Acetylation & Co: an expanding repertoire of histone acylations regulates chromatin and transcription

Department of Molecular and Cell Biology, University of Leicester, Leicester, U.K.

Find articles by Claire E. Barnes

David M. English

Department of Molecular and Cell Biology, University of Leicester, Leicester, U.K.

Find articles by David M. English

Shaun M. Cowley

Department of Molecular and Cell Biology, University of Leicester, Leicester, U.K.

Find articles by Shaun M. Cowley Department of Molecular and Cell Biology, University of Leicester, Leicester, U.K. Correspondence: Shaun M. Cowley (ku.ca.el@75cms) Received 2019 Jan 4; Revised 2019 Mar 8; Accepted 2019 Mar 12. Copyright © 2019 The Author(s).

This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).

Abstract

Packaging the long and fragile genomes of eukaryotic species into nucleosomes is all well and good, but how do cells gain access to the DNA again after it has been bundled away? The solution, in every species from yeast to man, is to post-translationally modify histones, altering their chemical properties to either relax the chromatin, label it for remodelling or make it more compact still. Histones are subject to a myriad of modifications: acetylation, methylation, phosphorylation, ubiquitination etc. This review focuses on histone acylations, a diverse group of modifications which occur on the ε-amino group of Lysine residues and includes the well-characterised Lysine acetylation. Over the last 50 years, histone acetylation has been extensively characterised, with the discovery of histone acetyltransferases (HATs) and histone deacetylases (HDACs), and global mapping experiments, revealing an association of hyperacetylated histones with accessible, transcriptionally active chromatin. More recently, there has been an explosion in the number of unique short chain ‘acylations’ identified by MS, including: propionylation, butyrylation, crotonylation, succinylation, malonylation and 2-hydroxyisobutyrylation. These novel modifications add a range of chemical environments to histones, and similar to acetylation, appear to accumulate at transcriptional start sites and correlate with gene activity.

Keywords: Acetylation, Chromatin, Histone, Transcription

Histone modification: crowbar and post-it note

Packaging DNA into nucleosomes helps protect the long fragile genomes of eukaryotic species. However, in doing so it presents a constant physical barrier to the protein machinery required for its replication, repair and transcription. Wrapped up tightly in its histone overcoat, how are cells able to gain access to the underlying DNA? Universally, in species as diverse as brewer’s yeast, fruit flies, worms and man, the answer is to chemically modify the histones to either help open up or compress the chromatin still further. And what a variety of modifications there are: acetylation, methylation, phosphorylation, ubiquitination, sumoylation, adenylation etc. So many that we would probably need another issue of Essays in Biochemistry to fit them all in and do them justice. This review will therefore focus on acetylation and a range of newly identified ‘acylations’ while recommending the reader to a number of other reviews which comprehensively cover these additional modifications [1–4].

Acylation (which includes acetylation) occurs on Lysine residues through the addition of an acyl group from an acyl-CoA donor to the ε-amino group of the Lysine side chain. Each of the core histones contains a globular core domain and a flexible N-terminal tail with an array of highly conserved Lysine residues, which being positively charged, have a natural affinity for both the DNA backbone and a negatively charged patch on neighbouring nucleosomes. The addition of the acyl group masks the positive charge on the Lysine residue, thereby reducing the affinity of the tail for chromatin, leaving the underlying DNA more exposed. This mechanism is amplified by the number of Lysine residues present in each N-terminal tail: 4 of the first 15 residues in histone H2A are Lysine (27%), 8 of 24 in H2B (33%), 8 of 36 in H3 (22%) and 5 of 20 in H4 (20%) ( Figure 1 ). The nucleosome has two copies of each core histone so that the 146 bp of DNA is surrounded by a flexible Lysine-rich environment, ripe for modification. It is worth bearing in mind that Lysine residues can also be methylated and ubiquitinated, which are two of the most abundant histone modifications. There is, therefore, a direct competition between different chemical modifications ( Figure 1 ). For example, if H3K27 is acetylated (H3K27ac – a marker of open chromatin), then it cannot be methylated (H3K27me3 – a marker of repressed chromatin), resulting in opposing transcriptional readouts [5]. Lysine acetylation plays a second role, in addition to changing the charge on the histone tail, it also functions as a binding site for proteins bearing a bromodomain (BD), which tend to be proteins with a pro-transcriptional function (discussed in detail below). Together, these activities promote a more open, less condensed form of chromatin that is transcriptionally permissive. Indeed, genes whose underlying chromatin is acylated are more likely to be transcribed [6–11].

An external file that holds a picture, illustration, etc. Object name is ebc-63-ebc20180061-g1.jpg

Schematic representation of histone modifications

Identified sites of histone acylation, methylation and phosphorylation in the four core histones are indicated. Amino acids located in the globular histone domains are shown in bold. Based on PTMs identified in [83,84,111].

Histone acetylation is often referred to as an ‘epigenetic’ modification, that is, a code which sits on top of the genetic code, regulating access to the DNA. However, there is much debate as to whether histone modifications are really epigenetic or not, since this characterisation implies at least a degree of heritability, i.e. that sites of histone acetylation are passed on from mother cell to daughter cell following division. Despite much effort in characterising the proteins which regulate acylation levels in cells, a mechanism for the stable inheritance of histone acetylation has remained elusive, suggesting that epigenetic is not quite the right term. Another argument against is that the half-life of acetyl-lysine (KAc) is typically 30 min to 2 h [12,13], so that if it does represent a code then it is a short-lived one. KAc is a highly dynamic modification being constantly added and removed to enable DNA accessibility when required. Indeed, there is evidence to suggest that cycles of acetylation and deacetylation are required for active gene transcription [14–16]. One way to view histone acylation is as part of a broad signalling mechanism [17], part crowbar (physically opening chromatin) and part Post-it Note (a temporary reminder that a job needs doing here), allowing DNA binding factors the access needed to get on with their job.

Histone acetylation promotes open chromatin and gene activation

Acetylation was the first identified histone acyl modification and is the most prevalent [18,19]. With remarkable prescience, Allfrey et al. [18] not only associated acetylation with facilitating RNA synthesis from nucleosomal DNA, but also hypothesised that histone acetylation was a mechanism for the dynamic regulation of gene transcription in vivo. Since then the association of increased levels of histone acetylation with transcriptional activation has been demonstrated by numerous studies [6,20,21]. Evidence for histone acetylation preceding transcription has been shown through reports of global increases in acetylation occurring prior to global increases in mRNA [20] and the association of acetylation at inducible gene promoters prior to their stimulation [21]. Histone acetylation as a potential causal agent for transcriptional activation was further substantiated by the identification of enzymes capable of catalysing the addition of and removal of acetyl groups, histone acetyltransferases (HATs) and histone deacetylases (HDACs) respectively [22–24]. The previous association of HATs and HDACs as transcriptional regulators further cemented the functional association of histone acetylation with transcription [25,26]. However, a word of caution, transcriptional regulation is more complex than acetylation = on and deacetylation = off, as several studies have indicated a requirement for HDAC activity in transcriptional activation [14,16,27–29]. The function of HATs and HDACs in the regulation of transcription is further complicated by their ever-expanding activities towards non-histone substrates as well as histones [12,30,31]. It may therefore be more apt to call these enzymes lysine acetyltransferases (KATs) and lysine deacetylases (KDACs) respectively, however for continuity with other literature we will continue to use HATs and HDACs throughout this review.

