Abstract
The association of histones with specific chaperone complexes is important for their folding, oligomerization, post-translational modification, nuclear import, stability, assembly and genomic localization. In this way, the chaperoning of soluble histones is a key determinant of histone availability and fate, which affects all chromosomal processes, including gene expression, chromosome segregation and genome replication and repair. Here, we review the distinct structural and functional properties of the expanding network of histone chaperones. We emphasize how chaperones cooperate in the histone chaperone network and via co-chaperone complexes to match histone supply with demand, thereby promoting proper nucleosome assembly and maintaining epigenetic information by recycling modified histones evicted from chromatin.
ToC blurb
Histone chaperones safeguard the chromatin template and shield histones from promiscuous interactions to ensure their proper storage, transport, post-translational modification, nucleosome assembly and turnover.
Nucleosomes (BOX 1), which restrict DNA accessibility, must be highly dynamic in terms of their positioning and state of assembly to allow access to the base read-out of DNA. The modular nature of nucleosomes provides functional complexity through the incorporation of histone variants and combinatorial post-translational modifications (PTMs), which in turn regulate gene expression and nuclear architecture. Beyond their nucleosomal context, histone proteins require buffering to prevent their aggregation and spurious inter actions with DNA. To meet this requirement, histones are escorted by histone chaperones1. These molecular chaperones guide multiple aspects of histone metabolism, including histone storage, transport, PTM, nucleosome assembly and histone turnover. Although histone chaperones share the common features of binding and shielding histones from promiscuous interactions, they are a diverse group of proteins with low or no sequence similarity and with distinct structural and functional properties.
Box 1. An overview of nucleosome architecture and assembly intermediates.
Eukaryotic DNA is packaged with proteins forming a beads-on-a-string array called chromatin, the primary unit of which is the nucleosome. The nucleosome architecture consists of an octameric configuration of histone fold proteins that wrap ~146 bp of DNA. Each histone protein has a histone fold domain consisting of three helices (α1–α3) linked together by short loops (L1 and L2) that allows for heterodimerization (H2A with H2B and H3 with H4) (see the figure)65. Histone fold dimers can further oligomerize through four-helix bundles65,66, with a four-helix bundle between α2–α3 helices of each copy of H3 forming the so-called H3–H4 tetramerization interface. The deposition of tetrameric H3–H4, forming the ‘tetrasome’, is thought to initiate nucleosome assembly as the H3–H4 tetramer occupies the central portion of DNA (the dyad position) in the nucleosome65. Nucleosome assembly is completed by the addition of two H2A–H2B dimers that wrap the remaining DNA at the entry and exit points of the nucleosome to form 1.7 left-handed super-helical turns of DNA65. Each H2A–H2B dimer associates with the H3–H4 tetramer via a four-helix bundle between H2B α2–α3 and H4 α2–α3 helices65,66. Structures of the H3–H4 tetramer and tetrasome are inferred from structures of the nucleosome65 and histone octamer66. Beyond their histone fold domains, canonical histones include the following features: H2A has a carboxy-terminal extension that includes two short helices (the H2A docking domain) and a short amino-terminal helix (H2A αN); H2B contains a C-terminal α-helix (H2B αC); H3 contains an N-terminal helix (H3 αN); and H2A, H2B, H3 and H4 all contain N-terminal tails. N-terminal histone tails are subject to a plethora of regulatory post-translational modifications, which influence nucleosome dynamics. Other important regulatory elements include the following features: the H2A docking domain, which locks the H3 αN helix in position between the entry/exit and dyad DNA turns65, which is otherwise structurally heterogeneous in the H3–H4 tetramer213; and the H2A–H2B acidic patch, which can bind the basic patch of the H4 N-terminal tail of an adjacent nucleosome86.

DNA replication, transcription and repair are all processes that involve chromatin disruption and restoration, requiring dynamic changes in chromatin assembly states to be coordinated with the DNA processing machinery2–4 (FIG. 1). The challenge of the histone chaperone network is to span these diverse cellular processes5,6, while distinguishing between canonical histones and replacement variants7,8, to meet the demand for de novo histone deposition and chromatin refurbishment throughout the cell cycle and in specialized chromosome domains (FIG. 1). Failure to regulate histone supply can jeopardize functional domains such as telomeres9 and centromeres10–12, alter the barrier for cellular reprogramming13,14 and challenge DNA replication15 and genome integrity16,17. As such, histone chaperones function to safeguard the chromatin template and to ensure that epigenetic information is maintained.
Figure 1. Overview of histone deposition mechanisms.
Newly synthesized histones are incorporated into chromatin via globally and locally acting mechanisms. A network of specialized histone chaperones controls histone delivery and deposition. The figure provides an overview of replication-coupled and replication-independent pathways that require the incorporation of newly synthesized canonical histones and replacement variants, together with parental histone recycling. The histone chaperones that are implicated in each process are listed; for definitions of histone chaperone abbreviations see TABLE 1.
In this Review, we first discuss the mechanistic aspects of histone chaperone function with reference to insights gained from structural studies, which have revealed the multiple properties that histone chaperones have acquired to carry out their functions. We then address how the histone chaperone network is built through histone-dependent co-chaperone complexes and histone-independent chaperone–chaperone interactions. Building on this molecular insight, we discuss how replication-coupled and replication-independent nucleosome assembly pathways combine to maintain chromatin integrity, which has broad implications for genome stability, epigenetic plasticity and disease. For in-depth discussions of variant-specific histone chaperones8,18 and of transcription- and repair-coupled histone dynamics2,4, we refer the reader to recent comprehensive reviews.
Mechanistic insights into chaperoning histones
Our understanding of histone chaperone biology has greatly benefited from structural analyses of histone chaperone complexes (FIG. 2). However, no single feature seems to demarcate a protein as a histone chaperone because both intrinsically disordered regions and structural folds are used to chaperone histones (FIG. 2; TABLE 1). Some structural folds are recognized specifically for their histone chaperone function, including the nucleosome assembly protein 1-like (NAP1-like) and nucleoplasmin (NPM) folds. Other protein folds are not restricted to chaperoning histones, including the immunoglobulin-like (Ig-like) domain, the tetratricopeptide repeat (TPR) domain, the WD40 repeat domain and the pleckstrin homology domain (FIG. 3). However, despite low levels of structural conservation among histone chaperones, some principles of histone chaperone function are beginning to emerge.
Figure 2. Structural features of histone–chaperone complexes.
Two-dimensional depictions of 3D structures in the RCSB Protein Data Bank (PDB). Protein secondary structures are represented by arrows (β-strands), rectangles (α-helices) and black lines (loops). In some cases, loops are represented as a thick line and free DNA-interaction surfaces are indicated by a broken line. For definitions of histone chaperone abbreviations see TABLE 1. A. Histone chaperones binding dimeric H3–H4: (part Aa) Asf1–H3–H4 (PDB identifier: 2HUE)22; (part Ab) HJURP–CENP-A–H4 (PDB ID: 3R45)29; (part Ac) Scm3–Cse4–H4 (PDB ID: 2YFV)30; and (part Ad) DAXX–H3.3–H4 (PDB ID: 4H9N)31, the location of the nucleosomal H3 αN helix (PDB ID: 1AOI)65 is indicated by broken lines. B. Histone chaperones binding tetrameric H3–H4: (part Ba) MCM2–H3–H4 (PDB ID: 5BNV)34 — note that the structure includes two MCM2 HBDs; (part Bb) SPT2–H3–H4 (PDB ID: 5BSA)33; and (part Bc) SPT16 middle domain (SPT16-M)–H3–H4 (PDB ID: 4Z2M)37. C. Histones chaperones binding dimeric H2A– or H2A.Z–H2B: (part Ca) Nap1–H2A–H2B (PDB ID: 5G2E)52; (part Cb) minimal binding domain (MBD) of ANP32E (PDB ID: 4CAY, 4NFT)68,69 and Swr1 (PDB ID: 4M6B)67 with a H2A.Z–H2B chimaera (H2A.Z αC helix extension indicated), and SPT16 with H2A–H2B (PDB ID: 4WNN)46; (part Cc) SPT16-M–H2B chimaera with H2A (PDB ID: 4KHA)41; (part Cd) YL1 in complex with H2A.Z–H2B (PDB ID: 5FUG, 5CHL)71,70; and (part Ce) Chz1–H2A.Z–H2B (PDB ID: 2JSS)82. D. Co-chaperone complexes: (part Da) MCM2–Asf1 chimaera bound to H3–H4 (PDB ID: 5BO0)34; (part Db) the Vps75–Asf1–H3–H4 co-chaperone complex51; (part Dc) UBN1–H3.3–H4–Asf1 (PDB ID: 4ZBJ)93; and (part Dd) MCM2-TONSL bound to a H3–H4 tetramer (PDB ID: 5JA4)84.
Table 1.
Details of histone chap erones
| Histone chaperone | Saccharomyces cerevisiae homologue (or homologues)* | Histone preference | Complex (or complexes) | Domains |
|---|---|---|---|---|
| Anti-silencing function 1A/B (ASF1A/B) | Anti-silencing function 1(Asf1) | H3.1–, H3.2–, H3.3–H4 | Multiple | ASF1, Ig-like, B-domain |
| Minichromosome maintenance protein 2 (MCM2) | Minichromosome maintenance protein 2 (Mcm2) | CENP-A–, H3.1–, H3.2–, H3.3–H4 | MCM2–7 complex | AAA+ ATPase, HBD |
| Tonsuku Like (TONSL) | ND | H3–H4 | MMS22–TONSL complex | ARD, TPR, LRR |
| Retinoblastoma associated protein 46 (RBAP46) | Histone acetyltransferase 2 (Hat2) | H3–H4 | HAT, HDAC, NuRF, NuRD, PRC2 | WD40 |
| Heat shock protein 90A/B (HSP90A and HSP90B) | Heat shock cognate (Hsc82), Heat shock protein 82 (Hsp82) | H1, H2A, H2B, H3, H4 | Heat shock family | HSP90, ATPase-like, TPR binding motif |
| Heat shock cognate 70 (HSC70) | Stress-seventy subfamily A (Ssa1, Ssa2, Ssa3, Ssa4) | H1, H2A, H2B, H3, H4 | Heat shock family | HSC70, ATPase domain, peptide binding domain |
| Somatic nuclear autoantigenic sperm protein (sNASP) | Hat1 interacting factor (Hif1) | H3.1–, H3.2–, H3.3–H4, H1 | H AT | TPR |
| Importin 4 (IPO4) | ND | H3.1–, H3.2–, H3.3–H4 | Nuclear import receptor | HEAT domain, Armadillo-type-fold |
| Suppressor of Ty 2 (SPT2) | Suppressor of Ty 2 (Spt2) | H3–H4 | ND | SPT2, HMG box |
| Suppressor of Ty 6 (SPT6) | Suppressor of Ty (Spt6) | H3–H4 | ND | SPT6 core domain, SH2 |
| ND | Regulator of Ty 1 transposition 106 (Rtt106) | H3–H4 | ND | Dimerization, pleckstrin homology |
| ND | Swi2/snf2-related 1 (Swr1) | H2A.Z–H2B | SWR-C | Swi2/Snf2-related helicase, Swr1-Z domain |
| Acidic-leucine-rich nuclear phosphoprotein 32E (ANP32E) | ND | H2A.Z–H2B | P400–TIP60 | LRR, ZID domain |
| ND | Chaperone for Htz1/H2A–H2B dimer 1 (Chz1) | H2A.Z–H2B | Delivery to SWR-C | H2A.Z-H2B binding |
| Protein YL1 (YL1), vacuolar protein sorting 72 homologue (VPS72) | SWr complex 2 (Swc2) | H2A.Z-H2B | SRCAP/SWR-C, P400–TIP60 | ZID domain |
| Holliday junction recognition protein (HJURP) | Suppressor of chromosome missegregation 3 (Scm3) | CENP-A-H4 (Cse4-H4) | Centromere assembly | Scm3/HJURP |
| Patient SE translocation (SET) | Vacuolar protein sorting 75 (Vps75) | H3–H4 | INHAT, Vps75–Rtt109 | NAP1-like |
| Nucleosome assembly protein 1-like (NAP1L1-6) | Nucleosome assembly protein (Nap1) | H2A–, H2A.Z–H2B, H3–H4, H1 | Nuclear import importin 9, Kap114 | NAP1-like |
| Nucleophosmin (NPM1) | ND | H3–H4, CENP-A–H4, H1 | SWAP | Nucleoplasmin |
| Nucleoplasmin 2 (NPM2) | ND | H2A–H2B | ND | Nucleoplasmin |
| Nucleoplasmin 3 (NPM3) | ND | ND | NPM1 | Nucleoplasmin, NPM1 binding |
| Nucleolin (NCL) | ND | H2A–H2B, H1 | SWAP | RNA binding, DNA helicase/ATPase |
| Chromatin assembly factor 1, subunit A (CHAF1A or p150) | Chromatin assembly complex 1 (Cac1) | H3.1–H4 | CAF1 | PIP-box, p60 binding, HP1-binding, WH domain |
| Chromatin assembly factor 1, subunit B (CHAF1B or p60) | Chromatin assembly complex 2 (Cac2) | H3.1–H4 | CAF1 | WD40, B-domain |
| Retinoblastoma associated protein 48 (RBAP48) | Multi-copy suppressor of IRA1 (Msi1) | H3.1–, H3.2–, H3.3–H4 | CAF1, HDAC, NuRF, NuRD, and PRC2 | WD40 |
| Histone regulation A (HIRA) | Histone regulation (Hir1, Hir2) | ND | HIRA/HIR | WD40, B-domain |
| Ubinuclein 1 (UBN1) | Histone periodic control (Hpc2) | H3.3–H4 | HIRA/HIR | UBN1, HRD |
| Calcineurin-binding protein cabin-1 (CABIN1) | Histone regulation (Hir3) | ND | HIRA/HIR | TPR |
| Suppressor of Ty 16 (SPT16) | Suppressor of Ty 16 (Spt16) | H2A–H2B, H3–H4 | FACT | Peptidase-like, pleckstrin homology |
| Structure-specific recognition protein 1 (SSRP1) | POl1 binding 3 (Pob3)–non-histone protein 6 (Nhp6A/B) | H2A–H2B, H3–H4 | FACT | Pleckstrin homology, HMG box |
| Death domain-associated protein 6 (DAXX) | ND | H3.3–H4 | DAXX–ATRX | HBD, four-helix bundle, SIM, PML targeting |
| Alpha-thalassaemia/mental retardation syndrome X-linked (ATRX) | ND | ND | DAXX–ATRX | Swi/Snf2-related helicase, ADD, HP1 binding, MeCP2 binding |
AAA+, ATPases associated with diverse cellular activities; ADD, ATRX-DNMT3-DNMT3L domain; ARD, ankyrin repeat domain; B-domain, Asf1 binding domain; CAF1, chromatin assembly factor 1 complex; FACT, facilitates chromatin transcription; HAT, histone acetyltransferase; HBD, histone-binding domain; HDAC, histone deacetylase; HEAT, domain also present in Huntington, Ef3, protein phosphatase 2A and Tor1; HIR, histone regulation complex (yeast); HIRA, histone regulation A complex (human); HMG-box, high mobility group box; HRD, Hpc2-related domain; HP1, heterochromatin protein 1; Ig-like, immunoglobulin-like; INHAT, inhibitor of acetyltransferase; Kap114, karyopherin 114; LRR, leucine-rich repeat; MCM2–7, the complex of minichromosome maintenance proteins 2, 3, 4, 5, 6 and 7; MeCP2, methyl-CpG-binding protein 2; MMS22L, methyl methanesulfonate sensitivity 22-like; NAP1-like, nucleosome assembly protein 1-like; ND, not defined; NuRD, nucleosome remodelling and histone deacetylase; NuRF, nucleosome remodelling factor; PIP-box, PCNA-interacting protein box; PML, promyelocytic leukaemia nuclear body; PRC2, polycomb repressive complex 2; Rtt109, regulator of Ty 1 transposition 109; SIM, SUMO-interacting motif; SH2, Src homology 2 domain; SRCAP, Snf-2-related CREB-binding protein activator protein; SWAP, switch-associated protein complex; SWR-C, Swr1 complex; Swi2/Snf2, switch/sucrose nonfermenting 2; TIP60, 60 kDa Tat-interactive protein; TPR, tetratricopeptide repeat; WD40, 40 amino acid repeat terminating in tryptophan-aspartic acid; WH, winged helix; ZID, H2A.Z-interacting domain.
