Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 May 17.
Published in final edited form as: Nat Rev Mol Cell Biol. 2017 May 17;18(7):407–422. doi: 10.1038/nrm.2017.26

Mechanisms of action and regulation of ATP-dependent chromatin-remodelling complexes

Cedric R Clapier 1, Janet Iwasa 2, Bradley R Cairns 1, Craig L Peterson 3
PMCID: PMC8127953  NIHMSID: NIHMS1695149  PMID: 28512350

Abstract

Cells utilize diverse ATP-dependent nucleosome remodelling complexes to carry out histone sliding, ejection or incorporation of histone variants, suggesting that different mechanisms of action are used by the various chromatin remodelling subfamilies. However, as all chromatin remodelling subfamilies contain an ATPase–translocase ‘motor’ that translocates DNA from a common location within the nucleosome. In this Review we discuss (and illustrate with animations) an alternative, unifying mechanism of chromatin remodelling, which is based on the regulation of DNA translocation. We propose the ‘hourglass’ model of remodeller function, in which each remodeller subfamily utilizes diverse specialized proteins and protein domains to assist in nucleosome targeting or to differentially detect nucleosome epitopes. These modules converge to regulate a common DNA translocation mechanism, to inform the conserved ATPase ‘motor’ on whether and how to apply DNA translocation, which together achieve the various outcomes of chromatin remodelling: nucleosome assembly, chromatin access and nucleosome editing.

Introduction

Chromatin-structure dynamics are important for the regulation of gene expression and chromosome function. The basic components of chromatin are nucleosomes, which comprise octamers of histone proteins around which DNA is wrapped. The majority of nucleosome assembly occurs during DNA replication, directly in the wake of the replisome [G], and involves the delivery of histones to nascent DNA by histone chaperones [G]. Chromatin dynamics involve the action of specialized ATP-dependent chromatin remodelling complexes (herein termed ‘remodellers’). Remodellers include enzymes that ensure proper density and spacing of nucleosomes and can also contribute to gene repression. Another set of remodellers cooperate with site-specific transcription factors and histone modification enzymes to move or eject histones to enable binding of transcription factor to DNA. Yet another set of remodellers is involved in creating specialized chromosomal regions where canonical histones [G] are replaced by histone variants [G]. Thus, genome-wide nucleosome occupancy and composition are tailored by specialized remodellers1,2. Genetic experiments have revealed ATP-dependent chromatin remodelling enzymes as essential regulators of nearly every chromosomal process, and their deregulation leads to a variety of diseases, including cancer1,3,4.

Phylogenetic and functional analyses have classified all remodeller ATPases within the RNA/DNA helicase ‘Superfamily 2’, which can be divided into four subfamilies of chromatin remodelling enzymes: imitation switch (ISWI), chromodomain helicase DNA-binding (CHD), Switch/sucrose non-fermentable (SWI/SNF) and INO8059 (BOX 1; Supplementary information S1 (figure)). As the most extensive structure and mechanism studies have been conducted on the yeast enzymes, we focus on the yeast members of each subfamily, but we also discuss studies on the related mammalian complexes. Given their differential effect on nucleosomes, early models suggested that each subfamily used unrelated enzymatic mechanisms to achieve nucleosome organization, disorganization, ejection, or changes in nucleosome composition. This review instead proposes a unifying central mechanism for all types of remodellers: ATP-dependent DNA translocation; that is, the ATP-dependent movement of DNA by an enzyme, which, for remodellers, involves moving the DNA along the histone surface. Evidence for this mechanism begins with the observation that the catalytic subunits of all remodellers have an ATPase domain and activities consistent with DNA translocation. This ATPase domain may thus be considered a common ‘motor’ that uses DNA translocation to break histone–DNA contacts and to propel DNA along the histone surface. As different remodellers prefer different nucleosomes as substrates and generate distinct remodelling outcomes, we discuss how DNA translocation can be regulated to achieve these outcomes through remodeller-specific domains and proteins, and their interactions with histone modifications and histone variants. In this review, we synthesize the available data and propose an ‘hourglass’ model for chromatin remodeller function —targeting specificity requires diverse remodeller-specific proteins, the function of which converge on a common DNA translocation mechanism, which is regulated by diverse remodeller-specific proteins to achieve different remodelling outcomes.

BOX 1. Remodeller classification and specialization.

Chromatin remodellers can be classified into four subfamilies: imitation switch (ISWI), chromodomain helicase DNA-binding (CHD), switch/sucrose non-fermentable (SWI/SNF) and INO80, on the basis of the similarities and differences in their catalytic ATPases10 (FIG. 1b) and associated subunits (Supplementary information S1 (figure)). In addition, most higher eukaryotes have multiple remodeller subtypes within each subfamily, to provide cell type-specific or developmentally-specific remodellers1,158,159. Furthermore, orphan remodellers, which do not belong to a subfamily (for example α-thalassemia/mental retardation syndrome X-linked (ATRX)21,157,160 and Cockayne syndrome group B (CSB)161,162) also exist, but are less well-characterized mechanistically, and are therefore not discussed here. See Supplementary information S1 (figure) for the composition of remodeller complexes and Supplementary information S2 (figure) for remodeller structures.

ISWI subfamily.

The ATPase domain of the ISWI subfamily remodellers contains two RecA-like lobes, which are separated by a small insertion sequence163,164 (FIG. 1b) and a carboxy-terminal HAND-SANT-SLIDE (HSS) domain165 that binds the unmodified histone H3 tail166 and the linker DNA flanking the nucleosome167. Two domains that flank the ATPase lobes (autoinhibitory N terminal (AutoN) and negative regulator of coupling (NegC)) regulate the activity of the ATPase domian33. Budding yeast have two highly similar ISWI ATPases, Isw1 and Isw2, which assemble into three distinct complexes43,168. Functionally, most ISWI-subfamily complexes assemble and regularly space nucleosomes to limit chromatin accessibility and gene expression; however a subset, such as the nucleosome remodelling factor (NURF) complex, have accessory subunits that confer access and that promote transcription169,170.

CHD subfamily.

The ATPase domain of CHD subfamily remodellers resembles that of ISWI remodellers48, but uniquely bears in its N terminus two signature, tandemly arranged chromodomains171,172 (FIG. 1b). Analogous to ISWI, the C terminus of CHD remodellers contains a NegC domain48 followed by a DNA-binding domain (DBD) comprised of only the SANT and SLIDE domains173. In Saccharomyces cerevisiae, Chd1 functions as a monomeric remodeller171, whereas CHD remodellers have diverse complexes in metazoans172,174 and conduct all three general remodelling processes: assembly (spacing nucleosomes)16, access (exposing promoters)175 and editing (incorporating histone H3.3)176. For example, yeast Chd1 mainly conducts chromatin assembly, whereas the metazoan nucleosome remodelling deacetylase (NuRD) complex remodeller helps repressors to bind chromatin and represses genes through its associated histone deacetylases174,177.

SWI/SNF subfamily.

The ATPase domain of SWI/SNF remodellers contains two RecA-like lobes that flank a small conserved insertion, an N-terminal helicase/SANT-associated (HSA) domain that binds actin and/or actin-related proteins178180 (ARPs, see Supplementary information S8 (box)), an adjacent post-HSA domain, AT-hooks and a C-terminal bromodomain (FIG. 1b). Organisms often have two similar SWI/SNF subtypes: in yeast the Sth1 and Snf2/Swi2 ATPases nucleate the complexes remodelling the structure of chromatin (RSC) and SWI/SNF complexes, respectively. SWI/SNF-subfamily complexes contain core subunits (for example, yeast Swi3, Swp73, Snf5, and ARPs). In higher eukaryotes, combinatorial construction of SWI/SNF complexes using different core paralogues results in tissue- and developmental-specific SWI/SNF subtypes2. SWI/SNF subfamily remodellers typically facilitate chromatin access (FIG. 1a), as they slide and eject nucleosomes, and are used for either gene activation or repression.

INO80 subfamily.

Two closely related ATPases, known as Ino80 and Swr1 in yeast, define this subfamily181,182. These proteins contain a variable, large insertion between the RecA-like lobes (FIG. 1b). This insertion is shorter in yeast (~250 amino acids) than in mammals (>1,000 amino acids) and binds a single heterohexameric ring of the helicase-related (AAA+-ATPase) ruvB-like protein 1 (Rvb1) and Rvb2102,183, one ARP and one YL-1 protein family member. The N terminus of INO80 subfamily ATPases contains an HSA domain that nucleates actin and ARPs103, which has been validated by structural studies of the yeast Swr1 HSA domain–actin–Arp4 module of SWR1C184. Thus, INO80-subfamily ATPases scaffold three modules: the amino-terminal, the Rvb1 and Rvb2, and the carboxy-terminal modules57,101,102 (Supplementary information S2 (Figure)). In higher eukaryotes, certain Swr1 complex subtypes also contain histone acetyltransferase (HAT) subunits (for example, human p400, whereas their yeast counterparts functionally interact with HATs58.

INO80-subfamily remodellers uniquely perform editing functions, although INO80C also conducts chromatin access and nucleosome spacing functions. The SWR1C, p400 and Snf-2-related CBP activator protein (SRCAP) complex subtypes replace canonical H2A–H2B dimers with H2A.Z histone variant-containing H2A.Z–H2B dimers, whereas INO80C can catalyse the reciprocal reaction19,91. Of note, the vertebrate p400 subtype may also replace H3.1 with the variant H3.3 (REF. 185). Finally, yeast INO80C removes the variant H2A.X, which probably underlies its DNA repair functions186,187 as well as its chromatin access and transcription activation functions188,189.

Functional taxonomy of remodellers

Chromatin remodellers can be classified by their phylogenetic relationships10 and/or by their different functionalities1,5. We first describe their functionalities and then integrate this with phylogenetic and subfamily designations. Most remodellers are specialized to preferentially conduct one of the following three functions: nucleosome assembly and organization, chromatin access, and nucleosome editing (installing or removing histone variants) (FIG. 1a).

Figure 1. Functions and domain organization of chromatin remodellers.

Figure 1.

a. Functional classification of remodellers. The ATPase–translocase subunit of all remodellers is depicted in pink; additional subunits are depicted in green (imitation switch(ISWI) and chromodomain helicase DNA-binding(CHD)), brown (switch/sucrose non-fermentable (SWI/SNF)) and blue (INO80). Nucleosome assembly: Particular ISWI and CHD subfamily remodellers participate in the random deposition of histones, the maturation of nucleosomes and their spacing. Chromatin access: Primarily SWI/SNF subfamily remodellers alter chromatin by repositioning nucleosomes, ejecting octamers or evicting histone dimers. Nucleosome editing: Remodellers of the INO80 subfamily (INO80C or Swr1 complex (SWR1C)) change nucleosome composition by exchanging canonical and variant histones, for example installing H2A.Z variants (yellow). We note that this functional classification is a simplification, as INO80C, the ISWI remodeller nucleosome remodelling factor (NURF) and certain CHD remodellers can promote chromatin access.

b. Domain organization of remodeller subfamilies. The ATPase–translocase domain (Tr) of all the remodellers is sufficient to carry out DNA translocation. It is comprised of two RecA-like lobes (lobe 1 and lobe 2, which are separated by a short or long (such as in the INO80 subfamily) insertion (grey)). Remodellers can be classified into four subfamilies based on the length and function of the insertion and on their domain organization.

c. ‘Inchworming’ mechanism of DNA translocation. An ATP binding–hydrolysis-dependent conformational cycle of the RecA-like lobes (‘mittens’) drives DNA translocation. Mittens are depicted closed when lobes have high affinity for DNA and open when lobes have low affinity for DNA (see also Supplementary information S3 (movie)). Although the DNA can be double stranded, only the tracking strand of DNA is depicted, along which the lobes move in the 3’–5’ direction (validated by single-stranded DNA studies53). Steps are depicted as sequential, but they may be concerted, and a similar model with mitten/lobe 1 being stationary is equally supported. The yellow arrows represent remodeller movement; the green arrow represents DNA translocation. The precise step in which inorganic phosphate (Pi) is released is not known. NegC* is a region with structural similarity to the ISWI negatice regulator of coupling (NegC) domain. AutoN, autoinhibitory N-terminal; Bromo, bromodomain; CHD, chromodomain; DBD, DNA-binding domain; HSA, helicase/SANT-associated; HSS, HAND-SANT-SLIDE; SnAC, Snf2 ATP coupling. Part c is adapted from by REF. 52, Cell Press.

Nucleosome assembly and organization.

Following replication, histone chaperones deliver histone complexes (H3–H4 tetramers and H2A–H2B dimers) to nascent DNA behind the replisome11, where ‘assembly’ remodellers such as the ISWI and CHD subfamily remodellers, have two functions (FIG. 1a): they first help the initial histone–DNA complexes (pre-nucleosomes) to mature into canonical octameric nucleosomes12,13, and they next form nucleosome arrays by spacing nucleosomes at relatively fixed distance apart1417. This assembly and spacing process also takes place during transcription at locations where nucleosomes have been dynamically ejected.

