Next Article in Journal
Mitochondrial Neurodegeneration: Lessons from Drosophila melanogaster Models
Next Article in Special Issue
Regulation of the SUV39H Family Methyltransferases: Insights from Fission Yeast
Previous Article in Journal
Therapeutic Potential of Phytocannabinoid Cannabigerol for Multiple Sclerosis: Modulation of Microglial Activation In Vitro and In Vivo
Previous Article in Special Issue
Exploring Genetic Interactions with Telomere Protection Gene pot1 in Fission Yeast
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Opposing Roles of FACT for Euchromatin and Heterochromatin in Yeast

Department of Chemistry, Faculty of Science, Hokkaido University, Sapporo 060-0810, Japan
*
Author to whom correspondence should be addressed.
Biomolecules 2023, 13(2), 377; https://doi.org/10.3390/biom13020377
Submission received: 23 January 2023 / Revised: 14 February 2023 / Accepted: 15 February 2023 / Published: 16 February 2023
(This article belongs to the Special Issue Yeast Models for Gene Regulation)

Abstract

:
DNA is stored in the nucleus of a cell in a folded state; however, only the necessary genetic information is extracted from the required group of genes. The key to extracting genetic information is chromatin ambivalence. Depending on the chromosomal region, chromatin is characterized into low-density “euchromatin” and high-density “heterochromatin”, with various factors being involved in its regulation. Here, we focus on chromatin regulation and gene expression by the yeast FACT complex, which functions in both euchromatin and heterochromatin. FACT is known as a histone H2A/H2B chaperone and was initially reported as an elongation factor associated with RNA polymerase II. In budding yeast, FACT activates promoter chromatin by interacting with the transcriptional activators SBF/MBF via the regulation of G1/S cell cycle genes. In fission yeast, FACT plays an important role in the formation of higher-order chromatin structures and transcriptional repression by binding to Swi6, an HP1 family protein, at heterochromatin. This FACT property, which refers to the alternate chromatin-regulation depending on the binding partner, is an interesting phenomenon. Further analysis of nucleosome regulation within heterochromatin is expected in future studies.

1. Introduction

Gene expression requires the binding of transcriptional activators that recognize and bind specific upstream activating sequences of DNA on the promoter of each gene or enhancer of distal sites [1,2,3]. In most cases, chromatin remodeling factors, histone chaperones, and histone acetyl-transferase complexes are recruited by gene-specific transcriptional activators to loosen the core promoter chromatin around the TATA-box and the transcriptional start site for the formation of a preinitiation complex with the general transcription factors, TFIIA, TFIIB, TFIID, TFIIE, TFIIF, TFIIH, and RNA polymerase II, adjacent to the +1 nucleosome [4,5,6,7,8]. The sequential scheme of this transcription initiation indicates the repressive feature of the chromatin structure itself against gene expression, and that the regulatory mechanism of relaxing and closing the chromatin structure is closely related to the regulation of gene expression [9,10]. The landscape of nucleosome occupancy is analyzed by ChIP-seq or MNase-seq, continuously revealing the genome-wide positioning of nucleosomes and profiling of gene expression [11,12,13,14]. In terms of chromatin structure, there is euchromatin with a loose chromatin structure and heterochromatin with a complex higher-order chromatin structure [15], with post-translational modifications of histone proteins playing a key role in maintaining each structure [16,17]. It has been reported that the SIR complex, Sir2, Sir3, and Sir4, is responsible for the heterochromatin formation of the MAT locus, subtelomeres, and rDNA regions depend on a DNA element called “silencer” in budding yeast [18,19]. The heterochromatin formed in this way is somewhat unique and differs from the heterochromatin in other eukaryotes, which is epigenetically formed and maintained by the post-translational modifications of histones. SIR complex-mediated silencing is assumed to be more stably maintained by DNA-binding factors. In addition, there have been no reports that FACT contributes to this SIR complex-dependent silencing of budding yeast heterochromatin so far. In fission yeast, heterochromatin is formed and maintained through high-histone H3K9 methylation and low-histone H3K4 methylation by histone H3K9 methyltransferase, histone H3K4 de-methylase, and HP1, as in mammal cells [20,21,22,23,24,25,26,27]. A histone H3K9 methyltransferase and HP1 family proteins are conserved as in other eukaryotes, and are responsible for the constitutive heterochromatin at the MAT locus, pericentromeres, and subtelomeres [21]. In addition, recent studies reported that H3K9me islands are scattered on chromosome arms [28,29]. The FACT complex has been shown to stabilize the constitutive heterochromatin by working in concert with HP1/Swi6 in fission yeast [30,31,32]. While the chromatin remodeling models of FACT for transcriptional stimulation on euchromatin have been proposed to date, a completely new mechanism of chromatin silencing by FACT that represses nucleosome dynamics on heterochromatin will become a subject of discussion.

2. FACT Plays Multifunctional Roles in Transcriptional Regulation

The FACT complex, a heterodimer of Spt16 and SSRP1, was isolated as an RNA polymerase II transcriptional elongation factor required for efficient chromatin transcription [33,34,35,36]. Unlike other chromatin remodeling factors, FACT has no ATPase domain and performs chromatin remodeling in an ATP-independent manner. Recent studies have reported that nucleosome regulation by interplay between FACT and Chd1, an ATP-dependent chromatin remodeling factor, plays an important role in RNA polymerase II transcriptional initiation at +1 nucleosome [37,38]. In addition to RNA polymerase II-mediated transcription, TFIIIC/TFIIIB may be involved in a physical interaction with the FACT complex for the RNA polymerase III transcribed tRNA genes in budding yeast [39]. The nucleosome-binding activity of human FACT is low in vitro, but through some destabilization in the contact between core histones and nucleosomal DNA, FACT starts interacting with the core region of the histone covered by DNA [40,41]. FACT is a highly conserved histone chaperone between divergent eukaryotic species (Figure 1).
In yeast, it consists of an Spt16/Pob3 heterodimer and the high-mobility group box (HMGB) protein Nhp6 [42,43,44,45]. Spt16 was originally isolated as CDC68, a gene responsible for causing G1 arrest [46]. Pob3 and Nhp6 are bipartite analogs of SSRP1 [47,48]. The intracellular roles of FACT in budding and fission yeast appear to be somewhat different. First, the copy numbers of the budding yeast NHP6 gene and the fission yeast nhp6+ gene are different: there are two copies of NHP6 in budding yeast, NHP6A and NHP6B [49], whereas there is a single copy of nhp6+ in fission yeast [50]. Second, SPT16 and POB3 are essential genes, whereas NHP6A/B are nonessential genes in budding yeast [44,51]. In fission yeast, spt16+ is an essential gene, whereas pob3+ and nhp6+ are nonessential genes [50], allowing the disruption of pob3+ and nhp6+, which are functional bipartite analogues of SSRP1 in multicellular organisms, exceptionally through eukaryotic species. Considering that the function of Spt16 is required even in the absence of Pob3 and Nhp6, it is likely that it exerts a certain chromatin regulatory role, which is specific to Spt16 in fission yeast. In suggesting a role for each FACT component, human SSRP1 has Spt16-dependent and independent roles in transcriptional regulation [52,53]. In budding yeast, Pob3 forms a stable heterodimer with Spt16 via their dimerization domains [43,48]. Biochemical studies have exhibited that Nhp6 plays an essential role in binding the Spt16/Pob3 heterodimer to the nucleosome [45], with the required amount of Nhp6 appearing to be stoichiometrically in excess to that of the Spt16/Pob3 heterodimer [54]. This might suggest that when the HMGB DNA-binding domain is fused to the FACT, as in SSRP1, it enhances nucleosome recognition for the efficient H2A/H2B dimer eviction from the nucleosome. Another FACT isoform, in which Pob3 and Nhp6 are expressed as a fusion SSRP1 protein, has also been analyzed in vivo. In budding yeast, Nhp6 is expressed from NHP6A and NHP6B, and strains in which both NHP6A and NHP6B are simultaneously disrupted showed growth defects. Under this condition, the expression of the POB3-NHP6 fusion gene was found to complement the growth defect shown by the nhp6ab∆ strain [47]. Moreover, the Pob3-Nhp6-fused FACT has been reported to be involved in nucleosome regulation, as indicated by biochemical analyses. Single or multiple HMGB modules were fused to Pob3 to mimic SSRP1 for evaluating its nucleosome-binding capacity. Human SSRP1 and a yeast Pob3-Nhp6 fusion both required free Nhp6 to support nucleosome reorganization. This result indicated that a single intrinsic DNA-binding HMGB was not sufficient for intact FACT nucleosome reorganizing activity, whereas triple HMGB modules at the C-terminus of Pob3 supported FACT activity without free Nhp6. However, this FACT variant was not efficiently released from nucleosomes, in turn exhibiting toxicity in yeast [55]. Recent cryo-EM structure analysis revealed that human Spt16 bound to histones in a subnucleosome and tethered H2A/H2B through its C-terminal acidic tail by acting as a placeholder for DNA, with no electron density being observed at the HMGB domain of SSRP1 [56]. Phosphorylation of the Spt16 C-terminal acidic tail is required for its binding to H2A/H2B in the nucleosome [57,58], suggesting the involvement of CKII [59,60]. FACT was also reported to displace H2A/H2B dimers from the nucleosome through the tandem PH domain of Spt16 and histone H3/H4-binding of the Spt16 peptidase-like domain with the help of Nhp6 [61,62,63]. Apart from its DNA-binding activity, cryo-EM analysis revealed that Nhp6 binds to both C-terminal acidic tails of Spt16 and Pob3 to unfold the FACT complex structure for the activation of efficient nucleosome reorganization [64]. These results suggested the importance of Nhp6 for chromatin remodeling [56].