Yeast and human homologue (or homologues) were identified using http://yeastmine.yeastgenome.org and functional evidence in the literature.
Figure 3. Other histone chaperone structures.
Two-dimensional depictions of 3D structures in the RCSB Protein Data Bank (PDB). Histone chaperones are shown in yellow. For definitions of histone chaperone abbreviations see TABLE 1. A | Immunoglobulin-like folds: (part Aa) Asf1 (PDB identifier: 1ROC)21, showing the location of B-domain interactions with Rad53 (PDB ID:2YGV)97, Cac2/CAF1 p60 (PDB ID: 2Z3F)94, Hip1/Hir1/HIRA (PDB ID: 2Z34, 2I32)94,95 and codanin 1 (REF. 96); and (part Ab) the YEATS domain of Yaf9 (PDB ID: 3FK3)24. B | Tetratricopeptide repeat of Hif1 (PDB ID: 4NQ0)27. C | Tandem pleckstrin homology domains (PH1 and PH2) of Rtt106 (PDB ID: 3GYP)39. D | Nucleosome assembly protein 1 (Nap1)-like folds of (part Da) the Nap1 dimer (PDB ID: 2AYU)49; (part Db) Vps75 tetramer (PDB ID: 5AGC)51; and Vps75–Rtt109 complexes with (part Dc) 2:1 (PDB ID: 3Q66)60 and (part Dd) 2:2 (PDB ID: 3Q35)58 stoichiometries. E | Nucleoplasmin (NPM) fold of Xenopus laevis NO38 (PDB ID: 1XB9)62. F | WD40 repeat domains of RBAP46 and RBAP48 proteins with the H4 α1 (PDB ID: 3C9C, 3CFV)74,75 and the H3 tail (PDB ID: 2YBA)77.
Shielding functional histone interfaces
One means of regulating nucleosome assembly is to shield the histone surfaces that associate with DNA and other histones in the nucleosome (BOX 1). Thermodynamic studies have indicated that only proper nucleosomal contacts between histones and DNA are able to compete with the interactions between histones and their chaperones19,20. Thus, in this way, histone chaperones can promote proper nucleosome formation by eliminating non-nucleosomal histone–DNA interactions19. Structural insights into the mode of interaction between histone chaperones and histones have revealed surprising aspects of histone chaperone biology.
The first histone chaperone complex to be crystallized was Asf1, which has an Ig-like fold21, in complex with H3–H4 (REFS 22,23). This structure showed that Asf1 binds H3–H4 dimers through association with the H3 α2–α3 helices (FIG. 2Aa), which constitute the H3–H4 tetramerization interface (BOX 1). Surprisingly, the structurally related YEATS domain24 (FIG. 3Ab) has been characterized as a reader module for the H3 tail25,26, whereas the unrelated TPR domain of sNASP (Hif1 in yeast)27 (FIG. 3B) also binds the H3 α3 helix28. HJURP29, Scm3 (REF. 30) and DAXX31,32 bind dimers of H3–H4 in a variant-specific manner (reviewed in REF. 18) but their histone-binding domains (HBDs) are structurally unrelated to those of Asf1 (FIG. 2A) and sNASP. Both HJURP and the yeast homologue Scm3 form coiled-coil interactions with the α2 helix of H3-like centromere protein A (CENP-A) (Cse4 in yeast) in a strikingly similar manner to the association of DAXX with the α2 helix of histone H4 (REF. 31) (FIG. 2A). Thus, there seems to be some intrinsic requirement to chaperone newly synthesized H3–H4 and variants thereof as dimers, and histone chaperones achieve this goal in many ways. However, although they chaperone the H3–H4 tetramerization interface, these histone chaperones do not efficiently shield the DNA-binding surface of H3–H4 (FIG. 2A). Therefore, dimer-specific chaperones probably either collaborate with additional chaperones to shield the DNA-binding surface of H3–H4 (in the case of Asf1) or participate directly in H3–H4 dimer deposition (in the case of HJURP, Scm3 and DAXX).
More recently, structures of the HBDs of MCM2 and SPT2 in complex with H3–H4 tetramers have been solved (FIG. 2Ba,b). These structures show that both chaperones simultaneously shield both the DNA-binding and H2A–H2B-binding interfaces of H3–H4 (REFS 33–35). In the case of MCM2, ASF1 can split the H3–H4 tetramer to form a complex in which MCM2 and ASF1 co-chaperone a H3–H4 dimer (REF. 34) (FIG. 2 Da). In the case of SPT2, the binding to a H3–H4 tetramer seems to be more direct through recognition of the H3–H4 tetramerization interface (FIG. 2Bb). However, the majority of contacts between SPT2 and the H3–H4 tetramer are asymmetrically loaded on one dimer of H3–H4, with a relatively small helical portion of SPT2 reaching across the tetramerization interface to capture the second dimer33 (FIG. 2Bb). Therefore, it remains possible that SPT2 also binds a H3–H4 dimer in a co- chaperone complex, analogous to that formed by MCM2 and ASF1 (REF. 34). In the absence of ASF1, H3–H4 is tetrameric and each dimer is bound by a MCM2 HBD. An intriguing possibility is therefore that one MCM2 may disengage the H3–H4 tetramer, revealing half of the DNA-binding surface that is required for deposition on DNA. The H3–H4 tetramer would thus remain tethered to MCM2 during the initial stages of DNA deposition; SPT2 might also deposit tetrameric H3–H4 via a similar mechanism.
The facilitates chromatin transcription (FACT) complex36 (TABLE 1) binds the lateral surface of H3–H4 tetramers through the middle domain of SPT16 (SPT16-M), which forms a tandem arrangement of pleckstrin homology domains37 (FIG. 2Bc). Interestingly, this binding mode enables SPT16-M to bind a partially disassembled nucleosome in which one H2A–H2B dimer is lost37 and, potentially, a H3–H4 tetramer in a co-chaperone complex with MCM2 (REFS 37,38). Whether the tandem pleckstrin homology domains of Rtt106 (Rtt106-M)39 (FIG. 3C), Pob3 (Pob3-M)40 and the Spt16–Pob3 dimer interface (Spt16-D–Pob3-N)41 recognize H3–H4 in a similar manner to SPT16-M is currently unclear. Rtt106-M exhibits specificity for H3–H4, and this domain recognizes histone H3 acetylated at Lys56 (H3K56ac)42,43. Dimerization is necessary for the function of Rtt106 in yeast44, but this process is mediated by a novel protein fold rather than its pleckstrin homology domain43. Histone H3–H4-binding activity is also found in the peptidase-like fold of SPT16 (SPT16-N)45 and SPT16 can also bind H2A–H2B36,41,46. Structural data on the mode of interaction between SPT16 and H2A–H2B seem to be contradictory46, with a conserved peptide motif of SPT16 (REF. 46) (FIG. 2Cb) and SPT16-M41 (FIG. 2Cc) both binding the same region of H2A–H2B. It is not yet clear whether these simultaneously incompatible binding modes reflect different functional states of the FACT complex. An important topic for future investigations will be to understand how the multiple histone-binding properties of the FACT complex might support the invasion and unravelling of nucleosomes and couple histone eviction with histone recycling (reviewed in REF. 47).
The NAP1-like fold is a constitutive homodimer with a headphone-like organization48–50 (FIG. 3D). Each NAP1-like fold dimer binds a single histone dimer — as has been shown for binding of Vps75 or Nap1 to H3–H4 (REF. 51) and for binding of Nap1 to H2A–H2B52 — in a manner that does not obscure the H3–H4 tetra-merization interface51. As such, Vps75 and Nap1 can bind H3–H4 tetramers53 or H3–H4 dimers in complex with Asf1 (REF. 51) (FIG. 2Db). The reported cellular functions of Vps75 and Nap1 are with H3–H4 and H2A–H2B, respectively, although both chaperones are capable of binding both H2A–H2B and H3–H4 in vitro54,55. Vps75 and, to a lesser extent, Nap1 are capable of forming tetrameric ring-like assemblies51,56 and they further oligomerize upon histone binding51,52,57. Tetramerization of Vps75 (FIG. 3Db) imparts a “self-chaperoning” function51,56 as it shields key binding surfaces for H3–H4 (REF. 51) (FIG. 2Db) and the acetyltransferase Rtt109 (REFS 58–60) (FIG. 3Dc,d). Histone binding reconfigures the Vps75 tetramer51 potentially providing a cooperative binding mechanism to accommodate a H3–H4 tetramer. Furthermore, Vps75 partially shields the DNA-binding surface of H3–H4, leaving the dyad DNA-binding surface exposed51, which may explain its activity in promoting tetrasome assembly53. Similarly, Nap1 shields the DNA-binding surfaces of H2A–H2B52 (FIG. 2Ca), possibly ensuring that proper contacts are made between H2A–H2B and the tetrasome before releasing the DNA-binding surface of H2A–H2B. NPM-fold histone chaperones also oligomerize, forming pentameric rings61,62 (FIG. 3E) that are capable of binding multiple copies of various histone subtypes (TABLE 1) and promote tetrasome assembly63,64.
Sequestering histones in non-nucleosomal conformations
In addition to shielding the functional interfaces of histones, histone chaperones can also trap histones in conformations that are not observed in the context of nucleosomes65. In doing so, these chaperones may promote the correct assembly of the histones onto nucleosomal DNA by directing the path of nucleosomal DNA towards the correct trajectory around the histones. One demonstration of this concept is the association of DAXX with H3.3–H4 (REFS 31,32). In this structure, the αN helix of H3.3 is rotated 180° compared with its location in the histone octamer66, which probably results in a significant reorientation of the H3.3 tail31,32 (FIG. 2Ad). As the H3 αN helix sits in a pivotal position in the nucleosome, making contacts with both dyad and entry/exit DNA (BOX 1), controlling the trajectory of this helix may be an important feature of nucleosome assembly and disassembly. The dynamics of the H2A docking domain, which supports the H3 αN helix in the nucleosome65, may also be under the control of histone chaperones. When bound by Swr1, ANP32E and YL1, an extra helical turn is observed in the carboxy-terminal helix of H2A.Z, which could potentially affect the trajectory of the H2A.Z docking domain67–71 (FIG. 2Cd). The interplay between the H3 αN helix and the H2A docking domain is also thought to be a key feature of FACT complex-mediated eviction of H2A–H2B from nucleosomes37.
In addition to the already mentioned interaction of ASF1 with the tetramerization interface of H3–H4 (REFS 22,23), ASF1 makes additional contacts with the C terminus of histone H4 in what has been termed a “strand capture” mechanism22 (FIG. 2Aa). In the nucleosome, the C-terminal β-strand of H4 associates in a parallel manner with a short β-strand of H2A65. When in complex with ASF1, the same region of H4 turns 90° in a hinge-like manner to associate in an anti-parallel fashion with a short β-strand at the C terminus of the ASF1 globular domain22,23. This ‘strand capture’ feature of ASF1 may facilitate nucleosome disassembly in the context of larger complexes, as ASF1 alone cannot disengage H3–H4 from DNA72. Interestingly, DAXX anchors the H4 C terminus in another conformation distinct to that observed in complex with ASF1 (REF. 31). It is not known how sequestration of the H4 C terminus influences nucleosome assembly or disassembly processes, but flexibility in the H4 C terminus must be an important feature of histone dynamics as the H4 G94P mutation, which restricts this mobility, compromises yeast viability and chromatin structure73.
In most of the crystal structures of histone–chaperone interactions, the histone fold is remarkably stable and similar to that observed in the nucleosome (BOX 1; FIG. 2). However, the interaction of the WD40 repeats of RBAP46 and RBAP48 with the α1 helix of H4 (REFS 74,75) (FIG. 3F) would require a significant conformational change within the histone fold domain of H3–H4. Evidence of such a conformational rearrangement has been reported and is proposed to promote the handover of H3–H4 from ASF1 to RBAP48 (REF. 76) in the chromatin assembly factor 1 (CAF1) complex. Interestingly, the WD40 repeats of RBAP46 and RBAP48 can also bind the H3 tail77 through a different interface (FIG. 3F). However, these binding abilities are not necessarily always used. For example, the interaction of RBAP46 with H4 α1 helix directs the histone acetyltransferase activity of the HAT1 complex towards non-nucleosomal H3–H4 (REF. 78). However, this interaction is blocked in the poly-comb repressive complex 2 (PRC2) methyltransferase, in which Su(z)12 occupies the H4 α1 helix-binding surface of RBAP46 and RBAP48 and in turn directs PRC2 to a nucleosomal substrate77. In addition, RBAP46 may be re-directed to bind CENP-A–H4 dimers through a co-chaperone interaction with HJURP, as has been observed for orthologous yeast proteins79,80.
Structural flexibility is not only restricted to the histones within histone–chaperone complexes. Indeed, the structures of MCM2 (REFS 34,35,81), SPT2 (REF. 33), Chz1 (REF. 82) and YL1 (REFS 70,71) (FIG. 2) imply that the chaper-one undergoes a marked transformation upon histone binding. Furthermore, the HBD of DAXX is predominantly unfolded in the absence of its histone cargo83; this is also expected to be the case for the minimal HBDs of SPT16, ANP32E and Swr1 (REFS 46,67–69) (FIG. 2Cb).
Co-chaperone relationships
Histone chaperones generally shield functional surfaces of histones without fully encasing the histone fold. The combination of multiple histone chaperones therefore has the potential to form a more complete shield around the histone fold as part of a co-chaperone complex. This concept is illustrated by the MCM2–H3–H4–ASF1 complex, in which MCM2 shields the DNA- and H2A–H2B-binding surfaces of H3–H4, whereas ASF1 occludes H3–H4 tetramerization34,81 (FIG. 2Da). Despite shielding many of the interfaces required for nucleosome assembly, the MCM2–H3–H4–ASF1 complex further promotes nucleosome assembly compared with ASF1–H3–H4 and MCM2–H3–H4 alone34. An explanation for this observation is not yet available, but it may be that there are synergistic effects of co-chaperoning during H3–H4 deposition. Interestingly, histones bound to MCM2 and ASF1 can also engage TONSL84, which has a dual function as a histone reader and a histone chaperone84,85. The ankyrin repeat domain (ARD) of TONSL recognizes the basic patch of the H4 tail (K16–K20) when K20 is unmethylated (H4K20me0)84 (FIG. 2Dd). The TONSL ARD holds the H4 tail perpendicular to the DNA-binding surface of the H3–H4 tetramer and can also bind nucleosomal H4 (REF. 84). As such, TONSL ARD recognition of the H4 tail may help to guide nucleosomal DNA across the H3–H4 tetramer during deposition and potentially counteract binding of the H4 tail to the acidic patch of H2A–H2B on neighbouring nucleosomes86 to prevent premature chromatin compaction84.
