8
,
e43008
.
Donovan
,
B. T.
,
Chen
,
H.
,
Eek
,
P.
,
Meng
,
Z.
,
Jipa
,
C.
,
Tan
,
S.
,
Bai
,
L.
and
Poirier
,
M. G.
(
2023
).
Basic helix-loop-helix pioneer factors interact with the histone octamer to invade nucleosomes and generate nucleosome-depleted regions
.
Mol. Cell
83
,
1251
-
1263.e6
.
Duan
,
J.
,
Rieder
,
L.
,
Colonnetta
,
M. M.
,
Huang
,
A.
,
Mckenney
,
M.
,
Watters
,
S.
,
Deshpande
,
G.
,
Jordan
,
W.
,
Fawzi
,
N.
and
Larschan
,
E.
(
2021
).
CLAMP and Zelda function together to promote Drosophila zygotic genome activation
.
eLife
10
,
e69937
.
Dufourt
,
J.
,
Trullo
,
A.
,
Hunter
,
J.
,
Fernandez
,
C.
,
Lazaro
,
J.
,
Dejean
,
M.
,
Morales
,
L.
,
Nait-Amer
,
S.
,
Schulz
,
K. N.
,
Harrison
,
M. M.
et al.
(
2018
).
Temporal control of gene expression by the pioneer factor Zelda through transient interactions in hubs
.
Nat. Commun.
9
,
5194
.
Dynlacht
,
B. D.
,
Attardi
,
L. D.
,
Admon
,
A.
,
Freeman
,
M.
and
Tjian
,
R.
(
1989
).
Functional analysis of NTF-1, a developmentally regulated Drosophila transcription factor that binds neuronal cis elements
.
Genes Dev.
3
,
1677
-
1688
.
Echigoya
,
K.
,
Koyama
,
M.
,
Negishi
,
L.
,
Takizawa
,
Y.
,
Mizukami
,
Y.
,
Shimabayashi
,
H.
,
Kuroda
,
A.
and
Kurumizaka
,
H.
(
2020
).
Nucleosome binding by the pioneer transcription factor OCT4
.
Sci. Rep.
10
,
11832
.
Espinola
,
S. M.
,
Götz
,
M.
,
Bellec
,
M.
,
Messina
,
O.
,
Fiche
,
J.-B.
,
Houbron
,
C.
,
Dejean
,
M.
,
Reim
,
I.
,
Cardozo Gizzi
,
A. M.
,
Lagha
,
M.
et al.
(
2021
).
Cis-regulatory chromatin loops arise before TADs and gene activation, and are independent of cell fate during early Drosophila development
.
Nat. Genet.
53
,
477
-
486
.
Fakhouri
,
T. H. I.
,
Stevenson
,
J.
,
Chisholm
,
A. D.
and
Mango
,
S. E.
(
2010
).
Dynamic chromatin organization during foregut development mediated by the organ selector gene PHA-4/FoxA
.
PLoS Genet.
6
,
e1001060
.
Fang
,
F.
,
Xu
,
Y.
,
Chew
,
K.-K.
,
Chen
,
X.
,
Ng
,
H.-H.
and
Matsudaira
,
P.
(
2014
).
Coactivators p300 and CBP maintain the identity of mouse embryonic stem cells by mediating long-range chromatin structure
.
Stem Cells
32
,
1805
-
1816
.
Fernandez Garcia
,
M.
,
Moore
,
C. D.
,
Schulz
,
K. N.
,
Alberto
,
O.
,
Donague
,
G.
,
Harrison
,
M. M.
,
Zhu
,
H.
and
Zaret
,
K. S.
(
2019
).
Structural features of transcription factors associating with nucleosome binding
.
Mol. Cell
75
,
921
-
932.e6
.
Ferrie
,
J. J.
,
Karr
,
J. P.
,
Tjian
,
R.
and
Darzacq
,
X.
(
2022
).
“Structure”-function relationships in eukaryotic transcription factors: the role of intrinsically disordered regions in gene regulation
.
Mol. Cell
82
,
3970
-
3984
.
Festuccia
,
N.
,
Dubois
,
A.
,
Vandormael-Pournin
,
S.
,
Gallego Tejeda
,
E.
,
Mouren
,
A.
,
Bessonnard
,
S.
,
Mueller
,
F.
,
Proux
,
C.
,
Cohen-Tannoudji
,
M.
and
Navarro
,
P.
(
2016
).
Mitotic binding of Esrrb marks key regulatory regions of the pluripotency network
.
Nat. Cell Biol.
18
,
1139
-
1148
.
Festuccia
,
N.
,
Owens
,
N.
,
Papadopoulou
,
T.
,
Gonzalez
,
I.
,
Tachtsidi
,
A.
,
Vandoermel-Pournin
,
S.
,
Gallego
,
E.
,
Gutierrez
,
N.
,
Dubois
,
A.
,
Cohen-Tannoudji
,
M.
et al.
(
2019
).
Transcription factor activity and nucleosome organization in mitosis
.
Genome Res.
29
,
250
-
260
.
Frederick
,
M. A.
,
Williamson
,
K. E.
,
Fernandez Garcia
,
M.
,
Ferretti
,
M. B.
,
McCarthy
,
R. L.
,
Donahue
,
G.
,
Luzete Monteiro
,
E.
,
Takenaka
,
N.
,
Reynaga
,
J.
,
Kadoch
,
C.
et al.
(
2023
).
A pioneer factor locally opens compacted chromatin to enable targeted ATP-dependent nucleosome remodeling
.
Nat. Struct. Mol. Biol.
30
,
31
-
37
.
Fuda
,
N. J.
,
Guertin
,
M. J.
,
Sharma
,
S.
,
Danko
,
C. G.
,
Martins
,
A. L.
,
Siepel
,
A.
and
Lis
,
J. T.
(
2015
).
GAGA factor maintains nucleosome-free regions and has a role in RNA polymerase II recruitment to promoters
.
PLoS Genet.
11
,
e1005108
.
Fyodorov
,
D. V.
,
Zhou
,
B.-R.
,
Skoultchi
,
A. I.
and
Bai
,
Y.
(
2018
).
Emerging roles of linker histones in regulating chromatin structure and function
.
Nat. Rev. Mol. Cell Biol.
19
,
192
-
206
.
Gaertner
,
B.
,
Johnston
,
J.
,
Chen
,
K.
,
Wallaschek
,
N.
,
Paulson
,
A.
,
Garruss
,
A. S.
,
Gaudenz
,
K.
,
De Kumar
,
B.
,
Krumlauf
,
R.
and
Zeitlinger
,
J.
(
2012
).
Poised RNA polymerase II changes over developmental time and prepares genes for future expression
.
Cell Rep.
2
,
1670
-
1683
.
Galili
,
N.
,
Davis
,
R. J.
,
Fredericks
,
W. J.
,
Mukhopadhyay
,
S.
,
Rauscher
,
F. J.
,
Emanuel
,
B. S.
,
Rovera
,
G.
and
Barr
,
F. G.
(
1993
).
Fusion of a fork head domain gene to PAX3 in the solid tumour alveolar rhabdomyosarcoma
.
Nat. Genet.
5
,
230
-
235
.
Gaskill
,
M. M.
,
Gibson
,
T. J.
,
Larson
,
E. D.
and
Harrison
,
M. M.
(
2021
).
GAF is essential for zygotic genome activation and chromatin accessibility in the early Drosophila embryo
.
eLife
10
,
e66668
.
Gaskill
,
M. M.
,
Soluri
,
I. V.
,
Branks
,
A. E.
,
Boka
,
A. P.
,
Stadler
,
M. R.
,
Vietor
,
K.
,
Huang
,
H.-Y. S.
,
Gibson
,
T. J.
,
Mukherjee
,
A.
,
Mir
,
M.
et al.
(
2023
).
Localization of the Drosophila pioneer factor GAF to subnuclear foci is driven by DNA binding and required to silence satellite repeat expression
.
Dev. Cell
58
,
1610
-
1624.e8
.
Gasperoni
,
J. G.
,
Fuller
,
J. N.
,
Darido
,
C.
,
Wilanowski
,
T.
and
Dworkin
,
S.
(
2022
).
Grainyhead-like (Grhl) target genes in development and cancer
.
IJMS
23
,
2735
.
Gassler
,
J.
,
Kobayashi
,
W.
,
Gáspár
,
I.
,
Ruangroengkulrith
,
S.
,
Mohanan
,
A.
,
Gómez Hernández
,
L.
,
Kravchenko
,
P.
,
Kümmecke
,
M.
,
Lalic
,
A.
,
Rifel
,
N.
et al.
(
2022
).
Zygotic genome activation by the totipotency pioneer factor Nr5a2
.
Science
378
,
1305
-
1315
.