Chromatin access.

Rendering the chromatin more accessible to proteins and RNA involves sliding nucleosomes along the DNA, evicting nucleosome components (such as H2A–H2B dimers) or ejecting full nucleosome (FIG. 1a). These functions are primarily (but not solely) carried out by SWI/SNF subfamily remodellers. Access remodellers can expose binding sites for transcription activators or repressors at gene promoters or enhancers18, and can increase accessibility for DNA repair and recombination factors. Whereas assembly remodellers promote gene silencing through the creation of tightly-packed nucleosome arrays, access remodellers usually promote gene expression by opening the chromatin for transcription factors.

Nucleosome editing.

Remodellers of the INO80 subfamily conduct the replication-independent removal of a particular histone within a nucleosome and its replacement with either a canonical or a variant histone (FIG. 1a). Common examples of editing include the replacement of canonical H2A or H3 with related variants, which is assisted by editing remodellers, such as yeast SWR1 complex (SWR1C) and mammalian Snf-2-related CBP activator protein (SRCAP) and p400 (REFS 1921). The inclusion of histone variants at a single nucleosome or an array of nucleosomes can affect factor recruitment, deterrence or activity.

Together, remodellers help to ensure dense nucleosome packaging (at steady state) in the vast majority of the genome, whilst also allowing factors rapid access to particular loci and acting to specialize chromosomal regions by inserting histone variants. These diverse and complex tasks involve different specialized remodeller complexes (BOX 1). Despite this diversity, we explore below the shared properties and mechanisms of action of the remodeller complexes and how they might be differentially regulated to achieve these different tasks.

A shared DNA translocation mechanism

Although they impact nucleosomes differently, all remodellers share particular properties, including: 1) a greater affinity for the nucleosome than for naked DNA; 2) a single catalytic subunit, which contains an ATPase domain that is split into two RecA-like lobes [G] (DExx, Lobe 1; HELICc, Lobe 2) (FIG. 1b and 1c); 3) domains and/or proteins that regulate the ATPase domain; and 4) domains and/or proteins for interaction with other chromatin proteins, chaperones, or site-specific transcription factors (FIG. 1b). Together, the first two shared properties listed above enable the remodellers to act on nucleosomes, whereas the latter two attributes enable selective action on particular nucleosomes at specific locations.

We discuss below the evidence that a particular enzymatic property, DNA translocation, is shared by all types of chromatin remodellers and provides the underlying force that is needed to break histone–DNA contacts, which each remodeller then tailors to achieve nucleosome repositioning and sliding, ejection or editing. This perspective has emerged from the accumulation of more than a decade of evidence from multiple laboratories. The evidence obtained from these studies can be tiered by the extent to which the assays directly test and support translocation: 1) direct assays, which provide real-time, single-molecule and/or visual observations of DNA translocation (for example optical or magnetic tweezers, and DNA curtains)2226; 2) indirect assays, which specifically test translocation (for example, triple-helix displacement, ‘tethered’ translocation, and nucleosomes with strand-specific gapped DNA)2735; and 3) indirect assays, which test the effect of remodellers on DNA or nucleosomes (for example, nucleosome sliding, facilitated access of factors to nucleosomal DNA, and changes in DNA topology) and which can be interpreted as being consistent with DNA translocation, or equally with other models, such as imposing DNA twist [G] or histone octamer conformational changes21,3646.

Perhaps the most thoroughly studied remodellers are those of the SWI/SNF-subfamily ATPases, for which all the assays listed above have been applied, resulting in strong cumulative evidence for DNA translocation. These studies include assessment of the major biophysical parameters of DNA translocation (velocity, processivity and force) through direct single molecule studies23,24. Although only a subset of the assays above have been applied to other remodellers (or their ATPase subunits), the data are similarly consistent with DNA translocation being a unifying mechanism of chromatin remodelling. Below, we explore how this unifying mechanism, when combined with regulation from associated subunits, can explain the remodelling outcomes of all the remodeller subtypes. We note that a unifying property still allows for other mechanistic contributions (for example, changes in octamer conformation) that might facilitate or inhibit the DNA translocation process or which may affect octamer stability.

The Mechanism of DNA Translocation.

Our current understanding of the mechanism of DNA translocation by remodellers is influenced by prior work on monomeric DNA (and RNA) helicases, which conduct ATP-dependent translocation by directional tracking along the phosphate backbone of one of the two DNA strands47. Helicases and remodeller ATPases are related enzymes that belong to a much larger family of ATP-dependent nucleic acid translocases — all of which have the shared property of using two RecA-like ‘lobes’ (DExx, Lobe1; HELICc, Lobe2) to conduct DNA translocation (FIG. 1b and Supplementary information S2 (figure)). Helicases insert a protein domain between the two DNA strands during translocation, causing DNA strand separation, whereas remodeller ATPases do not.

Multiple structures of monomeric DNA/RNA helicases are available47, which are similar to the three currently-available remodeller ATPase structures: Saccharomyces cerevisiae Chd148 (Supplementary information S2 (figure)), and Myceliophthora thermophila Snf249 and ISWI50. As none of these three structures contains DNA, our understanding relies more on structures of monomeric DNA helicases–translocases, which reveal the existence of a DNA-binding cleft and a site for ATP binding and hydrolysis between the RecA-like lobes, which together constitute the DNA translocation motor. Both lobe bind to the same strand of DNA, with one lobe slightly ahead of the other. This enables an ‘inchworming’ mechanism of unidirectional movement that involed the reciprocal action of the two RecA-like lobes that sequentially bind to and release DNA — like two ‘mittens’ (FIG. 1c; Supplementary information S3 (movie)) apparently moving 1–2 bp of DNA per cycle of ATP binding–hydrolysis–release24,25,47,51,52. Within the SWI/SNF-subfamily enzymes, these lobes track along one of the two strands of the DNA, moving in the 3’–5’ direction following the phosphate backbone, as gaps limited even to a single base pair can greatly impair tracking and translocation28,53. A change in perspective reveals how this property is applied to the nucleosome: when the ‘translocating’ enzyme is fixed, the DNA appears to be propelled by the enzyme, causing one helical rotation of DNA for every ~10bp that is translocated (Supplementary information S3 (movie)).

Implementation of DNA translocation on the nucleosome.

We explore below how the enzyme, if it is anchored to a fixed position on the histone octamer, can confer both translocation and twist to the DNA in order to break histone DNA contacts. We provide a biophysical perspective on nucleosomes and histone–DNA contacts in BOX 2, complemented by an animation of nucleosome structure (Supplementary information S4 (movie)). An important aspect to consider is where the translocating enzyme engages the histone octamer and nucleosomal DNA. At present, we lack a high-resolution structure of a remodeller bound to a nucleosome with the ATPase engaged. Early models of remodeller–nucleosome engagement and translocation by remodellers stipulated a processive movement of the remodeller along the DNA, involving the peeling off of DNA while traveling around the octamer, much like a RNA or DNA polymerase54, as well as the use of translocation from outside the nucleosome to push linker DNA into the nucleosome28,31. However, many studies (including hydroxyl-radical footprinting, crosslinking, DNaseI hypersensitivity, fluorescence resonance energy transfer (FRET) and comparative analysis of mononucleosomes containing or lacking strand-specific DNA gaps) now support an alternative model, in which the RecA-like lobes (the translocase domain) bind to chromatin within the nucleosome at superhelical location 2 (SHL2), which is two DNA helical turns away from the nucleosome dyad [G],27,43,44,53,55. Importantly, the position of the translocase domain remains fixed on the octamer, probably through the use of a histone-binding domain (HBD). For SWI/SNF, a clear HBD is located in the carboxy-terminus56 (FIG. 1b), known as Snf2 ATP coupling (SnAC), which helps to maintain octamer attachment during forcible DNA translocation, a property that is likely to be necessary for nucleosome ejection. For ISWI and CHD, moderate sliding activity is observed with derivatives that lack the carboxy-terminus, suggesting that an HBD resides within the translocase domain or amino-termini of these proteins33,48. Although isolated INO80-subfamily ATPases have not been tested, combined results from SWI/SNF, ISWI and CHD subfamily enzymes converge on a unifying concept: the implementation of DNA translocation from a fixed location, two helical turns away from the nucleosome dyad.

BOX 2. Nucleosome structure and thermodynamics.

Understanding chromatin-remodelling mechanisms requires an appreciation of the biophysical features and challenges that are presented by their substrate, the nucleosome, which consists of a histone octamer wrapped by 147bp of DNA. The canonical octamer is composed of two H3–H4 dimers that form the central tetramer, which is then capped on each end by an H2A–H2B dimer190. Together, they form an interlocked right-handed helical staircase upon which the DNA climbs191. Positively-charged amino acids facing outwards from the histone staircase contact the negatively-charged phosphate backbone of the DNA (Supplementary information S4 (movie)). An important nucleosomal landmark is the dyad, which is the central location where the nucleosome shows two-fold rotational symmetry; the dyad axis that defines this feature is depicted with a grey stippled bar in all the figures and animations.

Although each single histone–DNA contact is fairly weak (~1 kcal/mole, requiring ~1 pN of force to disrupt), the sum of all the 14 histone–DNA contacts that are normally present on the nucleosome confers considerable positional stability (~12–14 kcal). Together, these histone–DNA contacts define the energetic barrier that chromatin remodellers must overcome. As ATP hydrolysis provides ~7.3 kcal/mole of free energy, remodellers must either break only a few histone–DNA contacts at a time or must accumulate the energy from more than one ATP hydrolysis event to yield a repositioned, ejected or edited nucleosome. Thus, rather than breaking all histone-DNA contacts simultaneously, the need for a lower input of energy can be envisioned through the sequential disruption of subsets of histone–DNA contacts, thereby creating a DNA ‘wave’ that propagates around the octamer surface by one-dimensional diffusion. This mechanism enables histone–DNA contacts to sequentially break and re-form along the length of the nucleosome, while maintaining most of the contacts at any given time, thereby lowering the total energy that is required for repositioning. Although nucleosomes bind more strongly to certain types of DNA sequences (those favoring left-handed curvature), the energy differences between favourable and unfavourable sequences are small relative to the energy provided by ATP hydrolysis; thus, remodellers can impart sufficient force to slide nucleosomes along any DNA sequence.

Within nucleosomes, the energy cost associated with wrapping DNA far beyond its ‘persistence length’ [G] is compensated by the energy gained through histone–DNA contacts. Thus, the nucleosome can be considered a loaded spring, with energy stored in DNA bending. Remodellers that conduct nucleosome ejection implement the eventual disruption of all histone–DNA contacts, which releases the bending energy in the DNA, and may therefore take advantage of the bending energy in the disruption process. By contrast, remodellers that only conduct nucleosome sliding may limit the disruption of histone–DNA contacts to avoid the binding energy falling below the stored bending energy, which would favour and cause ejection.

Once bound to the SHL2 position on the octamer, studies suggest that the translocase domain carries out directional DNA translocation by pulling in DNA from the proximal side of the nucleosome (the DNA entry site, which is ~50 bp from the translocase) and pumping it towards the distal side (the DNA exit site, which is ~97bp from the translocase)57,58 (FIG. 2 and 3). This DNA propelling action involves the reciprocal action of the two RecA-like subdomains, moving 1–2 bp of DNA per cycle of ATP binding–hydrolysis–release24,51 (Supplementary information S5, S6 (movies)). This translocation creates DNA twist of opposite polarity on each side of the translocase domain: the proximal side is under-twisted and lacks sufficient DNA, whereas the distal side is over-twisted and contains excess DNA. On the proximal side, the translocase action breaks histone–DNA contacts, drawing DNA from the proximal linker into the nucleosome. Viewed from the location of the translocase domain, this DNA segment can propagate in a wave-like manner towards the distal exit site of the nucleosome by diffusion, with histone–DNA contacts broken at the leading edge of the wave and reforming at the lagging edge of the wave. On arrival at the distal linker, the DNA twist is resolved, and the linker is extended by 1–2 bp, resulting in the overall ‘sliding’ of the histone octamer 1–2 bp along the DNA. Through iteration, ATP hydrolysis cycles lead to additional directional displacement. This model has been termed ‘wave-ratchet-wave’ to denote the movement of DNA both towards and away from the internal translocase domain53. These ‘waves’ can be as small as 1bp24,25 (Supplementary information S5 (movie)), or can instead involve more than 1bp on the proximal side of the translocase domain51 (Supplementary information S6 (movie)), especially if DNA movement is constrained (Supplementary information S7 (movie)). Of note, the translocase domain also functions as an internal ratchet to ensure directional movement of the DNA. We explore below how remodeller subfamilies regulate DNA translocation differently to achieve their specialized outcomes.

Figure 2. Model of regulation of DNA translocation leading to precise nucleosome spacing by ISWI subfamily remodellers.