Highlighting its central role in chaperoning H3–H4, ASF1 forms a large number of co-chaperone complexes with H3.1, H3.2, H3.3 and one or more of the chaperones Vps75, sNASP, RBAP46, RBAP48, MCM2 and the reader and chaperone TONSL51,76,84,85,87–91 (FIG. 4). Curiously, sNASP, RBAP46 and ASF1 seem to be capable of engaging histones H3–H4 in a single co-chaperone complex87,89–91. However, ASF1 and sNASP may have overlapping binding sites at the H3–H4 tetramerization interface22,23,28. Therefore, for sNASP and ASF1 to engage histones at the same time (the interaction between ASF1 and sNASP is histone dependent87), additional H3–H4-binding sites may be present in sNASP to allow the partial handover of the H3–H4 tetramerization interface to ASF1. As such, the interplay among sNASP, ASF1 and H3–H4 would benefit from further biochemical and structural characterization. In addition, sNASP forms a complex with HSP90 (REF. 90), a molecular chaperone that assists in the folding of many proteins presented to HSP90 by TPR domain-containing co-chaperones92. Thus, sNASP may collaborate with HSP90 to assemble the H3–H4 dimer before participating in other co-chaperone complexes.
Figure 4. The histone supply network.
After synthesis, histones H3–H4 and H2A–H2B engage with multiple different chaperones and enzymes on their way to chromatin. ASF1 is a central chaperone in the delivery of newly synthesized H3.1–, H3.2– or H3.3–H4 dimers. Histone-free ASF1 is phosphorylated by tousled-like kinases (TLKs), promoting histone binding. Histones bound by ASF1 are engaged in multiple different co-chaperone complexes, which further shields their functional interfaces and facilitates their acetylation, nuclear import and storage. Co-chaperoning may be a general paradigm, supporting branching of the pathway and the modification of histones ‘on the go’. ASF1 shuttles H3.1/H3.2–H4 and H3.3–H4 dimers to the CAF1 and HIRA complexes for replication-coupled and replication-independent deposition, respectively. H3.3–H4 dimers are also deposited by DAXX–ATRX, whereas the H2A–H2B supply is handled by NAP1 and FACT. Histone chaperones also have an important role in the feedback regulation of histone supply. Although several histone chaperones are implicated in H3–H4 delivery with ASF1, it is still unclear whether other chaperones collaborate with DAXX and NAP1 in the delivery of H3.3–H4 and H2A–H2B dimers, respectively. See TABLE 1 for definitions of histone chaperone abbreviations.
Other chaperone–chaperone interactions are direct and independent of histones and as such fall outside the definition of a co-chaperone relationship; however, these interactions are nevertheless important. Histone-independent interactions between chaperones may function to regulate the handover of histones from one chaperone to another. These interactions could also help to recruit histone chaperones with different abilities to sites of nucleosome assembly or disassembly and chromatin remodelling. The interactions of ASF1 with RBAP48 (REF. 87), and probably UBN1 (REF. 93) (FIG. 2Dc), are histone dependent. By contrast, other subunits of the CAF1 complex and the histone regulation complex (HIRA/HIR; see TABLE 1) bind to ASF1 directly via a small epitope termed the B-domain, which interacts with the β-sheet sandwich on the opposite side of H3–H4 (REFS 21,94,95) (FIG. 3Aa; TABLE 1). Other factors, including codanin 1 (REF. 96) and Rad53 (REF. 97), also bind ASF1 via B-domains, which suggests that these factors might compete with CAF1 and HIRA complexes in the histone delivery pathway. However, it remains a key open question how the interplay between histone-dependent and chaperone–chaperone interactions facilitates the presumed handover of histones from ASF1 to the CAF1 or HIRA complex.
Chaperone networks in histone supply
During chromatin assembly and histone turnover, new histones need to be synthesized, processed and delivered to specific sites in the genome. This process is orchestrated by histone chaperones, which handle histones from the time of synthesis in the cytoplasm to their delivery to chromatin.
Matching histone supply with demand
Several layers of regulation govern the histone supply chain to meet the demand for nucleosome assembly at any given time while keeping the pool of soluble histones at a minimum. This regulation is most apparent during chromatin replication, when the histone content of the cell is doubled. The high level of new histone synthesis required to support chromatin replication is matched by high expression rates of the canonical histones genes, which are present in multiple copies in metazoans (reviewed in REF. 7). In budding yeast, an attractive negative feedback model has been proposed whereby histone chaperones monitor the level of soluble histones to dynamically control histone gene expression (reviewed in REF. 98). According to this model, the yeast histone regulation complex (Hir), Asf1 and Rtt106 assemble at a cis-regulatory element and, presumably in response to high levels of soluble histones, form a repressive chromatin structure that silences histone genes beyond S phase98. If the capacity of the histone chaperones is saturated, the Rad53 checkpoint kinase targets excess histones for ubiquitylation and proteasomal degradation17,99 through a mechanism that may involve its ability to bind histone-free Asf1 (REF. 97). Feedback regulation remains to be explored in other organisms, but there are indications that the roles of HIRA and ASF1 in histone gene regulation may be conserved100,101.
S phase active kinases also play a role in the spatio-temporal regulation of histone provision and nucleosome assembly. ASF1 phosphorylation by TLK1 and TLK2 promotes histone binding and interaction with downstream chaperones102,103 to ensure efficient histone supply during DNA replication. Phosphorylation of CAF1 p150 by the CDC7–DBF4 kinase disrupts dimerization, which in turn facilitates binding to PCNA104 and chromatin assembly. By contrast, CDK1- and CDK2-mediated phosphorylation of HJURP inhibits its recruitment to centromeres and prevents premature CENP-A incorporation105. These examples are probably just the tip of the iceberg, and histone chaperone PTMs are likely to emerge as an important means to regulate histone handover, deposition and eviction.
Chaperone functions in histone supply
Isolation and proteomic analyses of histone chaperone complexes have revealed that histones engage with multiple chaperones on their way to chromatin85,87,89,90,106–108 (FIG. 4). The principles of how chaperones function together in a network to optimize histone delivery and storage are beginning to emerge5, and a key feature seems to be that multiple chaperones can engage with a single histone dimer, as highlighted in FIG. 4. Soluble H3 and unacetylated H4 have been found together with heat shock proteins HSC70 and HSP90 (REFS 90,109), which suggests that these are early acting factors that assist histone folding before RBAP46–HAT1-mediated diacetylation of H4K5 and H4K12 (H4K5acK12ac)110 (FIG. 4). RBAP46–HAT1 forms a conserved histone-dependent complex with sNASP90,110, which functions in a storage capacity for soluble H3–H4 and counteracts histone degradation through HSC70- and HSP90-mediated autophagy111. The RBAP46–HAT1–H3–H4–sNASP co-chaperone complex is present in both the cytoplasm and the nucleus85,87,90,112, which suggests that the ability to shuttle histones into storage and maintain the H4K5acK12ac mark is important in both cellular compartments.
ASF1 forms a complex with H3–H4 and importin 4 (REFS 87,90,109), which suggests that ASF1 binds histones in the cytoplasm and transports them to the nucleus for handover of H3.1/H3.2–H4 to the CAF1 complex and H3.3–H4 to the HIRA complex for deposition88,95,107,113–117. The major function of ASF1 is presumably to prevent premature tetramerization of H3–H4 (REFS 22,23) while allowing H3–H4 dimers107,118 in transit to engage with additional chaperones87,90 (FIG. 4). ASF1 can bind H3–H4 within the RBAP46–HAT1–H3–H4–sNASP co-chaperone complex91, which may provide ASF1 with access to both the cytoplasmic and nuclear stores of H3–H4. Most vertebrates have two ASF1 paralogues, ASF1A and ASF1B, which differ primarily in their amino- and carboxy-terminal regions outside the core histone-binding domain95,119. ASF1A is ubiquitously expressed in all tissues and throughout the cell cycle119, whereas ASF1B is expressed in S phase in an E2F-dependent manner120. ASF1A preferentially interacts with the HIRA complex, whereas ASF1B binds preferentially to the CAF1 complex95,121. However, ASF1A and ASF1B can function redundantly in chromatin replication because the depletion of both paralogues (in contrast to single depletion) arrests DNA replication in human cell lines88,121, similar to depletion of CAF1 (REFS 122–124). This effect phenocopies the inhibition of histone biosynthesis15, which argues that ASF1 (ASF1A and ASF1B) and CAF1 have non-redundant functions in replication-coupled histone provision and deposition.
New H3–H4 dimers are also found in a large co-chaperone complex that includes ASF1, MCM2 and TONSL–MMS22L84,87,89. Binding to MCM2 can stabilize H3–H4, but it is not required for the transfer of histones to CAF1 for replication-coupled histone deposition34. MCM2 may therefore provide an additional storage site for new H3–H4 dimers both in solution (FIG. 4) and as part of inactive chromatin-bound MCM2–7 hexamers34,87. TONSL, which, together with MMS22L, promotes homologous recombination125, specifically binds new histones by recognizing the H4 tail unmodified at K20 (H4K20me0)84. However, whether TONSL function is required for histone delivery or deposition remains unknown. Nevertheless, given its role as a histone reader, TONSL may utilize the new H3–H4 delivery pathway to be loaded onto replicated DNA where it can promote error-free DNA repair in case a lesion is encountered by the replication fork84.
Similar to H3–H4, several histone chaperones are implicated in the nuclear import and delivery of H2A–H2B dimers (FIG. 4), but the chaperone network for these histones is less well understood. Notable histone chaperones include Nap1 (REF. 55), Chz1 (REF. 106), ANP32E68,69, YL1 (REFS 70,71) and FACT36. Nap1 is a multifunctional chaperone that shields the DNA-binding interfaces of histones52 and prevents unscheduled accumulation of H2A–H2B on DNA19. Nap1 contributes to the nuclear import of both H2A–H2B and H2A.Z–H2B, together with Kap114 (importin 9)126, and Nap1 probably delivers H2A–H2B dimers to sites of both ongoing transcription and DNA replication. Chz1 is localized predominantly in the nucleus and specifically chaperones H2A.Z–H2B dimers in yeast106. Interestingly, Chz1 has overlapping functions with Nap1 in handling the H2A.Z variant106, highlighting a degree of redundancy in the H2A–H2B delivery network. Furthermore, if Chz1 and Nap1 are absent, FACT is able to substitute these two chaperones to compensate and deliver H2A.Z–H2B to the Swr1 remodelling complex (SWR-C) for deposition106. In human cells, H2A.Z–H2B is handled by the variant-specific chaperones ANP32E68,69 and YL1 (REFS 70,71); of these two Swr1-related complexes in mammals, YL1 functions in both SRCAP and P400–TIP60 (REF. 127), whereas ANP32E seems to be specific for p400–TIP60 (REF. 69).
Histone modifications during supply
Newly synthesized histones can be modified by various PTMs, which subsequently can affect chaperone binding, histone deposition and final chromatin state. The H4K5acK12ac mark is both highly conserved128 and abundant (being present on ~80% of total soluble H4 (REF. 129) and ~98% of ASF1-bound H4 (REF. 87) in HeLa cells). Paradoxically, the exact function of this mark is unclear, although it has been implicated in nuclear import, replication-coupled nucleosome assembly, chromatin maturation and replication fork repair (reviewed in REFS 5,128). In yeast, H3K56ac is a highly abundant mark on newly synthesized histones130. H3K56ac is imposed by Rtt109 on Asf1-bound histones (reviewed in REF. 54) and it promotes association with the downstream chaperones Rtt106 and CAF1 and thus facilitates nucleosome assembly42,43,131. This function does not seem to be conserved in mammalian cells, as H3K56ac is only present on ~1% of newly synthesized H3 before87 and after deposition132,133. In yeast, H3 ubiquitylation by Rtt101–Mms1 can also promote histone handover from Asf1 to downstream chaperones134 by reducing the affinity of H3 for Asf1. Similarly, the ubiquitylation of CENP-A by CUL4–RBX1–COPS8 promotes its interaction with HJURP and facilitates the deposition of CENP-A at centromeres135.
A key consideration regarding the PTM of soluble histones is whether these modifications are universally imposed during delivery or are restricted to histones assembled into particular chromatin domains. In metazoans, H3 can be monomethylated at Lys9 (REF. 129) during translation136 and replication137, and only a fraction of newly synthesized histones carry this modification (5–30%)87,129,132. By functioning as a precursor for H3K9me3, monomethylation may contribute to heterochromatin establishment (reviewed in REF. 3); however, it is unknown whether monomethylation has a function in histone delivery or how it is targeted to heterochromatin sites.
Chaperones in histone deposition and recycling
The activity of histone chaperones is measured by their ability to mediate the deposition of histones onto DNA in vitro. However, in vivo, some chaperones are restricted to handling histones at earlier stages in the delivery pathway and only a few chaperones are currently recognized as actual deposition factors. It remains unclear what functional properties are required to assemble nucleosomes, but shared features of the chaperones involved in this process include specificity for histone variants (reviewed in REF. 18) and recruitment to chromatin through a mechanism that is linked to their mode of action.
De novo histone deposition
Nucleosome assembly factors are recruited to their sites of action through highly distinct mechanisms that can be coupled to a process (such as DNA replication, repair or transcription) or linked to chromatin features (such as histone marks, transcription factors or free DNA). For example, the CAF1 complex mediates replication-coupled nucleosome assembly of H3.1/H3.2–H4 by associating with the replication machinery through its largest subunit (p150), which binds the homotrimeric sliding-clamp PCNA138 (FIG. 5a). In this way, newly synthesized DNA is almost immediately assembled into nucleosomes in a manner that makes the processes of DNA replication and nucleosome assembly inter-dependent15,139. The interaction with PCNA can also recruit CAF1 to assemble nucleosomes at sites of DNA repair140,141 (FIG. 5b), although CAF1 has been observed to function independently of PCNA and DNA synthesis during double-stranded break repair142,143. In yeast, Rtt106 can also mediate replication-coupled H3–H4 deposition144 (FIG. 5a), but the recruitment mechanism for Rtt106 is less well defined and may involve concomitant interactions with DNA39, Cac1 (CAF1 p150)144 and FACT145.
Figure 5. Recruitment of histone chaperones to chromatin.
Histone chaperones recognize specific factors or features of chromatin to target canonical and variant histones to designated genomic loci. The CAF1 complex binds to PCNA to promote deposition during DNA replication (part a) and DNA repair (part b), whereas the HIRA complex may be recruited by naked DNA (part b). At transcription start sites (TSSs), Swr1 as part of the SWR-C complex recognizes the nucleosome-depleted region and exchanges H2A–H2B for H2A.Z–H2B at the −1 and +1 nucleosome (part c). Furthermore, the HIRA complex interacts with RNA polymerase II (Pol II) and transcription factors, which could facilitate H3.3–H4 deposition in gene bodies and at promoters, respectively. SPT2, SPT6 and potentially Vps75 mediate recycling of H3–H4 during ongoing transcription and SPT6 can be recruited by binding the phosphorylated carboxy-terminal repeat domain of Pol II. NAP1 and FACT are also recruited to transcription sites and can facilitate histone H2A–H2B eviction. In heterochromatin (telomeres, pericentromeres and repetitive elements), DAXX–ATRX is recruited through ATRX-mediated binding to H3K9me3 (part d). At centromeres, HJURP recruitment requires priming by the MIS18 complex together with core centromeric components (part e). For definitions of histone chaperone abbreviations, see TABLE 1.
