Skip to main content

Transcriptional silencing in Saccharomyces cerevisiae: known unknowns

Abstract

Transcriptional silencing in Saccharomyces cerevisiae is a persistent and highly stable form of gene repression. It involves DNA silencers and repressor proteins that bind nucleosomes. The silenced state is influenced by numerous factors including the concentration of repressors, nature of activators, architecture of regulatory elements, modifying enzymes and the dynamics of chromatin.Silencers function to increase the residence time of repressor Sir proteins at silenced domains while clustering of silenced domains enables increased concentrations of repressors and helps facilitate long-range interactions. The presence of an accessible NDR at the regulatory regions of silenced genes, the cycling of chromatin configurations at regulatory sites, the mobility of Sir proteins, and the non-uniform distribution of the Sir proteins across the silenced domain, all result in silenced chromatin that only stably silences weak promoters and enhancers via changes in transcription burst duration and frequency.These data collectively suggest that silencing is probabilistic and the robustness of silencing is achieved through sub-optimization of many different nodes of action such that a stable expression state is generated and maintained even though individual constituents are in constant flux.

The DNA in a eukaryotic nucleus is wrapped around histones to form nucleosomes. The interplay between non-histone proteins and nucleosomal filaments leads to stable programs of gene expression [1] resulting in a continuum of expression levels [2,3,4]. Transcriptional activation involves sequence specific transcription factors (TFs) and general transcription factors (GTFs) [5, 6]. The former bind upstream activating sequence (UAS) enhancers to regulate transcription while the GTFs bind sequences in the core promoter to initiate transcription. Transcriptional repression refers to an inactive state dependent upon the continual presence of the repressing signal and specific repressor proteins that often bind specific sequences [7, 8]. Silencing is a position-independent persistent and stable form of repression that requires proteins that bind to DNA sequences called silencers as well as repressor proteins that interact with nucleosomes to create a structure that silences multiple genes with diverse regulatory elements and once established the expression state is maintained and propagated with high fidelity [9].

The aim of this review is to highlight recent advances and key outstanding questions on gene silencing in Saccharomyces cerevisiae. It is not a comprehensive review of the subject and the reader is directed to several excellent comprehensive reviews on this subject [9,10,11,12,13]. There are many parallels between Sir mediated silencing in S. cerevisiae and HP1 and Polycomb mediated silencing in other eukaryotes and some of these similarities will be highlighted in the course of this review (also see [14,15,16,17,18]).

What is Silencing?

Based on recent and prior results:

1) Silencers are critical in establishing and maintaining silencing. Their function is to increase the concentration of the repressor Sir proteins in the vicinity of the silenced chromatin domains.

2) The regulatory elements involved in gene activation are key arbiters that determine the effectiveness of silencing and only weak constitutively active genes are stably and effectively silenced.

3) Silencing is the result of a web of multivalent stereo-specific interactions between Sir proteins, silencer bound proteins and nucleosomes. These interactions create a chromatin environment that reduces nucleosome mobility and hinders the rate of factor access with attendant negative effects on transcription.

4) Silencing is influenced by clustering of silenced domains. The formation of membrane-less compartments increases Sir protein concentration; helps mediate long-range Sir-nucleosome interactions and possibly hinders the ability of cofactors to function.

5) Silencing is an energy dissipating, non-equilibrium, stochastic probabilistic process driven by constituents that are constantly in flux. The process is governed in part by concentration, affinity and residence times of many different individual components. These different parameters together determine the silenced state of a gene.

6) The stability of the silenced state is likely to be governed by diffuse forms of “co-operativity” resulting in elevated avidity of the Sir proteins across the entire domain.

Silencing has been investigated using a myriad of approaches, each of which interrogates distinct and specific aspects of silencing. The individual vignettes integrated together generate a picture of the process. The principles of silenced chromatin gleaned and enunciated in this review, the known unknowns highlighted and the models proposed to explain silenced chromatin are focused on S. cerevisiae silencing. The experiments and the models proposed often rely on studies on facultative and constitutive silenced chromatin in other eukaryotes and are also arguably relevant for studies in these organisms.

Transcription activation

Gene regulation is mediated by proteins that bind specific DNA sequences as well as proteins that interact with and alter the chromatin template and involves tens of molecular reactions. Transcription of a gene is initiated when a site in the enhancer is bound by TFs. Understanding how TFs find and bind DNA is necessary for the understanding of silencing since silencers are bound by proteins that follow the same rules as TF binding to enhancers. Most TFs recognize and bind short DNA sequences and explore the genome through transient interactions until they find and bind a sequence with high affinity for significant periods resulting in “sequence-specific” binding [19,20,21,22,23,24,25].

TFs bind DNA sites in the genome with varying strengths, specificities and lengths of time and these interactions then result in functional consequences. Their function is dependent upon their binding domains, their ability to phase separate into condensates, the number of high affinity binding sites for TFs, the density of binding sites, cooperativity of binding and the ability of TFs to interact with multiple components such as other TFs, cofactors, histone modifiers and chromatin remodellers. TF binding can thus be regulated by altering their concentration, biochemical characteristics or binding affinities or by altering their access to binding sites. The residence times of TFs at these elements are also influenced by other factors such as adjacent flanking sequences, chromatin dynamics and silencing [24, 26,27,28,29,30,31,32,33,34].

While sequence specific interactions are critical for gene regulation, non-specific interactions also play a supporting but important role. The association and dissociation of TFs between clusters of binding sites can trap TFs in a specific region of the nucleus creating membrane-less nuclear compartments with high local concentrations of TFs [26, 35]. The presence of elevated levels of specific proteins in the nucleoplasm can also lead to negative effects by inducing unbinding of TFs via facilitated dissociation akin to squelching [36], a mechanism that could play a role in gene silencing.

The transient interactions between TFs and DNA result in functional consequences. There is a temporal coupling between binding events at the enhancer and the promoter but the relationship is not simple [37,38,39]. TFs bind their sites transiently in chromatin and recruit chromatin modifiers and remodellers to stabilize open sites [40]. Stable TF and mediator binding then create temporal windows during which information is transferred, possibly through allostery, to the GTFs bound to the core promoter resulting in transcription initiation and convoys of RNA polymerase [41]. The length of a transcription burst, how often a burst of transcription occurs and how many transcripts are produced in a burst are regulated. In broad terms, the frequency of TF binding to an enhancer affects transcription burst frequency while core promoter sequences affect burst duration by modulating residence times of GTFs [37, 42, 43]. There are different individual binding timescales that collectively determine bursting. These include TF binding and unbinding, mediator dynamics, chromatin modifier and remodeller binding dynamics, nucleosome cycling and GTF binding and release amongst others [39]. Furthermore, the different steps do not have identical reaction rates and this coupled with fluctuations in concentrations of factors generates transcription initiation events that fluctuate. Some genes like housekeeping genes are highly transcribed, more or less continuously, while others are rarely transcribed or transcribed only under inducing conditions in bursts [44]. Within this framework, transcriptional silencing could be visualized as the ultimate endpoint of burst regulation resulting in a completely inactive state that dampens some or all of these steps and persists across multiple cell cycles (Figs. 1 and 7) but questions remain regarding the identity of the specific step/s and the mechanism by which silencing is affected.

Current data suggest that a simple thermodynamic equilibrium framework is not sufficient to fully describe gene regulation [45, 46] and nucleosome dynamics play an important role in this process [37, 47]. The affinity of a nucleosome for a site is typically greater than the affinity of the TF for that site [27, 48, 49]. The DNA wrapped around nucleosomes also constantly breathes via thermal motion creating temporal windows of accessibility and inaccessibility for TF binding [40, 50,51,52,53].

There is a high degree of nucleosome organization at regulatory regions with well-positioned + 1 and − 1 nucleosomes flanking a nucleosome depleted region (NDR). The NDR encompasses the UAS enhancer and core promoter [54,55,56,57] and the NDR is actively generated and maintained by ATP expending chromatin remodelers aided by post translational modifications of histones [6, 31, 46, 49, 58,59,60,61,62,63,64]. The precise location of the + 1-nucleosome adjacent to the NDR influences transcription initiation [65] and is likely an important arbiter of transcription bursting. If the rate of sliding/removal of nucleosomes or the rate of unwrapping of the DNA from the nucleosome surface is decreased, then nucleosomes hinder accessibility of sites even when TFs are present and the probability of transcription corresponds in part to the rate of TF binding site exposure in nucleosomes and it is likely that silencing affects these rates.

In a population of cells there exist different nucleosome configurations over the regulatory regions of genes; some that are conducive for transcription and others not (Fig. 2). Surprisingly, all possible nucleosome configurations are observed in the population regardless of the gene’s average expression state [47, 66]. The gene cycles through different nucleosome configurations and dwells for different lengths of time in each configuration. The key difference is the frequency of the different nucleosome configurations. The fully nucleosomal configuration is more often found under repressing conditions (right panels in Fig. 2) while very few gene regulatory regions are fully nucleosomal under activating conditions (left panels in Fig. 2). TFs and chromatin remodelers regulate transcription by changing the probability distribution of different nucleosomal configurations at regulatory sites and the Sir proteins may function to alter the frequency of these probability distributions during silencing.

Some genes such as housekeeping genes are regulated in a deterministic (less noisy) manner while others utilize opportunistic mechanisms [67]. Some TFs called general regulatory factors function deterministically- they are constitutively present at high concentrations and recruit chromatin remodellers to reposition and mobilize nucleosomes (see top panel in Fig. 2). In a population of cells, they increase the likelihood of chromatin configurations that favour high levels of transcription burst frequency and duration (see [68] for a detailed description). Inducible genes are regulated by TFs whose biochemical characteristics and concentration in the nucleus change under inducing conditions and these TFs then alter the nucleosome configurations using varied mechanisms [5, 6] (middle panel in Fig. 2). The end point of this is a change in the probability distribution of active versus inactive nucleosome configurations leading to sustained bursts of transcription under inducing conditions [47]. The final class of genes are entirely opportunistic- they may not rely on TFs [67] but on stochastic removal of nucleosomes from core promoters by free floating chromatin remodellers and histone modifiers allowing PIC formation and these genes are the weakest in terms of transcription output (bottom panel in Fig. 2). Thus, each class of genes has different probabilistic expression profiles dependent on the frequency of the different nucleosomal architectures that are generated by chromatin remodellers and histone modifiers and the responses of these genes to silencing factors is correspondingly different [69].

Silencer activity

Silencers are essential for silencing [70] and possess a robust NDR bound by combinations of multi-functional DNA binding proteins ORC, Rap1, Abf1 and Sum1 [71, 72]. A signature function of silencers is the “recruitment” of the Sir2, Sir3, Sir4 proteins to the silenced domain [73,74,75,76,77,78] and possibly in channeling chromatin remodelers to generate evenly spaced nucleosomes thus creating a directionality in the spread of silencing [79,80,81]. Once Sir2 and Sir4 are recruited to the silencers, Sir2 then deacetylates histones in adjacent nucleosomes enabling Sir3 and Sir4 to bind to the hypoacetylated evenly positioned nucleosomes [76, 77, 82,83,84]. Repetition of this process leads to Sir protein spreading creating a silenced domain [9, 13, 71].

We understand the function of the silencers but several questions persist regarding their mode of action. How the specific configuration of sites and proteins at silencers result in silencing as opposed to gene activation remains unclear. Like models for enhancer function, do silencers and their associated proteins function like billboards performing distinct, independent roles or do they function by forming a large complex analogous to the enhanceosome [85, 86]? Whether silencer bound proteins are regulated during silencing is also not known [87, 88] and whether their binding to the silencer depends upon cofactors such as Sir1, chromatin modifying and remodeling factors is also not known.