Within the context of chromatin, hyperacetylation of histone tails reduces the thermal stability of nucleosomes and a H4 histone tail–DNA complex [32,33], as well as increasing susceptibility to DNase I digestion [34–36]. High levels of histone acetylation have also been associated with a reduction in the formation of higher order or compacted chromatin structures [37,38], with H4K16ac having been linked to a failure of the formation of the 30-nm chromatin fibre and associated with transcriptionally active chromatin fractions [39]. The crystal structure of a nucleosome indicates an interaction occurs between H4K16 and a highly acidic patch on the H2A/H2B dimer of adjacent nucleosomes [40]. Modelling of the interaction of the H4 tail with adjacent nucleosomes has suggested the acetylation of H4K16 impairs and weakens the internucleosomal interaction of the H4 tail with the acidic H2A/H2B patch [41]. Acetylation also occurs within the core globular domains of histones (see Figure 1 ) [42,43]. A number of these modifications are situated at histone–DNA interacting regions and have the potential to modulate their interaction [44,45]. Indeed, alteration of DNA–histone interactions has been demonstrated, with acetylation of K115, K122 and K64 reducing DNA-binding affinity and increasing nucleosome mobility [46,47].

Site-specific lysine acetylation regulates gene expression

A plethora of specific KAc sites have been identified in each of the core histones (aided by high-resolution MS) that occur predominantly in the N-terminal tails (shown in Figure 1 ) [42,48,49]. A key development to investigating their function has been the generation of a range of antibodies which recognise specific histone KAc modifications [50]. In particular, they have allowed ChIP followed by next-generation sequencing (ChIP-seq) studies to map individual KAc sites across the genome, revealing histone acetylation in the proximity (1–2 kb) of transcription start sites (TSS) of actively transcribed genes [7–9]. However, there is some variation in the localisation, with H3K9ac and H3K27ac highly correlated to the TSS of active genes, whereas H4K12ac and H4K16ac are present at both TSS and along the gene body [8]. Distal regulatory regions (e.g. enhancers) have also been correlated with increased levels of H3K27ac [51,52] and localisation of the HAT, p300, in combination with mono-methylation of H3K4 [51,53,54]. However, the presence of specific histone modifications at enhancers is more complex, as recently H3K4 mono-, di- and tri-methylation have all been identified at enhancers and shown to correlate with enhancer RNA (eRNA) transcript levels (with H3K4me3 at sites of highest eRNA transcription) [55]. Furthermore, H3K16ac and the acetylation of globular domain residues H3K122 and H3K64 have also been associated with enhancers, which often lack H3K27ac (a classical mark of active enhancers) [47,56,57].

Mammalian cells contain numerous gene copies for each of the core histones [58] making mutagenesis studies of individual Lysine residues technically problematic. However, mutational analyses in yeast histone H4 demonstrated that Lys→Arg mutations at positions H4K5, K8 and K12 had additive effects upon gene expression changes, whereas K16R showed a greater individual effect, indicating a functionally distinct role for H4K16ac at least at a subset of genes [59]. Studies in mammalian systems have largely utilised in vitro transcriptional assays with recombinant nucleosomes to examine the function of specific histone acetylation modifications in gene regulation. For example, H3K14ac has been shown to be required for promoter nucleosome disassembly through Nap1 with an acetyl blocking H3K14R mutant preventing transcription and nucleosome eviction [60]. Recently, H3K9ac has been linked to the recruitment of the super elongation complex (SEC) to promote RNA pol II pause release, with mutation of Lysine to Arginine resulting in decreased transcription (due to increased RNA pol II pausing) [61]. The acetylation of globular domain residues, H3K56, H3K115 and H3K122, was investigated in Drosophila through acetyl mimicking (Gln) or acetyl blocking (Arg), mutations and demonstrated varied effects upon development, indicating potentially distinct functions of these modifications [62]. The acetylation of globular domain residues H3K122 and H3K64 has also been studied in mammalian systems, with overexpression of the acetyl mimicking H3K64Q resulting in increased gene expression and H3K122Q showing increased transcriptional activation in in vitro transcriptional assays [47,63]. The mechanism by which acetylation of H3K122 and H3K64 are proposed to increase transcription levels is by reducing DNA–histone interactions, which results in increased eviction of nucleosomes from the DNA [47,63].

Reading the runes: recognising and deciphering the pattern of histone acylations

Recognising the 50+ specific sites of histone acetylation ( Figure 1 ) is essential for the propagation of the ‘signal’ to downstream processes and functions. The major protein domain associated with KAc binding is the BD, although two other domains, the double PHD finger domain and the YEATS domain, are also capable of recognising KAc residues on histones [64–67]. In mammalian species approximately 61 BDs in 46 proteins have been identified, which include histone modifying enzymes and chromatin remodelling complexes, further indicating the close association between KAc recognition and chromatin state. A non-specific DNA binding capacity has also been identified in several BDs indicating a potential mechanism for enhancing and stabilising BD–chromatin interactions [68,69]. BDs are also implicated in additional functions such as non-histone KAc recognition. BRD3 for example, has been shown to bind to an acetylated form of the transcription factor, Gata1 [70]; and the second BD of BRD4 has the capacity to interact with acetylated cyclin T1 (a core component of P-TEFb), although this interaction alone is not sufficient for full activation of P-TEFb dependent transcription [71]. Inhibition of the p300/CBP BD caused no overall change in the localisation of p300, but reduced the levels of H3K27ac at enhancers indicating a role for the BD in regulating the catalytic activity towards H3K27 [72]. A key concept, highlighted throughout the study of chromatin PTMs, is the high degree of cross-talk between different modifications. For example, the tandem Tudor domain of SGF29 is essential for the recruitment of the SAGA HAT complex to sites of H3K4me3, which enables the processive acetylation of H3 tails [73,74]. Phosphorylation of H3S10 promotes GCN5-mediated acetylation of H3K14 through enhanced binding of GCN5 to the H3 tail [75,76]. H3S10 phosphorylation has also been shown to recruit the HAT, MOF, via the adaptor protein 14-3-3 and this phosphorylation-dependent recruitment is required for the acetylation of H4K16 [77].

A constellation of novel acylations

As discussed above, histone acetylation was discovered in the 1960s and has been characterised extensively over the last 50 years with the discovery of HATs and HDACs, and the global mapping of these modifications across the genome. Indeed it would have been reasonable to conclude that’s all she wrote in relation to histone acylation. However, over the last few years there has been an explosion in the number of unique short-chain histone Lysine ‘acylations’ identified by MS, these include: propionylation (KPr) [78], butyrylation (KBu) [78], crotonylation (KCr) [42], succinylation (KSucc) [79], malonylation (KMal) [79], 2-hydroxyisobutyrylation (KHib) [80], glutarylation (KGlu) [81], β-hydroxybutyrylation (KBhb) [82], and most recently benzoylation (KBz) [83] (summarised in Table 1 ). These modifications arise from their corresponding acyl-CoAs (e.g. propionyl-CoA, crotonyl-CoA etc.) and have different chemical properties. Hydrophobic groups (KPr, KBu, KCr and KBz) neutralise the positive charge of lysine residues (like acetylation), the acidic groups (KSucc, KMal and KGlu) change the positive charge to a negative charge, while polar groups (KHib and KBhb) allow hydrogen bond formation with interacting molecules. KBz stands out as the only known histone PTM with an aromatic acyl group, while KCr is planar and KBhb and KHib are branched (for a more comprehensive review of histone acylations, see [84]). An initial question to arise from the discovery of these modifications is whether they are ‘written’ and ‘erased’ by the same HATs and HDACs that control acetylation. A variety of in vivo and in vitro studies have demonstrated that the known HAT and HDAC families have wide-ranging acylation and deacylation capabilities (summarised in Table 1 ). The results of these studies suggest that HATs/HDACs show limited specificity for acylations, exemplified by the wide range of p300/CBP activities [10,85]. To date, no enzymes in addition to HATs and HDACs have been shown to be responsible for directly adding or removing these acylations. However, other enzymes do play roles in regulating their levels, for example the α-KGDH complex increases the local concentration of succinyl-CoA allowing GCN5 to succinylate H3K79 [86]. While CDYL acts as crotonyl-CoA hydratase, negatively regulating histone KCr [87], indicating further complexity in the regulation of histone acylations.