Replication-independent incorporation of newly synthesized H3.3 may occur through a histone exchange reaction or via a gap-filling mechanism. Analysis of H3.3 occupancy on DNA by chromatin immunoprecipitation followed by sequencing (ChIP-seq) shows that this histone variant is enriched at cis-regulatory elements, gene bodies and telomeres, with turnover rates being highest at active promoters and enhancers (reviewed in REFS 2,146). The underlying mechanism of H3.3 turnover depends on the genomic location and is governed by the HIRA complex in transcribed regions and promoters and by DAXX–ATRX in heterochromatin at pericentromeres and telomeres108,147–152 (FIG. 5c,d). The HIRA complex has multiple modes of recruitment through interactions with RNA polymerase II150,151 and transcriptional regulators150,152,153 (FIG. 5c). In addition, the HIRA complex performs a gap-filling function that may involve interactions with naked DNA151. Although further experiments are required to resolve the role of DNA binding in the functions of the HIRA complex, this mode of recruitment could explain HIRA-mediated deposition of H3.3–H4 at DNA repair sites154 (FIG. 5b). The histone chaperone DAXX mediates H3.3–H4 deposition in heterochromatic regions via interactions with the SWI/SNF-like chromatin remodeller ATRX108,147,149. ATRX specifically recognizes H3K9me3 through its ADD (ATRX-DNMT3-DNMT3L) domain155–157, providing a means to direct DAXX–ATRX-dependent H3.3–H4 deposition to heterochromatic loci (FIG. 5d). How histones H3.3–H4 are delivered to ATRX is currently unclear, but the presence of DAXX in the cytosolic fraction and its interaction with RBAP46 or RBAP48 (REF. 108) suggest that it could chaperone H3.3–H4 during nuclear import.
Directed deposition of CENP-A–H4 is required for centromere function and for faithful chromosome segregation (reviewed in REF. 10). In budding yeast, a single Cse4 (CENP-A)-containing nucleosome is found at a sequence-defined centromere. During DNA replication, the old Cse4-containing nucleosome is removed and a new one is installed by the Scm3 chaper-one158–163. In fission yeast and metazoans, the presence of CENP-A-containing chromatin, rather than the DNA sequence, defines the centromere and, in this case, the old CENP-A is recycled during DNA replication164,165. New CENP-A is not incorporated in a replication-dependent manner; rather, it is deposited by HJURP through an exchange reaction before the next round of DNA replication (to counteract the dilution with H3–H4 taking place during replication)165–168. Scm3 is recruited to DNA in a sequence-dependent manner and stays with the CENP-A-containing nucleosome throughout the cell cycle158,163, whereas HJURP-dependent deposition of CENP-A is cell cycle regulated105,166,167 and requires several constitutive centromere components as well the MIS18 complex (FIG. 5e) and RBAP46 or RBAP48 (reviewed in REF. 10). Given that nucleosomal CENP-A itself directs the recruitment of centromeric core proteins and MIS18 (REFS 169,170), this process represents a self-sustaining epigenetic mechanism for centromere maintenance.
Although the principles of histone variant specificity and recruitment of chaperones are emerging, understanding the actual histone deposition process remains a major challenge. However, H2A.Z–H2B deposition by the SWR-C remodelling complex is particularly well understood, as this process has been extensively characterized by biochemical investigations67,171–174. SWR-C catalyses a two-step, ATP-dependent histone exchange reaction triggered by the presence of H2A-containing nucleosomes and non-nucleosomal H2A.Z174. This process involves two distinct chaperone activities of SWR-C, provided by the catalytic subunit Swr1 (REFS 67,172) and the accessory subunit Swc2 (REF. 171) (YL1 (REFS 70,71)). Interestingly SWR-C function is directed to nucleosomes flanking the transcription start site due to the preference of SWR-C for nucleosomes with long linker DNA175,176, which is exposed in the nucleosome-free region (FIG. 5c). The coupling of histone chaperone activity and chromatin remodelling activity is also observed for the DAXX–ATRX complex. However, the exchange of nucleosomal H3–H4 for soluble H3.3–H4 would require extensive remodelling of the nucleosome to access the central H3–H4 tetramer. It will therefore be important to investigate whether H3.3–H4 is deposited by DAXX–ATRX through an exchange reaction or a gap-filling mechanism.
Deposition of H3–H4 to form the tetrasome could foreseeably proceed in two ways: the deposition of a pre-assembled H3–H4 tetramer on DNA or by the sequential deposition of individual dimers on DNA. As H3–H4 is mainly dimeric at physiological-like salt concentrations72, stabilization of the H3–H4 tetramer is a necessary requirement for H3–H4 to be deposited as a tetramer. Several chaperones have the ability to bind H3–H4 tetramers, including CAF1 (REFS 177,178), MCM2 (REFS 34,35), SPT2 (REF. 33), Rtt106 (REFS 43,44), Vps75 (REFS 51,53), Nap1 (REF. 53) and FACT37. By contrast, Scm3 mediates assembly of a Cse4–H4 tetrasome through a mechanism involving the consecutive deposition of Cse4–H4 dimers179. HJURP also handles CENP-A–H4 dimers, but through dimerization has the capacity to chaperone two CENP-A–H4 dimers, which may facilitate CENP-A–H4 tetramer formation during deposition180.
Once histone chaperones are recruited to their site of action, their histone variant specificity directs the genomic localization of these histone subtypes8,18. Mutation and domain-swapping experiments have helped to identify the key features of histone variants that specify their assembly pathway and genomic localization167,181,182. Although some chaperones have a natural preference for one histone variant, an imbalance in the chaperone network may skew the system. For example, in the absence of DAXX, CAF1 is able to bind H3.3–H4 (REF. 108), and upon CENP-A overexpression, DAXX can bind CENP-A–H4 and mediates its incorporation at ectopic sites through an exchange reaction generating heterotypic H3.3- and CENP-A-containing nucleosomes12.
Histone recycling
The recycling of modified histones that are evicted during replication, transcription and repair could be key to maintaining the epigenetic state of the locus, and histone chaperones work in an integrated manner with DNA processing machineries to ensure this process. Here, we focus on histone recycling during DNA replication and transcription, as these processes are the best characterized.
The kinetics of histone turnover highlight enhancers and promoters as major sites of histone replacement, whereas gene bodies have slower rates of histone turnover (reviewed in REF. 2). Asf1 can aid histone eviction at promoters and in coding regions183–185. In moderately transcribed genes, a single H2A–H2B dimer may be displaced per nucleosome, whereas highly transcribed genes are characterized by more pronounced nucleosome disruption (reviewed in REF. 2). Displacement of a H2A–H2B dimer can be facilitated by FACT36, and the resultant hexasome can be maintained by FACT36 or Nap1 (REFS 52,186). The loss of either chaperone leads to the depletion of histones in transcribed regions187–189. The re-establishment of chromatin structure after the passage of RNA polymerase II prevents the initiation of transcription from cryptic promoters190,191, and a substantial number of histone chaperones, including SPT2, SPT6, Rtt106, FACT, Vps75, Asf1 and HIRA2,185,192, are implicated in this process (reviewed in REF. 2) (FIG. 5c). SPT6 is a H3–H4 chaperone193 that is coupled to the transcription machinery through interaction with the phosphorylated form of the RNA polymerase II C-terminal domain (reviewed in REF. 2). SPT2 chaperones H3–H4 tetramers33,194 and has overlapping functions with SPT6 (REF. 195), which suggests that H3 and H4 can be recycled as tetramers during transcription. In addition, Nap1-like chaperones (such as Vps75) show genetic interactions with transcription elongation factors192,196 and could also contribute to the recycling of H3–H4 tetramers51,53,56. The FACT complex also has the capacity to handle H3–H4 tetramers37 and is an additional candidate for mediating H3–H4 recycling during transcription. Understanding the division of labour among these many chaperones, whether they act together or in distinct settings, is an important challenge for future research.
The relative balance between histone turnover (eviction and incorporation of newly synthesized histones) and histone recycling is also influenced by histone modifications. H3K56ac can promote histone turnover197, whereas co-transcriptional methylation of H3K36 reduces histone exchange and favours the retention of old histones198. Furthermore, transcription-coupled mono-ubiquitylation of H2BK120 (K123 in yeast) cooperates with FACT in promoting transcription and ensuring the re- assembly of nucleosomes in the wake of the polymerase199,200. Interestingly, H2B monoubiquitylation is also present at replication origins in yeast, where it may aid progression of the replication fork and chromatin assembly201.
During DNA replication, nucleosomes are disrupted ahead of the replicative helicase and H2A–H2B dimers and H3–H4 tetramers are recycled to the daughter strands (reviewed in REFS 3,202) (FIG. 6). The force of the progressing replicative helicase, composed of CDC45, MCM2–7 and GINS (the CMG complex), may be sufficient to trigger nucleosome disruption. Meanwhile, the FACT complex associates with the CMG helicase and is thus in a good position to aid nucleosome disruption38,203–205 (FIG. 6). In yeast, approximately one FACT complex is present for every five nucleosomes, which could facilitate efficient nucleosome disassembly and recycling206. A quantitative analysis of new and old histones in nascent chromatin has shown that histone recycling is highly efficient, resulting in a 1:1 ratio of new and old H3–H4 and H2A–H2B histones on replicated DNA132. The histone variants H3.3 and H2A.X are recycled with similar efficiency as their canonical counterparts132, whereas H2A.Z is lost at least partially upon replication fork passage132,207. The segregation of new and old histones on nascent DNA has been intensely studied5,202, and current data support the segregation of new and old H3.1–H4 dimers into separate nucleosomes post-replication208. This finding suggests that H3.1–H4 tetramers are preserved during recycling. By contrast, the transfer of the two nucleosomal H2A–H2B dimers seems to be uncoupled, with nucleosomes in post-replicative chromatin containing a mixture of new and old H2A–H2B dimers associated with new or old H3.1–H4 tetramers208.
Figure 6. Parental histone recycling during DNA replication.
Evicted parental histones are randomly segregated to daughter DNA strands. The FACT complex may contribute to nucleosome disruption and histone recycling through interaction with the CMG complex (which includes CDC45–MCM2–7–GINS), polymearase α(polα) and RPA. MCM2, which is part of the CMG helicase complex, provides a binding platform for evicted H3–H4 tetramers and may facilitate their recycling directly or in collaboration with FACT or as dimers with ASF1, which splits H3–H4 tetramers and forms a co-chaperone complex with MCM2. It remains unknown whether other components of the DNA replication machinery have histone chaperone activity and whether deposition of old and new histones occurs by separate pathways. For definitions of histone chaperone abbbreviations, see TABLE 1.
Efficient re-incorporation of old histones behind the replication fork probably requires integration of histone recycling with DNA replication, as best illustrated by the chaperone activity of MCM2 (REFS 34,35,89), which is part of the CMG helicase. The flexible N-terminal tail of MCM2 contains a highly conserved HBD34,35,38,209, which can chaperone210 histones in the context of soluble and chromatin-bound MCM2, including inactive MCM2–7 double hexamers and active CMG helicases34,38,87,89 (FIG. 6). MCM2 chaperones a H3–H4 tetramer in vitro34,35 and in vivo34 by mimicking nucleosomal DNA. This mode of interaction renders MCM2 able to chaperone all H3 variants, including CENP-A34, which allows the CMG complex to handle old evicted H3–H4 tetramers genome-wide. The CMG complex could thus function as a platform for retaining evicted histones at the replication fork34; however, it remains less clear how these histones are transferred to newly synthesized DNA. The MCM2 tail may directly facilitate the re-deposition of histones; yet, this process might create a bias towards segregation onto the lagging DNA strand, as structural studies suggest that the N-terminal region of MCM2 is located close to polymerase-α211. Thus, other histone-binding platforms may exist within the DNA replication machinery or soluble histone chaperones may mediate histone transfer. ASF1 interacts with MCM2 within the CMG complex at both active and stalled replication forks in a histone-dependent manner34,87,89, which may facilitate histone transfer in a reaction involving splitting of the MCM2-bound H3–H4 tetramer34,89 (FIG. 6). However, given that old and new H3.1–H4 dimers generally do not mix during DNA replication208, old H3.1–H4 dimers would have to be maintained as tetramers or channelled as dimers into a deposition pathway specific to old histones. Alternatively, the role of ASF1 could be to ensure that evicted histones are chaperoned when the replication fork stalls and recycling is interrupted87,208. The FACT complex is recruited to active replication forks38,40,203–205,212 through an apparent multitude of interactions that include polymerase-α212, replication protein A (RPA)40, MCM4 (REF. 205) as well as histone-dependent inter actions with MCM2 (REF. 38) (FIG. 6). Alignment of MCM2–H3–H4 (REFS 34,35) and SPT16-M–H3–H4 (REF. 37) structures support that FACT and MCM2 may collaborate in a co-chaperone complex37,38 during parental histone transfer. It is possible that one old H2A–H2B dimer is evicted and the resultant hexasome can be recycled. In this instance, if MCM2 and FACT collaborate during histone transfer, additional H2A–H2B binding surfaces such as those present in the FACT complex36,41,47 would need to be used. Old histones maintain their PTMs during recycling132 and these PTMs may, via recruitment of their cognate enzymes, function as a blueprint for the modification of neighbouring new histones (reviewed in REF. 3) to maintain epigenetic states. Thus, understanding histone recycling and whether the process can be challenged by developmental cues and/or cellular stresses therefore represents one key to understand epigenetic cellular memory.
Concluding remarks
The chromatin landscape must remain flexible to enable the regulation of gene expression and programmed changes in cell identity to occur while also protecting DNA from deleterious events. The functions of histone chaperones are crucial for allowing dynamic accessibility to genomic loci, which underscores the importance of unveiling the modes of action and biological functions of this large and diverse group of proteins. As discussed in this Review, key questions regarding nucleosome dynamics and histone chaperone function remain unresolved. For example, what is the biological significance of co-chaperone relationships? Is it important to shield the entire histone fold dimer? How do histone chaperones mediate nucleosome assembly and disassembly in vivo? How is chaperone function integrated with chromatin remodelling and the larger machineries that are involved in DNA transcription, replication and repair? Further structural analyses of histone chaperones, in the context of larger protein assemblies and of the deposition and disassembly machineries, by crystallography and cryoelectron microscopy should provide an exciting new entry point to answer these long-standing questions. Furthermore, genome editing holds great promise for translating basic molecular understanding into a broader biological context, which should reveal the importance of the individual chaperones in histone logistics.
Key points.
Chromatin integrity and functionality is governed by the controlled assembly and disassembly of nucleosomes.
An elaborate histone chaperone network governs histone provision, chromatin assembly, histone recycling and histone turnover.
Histone chaperone networks operate through histone-dependent co-chaperone interactions and direct chaperone–chaperone contacts.
The mode of action of histone chaperones is interpreted from structural and biochemical studies of histone–chaperone complexes.
Key molecular functions of histone chaperones include the shielding of functional histone interfaces and trapping histones in non-nucleosomal conformations.
The integration of histone chaperone function across DNA metabolic processes acts to maintain genome and epigenome integrity.