Like enhancers, different silencers have varying strengths [89, 90]. This is likely due to the differences in the binding parameters of silencer bound proteins with one another, with DNA and with the Sir proteins. While silencer bound proteins are undoubtedly subject to the laws of thermodynamics, surprisingly little is known about the affinities, residence times and cooperativity of these proteins at silencers. The binding of ORC to the silencers appears to be stronger than its interaction with euchromatic origins of replication [91] while measurements of residence time of Rap1 show faster turnover at silenced telomeres compared to sites at enhancers [92]. Thus, while additional research is necessary, one model is that differences in thermodynamic and kinetic parameters result in differences in the amount of Sir proteins retained/recruited in the vicinity of the silencers leading to differences in silencer strengths.

In vertebrates, DNA sequence elements functioning as silencers have not been identified for HP1 or polycomb mediated silencing but silencer like elements have been annotated and analyzed in S. pombe [93, 94] and Drosophila [95]. In the instances where silencers have been demarcated, the functions ascribed to these DNA elements are analogous to the roles ascribed to silencers in S. cerevisiae (reviewed in [15, 96,97,98,99,100,101]). Like Sir mediated silencing, nucleation and spread of heterochromatin in other eukaryotes involves targeting methylases to methylate histones in nucleosomes followed by binding of the methylases and repressor proteins to the methylated histone leading to the spread of silenced chromatin [102, 103].

The nature of silenced chromatin domains

Chromatin plays a critical role in gene silencing. One defining property of silenced chromatin is that it is marked by hypoacetylated histones H3 and H4 that are necessary for Sir protein binding (reviewed in [13]). Consistent with this, ChIP-seq mapping of H4K16 acetylation reveals uniformly hypoacetylated histones across entire silenced domains [104,105,106]. The silenced domain is also marked by evenly spaced nucleosomes with long linker DNAs. This nucleosome configuration is silencer dependent and disrupted in Sir mutants suggesting that Sir-nucleosome interactions play a role in organizing/stabilizing nucleosome configurations [81, 107, 108]. Surprisingly, the regulatory regions of silenced genes have accessible NDRs [109, 110]. Furthermore, ChIP indicate that the Sir proteins are highly enriched at the silencers with reduced levels at the enhancers/promoters of the silenced genes [104,105,106, 111] (Fig. 3). Alternative methods mapping dynamic interactions also show localized peaks of Sir proteins at silencers [110]. Consonant with these are studies which show highly localized peaks of Sir protein at sub-telomeric sites with flanking nucleosomes bound by Abf1 and Reb1 creating a trinucleosome repressed domain [67, 112] consistent with early data showing discontinuous silencing domains [113,114,115] though it is unclear if this architecture is present in all cells or is a population average image derived from heterogeneity of binding.

Data show that the Sir proteins are mobile in the nucleus [70, 82, 116, 117] and constantly exchange between sites. The peaks of Sir proteins at the silencers suggest that their residence time at silencers is likely to be high while the lower levels within the silenced domain suggest that their binding to nucleosomes is more transient. This raises the question of whether stable Sir protein binding to nucleosomes is necessary to block transcription or whether deacetylation of histones and/or preventing movement of nucleosomes across the entire domain is the key driver in silencing and illustrates the importance of considering binding kinetics. While the binding affinities of Sir3 for acetylated and unacetylated nucleosomes are known [118], we do not have information on affinities of the other Sir proteins or the residence time of the Sir proteins bound to chromatin. To gain a more granular picture, it is necessary to compare residence times and binding constants of Sir proteins at silencers versus nucleosomes as well as the turnover rates of histone modifications.

The end point of Sir protein binding to nucleosomes is the inhibition of transcription. A classical study measuring accessibility of chromatin to a DAM methylase enzyme demonstrated that the Sir proteins reduced access to DNA [119] though a recent study failed to observe significant reduction in accessibility at silenced chromatin to other probes [120]. In vitro studies of Sir bound chromatin also highlight reduced accessibility to various enzymatic probes [118, 121, 122] (see [123] for detailed discussion on chromatin accessibility assays). While these studies showed reduced accessibility but not inaccessibility, they highlight the dynamic but restrictive nature of silenced chromatin. It should be noted that most of the in vitro studies were performed in the absence of active chromatin modifying and remodeling complexes under conditions that favor stable binding of Sir proteins to nucleosomes.

Based on these studies simple steric hindrance models were proposed where Sir proteins bound to key nucleosomes stereospecifically blocked the binding of TFs or GTFs to regulatory sequences. Consistent with these models, ChIP mapping experiments with different reporter genes showed that silenced chromatin restricts different transcription proteins from binding at different enhancers and promoters [124,125,126,127,128].

The question then is how the Sir proteins block multiple factors and processes via a single binding mode. The presence of an accessible NDR at the regulatory regions of silenced genes, the cycling of different chromatin configurations at regulatory sites, the mobility of Sir proteins, and the non-uniform distribution of the Sir proteins across the silenced domain raise questions regarding these simple steric hindrance models for silencing. Gene activation is a multi-step dynamic process and different gene regulatory elements and their cognate TFs use different cofactors in different temporal order to generate a PIC and initiate transcription [5, 6]. It is likely that the Sir proteins alter specific rate-limiting steps in this process to mediate silencing. One possibility is that the Sir proteins alter a common early step while being agnostic to the diversity of downstream effector proteins. One early common step in gene activation is the mobilization of nucleosomes from regulatory sequences prior to transcription. Sir proteins could directly or indirectly affect the rate of sliding and/or removal of nucleosomes from regulatory sites. This would alter the probability distribution of active versus repressed nucleosomal configurations which would then have downstream effects on TF binding, PIC formation or post initiation events. While replication-independent histone exchange is infrequent and affects only 1 to 10% of nucleosomes [64], there is a correlation between the level of exchange and RNA polymerase levels [129]. Consistent with this, silenced heterochromatic genes have lower histone exchange compared to active euchromatic genes [130] though the promoters of silenced genes have increased histone turnover compared to the coding regions of these genes (which might help explain the NDR at these loci). In addition, an in vitro study suggests that the Swi/Snf chromatin remodeller interacts with Sir3 and this interaction is necessary for the eviction of Sir3 from reconstituted heterochromatin [131]. A recent study also showed that reducing nucleosome density and increasing the degree of freedom for nucleosome movement destabilized the silenced state [132]. While in vivo data with remodeller mutants could be due to indirect pleiotropic effects a recent analysis of chromatin remodeler mutants with silenced reporter genes [69] was consistent with and as such supportive of this model.

It will be interesting to know the frequency distributions of different nucleosomal configurations when a gene is active versus silent as well as the precise positioning of nucleosomes and the role of chromatin remodellers in this process. Knowing the binding affinities and residence time of the GTF/TFs at silenced enhancers and promoters would also be informative and allow the testing of different models.

This model builds on pioneering studies in other eukaryotes showing that polycomb and trithorax complexes colocalize [133] and facultative heterochromatin mediated repression is mediated in part by inhibiting chromatin remodellers and histone acetylases [15, 31, 134,135,136,137,138]. Similar mechanisms have also been shown to operate for constitutive HP1 containing heterochromatin. The turnover dynamics of nucleosomes are conserved across sites in S. pombe and Drosophila [139,140,141,142] raising the possibility that different silencing proteins (and different chromatin modifications) might utilize similar mechanisms to mediate silencing in all eukaryotes.

Sir protein structures and models for gene silencing

The next question is how the Sir proteins alter nucleosome dynamics. Sir proteins bind nucleosomes and each other and the nature of this stereospecific binding is central to silencing. The web of multivalent interactions (Fig. 4) suggests coordinated and/or sequential interactions between Sir proteins and their partners in silencing.

The structure of several Sir protein domains has been solved though the structure of the holocomplex has not yet been determined [122, 143,144,145,146,147,148,149,150,151,152]. These structural studies highlight possibilities by which Sir binding to nucleosomes could affect chromatin configurations and silencing. Sir3 contains a winged helix domain which is important for Sir3-Sir3 interactions that are necessary for its binding to nucleosomes [122]. Structural studies on the H1 winged helix-nucleosome interactions are suggestive of how Sir protein binding might play a role in nucleosome mediated silencing [153, 154]. Analogous to the H1-nucleosome structures, the binding of the Sir3 winged helix to the nucleosome and the induced clamping of the H4 tail with intranucleosomal DNA could reduce nucleosome mobility thus indirectly restricting access of the transcription machinery (Fig. 5). In this model Sir protein binding would raise the energetic barrier for the opening of chromatin thus inhibiting transcription. Second, Sir2 mediated histone deacetylation is likely to strengthen histone-DNA interactions thus reducing breathing of nucleosomal DNA [156] and thus reducing the rate of chromatin opening. The reduced breathing of the clamped nucleosomal DNA would close a window of opportunity for TF binding. Third, unacetylated tails have a reduced affinity for chromatin remodeler binding thus discouraging nucleosome movement [157]. Sir protein binding to nucleosomal linker DNA [110] results in even spacing and positioning of nucleosomes with long linkers [81, 107, 108] which would likely help in nucleosome stacking and interdigitation of the chromatin filament [158] thus enabling internucleosomal interactions and chromatin compaction [159,160,161] which could reduce TF access to DNA.

There are two classes of BAH domains in yeast- Sir3 like and RSC like [155] and these domains are involved in protein-nucleosome interactions. Structure determinations show that Rsc binds nucleosomes at sites close to those occupied by the Sir3 BAH domain [148, 155, 157, 162,163,164] (Fig. 5). A higher koff for Rsc compared to Sir3 (whose koff is currently unknown) could therefore favor a Sir3-nucleosome bound state at genes undergoing silencing thus highlighting another molecular mechanism that could influence chromatin remodeler mediated nucleosome mobility.

BAH domains are also necessary for H3K27me3 recognition in nucleosomes by polycomb proteins [165] while interactions between HP1 and histone H1 have been shown to function in HP1 mediated silencing [18]. All these models are speculative and in need to testing. Thus, while the structures of Sir protein domains have been informative, the structure of a Sir holocomplex with oligonucleosomes would be illuminating. Biochemical reconstitutions of silenced chromatin in the presence of silencers, sub-saturating levels of the Sir proteins and chromatin with different repeat lengths would also be informative.

The role of Sir protein compartments in silencing

Silenced domains localize to the nuclear periphery and are anchored by numerous factors (reviewed in [166,167,168,169,170,171]) leading to the tethering of chromatin fibers and the clustering of these domains. While differential properties of chromatin have been suggested to promote phase separation and compartmentalization of the nucleoplasm [172], there is as yet no evidence of Sir proteins forming liquid droplets. Clustering of domains creates silencing compartments enriched for Sir proteins (reviewed in [12, 168, 169, 173, 174]) and this is important for transcription repression [175]. There are approximately 57,000 nucleosomes in yeast [57] and roughly 10% of these are hypoacetylated [176]. Given the sizes of silenced regions, one can estimate that ~ 2/3 of the hypoacetylated nucleosomes are present at silenced domains. High-throughput studies suggest that there are ~ 1400 molecules of Sir3 and Sir4 protein in a typical yeast cell [177] which is significantly fewer than the number of unacetylated nucleosomes. Consistent with this, analysis has shown that Sir proteins are a limiting component of stable silencing [83, 90, 156, 178, 179].

Clustering of silenced domains would aid in trapping of Sir proteins amongst the clustered loci (pinball effect) creating a compartment with elevated concentrations of Sir proteins which would then in turn influence binding equilibria [111, 169, 180] (Fig. 6). It is informative to consider a hypothetical scenario. If one Sir protein is present in a yeast nucleus with a diameter of 2 μm (volume ~ 4 μm [3] or 4 fl.), this is equal to a concentration of ~ 2 nM. If the nucleus is compartmentalized and if one were to restrict one Sir protein to a compartment that is 10% of the diameter of the typical nucleus, the Sir protein concentration would increase nearly 1000-fold providing a powerful way to increase the effective concentration. Recent data show that one of the earliest steps in the establishment of silencing involves the perinuclear anchoring of silenced loci that aids in the accumulation of Sir proteins which reinforced anchoring culminating in a self-reinforcing loop in silencing [181].