Table 1

Summary of alternative histone acyl modifications
Acyl group structurePropertiesHATsHDACs
An external file that holds a picture, illustration, etc. Object name is ebc-63-ebc20180061-i1.jpgHydrophobicp300/CBP [79,93], PCAF [112], GCN5 [93,113], MOF, HBO1, MOZ [114]SIRT1/2/3 [115]
An external file that holds a picture, illustration, etc. Object name is ebc-63-ebc20180061-i2.jpgHydrophobicp300/CBP [78,93], PCAF [93], GCN5 [93,113]SIRT1/2/3 [115]
An external file that holds a picture, illustration, etc. Object name is ebc-63-ebc20180061-i3.jpgHydrophobicp300/CBP [10], MOF [95]HDACs 1/2/3 [96,116–118], SIRT1/2/3 [119]
An external file that holds a picture, illustration, etc. Object name is ebc-63-ebc20180061-i4.jpgHydrophobic SIRT2 [83]
An external file that holds a picture, illustration, etc. Object name is ebc-63-ebc20180061-i5.jpgAcidic SIRT5 [120]
An external file that holds a picture, illustration, etc. Object name is ebc-63-ebc20180061-i6.jpgAcidicGCN5 [86]SIRT5 [120,121], SIRT7 [99]
An external file that holds a picture, illustration, etc. Object name is ebc-63-ebc20180061-i7.jpgAcidicp300 [81]SIRT5 [81]
An external file that holds a picture, illustration, etc. Object name is ebc-63-ebc20180061-i8.jpgPolarp300 [94], Tip60 [122]HDACs 1/2/3 (largely 2 and 3) [80,122]
An external file that holds a picture, illustration, etc. Object name is ebc-63-ebc20180061-i9.jpgPolarp300 [85]

The chemical structures and properties of acyl groups and the enzymes which are currently known to catalyse the addition or removal of these from histones are shown.

A further point for debate is whether these modifications arise due to the chemical reactivity of acyl-CoAs without the need for enzymatic catalysis, as has been seen in the favourable conditions of the mitochondria (alkaline pH and high acyl-CoA concentrations) [88,89]. Recently, Simithy et al. [19] tried to address this and determined that all the histone-acyl modifications they studied could occur chemically (although the acyl-CoA concentrations used were far above physiological levels) as well as through HAT activity, albeit with decreased efficiencies for the acidic, branched or planar acyl-CoAs. Interestingly, the HATs tested showed a specificity for N-terminal histone tails whereas, sites of chemical addition were closer to the C-terminus [19]. The formation of high-energy cyclic anhydride intermediates that rapidly acylate proteins by succinyl-CoA and glutaryl-CoA suggests that KSucc and KGlu may occur non-enzymatically and perhaps this mechanism may be more prevalent for certain acylations than others [90]. Further studies are still required to determine the balance between chemical acylation and HAT activity. The levels of acyl-CoAs reflect the metabolic status of the cell [91] and studies have shown that altering acyl-CoA concentrations can modify the levels of histone acylations [10,19,92], highlighting an interesting link between cell metabolism and chromatin modifications. The relative abundance of each acyl-CoA and differences in the relative levels of these between cell types [19] may therefore, at least partially, regulate the abundance of different chromatin acylation marks.

Physiological roles of diverse acylations

A number of studies have shown that similar to acetylation, many of the alternative acyl marks are found at transcriptionally active regions of the genome. For instance, H3K9Bhb is enriched at gene promoters in mouse liver tissue [82], suggesting a role in transcriptional regulation. KCr, KBu, KPr and KHib have all been shown to directly stimulate transcription to a similar (or even greater) extent than acetylation, using cell-free assays [10,11,93,94]. As many of these marks occur at the same sites as acetylation, it raises the question as to whether their functions overlap or diverge. CBP/p300 mutants lacking acetyltransferase but retaining crotonyltransferase activity were still able to enhance transcription, suggesting a role for crotonylation in enhancing transcription [95]. Further studies indicate that crotonylation plays a role in maintaining the pluripotent state in mouse embryonic stem cells [96], assists histone replacement during spermatogenesis [87] and can reverse HIV latency [97], potentially providing a therapeutic opportunity. One recent study has suggested that malonylation of yeast H2A may lead to a chromosome segregation defect [98]. Hypersuccinylation, achieved by both the depletion of SIRT7 and succinate dehydrogenase, results in defects in DNA repair [99,100]. These studies highlight the ever-expanding functions of histone lysine acylations.

As discussed above in relation to KAc, ‘readers’ of the histone modification play a critical role in interpreting and propagating the signal and the same appears to be the case for the newly identified acylations. Studies using known KAc reader proteins have identified a range of binding capabilities for the newer acyl marks. Human BDs have a general capability to bind KPr (as the hydrocarbon chain of KPr is only one carbon longer than KAc) but not KCr or KBu (other than BRD9, CECR2 and TAF1) [101]. An investigation of acyl marks at H4K5/K8 showed that the first BD (BD1) of BRDT binds both H4K5AcK8Ac and H4K5AcK8Bu; however, binding is abolished by butyrylation at H4K5 [11], suggesting a competition between acylations. It was subsequently shown that the binding of BRDT and BRD4-BD1 to H4K5 was enhanced by any acylation at H4K8 [102]. The Double PHD Finger (DPF) domains of two HATs MOZ and MORF have been shown to bind preferentially to H3K14Cr and H3K14Bu respectively [103,104]. The ability for these HATs to bind different acyl marks is thought to be vital for the spread of histone acylation which is proposed to help form and maintain open chromatin. In addition to BDs, proteins containing a YEATS domain, have shown a general preference for KCr over KAc. The AF9 YEATS domain can bind to several KCr marks on histone H3 (K9, K18, K27) as well as KPr and KBu marks [105], while Taf14 binds preferentially to H3K9Cr [106] and YEATS2 to H3K27Cr [107]. Using a mutated version of TAF14 designed to selectively bind H3K9Cr over H3K9Ac; Klein et al. [108] were able to show that there may be a differential requirement of H3K9Ac and H3K9Cr in the expression of TAF14-regulated genes. More recently it has been suggested that the YEATS domain of GAS41 is a pH-dependent reader of H3K122Succ [109], hinting that there may be a further expansion of the reading capabilities of the YEATS domain family (reviewed in [110]). YEATS and PHD finger domains selecting for alternative acyl marks over acetylation is an exciting discovery, and though there is clearly far more to be discovered in terms of acyl readers, it may be the case that the recruitment of distinct readers by different acyl modifications leads to different transcriptional readouts.

Summary

Within the last decade, multiple histone acylation marks have been discovered that have distinct characteristics in addition to those of histone acetylation.

These marks appear to be regulated by HATs and HDACs, but the contribution of non-enzymatic acylation cannot be discounted.

There are clear links with the newer acylations and active transcription, which recruit diverse reader proteins that are able to further modify or remodel chromatin.

There is clearly still a lot to discover in terms of the function and physiological relevance of these diverse chemical modifications and (no pun intended) this marks a very exciting time in the study of chromatin biology.