Gentsch
,
G. E.
,
Spruce
,
T.
,
Owens
,
N. D. L.
and
Smith
,
J. C.
(
2019
).
Maternal pluripotency factors initiate extensive chromatin remodelling to predefine first response to inductive signals
.
Nat. Commun.
10
,
4269
.
Gibson
,
T. J.
,
Larson
,
E. D.
and
Harrison
,
M. M.
(
2024
).
Protein-intrinsic properties and context-dependent effects regulate pioneer factor binding and function
.
Nat. Struct. Mol. Biol.
31
,
548
-
558
.
Gottesfeld
,
J.
(
1997
).
Mitotic repression of the transcriptional machinery
.
Trends Biochem. Sci.
22
,
197
-
202
.
Gouhier
,
A.
,
Dumoulin-Gagnon
,
J.
,
Lapointe-Roberge
,
V.
,
Harris
,
J.
,
Balsalobre
,
A.
and
Drouin
,
J.
(
2024
).
Pioneer factor Pax7 initiates two-step cell-cycle-dependent chromatin opening
.
Nat. Struct. Mol. Biol.
31
,
92
-
101
.
Gualdi
,
R.
,
Bossard
,
P.
,
Zheng
,
M.
,
Hamada
,
Y.
,
Coleman
,
J. R.
and
Zaret
,
K. S.
(
1996
).
Hepatic specification of the gut endoderm in vitro: cell signaling and transcriptional control
.
Genes Dev.
10
,
1670
-
1682
.
Guan
,
R.
,
Lian
,
T.
,
Zhou
,
B.-R.
and
Bai
,
Y.
(
2023
).
Structural mechanism of LIN28B nucleosome targeting by OCT4 for pluripotency
.
Mol. Cell
83
,
1970
-
1982.e6
.
Guo
,
Y.
,
Kitano
,
T.
,
Inoue
,
K.
,
Murano
,
K.
,
Hirose
,
M.
,
Li
,
T. D.
,
Sakashita
,
A.
,
Ishizu
,
H.
,
Ogonuki
,
N.
,
Matoba
S.
et al.
(
2024
).
Obox4 promotes zygotic genome activation upon loss of Dux
.
eLife
13
,
e95856
. [ePub ahead of print]
Hao
,
Y.
,
Han
,
L.
,
Wu
,
A.
and
Bochkis
,
I. M.
(
2024
).
Pioneer factor Foxa2 mediates chromatin conformation changes for activation of bile acid targets of FXR
.
Cell. Mol. Gastroenterol. Hepatol.
17
,
237
-
249
.
Hayward-Lara
,
G.
,
Fischer
,
M. D.
and
Mir
,
M.
(
2024
).
Dynamic microenvironments shape nuclear organization and gene expression
.
Curr. Opin. Genet. Dev.
86
,
102177
.
Heinz
,
S.
,
Romanoski
,
C. E.
,
Benner
,
C.
and
Glass
,
C. K.
(
2015
).
The selection and function of cell type-specific enhancers
.
Nat. Rev. Mol. Cell Biol.
16
,
144
-
154
.
Hendrickson
,
P. G.
,
Doráis
,
J. A.
,
Grow
,
E. J.
,
Whiddon
,
J. L.
,
Lim
,
J.-W.
,
Wike
,
C. L.
,
Weaver
,
B. D.
,
Pflueger
,
C.
,
Emery
,
B. R.
,
Wilcox
,
A. L.
et al.
(
2017
).
Conserved roles of mouse DUX and human DUX4 in activating cleavage-stage genes and MERVL/HERVL retrotransposons
.
Nat. Genet.
49
,
925
-
934
.
Horisawa
,
K.
and
Suzuki
,
A.
(
2023
).
The role of pioneer transcription factors in the induction of direct cellular reprogramming
.
Regen. Ther.
24
,
112
-
116
.
Hosokawa
,
H.
,
Ungerbäck
,
J.
,
Wang
,
X.
,
Matsumoto
,
M.
,
Nakayama
,
K. I.
,
Cohen
,
S. M.
,
Tanaka
,
T.
and
Rothenberg
,
E. V.
(
2018
).
Transcription factor PU.1 represses and activates gene expression in early T cells by redirecting partner transcription factor binding
.
Immunity
48
,
1119
-
1134.e7
.
Hovland
,
A. S.
,
Bhattacharya
,
D.
,
Azambuja
,
A. P.
,
Pramio
,
D.
,
Copeland
,
J.
,
Rothstein
,
M.
and
Simoes-Costa
,
M.
(
2022
).
Pluripotency factors are repurposed to shape the epigenomic landscape of neural crest cells
.
Dev. Cell
57
,
2257
-
2272.e5
.
Hsu
,
H.-T.
,
Chen
,
H.-M.
,
Yang
,
Z.
,
Wang
,
J.
,
Lee
,
N. K.
,
Burger
,
A.
,
Zaret
,
K.
,
Liu
,
T.
,
Levine
,
E.
and
Mango
,
S. E.
(
2015
).
Recruitment of RNA polymerase II by the pioneer transcription factor PHA-4
.
Science
348
,
1372
-
1376
.
Huang
,
P.
,
He
,
Z.
,
Ji
,
S.
,
Sun
,
H.
,
Xiang
,
D.
,
Liu
,
C.
,
Hu
,
Y.
,
Wang
,
X.
and
Hui
,
L.
(
2011
).
Induction of functional hepatocyte-like cells from mouse fibroblasts by defined factors
.
Nature
475
,
386
-
389
.
Huang
,
X.
,
Park
,
K.
,
Gontarz
,
P.
,
Zhang
,
B.
,
Pan
,
J.
,
McKenzie
,
Z.
,
Fischer
,
L. A.
,
Dong
,
C.
,
Dietmann
,
S.
,
Xing
,
X.
et al.
(
2021
).
OCT4 cooperates with distinct ATP-dependent chromatin remodelers in naïve and primed pluripotent states in human
.
Nat. Commun.
12
,
5123
.
Hug
,
C. B.
,
Grimaldi
,
A. G.
,
Kruse
,
K.
and
Vaquerizas
,
J. M.
(
2017
).
chromatin architecture emerges during zygotic genome activation independent of transcription
.
Cell
169
,
216
-
228.e19
.
Isaac
,
R. S.
,
Jiang
,
F.
,
Doudna
,
J. A.
,
Lim
,
W. A.
,
Narlikar
,
G. J.
and
Almeida
,
R.
(
2016
).
Nucleosome breathing and remodeling constrain CRISPR-Cas9 function
.
eLife
5
,
e13450
.
Iwafuchi
,
M.
,
Cuesta
,
I.
,
Donahue
,
G.
,
Takenaka
,
N.
,
Osipovich
,
A. B.
,
Magnuson
,
M. A.
,
Roder
,
H.
,
Seeholzer
,
S. H.
,
Santisteban
,
P.
and
Zaret
,
K. S.
(
2020
).
Gene network transitions in embryos depend upon interactions between a pioneer transcription factor and core histones
.
Nat. Genet.
52
,
418
-
427
.
Iwafuchi-Doi
,
M.
,
Donahue
,
G.
,
Kakumanu
,
A.
,
Watts
,
J. A.
,
Mahony
,
S.
,
Pugh
,
B. F.
,
Lee
,
D.
,
Kaestner
,
K. H.
and
Zaret
,
K. S.
(
2016
).
The pioneer transcription factor FoxA maintains an accessible nucleosome configuration at enhancers for tissue-specific gene activation
.
Mol. Cell
62
,
79
-
91
.
Jacobs
,
J.
,
Atkins
,
M.
,
Davie
,
K.
,
Imrichova
,
H.
,
Romanelli
,
L.
,
Christiaens
,
V.
,
Hulselmans
,
G.
,
Potier
,
D.
,
Wouters
,
J.
,
Taskiran
,
I. I.
et al.
(
2018
).
The transcription factor Grainy head primes epithelial enhancers for spatiotemporal activation by displacing nucleosomes
.
Nat. Genet.
50
,
1011
-
1020
.
Ji
,
S.
,
Chen
,
F.
,
Stein
,
P.
,
Wang
,
J.
,
Zhou
,
Z.
,
Wang
,
L.
,
Zhao
,
Q.
,
Lin
,
Z.
,
Liu
,
B.
,
Xu
,
K.
et al.
(
2023
).
OBOX regulates mouse zygotic genome activation and early development
.
Nature
620
,
1047
-
1053
.
Ji
,
D.
,
Shao
,
C.
,
Yu
,
J.
,
Hou
,
Y.
,
Gao
,
X.
,
Wu
,
Y.
,
Wang
,
L.
and
Chen
,
P.
(
2024
).
FOXA1 forms biomolecular condensates that unpack condensed chromatin to function as a pioneer factor
.