Clustering could therefore alter thermodynamic parameters and Sir concentration is likely to be one important parameter. Measurements in mammalian cells show an approximately two-fold difference in nucleosome density between euchromatin and heterochromatin [182] but this leads to subtle effects on molecular diffusion and movement of transcription factors [183, 184] and it is therefore unlikely that silencing can be solely described by these values alone especially given the non-equilibrium nature of gene regulation and the relative immobility of tethered chromatin fibers.

Alterations in concentration will also not alter stereospecific interactions between Sir proteins and nucleosomes but would create a “circe” effect [180]. Increased Sir concentrations would reduce the search times required for Sir proteins to find, deacetylate and bind unacetylated nucleosomes (kon) and would alter the probability distribution of hypoacetylated Sir bound nucleosomes. Nucleosomes that localize within clouds of high Sir protein concentration would have a higher probability of becoming and remaining hypoacetylated. The concentration dependent macromolecular crowding might also alter TF search times and Kon thus hindering TF and chromatin remodeler access [183,184,185,186,187,188]. This is consistent with studies where merely tethering a locus to the nuclear periphery results in Sir mediated repression of reporter genes [189, 190]. The increased concentration may aid in cooperativity between Sir proteins though no Hill coefficient measurements have been reported for silencing components. While clustering should not alter the residence times of the Sir proteins bound to nucleosomes, the higher concentration of free-floating Sir proteins in the nucleoplasm might also facilitate dissociation of TFs from UAS enhancers thus indirectly increasing the effectiveness of silencing.

The clustering of silenced domains also brings distant sequences into close three-dimensional proximity [111, 191, 192] and this helps in Sir protein dependent long-range internucleosomal interactions [122, 159, 193]. Recent analysis with specific mutant alleles showed that Sir3-Sir3 internucleosomal interactions promoted long-range chromatin contacts most likely via the winged helix domain of Sir3 [161]. The three-dimensional web of interactions between silencers, Sir proteins and nucleosomes would create a dynamic 3D structure with a self-enforcing loop that stabilizes the silenced state [194, 195].

The importance of subnuclear compartments in silencing is highly conserved across eukaryotes. Polycomb and Swi6 mediated compaction and phase separation aids in H3K27me3 and H3K9me3 mediated facultative and constitutive heterochromatin [141, 196,197,198,199,200,201,202,203] highlighting some commonalities in processes between different heterochromatic systems.

Differential susceptibility of enhancers and core promoters to silencing

Silencing has traditionally been considered a gene agnostic phenomenon that could silence most genes. However, recent analysis of native genes at silenced loci using mRNA-Seq showed that only a few genes at native telomeres are silenced while many others resist/escape silencing [104,105,106]. Detailed analysis of discontinuous silencing at native telomeres showed that a dynamic competition between TFs and histone deacetylation by Sir2 determines the extent of silencing at these sites [112]. Recent studies measuring silencing of a set of gene regulatory elements using multiple reporter assays including directly visualizing transcription in single cells also showed that Sir proteins stably silence only weak and uninduced stress response regulatory elements but are unable to stably repress strong housekeeping gene regulatory elements and they do so by altering transcription bursting [69].

Chromatin configurations at regulatory sites constantly cycle between open and closed states. The enhancers that are silenced have weak or non-existent TF binding sites with low burst frequency and duration and thus succumb to silencing - monostable repression. For activation, these genes utilize opportunistic mechanisms, where chance opening of the regulatory sequences by nucleosome sliding/movement by free floating chromatin remodellers allows for the transcription machinery to initiate transcription. The presence of Sir proteins bound to nucleosomes would reduce the mobility of nucleosomes and in a population of cells these genes would predominantly be packaged in chromatin configurations that disfavor transcription thus reducing the probability of gene activation (Fig. 2). In this statistical scenario, the Sir proteins would function not by increased condensation of chromatin or physically blocking a factor from binding but by reducing the rate of chromatin opening.

Strong housekeeping genes resist silencing and remain active- monostable activation. These strong constitutively active regulatory elements bound by general regulatory TFs, with their attendant biochemical properties, exhibit high burst frequency and duration and remain active [69]. The ability of these genes to resist silencing (monostable activation) is likely determined by varied criteria such as the type of TFs, the number of binding sites for the TFs, the concentrations and the binding characteristics of these proteins and their ability to recruit chromatin remodelers and modifying enzymes [68, 112]. These factors together likely affect the probability distribution of active versus inactive chromatin configurations to favor the accessible chromatin states. This may also be the underlying reason why heterochromatin barrier insulators are populated by such sequences [115, 204,205,206].

Under certain circumstances, bistable expression states arise for genes undergoing silencing [207,208,209]. Bimodal expression patterns in genetically identical cells arise from direct competition between TFs and repressors for binding to the same site [210]. Bimodal expression states become bistable (expression states that persist through cell division) due to either positive feedback or double negative feedback coupled with cooperativity (non-linearity) [211, 212]. Bistable expression states arise at silenced loci under specific conditions- when mutations in the silencers weaken silencing, when a relatively strong activator binds the UAS enhancer of a reporter gene or when a silenced domain is transposed to a euchromatic site in the nucleus where Sir protein concentrations are lower [69, 70, 207,208,209].

In the context of bistable expression states, one question that naturally arises is how the active state persists in competition with silencing (and vice versa) given that activation is a multi-step probabilistic process where each step is naturally inefficient, stochastic and transient (bursty) [37, 213, 214]. One possibility the active state is stably maintained is because the establishment of the silenced state is an inherently slower process, compared to activation, and requires an extended period of transcription inactivity to fully form [181, 215]. A gene that is repeatedly activated (even for short bursts) would prevent a stable silent state from forming (Fig. 7).

Another reason for the observed stability is that the silenced domains exhibit hysteresis (see [216,217,218] for a description of hysteresis). The response of the silenced domain to changes in histone acetylation depends on whether the system is in its off or its on state and Sir2 enzymatic activity is the driver of the hysteresis effect [219, 220]. Recent work shows that if the locus is silent, then 75% of nucleosomes need to become acetylated (acquire an acetyl mimic histone H4 (K16Q)) before the system loses silencing [90]. If the locus is active (because of the presence of the histone H4K16Q allele), then greater than 75% of nucleosomes must acquire unacetylated histones (wild-type H4K16) for an extended period of time to establish silencing (Ken Wu and RTK unpublished results). This Sir2 and histone acetylation mediated hysteresis effect could therefore also contribute towards the observed stability of the silenced state (Fig. 8).

The observations that Sir proteins can only stably silence a subset of genes is also reminiscent of earlier observations with polycomb proteins [221]. These proteins were shown to only stably silence specific promoters and most importantly the primary function of these proteins was the stable maintenance of the inactive state after transcription of the gene had ceased (reviewed in [15]). Recent studies with synthetic silencing systems mirror these conclusions [222, 223].

Domain wide regulation of stable heritable silencing

One defining property of silencing (albeit of weakly transcribing genes) is stable maintenance of the silenced state and the high fidelity of inheritance of this state once it is established [224, 225]. The silent domain is a stable structure created by Sir proteins in constant flux. Measurements of the stability of silencing showed that in wild type cells, silencing at HML is stochastically lost in one out of every 1000 cells with a similar value at HMR [224].

Interestingly, once the silenced domain is established it can tolerate significant fluctuations in the levels of Sir proteins and nucleosomal acetylation without loss of silencing. Silencing is only lost when approximately 75% of the nucleosomes across the silenced domain acquire H4K16 acetyl like marks [90]. While it is possible that the deacetylation of a single nucleosome is important in silencing, these data argue against this. Interestingly, simple calculations based on these data highlight the possibility that a rather small (less than two-fold) reduction in the ability of acetyltransferases to acetylate a nucleosome, spread across a domain of 20 nucleosomes, might be enough to establish a silent domain [90]. What these data also suggest is that the nucleosomes across the clustered silent domains function together where each nucleosome bound by Sir proteins acts as weak point silencer helping maintain transient elevated local concentrations of Sir proteins. The many multivalent interactions between Sir proteins and nucleosomes would create a 3D-mesh of interactions (Fig. 4) and each interaction would create a reinforcing feedback loop where Sir2 deacetylation of histones would facilitate Sir3 binding which in turn could facilitate interactions with Sir4 and its interacting partner Sir2 [211, 212]. Once a critical mass of Sir proteins bound to nucleosomes is achieved, the system becomes self-sustaining and stable. Thus, while the individual Sir-nucleosomes interactions are weak and therefore transient, the overall system would exhibit stability so long as multiple nucleosomal sites remain bound by Sir proteins for sufficiently long periods of time. In this system the transient removal of Sir proteins from a single nucleosome, even one over a key regulatory sequence would be unlikely to initiate or allow persistent and sustained gene activity.

In conclusion, silencing is a weak form of repression and there are numerous factors that together result in a stable silenced state. Silenced chromatin only efficiently represses weak enhancers and promoters. The silencer and silencer bound proteins are necessary for the efficient maintenance of the silent state. They function by increasing the local concentration of the Sir proteins. The clustering of silenced loci is important in this process. The data suggest that the silenced state depends on local Sir-nucleosome interactions as well as a domain wide web of interactions analogous to what has been proposed for phase separation in gene activation [26]. The stability of the final state is likely influenced by numerous factors such as the concentration of the Sir proteins, transcription activators, architecture of gene enhancers and promoters, histone modifying enzymes, the positioning and dynamics of nucleosomes over regulatory sequences and modifications of histone residues. The robustness of silencing is achieved through sub-optimization of many different factors such that in the presence of all these structures and processes a stable expression state is generated and maintained even though individual constituents are in constant flux.

Fig. 1
figure 1

A schematic of different transcription initiation frequencies and burst durations. Green trails represent transcription events. The number and clustering of transcription trails reflect bursting frequency and duration. X-axis reflects time

Fig. 2
figure 2

Schematic of nucleosomes cycling through different active and repressed chromatin configurations for different classes of UAS enhancers. Full circles indicate nucleosome presence, dashed circles indicate nucleosome depletion. Core promoters are shown as black rectangles while gray rectangles highlight TF binding sites in UAS enhancers

Fig. 3
figure 3

Schematic of the distribution of Sir proteins and TFs at silenced loci with different classes of UAS enhancer/core promoters. Red dots represent Sir proteins and peaks of distribution, green ovals represent TFs, blue squares indicate free-floating chromatin remodellers. Vertical black bars represent silencers and gray boxes represent enhancers and black boxes represent promoters

Fig. 4
figure 4

Schematic of the numerous multivalent interactions between the different factors involved in transcriptional silencing. Red arrows depict interactions involved in silencing while green arrows depict interactions involved in activation

Fig. 5
figure 5

Molecular structures of various nucleosome bound and free protein complexes. Histones H3 and H4 are coloured dark and light blue while the histones H2A and H2B are coloured dark and light green. A Structure of a nucleosome bound by the winged helix (in red) of histone H1. B Structure of the winged helix domain of Sir3. C Structure of a nucleosome bound by the Sir3 BAH domain (in red). D Structure of a nucleosome bound by the RSC complex (BAH domain containing subunits of RSC are shown in red)

Fig. 6
figure 6

Schematic of the effects of clustering of the silenced chromatin domains. Red dots represent Sir proteins, green ovals represent TFs, rectangles represent silencers. Trapezoids represent enhancers

Fig. 7
figure 7

Relationship between stability of silencing and the inheritance of expression states through the cell cycle. Red lines highlight silenced state while the green lines reflect the active state. Green trails indicate transcription. Blue line represents phases of the cell cycle

Fig. 8
figure 8

Hysteresis in gene silencing. An active domain requires deacetylation of 75% of the nucleosomes to become silenced. A silent domain requires acetylation of 75% or more nucleosomes to lose silencing. Red circles denote unacetylated histones and green circles denote acetylated histones. Moon shapes denote Sir proteins; gray rectangles are silencers and black rectangles are enhancers and promoters

Data availability

No datasets were generated or analysed during the current study.

References

  1. Kim S, Wysocka J. Deciphering the multi-scale, quantitative cis-regulatory code. Mol Cell. 2023;83:373–92. https://doi.org/10.1016/j.molcel.2022.12.032.