Acknowledgments

We thank members of the Cowley lab for useful discussions and comments on the manuscript. We apologise to those whose work could not be cited within this review due to space limitations or our own oversights.

Abbreviations

BDbromodomain
eRNAenhancer RNA
HAThistone acetyltransferase
HDAChistone deacetylase
KAcacetyl-lysine
KBhbβ-hydroxybutyryl-lysine
KBubutyryl-lysine
KBzbenzoyl-lysine
KCrcrotonyl-lysine
KGluglutaryl-lysine
KHib2-hydroxyisobutyryl-lysine
KMalmalonyl-lysine
KPrpropionyl-lysine
KSuccsuccinyl-lysine
PHDPlant homeodomain
PTMPost-translational modification
TSStranscription start site

Competing interests

The authors declare that there are no competing interests associated with the manuscript.

Funding

This work was supported by Studentships from BBSRC-MIBTP (to C.E.B) and MRC-IMPACT Ph.D. (D.M.E) programmes respectively; a senior non-clinical fellowship from MRC [grant number MR/J009202/1 (to S.M.C.)]; and BBSRC project grants [grant numbers BB/N002954/1, BB/P021689/1 (to S.M.C.)].

References

1. Bannister A.J. and Kouzarides T. (2011) Regulation of chromatin by histone modifications . Cell Res. 21 , 381–395 10.1038/cr.2011.22 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

2. Kouzarides T. (2007) Chromatin modifications and their function . Cell 128 , 693–705 10.1016/j.cell.2007.02.005 [PubMed] [CrossRef] [Google Scholar]

3. Izzo A. and Schneider R. (2010) Chatting histone modifications in mammals . Brief Funct. Genomics 9 , 429–443 10.1093/bfgp/elq024 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

4. Banerjee T. and Chakravarti D. (2011) A peek into the complex realm of histone phosphorylation . Mol. Cell. Biol. 31 , 4858–4873 10.1128/MCB.05631-11 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

5. Tie F., Banerjee R., Stratton C.A., Prasad-Sinha J., Stepanik V., Zlobin A.. et al. (2009) CBP-mediated acetylation of histone H3 lysine 27 antagonizes Drosophila polycomb silencing . Development 136 , 3131–3141 10.1242/dev.037127 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

6. Hebbes T.R., Thorne A.W. and Cranerobinson C. (1988) A direct link between core histone acetylation and transcriptionally active chromatin . EMBO J. 7 , 1395–1402 10.1002/j.1460-2075.1988.tb02956.x [PMC free article] [PubMed] [CrossRef] [Google Scholar]

7. Schubeler D., MacAlpine D.M., Scalzo D., Wirbelauer C., Kooperberg C., van Leeuwen F.. et al. (2004) The histone modification pattern of active genes revealed through genome-wide chromatin analysis of a higher eukaryote . Genes Dev. 18 , 1263–1271 10.1101/gad.1198204 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

8. Wang Z.B., Zang C.Z., Rosenfeld J.A., Schones D.E., Barski A., Cuddapah S.. et al. (2008) Combinatorial patterns of histone acetylations and methylations in the human genome . Nat. Genet. 40 , 897–903 10.1038/ng.154 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

9. Heintzman N.D., Stuart R.K., Hon G., Fu Y.T., Ching C.W., Hawkins R.D.. et al. (2007) Distinct and predictive chromatin signatures of transcriptional promoters and enhancers in the human genome . Nat. Genet. 39 , 311–318 10.1038/ng1966 [PubMed] [CrossRef] [Google Scholar]

10. Sabari B.R., Tang Z.Y., Huang H., Yong-Gonzalez V., Molina H., Kong H.E.. et al. (2015) Intracellular crotonyl-CoA stimulates transcription through p300-catalyzed histone crotonylation . Mol. Cell 58 , 203–215 10.1016/j.molcel.2015.02.029 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

11. Goudarzi A., Zhang D., Huang H., Barral S., Kwon O.K., Qi S.K.. et al. (2016) Dynamic competing histone H4 K5K8 acetylation and butyrylation are hallmarks of highly active gene promoters . Mol. Cell 62 , 169–180 10.1016/j.molcel.2016.03.014 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

12. Weinert B.T., Narita T., Satpathy S., Srinivasan B., Hansen B.K., Schölz C.. et al. (2018) Time-resolved analysis reveals rapid dynamics and broad scope of the CBP/p300 acetylome . Cell 174 , 231–244.e212 10.1016/j.cell.2018.04.033 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

13. Zheng Y.P., Thomas P.M. and Kelleher N.L. (2013) Measurement of acetylation turnover at distinct lysines in human histones identifies long-lived acetylation sites . Nat. Commun. 4 , 2203, , 10.1038/ncomms3203 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

14. Wang A., Kurdistani S.K. and Grunstein M. (2002) Requirement of Hos2 histone deacetylase for gene activity in yeast . Science 298 , 1412–1414 10.1126/science.1077790 [PubMed] [CrossRef] [Google Scholar]

15. Metivier R., Penot G., Hubner M.R., Reid G., Brand H., Kos M.. et al. (2003) Estrogen receptor-alpha directs ordered, cyclical, and combinatorial recruitment of cofactors on a natural target promoter . Cell 115 , 751–763 10.1016/S0092-8674(03)00934-6 [PubMed] [CrossRef] [Google Scholar]

16. Wang Z.B., Zang C.Z., Cui K.R., Schones D.E., Barski A., Peng W.Q.. et al. (2009) Genome-wide mapping of HATs and HDACs reveals distinct functions in active and inactive genes . Cell 138 , 1019–1031 10.1016/j.cell.2009.06.049 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

17. Schreiber S.L. and Bernstein B.E. (2002) Signaling network model of chromatin . Cell 111 , 771–778 10.1016/S0092-8674(02)01196-0 [PubMed] [CrossRef] [Google Scholar]

18. Allfrey V.G., Faulkner R. and Mirsky A.E. (1964) Acetylation and methylation of histones and their possible role in the regulation of rna synthesis . Proc. Natl. Acad. Sci. U.S.A. 51 , 786–794 10.1073/pnas.51.5.786 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

19. Simithy J., Sidoli S., Yuan Z.F., Coradin M., Bhanu N.V., Marchione D.M.. et al. (2017) Characterization of histone acylations links chromatin modifications with metabolism . Nat. Commun. 8 , 1141 10.1038/s41467-017-01384-9 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

20. Pogo B.G., Allfrey V.G. and Mirsky A.E. (1966) RNA synthesis and histone acetylation during the course of gene activation in lymphocytes . Proc. Natl. Acad. Sci. U.S.A. 55 , 805–812 10.1073/pnas.55.4.805 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

21. Clayton A.L., Hebbes T.R., Thorne A.W. and Crane-Robinson C. (1993) Histone acetylation and gene induction in human cells . FEBS Lett. 336 , 23–26 10.1016/0014-5793(93)81601-U [PubMed] [CrossRef] [Google Scholar]

22. Taunton J., Hassig C.A. and Schreiber S.L. (1996) A mammalian histone deacetylase related to the yeast transcriptional regulator Rpd3p . Science 272 , 408–411 10.1126/science.272.5260.408 [PubMed] [CrossRef] [Google Scholar]

23. Kleff S., Andrulis E.D., Anderson C.W. and Sternglanz R. (1995) Identification of a gene encoding a yeast histone H4 acetyltransferase . J. Biol. Chem. 270 , 24674–24677 10.1074/jbc.270.42.24674 [PubMed] [CrossRef] [Google Scholar]