Figure 2.

a. The nucleosome is shown with the left-handed wrapping of DNA around the histone octamer (grey transparent cylinder). The DNA colour changes from light green to dark green when passing the nucleosome dyad (from the proximal side to the distal side), to distinguish the second gyre [G]. The location where a remodeller ATPase–translocase (Tr) domain will bind on the nucleosome, as well as one H4 tail, is also depicted. A red circle serves as a reference point to trace DNA translocation.

b. Schematic of the domain architecture of Drosophila melanogaster imitation switch (ISWI). Regulation of D. melanogaster ISWI: in the left-hand diagram, the autoinhibitory N-terminal (AutoN) and negative regulator of coupling (NegC) domains inhibit the ATPase activity and coupling of the translocase domain, respectively, thus inactivating DNA translocation. In the right-hand diagram, ISWI is activated by a double ‘inhibition of inhibition’: by the Arg17–Arg19 patch of the histone H4 tail and by linker DNA, which antagonize AutoN and NegC, thereby increasing ATPase and coupling, respectively, and activating DNA translocation.

c. Precise spacing of nucleosomes by ISWI subfamily remodellers. The ISWI remodeller interacts with the nucleosome two DNA helical turns away from the dyad via its ATPase–translocase domain (Tr) and is anchored to the surface of the octamer by a histone-binsing domain (HBD). The H4 tail binds to the remodeller, releasing AutoN inhibition and activating ATPase activity. In parallel, the HAND-SANT-SLIDE (HSS) domain binds to linker DNA, releasing NegC inhibition and reinstating efficient coupling (State 1). The translocase domain translocates 3bp of DNA along the surface of the nucleosome (State 1 to State 2), generating DNA tension (orange) on both the proximal and the distal sides of the translocase domain, owing to the lack and excess of DNA, respectively (State 2). On the distal side, DNA tension is resolved (restoring green DNA colour) by one-dimensional diffusion of the 3bp of excess DNA (implemented as three 1bp steps), which moves around the distal side (2nd half) of the nucleosome in the form of a small wave, and then resolves in the distal linker (State 2 to State 3). On the proximal side, DNA tension is resolved by DNA from the proximal linker entering the nucleosome, 3bp at a time, which requires the release of the HSS binding (State 3 to State 4). Consequently, the adjacent nucleosome approaches by 3bp (State 4) and the HSS re-binds linker DNA in its new position (State 5). This process is reiterated, resulting progressive approach of the adjacent nucleosome (State 5 to State 6), until the approaching octamer prevents the HSS from re-binding the linker DNA, thereby reinstating inhibition by NegC (State 6). DNA translocation then ceases, and the remodeller is released from the nucleosome (not shown), setting a precise inter-nucleosome distance. Multiple occurrences of this process result in nucleosome arrays with precise, regular spacing (Supplementary information S7 (movie)).

Figure 3. Models of nucleosome ejection by SWI/SNF subfamily remodellers.

Figure 3.

Figure 3.

a. Regulation of DNA translocation leading to nucleosome ejection by switch/sucrose non-fermentable (SWI/SNF) subfamily remodellers. The actin-related protein (ARP) module regulates both the ATPase and the coupling activities of the ATPase-translocase domain (Tr), which interacts with the nucleosome two helical turns away from the dyad (State 1), anchors to the surface of the octamer via a histone-binding domain (HBD) and translocates 1–2 bp of DNA along the surface of the nucleosome (State 1 to State 2), thereby generating DNA tension on both the proximal and the distal sides of the translocase domain. Low to moderate ATPase activity and coupling leads to weak DNA translocation and low DNA tension (orange, State2a) that are resolved by sliding (restoring green DNA colour, State 3a). Continued DNA translocation results in continued sliding, the progressive displacement of the histone octamer with respect to the DNA (State 4a). Alternatively, SWI/SNF can generate high coupling (by the helicase/SANT-associated (HSA)-Arp7-Arp9 module) and high ATPase activity (by post-HSA), leading to strong DNA translocation and high DNA tension and disruption of histone-DNA contacts (red, State 2b), which results in histone ejection (State 3b; see Supplementary information S9 (movie)).

b. Nucleosome ejection by spooling during remodelling by SWI/SNF subfamily remodellers. As shown in part a, the ATPase–translocase domain of the SWI/SNF subfamily remodellers can generate low-to-moderate DNA tension (State 1 to State 2) that is resolved by sliding. By iterations, continued DNA translocation and sliding lead to the approach of an adjacent octamer (State 3) that ultimately collides with the remodeller-bound nucleosome, resulting in the DNA being peeled of the adjacent octamer (State 4) and eventually in the ejection of the adjacent nucleosome by ‘spooling’ (State 5).

Differential regulation of translocation

In isolation, several remodeller ATPase domains are intrinsically active, although not necessarily maximally active, and it is becoming clear that the regulation of ATPase activity is carried out by domains that flank the ATPase domain and/or by associated proteins, through one of three modes: ‘gating’, ATP turnover, or ‘coupling’. A fourth mode, regulated affinity, is related to remodeller targeting and substrate selectivity and is not discussed here, except when it also affects enzymatic activity.

‘Gating’ refers to the ability of the substrate nucleosome to access the ATPase active site. A prominent example involves the yeast Chd1 remodeller, which has N-terminal chromodomains positioned to interfere with the path of DNA through the DNA-binding cleft, thereby requiring a conformational change to allow DNA binding48. An alternative form of gating regulates the Swi2/Snf2-subfamily ATPases and involves the non-productive positioning of the two RecA-like lobes without DNA, which requires a regulated conformational change to form a proper cleft for DNA interaction and ATP hydrolysis49.

The regulation of ATP turnover determines how quickly the enzyme conducts the ATP-binding–hydrolysis driven conformation cycle — the successive conformations that an enzyme adopts during the ATP hydrolysis cycle — and thus how fast the RecA-like ‘mittens’ move along DNA (Supplementary information S3, S5, S6 (movies)).

‘Coupling’ refers to the efficiency of this process — the probability that ATP hydrolysis will result in DNA translocation (or the amount of DNA translocated per ATP hydrolysis). Thus, coupling can represent the ability of the mittens to properly grip and release during the conformation cycle, and may help to implement force. For each remodeller subtype, we describe below the domains and proteins that are known to regulate gating, ATPase activity or coupling, and how they collaborate to determine the remodelling outcome.

Mechanisms and regulation of chromatin assembly.

Nucleosome assembly and spacing is primarily conducted by ISWI and CHD subfamily remodellers, although sliding and spacing activity has also been reported for an INO80 subfamily member59. Replication presents a challenge to chromatin organization as nucleosomes are initially deposited randomly in the wake of DNA replication. First, ‘pre-nucleosomes’ form, which are histone octamers upon which the DNA is not fully or properly wrapped (~80bp rather than 147bp)12,13. The remodeller may assist in the formation of fully mature and wrapped nucleosomes, and may subsequently use ATP-dependent nucleosome sliding activities to achieve ordered spacing12,13 (FIG. 1a).

The key to spacing involves the separate regulation of the ATPase and the ‘coupling’ components of DNA translocation. ISWI and CHD remodellers each use a DNA-binding domain (DBD), HAND-SANT-SLIDE (HSS) and DBD, respectively (FIG. 1b) as a ‘molecular ruler’ that binds linker (extranucleosomal) DNA and that measures the distance between nucleosomes60. HSS binding to linker DNA leads to: 1) the binding of the ATPase domain on the proximal side, two helical turns from the nucleosome dyad (FIG. 2a) and 2) the promotion of ‘coupling’, by antagonizing the negative regulator of coupling (NegC) domain, which physically bridges the two RecA-like lobes, and negatively regulates coupling. Before nucleosome binding, NegC prevents productive DNA translocation by uncoupling ATP hydrolysis from translocation, imparting a form of autoinhibition33 (FIG. 2b). Binding of the HSS to linker DNA relieves NegC autoinhibition, thereby restoring coupling and enabling DNA translocation (FIG. 2c, see state 1 and state 5). This interplay and inhibition of the inhibition of catalytic activity together provide an alternative regulatory mechanism to the previously suggested ‘power stroke’ models in which the HSS pushes DNA into the nucleosome33,61. The active translocase domain pulls in DNA, causing tension on both sides of the domain (FIG. 2c, state 2). This tension is resolved in two temporal phases: on the distal side, tension is resolved by ‘wave propagation’ as described above (FIG 2c, state 3). On the proximal side, the tension constrained between the translocase and HSS domain is retained until the HSS releases the linker DNA (FIG. 2c, state 4). In fact, single molecule FRET experiments have suggested that the HSS remains bound during the translocation of 3bp of DNA. This results in 3bp of DNA exiting first from the distal side of the nucleosome via wave propagation, 1bp at a time. Next, the 3bp of undertwisted DNA is relaxed and drawn into the nucleosome51, which we interpret as resulting from the HSS domain releasing this undertwisted DNA (FIG. 2c, state 4; Supplementary information S7 (movie)). Thus ~1bp of sliding occurs per ATP hydrolysis, but is executed in 3bp successive steps. Iterations of this cycle will draw the adjacent nucleosome ever nearer (FIG 2c, state 5 to 6), progressively shortening the linker DNA, until the adjacent nucleosome interferes with binding by the HSS through steric hindrance (FIG. 2c, state 6). When the HSS can no longer re-bind linker DNA, it fails to antagonize NegC and the translocation ceases (FIG. 2c, state 6), leaving the adjacent nucleosome at a fixed distance from the substrate nucleosome. Sequential application of the spacing process to all nucleosomes on the template will produce an array in which all nucleosomes are separated by the same distance. This entire process is depicted in Supplementary information S7 (movie).

Of note, the ISWI complex ATP-utilising chromatin remodelling and assembly factor (ACF) contains a protein (Acf1) that extends the length of the DNA bound by the HSS and the DBD domains and thus yields a nucleosome array with a longer average linker16, in keeping with the model described above. Moreover, Acf1 contributes to the mechanism of sensing the length of linker DNA through an interesting interplay with the histone H4 tail and the autoinhibitory N-terminal (AutoN) domain (see below)62. We note that a DNA-bound transcription factor (or a strongly positioned nucleosome) can function as a boundary element, against which unidirectional sliding and spacing events occur, thereby facilitating the formation of an array of regularly spaced nucleosomes46,6366.

In addition to regulated coupling, regulated ATPase activity is also a feature of ISWI. A small basic patch (Arg residues 17–19) of the histone H4 tail activates chromatin assembly remodellers such as ISWI6770 by helping to orient the ATPase domain on the nucleosome to two helical turns from the dyad71 and inducing an allosteric change in the ATP-binding pocket72. The N-terminus of ISWI contains a ‘mimic’ of the H4 tail basic patch known as AutoN (FIG. 1b), which inhibits ATPase activity33 and provides a form of autoinhibition (analogous to that of NegC) that is antagonized by the authentic H4 tail basic patch. Of note, the human ACF1 protein also contributes allosterically to this regulation when not binding linker DNA by sequestering the histone H4 tail62. Thus, ISWI activity is regulated at two levels, ATPase and coupling, with both involve the relief of intrinsic autoinhibition by nucleosomal epitopes33. The recently published structure of M. thermophila ISWI50 both strongly supports and extends prior biochemical and genetic research. This work confirms that the AutoN domain and the H4 tail basic patch exhibit mutually exclusive binding to the surface of lobe 2, and the AutoN domain has been shown to extend past its original length to include an additional inhibitory element. In parallel, NegC was confirmed to act allosterically to respond to extranucleosomal DNA length. Thus, ISWI is a tightly and dynamically regulated motor; it is held in check by autoinhibition, but, once relieved of its intrinsic brakes, it functions efficiently as an autonomous nucleosome remodeller33,73.

Studies suggest a similar mechanism of nucleosome spacing by the yeast CHD subfamily, based in part on their NegC and similar HSS DNA-binding domains16. However, CHD remodellers bear one or more chromodomains in their N-terminus rather than a clear AutoN domain, suggesting slightly different modes of regulation. Of note, the CHD subfamily has greatly expanded in mammals, but has had more limited mechanistic characterization.

Interestingly, certain ISWI complexes (for example, human SNF2H (also known as SMARCA5)) can operate via a 2:1 rather than a 1:1 remodeller:nucleosome stoichiometry31,74. The 2:1 ratio involves a second SNF2H complex binding in a symmetrical position on the opposite side of the nucleosome, without a steric clash with the first complex. As the DNA translocation mechanism is directional (propelling towards the dyad), two SNF2H complexes on opposite sides may alternate in activity, enabling octamer movement in alternative directions74. Furthermore, SNF2H dimerization enhances remodelling, and a conformational change in SNF2H that occurs upon ATP analogue binding may affect the sensing of linker DNA length and translocation75. This work is consistent with the regulatory mechanisms described above, as ISWI conformational changes during an ATP hydrolysis cycle may affect HSS–linker interactions.

Mechanisms and regulation of chromatin access.