    Article  CAS  PubMed  Google Scholar 

  2. Pelechano V, Chavez S, Perez-Ortin JE. A complete set of nascent transcription rates for yeast genes. PLoS ONE. 2010;5:e15442. https://doi.org/10.1371/journal.pone.0015442.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Miller C, Schwalb B, Maier K, Schulz D, Dumcke S, Zacher B, Mayer A, Sydow J, Marcinowski L, Dolken L, et al. Dynamic transcriptome analysis measures rates of mRNA synthesis and decay in yeast. Mol Syst Biol. 2011;7:458. https://doi.org/10.1038/msb.2010.112.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Gasch AP, Spellman PT, Kao CM, Carmel-Harel O, Eisen MB, Storz G, Botstein D, Brown PO. Genomic expression programs in the response of yeast cells to environmental changes. Mol Biol Cell. 2000;11:4241–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Hahn S, Young ET. Transcriptional regulation in Saccharomyces cerevisiae: transcription factor regulation and function, mechanisms of initiation, and roles of activators and coactivators. Genetics. 2011;189:705–36. https://doi.org/10.1534/genetics.111.127019.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Rando OJ, Winston F. Chromatin and transcription in yeast. Genetics. 2012;190:351–87. https://doi.org/10.1534/genetics.111.132266.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Gaston K, Jayaraman PS. Transcriptional repression in eukaryotes: repressors and repression mechanisms. Cell Mol Life Sci. 2003;60:721–41. https://doi.org/10.1007/s00018-003-2260-3.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  8. Arnosti DN. Soft repression and chromatin modification by conserved transcriptional corepressors. Enzymes. 2023;53:69–96. https://doi.org/10.1016/bs.enz.2023.08.001.

    Article  PubMed  Google Scholar 

  9. Gartenberg MR, Smith JS. The nuts and bolts of transcriptionally silent chromatin in Saccharomyces cerevisiae. Genetics. 2016;203:1563–99. https://doi.org/10.1534/genetics.112.145243.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Allshire RC, Madhani HD. Ten principles of heterochromatin formation and function. Nat Rev Mol Cell Biol. 2018;19:229–44. https://doi.org/10.1038/nrm.2017.119.

    Article  CAS  PubMed  Google Scholar 

  11. Bi X. Heterochromatin structure: lessons from the budding yeast. IUBMB Life. 2014;66:657–66. https://doi.org/10.1002/iub.1322.

    Article  CAS  PubMed  Google Scholar 

  12. Kueng S, Oppikofer M, Gasser SM. SIR proteins and the assembly of silent chromatin in budding yeast. Annu Rev Genet. 2013;47:275–306. https://doi.org/10.1146/annurev-genet-021313-173730.

    Article  CAS  PubMed  Google Scholar 

  13. Grunstein M, Gasser SM. Epigenetics in Saccharomyces cerevisiae. Cold Spring Harb Perspect Biol. 2013;5. https://doi.org/10.1101/cshperspect.a017491.

  14. Trojer P, Reinberg D. Facultative heterochromatin: is there a distinctive molecular signature? Mol Cell. 2007;28:1–13.

    Article  CAS  PubMed  Google Scholar 

  15. Schuettengruber B, Bourbon HM, Di Croce L, Cavalli G. Genome regulation by Polycomb and trithorax: 70 years and counting. Cell. 2017;171:34–57. https://doi.org/10.1016/j.cell.2017.08.002.

    Article  CAS  PubMed  Google Scholar 

  16. Blackledge NP, Klose RJ. The molecular principles of gene regulation by polycomb repressive complexes. Nat Rev Mol Cell Biol. 2021;22:815–33. https://doi.org/10.1038/s41580-021-00398-y.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Uckelmann M, Davidovich C. Chromatin compaction by polycomb group proteins revisited. Curr Opin Struct Biol. 2024;86:102806. https://doi.org/10.1016/j.sbi.2024.102806.

    Article  CAS  PubMed  Google Scholar 

  18. Kumar A, Kono H. Heterochromatin protein 1 (HP1): interactions with itself and chromatin components. Biophys Rev. 2020;12:387–400. https://doi.org/10.1007/s12551-020-00663-y.

    Article  PubMed  PubMed Central  Google Scholar 

  19. von Hippel PH. From simple DNA-protein interactions to the macromolecular machines of gene expression. Annu Rev Biophys Biomol Struct. 2007;36:79–105. https://doi.org/10.1146/annurev.biophys.34.040204.144521.

    Article  CAS  Google Scholar 

  20. Biggin MD. Animal transcription networks as highly connected, quantitative continua. Dev Cell. 2011;21:611–26. https://doi.org/10.1016/j.devcel.2011.09.008.

    Article  CAS  PubMed  Google Scholar 

  21. Wagh K, Stavreva DA, Upadhyaya A, Hager GL. Transcription factor dynamics: one molecule at a time. Annu Rev Cell Dev Biol. 2023;39:277–305. https://doi.org/10.1146/annurev-cellbio-022823-013847.

    Article  CAS  PubMed  Google Scholar 

  22. Larson DR. What do expression dynamics tell us about the mechanism of transcription? Curr Opin Genet Dev. 2011;21:591–9. https://doi.org/10.1016/j.gde.2011.07.010.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Suter DM. Transcription factors and DNA play hide and seek. Trends Cell Biol. 2020;30:491–500. https://doi.org/10.1016/j.tcb.2020.03.003.

    Article  CAS  PubMed  Google Scholar 

  24. Jana T, Brodsky S, Barkai N. Speed-specificity trade-offs in the transcription factors search for their genomic binding sites. Trends Genet. 2021;37:421–32. https://doi.org/10.1016/j.tig.2020.12.001.

    Article  CAS  PubMed  Google Scholar 

  25. Carminati M, Vecchia L, Stoos L, Thoma NH. Pioneer factors: emerging rules of engagement for transcription factors on chromatinized DNA. Curr Opin Struct Biol. 2024;88:102875. https://doi.org/10.1016/j.sbi.2024.102875.

    Article  CAS  PubMed  Google Scholar 

  26. Hnisz D, Shrinivas K, Young RA, Chakraborty AK, Sharp PA. Phase Sep Model Transcriptional Control Cell. 2017;169:13–23. https://doi.org/10.1016/j.cell.2017.02.007.

    Article  CAS  Google Scholar 

  27. Weingarten-Gabbay S, Segal E. The grammar of transcriptional regulation. Hum Genet. 2014;133:701–11. https://doi.org/10.1007/s00439-013-1413-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Friedman N, Rando OJ. Epigenomics and the structure of the living genome. Genome Res. 2015;25:1482–90. https://doi.org/10.1101/gr.190165.115.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Lai WKM, Pugh BF. Understanding nucleosome dynamics and their links to gene expression and DNA replication. Nat Rev Mol Cell Biol. 2017;18:548–62. https://doi.org/10.1038/nrm.2017.47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Cramer P. Organization and regulation of gene transcription. Nature. 2019;573:45–54. https://doi.org/10.1038/s41586-019-1517-4.

    Article  CAS  PubMed  Google Scholar 

  31. Brahma S, Henikoff S. Epigenome Regulation by Dynamic Nucleosome Unwrapping. Trends Biochem Sci. 2020;45:13–26. https://doi.org/10.1016/j.tibs.2019.09.003.

    Article  CAS  PubMed  Google Scholar 

  32. de Jonge WJ, Patel HP, Meeussen JVW, Lenstra TL. Following the tracks: how transcription factor binding dynamics control transcription. Biophys J. 2022;121:1583–92. https://doi.org/10.1016/j.bpj.2022.03.026.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Boeger H. Kinetic proofreading. Annu Rev Biochem. 2022;91:423–47. https://doi.org/10.1146/annurev-biochem-040320-103630.

    Article  CAS  PubMed  Google Scholar 

  34. Brodsky S, Jana T, Barkai N. Order through disorder: the role of intrinsically disordered regions in transcription factor binding specificity. Curr Opin Struct Biol. 2021;71:110–5. https://doi.org/10.1016/j.sbi.2021.06.011.

    Article  CAS  PubMed  Google Scholar 

  35. Darzacq X, Tjian R. Weak multivalent biomolecular interactions: a strength versus numbers tug of war with implications for phase partitioning. RNA. 2022;28:48–51. https://doi.org/10.1261/rna.079004.121.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Koşar Z, Erbaş A. Can the concentration of a transcription factor affect gene expression? Front Soft Matter. 2022;2. https://doi.org/10.3389/frsfm.2022.914494.

  37. Rodriguez J, Larson DR. Transcription in living cells: Molecular mechanisms of bursting. Annu Rev Biochem. 2020;89:189–212. https://doi.org/10.1146/annurev-biochem-011520-105250.

    Article  CAS  PubMed  Google Scholar 

  38. Panigrahi A, O’Malley BW. Mechanisms of enhancer action: the known and the unknown. Genome Biol. 2021;22:108. https://doi.org/10.1186/s13059-021-02322-1.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Lammers NC, Kim YJ, Zhao J, Garcia HG. A matter of time: using dynamics and theory to uncover mechanisms of transcriptional bursting. Curr Opin Cell Biol. 2020;67:147–57. https://doi.org/10.1016/j.ceb.2020.08.001.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Michael AK, Thoma NH. Reading the chromatinized genome. Cell. 2021;184:3599–611. https://doi.org/10.1016/j.cell.2021.05.029.

    Article  CAS  PubMed  Google Scholar 

  41. Wang Y, Ni T, Wang W, Liu F. Gene transcription in bursting: a unified mode for realizing accuracy and stochasticity. Biol Rev Camb Philos Soc. 2019;94:248–58. https://doi.org/10.1111/brv.12452.

    Article  PubMed  Google Scholar 

  42. Tunnacliffe E, Chubb JR. What is a Transcriptional Burst? Trends Genet. 2020;36:288–97. https://doi.org/10.1016/j.tig.2020.01.003.

    Article  CAS  PubMed  Google Scholar 

  43. Otto G. Enhancers and promoters regulate burst kinetics. Nat Rev Mol Cell Biol. 2019;20:134–5. https://doi.org/10.1038/s41580-019-0100-z.

    Article  CAS  PubMed  Google Scholar 

  44. Corrigan AM, Tunnacliffe E, Cannon D, Chubb JR. A continuum model of transcriptional bursting. Elife. 2016;5. https://doi.org/10.7554/eLife.13051.

  45. Eck E, Liu J, Kazemzadeh-Atoufi M, Ghoreishi S, Blythe SA, Garcia HG. Quantitative dissection of transcription in development yields evidence for transcription-factor-driven chromatin accessibility. Elife. 2020;9. https://doi.org/10.7554/eLife.56429.

  46. Estrada J, Wong F, DePace A, Gunawardena J. Information Integration and Energy expenditure in Gene Regulation. Cell. 2016;166:234–44. https://doi.org/10.1016/j.cell.2016.06.012.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Boeger H, Shelansky R, Patel H, Brown CR. From structural variation of gene molecules to Chromatin Dynamics and Transcriptional Bursting. Genes (Basel). 2015;6:469–83. https://doi.org/10.3390/genes6030469.

    Article  CAS  PubMed  Google Scholar 

  48. Miller JA, Widom J. Collaborative competition mechanism for gene activation in vivo. Mol Cell Biol. 2003;23:1623–32. https://doi.org/10.1128/MCB.23.5.1623-1632.2003.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Struhl K, Segal E. Determinants of nucleosome positioning. Nat Struct Mol Biol. 2013;20:267–73. https://doi.org/10.1038/nsmb.2506.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  50. Tims HS, Gurunathan K, Levitus M, Widom J. Dynamics of nucleosome invasion by DNA binding proteins. J Mol Biol. 2011;411:430–48. https://doi.org/10.1016/j.jmb.2011.05.044.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Bowman GD, Poirier MG. Post-translational modifications of histones that influence nucleosome dynamics. Chem Rev. 2015;115:2274–95. https://doi.org/10.1021/cr500350x.

    Article  CAS  PubMed  Google Scholar 

  52. Luo Y, North JA, Rose SD, Poirier MG. Nucleosomes accelerate transcription factor dissociation. Nucleic Acids Res. 2014;42:3017–27. https://doi.org/10.1093/nar/gkt1319.