24. Brownell J.E., Zhou J., Ranalli T., Kobayashi R., Edmondson D.G., Roth S.Y.. et al. (1996) Tetrahymena histone acetyltransferase A: a homolog to yeast Gcn5p linking histone acetylation to gene activation . Cell 84 , 843–851 10.1016/S0092-8674(00)81063-6 [PubMed] [CrossRef] [Google Scholar]

25. Vidal M. and Gaber R.F. (1991) RPD3 encodes a second factor required to achieve maximum positive and negative transcriptional states in Saccharomyces cerevisiae . Mol. Cell. Biol. 11 , 6317–6327 10.1128/MCB.11.12.6317 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

26. Georgakopoulos T. and Thireos G. (1992) Two distinct yeast transcriptional activators require the function of the GCN5 protein to promote normal levels of transcription . EMBO J. 11 , 4145–4152 10.1002/j.1460-2075.1992.tb05507.x [PMC free article] [PubMed] [CrossRef] [Google Scholar]

27. Bernstein B.E., Tong J.K. and Schreiber S.L. (2000) Genomewide studies of histone deacetylase function in yeast . Proc. Natl. Acad. Sci. U.S.A. 97 , 13708–13713 10.1073/pnas.250477697 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

28. Kim Y.J., Greer C.B., Cecchini K.R., Harris L.N., Tuck D.P. and Kim T.H. (2013) HDAC inhibitors induce transcriptional repression of high copy number genes in breast cancer through elongation blockade . Oncogene 32 , 2828–2835 10.1038/onc.2013.32 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

29. Greer C.B., Tanaka Y., Kim Y.J., Xie P., Zhang M.Q., Park I.H.. et al. (2015) Histone deacetylases positively regulate transcription through the elongation machinery . Cell Rep. 13 , 1444–1455 10.1016/j.celrep.2015.10.013 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

30. Scholz C., Weinert B.T., Wagner S.A., Beli P., Miyake Y., Qi J.. et al. (2015) Acetylation site specificities of lysine deacetylase inhibitors in human cells . Nat. Biotechnol. 33 , 415–U136 10.1038/nbt.3130 [PubMed] [CrossRef] [Google Scholar]

31. Choudhary C., Kumar C., Gnad F., Nielsen M.L., Rehman M., Walther T.C.. et al. (2009) Lysine acetylation targets protein complexes and co-regulates major cellular functions . Science 325 , 834–840 10.1126/science.1175371 [PubMed] [CrossRef] [Google Scholar]

32. Hong L., Schroth G.P., Matthews H.R., Yau P. and Bradbury E.M. (1993) Studies of the DNA binding properties of histone H4 amino terminus. Thermal denaturation studies reveal that acetylation markedly reduces the binding constant of the H4 “tail” to DNA . J. Biol. Chem. 268 , 305–314 [PubMed] [Google Scholar]

33. Yau P., Thorne A.W., Imai B.S., Matthews H.R. and Bradbury E.M. (1982) Thermal denaturation studies of acetylated nucleosomes and oligonucleosomes . Eur. J. Biochem. 129 , 281–288 10.1111/j.1432-1033.1982.tb07050.x [PubMed] [CrossRef] [Google Scholar]

34. Simpson R.T. (1978) Structure of chromatin containing extensively acetylated H3 and H4 . Cell 13 , 691–699 10.1016/0092-8674(78)90219-2 [PubMed] [CrossRef] [Google Scholar]

35. Vidali G., Boffa L.C., Bradbury E.M. and Allfrey V.G. (1978) Butyrate suppression of histone deacetylation leads to accumulation of multiacetylated forms of histones h-3 and h-4 and increased DNASE-i sensitivity of associated DNA sequences . Proc. Natl. Acad. Sci. U.S.A. 75 , 2239–2243 10.1073/pnas.75.5.2239 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

36. Hebbes T.R., Clayton A.L., Thorne A.W. and Cranerobinson C. (1994) Core histone hyperacetylation co-maps with generalized DNASE-i sensitivity in the chicken beta-globin chromosomal domain . EMBO J. 13 , 1823–1830 10.1002/j.1460-2075.1994.tb06451.x [PMC free article] [PubMed] [CrossRef] [Google Scholar]

37. Ridsdale J.A., Hendzel M.J., Delcuve G.P. and Davie J.R. (1990) Histone acetylation alters the capacity of the H1 histones to condense transcriptionally active competent chromatin . J. Biol. Chem. 265 , 5150–5156 [PubMed] [Google Scholar]

38. Tse C., Sera T., Wolffe A.P. and Hansen J.C. (1998) Disruption of higher-order folding by core histone acetylation dramatically enhances transcription of nucleosomal arrays by RNA polymerase III . Mol. Cell. Biol. 18 , 4629–4638 10.1128/MCB.18.8.4629 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

39. Shogren-Knaak M., Ishii H., Sun J.M., Pazin M.J., Davie J.R. and Peterson C.L. (2006) Histone H4-K16 acetylation controls chromatin structure and protein interactions . Science 311 , 844–847 10.1126/science.1124000 [PubMed] [CrossRef] [Google Scholar]

40. Luger K., Mader A.W., Richmond R.K., Sargent D.F. and Richmond T.J. (1997) Crystal structure of the nucleosome core particle at 2.8 angstrom resolution . Nature 389 , 251–260 10.1038/38444 [PubMed] [CrossRef] [Google Scholar]

41. Zhang R.T., Erler J. and Langowski J. (2017) Histone acetylation regulates chromatin accessibility: role of H4K16 in inter-nucleosome interaction . Biophys. J. 112 , 450–459 10.1016/j.bpj.2016.11.015 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

42. Tan M.J., Luo H., Lee S., Jin F.L., Yang J.S., Montellier E.. et al. (2011) Identification of 67 histone marks and histone lysine crotonylation as a new type of histone modification . Cell 146 , 1015–1027 10.1016/j.cell.2011.08.008 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

43. Zhang L.W., Eugeni E.E., Parthun M.R. and Freitas M.A. (2003) Identification of novel histone post-translational modifications by peptide mass fingerprinting . Chromosoma 112 , 77–86 10.1007/s00412-003-0244-6 [PubMed] [CrossRef] [Google Scholar]

44. Freitas M.A., Sklenar A.R. and Parthun M.R. (2004) Application of mass spectrometry to the identification and quantification of histone post-translational modifications . J. Cell. Biochem. 92 , 691–700 10.1002/jcb.20106 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

45. Cosgrove M.S., Boeke J.D. and Wolberger C. (2004) Regulated nucleosome mobility and the histone code . Nat. Struct. Mol. Biol. 11 , 1037–1043 10.1038/nsmb851 [PubMed] [CrossRef] [Google Scholar]

46. Manohar M., Mooney A.M., North J.A., Nakkula R.J., Picking J.W., Edon A.. et al. (2009) Acetylation of histone H3 at the nucleosome dyad alters DNA-histone binding . J. Biol. Chem. 284 , 23312–23321 10.1074/jbc.M109.003202 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

47. Di Cerbo V., Mohn F., Ryan D.P., Montellier E., Kacem S., Tropberger P.. et al. (2014) Acetylation of histone H3 at lysine 64 regulates nucleosome dynamics and facilitates transcription . Elife 3 , 10.7554/eLife.01632 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

48. Beck H.C., Nielsen E.C., Matthiesen R., Jensen L.H., Sehested M., Finn P.. et al. (2006) Quantitative proteomic analysis of posttranslational modifications of human histones . Mol. Cell. Proteomics 5 , 1314–1325 10.1074/mcp.M600007-MCP200 [PubMed] [CrossRef] [Google Scholar]