Acknowledgments
The authors apologize for the many studies in the histone chaperone field that were unable to be cited owing to space restrictions. The authors thank C. Alabert, A. Bowman and Z. Jasencakova for useful comments on the manuscript and for help with figure design. D.J.P. is supported by funds from the Leukemia and Lymphoma Society and STARR Foundation and by the Memorial Sloan-Kettering Cancer Core Grant (P30 CA008748). A.G. is an EMBO Young Investigator and her research is supported by the European Research Council (ERC StG, no. 281765), the Danish National Research Foundation to the Center for Epigenetics (DNRF82), the Danish Cancer Society, the Danish Medical Research Council, the Novo Nordisk Foundation and the Lundbeck Foundation.
Glossary
- Histone chaperones
Defined here as proteins that handle non-nucleosomal histones in vivo and mediate the assembly of nucleosomes from isolated histones and DNA in vitro.
- Histone storage
The sequestration of histones in the soluble fraction of the cell that prevents their degradation.
- Histone turnover
The eviction of nucleosomal histones, followed by deposition of new histones at the same genomic loci.
- Histone chaperone network
The integration of histone chaperone functions to support histone dynamics across various cellular processes.
- Canonical histones
Core histone subtypes (H3.1, H3.2, H4, H2A and H2B) that are expressed in S phase of the cell cycle and mainly incorporated into nucleosomes in a DNA replication-dependent manner.
- Replacement variants
Histone subtypes (such as H3.3, CENP-A, and H2A.Z) incorporated into nucleosomes via DNA replication-independent pathways and for which expression is not restricted to S phase.
- De novo histone deposition
Incorporation of newly synthesized histones into chromatin.
- Co-chaperone
Here defined as a complex containing two or more histone chaperones brought together in a histone-dependent manner.
- Epigenetic plasticity
Heritable information other than DNA sequence that maintains cellular traits while also being subject to change without said changes being permanent.
- H3K56ac
A mark of newly synthesized H3–H4 in yeast, catalysed by Rtt109 in an Asf1-dependent manner, that promotes replication-dependent histone deposition.
- Histone recycling
Re-deposition of histones evicted from chromatin by cellular processes that require access to the DNA template
- Dyad DNA
The dyad position locates the pseudo axis of symmetry, which coincides with the central base pair (or pairs) of nucleosomal DNA and the H3–H4 tetramerization interface, around which the nucleosome can be rotated 180° and map back onto itself
- Tetrasome
Thought to be the first assembly intermediate during nucleosome assembly, the tetrasome is the product of the deposition of a H3–H4 tetramer on DNA
- RBAP46 and RBAP48
Histone chaperone homologues that are almost identical and seem to be interchangeable in most of their chromatin-modifying complexes, apart from HAT1 (RBAP46) and CAF1 (RBAP48).
- Histone reader
A protein that binds to histones in a post-translational modification-dependent manner.
- H4K20me0
Histone H4 unmethylated at lysine 20 (H4K20me0); a signature of newly synthesized histones that marks post-replicative chromatin until G2/M phase of the cell cycle, when H4K20 methylation is established on those new histones.
- Soluble histones
Non-nucleosomal histones.
- H4K5acK12ac
Highly conserved diacetylation mark, catalysed by RBAP46–HAT1, that marks newly synthesized histone H4 before deposition.
- Histone exchange
The replacement of nucleosomal histones with the corresponding canonical histones (H2A–H2B, H3–H4) or replacement variants (H2A.Z–H2B, H3.3–H4).
- Hexasome
A nucleosome intermediate generated by either the loss of one H2A–H2B dimer from the nucleosome or the addition of one H2A–H2B dimer to the H3–H4 tetrasome.
Biographies
Colin M. Hammond is a postdoctoral fellow in the laboratory of Anja Groth at the Biotech Research and Innovation Centre (BRIC), University of Copenhagen, Denmark. His research focuses on histone chaperone function during DNA replication. He gained his Ph.D. investigating mechanistic aspects of histone chaperone biology using structural and biochemical approaches in the Tom Owen-Hughes laboratory at the College of Life Sciences, University of Dundee, UK.
Caroline B. Strømme carried out her Ph.D. in the laboratory of Anja Groth at the Biotech Research and Innovation Centre (BRIC), University of Copenhagen, Denmark. She investigated histone chaperone functions within the replicative helicase to understand mechanisms of histone dynamics at replication forks.
Hongda Huang received his Ph.D. in biology from the University of Science and Technology of China and is now a research associate in Dinshaw J. Patel’s laboratory at the Memorial Sloan-Kettering Cancer Center in New York, USA. His research focuses on the biochemical and structural mechanisms of disease-related protein complexes, in particular the histone chaperones.
Dinshaw J. Patel is Member and Abby Rockefeller Mauze Chair in experimental therapeutics at the Memorial Sloan-Kettering Cancer Center in New York, USA. His research interests are in RNA-mediated gene regulation including CRISPR–Cas systems, histone and DNA methylation-mediated epigenetic regulation including histone chaperones, and metazoan cytoplasmic nucleic acid sensors that trigger the interferon response.
Anja Groth is a professor and principal investigator at the Biotech Research and Innovation Centre (BRIC), University of Copenhagen, Denmark. Her research aims to understand how cells ensure faithful transmission of both genetic and epigenetic information during cell division, with a special focus on chromatin replication, histone chaperones, and epigenome and genome integrity.
Footnotes
Competing interests statement
The authors declare competing financial interests: C.M.H., H.H., D.J.P. and A.G. are named inventors on a patent application covering the discoveries in Saredi et al. H4K20me0 marks post-replicative chromatin and recruits the TONSL–MMS22L DNA repair complex. Nature 534, 714–718 (2016).
DATABASES
RSCB Protein Data Bank: http://www.rcsb.org/pdb/home/home.do
SGD YeastMine: http://yeastmine.yeastgenome.org
ALL LINKS ARE ACTIVE IN THE ONLINE PDF
Subject categories
Biological sciences / Molecular biology / Chromatin / Nucleosomes [URI /631/337/100/1701]
Biological sciences / Molecular biology / Chromatin / Histone variants [URI /631/337/100/2286]
Biological sciences / Molecular biology / DNA replication [URI /631/337/151]
Biological sciences / Molecular biology / Protein folding / Chaperones [URI /631/337/470/1981]
References
- 1.Laskey R, Honda B, Mills A, Finch J. Nucleosomes are assembled by an acidic protein which binds histones and transfers them to DNA. Nature. 1978;275:416–420. doi: 10.1038/275416a0. The first description of histone chaperone function elucidated by classical biochemical approaches probing the physicochemical nature of the association between nucleoplasmin and histones. [DOI] [PubMed] [Google Scholar]
- 2.Venkatesh S, Workman JL. Histone exchange, chromatin structure and the regulation of transcription. Nat Rev Mol Cell Biol. 2015;16:178–189. doi: 10.1038/nrm3941. [DOI] [PubMed] [Google Scholar]
- 3.Alabert C, Groth A. Chromatin replication and epigenome maintenance. Nat Rev Mol Cell Biol. 2012;13:153–167. doi: 10.1038/nrm3288. [DOI] [PubMed] [Google Scholar]
- 4.Adam S, Dabin J, Polo SE. Chromatin plasticity in response to DNA damage: the shape of things to come. DNA Repair. 2015;32:120–126. doi: 10.1016/j.dnarep.2015.04.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Annunziato AT. Assembling chromatin: the long and winding road. Biochim Biophys Acta. 2013;1819:196–210. doi: 10.1016/j.bbagrm.2011.07.005. [DOI] [PubMed] [Google Scholar]
- 6.Gurard-Levin ZA, Quivy JP, Almouzni G. Histone chaperones: assisting histone traffic and nucleosome dynamics. Annu Rev Biochem. 2014;83:487–517. doi: 10.1146/annurev-biochem-060713-035536. [DOI] [PubMed] [Google Scholar]
- 7.Marzluff WF, Wagner EJ, Duronio RJ. Metabolism and regulation of canonical histone mRNAs: life without a poly(A) tail. Nat Rev Genet. 2008;9:843–854. doi: 10.1038/nrg2438. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Henikoff S, Smith MM. Histone variants and epigenetics. Cold Spring Harb Perspect Biol. 2015;7:a019364. doi: 10.1101/cshperspect.a019364. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.O’Sullivan RJ, et al. Rapid induction of alternative lengthening of telomeres by depletion of the histone chaperone ASF1. Nat Struct Mol Biol. 2014;21:167–174. doi: 10.1038/nsmb.2754. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Muller S, Almouzni G. A network of players in H3 histone variant deposition and maintenance at centromeres. Biochim Biophys Acta. 2014;1839:241–250. doi: 10.1016/j.bbagrm.2013.11.008. [DOI] [PubMed] [Google Scholar]
- 11.Heun P, et al. Mislocalization of the Drosophila centromere-specific histone CID promotes formation of functional ectopic kinetochores. Dev Cell. 2006;10:303–315. doi: 10.1016/j.devcel.2006.01.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Lacoste N, et al. Mislocalization of the centromeric histone variant CenH3/CENP-A in human cells depends on the chaperone DAXX. Mol Cell. 2014;53:631–644. doi: 10.1016/j.molcel.2014.01.018. [DOI] [PubMed] [Google Scholar]
- 13.Ishiuchi T, et al. Early embryonic-like cells are induced by downregulating replication-dependent chromatin assembly. Nat Struct Mol Biol. 2015;22:662–671. doi: 10.1038/nsmb.3066. [DOI] [PubMed] [Google Scholar]
- 14.Cheloufi S, et al. The histone chaperone CAF-1 safeguards somatic cell identity. Nature. 2015;528:218–224. doi: 10.1038/nature15749. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Mejlvang J, et al. New histone supply regulates replication fork speed and PCNA unloading. J Cell Biol. 2014;204:29–43. doi: 10.1083/jcb.201305017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Meeks-Wagner D, Hartwell LH. Normal stoichiometry of histone dimer sets is necessary for high fidelity of mitotic chromosome transmission. Cell. 1986;44:43–52. doi: 10.1016/0092-8674(86)90483-6. [DOI] [PubMed] [Google Scholar]
- 17.Gunjan A, Verreault AA. Rad53 kinase-dependent surveillance mechanism that regulates histone protein levels in S. cerevisiae. Cell. 2003;115:537–549. doi: 10.1016/s0092-8674(03)00896-1. Shows that the checkpoint kinase Rad53 monitors the level of soluble histones and that accumulation of excess histones caused by mutation of Rad53 jeopardizes genome stability. [DOI] [PubMed] [Google Scholar]
- 18.Mattiroli F, D’Arcy S, Luger K. The right place at the right time: chaperoning core histone variants. EMBO Rep. 2015;16:1454–1466. doi: 10.15252/embr.201540840. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Andrews AJ, Chen X, Zevin A, Stargell LA, Luger K. The histone chaperone Nap1 promotes nucleosome assembly by eliminating nonnucleosomal histone DNA interactions. Mol Cell. 2010;37:834–842. doi: 10.1016/j.molcel.2010.01.037. Thermodynamic study detailing the role of Nap1 in nucleosome assembly, showing how histone chaperones can buffer interactions between histones and DNA. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Andrews AJ, Downing G, Brown K, Park YJ, Luger KA. Thermodynamic model for Nap1–histone Interactions. J Biol Chem. 2008;283:32412–32418. doi: 10.1074/jbc.M805918200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Daganzo SM, et al. Structure and function of the conserved core of histone deposition protein Asf1. Curr Biol. 2003;13:2148–2158. doi: 10.1016/j.cub.2003.11.027. [DOI] [PubMed] [Google Scholar]
- 22.English CM, Adkins MW, Carson JJ, Churchill ME, Tyler JK. Structural basis for the histone chaperone activity of Asf1. Cell. 2006;127:495–508. doi: 10.1016/j.cell.2006.08.047. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Natsume R, et al. Structure and function of the histone chaperone CIA/ASF1 complexed with histones H3 and H4. Nature. 2007;446:338–341. doi: 10.1038/nature05613. References 22 and 23 provide the first co-crystal structure of a histone chaperone in complex with its histone cargo, providing a molecular description of how Asf1 chaperones an H3–H4 dimer. [DOI] [PubMed] [Google Scholar]
- 24.Wang AY, et al. Asf1-like structure of the conserved Yaf9 YEATS domain and role in H2A.Z deposition and acetylation. Proc Natl Acad Sci USA. 2009;106:21573–21578. doi: 10.1073/pnas.0906539106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Shanle EK, et al. Association of Taf14 with acetylated histone H3 directs gene transcription and the DNA damage response. Genes Dev. 2015;29:1795–1800. doi: 10.1101/gad.269977.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Li Y, et al. AF9 YEATS domain links histone acetylation to DOT1L-mediated H3K79 methylation. Cell. 2014;159:558–571. doi: 10.1016/j.cell.2014.09.049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Liu H, et al. Structural insights into yeast histone chaperone Hif1: a scaffold protein recruiting protein complexes to core histones. Biochem J. 2014;462:465–473. doi: 10.1042/BJ20131640. [DOI] [PubMed] [Google Scholar]
- 28.Bowman A, et al. The histone chaperone sNASP binds a conserved peptide motif within the globular core of histone H3 through its TPR repeats. Nucleic Acids Res. 2015;44:3105–3117. doi: 10.1093/nar/gkv1372. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Hu H, et al. Structure of a CENP-A-histone H4 heterodimer in complex with chaperone HJURP. Genes Dev. 2011;25:901–906. doi: 10.1101/gad.2045111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Cho USS, Harrison SC. Recognition of the centromere-specific histone Cse4 by the chaperone Scm3. Proc Natl Acad Sci USA. 2011;108:9367–9371. doi: 10.1073/pnas.1106389108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Elsässer S, et al. DAXX envelops a histone H3.3–H4 dimer for H3.3-specific recognition. Nature. 2012;491:560–565. doi: 10.1038/nature11608. Defines the structural specificity of DAXX for H3.3–H4 and describes the sequestration of the H3 αN helix and H4 C terminus in non-nucleosomal conformations. This study uses a novel approach to reconstitute the histone chaperone complex from unfolded components. See also REF. 32. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Liu CP, et al. Structure of the variant histone H3.3–H4 heterodimer in complex with its chaperone DAXX. Nat Struct Mol Biol. 2012;19:1287–1292. doi: 10.1038/nsmb.2439. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Chen S, et al. Structure–function studies of histone H3/H4 tetramer maintenance during transcription by chaperone Spt2. Genes Dev. 2015;29:1326–1340. doi: 10.1101/gad.261115.115. Presents the structure of human SPT2 bound to a H3–H4 tetramer and shows that histone-binding mutants of Spt2 in yeast phenocopy Spt2 deletion in failing to suppress cryptic transcription. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Huang H, et al. A unique binding mode enables MCM2 to chaperone histones H3–H4 at replication forks. Nat Struct Mol Biol. 2015;22:618–626. doi: 10.1038/nsmb.3055. Provides the crystal structure of MCM2 with a H3–H4 tetramer and shows that MCM2 in vivo chaperones H3–H4 alone and as part of the replicative helicase. Also provides the first structure of a co-chaperone complex, the MCM2–H3–H4–ASF1 complex. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Richet N, et al. Structural insight into how the human helicase subunit MCM2 may act as a histone chaperone together with ASF1 at the replication fork. Nucleic Acids Res. 2015;43:1905–1917. doi: 10.1093/nar/gkv021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Belotserkovskaya R, et al. FACT facilitates transcription-dependent nucleosome alteration. Science. 2003;301:1090–1093. doi: 10.1126/science.1085703. Describes the histone chaperone function of the FACT complex and its ability to promote H2A–H2B eviction and transcription through chromatin. [DOI] [PubMed] [Google Scholar]