    Article  CAS  PubMed  Google Scholar 

  53. Donovan BT, Luo Y, Meng Z, Poirier MG. The nucleosome unwrapping free energy landscape defines distinct regions of transcription factor accessibility and kinetics. Nucleic Acids Res. 2023;51:1139–53. https://doi.org/10.1093/nar/gkac1267.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Kubik S, Bruzzone MJ, Shore D. (2017). Establishing nucleosome architecture and stability at promoters: Roles of pioneer transcription factors and the RSC chromatin remodeler. Bioessays 39. https://doi.org/10.1002/bies.201600237

  55. Chereji RV, Clark DJ. Major determinants of Nucleosome Positioning. Biophys J. 2018;114:2279–89. https://doi.org/10.1016/j.bpj.2018.03.015.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Jansen A, Verstrepen KJ. Nucleosome positioning in Saccharomyces cerevisiae. Microbiol Mol Biol Rev. 2011;75:301–20. https://doi.org/10.1128/MMBR.00046-10.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Oberbeckmann E, Wolff M, Krietenstein N, Heron M, Ellins JL, Schmid A, Krebs S, Blum H, Gerland U, Korber P. Absolute nucleosome occupancy map for the Saccharomyces cerevisiae genome. Genome Res. 2019;29:1996–2009. https://doi.org/10.1101/gr.253419.119.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Becker PB, Workman JL. Nucleosome remodeling and epigenetics. Cold Spring Harb Perspect Biol. 2013;5. https://doi.org/10.1101/cshperspect.a017905.

  59. Kubik S, Bruzzone MJ, Challal D, Dreos R, Mattarocci S, Bucher P, Libri D, Shore D. Opposing chromatin remodelers control transcription initiation frequency and start site selection. Nat Struct Mol Biol. 2019;26:744–54. https://doi.org/10.1038/s41594-019-0273-3.

    Article  CAS  PubMed  Google Scholar 

  60. Ganguli D, Chereji RV, Iben JR, Cole HA, Clark DJ. RSC-dependent constructive and destructive interference between opposing arrays of phased nucleosomes in yeast. Genome Res. 2014;24:1637–49. https://doi.org/10.1101/gr.177014.114.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Kubik S, O’Duibhir E, de Jonge WJ, Mattarocci S, Albert B, Falcone JL, Bruzzone MJ, Holstege FCP, Shore D. Sequence-Directed Action of RSC Remodeler and General Regulatory Factors modulates + 1 nucleosome position to facilitate transcription. Mol Cell. 2018;71:89–e102105. https://doi.org/10.1016/j.molcel.2018.05.030.

    Article  CAS  PubMed  Google Scholar 

  62. North JA, Javaid S, Ferdinand MB, Chatterjee N, Picking JW, Shoffner M, Nakkula RJ, Bartholomew B, Ottesen JJ, Fishel R, Poirier MG. Phosphorylation of histone H3(T118) alters nucleosome dynamics and remodeling. Nucleic Acids Res. 2011;39:6465–74. https://doi.org/10.1093/nar/gkr304.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Kim JM, Visanpattanasin P, Jou V, Liu S, Tang X, Zheng Q, Li KY, Snedeker J, Lavis LD, Lionnet T, Wu C. Single-molecule imaging of chromatin remodelers reveals role of ATPase in promoting fast kinetics of target search and dissociation from chromatin. Elife. 2021;10. https://doi.org/10.7554/eLife.69387.

  64. Yaakov G, Jonas F, Barkai N. Measurement of histone replacement dynamics with genetically encoded exchange timers in yeast. Nat Biotechnol. 2021;39:1434–43. https://doi.org/10.1038/s41587-021-00959-8.

    Article  CAS  PubMed  Google Scholar 

  65. Abril-Garrido J, Dienemann C, Grabbe F, Velychko T, Lidschreiber M, Wang H, Cramer P. Structural basis of transcription reduction by a promoter-proximal + 1 nucleosome. Mol Cell. 2023;83:1798–e18091797. https://doi.org/10.1016/j.molcel.2023.04.011.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Brown CR, Mao C, Falkovskaia E, Jurica MS, Boeger H. Linking stochastic fluctuations in chromatin structure and gene expression. PLoS Biol. 2013;11:e1001621. https://doi.org/10.1371/journal.pbio.1001621.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Rossi MJ, Kuntala PK, Lai WKM, Yamada N, Badjatia N, Mittal C, Kuzu G, Bocklund K, Farrell NP, Blanda TR, et al. A high-resolution protein architecture of the budding yeast genome. Nature. 2021;592:309–14. https://doi.org/10.1038/s41586-021-03314-8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Stoeber S, Godin H, Xu C, Bai L. Pioneer factors: nature or nurture? Crit Rev Biochem Mol Biol. 2024;1–15. https://doi.org/10.1080/10409238.2024.2355885.

  69. Wu K, Dhillon N, Bajor A, Abrahamsson S, Kamakaka RT. Yeast heterochromatin stably silences only weak regulatory elements by altering burst duration. Cell Rep. 2024;43:113983. https://doi.org/10.1016/j.celrep.2024.113983.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Cheng TH, Gartenberg MR. Yeast heterochromatin is a dynamic structure that requires silencers continuously. Genes Dev. 2000;14:452–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Rusche LN, Kirchmaier AL, Rine J. The establishment, inheritance, and function of silenced chromatin in Saccharomyces cerevisiae. Annu Rev Biochem; 2003.

  72. Irlbacher H, Franke J, Manke T, Vingron M, Ehrenhofer-Murray AE. Control of replication initiation and heterochromatin formation in Saccharomyces cerevisiae by a regulator of meiotic gene expression. Genes Dev. 2005;19:1811–22. https://doi.org/10.1101/gad.334805.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Chien CT, Buck S, Sternglanz R, Shore D. Targeting of SIR1 protein establishes transcriptional silencing at HM loci and telomeres in yeast. Cell. 1993;75:531–41. https://doi.org/10.1016/0092-8674(93)90387-6.

    Article  CAS  PubMed  Google Scholar 

  74. Fox CA, Ehrenhofer-Murray AE, Loo S, Rine J. The origin recognition complex, SIR1, and the S phase requirement for silencing. Science. 1997;276:1547–51.

    Article  CAS  PubMed  Google Scholar 

  75. Dhillon N, Kamakaka RT. A histone variant, Htz1p, and a Sir1p-like protein, Esc2p, mediate silencing at HMR. Mol Cell. 2000;6:769–80. https://doi.org/10.1016/s1097-2765(00)00076-9.

    Article  CAS  PubMed  Google Scholar 

  76. Rusche LN, Kirchmaier AL, Rine J. Ordered nucleation and spreading of silenced chromatin in Saccharomyces cerevisiae. Mol Biol Cell. 2002;13:2207–22. https://doi.org/10.1091/mbc.e02-03-0175.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Luo K, Vega-Palas MA, Grunstein M. Rap1-Sir4 binding independent of other Sir, yKu, or histone interactions initiates the assembly of telomeric heterochromatin in yeast. Genes Dev. 2002;16:1528–39. https://doi.org/10.1101/gad.988802.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Saxton DS, Rine J. Distinct silencer states generate epigenetic states of heterochromatin. Mol Cell. 2022;82:3566–e35793565. https://doi.org/10.1016/j.molcel.2022.08.002.

    Article  CAS  PubMed  Google Scholar 

  79. Bi X, Braunstein M, Shei GJ, Broach JR. The yeast HML I silencer defines a heterochromatin domain boundary by directional establishment of silencing. Proc Natl Acad Sci U S A. 1999;96:11934–9. https://doi.org/10.1073/pnas.96.21.11934.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Krietenstein N, Wal M, Watanabe S, Park B, Peterson CL, Pugh BF, Korber P. Genomic Nucleosome Organization Reconstituted with pure proteins. Cell. 2016;167:709–e721712. https://doi.org/10.1016/j.cell.2016.09.045.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Zou Y, Yu Q, Bi X. Asymmetric positioning of nucleosomes and directional establishment of transcriptionally silent chromatin by Saccharomyces cerevisiae silencers. Mol Cell Biol. 2006;26:7806–19. https://doi.org/10.1128/MCB.01197-06.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  82. Hecht A, Laroche T, Strahl-Bolsinger S, Gasser SM, Grunstein M. Histone H3 and H4 N-termini interact with SIR3 and SIR4 proteins: a molecular model for the formation of heterochromatin in yeast. Cell. 1995;80:583–92.

    Article  CAS  PubMed  Google Scholar 

  83. Hecht A, Strahl-Bolsinger S, Grunstein M. Spreading of transcriptional repressor SIR3 from telomeric heterochromatin. Nature. 1996;383:92–6. https://doi.org/10.1038/383092a0.

    Article  CAS  PubMed  Google Scholar 

  84. Strahl-Bolsinger S, Hecht A, Luo K, Grunstein M. SIR2 and SIR4 interactions differ in core and extended telomeric heterochromatin in yeast. Genes Dev. 1997;11:83–93. https://doi.org/10.1101/gad.11.1.83.

    Article  CAS  PubMed  Google Scholar 

  85. Arnosti DN, Kulkarni MM. Transcriptional enhancers: Intelligent enhanceosomes or flexible billboards? J Cell Biochem. 2005;94:890–8. https://doi.org/10.1002/jcb.20352.

    Article  CAS  PubMed  Google Scholar 

  86. Panne D. The enhanceosome. Curr Opin Struct Biol. 2008;18:236–42.

    Article  CAS  PubMed  Google Scholar 

  87. Li X, Gerber SA, Rudner AD, Beausoleil SA, Haas W, Villen J, Elias JE, Gygi SP. Large-scale phosphorylation analysis of alpha-factor-arrested Saccharomyces cerevisiae. J Proteome Res. 2007;6:1190–7. https://doi.org/10.1021/pr060559j.

    Article  CAS  PubMed  Google Scholar 

  88. Albuquerque CP, Smolka MB, Payne SH, Bafna V, Eng J, Zhou H. A multidimensional chromatography technology for in-depth phosphoproteome analysis. Mol Cell Proteom. 2008;7:1389–96. https://doi.org/10.1074/mcp.M700468-MCP200.

    Article  CAS  Google Scholar 

  89. Shei GJ, Broach JR. Yeast silencers can act as orientation-dependent gene inactivation centers that respond to environmental signals. Mol Cell Biol. 1995;15:3496–506.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Wu K, Dhillon N, Du K, Kamakaka RT. Measuring the buffering capacity of gene silencing in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A. 2021;118. https://doi.org/10.1073/pnas.2111841118.

  91. Palacios DeBeer MA, Muller U, Fox CA. Differential DNA affinity specifies roles for the origin recognition complex in budding yeast heterochromatin. Genes Dev. 2003;17:1817–22.

    Article  PubMed  PubMed Central  Google Scholar 

  92. Lickwar CR, Mueller F, Hanlon SE, McNally JG, Lieb JD. Genome-wide protein-DNA binding dynamics suggest a molecular clutch for transcription factor function. Nature. 2012;484:251–5. https://doi.org/10.1038/nature10985.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Jia S, Noma K, Grewal SI. RNAi-independent heterochromatin nucleation by the stress-activated ATF/CREB family proteins. Science. 2004;304:1971–6. https://doi.org/10.1126/science.1099035.

    Article  CAS  PubMed  Google Scholar 

  94. Kanoh J, Sadaie M, Urano T, Ishikawa F. Telomere binding protein Taz1 establishes Swi6 heterochromatin independently of RNAi at telomeres. Curr Biol. 2005;15:1808–19. https://doi.org/10.1016/j.cub.2005.09.041.

    Article  CAS  PubMed  Google Scholar 

  95. Simon J, Chiang A, Bender W, Shimell MJ, O’Connor M. Elements of the Drosophila bithorax complex that mediate repression by polycomb group products. Dev Biol. 1993;158:131–44. https://doi.org/10.1006/dbio.1993.1174.