49. Tweedie-Cullen R.Y., Brunner A.M., Grossmann J., Mohanna S., Sichau D., Nanni P.. et al. (2012) Identification of combinatorial patterns of post-translational modifications on individual histones in the mouse brain . PLoS ONE 7 , e36980 10.1371/journal.pone.0036980 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

50. Turner B.M. and Fellows G. (1989) Specific antibodies reveal ordered and cell-cycle-related use of histone-H4 acetylation sites in mammalian cells . Eur. J. Biochem. 179 , 131–139 10.1111/j.1432-1033.1989.tb14530.x [PubMed] [CrossRef] [Google Scholar]

51. Creyghton M.P., Cheng A.W., Welstead G.G., Kooistra T., Carey B.W., Steine E.J.. et al. (2010) Histone H3K27ac separates active from poised enhancers and predicts developmental state . Proc. Natl. Acad. Sci. U.S.A. 107 , 21931–21936 10.1073/pnas.1016071107 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

52. Rada-Iglesias A., Bajpai R., Swigut T., Brugmann S.A., Flynn R.A. and Wysocka J. (2011) A unique chromatin signature uncovers early developmental enhancers in humans . Nature 470 , 279–+ 10.1038/nature09692 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

53. Visel A., Blow M.J., Li Z.R., Zhang T., Akiyama J.A., Holt A.. et al. (2009) ChIP-seq accurately predicts tissue-specific activity of enhancers . Nature 457 , 854–858 10.1038/nature07730 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

54. Heintzman N.D., Hon G.C., Hawkins R.D., Kheradpour P., Stark A., Harp L.F.. et al. (2009) Histone modifications at human enhancers reflect global cell-type-specific gene expression . Nature 459 , 108–112 10.1038/nature07829 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

55. Henriques T., Scruggs B.S., Inouye M.O., Muse G.W., Williams L.H., Burkholder A.B.. et al. (2018) Widespread transcriptional pausing and elongation control at enhancers . Genes Dev. 32 , 26–41 10.1101/gad.309351.117 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

56. Taylor G. C.A., Eskeland R., Hekimoglu-Balkan B., Pradeepa M.M. and Bickmore W.A. (2013) H4K16 acetylation marks active genes and enhancers of embryonic stem cells, but does not alter chromatin compaction . Genome Res. 23 , 2053–2065 10.1101/gr.155028.113 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

57. Pradeepa M.M., Grimes G.R., Kumar Y., Olley G., Taylor G. C.A., Schneider R.. et al. (2016) Histone H3 globular domain acetylation identifies a new class of enhancers . Nat. Genet. 48 , 681–+ 10.1038/ng.3550 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

58. Marzluff W.F., Gongidi P., Woods K.R., Jin J. and Maltais L.J. (2002) The human and mouse replication-dependent histone genes . Genomics 80 , 487–498 10.1006/geno.2002.6850 [PubMed] [CrossRef] [Google Scholar]

59. Dion M.F., Altschuler S.J., Wu L.F. and Rando O.J. (2005) Genomic characterization reveals a simple histone H4 acetylation code . Proc. Natl. Acad. Sci. U.S.A. 102 , 5501–5506 10.1073/pnas.0500136102 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

60. Luebben W.R., Sharma N. and Nyborg J.K. (2010) Nucleosome eviction and activated transcription require p300 acetylation of histone H3 lysine 14 . Proc. Natl. Acad. Sci. U.S.A. 107 , 19254–19259 10.1073/pnas.1009650107 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

61. Gates L.A., Shi J.J., Rohira A.D., Feng Q., Zhu B.K., Bedford M.T.. et al. (2017) Acetylation on histone H3 lysine 9 mediates a switch from transcription initiation to elongation . J. Biol. Chem. 292 , 14456–14472 10.1074/jbc.M117.802074 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

62. Graves H.K., Wang P.P., Lagarde M., Chen Z.H. and Tyler J.K. (2016) Mutations that prevent or mimic persistent post-translational modifications of the histone H3 globular domain cause lethality and growth defects in Drosophila . Epigenetics Chromatin 9 , 9, 10.1186/s13072-016-0059-3 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

63. Tropberger P., Pott S., Keller C., Kamieniarz-Gdula K., Caron M., Richter F.. et al. (2013) Regulation of transcription through acetylation of H3K122 on the lateral surface of the histone octamer . Cell 152 , 859–872 10.1016/j.cell.2013.01.032 [PubMed] [CrossRef] [Google Scholar]

64. Lange M., Kaynak B., Forster U.B., Tonjes M., Fischer J.J., Grimm C.. et al. (2008) Regulation of muscle development by DPF3, a novel histone acetylation and methylation reader of the BAF chromatin remodeling complex . Genes Dev. 22 , 2370–2384 10.1101/gad.471408 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

65. Qiu Y., Liu L., Zhao C., Han C.C., Li F.D., Zhang J.H.. et al. (2012) Combinatorial readout of unmodified H3R2 and acetylated H3K14 by the tandem PHD finger of MOZ reveals a regulatory mechanism for HOXA9 transcription . Genes Dev. 26 , 1376–1391 10.1101/gad.188359.112 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

66. Li Y.Y., Wen H., Xi Y.X., Tanaka K., Wang H.B., Peng D.N.. et al. (2014) AF9 YEATS domain links histone acetylation to DOT1L-mediated H3K79 methylation . Cell 159 , 558–571 10.1016/j.cell.2014.09.049 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

67. Dhalluin C., Carlson J.E., Zeng L., He C., Aggarwal A.K. and Zhou M.M. (1999) Structure and ligand of a histone acetyltransferase bromodomain . Nature 399 , 491–496 10.1038/20974 [PubMed] [CrossRef] [Google Scholar]

68. Miller T. C.R., Simon B., Rybin V., Grotsch H., Curtet S., Khochbin S.. et al. (2016) A bromodomain-DNA interaction facilitates acetylation-dependent bivalent nucleosome recognition by the BET protein BRDT . Nat. Commun. 7 , 1385 10.1038/ncomms13855 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

69. Morrison E.A., Sanchez J.C., Ronan J.L., Farrell D.P., Varzavand K., Johnson J.K.. et al. (2017) DNA binding drives the association of BRG1/hBRM bromodomains with nucleosomes . Nat. Commun. 8 , 16080 10.1038/ncomms16080 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

70. Lamonica J.M., Deng W.L., Kadauke S., Campbell A.E., Gamsjaeger R., Wang H.X.. et al. (2011) Bromodomain protein Brd3 associates with acetylated GATA1 to promote its chromatin occupancy at erythroid target genes . Proc. Natl. Acad. Sci. U.S.A. 108 , E159–E168 10.1073/pnas.1102140108 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

71. Schroder S., Cho S.Y., Zeng L., Zhang Q., Kaehlcke K., Mak L.. et al. (2012) Two-pronged Binding with bromodomain-containing protein 4 liberates positive transcription elongation factor b from inactive ribonucleoprotein complexes . J. Biol. Chem. 287 , 1090–1099 10.1074/jbc.M111.282855 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

72. Raisner R., Kharbanda S., Jin L.Y., Jeng E., Chan E., Merchant M.. et al. (2018) Enhancer activity requires CBP/P300 bromodomain-dependent histone H3K27 acetylation . Cell Rep. 24 , 1722–1729 10.1016/j.celrep.2018.07.041 [PubMed] [CrossRef] [Google Scholar]

73. Vermeulen M., Eberl H.C., Matarese F., Marks H., Denissov S., Butter F.. et al. (2010) Quantitative interaction proteomics and genome-wide profiling of epigenetic histone marks and their readers . Cell 142 , 967–980 10.1016/j.cell.2010.08.020 [PubMed] [CrossRef] [Google Scholar]