3. Working Models of FACT for Nucleosome Dynamics in Fission Yeast

Various molecular models have been proposed to explain the means by which FACT transforms the nucleosome [57,64,65], but few molecular models have been proposed for how FACT regulates chromatin silencing. In the case of the fission yeast FACT, the histone H3/H4-binding activity of the peptidase-like domain at the N-terminus of Spt16, the histone H2A/H2B chaperone activity of the tandem PH domain in the central region of Spt16, the histone H3/H4-binding activity of the tandem PH domain of Pob3, and the DNA-binding activity of Nhp6 are thought to play key roles in nucleosome recognition [30,61,62,66]. Accordingly, two different models by which fission yeast FACT binds to the mononucleosome or dinucleosome is shown in Figure 2. In the case of binding to the mononucleosome, the peptidase-like domain of Spt16 and the tandem PH domain of Pob3 bind to the two histone H3/H4 dimers present in the mononucleosome via their dimerization domains, respectively (Figure 2A).
After binding of the Spt16/Pob3 heterodimer to the mononucleosome, the acidic tail at the C-terminus of the two proteins competes with the nucleosomal DNA on the surface of histone H2A/H2B. Following this competition, Nhp6 binds to the fluctuated DNA and promotes the divergence of histone H2A/H2B from DNA in the nucleosome, with the tandem PH domain of Spt16 depositing the histone H2A/H2B dimer from the octasome and transforming it to a hexasome or tetrasome [67]. Studies have already reported the histone chaperone activity of both human and yeast FACT for histone H2A/H2B [68], suggesting the induction of a transient dynamic change in chromatin regulation by a similar process of nucleosome conformational change. Meanwhile, the histone-binding properties of the peptidase-like domain of Spt16 and the tandem PH domain of Pob3 have suggested their binding to the dinucleosome (Figure 2B). In this case, the peptidase-like domain of Spt16 and the tandem PH domain of Pob3 are expected to act separately on two neighboring nucleosomes to bridge them, such as HP1; however, the mechanism by which the tandem PH domain of Spt16 acts on histone H2A/H2B in the nucleosome remains undetermined. As we currently lack any biochemical or structural data on the mechanism of action of FACT for the dinucleosome, therefore the histone H2A/H2B chaperone activity of FACT should be determined in the dinucleosome regulatory case.

4. SBF Recruits FACT for Promoter Chromatin Activation in Budding Yeast

Previous studies have revealed that FACT dynamically alters the chromatin structure, transiently evicting nucleosomes for the passage of RNA polymerase II in vivo [63]. Following nucleosome eviction and the passage of RNA polymerase II, FACT deposits nucleosomes to close the transiently loosened chromatin structure [69,70]. In addition to functioning as a transcriptional elongation factor, previous studies have demonstrated that FACT binds to the G1/S START transcription factors SBF and MBF (Figure 3A), which are analogs of mammal E2F in budding yeast. SBF enters the nucleus in late M/early G1 phase [71,72,73] and binds to the G1 gene promoters. In turn, the SBF-recruited FACT transiently evicts nucleosomes from the promoter of G1/S regulatory genes before initiation of transcription by RNA polymerase II [74,75,76]. In budding yeast, SBF acts at the START checkpoint in the G1/S phase, regulating the expression of the CLN1 and CLN2 genes [77,78,79] (Figure 3A). Cell cycle-related gene transcription is regulated by the competition between positive and negative regulators of chromatin. In early G1 phase, the activity of SBF is repressed by Whi5 after binding to the “CACGAAAA” promoter element in the UAS until it is activated to initiate transcription at the proper timing. Whi5 recruits the histone deacetyl-transferase complex Rpd3(L) to keep the promoter chromatin in a silent state [74]. Cyclin kinase Cdk1/Cln1,2 phosphorylates Whi5 during the progression of the G1 phase to remove it from SBF [80], which is then converted to its activated state [81,82]. After the removal of Whi5 and Rpd3(L) from the promoter, the SBF-recruited FACT functions to change the promoter chromatin structure [74]; however, the detailed molecular mechanism by which FACT recognizes SBF/MBF remains undetermined. The expression timing of G1 genes during G1 phase is also different, with variations observed during the transition from early G1 to late G1/S. Even though they are regulated by the same transcription factor, SBF/MBF, it is assumed that the reason for this is the differences in the chromatin structure of the promoters of each gene. In addition to the G1 cyclin gene, the chromatin structure of the homothallic switching (HO) gene promoter is regulated by SFB and FACT (Figure 3B). The HO gene on chromosome IV encodes the Ho endonuclease.
Budding yeast strains commonly used in laboratories contain a mutation that results in defective nuclease activity in vitro and in vivo [83]. Wildtype Ho endonuclease induces a double-strand break that targets a DNA element in the MAT locus, causing mating-type switching via gene conversion of the MAT decision cassette [84,85,86,87]. This phenomenon occurs asymmetrically during cell division, with expression of the HO gene in the mother cell and transcriptional repression of the HO gene in the daughter cell. This asymmetry is generated by Ash1, a component of Rpd3(L) HDAC [88,89] that is mainly expressed in daughter cells, and binds to the promoter of the HO gene, thereby strongly repressing transcription [76,90,91,92,93,94]. In addition to this asymmetric expression, the expression of HO needs to be strictly regulated in the mother cell. The promoter structure of the HO gene is relatively long and complex compared with that of common yeast genes, and consists of two sequential regions, URS1 and URS2 (Figure 3B). The combination of URS1 and URS2, approximately 1.0 kbp each, regulates cell cycle-dependent transcriptional initiation [75,95,96,97,98]. URS1 contains two binding sites for Swi5, which is expressed at the M/G1 phase boundary. Swi5 is phosphorylated by the Cdk1 kinase and is transported into the nucleus from the end of M to the early G1 phase [99,100,101]. An SBF-binding site has also been identified at the 3′ side of URS1; however, mutation of this SBF-binding sequence does not affect the expression of the HO gene, suggesting that it is not a functionally essential element [76]. The SAGA complex, Swi/Snf complex, and SRB mediator complex are then recruited onto URS1 by Swi5 to loosen the chromatin structure of URS1. In turn, SBF and FACT loosen the chromatin structure from the 3′ side of URS1 to the 5′ side of URS2 and recruit additional SBF activators for the recruitment of the SAGA complex, Swi/Snf complex, and SRB mediator complex downstream of URS2 and TATA-box [74,75,95,96]. FACT is assumed to be the factor that causes the nucleosome eviction from the 3′ side of URS1 to the 5′ side of URS2 in the sequence of chromatin conformational changes involved in this transcriptional activation.

5. Wave of Nucleosome Eviction, and the Site Where FACT Functions in HO Promoter in Budding Yeast

A schematic representation of the FACT working region in the HO promoter and the dynamic changes in chromatin along cell cycle progression is shown in Figure 4. In wildtype budding yeast strains, nucleosome eviction at URS1 is triggered by the Swi5 activator, and the recruited SAGA complex, Swi/Snf complex, and SRB mediators. FACT-induced nucleosome eviction is then triggered from downstream URS1 to upstream URS2, with the wave of nucleosome eviction being propagated downstream to URS2 and the core promoter, where the chromatin around the TATA-box is finally opened to promote the transcription of the HO gene by RNA polymerase II (Figure 4A). The nucleosomes of URS1 are quickly repositioned, presumably due to polyubiquitination and proteolysis of Swi5 [100]. However, in the FACT mutant yeast strain, nucleosomes are evicted from URS1 by the Swi5 activator and the recruited SAGA complex, Swi/Snf complex, and SRB mediators, as in wildtype, but this eviction is not propagated downstream from the URS2 to the TATA-box during cell cycle progression (Figure 4B). ChIP analysis of FACT exhibited a biased binding pattern to the upstream of URS2 of the HO gene promoter, suggesting that FACT does not bind to the promoter solely through the SBF recruitment [75]. The reason for this biased promoter-binding activity of FACT remains unclear, and there might be a characteristic chromatin structure upstream of URS2 that FACT prefers.