- 37.Tsunaka Y, Fujiwara Y, Oyama T, Hirose S, Morikawa K. Integrated molecular mechanism directing nucleosome reorganization by human FACT. Genes Dev. 2016;30:673–686. doi: 10.1101/gad.274183.115. Elucidates structurally how SPT16 can bind a H3–H4 tetramer and provides insights into the mechanism of nucleosome disruption by the FACT complex. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Foltman M, et al. Eukaryotic replisome components cooperate to process histones during chromosome replication. Cell Rep. 2013;3:892–904. doi: 10.1016/j.celrep.2013.02.028. [DOI] [PubMed] [Google Scholar]
- 39.Liu Y, et al. Structural analysis of Rtt106p reveals a DNA binding role required for heterochromatin silencing. J Biol Chem. 2010;285:4251–4262. doi: 10.1074/jbc.M109.055996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.VanDemark A, et al. The structure of the yFACT Pob3-M domain, its interaction with the DNA replication factor RPA, and a potential role in nucleosome deposition. Mol Cell. 2006;22:363–374. doi: 10.1016/j.molcel.2006.03.025. [DOI] [PubMed] [Google Scholar]
- 41.Hondele M, et al. Structural basis of histone H2A–H2B recognition by the essential chaperone FACT. Nature. 2013;499:111–114. doi: 10.1038/nature12242. [DOI] [PubMed] [Google Scholar]
- 42.Li Q, et al. Acetylation of histone H3 lysine 56 regulates replication-coupled nucleosome assembly. Cell. 2008;134:244–255. doi: 10.1016/j.cell.2008.06.018. Shows that H3K56 acetylation promotes binding of histones to Rtt106 and the CAF1 complex and determines that the tandem pleckstrin homology domain of Rtt106 has specificity for H3K56ac. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Su D, et al. Structural basis for recognition of H3K56-acetylated histone H3–H4 by the chaperone Rtt106. Nature. 2012;483:104–107. doi: 10.1038/nature10861. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Fazly A, et al. Histone chaperone Rtt106 promotes nucleosome formation using (H3–H4)2 tetramers. J Biol Chem. 2012;287:10753–10760. doi: 10.1074/jbc.M112.347450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Stuwe T, et al. The FACT Spt16 “peptidase” domain is a histone H3–H4 binding module. Proc Natl Acad Sci USA. 2008;105:8884–8889. doi: 10.1073/pnas.0712293105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Kemble DJ, McCullough LL, Whitby FG, Formosa T, Hill CP. FACT disrupts nucleosome structure by binding H2A–H2B with conserved peptide motifs. Mol Cell. 2015;60:294–306. doi: 10.1016/j.molcel.2015.09.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Formosa T. The role of FACT in making and breaking nucleosomes. Biochim Biophys Acta. 2012;1819:247–255. doi: 10.1016/j.bbagrm.2011.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Selth L, Svejstrup J. Vps75, a new yeast member of the NAP histone chaperone family. J Biol Chem. 2007;282:12358–12362. doi: 10.1074/jbc.C700012200. [DOI] [PubMed] [Google Scholar]
- 49.Park YJ, Luger K. The structure of nucleosome assembly protein 1. Proc Natl Acad Sci USA. 2006;103:1248–1253. doi: 10.1073/pnas.0508002103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Muto S, et al. Relationship between the structure of SET/TAF-Iβ/INHAT and its histone chaperone activity. Proc Natl Acad Sci USA. 2007;104:4285–4290. doi: 10.1073/pnas.0603762104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Hammond CM, et al. The histone chaperone Vps75 forms multiple oligomeric assemblies capable of mediating exchange between histone H3–H4 tetramers and Asf1–H3–H4 complexes. Nucleic Acids Res. 2016;44:6157–6172. doi: 10.1093/nar/gkw209. Describes how Vps75 forms a co-chaperone complex with Asf1 and histones, and provides the first description of a self-chaperoning mechanism for histone chaperones by demonstrating that the histone-binding surface is shielded in the inactive Vps75 tetramer. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Aguilar-Gurrieri C, et al. Structural evidence for Nap1-dependent H2A–H2B deposition and nucleosome assembly. EMBO J. 2016;35:1465–1482. doi: 10.15252/embj.201694105. Provides the crystal structure of Nap1 with H2A–H2B, showing how Nap1 can buffer H2A–H2B interactions with DNA, and (together with reference 51) sets the precedence for Nap1-like proteins binding directly to a histone dimer. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Bowman A, et al. The histone chaperones Nap1 and Vps75 bind histones H3 and H4 in a tetrameric conformation. Mol Cell. 2011;41:398–408. doi: 10.1016/j.molcel.2011.01.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.D’Arcy S, Luger K. Understanding histone acetyltransferase Rtt109 structure and function: how many chaperones does it take? Curr Opin Struct Biol. 2011;21:728–734. doi: 10.1016/j.sbi.2011.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Zlatanova J, Seebart C, Tomschik M. Nap1: taking a closer look at a juggler protein of extraordinary skills. FASEB J. 2007;21:1294–1310. doi: 10.1096/fj.06-7199rev. [DOI] [PubMed] [Google Scholar]
- 56.Bowman A, et al. The histone chaperones Vps75 and Nap1 form ring-like, tetrameric structures in solution. Nucleic Acids Res. 2014;42:6038–6051. doi: 10.1093/nar/gku232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Newman ER, et al. Large multimeric assemblies of nucleosome assembly protein and histones revealed by small-angle X-ray scattering and electron microscopy. J Biol Chem. 2012;287:26657–26665. doi: 10.1074/jbc.M112.340422. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Tang Y, et al. Structure of the Rtt109–AcCoA/Vps75 complex and implications for chaperone-mediated histone acetylation. Structure. 2011;19:221–231. doi: 10.1016/j.str.2010.12.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Kolonko EM, et al. Catalytic activation of histone acetyltransferase Rtt109 by a histone chaperone. Proc Natl Acad Sci USA. 2010;107:20275–20280. doi: 10.1073/pnas.1009860107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Su D, et al. Structure and histone binding properties of the Vps75–Rtt109 chaperone–lysine acetyltransferase complex. J Biol Chem. 2011;286:15625–15629. doi: 10.1074/jbc.C111.220715. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Dutta S, et al. The crystal structure of nucleoplasmin-core: implications for histone binding and nucleosome assembly. Mol Cell. 2001;8:841–853. doi: 10.1016/s1097-2765(01)00354-9. [DOI] [PubMed] [Google Scholar]
- 62.Namboodiri VM, Akey IVV, Schmidt-Zachmann MS, Head JF, Akey CW. The structure and function of Xenopus NO38-core, a histone chaperone in the nucleolus. Structure. 2004;12:2149–2160. doi: 10.1016/j.str.2004.09.017. [DOI] [PubMed] [Google Scholar]
- 63.Fernández-Rivero N, et al. A quantitative characterization of nucleoplasmin/histone complexes reveals chaperone versatility. Sci Rep. 2016;6:32114. doi: 10.1038/srep32114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Ramos I, et al. Nucleoplasmin binds histone H2A–H2B dimers through its distal face. J Biol Chem. 2010;285:33771–33778. doi: 10.1074/jbc.M110.150664. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Luger K, Mäder A, Richmond R, Sargent D, Richmond T. Crystal structure of the nucleosome core particle at 2.8 Å resolution. Nature. 1997;389:251–260. doi: 10.1038/38444. This seminal work provides a detailed molecular description of the nucleosome that has inspired many young scientists to enter the chromatin field. [DOI] [PubMed] [Google Scholar]
- 66.Arents G, Burlingame RW, Wang BC, Love WE, Moudrianakis EN. The nucleosomal core histone octamer at 3.1 Å resolution: a tripartite protein assembly and a left-handed superhelix. Proc Natl Acad Sci USA. 1991;88:10148–10152. doi: 10.1073/pnas.88.22.10148. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Hong J, et al. The catalytic subunit of the SWR1 remodeler is a histone chaperone for the H2A.Z–H2B dimer. Mol Cell. 2014;53:498–505. doi: 10.1016/j.molcel.2014.01.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Mao Z, et al. Anp32e, a higher eukaryotic histone chaperone directs preferential recognition for H2A.Z. Cell Res. 2014;24:389–399. doi: 10.1038/cr.2014.30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Obri A, et al. ANP32E is a histone chaperone that removes H2A.Z from chromatin. Nature. 2014;505:648–653. doi: 10.1038/nature12922. References 68 and 69 identify the first H2AZ-specific histone chaperone in mammalian cells, ANP32E, providing structural details of the interaction and describing a specific role in H2AZ–H2B eviction. [DOI] [PubMed] [Google Scholar]
- 70.Liang X, et al. Structural basis of H2A.Z recognition by SRCAP chromatin-remodeling subunit YL1. Nat Struct Mol Biol. 2016;23:317–323. doi: 10.1038/nsmb.3190. [DOI] [PubMed] [Google Scholar]
- 71.Latrick CM, et al. Molecular basis and specificity of H2A.Z–H2B recognition and deposition by the histone chaperone YL1. Nat Struct Mol Biol. 2016;23:309–316. doi: 10.1038/nsmb.3189. [DOI] [PubMed] [Google Scholar]
- 72.Donham D, Scorgie J, Churchill M. The activity of the histone chaperone yeast Asf1 in the assembly and disassembly of histone H3/H4–DNA complexes. Nucleic Acids Res. 2011;39:5449–5458. doi: 10.1093/nar/gkr097. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Chavez MS, et al. The conformational flexibility of the C-terminus of histone H4 promotes histone octamer and nucleosome stability and yeast viability. Epigenetics Chromatin. 2012;5:5. doi: 10.1186/1756-8935-5-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Song JJ, Garlick JD, Kingston RE. Structural basis of histone H4 recognition by p55. Genes Dev. 2008;22:1313–1318. doi: 10.1101/gad.1653308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Murzina NV, et al. Structural basis for the recognition of histone H4 by the histone-chaperone RbAp46. Structure. 2008;16:1077–1085. doi: 10.1016/j.str.2008.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Zhang W, et al. Structural plasticity of histones H3–H4 facilitates their allosteric exchange between RbAp48 and ASF1. Nat Struct Mol Biol. 2013;20:29–35. doi: 10.1038/nsmb.2446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Schmitges FW, et al. Histone methylation by PRC2 is inhibited by active chromatin marks. Mol Cell. 2011;42:330–341. doi: 10.1016/j.molcel.2011.03.025. [DOI] [PubMed] [Google Scholar]
- 78.Li Y, et al. Hat2p recognizes the histone H3 tail to specify the acetylation of the newly synthesized H3/H4 heterodimer by the Hat1p/Hat2p complex. Genes Dev. 2014;28:1217–1227. doi: 10.1101/gad.240531.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.An S, Kim H, Cho US. Mis16 independently recognizes histone H4 and the CENP-ACnp1-specific chaperone Scm3sp. J Mol Biol. 2015;427:3230–3240. doi: 10.1016/j.jmb.2015.08.022. [DOI] [PubMed] [Google Scholar]
- 80.Furuyama T, Dalal Y, Henikoff S. Chaperone-mediated assembly of centromeric chromatin in vitro. Proc Natl Acad Sci USA. 2006;103:6172–6177. doi: 10.1073/pnas.0601686103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Wang H, Wang M, Yang N, Xu RM. Structure of the quaternary complex of histone H3–H4 heterodimer with chaperone ASF1 and the replicative helicase subunit MCM2. Protein Cell. 2015;6:693–697. doi: 10.1007/s13238-015-0190-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Zhou Z, et al. NMR structure of chaperone Chz1 complexed with histones H2A.Z–H2B. Nat Struct Mol Biol. 2008;15:868–869. doi: 10.1038/nsmb.1465. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.DeNizio JE, Elsässer SJ, Black BE. DAXX co-folds with H3.3/H4 using high local stability conferred by the H3.3 variant recognition residues. Nucleic Acids Res. 2014;42:4318–4331. doi: 10.1093/nar/gku090. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Saredi G, et al. H4K20me0 marks post-replicative chromatin and recruits the TONSL–MMS22L DNA repair complex. Nature. 2016;534:714–718. doi: 10.1038/nature18312. Reveals a dual histone chaperone and reader function of TONSL, binding H4 tails that are unmethylated at K20 on soluble and nucleosomal histones. H4K20me0 is specific to newly synthesized histones, allowing recruitment of the TONSL–MMS22L repair complex to post-replicative chromatin. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Campos EI, et al. Analysis of the histone H3. interactome: a suitable chaperone for the right event. Mol Cell. 2015;60:697–709. doi: 10.1016/j.molcel.2015.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Kalashnikova AA, Porter-Goff ME, Muthurajan UM, Luger K, Hansen JC. The role of the nucleosome acidic patch in modulating higher order chromatin structure. J R Soc Interface. 2013;10:20121022. doi: 10.1098/rsif.2012.1022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Jasencakova Z, et al. Replication stress interferes with histone recycling and predeposition marking of new histones. Mol Cell. 2010;37:736–743. doi: 10.1016/j.molcel.2010.01.033. This mass spectrometry analysis of modifications on ASF1-bound histones demonstrates that replication stress induces K9me1 accumulation on H3 and impairs recycling of old H3–H4. [DOI] [PubMed] [Google Scholar]
- 88.Groth A, et al. Human Asf1 regulates the flow of S phase histones during replicational stress. Mol Cell. 2005;17:301–311. doi: 10.1016/j.molcel.2004.12.018. [DOI] [PubMed] [Google Scholar]
- 89.Groth A, et al. Regulation of replication fork progression through histone supply and demand. Science. 2007;318:1928–1931. doi: 10.1126/science.1148992. Identifies the H3–H4-dependent interaction between ASF1 and the MCM2–7 helicase and reveals the requirement of ASF1 for DNA unwinding, coupling histone chaperone function to DNA replication. [DOI] [PubMed] [Google Scholar]