Chromatin access can be accomplished by specialized remodellers in all four subfamilies, but is most strongly associated with SWI/SNF remodellers. Biochemical and structural studies with SWI/SNF-subfamily remodellers strongly support a 1:1 remodeller:nucleosome stoichiometry, and reveal a striking binding pocket with almost perfect mononucleosome dimensions7682 (Supplementary information S2 (figure)). Access of the nucleosome to this pocket may involve conformational changes and can be regulated by histone tail modifications76,79,80,82, which might be recognized by one of many histone-interacting domains in SWI/SNF complexes. We consider this to be consistent with the idea of histone modifications ‘gating’ access to the pocket.

A key question about SWI/SNF remodellers is how sliding versus ejection is regulated. Recent work has identified several DNA translocation regulators: three domains that reside on the catalytic subunit (helicase/SANT-associated (HSA), post-HSA and Protrusion 1, which is located within the insertion between the two RecA-like lobes), and two actin-related proteins (ARPs) that directly bind to the HSA domain35 (FIG. 3a). ARPs are important functional members of two of the four remodeller subfamilies, SWI/SNF and INO80 (Supplementary information S8 (box)). In the yeast SWI/SNF subfamily, the associated ARPs facilitate sliding and enable ejection35. Mechanistically, these ARPs greatly improve ‘coupling’ by folding back and interacting functionally with Protrusion 1 within the translocase domain (FIG. 3a). By contrast, the post-HSA domain has no effect on coupling, but instead strongly inhibits ATP hydrolysis. These ATPase and coupling activities are integrated by the translocase domain to determine the velocity and efficiency of DNA translocation35. With low to moderate DNA translocation efficiency, proportional nucleosome sliding occurs, whereas with higher DNA translocation efficiency, histone ejection occurs35 (FIG. 3a; Supplementary information S9 (movie)). Of note, direct interaction of the post-HSA region with the Protrusion 1 domain was demonstrated in a recent structural study49, reinforcing the notion that interaction of the ARP/HSA/post-HSA module with Protrusion 1 is a conserved feature that is used for ATPase and/or coupling regulation.

We propose two not mutually exclusive models to explain how DNA translocation could lead to nucleosome ejection. In the first model, the disruption of multiple histone–DNA contacts by efficient and forcible DNA translocation might directly facilitate histone loss (FIG. 3a), and might additionally enable histone chaperones and/or specialized proteins on particular remodellers to gain access to assist in the removal of the underlying histones. Considering the position of the translocase domain, the H2A–H2B dimer should be the most susceptible to ejection, and indeed, SWI/SNF remodellers may destabilize and remove H2A–H2B dimers as the first step in ejection8386. Of note, the use of DNA translocation by SWI/SNF for the removal of H2A–H2B dimers may be a feature that is shared with editing remodellers, although only editing remodellers replace canonical H2A—H2B with histone variant dimers (see below).

In the second model, the nucleosome adjacent to the one that is undergoing remodelling is the nucleosome that will be ejected, rather than the nucleosome that is bound by the remodeller. In this case, the act of processive DNA translocation on the bound nucleosome initially draws the available linker DNA into the bound nucleosome and, when linker DNA availability is exhausted, the remodeller ‘spools’ the DNA off of the adjacent nucleosome, which leads to octamer ejection87,88 (FIG. 3b). This model is supported by studies of yeast SWI/SNF using templates with two nucleosomes89,90. Of note, SWI/SNF remodellers display higher levels of DNA translocation efficiency [G] than ISWI remodellers, which may explain why only SWI/SNF remodellers can eject nucleosomes35. Finally, a key unanswered question is how SWI/SNF remodellers are instructed to implement high levels of DNA translocation, which enables nucleosome ejection.

Regulation of nucleosome editing.

Nucleosome editing, which involves the incorporation or removal of histone variants, is mostly carried out by members of the INO80-subfamily; nucleosome editing allows construction of specialized chromatin regions in a replication-independent manner. Among the key variants to be incorporated is the H2A variant, H2A.Z. Elegant work has shown that the yeast INO80C subtype SWR1C complex removes canonical H2A–H2B dimers and replaces them with H2A.Z–H2B dimers19. Of note, SWR1C does not slide nucleosomes34, whereas INO80C can slide nucleosomes, as well as catalyze the eviction and replacement of H2A.Z–H2B dimers9193. Thus, the INO80C subtype has both ‘access’ and ‘editing’ activities. Of note, in yeast, an H2A.Z-contaning nucleosome is less stable than a canonical nucleosome94. This property may be used for regulating genes and heterochromatin properties9597, but is not discussed further here.

A remodeller specialized for ‘editing’ must be able to stabilize the hexasome [G] during remodelling and incorporate a specific variant dimer to generate a variant-containing nucleosome. Thus, ‘editing’ remodellers must discriminate between substrate and product nucleosomes. For example, yeast SWR1C contains proteins specialized in H2A.Z–H2B recognition98,99, and conducts dimer replacement in a stepwise and unidirectional manner, one dimer at a time, first generating heterotypic nucleosomes (with one canonical H2A–H2B dimer and one variant H2A.Z–H2B dimer), and then generating homotypic H2A.Z nucleosomes100. However, as SWR1C binds with comparable affinity to both H2A- and H2A.Z-containing nucleosomes, substrate discrimination is complex34. We hypothesize that the translocase domain of SWR1C only interacts productively with an H2A-containing nucleosome (or heterotypic nucleosome), and not with a homotypic H2A.Z-containing nucleosome, thereby specifying H2A–H2B dimer replacement (FIG. 4). Interestingly, as ATPase activity is stimulated by H2A nucleosomes, and is activated further by the addition of free H2A.Z–H2B dimers, free H2A.Z–H2B acts as both a regulator and a substrate molecule100. With regard to ‘gating’, H2A.Z–H2B, or a SWR1C subunit that recognizes H2A.Z, may ‘gate’ the interaction of the translocase domain with nucleosomal DNA near the dyad. Domain swapping experiments have identified H2A.Z residues that block dimer exchange by SWR1C, providing candidates for a gating domain34.

Figure 4. Model of histone exchange by the remodeller SWR1C.

Figure 4.

As in the case of imitation switch (ISWI) and switch/sucrose non-fermentable (SWI/SNF) subfamilies, the ATPase–translocase domain (Tr) of the yeast INO80 subfamily remodeller Swr1 complex (SWR1C) interacts with a nucleosome two helical turns away from the dyad (State 1), is anchored to the surface of the octamer via a histone-binding domain (HBD) and strongly translocates 1–2 bp of DNA at the surface of the nucleosome (State 1 to State 2), thereby generating high DNA tension (red DNA, State 2). On the distal side, resolution of DNA tension by DNA propagation may be prevented by the binding of the HBD (State 3). On the proximal side, DNA tension is resolved by the rupture of several upstream histone–DNA contacts (restoring green DNA colour, State 3), allowing the release of one canonical H2A–H2B dimer and the loading of a variant H2AZ–H2B dimer assisted by the Swc2 and/or Swr1 subunits (State 3 to State 4). Histone-DNA contacts are restored following the histone exchange (State 5; see Supplementary information S10 (movie)).

Structural studies of yeast INO80C and SWR1C have revealed their similar architecture, which is characterized by a compact ‘head’ domain containing ruvB-like proteins 1 and 2 (Rvb1–Rvb2) in a heterohexameric ring, and a flexible ‘tail’ domain57,101,102 (Supplementary information S2 (figure)). Both enzymes show remarkable structural heterogeneity, with a continuum of extended and compact forms. The INO80C complex may adopt an extended form that sandwiches a nucleosome between the ‘tail’ and the ‘head’101 domains (Supplementary information S2 (figure)), whereas the compact form blocks nucleosome binding, potentially providing an additional ‘gating’ mode of regulation.

Similarly to SWI/SNF subfamily remodellers, INO80 subtypes use ARPs for nucleosome editing103 (Supplementary information S8 (box)). In the case of yeast INO80C, one Arp8–Arp4–actin module binds to the HSA domain and promotes high affinity nucleosome binding102 (Supplementary information S2 (figure)). A second ARP-containing module, Ies6–Arp5, interacts with the ATPase insertion domain. The Ies2 subunit facilitates Ies6–Arp5 interaction with and activation of the Ino80 ATPase104,105. This module also promotes nucleosome binding105, with Arp5 coupling ATPase activity with nucleosome mobilization104106. Research on the reconstituted human INO80C core complex confirms the roles described above for the human homologues of Ies2 (INO80B) and Arp5–Ies6 (INO80C-ARP5)107. Finally, human Ino80 has a C-terminal domain that has the hallmarks of NegC, as its removal markedly increases the ATPase activity105.

Yeast SWR1C also seems to use DNA translocation to facilitate H2A–H2B removal and subsequent H2A.Z–H2B deposition; SWR1C interacts with DNA approximately two helical turns from the nucleosome dyad, and histone exchange is blocked by an adjacent 2bp gap, which is consistent with an essential role for DNA translocation in nucleosome editing34. Of note, the amount of translocation involved may not be extensive, as a DNA gap located 3–4 helical turns from the nucleosome dyad does not inhibit SWR1C34. As SWR1C does not reposition the H2A.Z-containing nucleosome product, how does SWR1C impose DNA translocation without sliding the nucleosome? We suggest that SWR1C functions like other remodellers: it engages nucleosomal DNA near the dyad (FIG. 4, state 1), but translocates only a few base pairs, thereby creating a region of localized DNA translocation and twist between the translocase and the proximal DNA entry–exit site (FIG. 4, state 2). However, unlike enzymes that catalyze sliding, SWR1C does not allow propagation of the twist and translocation ‘wave’ to the distal side of the nucleosome, but instead ‘holds’ this torsion in order to selectively destabilize DNA that is bound to the adjacent H2A–H2B dimer (FIG. 4, state 3). Following H2A exchange with H2A.Z, the DNA would then re-wrap, resulting in a nucleosomal product with an unchanged translational position (FIG. 4, state 5; Supplementary information S10 (movie)).

Regulation by targeting epitopes

We discuss below how nucleosomal epitopes, transcription factors and chromatin factors interact with remodeller domains to help to target and also to regulate chromatin remodellers.

Remodeller targeting and regulation by histone modifications and variants.

Multiple studies have linked histone modifications to the targeting and regulation of chromatin remodellers108. Remodellers of all four subtypes contain proteins and/or domains that bind to histone modifications (for example, bromodomains, bromo-adjacent homology (BAH) domains, chromodomains, plant homeodomain (PHD) domains, Pro-Trp-Trp-Pro (PWWP) domains and tryptophan-aspartic acid (WD40) domains). Furthermore, chromatin-modifying enzymes and complexes cooperate with remodellers, especially at environmentally-regulated genes109. For example, the yeast Spt-Ada-Gcn5-acetyltransferase (SAGA) complex commonly functions with SWI/SNF to promote gene activity, and can assist SWI/SNF recruitment to gene promoters110.

In principle, a histone modification (or variant) could influence (either positively or negatively) remodeller targeting or activity. Indeed, the remodeller domains and motifs listed above selectively engage nucleosomes that have particular modifications, and can enhance their affinity (targeting) or activity. To affect activity, the histone modification or variant should subsequently regulate the functions of the ATPase domain or the histone chaperones. To affect targeting, histone modifications are predicted to function as combinations of modifications, or cooperate with gene-specific activators and repressors111, as isolated remodeller motif–modification interactions are typically of low-to-moderate affinity (mid-to-high μM range)112119. Another aspect of selectivity is how remodellers avoid binding and/or acting on the nucleosomes that they should normally avoid, which may involve modifications that prevent remodeller binding (or, alternatively, ‘shielding’ the nucleosome by recruiting another factor120), or inhibiting remodeller activity through an allosteric mechanism.

Interactions exist between particular histone modifications and all four remodellers subfamilies; below, we provide a selective list. In the ISWI subfamily, a PHD or a PWWP domain is used for targeting methylated histones: the Drosophila melanogaster nucleosome remodelling factor (NURF) remodelling complex binds to trimethylated histone H3 Lys4 (H3K4me3) (REFS 113, 121, 122), and the yeast ISW1b complex binds to H3K36me3 (REF. 123). In the CHD subfamily, the chromodomain or chromodomains can similarly bind to methylated histones124,125. In the SWI/SNF subfamily, the bromodomain located at the C-terminus of the yeast Snf2/Swi2 ATPase promotes the targeting of SWI/SNF activity to nucleosomes that are acetylated on histone H3 (REFS 111,112,126128). The yeast Snf2/Swi2 bromodomain can also regulate the binding to and remodelling of acetylated nucleosomes by SWI/SNF through an intramolecular interaction with an acetylated residue within the Swi2/Snf2 subunit129,130. Similarly, the tandem bromodomain located on the yeast remodel the structure of chromatin (RSC) complex subunit 4 (Rsc4) shows specificity for binding to acetylated histone H3 Lys14 (H3K14ac), and this single acetylated residue is sufficient to enhance RSC binding to nucleosomes in vitro131. In the yeast INO80 subfamily, the double bromodomain-containing subunit bromodomain-containing factor 1 (Bdf1) promotes H2A.Z deposition by SWR1C on nucleosomes acetylated on either histone H4 or H2A132.