6. FACT-Dependent Heterochromatic Silencing in Fission Yeast

Similar to many other eukaryotes, heterochromatin in fission yeast is formed in a histone H3K9 methylation-dependent manner. The mechanism of formation of constitutive heterochromatin at centromeres, subtelomeres, and the MAT locus is very complex [21,102,103,104], with the molecular mechanism of heterochromatin formation being distinct in these three regions [105]. Histone H3K9 methylation-dependent higher-order chromatin structures cannot be stably maintained unless the various effector factors function at the correct timing. Methylation of histone H3K9 is the most important factor for heterochromatin formation. Although multicellular eukaryotes possess multiple histone H3K9 methyltransferases, in fission yeast, Clr4 is the sole source of methylase of histone H3K9 via its SET domain [106]. A recent study revealed that automethylation of Clr4 stimulates its enzymatic activity and maintains its epigenetic stability [107]. At least two recruitment mechanisms are known for Clr4 in the establishment of pericentromeric heterochromatin. One is the direct association of Clr4 with HP1/Swi6 [108], and the other is an RNAi-dependent recruitment onto heterochromatin [109,110] (Figure 5). Although heterochromatin formation and transcription of noncoding RNAs (ncRNAs) sound contradictory, HP1/Swi6 is strongly bound by Epe1, a JmjC protein [111,112]. Epe1 carries the acidic activation domain at the N-terminus and stimulates the transcription of heterochromatic ncRNAs by RNA polymerase II [112]. This transcription in the heterochromatin is assumed to be slow and suspendable and creates a scaffold retaining the nascent ncRNA on heterochromatin for RNAi-related effectors on heterochromatin [113,114,115,116].
In addition to the SET domain, Clr4 itself also has a chromodomain (CD) at its N-terminus that recognizes histone H3K9me, and following recognition exerts its self-propagation ability to methylate H3K9 on the adjacent nucleosome [25,117]. Swi6 and Chp2 are known as fission yeast HP1, which bind to H3K9me-containing nucleosomes, forming homodimers via their chromo-shadow domain (CSD) [118,119]. Heterochromatin is stably maintained by two homodimers, Swi6 and Chp2 [120,121], which attract different silencing effectors [122,123,124], with Swi6 being a more versatile HP1, potentially important for the formation and maintenance of stable heterochromatin. In addition to the different roles of the two HP1 family proteins in fission yeast, post-translational modifications of HP1 also affect HP1 heterochromatin formation. For instance, Swi6 has been reported to be phosphorylated, and mutations at the phosphorylation site were reported to disrupt heterochromatin formation [125,126,127].
Genetic analysis and ChIP-qPCR showed that fission yeast strains lacking pob3+ (pob3∆) had comparable levels of histone H3K9 methylation and Swi6 localization in the heterochromatic region to those of the wildtype strain, but with high levels of heterochromatic expression of ncRNAs. Phenotypic analysis of the pob3∆ strain indicated that heterochromatic silencing was defective in heterochromatin without a significant loss of levels of histone H3K9 methylation and HP1/Swi6-binding [30,128]. ChIP analysis of Spt16 exhibited that the binding level of Spt16 to the heterochromatic region in the pob3∆ strain was decreased to half that of the wildtype, suggesting the existence of a Pob3-independent recruitment mechanism of Spt16 onto heterochromatin. Genetic analysis also revealed that pob3∆swi6∆ double disruption exhibited an additive silencing defect compared with that shown in each of the pob3∆ and swi6∆ single-mutant strains [30]. Therefore, we assumed that the recruitment of Spt16 onto heterochromatin is partially dependent on Swi6. To test this hypothesis, we performed biochemical analysis using recombinant Spt16 and the fission yeast HP1 family, Swi6 and Chp2. We found that the peptidase-like domain of Spt16 directly binds to the dimerized chromo-shadow domain (CSD) of Swi6, but not to Chp2-CSD [30,120,129]. Although the “PxVxL/I” hydrophobic amino acid sequence of the CSD-binding motif is necessary for stable CSD-binding [130,131,132], this motif is not conserved in the peptidase-like domain of Spt16. Further Spt16-Swi6-binding experiments revealed that the binding of Spt16 and Swi6 was easily compromised by increasing the salt concentration in the binding buffer in vitro, suggesting the existence of a novel binding mode between Spt16 and Swi6.
To elucidate this binding mode between Spt16 and Swi6, we carefully compared the primary sequences of Swi6 and Chp2 in the CSD and found a difference in the β1–β2 connecting loop, which forms the protrusion in the CSD homodimer [30]. We identified a charge-biased “RKDD” cluster in the β1–β2 connecting loop of Swi6-CSD, but no such charge-biased cluster in the β1–β2 connecting loop of Chp2-CSD. Other HP1 family proteins were also examined for the presence of charge-biased clusters in the β1–β2 connecting loop, but no charge-biased clusters were found. A physical interaction between HP1c and SSRP1 for transcriptional activation on euchromatin has been reported in Drosophila melanogaster [133], but physical interactions between HP1 and Spt16 have not been reported in other species so far. The charge-biased cluster in the β1–β2 connecting loop of the Swi6-CSD might be a specific property of the formation of heterochromatin in fission yeast. These data have suggested that the “RKDD” sequence in the β1–β2 connecting loop of Swi6-CSD might function as a binding surface of the peptidase-like domain of Spt16. To this end, a recombinant Swi6-4A mutant (“RKDD” to “AAAA”) was used to test the binding activity of Spt16 in vitro. As expected, Swi6-4A lost its ability to bind to the peptidase-like domain of Spt16 [30]. Interestingly, heterochromatin was significantly disordered in Swi6-4A mutant fission yeast. Although FACT targets the β1–β2 connecting loop of Swi6, other effectors might bind to the “RKDD” sequence in the β1–β2 connecting loop of Swi6-CSD in fission yeast. This scenario requires further examination in future heterochromatin studies. To evaluate the effect of eliminating FACT in heterochromatin, a peptidase-like domain-truncated Spt16 was expressed in a pob3∆ strain. Different from the pob3∆ strain, a major decrease in the levels of histone H3K9 methylation and Swi6-binding was observed [30]. This indicated that FACT plays a critical role in the establishment and maintenance of heterochromatin.

7. Mechanism of Action of FACT on Nucleosomes within Heterochromatin for Formation and Maintenance of Heterochromatin in Fission Yeast

Our analysis revealed the molecular mechanism by which FACT is recruited onto heterochromatin for the dynamic regulation of the H2A/H2B dimer and optimal management of nucleosomes in heterochromatin [30]. The H2A.Z/H2B dimer contributes to a certain extent to this heterochromatic silencing [134]. Other groups have reported that FACT strongly suppresses histone turnover [31,32,135], and the mechanism by which FACT regulates nucleosomes in heterochromatin is an important aspect to be considered. To predict the means by which FACT regulates heterochromatic nucleosomes in fission yeast, we proposed a hypothetical model, as shown in Figure 6.
Within the heterochromatin, two nucleosomes are bridged by the HP1/Swi6 heterodimer via histone H3K9me (Figure 6A). Structural analysis revealed that it is not sufficient to enumerate the dinucleosome with a HP1 homodimer alone, and the linker DNA is exposed to allow the access of silencing effectors [136]. In fission yeast, the Swi6-CSD homodimer physically binds to the peptidase-like domain of Spt16 to recruit FACT onto the heterochromatin [30]. Concomitantly, the tandem PH domain of Pob3 binds to the H3/H4 dimer in a heterochromatic nucleosome (Figure 6B). After the recruitment of FACT by Swi6, the peptidase-like domain of Spt16, which recognized Swi6-CSD, shifts its scaffold to histone H3/H4 in the nucleosome [30]. This scaffold shift can be divided into two modes: binding to the dinucleosome as a bridge (Figure 6C) and binding to the mononucleosome (Figure 6D), shown as euchromatic nucleosome regulation in Figure 2. In the heterochromatic dinucleosome bridging model, FACT helps Swi6 enhance nucleosome condensation and the formation of a higher-order chromatin structure (Figure 6C). In the heterochromatic mononucleosome-binding model, the peptidase-like domain of Spt16 and the tandem PH domain of Pob3 clip the histone H3/H4 tetramer in the heterochromatic nucleosome to stably tether FACT on a single nucleosome (Figure 6D). Under the stable FACT-nucleosome binding, deposition and reposition of the histone H2A/H2B dimer occur via the chaperon activity of the Spt16 tandem PH domain [61]. Histone H2A/H2B ChIP analysis in the pob3∆ strain indicated that reposition of the H2A/H2B dimer was dependent on the recognition of the stable H3/H4 dimer in nucleosomes by the tandem PH domain of Pob3 [30].

8. Summary and Perspective

FACT is an important transcription stimulating factor that transiently relaxes the chromatin structure through its histone H2A/H2B chaperone activity. The transcription-activating properties of FACT are especially demonstrated in START, the checkpoint of the G1/S phase, through binding to SBF/MBF in budding yeast [74,75,76]. However, the mechanism by which FACT contributes to START activation by the mammalian E2F family remains unclear. In addition to transcriptional activation in budding yeast, studies using fission yeast have shown that FACT contributes to the formation and maintenance of histone H3K9-mediated heterochromatin [30]. In particular, the molecular mechanism by which FACT is recruited onto heterochromatin was previously analyzed in detail by the authors, and the means by which FACT regulates nucleosomes in heterochromatin will be a subject of future research.