    Article  CAS  PubMed  Google Scholar 

  96. Simon JA, Tamkun JW. Programming off and on states in chromatin: mechanisms of polycomb and trithorax group complexes. Curr Opin Genet Dev. 2002;12:210–8.

    Article  CAS  PubMed  Google Scholar 

  97. Entrevan M, Schuettengruber B, Cavalli G. Regulation of Genome Architecture and function by Polycomb Proteins. Trends Cell Biol. 2016;26:511–25. https://doi.org/10.1016/j.tcb.2016.04.009.

    Article  CAS  PubMed  Google Scholar 

  98. Simon J. Locking in stable states of gene expression: transcriptional control during Drosophila development. Curr Opin Cell Biol. 1995;7:376–85.

    Article  CAS  PubMed  Google Scholar 

  99. Danzer JR, Wallrath LL. Mechanisms of HP1-mediated gene silencing in Drosophila. Development. 2004;131:3571–80. https://doi.org/10.1242/dev.01223.

    Article  CAS  PubMed  Google Scholar 

  100. Blackledge NP, Farcas AM, Kondo T, King HW, McGouran JF, Hanssen LLP, Ito S, Cooper S, Kondo K, Koseki Y, et al. Variant PRC1 complex-dependent H2A ubiquitylation drives PRC2 recruitment and polycomb domain formation. Cell. 2014;157:1445–59. https://doi.org/10.1016/j.cell.2014.05.004.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Grewal SI, Jia S. Heterochromatin revisited. Nat Rev Genet. 2007;8:35–46. https://doi.org/10.1038/nrg2008.

    Article  CAS  PubMed  Google Scholar 

  102. Al-Sady B, Madhani HD, Narlikar GJ. Division of labor between the chromodomains of HP1 and Suv39 methylase enables coordination of heterochromatin spread. Mol Cell. 2013;51:80–91. https://doi.org/10.1016/j.molcel.2013.06.013.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Zhang K, Mosch K, Fischle W, Grewal SI. Roles of the Clr4 methyltransferase complex in nucleation, spreading and maintenance of heterochromatin. Nat Struct Mol Biol. 2008;15:381–8. https://doi.org/10.1038/nsmb.1406.

    Article  CAS  PubMed  Google Scholar 

  104. Ellahi A, Thurtle DM, Rine J. The chromatin and Transcriptional Landscape of native Saccharomyces cerevisiae telomeres and Subtelomeric domains. Genetics. 2015;200:505–21. https://doi.org/10.1534/genetics.115.175711.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  105. Thurtle DM, Rine J. The molecular topography of silenced chromatin in Saccharomyces cerevisiae. Genes Dev. 2014;28:245–58. https://doi.org/10.1101/gad.230532.113.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Hocher A, Ruault M, Kaferle P, Descrimes M, Garnier M, Morillon A, Taddei A. Expanding heterochromatin reveals discrete subtelomeric domains delimited by chromatin landscape transitions. Genome Res. 2018;28:1867–81. https://doi.org/10.1101/gr.236554.118.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Weiss K, Simpson RT. High-resolution structural analysis of chromatin at specific loci: Saccharomyces cerevisiae silent mating type locus HMLalpha. Mol Cell Biol. 1998;18:5392–403.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Ravindra A, Weiss K, Simpson RT. High-resolution structural analysis of chromatin at specific loci: Saccharomyces cerevisiae silent mating-type locus HMRa. Mol Cell Biol. 1999;19:7944–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Chereji RV, Ocampo J, Clark DJ. MNase-Sensitive complexes in yeast: nucleosomes and non-histone barriers. Mol Cell. 2017;65:565–e577563. https://doi.org/10.1016/j.molcel.2016.12.009.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Brothers M, Rine J. Distinguishing between recruitment and spread of silent chromatin structures in Saccharomyces cerevisiae. Elife. 2022;11. https://doi.org/10.7554/eLife.75653.

  111. Valenzuela L, Dhillon N, Dubey RN, Gartenberg MR, Kamakaka RT. Long-range communication between the silencers of HMR. Mol Cell Biol. 2008;28:1924–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  112. Bauer SL, Grochalski TNT, Smialowska A, Astrom SU. Sir2 and Reb1 antagonistically regulate nucleosome occupancy in subtelomeric X-elements and repress TERRAs by distinct mechanisms. PLoS Genet. 2022;18:e1010419. https://doi.org/10.1371/journal.pgen.1010419.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Lebrun E, Revardel E, Boscheron C, Li R, Gilson E, Fourel G. Protosilencers in Saccharomyces cerevisiae subtelomeric regions. Genetics. 2001;158:167–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Fourel G, Revardel E, Koering CE, Gilson E. Cohabitation of insulators and silencing elements in yeast subtelomeric regions. Embo J. 1999;18:2522–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Fourel G, Magdinier F, Gilson E. Insulator dynamics and the setting of chromatin domains. BioEssays. 2004;26:523–32.

    Article  CAS  PubMed  Google Scholar 

  116. Buck SW, Shore D. Action of a RAP1 carboxy-terminal silencing domain reveals an underlying competetion between HMR and telomeres in yeast. Genes Dev. 1995;9:370–84.

    Article  CAS  PubMed  Google Scholar 

  117. Maillet L, Boscheron C, Gotta M, Marcand S, Gilson E, Gasser SM. Evidence for silencing compartments within the yeast nucleus: a role for telomere proximity and sir protein concentration in silencer-mediated repression. Genes Dev. 1996;10:1796–811. https://doi.org/10.1101/gad.10.14.1796.

    Article  CAS  PubMed  Google Scholar 

  118. Johnson A, Li G, Sikorski TW, Buratowski S, Woodcock CL, Moazed D. Reconstitution of heterochromatin-dependent transcriptional gene silencing. Mol Cell. 2009;35:769–81. https://doi.org/10.1016/j.molcel.2009.07.030.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Gottschling DE. Telomere-proximal DNA in Saccharomyces cerevisiae is refractory to methyltransferase activity in vivo. Proc Natl Acad Sci U S A. 1992;89:4062–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Marr LT, Ocampo J, Clark DJ, Hayes JJ. Global histone protein surface accessibility in yeast indicates a uniformly loosely packed genome with canonical nucleosomes. Epigenetics Chromatin. 2021;14. https://doi.org/10.1186/s13072-020-00381-5.

  121. Ghidelli S, Donze D, Dhillon N, Kamakaka RT. Sir2p exists in two nucleosome-binding complexes with distinct deacetylase activities. Embo J. 2001;20:4522–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Oppikofer M, Kueng S, Keusch JJ, Hassler M, Ladurner AG, Gut H, Gasser SM. Dimerization of Sir3 via its C-terminal winged helix domain is essential for yeast heterochromatin formation. EMBO J. 2013;32:437–49. https://doi.org/10.1038/emboj.2012.343.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Mansisidor AR, Risca VI. Chromatin accessibility: methods, mechanisms, and biological insights. Nucleus. 2022;13:236–76. https://doi.org/10.1080/19491034.2022.2143106.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Chen L, Widom J. Mechanism of transcriptional silencing in yeast. Cell. 2005;120:37–48.

    Article  CAS  PubMed  Google Scholar 

  125. Steakley DL, Rine J. On the mechanism of gene silencing in Saccharomyces cerevisiae. G3 (Bethesda). 2015;5:1751–63. https://doi.org/10.1534/g3.115.018515.

    Article  CAS  PubMed  Google Scholar 

  126. Aparicio OM, Gottschling DE. Overcoming telomeric silencing: a trans-activator competes to establish gene expression in a cell cycle-dependent way. Genes Dev. 1994;8:1133–46.

    Article  CAS  PubMed  Google Scholar 

  127. Wang X, Bryant G, Zhao A, Ptashne M. Nucleosome avidities and transcriptional silencing in yeast. Curr Biol. 2015;25:1215–20. https://doi.org/10.1016/j.cub.2015.03.004.

    Article  CAS  PubMed  Google Scholar 

  128. Sekinger EA, Gross DS. Silenced chromatin is permissive to activator binding and PIC recruitment. Cell. 2001;105:403–14.

    Article  CAS  PubMed  Google Scholar 

  129. Rufiange A, Jacques PE, Bhat W, Robert F, Nourani A. Genome-wide replication-independent histone H3 exchange occurs predominantly at promoters and implicates H3 K56 acetylation and Asf1. Mol Cell. 2007;27:393–405.

    Article  CAS  PubMed  Google Scholar 

  130. Dion MF, Kaplan T, Kim M, Buratowski S, Friedman N, Rando OJ. Dynamics of replication-independent histone turnover in budding yeast. Science. 2007;315:1405–8.

    Article  CAS  PubMed  Google Scholar 

  131. Manning BJ, Peterson CL. Direct interactions promote eviction of the Sir3 heterochromatin protein by the SWI/SNF chromatin remodeling enzyme. Proc Natl Acad Sci U S A. 2014;111:17827–32. https://doi.org/10.1073/pnas.1420096111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Saxton DS, Rine J. Nucleosome positioning regulates the Establishment, Stability, and inheritance of Heterochromatin in Saccharomyces cerevisiae. Proc Natl Acad Sci U S A. 2020;117:27493–501. https://doi.org/10.1073/pnas.2004111117.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Enderle D, Beisel C, Stadler MB, Gerstung M, Athri P, Paro R. Polycomb preferentially targets stalled promoters of coding and noncoding transcripts. Genome Res. 2011;21:216–26. https://doi.org/10.1101/gr.114348.110.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Shao Z, Raible F, Mollaaghababa R, Guyon JR, Wu CT, Bender W, Kingston RE. Stabilization of chromatin structure by PRC1, a polycomb complex. Cell. 1999;98:37–46.

    Article  CAS  PubMed  Google Scholar 

  135. Kadoch C, Williams RT, Calarco JP, Miller EL, Weber CM, Braun SM, Pulice JL, Chory EJ, Crabtree GR. Dynamics of BAF-Polycomb complex opposition on heterochromatin in normal and oncogenic states. Nat Genet. 2017;49:213–22. https://doi.org/10.1038/ng.3734.

    Article  CAS  PubMed  Google Scholar 

  136. Tamkun JW, Deuring R, Scott MP, Kissinger M, Pattatucci AM, Kaufman TC, Kennison JA. Brahma: a regulator of Drosophila homeotic genes structurally related to the yeast transcriptional activator SNF2/SWI2. Cell. 1992;68:561–72. https://doi.org/10.1016/0092-8674(92)90191-e.

    Article  CAS  PubMed  Google Scholar 

  137. Tie F, Banerjee R, Stratton CA, Prasad-Sinha J, Stepanik V, Zlobin A, Diaz MO, Scacheri PC, Harte PJ. CBP-mediated acetylation of histone H3 lysine 27 antagonizes Drosophila polycomb silencing. Development. 2009;136:3131–41. https://doi.org/10.1242/dev.037127.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Seman M, Levashkevich A, Larkin A, Huang F, Ragunathan K. Uncoupling the distinct functions of HP1 proteins during heterochromatin establishment and maintenance. Cell Rep. 2023;42:113428. https://doi.org/10.1016/j.celrep.2023.113428.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Garcia JF, Dumesic PA, Hartley PD, El-Samad H, Madhani HD. Combinatorial, site-specific requirement for heterochromatic silencing factors in the elimination of nucleosome-free regions. Genes Dev. 2010;24:1758–71. https://doi.org/10.1101/gad.1946410.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Deal RB, Henikoff JG, Henikoff S. Genome-wide kinetics of nucleosome turnover determined by metabolic labeling of histones. Science. 2010;328:1161–4. https://doi.org/10.1126/science.1186777.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Holla S, Dhakshnamoorthy J, Folco HD, Balachandran V, Xiao H, Sun LL, Wheeler D, Zofall M, Grewal SIS. Positioning Heterochromatin at the Nuclear Periphery suppresses histone turnover to promote epigenetic inheritance. Cell. 2020;180:150–e164115. https://doi.org/10.1016/j.cell.2019.12.004.