In addition to roles in targeting, histone modifications and variants also regulate remodelling activities. For example, nucleosomes containing the H2A.Z variant stimulate the ATPase and remodelling activities of ISWI, but not the activities of SWI/SNF subfamily members133. For the SWI/SNF subfamily members, site-specific H3 acetylation was shown to enhance the remodelling activities of yeast RSC and SWI/SNF, without increasing their nucleosome binding affinities111,126. Intriguingly, acetylated H3 peptides open the nucleosome binding cavity within RSC80. The enhancement of remodelling activity required the Snf2/Swi2 bromodomain111, implying a regulatory role beyond targeting for bromodomain–histone tail interactions.

The H4 tail stimulates the D. melanogaster ISWI ATPase and remodelling activity6769, and most studies have suggested that acetylation of the H4 tail attenuates this stimulation33,50,126,134,135. Acetylation of H4K12 or H4K16 may weaken the ability of the H4 tail to compete with the AutoN domain. Of note, H4K16ac does not inactivate ISWI, but instead the ATP-dependent mobilization of nucleosomes is slowed by ~1.5-fold to 4-fold50,126,134,135. Furthermore, a recent study has suggested that the ability of D. melanogaster ISWI to properly space nucleosome arrays is not sensitive to H4K16ac136.

In addition to those in histone tails, many post-translational modifications occur on the globular surface of the nucleosome. Of particular interest are two lysine residues (sometimes targeted for acetylation), H3K56 and H3K64, which are located on the lateral surface where DNA enters the nucleosome. In yeast, H3K56ac enhances the ability of SWI/SNF and RSC to mobilize nucleosomes in cis137, whereas H3K64ac enhances the sliding activity of Chd1 but not RSC138. How these two acetylation marks can have distinct effects on different remodelling enzymes is not yet clear. H3K56ac also has a striking effect on the nucleosome editing activities of yeast INO80 subfamily members — it enhances the editing activity of INO80C but disrupts nucleosome discrimination by SWR1C, thereby allowing both H2A- and H2A.Z-containing nucleosomes to stimulate its ATPase activity94. How H3K56ac disrupts nucleosome discrimination is not known, although it has been suggested to function together with the H2A.Z-binding subunit, Swc2 (REF.94).

Remodeller targeting and regulation by DNA-binding proteins.

Alongside their important roles in recruiting remodellers and mediating enhancer-promoter looping139, transcription activators can directly interact with remodellers and can affect remodelling activities; for example, interactions between yeast SWI/SNF and a DNA-bound activator promotes nucleosome eviction in vitro140142. We speculate that – beyond recruitment – DNA site-specific transcription activators and repressors (in combination with histone modifications and variants) interact in a specific manner with remodeller domains and proteins, regulate DNA translocation parameters, and influence both remodelling efficiency and outcomes143. Therefore, the cell type-specific compositional diversity of remodellers may exist to interact with the cell type-specific activator and repressor repertoire to regulate remodeller targeting and remodelling activities.

The Hourglass Model

Considering the current data, we suggest that ATP-dependent chromatin remodellers align with the following conceptual framework: 1) compositional diversity, to specify both targeting and regulation; 2) a unifying mechanism, that involves a histone-anchored ATPase that conducts directional DNA translocation two helical turns from the dyad; and 3) functional specialization through proteins and domains that regulate how DNA translocation generates the alternative outcomes of chromatin assembly, access and editing. This framework resembles an ‘hourglass’, where the narrow neck represents the aspect of least variance among remodellers — the unifying implementation of histone-anchored DNA translocation within the nucleosome (FIG. 5).

Figure 5.

Figure 5.

The hourglass model of chromatin remodelling. Remodeller diversity (top) serves a variety of nucleosomal processes (bottom), but all might funnel through a unifying mechanism: an ATPase (‘motor’) subunit that is anchored to the histone octamer two helical turns away from the dyad that carries out DNA translocation (centre). At the top of the hourglass, the compositional diversity of remodellers enables specific interactions with particular transcription factors and/or histone modifications to specify targeting. The funnel depicts the implementation of ATP-dependent DNA translocation by the translocase (Tr) domain from a fixed location at the nucleosome, anchored by a histone-binding domain (HBD). At the bottom of the hourglass, the various nucleosome remodelling outcomes are depicted (assembly, access or editing), which are achieved through separate processes. Each process involves particular regulatory domains and/or proteins (blue boxes) on each remodeller, and their interactions with specific transcription factors and chromatin features, such as histone modifications, variants and linker DNA (green boxes). ARPs, actin-related proteins; AutoN, autoinhibitory N-terminal; CHD, chromodomain helicase DNA-binding; HSS, HAND-SANT-SLIDE; NegC, negative regulator of coupling.

This model also provides an intuitive path for the evolution of chromatin remodellers. An ancient protein that moved along DNA could have evolved into a DNA translocase within a chromatin remodeller by acquiring histone anchoring capacity, which enables the propelling of DNA around nucleosomes — placing it at the narrow part of the hourglass. From there, evolution could have then shaped how the ATPase is regulated, with autoinhibition representing a particularly common mode33,144,145. In this case, domains that flank the ATPase domain inhibit ATPase activity or efficiency, which is reversed by particular histone epitopes or associated proteins, thereby enabling DNA translocation. Intuitively, these ATPase regulatory domains could have then co-evolved with histone modifications and variants to regulate and to specialize the enzyme — and to have co-evolved further with DNA-binding transcription activators and repressors to regulate targeting — resulting in the expansive remodeller families and remodeller functions that we have now.

Perspectives and future directions

As discussed above, many remodellers are built in a modular mode, with cell-type specific subunits that tailor interactions with and regulation by cell-type specific activators and repressors. Going forwards, we envision much attention being paid to how these specific interactions provide a varied targeting repertoire, and also how they enable particular remodeller outcomes at specific locations. An additional question is how certain factors promote novel remodelling activities; for example, how SWI/SNF evicts the heterochromatin protein Sir3 from nucleosomes146. Metazoan SWI/SNF remodellers can evict Polycomb repressive complex 1 (PRC1) and PRC2147, but can also function synergistically with them in vivo to achieve cellular states such as pluripotency148. Remodeller activities may be further influenced by remodeller subunits that recognize histone modifications and variants. For SWI/SNF and INO80 subfamily remodellers, the roles of ARPs will be of particular interest based on their demonstrated ability to control ATPase activity, coupling, or both35,107, as well as their known interactions with histones. For the INO80 subfamily in particular, the precise mechanistic roles of the enigmatic Rvb AAA+ ATPases and their relationship to the remodeller ATPase cycle is of great interest as they provide an additional ATP-dependent activity. Although the composition of yeast ISWI and CHD subfamily remodellers is simple (Supplementary information S1 (figure)), it is much more complex in mammals, with a wide variety of associated proteins and candidate regulatory domains flanking the ATPase domain, making them ripe for future studies.

The field will greatly benefit from high-resolution structural studies of the ATPase–translocase domain interacting with regulatory domains or proteins and the nucleosome, along with mutant versions that confer conformational changes that inform about regulatory mechanisms. The characterization of the disease-causing mutations of remodeller subunits – in particular, those mutations that cause cancer and developmental disorders – will continue to draw much effort. As a first step, it will be important to characterize remodeller assembly in the cell of origin and determine whether a mutation solely affects remodeller targeting, or remodeller activity, or both, as well as its effect on transcription and genome stability.

Additional subjects for future study include: how histone chaperones cooperate with remodellers to install or evict specific histones11,149,150, how remodellers from different subfamilies may synergize or may antagonize each other to dynamically remodel chromatin states123,151,152, and how non-coding RNAs might influence remodeller targeting and regulation153156. Furthermore, we currently understand only a portion of the mechanistic details of orphan remodellers, such as the autoinhibitory mechanism that regulates the ATPase activity of the human transcription-coupled nucleotide excision repair ERCC6 (REFS 144, 145), and the use of histone chaperones to specify interaction of the human α-thalassemia/mental retardation syndrome X-linked (ATRX) remodeller with H3.3 (REFS 21, 157). It will be of great interest to determine whether these orphan enzymes share similar mechanistic and regulatory principles with the four main remodeller subfamilies.

We conclude this review by emphasizing that the genomic chromatin landscape is sculpted by remodellers that do not simply conduct a standard task once targeted, but instead function as ‘smart’ machines that are informed by their nucleosome substrate and resident proteins about whether and how to remodel, through the regulation of their common ATPase–DNA translocase activity.

Supplementary Material

Supp text
FigS1
FigS2
Animation01
Download video file (28.9MB, mov)
Animation02
Download video file (14.8MB, mov)
Animation03a
Download video file (22.4MB, mov)
Animation03b
Download video file (42.1MB, mov)
Animation04
Download video file (33.8MB, mov)
Animation05
Download video file (46.1MB, mov)
Animation06
Download video file (9.6MB, mov)

Online Summary:

  • Dynamic regulation of chromatin involves four subfamilies of ATP-dependent nucleosome remodelling complexes: ISWI, CHD (chromodomain), SWI/SNF and INO80. Each subfamily is specialized to preferentially achieve particular chromatin outcomes: assembly, access or editing.

  • Diversity in the protein composition of remodellers enables their specific interaction with particular transcription activators, repressors, and histone modifications, which together specify targeting.

  • Although diverse in protein composition, all remodellers have a similar ATPase ‘motor’ that translocates DNA from a common location within the nucleosome, which breaks histone-DNA contacts.

  • The diverse specialized proteins and domains in each remodeller subfamily are involved also in detecting nucleosome epitopes, which differentially regulate the conserved ATPase-translocase motor to achieve the various chromatin remodelling outcomes.

  • We propose an ‘hourglass’ model of chromatin remodelling that involves convergence on a DNA translocation mechanism, which is preceded and followed by remodeller diversity, for differential remodeller targeting and remodelling outcomes, respectively.

  • Remodellers are emerging as ‘smart’ machines that are informed whether and how to utilize DNA translocation to conduct chromatin remodelling.

GLOSSARY:

Replisome

A large protein complex that carries out the DNA replication process, from the unwinding of double-stranded DNA to strand duplication by DNA synthesis.

Histone chaperones

Proteins that bind to free histones, prevent histone aggregation and that can promote either nucleosome assembly or disassembly.

Canonical histones

The four core histones (H2A, H2B, H3 and H4) that are most commonly assembled into nucleosomes during replication and that constitute almost all of the nucleosomes across the genome.

Histone variants

Differ by few amino-acids from canonical histones and are expressed at low-to-moderate levels and typically inserted into nucleosomes independently of replication; they create specific chromatin regions and functions.

RecA-like lobes

Protein domains of helicases and remodellers, similar in structure and sequence to the ATPase domain of the Escherichia coli DNA-binding protein RecA.

DNA twist

A measure of the extent of helical winding of the DNA strands around each other, along their common axis. Often expressed as the number of base pairs of DNA per helical turn in B-form DNA.

Persistence length

A mechanical property of polymer stiffness, which for DNA is approximately 100bp.

Nucleosome dyad

A pseudo-two-fold symmetry element of the nucleosome core particle.

Gyre

In the context of a nucleosome, refers to one DNA wrap around the surface of the octamer.

DNA translocation efficiency

Quantified by measuring coupling, it describes the amount of DNA that is translocated per ATP hydrolysis and/or the probability that the enzyme conducts a DNA translocation step per ATP hydrolysis cycle.

Hexasome

Nucleosome that lacks one histone H2A–H2B dimer.

BIOGRAPHIES:

Cedric R. Clapier is a Research Assistant Professor in the Department of Oncological Sciences at the University of Utah School of Medicine. He was a graduate student at EMBL with Peter Becker where he discovered the first direct regulation of a chromatin remodeller (ISWI) by a nucleosome feature, an EMBO postdoctoral fellow with Christoph Müller (for ISWI structural studies), and a faculty associate with Brad Cairns on the regulation of chromatin remodelling. His research continues to focus on the mechanistic characterization of the regulatory architecture and logic of chromatin remodellers and their alterations in cancers.

Janet Iwasa is a Research Assistant Professor in the Biochemistry Department at the University of Utah. Her broad goal is to create accurate and compelling molecular visualizations that will support research, learning and scientific communication.

Bradley R. Cairns is Professor and Chair of the Department of Oncological Sciences, an Investigator with HHMI, and the Sr. Director of Basic Science at Huntsman Cancer Institute, University of Utah. His graduate work with Roger Kornberg (Stanford, Nobel Laureate) included yeast SWI/SNF and RSC complex purification, and his postdoctoral work with Fred Winston (Harvard Medical School) included SWI/SNF and RSC genetics and function. His lab now works broadly on chromatin-transcription relationships: basic, developmental and cancer-related.