Author Contributions

Conceptualization, S.T.; Writing—Original Draft Preparation, S.T.; Writing—Review and Editing, S.T. and Y.M.; Figure Preparation, S.T.; Funding Acquisition, S.T. and Y.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by KAKENHI (17K07496) from JSPS, MEXT KAKENHI (221S0002; A Grant-in-Aid for Scientific Research on Innovative Areas “Transcription Cycle”) from the Ministry of Education, Culture, Sports, Science, and Technology of Japan to S.T., and a Grant-in-Aid for Scientific Research on Priority Areas from the Ministry of Education, Culture, Sports, Science, and Technology of Japan (19H05742) and a Grant-in-Aid for Scientific Research (A) from JSPS (19H00973) to Y.M.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We thank our laboratory members for many helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zabidi, M.A.; Stark, A. Regulatory Enhancer-Core-Promoter Communication via Transcription Factors and Cofactors. Trends Genet. 2016, 32, 801–814. [Google Scholar] [CrossRef]
  2. Malik, S.; Roeder, R.G. The metazoan Mediator co-activator complex as an integrative hub for transcriptional regulation. Nat. Rev. Genet. 2010, 11, 761–772. [Google Scholar] [CrossRef] [Green Version]
  3. Roy, A.L.; Sen, R.; Roeder, R.G. Enhancer-promoter communication and transcriptional regulation of Igh. Trends Immunol. 2011, 32, 532–539. [Google Scholar] [CrossRef] [Green Version]
  4. Wang, H.; Schilbach, S.; Ninov, M.; Urlaub, H.; Cramer, P. Structures of transcription preinitiation complex engaged with the +1 nucleosome. Nat. Struct. Mol. Biol. 2023, 30, 226–232. [Google Scholar] [CrossRef]
  5. Schier, A.C.; Taatjes, D.J. Structure and mechanism of the RNA polymerase II transcription machinery. Genes Dev. 2020, 34, 465–488. [Google Scholar] [CrossRef] [Green Version]
  6. Hahn, S. Structure and mechanism of the RNA polymerase II transcription machinery. Nat. Struct. Mol. Biol. 2004, 11, 394–403. [Google Scholar] [CrossRef] [Green Version]
  7. Svaren, J.; Horz, W. Transcription factors vs. nucleosomes: Regulation of the PHO5 promoter in yeast. Trends Biochem. Sci. 1997, 22, 93–97. [Google Scholar] [CrossRef]
  8. Roeder, R.G. Role of general and gene-specific cofactors in the regulation of eukaryotic transcription. Cold Spring Harb. Symp. Quant. Biol. 1998, 63, 201–218. [Google Scholar] [CrossRef]
  9. Lee, C.K.; Shibata, Y.; Rao, B.; Strahl, B.D.; Lieb, J.D. Evidence for nucleosome depletion at active regulatory regions genome-wide. Nat. Genet. 2004, 36, 900–905. [Google Scholar] [CrossRef]
  10. Wu, C.; Wong, Y.C.; Elgin, S.C. The chromatin structure of specific genes: II. Disruption of chromatin structure during gene activity. Cell 1979, 16, 807–814. [Google Scholar] [CrossRef]
  11. Jiang, C.; Pugh, B.F. Nucleosome positioning and gene regulation: Advances through genomics. Nat. Rev. Genet. 2009, 10, 161–172. [Google Scholar] [CrossRef] [Green Version]
  12. Chereji, R.V.; Ocampo, J.; Clark, D.J. MNase-Sensitive Complexes in Yeast: Nucleosomes and Non-histone Barriers. Mol. Cell 2017, 65, 565–577.e563. [Google Scholar] [CrossRef] [Green Version]
  13. Givens, R.M.; Lai, W.K.; Rizzo, J.M.; Bard, J.E.; Mieczkowski, P.A.; Leatherwood, J.; Huberman, J.A.; Buck, M.J. Chromatin architectures at fission yeast transcriptional promoters and replication origins. Nucleic Acids Res. 2012, 40, 7176–7189. [Google Scholar] [CrossRef]
  14. Lantermann, A.B.; Straub, T.; Stralfors, A.; Yuan, G.C.; Ekwall, K.; Korber, P. Schizosaccharomyces pombe genome-wide nucleosome mapping reveals positioning mechanisms distinct from those of Saccharomyces cerevisiae. Nat. Struct. Mol. Biol. 2010, 17, 251–257. [Google Scholar] [CrossRef]
  15. Morrison, O.; Thakur, J. Molecular Complexes at Euchromatin, Heterochromatin and Centromeric Chromatin. Int. J. Mol. Sci. 2021, 22, 6922. [Google Scholar] [CrossRef]
  16. Shahid, Z.; Simpson, B.; Miao, K.H.; Singh, G. Genetics, Histone Code; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  17. Jenuwein, T.; Allis, C.D. Translating the histone code. Science 2001, 293, 1074–1080. [Google Scholar] [CrossRef] [Green Version]
  18. Hickman, M.A.; Froyd, C.A.; Rusche, L.N. Reinventing heterochromatin in budding yeasts: Sir2 and the origin recognition complex take center stage. Eukaryot. Cell 2011, 10, 1183–1192. [Google Scholar] [CrossRef] [Green Version]
  19. Guarente, L. Diverse and dynamic functions of the Sir silencing complex. Nat. Genet. 1999, 23, 281–285. [Google Scholar] [CrossRef]
  20. Bi, X. Heterochromatin structure: Lessons from the budding yeast. IUBMB Life 2014, 66, 657–666. [Google Scholar] [CrossRef]
  21. Grewal, S.I.; Jia, S. Heterochromatin revisited. Nat. Rev. Genet. 2007, 8, 35–46. [Google Scholar] [CrossRef]
  22. Noma, K.; Allis, C.D.; Grewal, S.I. Transitions in distinct histone H3 methylation patterns at the heterochromatin domain boundaries. Science 2001, 293, 1150–1155. [Google Scholar] [CrossRef]
  23. Nakayama, J.; Rice, J.C.; Strahl, B.D.; Allis, C.D.; Grewal, S.I. Role of histone H3 lysine 9 methylation in epigenetic control of heterochromatin assembly. Science 2001, 292, 110–113. [Google Scholar] [CrossRef] [Green Version]
  24. Stewart, M.D.; Li, J.; Wong, J. Relationship between histone H3 lysine 9 methylation, transcription repression, and heterochromatin protein 1 recruitment. Mol. Cell Biol. 2005, 25, 2525–2538. [Google Scholar] [CrossRef] [Green Version]
  25. Cutter DiPiazza, A.R.; Taneja, N.; Dhakshnamoorthy, J.; Wheeler, D.; Holla, S.; Grewal, S.I.S. Spreading and epigenetic inheritance of heterochromatin require a critical density of histone H3 lysine 9 tri-methylation. Proc. Natl. Acad. Sci. USA 2021, 118, e2100699118. [Google Scholar] [CrossRef]
  26. Peters, A.H.; Mermoud, J.E.; O’Carroll, D.; Pagani, M.; Schweizer, D.; Brockdorff, N.; Jenuwein, T. Histone H3 lysine 9 methylation is an epigenetic imprint of facultative heterochromatin. Nat. Genet. 2002, 30, 77–80. [Google Scholar] [CrossRef]
  27. Li, F.; Huarte, M.; Zaratiegui, M.; Vaughn, M.W.; Shi, Y.; Martienssen, R.; Cande, W.Z. Lid2 is required for coordinating H3K4 and H3K9 methylation of heterochromatin and euchromatin. Cell 2008, 135, 272–283. [Google Scholar] [CrossRef] [Green Version]
  28. Zofall, M.; Yamanaka, S.; Reyes-Turcu, F.E.; Zhang, K.; Rubin, C.; Grewal, S.I. RNA elimination machinery targeting meiotic mRNAs promotes facultative heterochromatin formation. Science 2012, 335, 96–100. [Google Scholar] [CrossRef]
  29. Cam, H.P.; Sugiyama, T.; Chen, E.S.; Chen, X.; FitzGerald, P.C.; Grewal, S.I. Comprehensive analysis of heterochromatin- and RNAi-mediated epigenetic control of the fission yeast genome. Nat. Genet. 2005, 37, 809–819. [Google Scholar] [CrossRef]
  30. Takahata, S.; Chida, S.; Ohnuma, A.; Ando, M.; Asanuma, T.; Murakami, Y. Two secured FACT recruitment mechanisms are essential for heterochromatin maintenance. Cell Rep. 2021, 36, 109540. [Google Scholar] [CrossRef]
  31. Holla, S.; Dhakshnamoorthy, J.; Folco, H.D.; Balachandran, V.; Xiao, H.; Sun, L.L.; Wheeler, D.; Zofall, M.; Grewal, S.I.S. Positioning Heterochromatin at the Nuclear Periphery Suppresses Histone Turnover to Promote Epigenetic Inheritance. Cell 2020, 180, 150–164.e115. [Google Scholar] [CrossRef]
  32. Murawska, M.; Greenstein, R.A.; Schauer, T.; Olsen, K.C.F.; Ng, H.; Ladurner, A.G.; Al-Sady, B.; Braun, S. The histone chaperone FACT facilitates heterochromatin spreading by regulating histone turnover and H3K9 methylation states. Cell Rep. 2021, 37, 109944. [Google Scholar] [CrossRef]
  33. LeRoy, G.; Orphanides, G.; Lane, W.S.; Reinberg, D. Requirement of RSF and FACT for transcription of chromatin templates in vitro. Science 1998, 282, 1900–1904. [Google Scholar] [CrossRef]