- 90.Campos E, et al. The program for processing newly synthesized histones H3.1 and H4. Nat Struct Mol Biol. 2010;17:1343–1351. doi: 10.1038/nsmb.1911. Characterizes distinct H3.1–H4 chaperone complexes present in the cytosolic fraction and identifies HSC70 and HSP90 as upstream chaperones. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Haigney A, Ricketts MD, Marmorstein R. Dissecting the molecular roles of histone chaperones in histone acetylation by type B histone acetyltransferases (HAT-B) J Biol Chem. 2015;290:30648–30657. doi: 10.1074/jbc.M115.688523. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.D’Andrea LD, Regan L. TPR proteins: the versatile helix. Trends Biochem Sci. 2003;28:655–662. doi: 10.1016/j.tibs.2003.10.007. [DOI] [PubMed] [Google Scholar]
- 93.Daniel Ricketts M, et al. Ubinuclein-1 confers histone H3.3-specific-binding by the HIRA histone chaperone complex. Nat Commun. 2015;6:7711. doi: 10.1038/ncomms8711. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Malay AD, Umehara T, Matsubara-Malay K, Padmanabhan B, Yokoyama S. Crystal structures of fission yeast histone chaperone Asf1 complexed with the Hip1 B-domain or the Cac2 C terminus. J Biol Chem. 2008;283:14022–14031. doi: 10.1074/jbc.M800594200. [DOI] [PubMed] [Google Scholar]
- 95.Tang Y, et al. Structure of a human ASF1a–HIRA complex and insights into specificity of histone chaperone complex assembly. Nat Struct Mol Biol. 2006;13:921–929. doi: 10.1038/nsmb1147. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Ask K, et al. Codanin-1, mutated in the anaemic disease CDAI, regulates Asf1 function in S-phase histone supply. EMBO J. 2012;31:2013–2023. doi: 10.1038/emboj.2012.55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Jiao Y, et al. Surprising complexity of the Asf1 histone chaperone–Rad53 kinase interaction. Proc Natl Acad Sci USA. 2012;109:2866–2871. doi: 10.1073/pnas.1106023109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Kurat CF, et al. Regulation of histone gene transcription in yeast. Cell Mol Life Sci. 2014;71:599–613. doi: 10.1007/s00018-013-1443-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Singh RK, Kabbaj MH, Paik J, Gunjan A. Histone levels are regulated by phosphorylation and ubiquitylation-dependent proteolysis. Nat Cell Biol. 2009;11:925–933. doi: 10.1038/ncb1903. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Nelson DM, et al. Coupling of DNA synthesis and histone synthesis in S phase independent of cyclin/cdk2 activity. Mol Cell Biol. 2002;22:7459–7472. doi: 10.1128/MCB.22.21.7459-7472.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Moshkin YM, et al. Histone chaperone ASF1 cooperates with the Brahma chromatin-remodelling machinery. Genes Dev. 2002;16:2621–2626. doi: 10.1101/gad.231202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Klimovskaia IM, et al. Tousled-like kinases phosphorylate Asf1 to promote histone supply during DNA replication. Nat Commun. 2014;5:3394. doi: 10.1038/ncomms4394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Sillje HH, Nigg EA. Identification of human Asf1 chromatin assembly factors as substrates of Tousled-like kinases. Curr Biol. 2001;11:1068–1073. doi: 10.1016/s0960-9822(01)00298-6. [DOI] [PubMed] [Google Scholar]
- 104.Gerard A, et al. The replication kinase Cdc7-Dbf4 promotes the interaction of the p150 subunit of chromatin assembly factor 1 with proliferating cell nuclear antigen. EMBO Rep. 2006;7:817–823. doi: 10.1038/sj.embor.7400750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Muller S, et al. Phosphorylation and DNA binding of HJURP determine its centromeric recruitment and function in CenH3CENP-A loading. Cell Rep. 2014;8:190–203. doi: 10.1016/j.celrep.2014.06.002. [DOI] [PubMed] [Google Scholar]
- 106.Luk E, et al. Chz1, a nuclear chaperone for histone H2AZ. Mol Cell. 2007;25:357–368. doi: 10.1016/j.molcel.2006.12.015. Identifies the first H2AZ-specific histone chaperone in yeast, Chz1, and defines the H2AZ interaction motif of Chz1. [DOI] [PubMed] [Google Scholar]
- 107.Tagami H, Ray-Gallet D, Almouzni G, Nakatani Y. Histone H3.1 and H3.3 complexes mediate nucleosome assembly pathways dependent or independent of DNA synthesis. Cell. 2004;116:51–61. doi: 10.1016/s0092-8674(03)01064-x. Identifies distinct complexes of H3.1 and H3.3 with CAF1 and HIRA complexes, respectively, and demonstrates that soluble H3–H4 are mainly found as dimers. [DOI] [PubMed] [Google Scholar]
- 108.Drane P, Ouararhni K, Depaux A, Shuaib M, Hamiche A. The death-associated protein DAXX is a novel histone chaperone involved in the replication-independent deposition of H3.3. Genes Dev. 2010;24:1253–1265. doi: 10.1101/gad.566910. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Alvarez F, et al. Sequential establishment of marks on soluble histones H3 and H4. J Biol Chem. 2011;286:17714–17721. doi: 10.1074/jbc.M111.223453. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Parthun MR. Histone acetyltransferase 1: more than just an enzyme? Biochim Biophys Acta. 2012;1819:256–263. doi: 10.1016/j.bbagrm.2011.07.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Cook AJ, Gurard-Levin ZA, Vassias I, Almouzni G. A specific function for the histone chaperone NASP to fine-tune a reservoir of soluble H3–H4 in the histone supply chain. Mol Cell. 2011;44:918–927. doi: 10.1016/j.molcel.2011.11.021. References 90 and 111 identify NASP as a H3–H4-specific chaperone, required for storage of H3–H4 and to prevent its degradation. [DOI] [PubMed] [Google Scholar]
- 112.Poveda A, et al. Hif1 is a component of yeast histone acetyltransferase B, a complex mainly localized in the nucleus. J Biol Chem. 2004;279:16033–16043. doi: 10.1074/jbc.M314228200. [DOI] [PubMed] [Google Scholar]
- 113.Green EM, et al. Replication-independent histone deposition by the HIR complex and Asf1. Curr Biol. 2005;15:2044–2049. doi: 10.1016/j.cub.2005.10.053. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Tyler JK, et al. The RCAF complex mediates chromatin assembly during DNA replication and repair. Nature. 1999;402:555–560. doi: 10.1038/990147. Identifies Asf1 as a H3–H4 chaperone that cooperates with the CAF1 complex in replication-coupled chromatin assembly. [DOI] [PubMed] [Google Scholar]
- 115.Ray-Gallet D, et al. HIRA is critical for a nucleosome assembly pathway independent of DNA synthesis. Mol Cell. 2002;9:1091–1100. doi: 10.1016/s1097-2765(02)00526-9. [DOI] [PubMed] [Google Scholar]
- 116.Mello JA, et al. Human Asf1 and CAF-1 interact and synergize in a repair-coupled nucleosome assembly pathway. EMBO Rep. 2002;3:329–334. doi: 10.1093/embo-reports/kvf068. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Tyler JK, et al. Interaction between the Drosophila CAF-1 and ASF1 chromatin assembly factors. Mol Cell Biol. 2001;21:6574–6584. doi: 10.1128/MCB.21.19.6574-6584.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Benson LJ, et al. Modifications of H3 and H4 during chromatin replication, nucleosome assembly, and histone exchange. J Biol Chem. 2006;281:9287–9296. doi: 10.1074/jbc.M512956200. [DOI] [PubMed] [Google Scholar]
- 119.Abascal F, et al. Subfunctionalization via adaptive evolution influenced by genomic context: the case of histone chaperones ASF1a and ASF1b. Mol Biol Evol. 2013;30:1853–1866. doi: 10.1093/molbev/mst086. [DOI] [PubMed] [Google Scholar]
- 120.Hayashi R, et al. Transcriptional regulation of human chromatin assembly factor ASF1. DNA Cell Biol. 2007;26:91–99. doi: 10.1089/dna.2006.0515. [DOI] [PubMed] [Google Scholar]
- 121.Corpet A, et al. Asf1b, the necessary Asf1 isoform for proliferation, is predictive of outcome in breast cancer. EMBO J. 2011;30:480–493. doi: 10.1038/emboj.2010.335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Quivy JP, Gerard A, Cook AJ, Roche D, Almouzni G. The HP1–p150/CAF-1 interaction is required for pericentric heterochromatin replication and S-phase progression in mouse cells. Nat Struct Mol Biol. 2008;15:972–979. doi: 10.1038/nsmb.1470. [DOI] [PubMed] [Google Scholar]
- 123.Hoek M, Stillman B. Chromatin assembly factor 1 is essential and couples chromatin assembly to DNA replication in vivo. Proc Natl Acad Sci USA. 2003;100:12183–12188. doi: 10.1073/pnas.1635158100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Nabatiyan A, Szuts D, Krude T. Induction of CAF-1 expression in response to DNA strand breaks in quiescent human cells. Mol Cell Biol. 2006;26:1839–1849. doi: 10.1128/MCB.26.5.1839-1849.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Piwko W, Buser R, Peter M. Rescuing stalled replication forks: MMS22L–TONSL, a novel complex for DNA replication fork repair in human cells. Cell Cycle. 2011;10:1703–1705. doi: 10.4161/cc.10.11.15557. [DOI] [PubMed] [Google Scholar]
- 126.Straube K, Blackwell JS, Jr, Pemberton LF. Nap1 and Chz1 have separate Htz1 nuclear import and assembly functions. Traffic. 2010;11:185–197. doi: 10.1111/j.1600-0854.2009.01010.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Cai Y, et al. The mammalian YL1 protein is a shared subunit of the TRRAP/TIP60 histone acetyltransferase and SRCAP complexes. J Biol Chem. 2005;280:13665–13670. doi: 10.1074/jbc.M500001200. [DOI] [PubMed] [Google Scholar]
- 128.Parthun MR. Hat1: the emerging cellular roles of a type B histone acetyltransferase. Oncogene. 2007;26:5319–5328. doi: 10.1038/sj.onc.1210602. [DOI] [PubMed] [Google Scholar]
- 129.Loyola A, Bonaldi T, Roche D, Imhof A, Almouzni G. PTMs on H3 variants before chromatin assembly potentiate their final epigenetic state. Mol Cell. 2006;24:309–316. doi: 10.1016/j.molcel.2006.08.019. This mass spectrometry analysis of histone modifications on soluble and nucleosomal H3.1–H4 and H3.3–H4 complexes identifies H3K9me1 as a predeposition mark that is important for establishing heterochromatin. [DOI] [PubMed] [Google Scholar]
- 130.Masumoto H, Hawke D, Kobayashi R, Verreault A. A role for cell-cycle-regulated histone H3 lysine 56 acetylation in the DNA damage response. Nature. 2005;436:294–298. doi: 10.1038/nature03714. Identifies H3K56 acetylation as a highly abundant mark on new soluble histones in yeast, defines the cell cycle dynamics of the mark and shows that it is required for genome stability. [DOI] [PubMed] [Google Scholar]
- 131.Chen CC, et al. Acetylated lysine 56 on histone H3 drives chromatin assembly after repair and signals for the completion of repair. Cell. 2008;134:231–243. doi: 10.1016/j.cell.2008.06.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Alabert C, et al. Two distinct modes for propagation of histone PTMs across the cell cycle. Genes Dev. 2015;29:585–590. doi: 10.1101/gad.256354.114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Xu F, Zhang K, Grunstein M. Acetylation in histone H3 globular domain regulates gene expression in yeast. Cell. 2005;121:375–385. doi: 10.1016/j.cell.2005.03.011. [DOI] [PubMed] [Google Scholar]
- 134.Han J, Zhang H, Wang Z, Zhou H, Zhang ZA. Cul4 E3 ubiquitin ligase regulates histone hand-off during nucleosome assembly. Cell. 2013;155:817–829. doi: 10.1016/j.cell.2013.10.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Niikura Y, et al. CENP-A K124 ubiquitylation is required for CENP-A deposition at the centromere. Dev Cell. 2015;32:589–603. doi: 10.1016/j.devcel.2015.01.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Rivera C, et al. Methylation of histone H3 lysine 9 occurs during translation. Nucleic Acids Res. 2015;43:9097–9106. doi: 10.1093/nar/gkv929. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Loyola A, et al. The HP 1α-CAF1-SetDB1-containing complex provides H3K9me1 for Suv39-mediated K9me3 in pericentric heterochromatin. EMBO Rep. 2009;10:769–775. doi: 10.1038/embor.2009.90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138.Shibahara K, Stillman B. Replication-dependent marking of DNA by PCNA facilitates CAF-1-coupled inheritance of chromatin. Cell. 1999;96:575–585. doi: 10.1016/s0092-8674(00)80661-3. Seminal paper showing that the CAF1 complex is recruited to replication forks through interaction with PCNA to mediate replication-coupled chromatin assembly. [DOI] [PubMed] [Google Scholar]
- 139.Smith DJ, Whitehouse I. Intrinsic coupling of lagging-strand synthesis to chromatin assembly. Nature. 2012;483:434–438. doi: 10.1038/nature10895. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Moggs JG, et al. A CAF-1–PCNA-mediated chromatin assembly pathway triggered by sensing DNA damage. Mol Cell Biol. 2000;20:1206–1218. doi: 10.1128/mcb.20.4.1206-1218.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Polo SE, Roche D, Almouzni G. New histone incorporation marks sites of UV repair in human cells. Cell. 2006;127:481–493. doi: 10.1016/j.cell.2006.08.049. [DOI] [PubMed] [Google Scholar]
- 142.Li X, Tyler JK. Nucleosome disassembly during human non-homologous end joining followed by concerted HIRA- and CAF-1-dependent reassembly. eLife. 2016;5:e15129. doi: 10.7554/eLife.15129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143.Brachet E, Beneut C, Serrentino ME, Borde V. The CAF-1 and Hir histone chaperones associate with sites of meiotic double-strand breaks in budding yeast. PLoS ONE. 2015;10:e0125965. doi: 10.1371/journal.pone.0125965. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Huang S, et al. Rtt106p is a histone chaperone involved in heterochromatin-mediated silencing. Proc Natl Acad Sci USA. 2005;102:13410–13415. doi: 10.1073/pnas.0506176102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Yang J, et al. The histone chaperone FACT contributes to DNA replication-coupled nucleosome assembly. Cell Rep. 2016;14:1128–1141. doi: 10.1016/j.celrep.2015.12.096. [DOI] [PubMed] [Google Scholar]
- 146.Huang C, Zhu B. H3.3 turnover: a mechanism to poise chromatin for transcription, or a response to open chromatin? Bioessays. 2014;36:579–584. doi: 10.1002/bies.201400005. [DOI] [PubMed] [Google Scholar]
- 147.Goldberg AD, et al. Distinct factors control histone variant H3.3 localization at specific genomic regions. Cell. 2010;140:678–691. doi: 10.1016/j.cell.2010.01.003. References 108 and 147 identify DAXX–ATRX as a H3.3-specific chaperone complex. DAXX–ATRX affinity tags endogenous H3.3 for genome-wide profiling, showing HIRA-dependent incorporation in genomic regions and ATRX-dependent incorporation in telomeres. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Elsasser SJ, Noh KM, Diaz N, Allis CD, Banaszynski LA. Histone H3.3 is required for endogenous retroviral element silencing in embryonic stem cells. Nature. 2015;522:240–244. doi: 10.1038/nature14345. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.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 USA. 2010;107:14075–14080. doi: 10.1073/pnas.1008850107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 150.Banaszynski LA, et al. Hira-dependent histone H3.3 deposition facilitates PRC2 recruitment at developmental loci in ES cells. Cell. 2013;155:107–120. doi: 10.1016/j.cell.2013.08.061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Ray-Gallet D, et al. Dynamics of histone H3 deposition in vivo reveal a nucleosome gap-filling mechanism for H3.3 to maintain chromatin integrity. Mol Cell. 2011;44:928–941. doi: 10.1016/j.molcel.2011.12.006. [DOI] [PubMed] [Google Scholar]