Craig L. Peterson was a postdoctoral fellow at UCSF with Ira Herskowitz (1988–1992) where he identified genes encoding SWI/SNF subunits. He is currently a Professor of Molecular Medicine at the University of Massachusetts Medical School. His research group continues to focus on the understanding how chromosome structure influences gene transcription, DNA replication and repair, with special emphasis on identifying and characterizing the cellular machines that control chromosome dynamics.

References

  • 1.Clapier CR & Cairns BR in Fundamentals of Chromatin (eds Workman JL & Abmayr SM) 69–146 (Springer, 2014). [Google Scholar]
  • 2.Ho L & Crabtree GR Chromatin remodelling during development. Nature 463, 474–484 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Wilson BG & Roberts CW SWI/SNF nucleosome remodellers and cancer. Nature reviews. Cancer 11, 481–492 (2011). [DOI] [PubMed] [Google Scholar]
  • 4.Lai AY & Wade PA Cancer biology and NuRD: a multifaceted chromatin remodelling complex. Nature reviews. Cancer 11, 588–596 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Clapier CR & Cairns BR The biology of chromatin remodeling complexes. Annu Rev Biochem 78, 273–304 (2009). [DOI] [PubMed] [Google Scholar]
  • 6.Bartholomew B Regulating the chromatin landscape: structural and mechanistic perspectives. Annu Rev Biochem 83, 671–696 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Narlikar GJ, Sundaramoorthy R & Owen-Hughes T Mechanisms and functions of ATP-dependent chromatin-remodeling enzymes. Cell 154, 490–503 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Hargreaves DC & Crabtree GR ATP-dependent chromatin remodeling: genetics, genomics and mechanisms. Cell Res 21, 396–420 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Becker PB & Workman JL Nucleosome remodeling and epigenetics. Cold Spring Harb Perspect Biol 5 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Flaus A, Martin DM, Barton GJ & Owen-Hughes T Identification of multiple distinct Snf2 subfamilies with conserved structural motifs. Nucleic Acids Res 34, 2887–2905 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Gurard-Levin ZA, Quivy JP & Almouzni G Histone chaperones: assisting histone traffic and nucleosome dynamics. Annu Rev Biochem 83, 487–517 (2014). [DOI] [PubMed] [Google Scholar]
  • 12.Torigoe SE, Urwin DL, Ishii H, Smith DE & Kadonaga JT Identification of a rapidly formed nonnucleosomal histone-DNA intermediate that is converted into chromatin by ACF. Mol Cell 43, 638–648 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Fei J et al. The prenucleosome, a stable conformational isomer of the nucleosome. Genes Dev 29, 2563–2575 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Corona DF et al. ISWI is an ATP-dependent nucleosome remodeling factor. Mol Cell 3, 239–245 (1999). [DOI] [PubMed] [Google Scholar]
  • 15.Varga-Weisz PD et al. Chromatin-remodelling factor CHRAC contains the ATPases ISWI and topoisomerase II [published erratum appears in Nature 1997 Oct 30;389(6654):1003]. Nature 388, 598–602 (1997). [DOI] [PubMed] [Google Scholar]
  • 16.Lusser A, Urwin DL & Kadonaga JT Distinct activities of CHD1 and ACF in ATP-dependent chromatin assembly. Nat Struct Mol Biol 12, 160–166 (2005). [DOI] [PubMed] [Google Scholar]
  • 17.Ito T, Bulger M, Pazin MJ, Kobayashi R & Kadonaga JT ACF, an ISWI-containing and ATP-utilizing chromatin assembly and remodeling factor. Cell 90, 145–155 (1997). [DOI] [PubMed] [Google Scholar]
  • 18.Boeger H, Griesenbeck J, Strattan JS & Kornberg RD Removal of promoter nucleosomes by disassembly rather than sliding in vivo. Mol Cell 14, 667–673 (2004). [DOI] [PubMed] [Google Scholar]
  • 19.Mizuguchi G et al. ATP-driven exchange of histone H2AZ variant catalyzed by SWR1 chromatin remodeling complex. Science 303, 343–348 (2004). This study demonstrates for the first time nucleosome editing by SWR1C involving H2A.Z.
  • 20.Ruhl DD et al. Purification of a human SRCAP complex that remodels chromatin by incorporating the histone variant H2A.Z into nucleosomes. Biochemistry 45, 5671–5677 (2006). [DOI] [PubMed] [Google Scholar]
  • 21.Goldberg AD et al. Distinct factors control histone variant H3.3 localization at specific genomic regions. Cell 140, 678–691 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Lia G et al. Direct observation of DNA distortion by the RSC complex. Mol Cell 21, 417–425 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Zhang Y et al. DNA translocation and loop formation mechanism of chromatin remodeling by SWI/SNF and RSC. Mol Cell 24, 559–568 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Sirinakis G et al. The RSC chromatin remodelling ATPase translocates DNA with high force and small step size. Embo J 30, 2364–2372 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Harada BT et al. Stepwise nucleosome translocation by RSC remodeling complexes. Elife 5, e10051 (2016). References 23–25 investigate the DNA step size occuring during DNA translocation by SWI/SNF remodellers.
  • 26.Blosser TR, Yang JG, Stone MD, Narlikar GJ & Zhuang X Dynamics of nucleosome remodelling by individual ACF complexes. Nature 462, 1022–1027 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Langst G & Becker PB ISWI induces nucleosome sliding on nicked DNA. Mol Cell 8, 1085–1092 (2001). [DOI] [PubMed] [Google Scholar]
  • 28.Saha A, Wittmeyer J & Cairns BR Chromatin remodeling by RSC involves ATP-dependent DNA translocation. Genes Dev 16, 2120–2134 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Whitehouse I, Stockdale C, Flaus A, Szczelkun MD & Owen-Hughes T Evidence for DNA Translocation by the ISWI Chromatin-Remodeling Enzyme. Mol Cell Biol 23, 1935–1945 (2003). References 28 and 29 demonstrated for the first time that chromatin remodellers act by DNA translocation.
  • 30.Zofall M, Persinger J & Bartholomew B Functional role of extranucleosomal DNA and the entry site of the nucleosome in chromatin remodeling by ISW2. Mol Cell Biol 24, 10047–10057 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Strohner R et al. A ‘loop recapture’ mechanism for ACF-dependent nucleosome remodeling. Nat Struct Mol Biol 12, 683–690 (2005). [DOI] [PubMed] [Google Scholar]
  • 32.Stockdale C, Flaus A, Ferreira H & Owen-Hughes T Analysis of nucleosome repositioning by yeast ISWI and Chd1 chromatin remodeling complexes. J Biol Chem 281, 16279–16288 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Clapier CR & Cairns BR Regulation of ISWI involves inhibitory modules antagonized by nucleosomal epitopes. Nature 492, 280–284 (2012). This work demonstrates that ISWI ATPase is an intrinsically active DNA translocase that is regulated by ‘inhibition of inhibition’ of both ATPase activity and coupling.
  • 34.Ranjan A et al. H2A histone-fold and DNA elements in nucleosome activate SWR1-mediated H2A.Z replacement in budding yeast. Elife 4, e06845 (2015). This work demonstrates that SWR1C interacts with nucleosomes at position SHL2 and that histone exchange requires DNA translocation.
  • 35.Clapier CR et al. Regulation of DNA Translocation Efficiency within the Chromatin Remodeler RSC/Sth1 Potentiates Nucleosome Sliding and Ejection. Mol Cell 62, 453–461 (2016). This study shows that nucleosome ejection by the Sth1 ATPase is achieved through upregulation of DNA translocation efficiency, and that actin-related proteins are required by the remodeller RSC for nucleosome ejection.
  • 36.Hamiche A, Sandaltzopoulos R, Gdula DA & Wu C ATP-dependent histone octamer sliding mediated by the chromatin remodeling complex NURF. Cell 97, 833–842 (1999). [DOI] [PubMed] [Google Scholar]
  • 37.Langst G, Bonte EJ, Corona DF & Becker PB Nucleosome movement by CHRAC and ISWI without disruption or trans-displacement of the histone octamer. Cell 97, 843–852 (1999). References 36 and 37 reveal the capacity of ISWI-subfamily remodellers to perform nucleosome sliding.
  • 38.Whitehouse I et al. Nucleosome mobilization catalysed by the yeast SWI/SNF complex. Nature 400, 784–787 (1999). This work showed for the first time nucleosome sliding by SWI/ SNF subfamily remodellers.
  • 39.Gavin I, Horn PJ & Peterson CL SWI/SNF chromatin remodeling requires changes in DNA topology. Mol Cell 7, 97–104 (2001). [DOI] [PubMed] [Google Scholar]
  • 40.Fyodorov DV & Kadonaga JT Dynamics of ATP-dependent chromatin assembly by ACF. Nature 418, 897–900 (2002). [DOI] [PubMed] [Google Scholar]
  • 41.Kassabov SR, Zhang B, Persinger J & Bartholomew B SWI/SNF unwraps, slides, and rewraps the nucleosome. Mol Cell 11, 391–403 (2003). [DOI] [PubMed] [Google Scholar]
  • 42.Fitzgerald DJ et al. Reaction cycle of the yeast Isw2 chromatin remodeling complex. Embo J 23, 3836–3843 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Kagalwala MN, Glaus BJ, Dang W, Zofall M & Bartholomew B Topography of the ISW2-nucleosome complex: insights into nucleosome spacing and chromatin remodeling. Embo J 23, 2092–2104 (2004). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Schwanbeck R, Xiao H & Wu C Spatial contacts and nucleosome step movements induced by the NURF chromatin remodeling complex. J Biol Chem 279, 39933–39941 (2004). References 43 and 44 define the binding of the ISWI subfamily remodellers to extranucleosomal DNA, and within the nucleosome two DNA helical turns from the dyad.
  • 45.Saha A, Wittmeyer J & Cairns BR Chromatin remodelling: the industrial revolution of DNA around histones. Nat Rev Mol Cell Biol 7, 437–447 (2006). [DOI] [PubMed] [Google Scholar]
  • 46.Yang JG, Madrid TS, Sevastopoulos E & Narlikar GJ The chromatin-remodeling enzyme ACF is an ATP-dependent DNA length sensor that regulates nucleosome spacing. Nat Struct Mol Biol 13, 1078–1083 (2006). [DOI] [PubMed] [Google Scholar]
  • 47.Singleton MR, Dillingham MS & Wigley DB Structure and mechanism of helicases and nucleic acid translocases. Annu Rev Biochem 76, 23–50 (2007). [DOI] [PubMed] [Google Scholar]
  • 48.Hauk G, McKnight JN, Nodelman IM & Bowman GD The chromodomains of the Chd1 chromatin remodeler regulate DNA access to the ATPase motor. Mol Cell 39, 711–723 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Xia X, Liu X, Li T, Fang X & Chen Z Structure of chromatin remodeler Swi2/Snf2 in the resting state. Nat Struct Mol Biol 23, 722–729 (2016). [DOI] [PubMed] [Google Scholar]
  • 50.Yan L, Wang L, Tian Y, Xia X & Chen Z Structure and regulation of the chromatin remodeller ISWI. Nature 540, 466–469 (2016). References 48–50 present the crystal structures of the Chd1, Snf2 and ISWI chromatin remodellers.
  • 51.Deindl S et al. ISWI Remodelers Slide Nucleosomes with Coordinated Multi-Base-Pair Entry Steps and Single-Base-Pair Exit Steps. Cell 152, 442–452 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Velankar SS, Soultanas P, Dillingham MS, Subramanya HS & Wigley DB Crystal structures of complexes of PcrA DNA helicase with a DNA substrate indicate an inchworm mechanism. Cell 97, 75–84 (1999). [DOI] [PubMed] [Google Scholar]
  • 53.Saha A, Wittmeyer J & Cairns BR Chromatin remodeling through directional DNA translocation from an internal nucleosomal site. Nat Struct Mol Biol 12, 747–755 (2005). [DOI] [PubMed] [Google Scholar]
  • 54.Lorch Y, Zhang M & Kornberg RD Histone octamer transfer by a chromatin-remodeling complex. Cell 96, 389–392 (1999). This work shows for the first time nucleosome ejection by SWI/ SNF remodellers.
  • 55.Zofall M, Persinger J, Kassabov SR & Bartholomew B Chromatin remodeling by ISW2 and SWI/SNF requires DNA translocation inside the nucleosome. Nat Struct Mol Biol 13, 339–346 (2006). References 53 and 55 demonstrate that DNA translocation occurs within the nucleosome.
  • 56.Sen P et al. The SnAC domain of SWI/SNF is a histone anchor required for remodeling. Mol Cell Biol 33, 360–370 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Nguyen VQ et al. Molecular architecture of the ATP-dependent chromatin-remodeling complex SWR1. Cell 154, 1220–1231 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Auger A et al. Eaf1 is the platform for NuA4 molecular assembly that evolutionarily links chromatin acetylation to ATP-dependent exchange of histone H2A variants. Mol Cell Biol 28, 2257–2270 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Udugama M, Sabri A & Bartholomew B The INO80 ATP-dependent chromatin remodeling complex is a nucleosome spacing factor. Mol Cell Biol 31, 662–673 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.McKnight JN, Jenkins KR, Nodelman IM, Escobar T & Bowman GD Extranucleosomal DNA binding directs nucleosome sliding by Chd1. Mol Cell Biol 31, 4746–4759 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Ludwigsen J, Klinker H & Mueller-Planitz F No need for a power stroke in ISWI-mediated nucleosome sliding. EMBO Rep 14, 1092–1097 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Hwang WL, Deindl S, Harada BT & Zhuang X Histone H4 tail mediates allosteric regulation of nucleosome remodelling by linker DNA. Nature 512, 213–217 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Tsukiyama T, Becker PB & Wu C ATP-dependent nucleosome disruption at a heat-shock promoter mediated by binding of GAGA transcription factor [see comments]. Nature 367, 525–532 (1994). [DOI] [PubMed] [Google Scholar]