74. Ringel A.E., Cieniewicz A.M., Taverna S.D. and Wolberger C. (2015) Nucleosome competition reveals processive acetylation by the SAGA HAT module . Proc. Natl. Acad. Sci. U.S.A. 112 , E5461–E5470 10.1073/pnas.1508449112 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

75. Cheung P., Tanner K.G., Cheung W.L., Sassone-Corsi P., Denu J.M. and Allis C.D. (2000) Synergistic coupling of histone H3 phosphorylation and acetylation in response to epidermal growth factor stimulation . Mol. Cell 5 , 905–915 10.1016/S1097-2765(00)80256-7 [PubMed] [CrossRef] [Google Scholar]

76. Clements A., Poux A.N., Lo W.S., Pillus L., Berger S.L. and Marmorstein R. (2003) Structural basis for histone and phosphohistone binding by the GCN5 histone acetyltransferase . Mol. Cell 12 , 461–473 10.1016/S1097-2765(03)00288-0 [PubMed] [CrossRef] [Google Scholar]

77. Zippo A., Serafini R., Rocchigiani M., Pennacchini S., Krepelova A. and Oliviero S. (2009) Histone crosstalk between H3S10ph and H4K16ac generates a histone code that mediates transcription elongation . Cell 138 , 1122–1136 10.1016/j.cell.2009.07.031 [PubMed] [CrossRef] [Google Scholar]

78. Chen Y., Sprung R., Tang Y., Ball H., Sangras B., Kim S.C.. et al. (2007) Lysine propionylation and butyrylation are novel post-translational modifications in histones . Mol. Cell. Proteomics 6 , 812–819 10.1074/mcp.M700021-MCP200 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

79. Xie Z.Y., Dai J. B.A., Dai L.Z., Tan M.J., Cheng Z.Y., Wu Y.M.. et al. (2012) Lysine succinylation and lysine malonylation in histones . Mol. Cell. Proteomics 11 , 100–107 10.1074/mcp.M111.015875 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

80. Dai L.Z., Peng C., Montellier E., Lu Z.K., Chen Y., Ishii H.. et al. (2014) Lysine 2-hydroxyisobutyrylation is a widely distributed active histone mark . Nat. Chem. Biol. 10 , 365–U373 10.1038/nchembio.1497 [PubMed] [CrossRef] [Google Scholar]

81. Tan M.J., Peng C., Anderson K.A., Chhoy P., Xie Z.Y., Dai L.Z.. et al. (2014) Lysine glutarylation is a protein posttranslational modification regulated by SIRT5 . Cell Metab. 19 , 605–617 10.1016/j.cmet.2014.03.014 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

82. Xie Z.Y., Zhang D., Chung D.J., Tang Z.Y., Huang H., Dai L.Z.. et al. (2016) Metabolic regulation of gene expression by histone lysine beta-hydroxybutyrylation . Mol. Cell 62 , 194–206 10.1016/j.molcel.2016.03.036 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

83. Huang H., Zhang D., Wang Y., Perez-Neut M., Han Z., Zheng Y.G.. et al. (2018) Lysine benzoylation is a histone mark regulated by SIRT2 . Nat. Commun. 9 , 11 10.1038/s41467-018-05567-w [PMC free article] [PubMed] [CrossRef] [Google Scholar]

84. Sabari B.R., Zhang D., Allis C.D. and Zhao Y.M. (2017) Metabolic regulation of gene expression through histone acylations . Nat. Rev. Mol. Cell Biol. 18 , 90–101 10.1038/nrm.2016.140 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

85. Kaczmarska Z., Ortega E., Goudarzi A., Huang H., Kim S., Marquez J.A.. et al. (2017) Structure of p300 in complex with acyl-CoA variants . Nat. Chem. Biol. 13 , 21–29 10.1038/nchembio.2217 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

86. Wang Y.G., Guo Y.R., Liu K., Yin Z., Liu R., Xia Y.. et al. (2017) KAT2A coupled with the alpha-KGDH complex acts as a histone H3 succinyltransferase . Nature 552 , 273–277 10.1038/nature25003 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

87. Liu S.M., Yu H.J., Liu Y.Q., Liu X.H., Zhang Y., Bu C.. et al. (2017) Chromodomain protein CDYL acts as a crotonyl-CoA hydratase to regulate histone crotonylation and spermatogenesis . Mol. Cell 67 , 853–866 10.1016/j.molcel.2017.07.011 [PubMed] [CrossRef] [Google Scholar]

88. Newman J.C., He W.J. and Verdin E. (2012) Mitochondrial protein acylation and intermediary metabolism: regulation by sirtuins and implications for metabolic disease . J. Biol. Chem. 287 , 42436–42443 10.1074/jbc.R112.404863 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

89. Wagner G.R. and Hirschey M.D. (2014) Nonenzymatic protein acylation as a carbon stress regulated by sirtuin deacylases . Mol. Cell 54 , 5–16 10.1016/j.molcel.2014.03.027 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

90. Wagner G.R., Bhatt D.P., O’Connell T.M., Thompson J.W., Dubois L.G., Backos D.S.. et al. (2017) A class of reactive Acyl-CoA species reveals the non-enzymatic origins of protein acylation . Cell Metab. 25 , 823–837 10.1016/j.cmet.2017.03.006 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

91. Dutta A., Abmayr S.M. and Workman J.L. (2016) Diverse activities of histone acylations connect metabolism to chromatin function . Mol. Cell 63 , 547–552 10.1016/j.molcel.2016.06.038 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

92. Wellen K.E., Hatzivassiliou G., Sachdeva U.M., Bui T.V., Cross J.R. and Thompson C.B. (2009) ATP-citrate lyase links cellular metabolism to histone acetylation . Science 324 , 1076–1080 10.1126/science.1164097 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

93. Kebede A.F., Nieborak A., Shahidian L.Z., Le Gras S., Richter F., Gomez D.A.. et al. (2017) Histone propionylation is a mark of active chromatin . Nat. Struct. Mol. Biol. 24 , 1048–+ 10.1038/nsmb.3490 [PubMed] [CrossRef] [Google Scholar]

94. Huang H., Tang S., Ji M., Tang Z.Y., Shimada M., Liu X.J.. et al. (2018) p300-mediated lysine 2-hydroxyisobutyrylation regulates glycolysis (vol 70, pg 663, 2018) . Mol. Cell 70 , 984–984 10.1016/j.molcel.2018.05.035 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

95. Liu X.G., Wei W., Liu Y.T., Yang X.L., Wu J., Zhang Y.. et al. (2017) MOF as an evolutionarily conserved histone crotonyltransferase and transcriptional activation by histone acetyltransferase-deficient and crotonyltransferase-competent CBP/p300 . Cell Discov. 3 , 17 10.1038/celldisc.2017.16 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

96. Wei W., Liu X.G., Chen J.W., Gao S.N., Lu L., Zhang H.F.. et al. (2017) Class I histone deacetylases are major histone decrotonylases: evidence for critical and broad function of histone crotonylation in transcription . Cell Res. 27 , 898–915 10.1038/cr.2017.68 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

97. Jiang G.C., Nguyen D., Archin N.M., Yukl S.A., Mendez-Lagares G., Tang Y.Y.. et al. (2018) HIV latency is reversed by ACSS2-driven histone crotonylation . J. Clin. Invest. 128 , 1190–1198 10.1172/JCI98071 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