Mol. Cell
84
,
244
-
260.e7
.
Jozwik
,
K. M.
and
Carroll
,
J. S.
(
2012
).
Pioneer factors in hormone-dependent cancers
.
Nat. Rev. Cancer
12
,
381
-
385
.
Jozwik
,
K. M.
,
Chernukhin
,
I.
,
Serandour
,
A. A.
,
Nagarajan
,
S.
and
Carroll
,
J. S.
(
2016
).
FOXA1 directs H3K4 monomethylation at enhancers via recruitment of the methyltransferase MLL3
.
Cell Rep.
17
,
2715
-
2723
.
Judd
,
J.
,
Duarte
,
F. M.
and
Lis
,
J. T.
(
2021
).
Pioneer-like factor GAF cooperates with PBAP (SWI/SNF) and NURF (ISWI) to regulate transcription
.
Genes Dev.
35
,
147
-
156
.
Kadauke
,
S.
,
Udugama
,
M. I.
,
Pawlicki
,
J. M.
,
Achtman
,
J. C.
,
Jain
,
D. P.
,
Cheng
,
Y.
,
Hardison
,
R. C.
and
Blobel
,
G. A.
(
2012
).
Tissue-specific mitotic bookmarking by hematopoietic transcription factor GATA1
.
Cell
150
,
725
-
737
.
Kawabe
,
Y.
,
Wang
,
Y. X.
,
McKinnell
,
I. W.
,
Bedford
,
M. T.
and
Rudnicki
,
M. A.
(
2012
).
Carm1 regulates Pax7 transcriptional activity through MLL1/2 recruitment during asymmetric satellite stem cell divisions
.
Cell Stem Cell
11
,
333
-
345
.
Kim
,
M.
and
McGinnis
,
W.
(
2011
).
Phosphorylation of Grainy head by ERK is essential for wound-dependent regeneration but not for development of an epidermal barrier
.
Proc. Natl. Acad. Sci. USA
108
,
650
-
655
.
Kim
,
K.-P.
,
Choi
,
J.
,
Yoon
,
J.
,
Bruder
,
J. M.
,
Shin
,
B.
,
Kim
,
J.
,
Arauzo-Bravo
,
M. J.
,
Han
,
D.
,
Wu
,
G.
,
Han
,
D. W.
et al.
(
2021
).
Permissive epigenomes endow reprogramming competence to transcriptional regulators
.
Nat. Chem. Biol.
17
,
47
-
56
.
King
,
H. W.
and
Klose
,
R. J.
(
2017
).
The pioneer factor OCT4 requires the chromatin remodeller BRG1 to support gene regulatory element function in mouse embryonic stem cells
.
eLife
6
,
e22631
.
Kornberg
,
R. D.
and
Thomas
,
J. O.
(
1974
).
Chromatin structure; oligomers of the histones
.
Science
184
,
865
-
868
.
Koromila
,
T.
,
Gao
,
F.
,
Iwasaki
,
Y.
,
He
,
P.
,
Pachter
,
L.
,
Gergen
,
J. P.
and
Stathopoulos
,
A.
(
2020
).
Odd-paired is a pioneer-like factor that coordinates with Zelda to control gene expression in embryos
.
eLife
9
,
e59610
.
Kouzarides
,
T.
(
2007
).
Chromatin modifications and their function
.
Cell
128
,
693
-
705
.
Kuznetsova
,
K.
,
Chabot
,
N. M.
,
Ugolini
,
M.
,
Wu
,
E.
,
Lalit
,
M.
,
Oda
,
H.
,
Sato
,
Y.
,
Kimura
,
H.
,
Jug
,
F.
and
Vastenhouw
,
N. L.
(
2023
).
Nanog organizes transcription bodies
.
Curr. Biol.
33
,
164
-
173.e5
.
Lambourne
,
L.
,
Mattioli
,
K.
,
Santoso
,
C.
,
Sheynkman
,
G.
,
Inukai
,
S.
,
Kaundal
,
B.
,
Berenson
,
A.
,
Spirohn-Fitzgerald
,
K.
,
Bhattacharjee
,
A.
,
Rothman
,
E.
et al.
(
2024
).
Widespread variation in molecular interactions and regulatory properties among transcription factor isoforms
.
BioRxiv
.
Larson
,
E. D.
,
Komori
,
H.
,
Gibson
,
T. J.
,
Ostgaard
,
C. M.
,
Hamm
,
D. C.
,
Schnell
,
J. M.
,
Lee
,
C.-Y.
and
Harrison
,
M. M.
(
2021
).
Cell-type-specific chromatin occupancy by the pioneer factor Zelda drives key developmental transitions in Drosophila
.
Nat. Commun.
12
,
7153
.
Lee
,
C. S.
,
Friedman
,
J. R.
,
Fulmer
,
J. T.
and
Kaestner
,
K. H.
(
2005
).
The initiation of liver development is dependent on Foxa transcription factors
.
Nature
435
,
944
-
947
.
Lee
,
W.
,
Tillo
,
D.
,
Bray
,
N.
,
Morse
,
R. H.
,
Davis
,
R. W.
,
Hughes
,
T. R.
and
Nislow
,
C.
(
2007
).
A high-resolution atlas of nucleosome occupancy in yeast
.
Nat. Genet.
39
,
1235
-
1244
.
Lee
,
M. T.
,
Bonneau
,
A. R.
,
Takacs
,
C. M.
,
Bazzini
,
A. A.
,
DiVito
,
K. R.
,
Fleming
,
E. S.
and
Giraldez
,
A. J.
(
2013
).
Nanog, Pou5f1 and SoxB1 activate zygotic gene expression during the maternal-to-zygotic transition
.
Nature
503
,
360
-
364
.
Lee
,
K.
,
Cho
,
H.
,
Rickert
,
R. W.
,
Li
,
Q. V.
,
Pulecio
,
J.
,
Leslie
,
C. S.
and
Huangfu
,
D.
(
2019
).
FOXA2 is required for enhancer priming during pancreatic differentiation
.
Cell Rep.
28
,
382
-
393.e7
.
Leichsenring
,
M.
,
Maes
,
J.
,
Mössner
,
R.
,
Driever
,
W.
and
Onichtchouk
,
D.
(
2013
).
Pou5f1 transcription factor controls zygotic gene activation in vertebrates
.
Science
341
,
1005
-
1009
.
Lemma
,
R. B.
,
Fleischer
,
T.
,
Martinsen
,
E.
,
Ledsaak
,
M.
,
Kristensen
,
V.
,
Eskeland
,
R.
,
Gabrielsen
,
O. S.
and
Mathelier
,
A.
(
2022
).
Pioneer transcription factors are associated with the modulation of DNA methylation patterns across cancers
.
Epigenet. Chromatin
15
,
13
.
Lerner
,
J.
,
Katznelson
,
A.
,
Zhang
,
J.
and
Zaret
,
K. S.
(
2023
).
Different chromatin-scanning modes lead to targeting of compacted chromatin by pioneer factors FOXA1 and SOX2
.
Cell Rep.
42
,
112748
.
Levo
,
M.
,
Raimundo
,
J.
,
Bing
,
X. Y.
,
Sisco
,
Z.
,
Batut
,
P. J.
,
Ryabichko
,
S.
,
Gregor
,
T.
and
Levine
,
M. S.
(
2022
).
Transcriptional coupling of distant regulatory genes in living embryos
.
Nature
605
,
754
-
760
.
Li
,
B.
,
Adams
,
C. C.
and
Workman
,
J. L.
(
1994
).
Nucleosome binding by the constitutive transcription factor Sp1
.
J. Biol. Chem.
269
,
7756
-
7763
.
Li
,
B.
,
Carey
,
M.
and
Workman
,
J. L.
(
2007
).
The role of chromatin during transcription
.
Cell
128
,
707
-
719
.
Li
,
X.-Y.
,
Harrison
,
M. M.
,
Villalta
,
J. E.
,
Kaplan
,
T.
and
Eisen
,
M. B.
(
2014
).
Establishment of regions of genomic activity during the Drosophila maternal to zygotic transition
.
eLife
3
,
e03737
.
Li
,
D.
,
Liu
,
J.
,
Yang
,
X.
,
Zhou
,
C.
,
Guo
,
J.
,
Wu
,
C.
,
Qin
,
Y.
,
Guo
,
L.
,
He
,
J.
,
Yu
,
S.
et al.
(
2017
).
Chromatin accessibility dynamics during iPSC reprogramming
.
Cell Stem Cell
21
,
819
-
833.e6
.
Li
,
G.
,
Hao
,
W.
and
Hu
,
W.
(
2020
).
Transcription factor PU.1 and immune cell differentiation
.
Int. J. Mol. Med.
46
,
1943
-
1950
.