  • 64.Kang JG, Hamiche A & Wu C GAL4 directs nucleosome sliding induced by NURF. Embo J 21, 1406–1413 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Nodelman IM et al. The Chd1 chromatin remodeler can sense both entry and exit sides of the nucleosome. Nucleic Acids Res 44, 7580–7591(2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Wiechens N et al. The Chromatin Remodelling Enzymes SNF2H and SNF2L Position Nucleosomes adjacent to CTCF and Other Transcription Factors. PLoS Genet 12, e1005940 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Clapier CR, Langst G, Corona DF, Becker PB & Nightingale KP Critical role for the histone H4 N terminus in nucleosome remodeling by ISWI. Mol Cell Biol 21, 875–883 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Hamiche A, Kang JG, Dennis C, Xiao H & Wu C Histone tails modulate nucleosome mobility and regulate ATP-dependent nucleosome sliding by NURF. Proc Natl Acad Sci U S A 98, 14316–14321 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Clapier CR, Nightingale KP & Becker PB A critical epitope for substrate recognition by the nucleosome remodeling ATPase ISWI. Nucleic Acids Res 30, 649–655 (2002). References 67–69 discover and characterize the activation of the remodeller ISWI by the histone H4 tail basic-patch.
  • 70.Fazzio TG, Gelbart ME & Tsukiyama T Two distinct mechanisms of chromatin interaction by the Isw2 chromatin remodeling complex in vivo. Mol Cell Biol 25, 9165–9174 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Dang W, Kagalwala MN & Bartholomew B Regulation of ISW2 by concerted action of histone H4 tail and extranucleosomal DNA. Mol Cell Biol 26, 7388–7396 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Racki LR et al. The histone H4 tail regulates the conformation of the ATP-binding pocket in the SNF2h chromatin remodeling enzyme. J Mol Biol 426, 2034–2044 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Mueller-Planitz F, Klinker H, Ludwigsen J & Becker PB The ATPase domain of ISWI is an autonomous nucleosome remodeling machine. Nat Struct Mol Biol 20, 82–89 (2013). [DOI] [PubMed] [Google Scholar]
  • 74.Racki LR et al. The chromatin remodeller ACF acts as a dimeric motor to space nucleosomes. Nature 462, 1016–1021 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Leonard JD & Narlikar GJ A nucleotide-driven switch regulates flanking DNA length sensing by a dimeric chromatin remodeler. Mol Cell 57, 850–859 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Asturias FJ, Chung WH, Kornberg RD & Lorch Y Structural analysis of the RSC chromatin-remodeling complex. Proc Natl Acad Sci U S A 99, 13477–13480 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Smith CL, Horowitz-Scherer R, Flanagan JF, Woodcock CL & Peterson CL Structural analysis of the yeast SWI/SNF chromatin remodeling complex. Nat Struct Biol 10, 141–145 (2003). [DOI] [PubMed] [Google Scholar]
  • 78.Leschziner AE, Lemon B, Tjian R & Nogales E Structural Studies of the Human PBAF Chromatin-Remodeling Complex. Structure (Camb) 13, 267–275 (2005). [DOI] [PubMed] [Google Scholar]
  • 79.Leschziner AE et al. Conformational flexibility in the chromatin remodeler RSC observed by electron microscopy and the orthogonal tilt reconstruction method. Proc Natl Acad Sci U S A 104, 4913–4918 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Skiniotis G, Moazed D & Walz T Acetylated histone tail peptides induce structural rearrangements in the RSC chromatin remodeling complex. J Biol Chem 282, 20804–20808 (2007). [DOI] [PubMed] [Google Scholar]
  • 81.Chaban Y et al. Structure of a RSC-nucleosome complex and insights into chromatin remodeling. Nat Struct Mol Biol 15, 1272–1277 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Dechassa ML et al. Architecture of the SWI/SNF-Nucleosome Complex. Mol Cell Biol 28, 6010–6021 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Bruno M, Flaus A & Owen-Hughes T Site-specific attachment of reporter compounds to recombinant histones. Methods Enzymol 375, 211–228 (2004). [DOI] [PubMed] [Google Scholar]
  • 84.Lorch Y, Maier-Davis B & Kornberg RD Chromatin remodeling by nucleosome disassembly in vitro. Proc Natl Acad Sci U S A 103, 3090–3093 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Yang X, Zaurin R, Beato M & Peterson CL Swi3p controls SWI/SNF assembly and ATP-dependent H2A-H2B displacement. Nat Struct Mol Biol 14, 540–547 (2007). [DOI] [PubMed] [Google Scholar]
  • 86.Rowe CE & Narlikar GJ The ATP-dependent remodeler RSC transfers histone dimers and octamers through the rapid formation of an unstable encounter intermediate. Biochemistry 49, 9882–9890 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Cairns BR Chromatin remodeling: insights and intrigue from single-molecule studies. Nat Struct Mol Biol 14, 989–996 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Boeger H, Griesenbeck J & Kornberg RD Nucleosome retention and the stochastic nature of promoter chromatin remodeling for transcription. Cell 133, 716–726 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Engeholm M et al. Nucleosomes can invade DNA territories occupied by their neighbors. Nat Struct Mol Biol 16, 151–158 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Dechassa ML et al. SWI/SNF has intrinsic nucleosome disassembly activity that is dependent on adjacent nucleosomes. Mol Cell 38, 590–602 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Papamichos-Chronakis M, Watanabe S, Rando OJ & Peterson CL Global regulation of H2A.Z localization by the INO80 chromatin-remodeling enzyme is essential for genome integrity. Cell 144, 200–213 (2011). This study demonstrates nucleosome editing by the remodeller INO80C and that INO80C prevents the mislocalization of H2A.Z outside gene promoters.
  • 92.Shen X, Ranallo R, Choi E & Wu C Involvement of actin-related proteins in ATP-dependent chromatin remodeling. Mol Cell 12, 147–155 (2003). [DOI] [PubMed] [Google Scholar]
  • 93.Jin C et al. H3.3/H2A.Z double variant-containing nucleosomes mark ‘nucleosome-free regions’ of active promoters and other regulatory regions. Nat Genet 41, 941–945 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Watanabe S, Radman-Livaja M, Rando OJ & Peterson CL A histone acetylation switch regulates H2A.Z deposition by the SWR-C remodeling enzyme. Science 340, 195–199 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Rege M et al. Chromatin Dynamics and the RNA Exosome Function in Concert to Regulate Transcriptional Homeostasis. Cell Rep 13, 1610–1622 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Zhang H, Roberts DN & Cairns BR Genome-Wide Dynamics of Htz1, a Histone H2A Variant that Poises Repressed/Basal Promoters for Activation through Histone Loss. Cell 123, 219–231 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Raisner RM et al. Histone Variant H2A.Z Marks the 5’ Ends of Both Active and Inactive Genes in Euchromatin. Cell 123, 233–248 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Wu WH et al. Swc2 is a widely conserved H2AZ-binding module essential for ATP-dependent histone exchange. Nat Struct Mol Biol 12, 1064–1071 (2005). [DOI] [PubMed] [Google Scholar]
  • 99.Hong J et al. The catalytic subunit of the SWR1 remodeler is a histone chaperone for the H2A.Z-H2B dimer. Mol Cell 53, 498–505 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Luk E et al. Stepwise histone replacement by SWR1 requires dual activation with histone H2A.Z and canonical nucleosome. Cell 143, 725–736 (2010). This study describes the stepwise replacement of H2A with H2A.Z in nucleosome by the remodeller SWR1C.
  • 101.Tosi A et al. Structure and subunit topology of the INO80 chromatin remodeler and its nucleosome complex. Cell 154, 1207–1219 (2013). [DOI] [PubMed] [Google Scholar]
  • 102.Watanabe S et al. Structural analyses of the chromatin remodelling enzymes INO80-C and SWR-C. Nat Commun 6, 7108 (2015). References 57, 101 and 102 present the latest structures of the remodellers INO80C and SWR1C.
  • 103.Szerlong H et al. The HSA domain binds nuclear actin-related proteins to regulate chromatin-remodeling ATPases. Nat Struct Mol Biol 15, 469–476 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Chen L et al. Subunit organization of the human INO80 chromatin remodeling complex: an evolutionarily conserved core complex catalyzes ATP-dependent nucleosome remodeling. J Biol Chem 286, 11283–11289 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Chen L, Conaway RC & Conaway JW Multiple modes of regulation of the human Ino80 SNF2 ATPase by subunits of the INO80 chromatin-remodeling complex. Proc Natl Acad Sci U S A 110, 20497–20502 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Yao W et al. Assembly of the Arp5 (Actin-related Protein) Subunit Involved in Distinct INO80 Chromatin Remodeling Activities. J Biol Chem 290, 25700–25709 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Willhoft O, Bythell-Douglas R, McCormack EA & Wigley DB Synergy and antagonism in regulation of recombinant human INO80 chromatin remodeling complex. Nucleic Acids Res 44, 8179–8188 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Suganuma T & Workman JL Signals and combinatorial functions of histone modifications. Annu Rev Biochem 80, 473–499 (2011). [DOI] [PubMed] [Google Scholar]
  • 109.Pollard KJ & Peterson CL Chromatin remodeling: a marriage between two families? Bioessays 20, 771–780 (1998). [DOI] [PubMed] [Google Scholar]
  • 110.Mitra D, Parnell EJ, Landon JW, Yu Y & Stillman DJ SWI/SNF binding to the HO promoter requires histone acetylation and stimulates TATA-binding protein recruitment. Mol Cell Biol 26, 4095–4110 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Chatterjee N et al. Histone H3 tail acetylation modulates ATP-dependent remodeling through multiple mechanisms. Nucleic Acids Res 39, 8378–8391 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Hassan AH et al. Function and selectivity of bromodomains in anchoring chromatin-modifying complexes to promoter nucleosomes. Cell 111, 369–379 (2002). [DOI] [PubMed] [Google Scholar]
  • 113.Li H et al. Molecular basis for site-specific read-out of histone H3K4me3 by the BPTF PHD finger of NURF. Nature 442, 91–95 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Song JJ, Garlick JD & Kingston RE Structural basis of histone H4 recognition by p55. Genes Dev 22, 1313–1318 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Flanagan JF et al. Double chromodomains cooperate to recognize the methylated histone H3 tail. Nature 438, 1181–1185 (2005). [DOI] [PubMed] [Google Scholar]
  • 116.Mansfield RE et al. Plant homeodomain (PHD) fingers of CHD4 are histone H3-binding modules with preference for unmodified H3K4 and methylated H3K9. J Biol Chem 286, 11779–11791 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Ruthenburg AJ et al. Recognition of a mononucleosomal histone modification pattern by BPTF via multivalent interactions. Cell 145, 692–706 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Tallant C et al. Molecular basis of histone tail recognition by human TIP5 PHD finger and bromodomain of the chromatin remodeling complex NoRC. Structure 23, 80–92 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Filippakopoulos P & Knapp S The bromodomain interaction module. FEBS Lett 586, 2692–2704 (2012). [DOI] [PubMed] [Google Scholar]
  • 120.Peterson CL Chromatin remodeling enzymes: taming the machines: Third in review series on chromatin dynamics. EMBO Rep 3, 319–322 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Santos-Rosa H et al. Methylation of histone H3 K4 mediates association of the Isw1p ATPase with chromatin. Mol Cell 12, 1325–1332 (2003). [DOI] [PubMed] [Google Scholar]
  • 122.Wysocka J et al. A PHD finger of NURF couples histone H3 lysine 4 trimethylation with chromatin remodelling. Nature 442, 86–90 (2006). [DOI] [PubMed] [Google Scholar]