    Article  CAS  PubMed  Google Scholar 

  142. Aygun O, Mehta S, Grewal SI. HDAC-mediated suppression of histone turnover promotes epigenetic stability of heterochromatin. Nat Struct Mol Biol. 2013;20:547–54. https://doi.org/10.1038/nsmb.2565.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Chang JF, Hall BE, Tanny JC, Moazed D, Filman D, Ellenberger T. Structure of the coiled-coil dimerization motif of Sir4 and its interaction with Sir3. Structure. 2003;11:637–49. https://doi.org/10.1016/s0969-2126(03)00093-5.

    Article  CAS  PubMed  Google Scholar 

  144. Min J, Landry J, Sternglanz R, Xu RM. Crystal structure of a SIR2 homolog-NAD complex. Cell. 2001;105:269–79.

    Article  CAS  PubMed  Google Scholar 

  145. Hsu HC, Wang CL, Wang M, Yang N, Chen Z, Sternglanz R, Xu RM. Structural basis for allosteric stimulation of Sir2 activity by Sir4 binding. Genes Dev. 2013;27:64–73. https://doi.org/10.1101/gad.208140.112.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Hou Z, Danzer JR, Fox CA, Keck JL. Structure of the Sir3 protein bromo adjacent homology (BAH) domain from S. Cerevisiae at 1.95 a resolution. Protein Sci. 2006;15:1182–6. https://doi.org/10.1110/ps.052061006.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Ehrentraut S, Hassler M, Oppikofer M, Kueng S, Weber JM, Mueller JW, Gasser SM, Ladurner AG, Ehrenhofer-Murray AE. Structural basis for the role of the Sir3 AAA + domain in silencing: interaction with Sir4 and unmethylated histone H3K79. Genes Dev. 2011;25:1835–46. https://doi.org/10.1101/gad.17175111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Armache KJ, Garlick JD, Canzio D, Narlikar GJ, Kingston RE. Structural basis of silencing: Sir3 BAH domain in complex with a nucleosome at 3.0 a resolution. Science. 2011;334:977–82. https://doi.org/10.1126/science.1210915.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Arnaudo N, Fernandez IS, McLaughlin SH, Peak-Chew SY, Rhodes D, Martino F. The N-terminal acetylation of Sir3 stabilizes its binding to the nucleosome core particle. Nat Struct Mol Biol. 2013;20:1119–21. https://doi.org/10.1038/nsmb.2641.

    Article  CAS  PubMed  Google Scholar 

  150. Wang F, Li G, Altaf M, Lu C, Currie MA, Johnson A, Moazed D. Heterochromatin protein Sir3 induces contacts between the amino terminus of histone H4 and nucleosomal DNA. Proc Natl Acad Sci U S A. 2013;110:8495–500. https://doi.org/10.1073/pnas.1300126110.

    Article  PubMed  PubMed Central  Google Scholar 

  151. Deshpande I, Keusch JJ, Challa K, Iesmantavicius V, Gasser SM, Gut H. The Sir4 H-BRCT domain interacts with phospho-proteins to sequester and repress yeast heterochromatin. EMBO J. 2019;38:e101744. https://doi.org/10.15252/embj.2019101744.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Liou GG, Tanny JC, Kruger RG, Walz T, Moazed D. Assembly of the SIR complex and its regulation by O-acetyl-ADP-ribose, a product of NAD-dependent histone deacetylation. Cell. 2005;121:515–27. https://doi.org/10.1016/j.cell.2005.03.035.

    Article  CAS  PubMed  Google Scholar 

  153. Zhou BR, Feng H, Kale S, Fox T, Khant H, de Val N, Ghirlando R, Panchenko AR, Bai Y. Distinct structures and dynamics of chromatosomes with different human Linker Histone Isoforms. Mol Cell. 2021;81:166–e182166. https://doi.org/10.1016/j.molcel.2020.10.038.

    Article  CAS  PubMed  Google Scholar 

  154. Dombrowski M, Engeholm M, Dienemann C, Dodonova S, Cramer P. Histone H1 binding to nucleosome arrays depends on linker DNA length and trajectory. Nat Struct Mol Biol. 2022;29:493–501. https://doi.org/10.1038/s41594-022-00768-w.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Jang S, Song JJ. The big picture of chromatin biology by cryo-EM. Curr Opin Struct Biol. 2019;58:76–87. https://doi.org/10.1016/j.sbi.2019.05.017.

    Article  CAS  PubMed  Google Scholar 

  156. Farris D, Saxton DS, Rine J. (2021). A novel allele of SIR2 reveals a heritable intermediate state of gene silencing. Genetics 218. https://doi.org/10.1093/genetics/iyab041

  157. Chambers AL, Pearl LH, Oliver AW, Downs JA. The BAH domain of Rsc2 is a histone H3 binding domain. Nucleic Acids Res. 2013;41:9168–82. https://doi.org/10.1093/nar/gkt662.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Aljahani A, Mauksch C, Oudelaar AM. The relationship between nucleosome positioning and higher-order genome folding. Curr Opin Cell Biol. 2024;89:102398. https://doi.org/10.1016/j.ceb.2024.102398.

    Article  CAS  PubMed  Google Scholar 

  159. Behrouzi R, Lu C, Currie MA, Jih G, Iglesias N, Moazed D. (2016). Heterochromatin assembly by interrupted Sir3 bridges across neighboring nucleosomes. Elife 5. https://doi.org/10.7554/eLife.17556

  160. Swygert SG, Senapati S, Bolukbasi MF, Wolfe SA, Lindsay S, Peterson CL. SIR proteins create compact heterochromatin fibers. Proc Natl Acad Sci U S A. 2018;115:12447–52. https://doi.org/10.1073/pnas.1810647115.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Ruault M, Scolari VF, Lazar-Stefanita L, Hocher A, Loiodice I, Koszul R, Taddei A. Sir3 mediates long-range chromosome interactions in budding yeast. Genome Res. 2021;31:411–25. https://doi.org/10.1101/gr.267872.120.

    Article  PubMed  PubMed Central  Google Scholar 

  162. Ye Y, Wu H, Chen K, Clapier CR, Verma N, Zhang W, Deng H, Cairns BR, Gao N, Chen Z. Structure of the RSC complex bound to the nucleosome. Science. 2019;366:838–43. https://doi.org/10.1126/science.aay0033.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Wagner FR, Dienemann C, Wang H, Stutzer A, Tegunov D, Urlaub H, Cramer P. Structure of SWI/SNF chromatin remodeller RSC bound to a nucleosome. Nature. 2020;579:448–51. https://doi.org/10.1038/s41586-020-2088-0.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Onishi M, Liou GG, Buchberger JR, Walz T, Moazed D. Role of the conserved Sir3-BAH domain in nucleosome binding and silent chromatin assembly. Mol Cell. 2007;28:1015–28. https://doi.org/10.1016/j.molcel.2007.12.004.

    Article  CAS  PubMed  Google Scholar 

  165. Fan H, Guo Y, Tsai YH, Storey AJ, Kim A, Gong W, Edmondson RD, Mackintosh SG, Li H, Byrum SD, et al. A conserved BAH module within mammalian BAHD1 connects H3K27me3 to polycomb gene silencing. Nucleic Acids Res. 2021;49:4441–55. https://doi.org/10.1093/nar/gkab210.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Palladino F, Laroche T, Gilson E, Pillus L, Gasser SM. The positioning of yeast telomeres depends on SIR3, SIR4, and the integrity of the nuclear membrane. Cold Spring Harb Symp Quant Biol. 1993;58:733–46.

    Article  CAS  PubMed  Google Scholar 

  167. Fabre E, Spichal M. (2014). Subnuclear Architecture of telomeres and subtelomeres in yeast. In SubTelomeres, E.J. Louis, and M.M. Becker, eds. (Springer), pp. 13–37.

  168. Taddei A, Hediger F, Neumann FR, Gasser SM. The function of nuclear architecture: a genetic approach. Annu Rev Genet. 2004;38:305–45.

    Article  CAS  PubMed  Google Scholar 

  169. Taddei A, Schober H, Gasser SM. The budding yeast nucleus. Cold Spring Harb Perspect Biol. 2010;2:a000612. https://doi.org/10.1101/cshperspect.a000612.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Oh J, Yeom S, Park J, Lee JS. The regional sequestration of heterochromatin structural proteins is critical to form and maintain silent chromatin. Epigenetics Chromatin. 2022;15. https://doi.org/10.1186/s13072-022-00435-w.

  171. Akhtar A, Gasser SM. The nuclear envelope and transcriptional control. Nat Rev Genet. 2007;8:507–17. https://doi.org/10.1038/nrg2122.

    Article  CAS  PubMed  Google Scholar 

  172. Tatarakis A, Behrouzi R, Moazed D. Evolving models of heterochromatin: from Foci to Liquid droplets. Mol Cell. 2017;67:725–7. https://doi.org/10.1016/j.molcel.2017.08.022.

    Article  CAS  PubMed  Google Scholar 

  173. Taddei A, Gartenberg MR, Neumann FR, Hediger F, Gasser SM. Multiple pathways tether telomeres and silent chromatin at the nuclear periphery: functional implications for sir-mediated repression. Novartis Found Symp. 2005;264:140–56. discussion 156–165, 227 – 130.

    Article  CAS  PubMed  Google Scholar 

  174. O’Kane CJ, Hyland EM. Yeast epigenetics: the inheritance of histone modification states. Biosci Rep. 2019;39. https://doi.org/10.1042/BSR20182006.

  175. Du M, Zhang Q, Bai L. Three distinct mechanisms of long-distance modulation of gene expression in yeast. PLoS Genet. 2017;13:e1006736. https://doi.org/10.1371/journal.pgen.1006736.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  176. Waterborg JH. Dynamics of histone acetylation in Saccharomyces cerevisiae. Biochemistry. 2001;40:2599–605. https://doi.org/10.1021/bi002480c.

    Article  CAS  PubMed  Google Scholar 

  177. Ho B, Baryshnikova A, Brown GW. Unification of protein abundance datasets yields a quantitative Saccharomyces cerevisiae Proteome. Cell Syst. 2018;6:192–e205193. https://doi.org/10.1016/j.cels.2017.12.004.

    Article  CAS  PubMed  Google Scholar 

  178. Renauld H, Aparicio OM, Zierath PD, Billington BL, Chhablani SK, Gottschling DE. Silent domains are assembled continuously from the telomere and are defined by promoter distance and strength, and by SIR3 dosage. Genes & Development; 1993. pp. 1133–45.

  179. Osborne EA, Dudoit S, Rine J. The establishment of gene silencing at single-cell resolution. Nat Genet. 2009;41:800–6. https://doi.org/10.1038/ng.402.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Gasser SM, Hediger F, Taddei A, Neumann FR, Gartenberg MR. The function of telomere clustering in yeast: the circe effect. Cold Spring Harb Symp Quant Biol. 2004;69:327–37. https://doi.org/10.1101/sqb.2004.69.327.

    Article  CAS  PubMed  Google Scholar 

  181. Loiodice I, Garnier M, Nikolov I, Taddei A. An Inducible System for silencing establishment reveals a stepwise mechanism in which anchoring at the Nuclear Periphery precedes heterochromatin formation. Cells. 2021;10:103390cells10112810.

    Article  Google Scholar 

  182. Rippe K. Dynamic organization of the cell nucleus. Curr Opin Genet Dev. 2007;17:373–80. https://doi.org/10.1016/j.gde.2007.08.007.

    Article  CAS  PubMed  Google Scholar 

  183. Bancaud A, Huet S, Daigle N, Mozziconacci J, Beaudouin J, Ellenberg J. Molecular crowding affects diffusion and binding of nuclear proteins in heterochromatin and reveals the fractal organization of chromatin. EMBO J. 2009;28:3785–98. https://doi.org/10.1038/emboj.2009.340.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Verschure PJ, van der Kraan I, Manders EM, Hoogstraten D, Houtsmuller AB, van Driel R. Condensed chromatin domains in the mammalian nucleus are accessible to large macromolecules. EMBO Rep. 2003;4:861–6. https://doi.org/10.1038/sj.embor.embor922.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Minton AP. How can biochemical reactions within cells differ from those in test tubes? J Cell Sci. 2006;119:2863–9. https://doi.org/10.1242/jcs.03063.