  34. Orphanides, G.; LeRoy, G.; Chang, C.H.; Luse, D.S.; Reinberg, D. FACT, a factor that facilitates transcript elongation through nucleosomes. Cell 1998, 92, 105–116. [Google Scholar] [CrossRef] [Green Version]
  35. Orphanides, G.; Wu, W.H.; Lane, W.S.; Hampsey, M.; Reinberg, D. The chromatin-specific transcription elongation factor FACT comprises human SPT16 and SSRP1 proteins. Nature 1999, 400, 284–288. [Google Scholar] [CrossRef]
  36. Belotserkovskaya, R.; Reinberg, D. Facts about FACT and transcript elongation through chromatin. Curr. Opin. Genet. Dev. 2004, 14, 139–146. [Google Scholar] [CrossRef]
  37. Farnung, L.; Ochmann, M.; Engeholm, M.; Cramer, P. Structural basis of nucleosome transcription mediated by Chd1 and FACT. Nat. Struct. Mol. Biol. 2021, 28, 382–387. [Google Scholar] [CrossRef]
  38. Jeronimo, C.; Angel, A.; Nguyen, V.Q.; Kim, J.M.; Poitras, C.; Lambert, E.; Collin, P.; Mellor, J.; Wu, C.; Robert, F. FACT is recruited to the +1 nucleosome of transcribed genes and spreads in a Chd1-dependent manner. Mol. Cell 2021, 81, 3542–3559.e3511. [Google Scholar] [CrossRef]
  39. Shukla, A.; Bhalla, P.; Potdar, P.K.; Jampala, P.; Bhargava, P. Transcription-dependent enrichment of the yeast FACT complex influences nucleosome dynamics on the RNA polymerase III-transcribed genes. RNA 2020, 27, 273–290. [Google Scholar] [CrossRef]
  40. Valieva, M.E.; Gerasimova, N.S.; Kudryashova, K.S.; Kozlova, A.L.; Kirpichnikov, M.P.; Hu, Q.; Botuyan, M.V.; Mer, G.; Feofanov, A.V.; Studitsky, V.M. Stabilization of Nucleosomes by Histone Tails and by FACT Revealed by spFRET Microscopy. Cancers 2017, 9, 3. [Google Scholar] [CrossRef] [Green Version]
  41. Tsunaka, Y.; Fujiwara, Y.; Oyama, T.; Hirose, S.; Morikawa, K. Integrated molecular mechanism directing nucleosome reorganization by human FACT. Genes Dev. 2016, 30, 673–686. [Google Scholar] [CrossRef] [Green Version]
  42. Wittmeyer, J.; Formosa, T. The Saccharomyces cerevisiae DNA polymerase alpha catalytic subunit interacts with Cdc68/Spt16 and with Pob3, a protein similar to an HMG1-like protein. Mol. Cell Biol. 1997, 17, 4178–4190. [Google Scholar] [CrossRef] [Green Version]
  43. Brewster, N.K.; Johnston, G.C.; Singer, R.A. Characterization of the CP complex, an abundant dimer of Cdc68 and Pob3 proteins that regulates yeast transcriptional activation and chromatin repression. J. Biol. Chem. 1998, 273, 21972–21979. [Google Scholar] [CrossRef] [Green Version]
  44. Wittmeyer, J.; Joss, L.; Formosa, T. Spt16 and Pob3 of Saccharomyces cerevisiae form an essential, abundant heterodimer that is nuclear, chromatin-associated, and copurifies with DNA polymerase alpha. Biochemistry 1999, 38, 8961–8971. [Google Scholar] [CrossRef]
  45. Formosa, T.; Eriksson, P.; Wittmeyer, J.; Ginn, J.; Yu, Y.; Stillman, D.J. Spt16-Pob3 and the HMG protein Nhp6 combine to form the nucleosome-binding factor SPN. EMBO J. 2001, 20, 3506–3517. [Google Scholar] [CrossRef] [PubMed]
  46. Rowley, A.; Singer, R.A.; Johnston, G.C. CDC68, a yeast gene that affects regulation of cell proliferation and transcription, encodes a protein with a highly acidic carboxyl terminus. Mol. Cell Biol. 1991, 11, 5718–5726. [Google Scholar] [CrossRef] [PubMed]
  47. Brewster, N.K.; Johnston, G.C.; Singer, R.A. A bipartite yeast SSRP1 analog comprised of Pob3 and Nhp6 proteins modulates transcription. Mol. Cell Biol. 2001, 21, 3491–3502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Formosa, T. FACT and the reorganized nucleosome. Mol. Biosyst. 2008, 4, 1085–1093. [Google Scholar] [CrossRef] [PubMed]
  49. Costigan, C.; Kolodrubetz, D.; Snyder, M. NHP6A and NHP6B, which encode HMG1-like proteins, are candidates for downstream components of the yeast SLT2 mitogen-activated protein kinase pathway. Mol. Cell Biol. 1994, 14, 2391–2403. [Google Scholar] [CrossRef] [PubMed]
  50. Kim, D.U.; Hayles, J.; Kim, D.; Wood, V.; Park, H.O.; Won, M.; Yoo, H.S.; Duhig, T.; Nam, M.; Palmer, G.; et al. Analysis of a genome-wide set of gene deletions in the fission yeast Schizosaccharomyces pombe. Nat. Biotechnol. 2010, 28, 617–623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Kolodrubetz, D.; Kruppa, M.; Burgum, A. Gene dosage affects the expression of the duplicated NHP6 genes of Saccharomyces cerevisiae. Gene 2001, 272, 93–101. [Google Scholar] [CrossRef]
  52. Li, Y.; Zeng, S.X.; Landais, I.; Lu, H. Human SSRP1 has Spt16-dependent and -independent roles in gene transcription. J. Biol. Chem. 2007, 282, 6936–6945. [Google Scholar] [CrossRef] [Green Version]
  53. Zhang, W.; Zeng, F.; Liu, Y.; Shao, C.; Li, S.; Lv, H.; Shi, Y.; Niu, L.; Teng, M.; Li, X. Crystal Structure of Human SSRP1 Middle Domain Reveals a Role in DNA Binding. Sci. Rep. 2015, 5, 18688. [Google Scholar] [CrossRef] [Green Version]
  54. Ruone, S.; Rhoades, A.R.; Formosa, T. Multiple Nhp6 molecules are required to recruit Spt16-Pob3 to form yFACT complexes and to reorganize nucleosomes. J. Biol. Chem. 2003, 278, 45288–45295. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. McCullough, L.L.; Connell, Z.; Xin, H.; Studitsky, V.M.; Feofanov, A.V.; Valieva, M.E.; Formosa, T. Functional roles of the DNA-binding HMGB domain in the histone chaperone FACT in nucleosome reorganization. J. Biol. Chem. 2018, 293, 6121–6133. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Stillman, D.J. Nhp6: A small but powerful effector of chromatin structure in Saccharomyces cerevisiae. Biochim. Biophys. Acta 2010, 1799, 175–180. [Google Scholar] [CrossRef] [Green Version]
  57. Mayanagi, K.; Saikusa, K.; Miyazaki, N.; Akashi, S.; Iwasaki, K.; Nishimura, Y.; Morikawa, K.; Tsunaka, Y. Structural visualization of key steps in nucleosome reorganization by human FACT. Sci. Rep. 2019, 9, 10183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Zheng, S.; Crickard, J.B.; Srikanth, A.; Reese, J.C. A highly conserved region within H2B is important for FACT to act on nucleosomes. Mol. Cell Biol. 2014, 34, 303–314. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Michl-Holzinger, P.; Obermeyer, S.; Markusch, H.; Pfab, A.; Ettner, A.; Bruckmann, A.; Babl, S.; Langst, G.; Schwartz, U.; Tvardovskiy, A.; et al. Phosphorylation of the FACT histone chaperone subunit SPT16 affects chromatin at RNA polymerase II transcriptional start sites in Arabidopsis. Nucleic Acids Res. 2022, 50, 5014–5028. [Google Scholar] [CrossRef]
  60. Li, Y.; Keller, D.M.; Scott, J.D.; Lu, H. CK2 phosphorylates SSRP1 and inhibits its DNA-binding activity. J. Biol. Chem. 2005, 280, 11869–11875. [Google Scholar] [CrossRef] [Green Version]
  61. Hondele, M.; Stuwe, T.; Hassler, M.; Halbach, F.; Bowman, A.; Zhang, E.T.; Nijmeijer, B.; Kotthoff, C.; Rybin, V.; Amlacher, S.; et al. Structural basis of histone H2A-H2B recognition by the essential chaperone FACT. Nature 2013, 499, 111–114. [Google Scholar] [CrossRef]
  62. Stuwe, T.; Hothorn, M.; Lejeune, E.; Rybin, V.; Bortfeld, M.; Scheffzek, K.; Ladurner, A.G. The FACT Spt16 “peptidase” domain is a histone H3-H4 binding module. Proc. Natl. Acad. Sci. USA 2008, 105, 8884–8889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Xin, H.; Takahata, S.; Blanksma, M.; McCullough, L.; Stillman, D.J.; Formosa, T. yFACT induces global accessibility of nucleosomal DNA without H2A-H2B displacement. Mol. Cell 2009, 35, 365–376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Sivkina, A.L.; Karlova, M.G.; Valieva, M.E.; McCullough, L.L.; Formosa, T.; Shaytan, A.K.; Feofanov, A.V.; Kirpichnikov, M.P.; Sokolova, O.S.; Studitsky, V.M. Electron microscopy analysis of ATP-independent nucleosome unfolding by FACT. Commun. Biol. 2022, 5, 2. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, P.; Yang, W.; Zhao, S.; Nashun, B. Regulation of chromatin structure and function: Insights into the histone chaperone FACT. Cell Cycle 2021, 20, 465–479. [Google Scholar] [CrossRef]