Li
,
L.
,
Lai
,
F.
,
Hu
,
X.
,
Liu
,
B.
,
Lu
,
X.
,
Lin
,
Z.
,
Liu
,
L.
,
Xiang
,
Y.
,
Frum
,
T.
,
Halbisen
,
M. A.
et al.
(
2023a
).
Multifaceted SOX2-chromatin interaction underpins pluripotency progression in early embryos
.
Science
382
,
eadi5516
.
Li
,
X.
,
Tang
,
X.
,
Bing
,
X.
,
Catalano
,
C.
,
Li
,
T.
,
Dolsten
,
G.
,
Wu
,
C.
and
Levine
,
M.
(
2023b
).
GAGA-associated factor fosters loop formation in the Drosophila genome
.
Mol. Cell
83
,
1519
-
1526.e4
.
Lian
,
T.
,
Guan
,
R.
,
Zhou
,
B.-R.
and
Bai
,
Y.
(
2023
).
The human LIN28B nucleosome is inherently pre-positioned for efficient binding of multiple OCT4s without H3 K27 acetylation
.
bioRxiv
.
Liang
,
H.-L.
,
Nien
,
C.-Y.
,
Liu
,
H.-Y.
,
Metzstein
,
M. M.
,
Kirov
,
N.
and
Rushlow
,
C.
(
2008
).
The zinc-finger protein Zelda is a key activator of the early zygotic genome in Drosophila
.
Nature
456
,
400
-
403
.
Liu
,
Z.
and
Kraus
,
W. L.
(
2017
).
Catalytic-independent functions of PARP-1 determine Sox2 pioneer activity at intractable genomic loci
.
Mol. Cell
65
,
589
-
603.e9
.
Liu
,
Y.
,
Pelham-Webb
,
B.
,
Di Giammartino
,
D. C.
,
Li
,
J.
,
Kim
,
D.
,
Kita
,
K.
,
Saiz
,
N.
,
Garg
,
V.
,
Doane
,
A.
,
Giannakakou
,
P.
et al.
(
2017
).
Widespread mitotic bookmarking by histone marks and transcription factors in pluripotent stem cells
.
Cell Rep.
19
,
1283
-
1293
.
Liu
,
H.
,
Tang
,
L.
,
Li
,
Y.
,
Xie
,
W.
,
Zhang
,
L.
,
Tang
,
H.
,
Xiao
,
T.
,
Yang
,
H.
,
Gu
,
W.
,
Wang
,
H.
et al.
(
2024
).
Nasopharyngeal carcinoma: current views on the tumor microenvironment's impact on drug resistance and clinical outcomes
.
Mol. Cancer
23
,
20
.
Lomaev
,
D.
,
Mikhailova
,
A.
,
Erokhin
,
M.
,
Shaposhnikov
,
A. V.
,
Moresco
,
J. J.
,
Blokhina
,
T.
,
Wolle
,
D.
,
Aoki
,
T.
,
Ryabykh
,
V.
,
Yates III
,
J. R.
et al.
(
2017
).
The GAGA factor regulatory network: identification of GAGA factor associated proteins
.
PLoS One
12
,
e0173602
.
Luo
,
W.
,
Li
,
S.
,
Peng
,
B.
,
Ye
,
Y.
,
Deng
,
X.
and
Yao
,
K.
(
2013
).
Embryonic stem cells markers SOX2, OCT4 and nanog expression and their correlations with epithelial-mesenchymal transition in nasopharyngeal carcinoma
.
PLoS One
8
,
e56324
.
Lupien
,
M.
,
Eeckhoute
,
J.
,
Meyer
,
C. A.
,
Wang
,
Q.
,
Zhang
,
Y.
,
Li
,
W.
,
Carroll
,
J. S.
,
Liu
,
X. S.
and
Brown
,
M.
(
2008
).
FoxA1 translates epigenetic signatures into enhancer-driven lineage-specific transcription
.
Cell
132
,
958
-
970
.
MacCarthy
,
C. M.
,
Huertas
,
J.
,
Ortmeier
,
C.
,
vom Bruch
,
H.
,
Tan
,
D. S.
,
Reinke
,
D.
,
Sander
,
A.
,
Bergbrede
,
T.
,
Jauch
,
R.
,
Schöler
,
H. R.
et al.
(
2022
).
OCT4 interprets and enhances nucleosome flexibility
.
Nucleic Acids Res.
50
,
10311
-
10327
.
Maresca
,
M.
,
van den Brand
,
T.
,
Li
,
H.
,
Teunissen
,
H.
,
Davies
,
J.
and
de Wit
,
E.
(
2023
).
Pioneer activity distinguishes activating from non–activating SOX2 binding sites
.
EMBO J.
42
,
e113150
.
Marro
,
S.
and
Yang
,
N.
(
2014
).
Transdifferentiation of mouse fibroblasts and hepatocytes to functional neurons
. In
Stem Cell Transcriptional Networks
(ed.
B. L.
Kidder
), pp.
237
-
246
.
New York, NY
:
Springer New York
.
Matoba
,
S.
,
Liu
,
Y.
,
Lu
,
F.
,
Iwabuchi
,
K. A.
,
Shen
,
L.
,
Inoue
,
A.
and
Zhang
,
Y.
(
2014
).
Embryonic development following somatic cell nuclear transfer impeded by persisting histone methylation
.
Cell
159
,
884
-
895
.
Matsui
,
S.
,
Granitto
,
M.
,
Buckley
,
M.
,
Ludwig
,
K.
,
Koigi
,
S.
,
Shiley
,
J.
,
Zacharias
,
W. J.
,
Mayhew
,
C. N.
,
Lim
,
H.-W.
and
Iwafuchi
,
M.
(
2024
).
Pioneer and PRDM transcription factors coordinate bivalent epigenetic states to safeguard cell fate
.
Mol. Cell
84
,
476
-
489.e10
.
Matsuoka
,
J.
,
Yashiro
,
M.
,
Sakurai
,
K.
,
Kubo
,
N.
,
Tanaka
,
H.
,
Muguruma
,
K.
,
Sawada
,
T.
,
Ohira
,
M.
and
Hirakawa
,
K.
(
2012
).
Role of the stemness factors Sox2, Oct3/4, and Nanog in gastric carcinoma
.
J. Surg. Res.
174
,
130
-
135
.
Mayran
,
A.
and
Drouin
,
J.
(
2018
).
Pioneer transcription factors shape the epigenetic landscape
.
J. Biol. Chem.
293
,
13795
-
13804
.
Mayran
,
A.
,
Khetchoumian
,
K.
,
Hariri
,
F.
,
Pastinen
,
T.
,
Gauthier
,
Y.
,
Balsalobre
,
A.
and
Drouin
,
J.
(
2018
).
Pioneer factor Pax7 deploys a stable enhancer repertoire for specification of cell fate
.
Nat. Genet.
50
,
259
-
269
.
Mayran
,
A.
,
Sochodolsky
,
K.
,
Khetchoumian
,
K.
,
Harris
,
J.
,
Gauthier
,
Y.
,
Bemmo
,
A.
,
Balsalobre
,
A.
and
Drouin
,
J.
(
2019
).
Pioneer and nonpioneer factor cooperation drives lineage specific chromatin opening
.
Nat. Commun.
10
,
3807
.
McCarthy
,
R. L.
,
Zhang
,
J.
and
Zaret
,
K. S.
(
2023
).
Diverse heterochromatin states restricting cell identity and reprogramming
.
Trends Biochem. Sci.
48
,
513
-
526
.
McDaniel
,
S. L.
,
Gibson
,
T. J.
,
Schulz
,
K. N.
,
Fernandez Garcia
,
M.
,
Nevil
,
M.
,
Jain
,
S. U.
,
Lewis
,
P. W.
,
Zaret
,
K. S.
and
Harrison
,
M. M.
(
2019
).
Continued activity of the pioneer factor Zelda is required to drive zygotic genome activation
.
Mol. Cell
74
,
185
-
195.e4
.
McKinnell
,
I. W.
,
Ishibashi
,
J.
,
Le Grand
,
F.
,
Punch
,
V. G. J.
,
Addicks
,
G. C.
,
Greenblatt
,
J. F.
,
Dilworth
,
F. J.
and
Rudnicki
,
M. A.
(
2008
).
Pax7 activates myogenic genes by recruitment of a histone methyltransferase complex
.
Nat. Cell Biol.
10
,
77
-
84
.
Meers
,
M. P.
,
Janssens
,
D. H.
and
Henikoff
,
S.
(
2019
).
Pioneer factor-nucleosome binding events during differentiation are motif encoded
.
Mol. Cell
75
,
562
-
575.e5
.
Miao
,
L.
,
Tang
,
Y.
,
Bonneau
,
A. R.
,
Chan
,
S. H.
,
Kojima
,
M. L.
,
Pownall
,
M. E.
,
Vejnar
,
C. E.
,
Gao
,
F.
,
Krishnaswamy
,
S.
,
Hendry
,
C. E.
et al.
(
2022
).
The landscape of pioneer factor activity reveals the mechanisms of chromatin reprogramming and genome activation
.
Mol. Cell
82
,
986
-
1002.e9
.
Michael
,
A. K.
,
Grand
,
R. S.
,
Isbel
,
L.
,
Cavadini
,
S.
,
Kozicka
,
Z.
,
Kempf
,
G.
,
Bunker
,
R. D.
,
Schenk
,
A. D.
,
Graff-Meyer
,
A.
,
Pathare
,
G. R.
et al.
(
2020
).
Mechanisms of OCT4-SOX2 motif readout on nucleosomes
.
Science
368
,
1460
-
1465
.
Minderjahn
,
J.
,
Schmidt
,
A.
,
Fuchs
,
A.
,
Schill
,
R.
,
Raithel
,
J.
,
Babina
,
M.
,
Schmidl
,
C.
,
Gebhard
,
C.
,
Schmidhofer
,
S.
,
Mendes
,
K.
et al.
(
2020
).
Mechanisms governing the pioneering and redistribution capabilities of the non-classical pioneer PU.1
.
Nat. Commun.
11
,
402
.
Mir
,
M.
,
Stadler
,
M. R.
,
Ortiz
,
S. A.
,
Hannon
,
C. E.
,
Harrison
,
M. M.
,
Darzacq
,
X.
and
Eisen
,
M. B.
(
2018
).
Dynamic multifactor hubs interact transiently with sites of active transcription in Drosophila embryos
.
eLife
7
,
e40497
.
Mivelaz
,
M.
,
Cao
,
A.-M.
,
Kubik
,
S.
,
Zencir
,
S.
,
Hovius
,
R.
,
Boichenko
,
I.
,
Stachowicz
,
A. M.
,
Kurat
,
C. F.
,
Shore
,
D.
and
Fierz
,
B.
(
2020
).
Chromatin fiber invasion and nucleosome displacement by the Rap1 transcription factor
.
Mol. Cell
77
,
488
-
500.e9
.
Mocciaro
,
E.
,
Runfola
,
V.
,
Ghezzi
,
P.
,
Pannese
,
M.
and
Gabellini
,
D.
(
2021
).
DUX4 role in normal physiology and in FSHD muscular dystrophy
.
Cells
10
,
3322
.
Morin
,
J. A.
,
Wittmann
,
S.
,
Choubey
,
S.
,
Klosin
,
A.
,
Golfier
,
S.
,
Hyman
,
A. A.
,
Jülicher
,
F.
and
Grill
,
S. W.
(
2022
).
Sequence-dependent surface condensation of a pioneer transcription factor on DNA
.
Nat. Phys.
18
,
271
-
276
.
Nakagawa
,
M.
,
Koyanagi
,
M.
,
Tanabe
,
K.
,
Takahashi
,
K.
,
Ichisaka
,
T.
,
Aoi
,
T.
,
Okita
,
K.
,
Mochiduki
,
Y.
,
Takizawa
,
N.
and
Yamanaka
,
S.
(
2008
).
Generation of induced pluripotent stem cells without Myc from mouse and human fibroblasts
.
Nat. Biotechnol.
26
,
101
-
106
.
Nakayama
,
T.
,
Shimojima
,
T.
and
Hirose
,
S.
(
2012
).
The PBAP remodeling complex is required for histone H3.3 replacement at chromatin boundaries and for boundary functions
.
Development
139
,
4582
-
4590
.
Naumova
,
N.
,
Imakaev
,
M.
,
Fudenberg
,
G.
,
Zhan
,
Y.
,
Lajoie
,
B. R.
,
Mirny
,
L. A.
and
Dekker
,
J.
(
2013
).
Organization of the mitotic chromosome
.
Science
342
,
948
-
953
.
Nevil
,
M.
,
Bondra
,
E. R.
,
Schulz
,
K. N.
,
Kaplan
,
T.
and
Harrison
,
M. M.
(
2017
).
Stable binding of the conserved transcription factor grainy head to its target genes throughout Drosophila melanogaster development
.
Genetics
205
,
605
-
620
.
Nevil
,
M.
,
Gibson
,
T. J.
,
Bartolutti
,
C.
,
Iyengar
,
A.
and
Harrison
,
M. M.
(
2020
).
Establishment of chromatin accessibility by the conserved transcription factor Grainy head is developmentally regulated
.
Development
147
,
dev185009
.
Nguyen
,
T.
,
Li
,
S.
,
Chang
,
J. T.-H.
,
Watters
,
J. W.
,
Ng
,
H.
,
Osunsade
,
A.
,
David
,
Y.
and
Liu
,
S.
(
2022
).
Chromatin sequesters pioneer transcription factor Sox2 from exerting force on DNA
.
Nat. Commun.
13
,
3988
.
Nüsslein-Volhard
,
C.
,
Wieschaus
,
E.
and
Kluding
,
H.
(
1984
).
Mutations affecting the pattern of the larval cuticle in Drosophila melanogaster : I. Zygotic loci on the second chromosome
.
Wilehm Roux Arch. Dev. Biol.
193
,
267
-
282
.
Ogiyama
,
Y.
,
Schuettengruber
,
B.
,
Papadopoulos
,
G. L.
,
Chang
,
J.-M.
and
Cavalli
,
G.
(
2018
).
Polycomb-dependent chromatin looping contributes to gene silencing during drosophila development
.
Mol. Cell
71
,
73
-
88.e5
.
Ohnishi
,
K.
,
Semi
,
K.
,
Yamamoto
,
T.
,
Shimizu
,
M.
,
Tanaka
,
A.
,
Mitsunaga
,
K.
,
Okita
,
K.
,
Osafune
,
K.
,
Arioka
,
Y.
,
Maeda
,
T.
et al.
(
2014
).
Premature termination of reprogramming in vivo leads to cancer development through altered epigenetic regulation
.
Cell
156
,
663
-
677
.
Onder
,
T. T.
,
Kara
,
N.
,
Cherry
,
A.
,
Sinha
,
A. U.
,
Zhu
,
N.
,
Bernt
,
K. M.
,
Cahan
,
P.
,
Mancarci
,
B. O.
,
Unternaehrer
,
J.
,
Gupta
,
P. B.
et al.
(
2012
).
Chromatin-modifying enzymes as modulators of reprogramming
.
Nature
483
,
598
-
602
.
Ozden
,
B.
,
Boopathi
,
R.
,
Barlas
,
A. B.
,
Lone
,
I. N.
,
Bednar
,
J.
,
Petosa
,
C.
,
Kale
,
S.
,
Hamiche
,
A.
,
Angelov
,
D.
,
Dimitrov
,
S.
et al.
(
2023
).
Molecular mechanism of nucleosome recognition by the pioneer transcription factor sox
.
J. Chem. Inf. Model.
63
,
3839
-
3853
.
Paltoglou
,
S.
,
Das
,
R.
,
Townley
,
S. L.
,
Hickey
,
T. E.
,
Tarulli
,
G. A.
,
Coutinho
,
I.
,
Fernandes
,
R.
,
Hanson
,
A. R.
,
Denis
,
I.
,
Carroll
,
J. S.
et al.
(
2017
).
Novel androgen receptor coregulator GRHL2 exerts both oncogenic and antimetastatic functions in prostate cancer
.
Cancer Res.
77
,
3417
-
3430
.
Paraiso
,
K. D.
,
Blitz
,
I. L.
,
Coley
,
M.
,
Cheung
,
J.
,
Sudou
,
N.
,
Taira
,
M.
and
Cho
,
K. W. Y.
(
2019
).
Endodermal maternal transcription factors establish super-enhancers during zygotic genome activation
.
Cell Rep.
27
,
2962
-
2977.e5
.
Park
,
S. H.
,
Fong
,
K.
,
Kim
,
J.
,
Wang
,
F.
,
Lu
,
X.
,
Lee
,
Y.
,
Brea
,
L. T.
,
Wadosky
,
K.
,
Guo
,
C.
,
Abdulkadir
,
S. A.
et al.
(
2021
).
Posttranslational regulation of FOXA1 by Polycomb and BUB3/USP7 deubiquitin complex in prostate cancer
.
Sci. Adv.
7
,
eabe2261
.
Păun
,
O.
,
Tan
,
Y. X.
,
Patel
,
H.
,
Strohbuecker
,
S.
,
Ghanate
,
A.
,
Cobolli-Gigli
,
C.
,
Llorian Sopena
,
M.
,
Gerontogianni
,
L.
,
Goldstone
,
R.