    Article  CAS  PubMed  Google Scholar 

  186. Kribelbauer JF, Rastogi C, Bussemaker HJ, Mann RS. Low-Affinity binding sites and the transcription factor specificity Paradox in Eukaryotes. Annu Rev Cell Dev Biol. 2019;35:357–79. https://doi.org/10.1146/annurev-cellbio-100617-062719.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Phair RD, Scaffidi P, Elbi C, Vecerova J, Dey A, Ozato K, Brown DT, Hager G, Bustin M, Misteli T. Global nature of dynamic protein-chromatin interactions in vivo: three-dimensional genome scanning and dynamic interaction networks of chromatin proteins. Mol Cell Biol. 2004;24:6393–402. https://doi.org/10.1128/MCB.24.14.6393-6402.2004.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  188. Dross N, Spriet C, Zwerger M, Muller G, Waldeck W, Langowski J. Mapping eGFP oligomer mobility in living cell nuclei. PLoS ONE. 2009;4:e5041. https://doi.org/10.1371/journal.pone.0005041.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Andrulis ED, Neiman AM, Zappulla DC, Sternglanz R. Perinuclear localization of chromatin facilitates transcriptional silencing. Nature. 1998;394:592–5. https://doi.org/10.1038/29100.

    Article  CAS  PubMed  Google Scholar 

  190. Taddei A, Van Houwe G, Nagai S, Erb I, van Nimwegen E, Gasser SM. The functional importance of telomere clustering: global changes in gene expression result from SIR factor dispersion. Genome Res. 2009;19:611–25. https://doi.org/10.1101/gr.083881.108.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Miele A, Bystricky K, Dekker J. (2009). Yeast silent mating type loci form heterochromatic clusters through silencer protein-dependent long-range interactions. PLoS Genet 5, e1000478.

  192. Kirkland JG, Kamakaka RT. Long-range heterochromatin association is mediated by silencing and double-strand DNA break repair proteins. J Cell Biol. 2013;201:809–26. https://doi.org/10.1083/jcb.201211105.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. McBryant SJ, Krause C, Woodcock CL, Hansen JC. The silent information regulator 3 protein, SIR3p, binds to chromatin fibers and assembles a hypercondensed chromatin architecture in the presence of salt. Mol Cell Biol. 2008;28:3563–72. https://doi.org/10.1128/MCB.01389-07.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Movilla Miangolarra A, Saxton DS, Yan Z, Rine J, Howard M. Two-way feedback between chromatin compaction and histone modification state explains Saccharomyces cerevisiae heterochromatin bistability. Proc Natl Acad Sci U S A. 2024;121:e2403316121. https://doi.org/10.1073/pnas.2403316121.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Meister P, Taddei A. Building silent compartments at the nuclear periphery: a recurrent theme. Curr Opin Genet Dev. 2013;23:96–103. https://doi.org/10.1016/j.gde.2012.12.001.

    Article  CAS  PubMed  Google Scholar 

  196. Vieux-Rochas M, Fabre PJ, Leleu M, Duboule D, Noordermeer D. Clustering of mammalian hox genes with other H3K27me3 targets within an active nuclear domain. Proc Natl Acad Sci U S A. 2015;112:4672–7. https://doi.org/10.1073/pnas.1504783112.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  197. Bantignies F, Roure V, Comet I, Leblanc B, Schuettengruber B, Bonnet J, Tixier V, Mas A, Cavalli G. Polycomb-dependent regulatory contacts between distant hox loci in Drosophila. Cell. 2011;144:214–26. https://doi.org/10.1016/j.cell.2010.12.026.

    Article  CAS  PubMed  Google Scholar 

  198. Wani AH, Boettiger AN, Schorderet P, Ergun A, Munger C, Sadreyev RI, Zhuang X, Kingston RE, Francis NJ. Chromatin topology is coupled to Polycomb group protein subnuclear organization. Nat Commun. 2016;7:10291. https://doi.org/10.1038/ncomms10291.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Grimaud C, Bantignies F, Pal-Bhadra M, Ghana P, Bhadra U, Cavalli G. RNAi components are required for Nuclear clustering of Polycomb Group Response Elements. Cell. 2006;124:957–71.

    Article  CAS  PubMed  Google Scholar 

  200. Risca VI, Denny SK, Straight AF, Greenleaf WJ. Variable chromatin structure revealed by in situ spatially correlated DNA cleavage mapping. Nature. 2017;541:237–41. https://doi.org/10.1038/nature20781.

    Article  CAS  PubMed  Google Scholar 

  201. Plys AJ, Davis CP, Kim J, Rizki G, Keenen MM, Marr SK, Kingston RE. Phase separation of polycomb-repressive complex 1 is governed by a charged disordered region of CBX2. Genes Dev. 2019;33:799–813. https://doi.org/10.1101/gad.326488.119.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Keenen MM, Brown D, Brennan LD, Renger R, Khoo H, Carlson CR, Huang B, Grill SW, Narlikar GJ, Redding S. HP1 proteins compact DNA into mechanically and positionally stable phase separated domains. Elife. 2021;10. https://doi.org/10.7554/eLife.64563.

  203. Yamada T, Fischle W, Sugiyama T, Allis CD, Grewal SI. The nucleation and maintenance of heterochromatin by a histone deacetylase in fission yeast. Mol Cell. 2005;20:173–85.

    Article  CAS  PubMed  Google Scholar 

  204. Donze D, Kamakaka RT. RNA polymerase III and RNA polymerase II promoter complexes are heterochromatin barriers in Saccharomyces cerevisiae. Embo J. 2001;20:520–31.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Raab JR, Kamakaka RT. Insulators and promoters: closer than we think. Nat Rev Genet. 2010;11:439–46. https://doi.org/10.1038/nrg2765.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  206. Fourel G, Lebrun E, Gilson E. Protosilencers as building blocks for heterochromatin. BioEssays. 2002;24:828–35.

    Article  CAS  PubMed  Google Scholar 

  207. Pillus L, Rine J. Epigenetic inheritance of transcriptional states in S. Cerevisiae. Cell. 1989;59:637–47.

    Article  CAS  PubMed  Google Scholar 

  208. Aparicio OM, Billington BL, Gottschling DE. Modifiers of position effect are shared between telomeric and silent mating-type loci in S. Cerevisiae. Cell. 1991;66:1279–87.

    Article  CAS  PubMed  Google Scholar 

  209. Sussel L, Vannier D, Shore D. Epigenetic switching of transcriptional states: cis- and trans-acting factors affecting establishment of silencing at the HMR locus in Saccharomyces cerevisiae. Mol Cell Biol. 1993;13:3919–28. https://doi.org/10.1128/mcb.13.7.3919-3928.1993.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Rossi FM, Kringstein AM, Spicher A, Guicherit OM, Blau HM. Transcriptional control: rheostat converted to on/off switch. Mol Cell. 2000;6:723–8. S1097-2765(00)00070 – 8 [pii].

    Article  CAS  PubMed  Google Scholar 

  211. Dodd IB, Micheelsen MA, Sneppen K, Thon G. Theoretical analysis of epigenetic cell memory by nucleosome modification. Cell. 2007;129:813–22. https://doi.org/10.1016/j.cell.2007.02.053.

    Article  CAS  PubMed  Google Scholar 

  212. Sneppen K, Dodd IB. Cooperative stabilization of the SIR complex provides robust epigenetic memory in a model of SIR silencing in Saccharomyces cerevisiae. Epigenetics. 2015;10:293–302. https://doi.org/10.1080/15592294.2015.1017200.

    Article  PubMed  PubMed Central  Google Scholar 

  213. Symmons O, Raj A. What’s luck got to do with it: single cells, multiple fates, and Biological Nondeterminism. Mol Cell. 2016;62:788–802. https://doi.org/10.1016/j.molcel.2016.05.023.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  214. Misteli T. The Self-Organizing genome: principles of genome. Archit Function Cell. 2020;183:28–45. https://doi.org/10.1016/j.cell.2020.09.014.

    Article  CAS  Google Scholar 

  215. Katan-Khaykovich Y, Struhl K. Heterochromatin formation involves changes in histone modifications over multiple cell generations. EMBO J. 2005;24:2138–49. https://doi.org/10.1038/sj.emboj.7600692.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Thomas P, Popovic N, Grima R. Phenotypic switching in gene regulatory networks. Proc Natl Acad Sci U S A. 2014;111:6994–9. https://doi.org/10.1073/pnas.1400049111.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Losick R, Desplan C. Stochasticity and cell fate. Science. 2008;320:65–8. https://doi.org/10.1126/science.1147888.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Ferrell JE Jr. Self-perpetuating states in signal transduction: positive feedback, double-negative feedback and bistability. Curr Opin Cell Biol. 2002;14:140–8. https://doi.org/10.1016/s0955-0674(02)00314-9.

    Article  CAS  PubMed  Google Scholar 

  219. Dayarian A, Sengupta AM. Titration and hysteresis in epigenetic chromatin silencing. Phys Biol. 2013;10. https://doi.org/10.1088/1478-3975/10/3/036005.

  220. Nagaraj VH, Mukhopadhyay S, Dayarian A, Sengupta AM. Breaking an epigenetic chromatin switch: curious features of hysteresis in Saccharomyces cerevisiae telomeric silencing. PLoS ONE. 2014;9:e113516. https://doi.org/10.1371/journal.pone.0113516.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Riising EM, Comet I, Leblanc B, Wu X, Johansen JV, Helin K. Gene silencing triggers polycomb repressive complex 2 recruitment to CpG islands genome wide. Mol Cell. 2014;55:347–60. https://doi.org/10.1016/j.molcel.2014.06.005.

    Article  CAS  PubMed  Google Scholar 

  222. Bintu L, Yong J, Antebi YE, McCue K, Kazuki Y, Uno N, Oshimura M, Elowitz MB. Dynamics of epigenetic regulation at the single-cell level. Science. 2016;351:720–4. https://doi.org/10.1126/science.aab2956.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  223. Lensch S, Herschl MH, Ludwig CH, Sinha J, Hinks MM, Mukund A, Fujimori T, Bintu L. (2022). Dynamic spreading of chromatin-mediated gene silencing and reactivation between neighboring genes in single cells. Elife 11. https://doi.org/10.7554/eLife.75115

  224. Dodson AE, Rine J. Heritable capture of heterochromatin dynamics in Saccharomyces cerevisiae. Elife. 2015;4:e05007. https://doi.org/10.7554/eLife.05007.

    Article  PubMed  PubMed Central  Google Scholar 

  225. Mano Y, Kobayashi TJ, Nakayama J, Uchida H, Oki M. Single cell visualization of yeast gene expression shows correlation of epigenetic switching between multiple heterochromatic regions through multiple generations. PLoS Biol. 2013;11:e1001601. https://doi.org/10.1371/journal.pbio.1001601.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We would like to thank J. Rine, J.T. Kadonaga, D.J. Clark, M. Gartenberg, G. Hartzog, S. Strome, M. Oki, H. Boeger, K. Wu, M. Peterson and F. Camakaka for comments and suggestions on the manuscript.

Funding

The writing of this manuscript was funded by the University of California COR CFA.

Author information

Authors and Affiliations

Authors

Contributions

ND wrote and edited the manuscript. RTK edited the manuscript and generated the figures.

Corresponding author

Correspondence to Rohinton T. Kamakaka.

Ethics declarations

Ethics approval and consent to participate

All appropriate ethical rules and regulations of the University of California were followed during the performance of the work described in this manuscript.

Consent for publication

All authors have consented to the material in this review and to submit the review for publication.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dhillon, N., Kamakaka, R.T. Transcriptional silencing in Saccharomyces cerevisiae: known unknowns. Epigenetics & Chromatin 17, 28 (2024). https://doi.org/10.1186/s13072-024-00553-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s13072-024-00553-7