  66. Zunder, R.M.; Antczak, A.J.; Berger, J.M.; Rine, J. Two surfaces on the histone chaperone Rtt106 mediate histone binding, replication, and silencing. Proc. Natl. Acad. Sci. USA 2012, 109, E144–E153. [Google Scholar] [CrossRef] [Green Version]
  67. Valieva, M.E.; Armeev, G.A.; Kudryashova, K.S.; Gerasimova, N.S.; Shaytan, A.K.; Kulaeva, O.I.; McCullough, L.L.; Formosa, T.; Georgiev, P.G.; Kirpichnikov, M.P.; et al. Large-scale ATP-independent nucleosome unfolding by a histone chaperone. Nat. Struct. Mol. Biol. 2016, 23, 1111–1116. [Google Scholar] [CrossRef] [Green Version]
  68. Winkler, D.D.; Luger, K. The histone chaperone FACT: Structural insights and mechanisms for nucleosome reorganization. J. Biol. Chem. 2011, 286, 18369–18374. [Google Scholar] [CrossRef] [Green Version]
  69. Jamai, A.; Puglisi, A.; Strubin, M. Histone chaperone spt16 promotes redeposition of the original h3-h4 histones evicted by elongating RNA polymerase. Mol. Cell 2009, 35, 377–383. [Google Scholar] [CrossRef]
  70. Voth, W.P.; Takahata, S.; Nishikawa, J.L.; Metcalfe, B.M.; Naar, A.M.; Stillman, D.J. A role for FACT in repopulation of nucleosomes at inducible genes. PLoS ONE 2014, 9, e84092. [Google Scholar] [CrossRef] [Green Version]
  71. Sidorova, J.M.; Mikesell, G.E.; Breeden, L.L. Cell cycle-regulated phosphorylation of Swi6 controls its nuclear localization. Mol. Biol. Cell 1995, 6, 1641–1658. [Google Scholar] [CrossRef] [Green Version]
  72. Moll, T.; Dirick, L.; Auer, H.; Bonkovsky, J.; Nasmyth, K. SWI6 is a regulatory subunit of two different cell cycle START-dependent transcription factors in Saccharomyces cerevisiae. J. Cell Sci. Suppl. 1992, 16, 87–96. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Bertoli, C.; Skotheim, J.M.; de Bruin, R.A. Control of cell cycle transcription during G1 and S phases. Nat. Rev. Mol. Cell Biol. 2013, 14, 518–528. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Takahata, S.; Yu, Y.; Stillman, D.J. The E2F functional analogue SBF recruits the Rpd3(L) HDAC, via Whi5 and Stb1, and the FACT chromatin reorganizer, to yeast G1 cyclin promoters. EMBO J. 2009, 28, 3378–3389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Takahata, S.; Yu, Y.; Stillman, D.J. FACT and Asf1 regulate nucleosome dynamics and coactivator binding at the HO promoter. Mol. Cell 2009, 34, 405–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Takahata, S.; Yu, Y.; Stillman, D.J. Repressive chromatin affects factor binding at yeast HO (homothallic switching) promoter. J. Biol. Chem. 2011, 286, 34809–34819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Costanzo, M.; Nishikawa, J.L.; Tang, X.; Millman, J.S.; Schub, O.; Breitkreuz, K.; Dewar, D.; Rupes, I.; Andrews, B.; Tyers, M. CDK activity antagonizes Whi5, an inhibitor of G1/S transcription in yeast. Cell 2004, 117, 899–913. [Google Scholar] [CrossRef] [Green Version]
  78. Wittenberg, C.; Reed, S.I. Cell cycle-dependent transcription in yeast: Promoters, transcription factors, and transcriptomes. Oncogene 2005, 24, 2746–2755. [Google Scholar] [CrossRef] [Green Version]
  79. Dirick, L.; Moll, T.; Auer, H.; Nasmyth, K. A central role for SWI6 in modulating cell cycle Start-specific transcription in yeast. Nature 1992, 357, 508–513. [Google Scholar] [CrossRef]
  80. Hasan, M.M.; Brocca, S.; Sacco, E.; Spinelli, M.; Papaleo, E.; Lambrughi, M.; Alberghina, L.; Vanoni, M. A comparative study of Whi5 and retinoblastoma proteins: From sequence and structure analysis to intracellular networks. Front. Physiol. 2013, 4, 315. [Google Scholar] [CrossRef] [Green Version]
  81. Wagner, M.V.; Smolka, M.B.; de Bruin, R.A.; Zhou, H.; Wittenberg, C.; Dowdy, S.F. Whi5 regulation by site specific CDK-phosphorylation in Saccharomyces cerevisiae. PLoS ONE 2009, 4, e4300. [Google Scholar] [CrossRef] [Green Version]
  82. de Bruin, R.A.; McDonald, W.H.; Kalashnikova, T.I.; Yates, J., 3rd; Wittenberg, C. Cln3 activates G1-specific transcription via phosphorylation of the SBF bound repressor Whi5. Cell 2004, 117, 887–898. [Google Scholar] [CrossRef] [Green Version]
  83. Nickoloff, J.A.; Chen, E.Y.; Heffron, F. A 24-base-pair DNA sequence from the MAT locus stimulates intergenic recombination in yeast. Proc. Natl. Acad. Sci. USA 1986, 83, 7831–7835. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Kostriken, R.; Strathern, J.N.; Klar, A.J.; Hicks, J.B.; Heffron, F. A site-specific endonuclease essential for mating-type switching in Saccharomyces cerevisiae. Cell 1983, 35, 167–174. [Google Scholar] [CrossRef] [PubMed]
  85. Nasmyth, K. Molecular analysis of a cell lineage. Nature 1983, 302, 670–676. [Google Scholar] [CrossRef]
  86. Klar, A.J.; Hicks, J.B.; Strathern, J.N. Directionality of yeast mating-type interconversion. Cell 1982, 28, 551–561. [Google Scholar] [CrossRef]
  87. Strathern, J.N.; Klar, A.J.; Hicks, J.B.; Abraham, J.A.; Ivy, J.M.; Nasmyth, K.A.; McGill, C. Homothallic switching of yeast mating type cassettes is initiated by a double-stranded cut in the MAT locus. Cell 1982, 31, 183–192. [Google Scholar] [CrossRef] [PubMed]
  88. Carrozza, M.J.; Florens, L.; Swanson, S.K.; Shia, W.J.; Anderson, S.; Yates, J.; Washburn, M.P.; Workman, J.L. Stable incorporation of sequence specific repressors Ash1 and Ume6 into the Rpd3L complex. Biochim. Biophys. Acta 2005, 1731, 77–87. [Google Scholar] [CrossRef] [PubMed]
  89. Sardiu, M.E.; Gilmore, J.M.; Carrozza, M.J.; Li, B.; Workman, J.L.; Florens, L.; Washburn, M.P. Determining protein complex connectivity using a probabilistic deletion network derived from quantitative proteomics. PLoS ONE 2009, 4, e7310. [Google Scholar] [CrossRef]
  90. Bohl, F.; Kruse, C.; Frank, A.; Ferring, D.; Jansen, R.P. She2p, a novel RNA-binding protein tethers ASH1 mRNA to the Myo4p myosin motor via She3p. EMBO J. 2000, 19, 5514–5524. [Google Scholar] [CrossRef] [Green Version]
  91. Darzacq, X.; Powrie, E.; Gu, W.; Singer, R.H.; Zenklusen, D. RNA asymmetric distribution and daughter/mother differentiation in yeast. Curr. Opin. Microbiol. 2003, 6, 614–620. [Google Scholar] [CrossRef] [Green Version]
  92. Yu, Y.; Yarrington, R.M.; Stillman, D.J. FACT and Ash1 promote long-range and bidirectional nucleosome eviction at the HO promoter. Nucleic Acids Res. 2020, 48, 10877–10889. [Google Scholar] [CrossRef] [PubMed]
  93. McBride, H.J.; Sil, A.; Measday, V.; Yu, Y.; Moffat, J.; Maxon, M.E.; Herskowitz, I.; Andrews, B.; Stillman, D.J. The protein kinase Pho85 is required for asymmetric accumulation of the Ash1 protein in Saccharomyces cerevisiae. Mol. Microbiol. 2001, 42, 345–353. [Google Scholar] [CrossRef] [PubMed]
  94. Cosma, M.P. Daughter-specific repression of Saccharomyces cerevisiae HO: Ash1 is the commander. EMBO Rep. 2004, 5, 953–957. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Cosma, M.P.; Tanaka, T.; Nasmyth, K. Ordered recruitment of transcription and chromatin remodeling factors to a cell cycle- and developmentally regulated promoter. Cell 1999, 97, 299–311. [Google Scholar] [CrossRef] [Green Version]
  96. Cosma, M.P.; Panizza, S.; Nasmyth, K. Cdk1 triggers association of RNA polymerase to cell cycle promoters only after recruitment of the mediator by SBF. Mol. Cell 2001, 7, 1213–1220. [Google Scholar] [CrossRef]
  97. Stillman, D.J.; Bankier, A.T.; Seddon, A.; Groenhout, E.G.; Nasmyth, K.A. Characterization of a transcription factor involved in mother cell specific transcription of the yeast HO gene. EMBO J. 1988, 7, 485–494. [Google Scholar] [CrossRef]
  98. Stillman, D.J. Dancing the cell cycle two-step: Regulation of yeast G1-cell-cycle genes by chromatin structure. Trends Biochem. Sci. 2013, 38, 467–475. [Google Scholar] [CrossRef] [Green Version]