,
Ang
,
S.-L.
et al.
(
2023
).
Pioneer factor ASCL1 cooperates with the mSWI/SNF complex at distal regulatory elements to regulate human neural differentiation
.
Genes Dev.
37
,
218
-
242
.
Peng
,
L.
,
Li
,
E.-M.
and
Xu
,
L.-Y.
(
2020
).
From start to end: phase separation and transcriptional regulation
.
Biochim. Biophys. Acta - Gene Regul. Mech.
1863
,
194641
.
Pfisterer
,
U.
,
Kirkeby
,
A.
,
Torper
,
O.
,
Wood
,
J.
,
Nelander
,
J.
,
Dufour
,
A.
,
Björklund
,
A.
,
Lindvall
,
O.
,
Jakobsson
,
J.
and
Parmar
,
M.
(
2011
).
Direct conversion of human fibroblasts to dopaminergic neurons
.
Proc. Natl. Acad. Sci. USA
108
,
10343
-
10348
.
Phanstiel
,
D. H.
,
Van Bortle
,
K.
,
Spacek
,
D.
,
Hess
,
G. T.
,
Shamim
,
M. S.
,
Machol
,
I.
,
Love
,
M. I.
,
Aiden
,
E. L.
,
Bassik
,
M. C.
and
Snyder
,
M. P.
(
2017
).
Static and dynamic DNA loops form AP-1-bound activation hubs during macrophage development
.
Mol. Cell
67
,
1037
-
1048.e6
.
Pineda
,
J. M. B.
and
Bradley
,
R. K.
(
2024
).
DUX4 is a common driver of immune evasion and immunotherapy failure in metastatic cancers
.
eLife
12
,
RP89017
.
Poirier
,
M. G.
,
Bussiek
,
M.
,
Langowski
,
J.
and
Widom
,
J.
(
2008
).
Spontaneous access to DNA target sites in folded chromatin fibers
.
J. Mol. Biol.
379
,
772
-
786
.
Price
,
R. M.
,
Budzyński
,
M. A.
,
Shen
,
J.
,
Mitchell
,
J. E.
,
Kwan
,
J. Z. J.
and
Teves
,
S. S.
(
2023
).
Heat shock transcription factors demonstrate a distinct mode of interaction with mitotic chromosomes
.
Nucleic Acids Res.
51
,
5040
-
5055
.
Raccaud
,
M.
,
Friman
,
E. T.
,
Alber
,
A. B.
,
Agarwal
,
H.
,
Deluz
,
C.
,
Kuhn
,
T.
,
Gebhardt
,
J. C. M.
and
Suter
,
D. M.
(
2019
).
Mitotic chromosome binding predicts transcription factor properties in interphase
.
Nat. Commun.
10
,
487
.
Rada-Iglesias
,
A.
,
Bajpai
,
R.
,
Swigut
,
T.
,
Brugmann
,
S. A.
,
Flynn
,
R. A.
and
Wysocka
,
J.
(
2011
).
A unique chromatin signature uncovers early developmental enhancers in humans
.
Nature
470
,
279
-
283
.
Raff
,
J. W.
,
Kellum
,
R.
and
Alberts
,
B.
(
1994
).
The Drosophila GAGA transcription factor is associated with specific regions of heterochromatin throughout the cell cycle
.
EMBO J.
13
,
5977
-
5983
.
Ramalingam
,
V.
,
Yu
,
X.
,
Slaughter
,
B. D.
,
Unruh
,
J. R.
,
Brennan
,
K. J.
,
Onyshchenko
,
A.
,
Lange
,
J. J.
,
Natarajan
,
M.
,
Buck
,
M.
and
Zeitlinger
,
J.
(
2023
).
Lola-I is a promoter pioneer factor that establishes de novo Pol II pausing during development
.
Nat. Commun.
14
,
5862
.
Reese
,
R. M.
,
Harrison
,
M. M.
and
Alarid
,
E. T.
(
2019
).
Grainyhead-like protein 2: the emerging role in hormone-dependent cancers and epigenetics
.
Endocrinology
160
,
1275
-
1288
.
Reese
,
R. M.
,
Helzer
,
K. T.
,
Allen
,
K. O.
,
Zheng
,
C.
,
Solodin
,
N.
and
Alarid
,
E. T.
(
2022
).
GRHL2 enhances phosphorylated estrogen receptor (ER) chromatin binding and regulates ER-mediated transcriptional activation and repression
.
Mol. Cell. Biol.
42
,
e00191
-
e00122
.
Reya
,
T.
,
Morrison
,
S. J.
,
Clarke
,
M. F.
and
Weissman
,
I. L.
(
2001
).
Stem cells, cancer, and cancer stem cells
.
Nature
414
,
105
-
111
.
Riggi
,
N.
,
Suvà
,
M.-L.
,
Suvà
,
D.
,
Cironi
,
L.
,
Provero
,
P.
,
Tercier
,
S.
,
Joseph
,
J.-M.
,
Stehle
,
J.-C.
,
Baumer
,
K.
,
Kindler
,
V.
et al.
(
2008
).
EWS-FLI-1 expression triggers a ewing's sarcoma initiation program in primary human mesenchymal stem cells
.
Cancer Res.
68
,
2176
-
2185
.
Riggi
,
N.
,
Knoechel
,
B.
,
Gillespie
,
S. M.
,
Rheinbay
,
E.
,
Boulay
,
G.
,
Suvà
,
M. L.
,
Rossetti
,
N. E.
,
Boonseng
,
W. E.
,
Oksuz
,
O.
,
Cook
,
E. B.
et al.
(
2014
).
EWS-FLI1 utilizes divergent chromatin remodeling mechanisms to directly activate or repress enhancer elements in ewing sarcoma
.
Cancer Cell
26
,
668
-
681
.
Rippe
,
K.
(
2022
).
Liquid–liquid phase separation in chromatin
.
Cold Spring Harb. Perspect. Biol.
14
,
a040683
.
Saigusa
,
S.
,
Tanaka
,
K.
,
Toiyama
,
Y.
,
Yokoe
,
T.
,
Okugawa
,
Y.
,
Ioue
,
Y.
,
Miki
,
C.
and
Kusunoki
,
M.
(
2009
).
Correlation of CD133, OCT4, and SOX2 in rectal cancer and their association with distant recurrence after chemoradiotherapy
.
Ann. Surg. Oncol.
16
,
3488
-
3498
.
Sardina
,
J. L.
,
Collombet
,
S.
,
Tian
,
T. V.
,
Gómez
,
A.
,
Di Stefano
,
B.
,
Berenguer
,
C.
,
Brumbaugh
,
J.
,
Stadhouders
,
R.
,
Segura-Morales
,
C.
,
Gut
,
M.
et al.
(
2018
).
Transcription factors drive Tet2-mediated enhancer demethylation to reprogram cell fate
.
Cell Stem Cell
23
,
727
-
741.e9
.
Schulz
,
K. N.
and
Harrison
,
M. M.
(
2019
).
Mechanisms regulating zygotic genome activation
.
Nat. Rev. Genet.
20
,
221
-
234
.
Schulz
,
K. N.
,
Bondra
,
E. R.
,
Moshe
,
A.
,
Villalta
,
J. E.
,
Lieb
,
J. D.
,
Kaplan
,
T.
,
McKay
,
D. J.
and
Harrison
,
M. M.
(
2015
).
Zelda is differentially required for chromatin accessibility, transcription factor binding, and gene expression in the early Drosophila embryo
.
Genome Res.
25
,
1715
-
1726
.
Seachrist
,
D. D.
,
Anstine
,
L. J.
and
Keri
,
R. A.
(
2021
).
FOXA1: a pioneer of nuclear receptor action in breast cancer
.
Cancers
13
,
5205
.
Sekiya
,
S.
and
Suzuki
,
A.
(
2011
).
Direct conversion of mouse fibroblasts to hepatocyte-like cells by defined factors
.
Nature
475
,
390
-
393
.
Sharma
,
R.
,
Choi
,
K.-J.
,
Quan
,
M. D.
,
Sharma
,
S.
,
Sankaran
,
B.
,
Park
,
H.
,
LaGrone
,
A.
,
Kim
,
J. J.
,
MacKenzie
,
K. R.
,
Ferreon
,
A. C. M.
et al.
(
2021
).
Liquid condensation of reprogramming factor KLF4 with DNA provides a mechanism for chromatin organization
.
Nat. Commun.
12
,
5579
.
Shevelyov
,
Y. Y.
(
2023
).
Interactions of chromatin with the nuclear lamina and nuclear pore complexes
.
Int. J. Mol. Sci.
24
,
15771
.
Sholl
,
L. M.
,
Barletta
,
J. A.
,
Yeap
,
B. Y.
,
Chirieac
,
L. R.
and
Hornick
,
J. L.
(
2010
).
Sox2 protein expression is an independent poor prognostic indicator in stage i lung adenocarcinoma
.
Am. J. Surg. Pathol.
34
,
1193
-
1198
.
Silvério-Alves
,
R.
,
Kurochkin
,
I.
,
Rydström
,
A.
,
Vazquez Echegaray
,
C.
,
Haider
,
J.
,
Nicholls
,
M.
,
Rode
,
C.
,
Thelaus
,
L.
,
Lindgren
,
A. Y.
,
Ferreira
,
A. G.
et al.
(
2023
).
GATA2 mitotic bookmarking is required for definitive haematopoiesis
.
Nat. Commun.
14
,
4645
.
Sincennes
,
M.-C.
,
Brun
,
C. E.
,
Lin
,
A. Y. T.
,
Rosembert
,
T.
,
Datzkiw
,
D.
,
Saber
,
J.
,
Ming
,
H.
,
Kawabe
,
Y.
and
Rudnicki
,
M. A.
(
2021
).
Acetylation of PAX7 controls muscle stem cell self-renewal and differentiation potential in mice
.
Nat. Commun.
12
,
3253
.
Sinha
,
K.
,
Bilokapic
,
S.
,
Du
,
Y.
,
Malik
,
D.
and
Halic
,
M.
(
2023
).
Histone modifications regulate pioneer transcription factor binding and cooperativity
.
Nature
619
,
378
-
384
.
Soares
,
M. A. F.
,
Soares
,
D. S.
,
Teixeira
,
V.
,
Heskol
,
A.
,
Bressan
,
R. B.
,
Pollard
,
S. M.
,
Oliveira
,
R. A.
and
Castro
,
D. S.
(
2021
).
Hierarchical reactivation of transcription during mitosis-to-G1 transition by Brn2 and Ascl1 in neural stem cells
.
Genes Dev.
35
,
1020
-
1034
.
Soluri
,
I. V.
,
Zumerling
,
L. M.
,
Payan Parra
,
O. A.
,
Clark
,
E. G.
and
Blythe
,
S. A.
(
2020
).
Zygotic pioneer factor activity of Odd-paired/Zic is necessary for late function of the Drosophila segmentation network
.
eLife
9
,
e53916
.
Soufi
,
A.
,
Donahue
,
G.
and
Zaret
,
K. S.
(
2012
).
Facilitators and impediments of the pluripotency reprogramming factors’ initial engagement with the genome
.
Cell
151
,
994
-
1004
.
Soufi
,
A.
,
Garcia
,
M. F.
,
Jaroszewicz
,
A.
,
Osman
,
N.
,
Pellegrini
,
M.
and
Zaret
,
K. S.
(
2015
).
Pioneer transcription factors target partial DNA motifs on nucleosomes to initiate reprogramming
.
Cell
161
,
555
-
568
.
Spelat
,
R.
,
Ferro
,
F.
and
Curcio
,
F.
(
2012
).
Serine 111 phosphorylation regulates OCT4A protein subcellular distribution and degradation
.
J. Biol. Chem.
287
,
38279
-
38288
.
Spitz
,
F.
and
Furlong
,
E. E. M.
(
2012
).
Transcription factors: from enhancer binding to developmental control
.
Nat. Rev. Genet.
13
,
613
-
626
.
Stergachis
,
A. B.
,
Debo
,
B. M.
,
Haugen
,
E.
,
Churchman
,
L. S.
and
Stamatoyannopoulos
,
J. A.
(
2020
).
Single-molecule regulatory architectures captured by chromatin fiber sequencing
.
Science
368
,
1449
-
1454
.
Stoeber
,
S.
,
Godin
,
H.
,
Xu
,
C.
and
Bai
,
L.
(
2024
).
Pioneer factors: nature or nurture?
Crit. Rev. Biochem. Mol. Biol.
22
,
1
-
15
.
Sun
,
Y.
,
Nien
,
C.-Y.
,
Chen
,
K.
,
Liu
,
H.-Y.
,
Johnston
,
J.
,
Zeitlinger
,
J.
and
Rushlow
,
C.
(
2015
).
Zelda overcomes the high intrinsic nucleosome barrier at enhancers during Drosophila zygotic genome activation
.
Genome Res.
25
,
1703
-
1714
.
Sundararajan
,
V.
,
Pang
,
Q. Y.
,
Choolani
,
M.
and
Huang
,
R. Y.-J.
(
2020
).
Spotlight on the granules (Grainyhead-like proteins) – from an evolutionary conserved controller of epithelial trait to pioneering the chromatin landscape
.
Front. Mol. Biosci.
7
,
213
.
Sunkel
,
B. D.
and
Stanton
,
B. Z.
(
2021
).
Pioneer factors in development and cancer
.
iScience
24
,
103132
.
Sutinen
,
P.
,
Rahkama
,
V.
,
Rytinki
,
M.
and
Palvimo
,
J. J.
(
2014
).
Nuclear mobility and activity of FOXA1 with androgen receptor are regulated by SUMOylation
.
Mol. Endocrinol.
28
,
1719
-
1728
.
Takahashi
,
K.
and
Yamanaka
,
S.
(
2006
).
Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors
.
Cell
126
,
663
-
676
.
Takaku
,
M.
,
Grimm
,
S. A.
,
Shimbo
,
T.
,
Perera
,
L.
,
Menafra
,
R.
,
Stunnenberg
,
H. G.
,
Archer
,
T. K.
,
Machida
,
S.
,
Kurumizaka
,
H.
and
Wade
,
P. A.
(
2016
).
GATA3-dependent cellular reprogramming requires activation-domain dependent recruitment of a chromatin remodeler
.
Genome Biol.
17
,
36
.
Tang
,
X.
,
Li
,
T.
,
Liu
,
S.
,
Wisniewski
,
J.
,
Zheng
,
Q.
,
Rong
,
Y.
,
Lavis
,
L. D.
and
Wu
,
C.
(
2022
).
Kinetic principles underlying pioneer function of GAGA transcription factor in live cells
.
Nat. Struct. Mol. Biol.
29
,
665
-
676
.
Taube
,
J. H.
,
Allton
,
K.
,
Duncan
,
S. A.
,
Shen
,
L.
and
Barton
,
M. C.
(
2010
).
Foxa1 functions as a pioneer transcription factor at transposable elements to activate afp during differentiation of embryonic stem cells
.
J. Biol. Chem.
285
,
16135
-
16144
.
Teves
,
S. S.
,
An
,
L.
,
Hansen
,
A. S.
,
Xie
,
L.
,
Darzacq
,
X.
and
Tjian
,
R.
(
2016
).
A dynamic mode of mitotic bookmarking by transcription factors
.
eLife
5
,
e22280
.
Theis
,
A.
and
Harrison
,
M. M.
(
2023
).
Reprogramming of three-dimensional chromatin organization in the early embryo
.
Curr. Opin. Struct. Biol.
81
,
102613
.
Ungerbäck
,
J.
,
Hosokawa
,
H.
,
Wang
,
X.
,
Strid
,
T.
,
Williams
,
B. A.
,
Sigvardsson
,
M.
and
Rothenberg
,
E. V.
(
2018
).
Pioneering, chromatin remodeling, and epigenetic constraint in early T-cell gene regulation by SPI1 (PU.1)
.
Genome Res.
28
,
1508
-
1519
.
Uv
,
A. E.
,
Harrison
,
E. J.
and
Bray
,
S. J.
(
1997
).
Tissue-specific splicing and functions of the Drosophila transcription factor Grainyhead
.
Mol. Cell. Biol.
17
,
6727
-
6735
.
Vanzan
,
L.
,
Soldati
,
H.
,
Ythier
,
V.
,
Anand
,
S.
,
Braun
,
S. M. G.
,
Francis
,
N.
and
Murr
,
R.
(
2021
).
High throughput screening identifies SOX2 as a super pioneer factor that inhibits DNA methylation maintenance at its binding sites
.
Nat. Commun.
12
,
3337
.
Vastenhouw
,
N. L.
,
Cao
,
W. X.
and
Lipshitz
,
H. D.
(
2019
).
The maternal-to-zygotic transition revisited
.
Development
146
,
dev161471
.
Veil
,
M.
,
Yampolsky
,
L. Y.
,
Grüning
,
B.
and
Onichtchouk
,
D.
(
2019
).
Pou5f3, SoxB1, and Nanog remodel chromatin on high nucleosome affinity regions at zygotic genome activation
.
Genome Res.
29
,
383
-
395
.
Venkatesan
,
K.
,
McManus
,
H. R.
,
Mello
,
C. C.
,
Smith
,
T. F.
and
Hansen
,
U.
(
2003
).