98. Ishiguro T., Tanabe K., Kobayashi Y., Mizumoto S., Kanai M. and Kawashima S.A. (2018) Malonylation of histone H2A at lysine 119 inhibits Bub1-dependent H2A phosphorylation and chromosomal localization of shugoshin proteins . Sci. Rep. 8 , 10 10.1038/s41598-018-26114-z [PMC free article] [PubMed] [CrossRef] [Google Scholar]

99. Li L., Shi L., Yang S.D., Yan R.R., Zhang D., Yang J.G.. et al. (2016) SIRT7 is a histone desuccinylase that functionally links to chromatin compaction and genome stability . Nat. Commun. 7 , 17, 10.1038/ncomms12235, [PMC free article] [PubMed] [CrossRef] [Google Scholar]

100. Smestad J., Erber L., Chen Y. and Maher L.J. (2018) Chromatin succinylation correlates with active gene expression and is perturbed by defective TCA cycle metabolism . iScience 2 , 63–75 10.1016/j.isci.2018.03.012 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

101. Flynn E.M., Huang O.W., Poy F., Oppikofer M., Bellon S.F., Tang Y.. et al. (2015) A subset of human bromodomains recognizes butyryllysine and crotonyllysine histone peptide modifications . Structure 23 , 1801–1814 10.1016/j.str.2015.08.004 [PubMed] [CrossRef] [Google Scholar]

102. Olp M.D., Zhu N. and Smith B.C. (2017) Metabolically derived lysine acylations and neighboring modifications tune the binding of the BET bromodomains to histone H4 . Biochemistry 56 , 5485–5495 10.1021/acs.biochem.7b00595 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

103. Xiong X.Z., Panchenko T., Yang S., Zhao S., Yan P.Q., Zhang W.H.. et al. (2016) Selective recognition of Histone crotonylation by double PHD fingers of MOZ and DPF2 . Nat. Chem. Biol. 12 , 1111–1118 10.1038/nchembio.2218 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

104. Klein B.J., Simithy J., Wang X.L., Ahn J., Andrews F.H., Zhang Y.. et al. (2017) Recognition of histone H3K14 acylation by MORF . Structure 25 , 650–654.e2 10.1016/j.str.2017.02.003 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

105. Li Y.Y., Sabari B.R., Panchenko T., Wen H., Zhao D., Guan H.P.. et al. (2016) Molecular coupling of histone crotonylation and active transcription by AF9 YEATS domain . Mol. Cell 62 , 181–193 10.1016/j.molcel.2016.03.028 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

106. Andrews F.H., Shinsky S.A., Shanle E.K., Bridgers J.B., Gest A., Tsun I.K.. et al. (2016) The Taf14 YEATS domain is a reader of histone crotonylation . Nat. Chem. Biol. 12 , U396–U398 10.1038/nchembio.2065 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

107. Zhao D., Guan H.P., Zhao S., Mi W.Y., Wen H., Li Y.Y.. et al. (2016) YEATS2 is a selective histone crotonylation reader . Cell Res. 26 , 629–632 10.1038/cr.2016.49 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

108. Klein B.J., Vann K.R., Andrews F.H., Wang W.W., Zhang J.B., Zhang Y.. et al. (2018) Structural insights into the pi-pi-pi stacking mechanism and DNA-binding activity of the YEATS domain . Nat. Commun. 9 , 11 10.1038/s41467-018-07072-6 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

109. Wang Y., Jin J., Chung M. W.H., Feng L., Sun H.Y. and Hao Q. (2018) Identification of the YEATS domain of GAS41 as a pH-dependent reader of histone succinylation . Proc. Natl. Acad. Sci. U.S.A. 115 , 2365–2370 10.1073/pnas.1717664115 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

110. Zhao D., Li Y.Y., Xiong X.Z., Chen Z.L. and Li H.T. (2017) YEATS Domain-A histone acylation reader in health and disease . J. Mol. Biol. 429 , 1994–2002 10.1016/j.jmb.2017.03.010 [PubMed] [CrossRef] [Google Scholar]

111. Huang H., Sabari B.R., Garcia B.A., Allis C.D. and Zhao Y. (2014) SnapShot: histone modifications . Cell 159 , 458.e451 10.1016/j.cell.2014.09.037 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

112. Leemhuis H., Packman L.C., Nightingale K.P. and Hollfelder F. (2008) The human histone acetyltransferase P/CAF is a promiscuous histone propionyltransferase . ChemBioChem 9 , 499–503 10.1002/cbic.200700556 [PubMed] [CrossRef] [Google Scholar]

113. Montgomery D.C., Sorum A.W. and Meier J.L. (2014) Chemoproteomic profiling of lysine acetyltransferases highlights an expanded landscape of catalytic acetylation . J. Am. Chem. Soc. 136 , 8669–8676 10.1021/ja502372j [PMC free article] [PubMed] [CrossRef] [Google Scholar]

114. Han Z., Wu H., Kim S., Yang X.K., Li Q.J., Huang H.. et al. (2018) Revealing the protein propionylation activity of the histone acetyltransferase MOF (males absent on the first) . J. Biol. Chem. 293 , 3410–3420 10.1074/jbc.RA117.000529 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

115. Feldman J.L., Baeza J. and Denu J.M. (2013) Activation of the protein deacetylase SIRT6 by long-chain fatty acids and widespread deacylation by mammalian sirtuins . J. Biol. Chem. 288 , 31350–31356 10.1074/jbc.C113.511261 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

116. Madsen A.S. and Olsen C.A. (2012) Profiling of substrates for zinc-dependent lysine deacylase enzymes: HDAC3 exhibits decrotonylase activity in vitro . Angew. Chem. Int. Ed. 51 , 9083–9087 10.1002/anie.201203754 [PubMed] [CrossRef] [Google Scholar]

117. Fellows R., Denizot J., Stellato C., Cuomo A., Jain P., Stoyanova E.. et al. (2018) Microbiota derived short chain fatty acids promote histone crotonylation in the colon through histone deacetylases . Nat. Commun. 9 , 15 10.1038/s41467-017-02651-5 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

118. Kelly R. D.W., Chandru A., Watson P.J., Song Y., Blades M., Robertson N.S.. et al. (2018) Histone deacetylase (HDAC) 1 and 2 complexes regulate both histone acetylation and crotonylation in vivo . Sci. Rep. 8 , 10 10.1038/s41598-018-32927-9 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

119. Bao X.C., Wang Y., Li X., Li X.M., Liu Z., Yang T.P.. et al. (2014) Identification of ‘erasers’ for lysine crotonylated histone marks using a chemical proteomics approach . Elife 3 , 18 10.7554/eLife.02999 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

120. Peng C., Lu Z.K., Xie Z.Y., Cheng Z.Y., Chen Y., Tan M.J.. et al. (2011) The first identification of lysine malonylation substrates and its regulatory enzyme . Mol. Cell. Proteomics 10 , 12 10.1074/mcp.M111.012658 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

121. Park J., Chen Y., Tishkoff D.X., Peng C., Tan M.J., Dai L.Z.. et al. (2013) SIRT5-mediated lysine desuccinylation impacts diverse metabolic pathways . Mol. Cell 50 , 919–930 10.1016/j.molcel.2013.06.001 [PMC free article] [PubMed] [CrossRef] [Google Scholar]

122. Huang H., Luo Z.Q., Qi S.K., Huang J., Xu P., Wang X.X.. et al. (2018) Landscape of the regulatory elements for lysine 2-hydroxyisobutyrylation pathway . Cell Res. 28 , 111–125 10.1038/cr.2017.149 [PMC free article] [PubMed] [CrossRef] [Google Scholar]