  99. Moll, T.; Tebb, G.; Surana, U.; Robitsch, H.; Nasmyth, K. The role of phosphorylation and the CDC28 protein kinase in cell cycle-regulated nuclear import of the S. cerevisiae transcription factor SWI5. Cell 1991, 66, 743–758. [Google Scholar] [CrossRef]
  100. Kishi, T.; Ikeda, A.; Koyama, N.; Fukada, J.; Nagao, R. A refined two-hybrid system reveals that SCF(Cdc4)-dependent degradation of Swi5 contributes to the regulatory mechanism of S-phase entry. Proc. Natl. Acad. Sci. USA 2008, 105, 14497–14502. [Google Scholar] [CrossRef] [Green Version]
  101. Dohrmann, P.R.; Butler, G.; Tamai, K.; Dorland, S.; Greene, J.R.; Thiele, D.J.; Stillman, D.J. Parallel pathways of gene regulation: Homologous regulators SWI5 and ACE2 differentially control transcription of HO and chitinase. Genes Dev. 1992, 6, 93–104. [Google Scholar] [CrossRef] [Green Version]
  102. Allshire, R.C.; Madhani, H.D. Ten principles of heterochromatin formation and function. Nat. Rev. Mol. Cell Biol. 2018, 19, 229–244. [Google Scholar] [CrossRef]
  103. Martienssen, R.; Moazed, D. RNAi and heterochromatin assembly. Cold Spring Harb. Perspect. Biol. 2015, 7, a019323. [Google Scholar] [CrossRef] [PubMed]
  104. Mizuguchi, T.; Barrowman, J.; Grewal, S.I. Chromosome domain architecture and dynamic organization of the fission yeast genome. FEBS Lett. 2015, 589, 2975–2986. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Oh, J.; Yeom, S.; Park, J.; Lee, J.S. The regional sequestration of heterochromatin structural proteins is critical to form and maintain silent chromatin. Epigenetics Chromatin 2022, 15, 5. [Google Scholar] [CrossRef]
  106. Padeken, J.; Methot, S.P.; Gasser, S.M. Establishment of H3K9-methylated heterochromatin and its functions in tissue differentiation and maintenance. Nat. Rev. Mol. Cell Biol. 2022, 23, 623–640. [Google Scholar] [CrossRef]
  107. Iglesias, N.; Currie, M.A.; Jih, G.; Paulo, J.A.; Siuti, N.; Kalocsay, M.; Gygi, S.P.; Moazed, D. Automethylation-induced conformational switch in Clr4 (Suv39h) maintains epigenetic stability. Nature 2018, 560, 504–508. [Google Scholar] [CrossRef] [PubMed]
  108. Haldar, S.; Saini, A.; Nanda, J.S.; Saini, S.; Singh, J. Role of Swi6/HP1 self-association-mediated recruitment of Clr4/Suv39 in establishment and maintenance of heterochromatin in fission yeast. J. Biol. Chem. 2011, 286, 9308–9320. [Google Scholar] [CrossRef] [Green Version]
  109. Bayne, E.H.; White, S.A.; Kagansky, A.; Bijos, D.A.; Sanchez-Pulido, L.; Hoe, K.L.; Kim, D.U.; Park, H.O.; Ponting, C.P.; Rappsilber, J.; et al. Stc1: A critical link between RNAi and chromatin modification required for heterochromatin integrity. Cell 2010, 140, 666–677. [Google Scholar] [CrossRef] [Green Version]
  110. He, C.; Pillai, S.S.; Taglini, F.; Li, F.; Ruan, K.; Zhang, J.; Wu, J.; Shi, Y.; Bayne, E.H. Structural analysis of Stc1 provides insights into the coupling of RNAi and chromatin modification. Proc. Natl. Acad. Sci. USA 2013, 110, E1879–E1888. [Google Scholar] [CrossRef] [Green Version]
  111. Ayoub, N.; Noma, K.; Isaac, S.; Kahan, T.; Grewal, S.I.; Cohen, A. A novel jmjC domain protein modulates heterochromatization in fission yeast. Mol. Cell Biol. 2003, 23, 4356–4370. [Google Scholar] [CrossRef] [Green Version]
  112. Zofall, M.; Grewal, S.I. Swi6/HP1 recruits a JmjC domain protein to facilitate transcription of heterochromatic repeats. Mol. Cell 2006, 22, 681–692. [Google Scholar] [CrossRef] [PubMed]
  113. Verdel, A.; Moazed, D. RNAi-directed assembly of heterochromatin in fission yeast. FEBS Lett. 2005, 579, 5872–5878. [Google Scholar] [CrossRef] [PubMed]
  114. Zofall, M.; Grewal, S.I. RNAi-mediated heterochromatin assembly in fission yeast. Cold Spring Harb. Symp. Quant. Biol. 2006, 71, 487–496. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Kato, H.; Goto, D.B.; Martienssen, R.A.; Urano, T.; Furukawa, K.; Murakami, Y. RNA polymerase II is required for RNAi-dependent heterochromatin assembly. Science 2005, 309, 467–469. [Google Scholar] [CrossRef]
  116. Sorida, M.; Hirauchi, T.; Ishizaki, H.; Kaito, W.; Shimada, A.; Mori, C.; Chikashige, Y.; Hiraoka, Y.; Suzuki, Y.; Ohkawa, Y.; et al. Regulation of ectopic heterochromatin-mediated epigenetic diversification by the JmjC family protein Epe1. PLoS Genet. 2019, 15, e1008129. [Google Scholar] [CrossRef] [Green Version]
  117. Zofall, M.; Sandhu, R.; Holla, S.; Wheeler, D.; Grewal, S.I.S. Histone deacetylation primes self-propagation of heterochromatin domains to promote epigenetic inheritance. Nat. Struct. Mol. Biol. 2022, 29, 898–909. [Google Scholar] [CrossRef]
  118. Eissenberg, J.C.; Elgin, S.C. HP1a: A structural chromosomal protein regulating transcription. Trends Genet. 2014, 30, 103–110. [Google Scholar] [CrossRef] [Green Version]
  119. Canzio, D.; Liao, M.; Naber, N.; Pate, E.; Larson, A.; Wu, S.; Marina, D.B.; Garcia, J.F.; Madhani, H.D.; Cooke, R.; et al. A conformational switch in HP1 releases auto-inhibition to drive heterochromatin assembly. Nature 2013, 496, 377–381. [Google Scholar] [CrossRef] [Green Version]
  120. Cowieson, N.P.; Partridge, J.F.; Allshire, R.C.; McLaughlin, P.J. Dimerisation of a chromo shadow domain and distinctions from the chromodomain as revealed by structural analysis. Curr. Biol. 2000, 10, 517–525. [Google Scholar] [CrossRef]
  121. Leopold, K.; Stirpe, A.; Schalch, T. Transcriptional gene silencing requires dedicated interaction between HP1 protein Chp2 and chromatin remodeler Mit1. Genes Dev. 2019, 33, 565–577. [Google Scholar] [CrossRef]
  122. Motamedi, M.R.; Hong, E.J.; Li, X.; Gerber, S.; Denison, C.; Gygi, S.; Moazed, D. HP1 proteins form distinct complexes and mediate heterochromatic gene silencing by nonoverlapping mechanisms. Mol. Cell 2008, 32, 778–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Fischer, T.; Cui, B.; Dhakshnamoorthy, J.; Zhou, M.; Rubin, C.; Zofall, M.; Veenstra, T.D.; Grewal, S.I. Diverse roles of HP1 proteins in heterochromatin assembly and functions in fission yeast. Proc. Natl. Acad. Sci. USA 2009, 106, 8998–9003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Sadaie, M.; Kawaguchi, R.; Ohtani, Y.; Arisaka, F.; Tanaka, K.; Shirahige, K.; Nakayama, J. Balance between distinct HP1 family proteins controls heterochromatin assembly in fission yeast. Mol. Cell Biol. 2008, 28, 6973–6988. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Shimada, A.; Dohke, K.; Sadaie, M.; Shinmyozu, K.; Nakayama, J.; Urano, T.; Murakami, Y. Phosphorylation of Swi6/HP1 regulates transcriptional gene silencing at heterochromatin. Genes Dev. 2009, 23, 18–23. [Google Scholar] [CrossRef] [Green Version]
  126. Eissenberg, J.C.; Ge, Y.W.; Hartnett, T. Increased phosphorylation of HP1, a heterochromatin-associated protein of Drosophila, is correlated with heterochromatin assembly. J. Biol. Chem. 1994, 269, 21315–21321. [Google Scholar] [CrossRef]
  127. Minc, E.; Allory, Y.; Worman, H.J.; Courvalin, J.C.; Buendia, B. Localization and phosphorylation of HP1 proteins during the cell cycle in mammalian cells. Chromosoma 1999, 108, 220–234. [Google Scholar] [CrossRef]
  128. Lejeune, E.; Bortfeld, M.; White, S.A.; Pidoux, A.L.; Ekwall, K.; Allshire, R.C.; Ladurner, A.G. The chromatin-remodeling factor FACT contributes to centromeric heterochromatin independently of RNAi. Curr. Biol. 2007, 17, 1219–1224. [Google Scholar] [CrossRef] [Green Version]
  129. Brasher, S.V.; Smith, B.O.; Fogh, R.H.; Nietlispach, D.; Thiru, A.; Nielsen, P.R.; Broadhurst, R.W.; Ball, L.J.; Murzina, N.V.; Laue, E.D. The structure of mouse HP1 suggests a unique mode of single peptide recognition by the shadow chromo domain dimer. EMBO J. 2000, 19, 1587–1597. [Google Scholar] [CrossRef] [Green Version]