Functional conservation between members of an ancient duplicated transcription factor family, LSF/Grainyhead
.
Nucleic Acids Res.
31
,
4304
-
4316
.
Vierbuchen
,
T.
,
Ostermeier
,
A.
,
Pang
,
Z. P.
,
Kokubu
,
Y.
,
Südhof
,
T. C.
and
Wernig
,
M.
(
2010
).
Direct conversion of fibroblasts to functional neurons by defined factors
.
Nature
463
,
1035
-
1041
.
Wang
,
S.
and
Samakovlis
,
C.
(
2012
).
Grainy head and its target genes in epithelial morphogenesis and wound healing
. In
Current Topics in Developmental Biology
(ed.
S.
Plaza
and
F.
Payre
), pp.
35
-
63
.
Elsevier
.
Wang
,
Z.-Y.
and
Yin
,
L.
(
2015
).
Estrogen receptor alpha-36 (ER-α36): a new player in human breast cancer
.
Mol. Cell. Endocrinol.
418
,
193
-
206
.
Wang
,
Z.
,
Zang
,
C.
,
Rosenfeld
,
J. A.
,
Schones
,
D. E.
,
Barski
,
A.
,
Cuddapah
,
S.
,
Cui
,
K.
,
Roh
,
T.-Y.
,
Peng
,
W.
,
Zhang
,
M. Q.
et al.
(
2008
).
Combinatorial patterns of histone acetylations and methylations in the human genome
.
Nat. Genet.
40
,
897
-
903
.
Wang
,
A.
,
Yue
,
F.
,
Li
,
Y.
,
Xie
,
R.
,
Harper
,
T.
,
Patel
,
N. A.
,
Muth
,
K.
,
Palmer
,
J.
,
Qiu
,
Y.
,
Wang
,
J.
et al.
(
2015
).
Epigenetic priming of enhancers predicts developmental competence of hESC-derived endodermal lineage intermediates
.
Cell Stem Cell
16
,
386
-
399
.
Wang
,
J.
,
Yu
,
H.
,
Ma
,
Q.
,
Zeng
,
P.
,
Wu
,
D.
,
Hou
,
Y.
,
Liu
,
X.
,
Jia
,
L.
,
Sun
,
J.
,
Chen
,
Y.
et al.
(
2021
).
Phase separation of OCT4 controls TAD reorganization to promote cell fate transitions
.
Cell Stem Cell
28
,
1868
-
1883.e11
.
Wapinski
,
O. L.
,
Vierbuchen
,
T.
,
Qu
,
K.
,
Lee
,
Q. Y.
,
Chanda
,
S.
,
Fuentes
,
D. R.
,
Giresi
,
P. G.
,
Ng
,
Y. H.
,
Marro
,
S.
,
Neff
,
N. F.
et al.
(
2013
).
Hierarchical mechanisms for direct reprogramming of fibroblasts to neurons
.
Cell
155
,
621
-
635
.
Wei
,
F.
,
Schöler
,
H. R.
and
Atchison
,
M. L.
(
2007
).
Sumoylation of Oct4 enhances its stability, DNA binding, and transactivation
.
J. Biol. Chem.
282
,
21551
-
21560
.
Wei
,
P.
,
Niu
,
M.
,
Pan
,
S.
,
Zhou
,
Y.
,
Shuai
,
C.
,
Wang
,
J.
,
Peng
,
S.
and
Li
,
G.
(
2014
).
Cancer stem-like cell: a novel target for nasopharyngeal carcinoma therapy
.
Stem Cell Res. Ther.
5
,
44
.
Wei
,
X.
,
Murphy
,
M. A.
,
Reddy
,
N. A.
,
Hao
,
Y.
,
Eggertsen
,
T. G.
,
Saucerman
,
J. J.
and
Bochkis
,
I. M.
(
2022
).
Redistribution of lamina-associated domains reshapes binding of pioneer factor FOXA2 in development of nonalcoholic fatty liver disease
.
Genome Res.
32
,
1981
-
1992
.
Whiddon
,
J. L.
,
Langford
,
A. T.
,
Wong
,
C.-J.
,
Zhong
,
J. W.
and
Tapscott
,
S. J.
(
2017
).
Conservation and innovation in the DUX4-family gene network
.
Nat. Genet.
49
,
935
-
940
.
Whitton
,
H.
,
Singh
,
L. N.
,
Patrick
,
M. A.
,
Price
,
A. J.
,
Osorio
,
F. G.
,
López-Otín
,
C.
and
Bochkis
,
I. M.
(
2018
).
Changes at the nuclear lamina alter binding of pioneer factor Foxa2 in aged liver
.
Aging Cell
17
,
e12742
.
Wilanowski
,
T.
,
Tuckfield
,
A.
,
Cerruti
,
L.
,
O'Connell
,
S.
,
Saint
,
R.
,
Parekh
,
V.
,
Tao
,
J.
,
Cunningham
,
J. M.
and
Jane
,
S. M.
(
2002
).
A highly conserved novel family of mammalian developmental transcription factors related to Drosophila grainyhead
.
Mech. Dev.
114
,
37
-
50
.
Williams
,
C. A. C.
,
Soufi
,
A.
and
Pollard
,
S. M.
(
2020
).
Post-translational modification of SOX family proteins: key biochemical targets in cancer?
Semin. Cancer Biol.
67
,
30
-
38
.
Wolf
,
B. K.
,
Zhao
,
Y.
,
McCray
,
A.
,
Hawk
,
W. H.
,
Deary
,
L. T.
,
Sugiarto
,
N. W.
,
LaCroix
,
I. S.
,
Gerber
,
S. A.
,
Cheng
,
C.
and
Wang
,
X.
(
2023
).
Cooperation of chromatin remodeling SWI/SNF complex and pioneer factor AP-1 shapes 3D enhancer landscapes
.
Nat. Struct. Mol. Biol.
30
,
10
-
21
.
Wolfrum
,
C.
,
Besser
,
D.
,
Luca
,
E.
and
Stoffel
,
M.
(
2003
).
Insulin regulates the activity of forkhead transcription factor Hnf-3β/Foxa-2 by Akt-mediated phosphorylation and nuclear/cytosolic localization
.
Proc. Natl. Acad. Sci. USA
100
,
11624
-
11629
.
Xu
,
C.
,
Kleinschmidt
,
H.
,
Yang
,
J.
,
Leith
,
E.
,
Johnson
,
J.
,
Tan
,
S.
,
Mahony
,
S.
and
Bai
,
L.
(
2023
).
Systematic dissection of sequence features affecting the binding specificity of a pioneer factor reveals binding synergy between FOXA1 and AP-1
.
bioRxiv
.
Yang
,
Y.
,
Gomez
,
N.
,
Infarinato
,
N.
,
Adam
,
R. C.
,
Sribour
,
M.
,
Baek
,
I.
,
Laurin
,
M.
and
Fuchs
,
E.
(
2023
).
The pioneer factor SOX9 competes for epigenetic factors to switch stem cell fates
.
Nat. Cell Biol.
25
,
1185
-
1195
.
Yu
,
Z.
,
Pestell
,
T. G.
,
Lisanti
,
M. P.
and
Pestell
,
R. G.
(
2012
).
Cancer stem cells
.
Int. J. Biochem. Cell Biol.
44
,
2144
-
2151
.
Young
,
J. M.
,
Whiddon
,
J. L.
,
Yao
,
Z.
,
Kasinathan
,
B.
,
Snider
,
L.
,
Geng
,
L. N.
,
Balog
,
J.
,
Tawil
,
R.
,
Van Der Maarel
,
S. M.
and
Tapscott
,
S. J.
(
2013
).
DUX4 binding to retroelements creates promoters that are active in FSHD muscle and testis
.
PLoS Genet.
9
,
e1003947
.
Zaret
,
K. S.
(
2020
).
Pioneer transcription factors initiating gene network changes
.
Annu. Rev. Genet.
54
,
367
-
385
.
Zhang
,
S.
,
Bell
,
E.
,
Zhi
,
H.
,
Brown
,
S.
,
Imran
,
S. A. M.
,
Azuara
,
V.
and
Cui
,
W.
(
2019
).
OCT4 and PAX6 determine the dual function of SOX2 in human ESCs as a key pluripotent or neural factor
.
Stem Cell Res. Ther.
10
,
122
.
Zhu
,
F.
,
Farnung
,
L.
,
Kaasinen
,
E.
,
Sahu
,
B.
,
Yin
,
Y.
,
Wei
,
B.
,
Dodonova
,
S. O.
,
Nitta
,
K. R.
,
Morgunova
,
E.
,
Taipale
,
M.
et al.
(
2018
).
The interaction landscape between transcription factors and the nucleosome
.
Nature
562
,
76
-
81
.

Competing interests

The authors declare no competing or financial interests.