  • 123.Smolle M et al. Chromatin remodelers Isw1 and Chd1 maintain chromatin structure during transcription by preventing histone exchange. Nat Struct Mol Biol 19, 884–892 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Sims RJ 3rd et al. Human but not yeast CHD1 binds directly and selectively to histone H3 methylated at lysine 4 via its tandem chromodomains. J Biol Chem 280, 41789–41792 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Watson AA et al. The PHD and chromo domains regulate the ATPase activity of the human chromatin remodeler CHD4. Journal of molecular biology 422, 3–17 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Ferreira H, Flaus A & Owen-Hughes T Histone modifications influence the action of Snf2 family remodelling enzymes by different mechanisms. J Mol Biol 374, 563–579 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Lee HS, Park JH, Kim SJ, Kwon SJ & Kwon J A cooperative activation loop among SWI/SNF, gamma-H2AX and H3 acetylation for DNA double-strand break repair. Embo J 29, 1434–1445 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Bennett G & Peterson CL SWI/SNF recruitment to a DNA double-strand break by the NuA4 and Gcn5 histone acetyltransferases. DNA Repair (Amst) 30, 38–45 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Kim JH, Saraf A, Florens L, Washburn M & Workman JL Gcn5 regulates the dissociation of SWI/SNF from chromatin by acetylation of Swi2/Snf2. Genes Dev 24, 2766–2771 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Dutta A et al. Swi/Snf dynamics on stress-responsive genes is governed by competitive bromodomain interactions. Genes Dev 28, 2314–2330 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.VanDemark AP et al. Autoregulation of the rsc4 tandem bromodomain by gcn5 acetylation. Mol Cell 27, 817–828 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Altaf M et al. NuA4-dependent acetylation of nucleosomal histones H4 and H2A directly stimulates incorporation of H2A.Z by the SWR1 complex. J Biol Chem 285, 15966–15977 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Goldman JA, Garlick JD & Kingston RE Chromatin remodeling by imitation switch (ISWI) class ATP-dependent remodelers is stimulated by histone variant H2A.Z. J Biol Chem 285, 4645–4651 (2010). This work was first to show a regulatory role for H2A.Z in remodelling by ISWI.
  • 134.Corona DF, Clapier CR, Becker PB & Tamkun JW Modulation of ISWI function by site-specific histone acetylation. EMBO Rep 3, 242–247 (2002). This work was first to demonstrate that chromatin remodellers can be regulated by a histone modification.
  • 135.Shogren-Knaak M et al. Histone H4-K16 acetylation controls chromatin structure and protein interactions. Science 311, 844–847 (2006). [DOI] [PubMed] [Google Scholar]
  • 136.Klinker H et al. ISWI remodelling of physiological chromatin fibres acetylated at lysine 16 of histone H4. PLoS One 9, e88411 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Neumann H et al. A method for genetically installing site-specific acetylation in recombinant histones defines the effects of H3 K56 acetylation. Mol Cell 36, 153–163 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Di Cerbo V et al. Acetylation of histone H3 at lysine 64 regulates nucleosome dynamics and facilitates transcription. Elife 3, e01632 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Yadon AN, Singh BN, Hampsey M & Tsukiyama T DNA looping facilitates targeting of a chromatin remodeling enzyme. Mol Cell 50, 93–103 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Owen-Hughes T & Workman JL Remodeling the chromatin structure of a nucleosome array by transcription factor-targeted trans-displacement of histones. Embo J 15, 4702–4712 (1996). [PMC free article] [PubMed] [Google Scholar]
  • 141.Gutierrez JL, Chandy M, Carrozza MJ & Workman JL Activation domains drive nucleosome eviction by SWI/SNF. Embo J 26, 730–740 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Yudkovsky N, Logie C, Hahn S & Peterson CL Recruitment of the SWI/SNF chromatin remodeling complex by transcriptional activators. Genes Dev 13, 2369–2374 (1999). References 140–142 present initial evidence that transcription activators can regulate chromatin remodelling.
  • 143.Li M et al. Dynamic regulation of transcription factors by nucleosome remodeling. Elife 4, e06249 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Lake RJ, Geyko A, Hemashettar G, Zhao Y & Fan HY UV-induced association of the CSB remodeling protein with chromatin requires ATP-dependent relief of N-terminal autorepression. Mol Cell 37, 235–246 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Wang L et al. Regulation of the Rhp26ERCC6/CSB chromatin remodeler by a novel conserved leucine latch motif. Proc Natl Acad Sci U S A 111, 18566–18571 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.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 111, 17827–17832 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Kia SK, Gorski MM, Giannakopoulos S & Verrijzer CP SWI/SNF mediates polycomb eviction and epigenetic reprogramming of the INK4b-ARF-INK4a locus. Mol Cell Biol 28, 3457–3464 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Ho L et al. esBAF facilitates pluripotency by conditioning the genome for LIF/STAT3 signalling and by regulating polycomb function. Nat Cell Biol 13, 903–913 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Venkatesh S & Workman JL Histone exchange, chromatin structure and the regulation of transcription. Nat Rev Mol Cell Biol 16, 178–189 (2015). [DOI] [PubMed] [Google Scholar]
  • 150.Li X & Tyler JK Nucleosome disassembly during human non-homologous end joining followed by concerted HIRA- and CAF-1-dependent reassembly. Elife 5, e15129 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 151.Parnell TJ, Schlichter A, Wilson BG & Cairns BR The chromatin remodelers RSC and ISW1 display functional and chromatin-based promoter antagonism. Elife 4, e06073 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Ocampo J, Chereji RV, Eriksson PR & Clark DJ The ISW1 and CHD1 ATP-dependent chromatin remodelers compete to set nucleosome spacing in vivo. Nucleic Acids Res 44, 4625–4635 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Mayer C, Neubert M & Grummt I The structure of NoRC-associated RNA is crucial for targeting the chromatin remodelling complex NoRC to the nucleolus. EMBO Rep 9, 774–780 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Onorati MC et al. The ISWI chromatin remodeler organizes the hsromega ncRNA-containing omega speckle nuclear compartments. PLoS Genet 7, e1002096 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Prensner JR et al. The long noncoding RNA SChLAP1 promotes aggressive prostate cancer and antagonizes the SWI/SNF complex. Nat Genet 45, 1392–1398 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Han P et al. A long noncoding RNA protects the heart from pathological hypertrophy. Nature 514, 102–106 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Lewis PW, Elsaesser SJ, Noh KM, Stadler SC & Allis CD Daxx is an H3.3-specific histone chaperone and cooperates with ATRX in replication-independent chromatin assembly at telomeres. Proc Natl Acad Sci U S A 107, 14075–14080 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Bao Y & Shen X SnapShot: Chromatin Remodeling Complexes. Cell 129, 632 (2007). [DOI] [PubMed] [Google Scholar]
  • 159.Lessard JA & Crabtree GR Chromatin Regulatory Mechanisms in Pluripotency. Annual Review of Cell and Developmental Biology 26, 503–532 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Law MJ et al. ATR-X syndrome protein targets tandem repeats and influences allele-specific expression in a size-dependent manner. Cell 143, 367–378 (2010). [DOI] [PubMed] [Google Scholar]
  • 161.Woudstra EC et al. A Rad26-Def1 complex coordinates repair and RNA pol II proteolysis in response to DNA damage. Nature 415, 929–933 (2002). [DOI] [PubMed] [Google Scholar]
  • 162.Citterio E et al. ATP-dependent chromatin remodeling by the Cockayne syndrome B DNA repair-transcription-coupling factor. Mol Cell Biol 20, 7643–7653 (2000). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Corona DF & Tamkun JW Multiple roles for ISWI in transcription, chromosome organization and DNA replication. Biochim Biophys Acta 1677, 113–119 (2004). [DOI] [PubMed] [Google Scholar]
  • 164.Yadon AN & Tsukiyama T SnapShot: Chromatin remodeling: ISWI. Cell 144, 453–453 e451 (2011). [DOI] [PubMed] [Google Scholar]
  • 165.Grune T et al. Crystal structure and functional analysis of a nucleosome recognition module of the remodeling factor ISWI. Mol Cell 12, 449–460 (2003). [DOI] [PubMed] [Google Scholar]
  • 166.Boyer LA, Latek RR & Peterson CL The SANT domain: a unique histone-tail-binding module? Nat Rev Mol Cell Biol 5, 158–163 (2004). [DOI] [PubMed] [Google Scholar]
  • 167.Dang W & Bartholomew B Domain architecture of the catalytic subunit in the ISW2-nucleosome complex. Mol Cell Biol 27, 8306–8317 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Vary JC Jr. et al. Yeast Isw1p forms two separable complexes in vivo. Mol Cell Biol 23, 80–91 (2003). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 169.Xiao H et al. Dual functions of largest NURF subunit NURF301 in nucleosome sliding and transcription factor interactions. Mol Cell 8, 531–543 (2001). [DOI] [PubMed] [Google Scholar]
  • 170.Hochheimer A, Zhou S, Zheng S, Holmes MC & Tjian R TRF2 associates with DREF and directs promoter-selective gene expression in Drosophila. Nature 420, 439–445 (2002). [DOI] [PubMed] [Google Scholar]
  • 171.Tran HG, Steger DJ, Iyer VR & Johnson AD The chromo domain protein chd1p from budding yeast is an ATP-dependent chromatin-modifying factor. Embo J 19, 2323–2331 (2000). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Kunert N & Brehm A Novel Mi-2 related ATP-dependent chromatin remodelers. Epigenetics 4, 209–211 (2009). [DOI] [PubMed] [Google Scholar]
  • 173.Ryan DP, Sundaramoorthy R, Martin D, Singh V & Owen-Hughes T The DNA-binding domain of the Chd1 chromatin-remodelling enzyme contains SANT and SLIDE domains. Embo J 30, 2596–2609 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Denslow SA & Wade PA The human Mi-2/NuRD complex and gene regulation. Oncogene 26, 5433–5438 (2007). [DOI] [PubMed] [Google Scholar]
  • 175.Murawska M & Brehm A CHD chromatin remodelers and the transcription cycle. Transcription 2, 244–253 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Konev AY et al. CHD1 motor protein is required for deposition of histone variant H3.3 into chromatin in vivo. Science 317, 1087–1090 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Allen HF, Wade PA & Kutateladze TG The NuRD architecture. Cell Mol Life Sci 70, 3513–3524 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Mohrmann L & Verrijzer CP Composition and functional specificity of SWI2/SNF2 class chromatin remodeling complexes. Biochim Biophys Acta 1681, 59–73 (2005). [DOI] [PubMed] [Google Scholar]
  • 179.Kasten MM, Clapier CR & Cairns BR SnapShot: Chromatin remodeling: SWI/SNF. Cell 144, 310 e311 (2011). [DOI] [PubMed] [Google Scholar]
  • 180.Schubert HL et al. Structure of an actin-related subcomplex of the SWI/SNF chromatin remodeler. Proc Natl Acad Sci U S A 110, 3345–3350 (2013). This work presented the first structure of an ARP module upon a remodeller HSA domain.
  • 181.Morrison AJ & Shen X Chromatin remodelling beyond transcription: the INO80 and SWR1 complexes. Nat Rev Mol Cell Biol 10, 373–384 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Bao Y & Shen X SnapShot: Chromatin remodeling: INO80 and SWR1. Cell 144, 158–158 e152 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Jha S & Dutta A RVB1/RVB2: running rings around molecular biology. Mol Cell 34, 521–533 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Cao T et al. Crystal structure of a nuclear actin ternary complex. Proc Natl Acad Sci U S A 113, 8985–8990 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Pradhan SK et al. EP400 Deposits H3.3 into Promoters and Enhancers during Gene Activation. Mol Cell 61, 27–38 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Papamichos-Chronakis M, Krebs JE & Peterson CL Interplay between Ino80 and Swr1 chromatin remodeling enzymes regulates cell cycle checkpoint adaptation in response to DNA damage. Genes Dev 20, 2437–2449 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 187.van Attikum H, Fritsch O & Gasser SM Distinct roles for SWR1 and INO80 chromatin remodeling complexes at chromosomal double-strand breaks. Embo J 26, 4113–4125 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Cai Y et al. YY1 functions with INO80 to activate transcription. Nat Struct Mol Biol 14, 872–874 (2007). [DOI] [PubMed] [Google Scholar]
  • 189.Wu S et al. A YY1-INO80 complex regulates genomic stability through homologous recombination-based repair. Nat Struct Mol Biol 14, 1165–1172 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Luger K, Mader AW, Richmond RK, Sargent DF & Richmond TJ Crystal structure of the nucleosome core particle at 2.8 A resolution [see comments]. Nature 389, 251–260 (1997). [DOI] [PubMed] [Google Scholar]
  • 191.Davey CA, Sargent DF, Luger K, Maeder AW & Richmond TJ Solvent mediated interactions in the structure of the nucleosome core particle at 1.9 a resolution. J Mol Biol 319, 1097–1113 (2002). [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supp text
FigS1
FigS2
Animation01
Download video file (28.9MB, mov)
Animation02
Download video file (14.8MB, mov)
Animation03a
Download video file (22.4MB, mov)
Animation03b
Download video file (42.1MB, mov)
Animation04
Download video file (33.8MB, mov)
Animation05
Download video file (46.1MB, mov)
Animation06
Download video file (9.6MB, mov)

RESOURCES