- 152.Pchelintsev NA, et al. Placing the HIRA histone chaperone complex in the chromatin landscape. Cell Rep. 2013;3:1012–1019. doi: 10.1016/j.celrep.2013.03.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Soni S, Pchelintsev N, Adams PD, Bieker JJ. Transcription factor EKLF (KLF1) recruitment of the histone chaperone HIRA is essential for β-globin gene expression. Proc Natl Acad Sci USA. 2014;111:13337–13342. doi: 10.1073/pnas.1405422111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Adam S, Polo SE, Almouzni G. Transcription recovery after DNA damage requires chromatin priming by the H3.3 histone chaperone HIRA. Cell. 2013;155:94–106. doi: 10.1016/j.cell.2013.08.029. [DOI] [PubMed] [Google Scholar]
- 155.Iwase S, et al. ATRX ADD domain links an atypical histone methylation recognition mechanism to human mental-retardation syndrome. Nat Struct Mol Biol. 2011;18:769–776. doi: 10.1038/nsmb.2062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Dhayalan A, et al. The ATRX-ADD domain binds to H3 tail peptides and reads the combined methylation state of K4 and K9. Hum Mol Genet. 2011;20:2195–2203. doi: 10.1093/hmg/ddr107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Eustermann S, et al. Combinatorial readout of histone H3 modifications specifies localization of ATRX to heterochromatin. Nat Struct Mol Biol. 2011;18:777–782. doi: 10.1038/nsmb.2070. [DOI] [PubMed] [Google Scholar]
- 158.Xiao H, et al. Nonhistone Scm3 binds to AT-rich DNA to organize atypical centromeric nucleosome of budding yeast. Mol Cell. 2011;43:369–380. doi: 10.1016/j.molcel.2011.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Stoler S, et al. Scm3, an essential Saccharomyces cerevisiae centromere protein required for G2/M progression and Cse4 localization. Proc Natl Acad Sci USA. 2007;104:10571–10576. doi: 10.1073/pnas.0703178104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Camahort R, et al. Scm3 is essential to recruit the histone h3 variant cse4 to centromeres and to maintain a functional kinetochore. Mol Cell. 2007;26:853–865. doi: 10.1016/j.molcel.2007.05.013. [DOI] [PubMed] [Google Scholar]
- 161.Shivaraju M, Camahort R, Mattingly M, Gerton JL. Scm3 is a centromeric nucleosome assembly factor. J Biol Chem. 2011;286:12016–12023. doi: 10.1074/jbc.M110.183640. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Mizuguchi G, Xiao H, Wisniewski J, Smith MM, Wu C. Nonhistone Scm3 and histones CenH3–H4 assemble the core of centromere-specific nucleosomes. Cell. 2007;129:1153–1164. doi: 10.1016/j.cell.2007.04.026. [DOI] [PubMed] [Google Scholar]
- 163.Wisniewski J, et al. Imaging the fate of histone Cse4 reveals de novo replacement in S phase and subsequent stable residence at centromeres. eLife. 2014;3:e02203. doi: 10.7554/eLife.02203. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Lando D, et al. Quantitative single-molecule microscopy reveals that CENP-ACnp1 deposition occurs during G2 in fission yeast. Open Biol. 2012;2:120078. doi: 10.1098/rsob.120078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Jansen LE, Black BE, Foltz DR, Cleveland DW. Propagation of centromeric chromatin requires exit from mitosis. J Cell Biol. 2007;176:795–805. doi: 10.1083/jcb.200701066. Introduces SNAP-tag pulse-labelling technology to track old and new CENP-A, showing efficient recycling of old CENP-A in S phase of the cell cycle complemented by de novo deposition of CENP-A in G1 phase. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Dunleavy EM, et al. HJURP is a cell-cycle-dependent maintenance and deposition factor of CENP-A at centromeres. Cell. 2009;137:485–497. doi: 10.1016/j.cell.2009.02.040. References 166 and 167 identify HJURP as a CENP-A-specific chaperone required for deposition of new CENP-A in early G1 phase. [DOI] [PubMed] [Google Scholar]
- 167.Foltz DR, et al. Centromere-specific assembly of CENP-a nucleosomes is mediated by HJURP. Cell. 2009;137:472–484. doi: 10.1016/j.cell.2009.02.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Pidoux AL, et al. Fission yeast Scm3: a CENP-A receptor required for integrity of subkinetochore chromatin. Mol Cell. 2009;33:299–311. doi: 10.1016/j.molcel.2009.01.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Westhorpe FG, Fuller CJ, Straight AF. A cell- free CENP-A assembly system defines the chromatin requirements for centromere maintenance. J Cell Biol. 2015;209:789–801. doi: 10.1083/jcb.201503132. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 170.Foltz DR, et al. The human CENP-A centromeric nucleosome-associated complex. Nat Cell Biol. 2006;8:458–469. doi: 10.1038/ncb1397. [DOI] [PubMed] [Google Scholar]
- 171.Wu WH, et al. Swc2 is a widely conserved H2AZ-binding module essential for ATP-dependent histone exchange. Nat Struct Mol Biol. 2005;12:1064–1071. doi: 10.1038/nsmb1023. [DOI] [PubMed] [Google Scholar]
- 172.Wu WH, et al. N terminus of Swr1 binds to histone H2AZ and provides a platform for subunit assembly in the chromatin remodeling complex. J Biol Chem. 2009;284:6200–6207. doi: 10.1074/jbc.M808830200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Mizuguchi G, et al. ATP-driven exchange of histone H2AZ variant catalyzed by SWR1 chromatin remodeling complex. Science. 2004;303:343–348. doi: 10.1126/science.1090701. Establishes an elegant in vitro assay to show that SWR1 mediates the ATP-dependent exchange of canonical H2A with the H2A. Z variant, paving the way for an in-depth understanding of SWR1 function. [DOI] [PubMed] [Google Scholar]
- 174.Luk E, et al. Stepwise histone replacement by SWR1 requires dual activation with histone H2A.Z and canonical nucleosome. Cell. 2010;143:725–736. doi: 10.1016/j.cell.2010.10.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Ranjan A, et al. Nucleosome-free region dominates histone acetylation in targeting SWR1 to promoters for H2A.Z replacement. Cell. 2013;154:1232–1245. doi: 10.1016/j.cell.2013.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Yen K, Vinayachandran V, Pugh BF. SWR-C and INO80 chromatin remodelers recognize nucleosome-free regions near + 1 nucleosomes. Cell. 2013;154:1246–1256. doi: 10.1016/j.cell.2013.08.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Liu WH, Roemer SC, Port AM, Churchill ME. CAF-1-induced oligomerization of histones H3/H4 and mutually exclusive interactions with Asf1 guide H3/H4 transitions among histone chaperones and DNA. Nucleic Acids Res. 2012;40:11229–11239. doi: 10.1093/nar/gks906. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Winkler D, Zhou H, Dar M, Zhang Z, Luger K. Yeast CAF-1 assembles histone (H3–H4)2 tetramers prior to DNA deposition. Nucleic Acids Res. 2012;40:10139–10149. doi: 10.1093/nar/gks812. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- 179.Dechassa ML, Wyns K, Luger K. Scm3 deposits a (Cse4–H4)2 tetramer onto DNA through a Cse4–H4 dimer intermediate. Nucleic Acids Res. 2014;42:5532–5542. doi: 10.1093/nar/gku205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Zasadzińska E, Barnhart-Dailey MC, Kuich HPJL, Foltz DR. Dimerization of the CENP-A assembly factor HJURP is required for centromeric nucleosome deposition. EMBO J. 2013;32:2113–2124. doi: 10.1038/emboj.2013.142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 181.Ahmad K, Henikoff S. The histone variant H3.3 marks active chromatin by replication-independent nucleosome assembly. Mol Cell. 2002;9:1191–1200. doi: 10.1016/s1097-2765(02)00542-7. [DOI] [PubMed] [Google Scholar]
- 182.Black BE, et al. Centromere identity maintained by nucleosomes assembled with histone H3 containing the CENP-A targeting domain. Mol Cell. 2007;25:309–322. doi: 10.1016/j.molcel.2006.12.018. [DOI] [PubMed] [Google Scholar]
- 183.Adkins MW, Howar SR, Tyler JK. Chromatin disassembly mediated by the histone chaperone Asf1 is essential for transcriptional activation of the yeast PHO5 and PHO8 genes. Mol Cell. 2004;14:657–666. doi: 10.1016/j.molcel.2004.05.016. [DOI] [PubMed] [Google Scholar]
- 184.Korber P, et al. The histone chaperone Asf1 increases the rate of histone eviction at the yeast PHO5 and PHO8 promoters. J Biol Chem. 2006;281:5539–5545. doi: 10.1074/jbc.M513340200. [DOI] [PubMed] [Google Scholar]
- 185.Schwabish MA, Struhl K. Asf1 mediates histone eviction and deposition during elongation by RNA polymerase II. Mol Cell. 2006;22:415–422. doi: 10.1016/j.molcel.2006.03.014. [DOI] [PubMed] [Google Scholar]
- 186.Kuryan BG, et al. Histone density is maintained during transcription mediated by the chromatin remodeler RSC and histone chaperone NAP1 in vitro. Proc Natl Acad Sci USA. 2012;109:1931–1936. doi: 10.1073/pnas.1109994109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Chen X, et al. The histone chaperone Nap1 is a major regulator of histone H2A–H2B dynamics at the inducible GAL locus. Mol Cell Biol. 2016;36:1287–1296. doi: 10.1128/MCB.00835-15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Jamai A, Puglisi A, Strubin M. Histone chaperone spt16 promotes redeposition of the original H3–H4 histones evicted by elongating RNA polymerase. Mol Cell. 2009;35:377–383. doi: 10.1016/j.molcel.2009.07.001. [DOI] [PubMed] [Google Scholar]
- 189.Voth WP, et al. A role for FACT in repopulation of nucleosomes at inducible genes. PLoS ONE. 2014;9:e84092. doi: 10.1371/journal.pone.0084092. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 190.Schwabish MA, Struhl K. Evidence for eviction and rapid deposition of histones upon transcriptional elongation by RNA polymerase II. Mol Cell Biol. 2004;24:10111–10117. doi: 10.1128/MCB.24.23.10111-10117.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Kaplan CD, Laprade L, Winston F. Transcription elongation factors repress transcription initiation from cryptic sites. Science. 2003;301:1096–1099. doi: 10.1126/science.1087374. [DOI] [PubMed] [Google Scholar]
- 192.Xue YMM, et al. Histone chaperones Nap1 and Vps75 regulate histone acetylation during transcription elongation. Mol Cell Biol. 2013;33:1645–1656. doi: 10.1128/MCB.01121-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Bortvin A, Winston F. Evidence that Spt6p controls chromatin structure by a direct interaction with histones. Science. 1996;272:1473–1476. doi: 10.1126/science.272.5267.1473. [DOI] [PubMed] [Google Scholar]
- 194.Osakabe A, et al. Vertebrate Spt2 is a novel nucleolar histone chaperone that assists in ribosomal DNA transcription. J Cell Sci. 2013;126:1323–1332. doi: 10.1242/jcs.112623. [DOI] [PubMed] [Google Scholar]
- 195.Nourani A, Robert F, Winston F. Evidence that Spt2/Sin1, an HMG-like factor, plays roles in transcription elongation, chromatin structure, and genome stability in S. accharomyces cerevisiae. Mol Cell Biol. 2006;26:1496–1509. doi: 10.1128/MCB.26.4.1496-1509.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Selth L, et al. An rtt109-independent role for vps75 in transcription-associated nucleosome dynamics. Mol Cell Biol. 2009;29:4220–4234. doi: 10.1128/MCB.01882-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Kaplan T, et al. Cell cycle- and chaperone-mediated regulation of H3K56ac incorporation in yeast. PLoS Genet. 2008;4:e1000270. doi: 10.1371/journal.pgen.1000270. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Venkatesh S, et al. Set2 methylation of histone H3 lysine 36 suppresses histone exchange on transcribed genes. Nature. 2012;489:452–455. doi: 10.1038/nature11326. [DOI] [PubMed] [Google Scholar]
- 199.Pavri R, et al. Histone H2B monoubiquitination functions cooperatively with FACT to regulate elongation by RNA polymerase II. Cell. 2006;125:703–717. doi: 10.1016/j.cell.2006.04.029. [DOI] [PubMed] [Google Scholar]
- 200.Fleming AB, Kao CF, Hillyer C, Pikaart M, Osley MA. H2B ubiquitylation plays a role in nucleosome dynamics during transcription elongation. Mol Cell. 2008;31:57–66. doi: 10.1016/j.molcel.2008.04.025. [DOI] [PubMed] [Google Scholar]
- 201.Trujillo KM, Osley MA. A role for H2B ubiquitylation in DNA replication. Mol Cell. 2012;48:734–746. doi: 10.1016/j.molcel.2012.09.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Annunziato AT. Split decision: what happens to nucleosomes during DNA replication? J Biol Chem. 2005;280:12065–12068. doi: 10.1074/jbc.R400039200. [DOI] [PubMed] [Google Scholar]
- 203.Alabert C, et al. Nascent chromatin capture proteomics determines chromatin dynamics during DNA replication and identifies unknown fork components. Nat Cell Biol. 2014;16:281–293. doi: 10.1038/ncb2918. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Gambus A, et al. GINS maintains association of Cdc45 with MCM in replisome progression complexes at eukaryotic DNA replication forks. Nat Cell Biol. 2006;8:358–366. doi: 10.1038/ncb1382. [DOI] [PubMed] [Google Scholar]
- 205.Tan BC, Chien CT, Hirose S, Lee SC. Functional cooperation between FACT and MCM helicase facilitates initiation of chromatin DNA replication. EMBO J. 2006;25:3975–3985. doi: 10.1038/sj.emboj.7601271. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206.McCullough L, Connell Z, Petersen C, Formosa T. The abundant histone chaperones Spt6 and FACT collaborate to assemble, inspect, and maintain chromatin structure in Saccharomyces cerevisiae. Genetics. 2015;201:1031–1045. doi: 10.1534/genetics.115.180794. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Nekrasov M, et al. Histone H2A.Z inheritance during the cell cycle and its impact on promoter organization and dynamics. Nat Struct Mol Biol. 2012;19:1076–1083. doi: 10.1038/nsmb.2424. [DOI] [PubMed] [Google Scholar]
- 208.Xu M, et al. Partitioning of histone H3–H4 tetramers during DNA replication-dependent chromatin assembly. Science. 2010;328:94–98. doi: 10.1126/science.1178994. This mass spectrometry analysis of tagged H3.1 and H3.3 mononucleosomes shows that new and old H3.1–H4 dimers (unlike H3.3–H4 dimers) do not mix but can be mixed with new and old H2A–H2B. [DOI] [PubMed] [Google Scholar]
- 209.Ishimi Y, Komamura Y, You Z, Kimura H. Biochemical function of mouse minichromosome maintenance 2 protein. J Biol Chem. 1998;273:8369–8375. doi: 10.1074/jbc.273.14.8369. [DOI] [PubMed] [Google Scholar]
- 210.Ishimi Y, Komamura-Kohno Y, Arai K, Masai H. Biochemical activities associated with mouse Mcm2 protein. J Biol Chem. 2001;276:42744–42752. doi: 10.1074/jbc.M106861200. [DOI] [PubMed] [Google Scholar]
- 211.Sun J, et al. The architecture of a eukaryotic replisome. Nat Struct Mol Biol. 2015;22:976–982. doi: 10.1038/nsmb.3113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 212.Wittmeyer J, Joss L, Formosa T. Spt16 and Pob3 of Saccharomyces cerevisiae form an essential, abundant heterodimer that is nuclear, chromatin-associated, and copurifies with DNA polymerase α. Biochemistry. 1999;38:8961–8971. doi: 10.1021/bi982851d. [DOI] [PubMed] [Google Scholar]
- 213.Bowman A, Ward R, El-Mkami H, Owen-Hughes T, Norman D. Probing the (H3–H4)2 histone tetramer structure using pulsed EPR spectroscopy combined with site-directed spin labelling. Nucleic Acids Res. 2010;38:695–707. doi: 10.1093/nar/gkp1003. Describes the structure of the H3–H4 tetramer and highlights the dynamic nature of the H3 αN helix in the absence of H2A–H2B. [DOI] [PMC free article] [PubMed] [Google Scholar]