  130. Thiru, A.; Nietlispach, D.; Mott, H.R.; Okuwaki, M.; Lyon, D.; Nielsen, P.R.; Hirshberg, M.; Verreault, A.; Murzina, N.V.; Laue, E.D. Structural basis of HP1/PXVXL motif peptide interactions and HP1 localisation to heterochromatin. EMBO J. 2004, 23, 489–499. [Google Scholar] [CrossRef] [Green Version]
  131. Liu, Y.; Qin, S.; Lei, M.; Tempel, W.; Zhang, Y.; Loppnau, P.; Li, Y.; Min, J. Peptide recognition by heterochromatin protein 1 (HP1) chromoshadow domains revisited: Plasticity in the pseudosymmetric histone binding site of human HP1. J. Biol. Chem. 2017, 292, 5655–5664. [Google Scholar] [CrossRef] [Green Version]
  132. Huang, Y.; Myers, M.P.; Xu, R.M. Crystal structure of the HP1-EMSY complex reveals an unusual mode of HP1 binding. Structure 2006, 14, 703–712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Kwon, S.H.; Florens, L.; Swanson, S.K.; Washburn, M.P.; Abmayr, S.M.; Workman, J.L. Heterochromatin protein 1 (HP1) connects the FACT histone chaperone complex to the phosphorylated CTD of RNA polymerase II. Genes Dev. 2010, 24, 2133–2145. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Tsukii, K.; Takahata, S.; Murakami, Y. Histone variant H2A.Z plays multiple roles in the maintenance of heterochromatin integrity. Genes Cells 2022, 27, 93–112. [Google Scholar] [CrossRef]
  135. Murawska, M.; Braun, S. Chaperoning heterochromatin: New roles of FACT in chromatin silencing. Trends Genet. 2022, 38, 646–649. [Google Scholar] [CrossRef] [PubMed]
  136. Machida, S.; Takizawa, Y.; Ishimaru, M.; Sugita, Y.; Sekine, S.; Nakayama, J.I.; Wolf, M.; Kurumizaka, H. Structural Basis of Heterochromatin Formation by Human HP1. Mol. Cell 2018, 69, 385–397.e388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Schematic structures of FACT complex in yeast and higher eukaryotes. N1: subdomain of peptidase-like domain 1, N2: subdomain of peptidase-like domain 2, DD: dimerization domain, Tandem PH: tandem pleckstrin homology domain, PH1: pleckstrin homology domain 1, PH2: pleckstrin homology domain 2, Acidic: acidic amino acid cluster, HMGB: high-mobility group box.
Figure 1. Schematic structures of FACT complex in yeast and higher eukaryotes. N1: subdomain of peptidase-like domain 1, N2: subdomain of peptidase-like domain 2, DD: dimerization domain, Tandem PH: tandem pleckstrin homology domain, PH1: pleckstrin homology domain 1, PH2: pleckstrin homology domain 2, Acidic: acidic amino acid cluster, HMGB: high-mobility group box.
Biomolecules 13 00377 g001
Figure 2. FACT working model for the mononucleosome and dinucleosome. PL: peptidase-like domain, TPH: tandem PH domain, DD: dimerization domain, A: acidic cluster. (A) Yeast FACT on mononucleosome. FACT functions as a histone chaperone in this scenario. (B) Yeast FACT on dinucleosome. Whether FACT functions as a histone chaperone in this scenario remains unknown.
Figure 2. FACT working model for the mononucleosome and dinucleosome. PL: peptidase-like domain, TPH: tandem PH domain, DD: dimerization domain, A: acidic cluster. (A) Yeast FACT on mononucleosome. FACT functions as a histone chaperone in this scenario. (B) Yeast FACT on dinucleosome. Whether FACT functions as a histone chaperone in this scenario remains unknown.
Biomolecules 13 00377 g002
Figure 3. Schematic diagrams of SBF-regulated gene promoter. UAS: upstream activating sequence, URS1: upstream regulatory sequence 1, URS2: upstream regulatory sequence 2. (A) Promoter structure of CLN1 and CLN2 genes. Three SBF-binding sites are annotated as arrows. Positive and negative regulators of gene expression are shown below and above the schematic of the gene structure, respectively. FACT is assigned an underline. (B) Promoter structure of the HO gene. Two Swi5-binding sites, one SBF-binding site, and one Ash1-binding site are annotated in URS1. Eight SBF-binding sites are annotated in URS2. Positive and negative regulators of gene expression are shown below and above the schematic of the gene structure, respectively. FACT is assigned an underline.
Figure 3. Schematic diagrams of SBF-regulated gene promoter. UAS: upstream activating sequence, URS1: upstream regulatory sequence 1, URS2: upstream regulatory sequence 2. (A) Promoter structure of CLN1 and CLN2 genes. Three SBF-binding sites are annotated as arrows. Positive and negative regulators of gene expression are shown below and above the schematic of the gene structure, respectively. FACT is assigned an underline. (B) Promoter structure of the HO gene. Two Swi5-binding sites, one SBF-binding site, and one Ash1-binding site are annotated in URS1. Eight SBF-binding sites are annotated in URS2. Positive and negative regulators of gene expression are shown below and above the schematic of the gene structure, respectively. FACT is assigned an underline.
Biomolecules 13 00377 g003
Figure 4. Wave of nucleosome deposition and reposition at the HO promoter during cell cycle progression. Based on GALp::CDC20 arrest/release experiment, nucleosome dynamics were analyzed by time-course histone H3 ChIP-qPCR during M to S phase transition. URS1: upstream regulatory sequence 1, URS2: upstream regulatory sequence 2. (A) Schematic representation of nucleosome dynamics along the HO promoter in wildtype cells. (B) Schematic representation of nucleosome dynamics along the HO promoter in FACT mutant cells.
Figure 4. Wave of nucleosome deposition and reposition at the HO promoter during cell cycle progression. Based on GALp::CDC20 arrest/release experiment, nucleosome dynamics were analyzed by time-course histone H3 ChIP-qPCR during M to S phase transition. URS1: upstream regulatory sequence 1, URS2: upstream regulatory sequence 2. (A) Schematic representation of nucleosome dynamics along the HO promoter in wildtype cells. (B) Schematic representation of nucleosome dynamics along the HO promoter in FACT mutant cells.
Biomolecules 13 00377 g004
Figure 5. Schematic diagrams of the FACT recruitment model at pericentromeric heterochromatin.
Figure 5. Schematic diagrams of the FACT recruitment model at pericentromeric heterochromatin.
Biomolecules 13 00377 g005
Figure 6. Hypothetical FACT working model on heterochromatic nucleosomes. PL: peptidase-like domain, TPH: tandem PH domain, DD: dimerization domain, A: acidic cluster, CD: chromodomain, CSD: chromo-shadow domain. (A) HP1/Swi6 bridge dinucleosome. (B) FACT is recruited by a Swi6-CSD homodimer via the peptidase-like domain of Spt16. (C) HP1/Swi6 and FACT cooperatively bridge dinucleosome. (D) FACT binds to heterochromatic mononucleosome.
Figure 6. Hypothetical FACT working model on heterochromatic nucleosomes. PL: peptidase-like domain, TPH: tandem PH domain, DD: dimerization domain, A: acidic cluster, CD: chromodomain, CSD: chromo-shadow domain. (A) HP1/Swi6 bridge dinucleosome. (B) FACT is recruited by a Swi6-CSD homodimer via the peptidase-like domain of Spt16. (C) HP1/Swi6 and FACT cooperatively bridge dinucleosome. (D) FACT binds to heterochromatic mononucleosome.
Biomolecules 13 00377 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Takahata, S.; Murakami, Y. Opposing Roles of FACT for Euchromatin and Heterochromatin in Yeast. Biomolecules 2023, 13, 377. https://doi.org/10.3390/biom13020377

AMA Style

Takahata S, Murakami Y. Opposing Roles of FACT for Euchromatin and Heterochromatin in Yeast. Biomolecules. 2023; 13(2):377. https://doi.org/10.3390/biom13020377

Chicago/Turabian Style

Takahata, Shinya, and Yota Murakami. 2023. "Opposing Roles of FACT for Euchromatin and Heterochromatin in Yeast" Biomolecules 13, no. 2: 377. https://doi.org/10.3390/biom13020377

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop