Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2012 Feb 6.
Published in final edited form as: Expert Rev Anticancer Ther. 2010 Jun;10(6):935–954. doi: 10.1586/era.10.62

Nonhistone protein acetylation as cancer therapy targets

Brahma N Singh 1,*, Guanghua Zhang 2,*, Yi L Hwa 3, Jinping Li 4, Sean C Dowdy 5, Shi-Wen Jiang 4,
PMCID: PMC3273412  NIHMSID: NIHMS215094  PMID: 20553216

Abstract

Acetylation and deacetylation are counteracting, post-translational modifications that affect a large number of histone and nonhistone proteins. The significance of histone acetylation in the modification of chromatin structure and dynamics, and thereby gene transcription regulation, has been well recognized. A steadily growing number of nonhistone proteins have been identified as acetylation targets and reversible lysine acetylation in these proteins plays an important role(s) in the regulation of mRNA stability, protein localization and degradation, and protein–protein and protein–DNA interactions. The recruitment of histone acetyltransferases (HATs) and histone deacetylases (HDACs) to the transcriptional machinery is a key element in the dynamic regulation of genes controlling cellular proliferation, differentiation and apoptosis. Many nonhistone proteins targeted by acetylation are the products of oncogenes or tumor-suppressor genes and are directly involved in tumorigenesis, tumor progression and metastasis. Aberrant activity of HDACs has been documented in several types of cancers and HDAC inhibitors (HDACi) have been employed for therapeutic purposes. Here we review the published literature in this field and provide updated information on the regulation and function of nonhistone protein acetylation. While concentrating on the molecular mechanism and pathways involved in the addition and removal of the acetyl moiety, therapeutic modalities of HDACi are also discussed.

Keywords: acetylation, cancer therapy, deacetylation, epigenetics, HDAC, HDAC inhibitor, histone deacetylase, nonhistone acetylation


Eukaryotic DNA, histones and histone-like proteins are assembled into chromatin, a highly organized, dynamic nucleoprotein complex that plays a significant role(s) in the regulation of cellular homeostasis [1,2]. The tails and the globular domains of nucleosomal histones can be modified by acetylation, phosphorylation, methylation, ubiquitination, sumoylation and, less commonly, by citrullination and ADP-ribosylation. These post-translational modifications can alter DNA–histone interactions or the binding of proteins, such as transcription factors, to chromatin [2,3]. Histone acetylations represent one of the best characterized post-translational modifications with profound functional implications for a wide range of cellular processes [4].

Acetylation levels of histone tails are maintained by the opposing, yet well balanced, activities of histone acetyltransferases (HATs) and histone deacetylases (HDACs). Histone hyperacetylation leads to a more open chromatin structure associated with active gene transcription [2,5]. This is due to reduced ionic interactions of the positively charged histone tails with the negatively charged DNA backbone and reduced internucleosomal interactions [2,6]. Remarkably, HDAC enzymes are not exclusively targeted towards histones. A steadily growing number of nonhistone proteins have been described to be subject to reversible acetylation by HATs and HDACs. Among these non-histone targets are transcription factors, hormone receptors, signal transducers, chaperones and proteins of the cytoskeleton [4]. Dynamic acetylation of nonhistone proteins modulates a wide variety of cellular events that are involved in many biological processes such as cell proliferation, cell survival and apoptosis [3,58]. Moreover, HATs and HDACs were found to be deregulated in cancer [4] and aberrant expression of HDACs has been observed in various tumor types [9,10]. It should be pointed out that a target protein may be acetylated by different HATs at different sites, which often exert divergent or opposite effects. For example, high-mobility group (HMG)-A proteins have been found to be integral components of enhanceo somes. Acetylation by p300/cAMP response element binding protein (CREB)-binding protein (CBP) leads to enhance osome disruption and hence, transcriptional repression [11], whereas acetylation of HMG-A by p300/CBP-associated factor (PCAF) results in enhanceosome assembly and transcription activation [12]. Therefore, an altered balance of protein acetylation appears to contribute to cell transformation and cancer development [13,14].

Acetylation of nonhistone proteins has been demonstrated to modulate protein functions by altering their stability, cellular localization and protein–nucleotide/protein–protein inter actions. Well-characterized targets of nonhistone acetylation include important cellular factors such as p53, nuclear factor-κB (NF-κB), p65, CBP, p300, STAT3, tubulin, PC4, GATA factors, nuclear receptors, c-Myc, hypoxia-inducible factor (HIF)-1α, FoxO1, heat-shock protein (Hsp)-90, HMG, E2F, MyoD, Bcr–Abl, the FLT3 kinase, c-Raf kinase and so on [4,14,15]. Interestingly, many nonhistone proteins targeted by acetylation are relevant for tumorigenesis, cancer cell proliferation and immune functions [1,2,14,16]. Therefore, re-expression of downregulated genes essential for growth arrest and cell death, as well as the alteration of aberrant acetylation patterns of nonhistone proteins, is considered a viable approach for cancer therapy [3,4].

While histone acetylation and its roles in transcription regulation have been covered by many excellent review articles, non-histone acetylation has not been thoroughly discussed. Here we provide an updated summary on the following topics concerning the regulation and function of nonhistone acetylation: the classification, expression and activity of HDAC enzymes; the well-characterized nonhistone acetylation targets and their impact on cell signaling, transcription and protein stability; HDAC involvement in cancer, apoptosis and cell cycle control; and HDAC inhibitors (HDACi) and their potential application in cancer treatment. Although this review focuses on nonhistone acetylation and individual acetylation targets are described separately, we should bear in mind that acetylation and deacetylation of histones as well as nonhistone proteins are regulated through an interconnected network and different epigenetic modifications actually function in a highly coordinated manner.

Acetylation & deacetylation

Nonhistone protein acetylation is a widespread phenomenon among eukaryotes. Co-translational Nα-terminal acetylation is one of the most frequent protein modifications, occurring on approximately 85% of eukaryotic proteins [3,10]. A less common, but perhaps more important, form of protein acetylation takes place post-translationally. Acetylation of several proteins is known to occur on the ε-amino group of the lysine residues. This type of acetylation of histones weakens the histone–DNA contact and lysine acetylation is highly reversible and tightly controlled by the two counteracting enzymatic activities: those of HAT and HDAC [3]. Factor acetyltransferases have been reported to acetylate transcriptional activators, coactivators, basal factors and nonhistone chromosomal proteins with remarkable substrate specificity similar to histone acetylation [7]. Early studies suggested that many lysine residues in histones are abundantly acetylated and that this alters the histone–histone interactions as well as histone–regulatory protein interactions [1,2,13,14]. These events lead to an open, more conducible chromatin structure, which facilitates the process of transcription [15]. In humans, there are 18 potential deacetylase enzymes that are capable of removing acetyl groups and maintaining the equilibrium of lysine acetylation in different proteins [4,16].

Classes & families of HDAC

Histone acetyltransferases exist in nearly all organisms and can be classified into four main classes and two families – the classical and silent information regulator 2 (Sir2)-related protein (sirtuin) families. This classification is based on sequence similarity to yeast deacetylases and their cofactor dependency. In humans, members of the classical family include HDAC-1, -2, -3 and -8 (class I), and HDAC-4, -5, -6, -7, -9 and -10 (class II) (Table 1) [2,17]. Class II HDACs share homology to yeast HDAC-1 and Hos3 and can be further subdivided into class IIa (HDAC-4, -5, -7 and -9) and class IIb HDACs (HDAC-6 and -10), which contain two catalytic sites. HDAC-11 shares sequence conservation with Rpd3 and Hda1 and is placed in class IV [18]. Expression of class II and IV HDACs is restricted to certain tissues, where they shuttle between the nucleus and cytoplasm. Class I, II and IV HDACs are Zn+-dependent, while the class III HDACs (SirT1–7), the homologs of the yeast SirT2 protein, require NAD+ for their activity [2].

Table 1.

Classification and characteristics of histone deacetylases.

Parameter Class I Class IIa Class IIb Class III and IV
Yeast HDAC RPD3 HDA1 HDA1 SirT2
Human HDAC HDAC-1–3, -8 HDAC-4, -5, -7, -9 HDAC-6, -10 SirT1–7, HDAC-11
Molecular
weight (kDa)
120–135 120–135 40–50 kDa
Size (no. of
amino acid
residues)
HDAC-1 (483), HDAC-2 (488),
HDAC-3 (428), HDAC-8 (377)
HDAC-4 (1084), HDAC-5 (1122),
HDAC-7 (855), HDAC-9 (1011)
HDAC-6 (1215),
HDAC-10 (669)
HDAC-11 (347)
Distribution Ubiquitous Brain, heart Testis, liver, kidney Unknown
Localization Nuclear Nuclear/
cytoplasmic
Mostly cytoplasmic Nuclear
Chromosomal
localization
HDAC-1 (1p34.1), HDAC-2
(6q21), HDAC-3 (5q31), HDAC-8
(Xq13)
HDAC-4 (2q37.2), HDAC-5
(17q21), HDAC-7 (12q13),
HDAC-9 (7p21-p15)
HDAC-6
(Xp11.22–23),
HDAC-10
(22q13.31-q13.33)
HDAC-11 (3p25.2)
Target
substrates
Histones, p53,
NF-κB
Histones Histones, tubulin,
Hsp
Histones, tubulin
p53, TAF
Catalytic sites One One HDAC-6, two and
HDAC-10, two
Two
Protein
complexes
NuRD, SIN3
Co-repressor N-CoR, SMRT N-CoR, SMRT
Interacting
proteins
Rb, p53, MyoD, NF-κB, SP11,
BRCA1, DNMT1, DNMT3A-B,
MBD2–3, MECP2, ATM
MEF2 Tubulin, Hsp p53
Co-factor Zn Zn Zn NAD+
Knockout
phenotype
HDAC-1, embryonic lethal,
increased histone acetylation,
increase in p21 and p27,
HDAC-2, cardiac defect
HDAC-4, defects in chondrocyte
differentiation
HDAC-5 and -9, cardiac defect
HDAC-7, maintenance of
vascular integrity, increase in
MMP10
Cellular
function
Participate in Sin3, NuRD,
Co-REST complex (HDAC-1, -2);
participate in SMRT/N-CoR
(HDAC-3)
Interaction with SMRT/N-CoR
and the co-repressors BcoR and
CtBP (HDAC-4)
Tubulin deacetylase
(HDAC-6)
Recruitment other
HDACs (HDAC-10)
Repression of p53 activity
Regulate the 3′-end
processing machinery of
mRNA
(SirT1 and SirT2)

CtBP: Carboxyl-terminal binding protein; HDAC: Histone deacetylase; Hsp: Heat-shock protein; MEF: Myocyte enhancer factor; N-CoR: Nuclear receptor co-repressor; NF-κB: Nuclear factor-κB; Rb: Retinoblastoma protein; SMRT: Silencing mediator for retinoid and thyroid hormone receptor.

While classical HDACs can be found in the cytosol and in the nucleus, no localization to mitochondria has been described so far [4,19]. Within the class I HDACs, HDAC-1, -2 and -8 are primarily found in the nucleus, whereas HDAC-3 is found in the nucleus, cytoplasm and cell membrane. Class II HDACs are able to shuttle in and out of the nucleus depending on different signals. HDAC-6 is a cytoplasmic, microtubule-associated enzyme.

Cellular functions & regulation of HDACs

Histone deacetylases often exist as components of multiprotein complexes such as the transcriptional co-repressors mSin3, nuclear receptor co-repressor (N-CoR) and silencing mediator for retinoid and thyroid hormone receptor (SMRT) [2022]. Using a library of fluorigenic tetrapeptide substrates, HDACs were ranked according to their substrate specificity: HDAC-8 > HDAC-1 > DAC-3 > HDAC-6 [21]. HDAC-1 and -2 are frequently found in complexes with Sir3, NuRD, N-CoR, mSirN3A, Ni-2/NRD and/or CoREST. These complexes are targeted to specific genomic regions by interaction with DNA-binding factors (Table 1). The following observations indicate diverse functions of the various HDACs [22]: different embryonic development stages have shown divergent HDAC-expression patterns, supporting a role for HDAC in embryogenesis and tissue differentiation; HDAC-3 modulates the functions of transcription factors such as TFII-1 and is critical for repression of multiple nuclear receptors (NRs); HDAC-1 interacts with MyoD and serves as a repressor for proliferating myoblasts; aberrant expression of HDAC-1 confers resistance to sodium butyrate-mediated apoptosis in melanoma cells through a p53-mediated pathway [23]. Johnson et al. observed that HDAC-1, -2 and -3 coimmunoprecipitated with the ATP-dependent chaperone protein Hsp-70 [24]. HDAC-6 deacetylates tubulin and modulates cell migration [25].

A series of HDAC-knockout models have been generated with various defects. Targeted disruption of HDAC-1 results in embryonic lethality and reduced proliferation despite increased expression of HDAC-2 and -3 [26]; HDAC-4-knockout mice display premature ossification due to the excessive proliferation of chondrocytes [27]. HDAC-4 acts as a repressor of chondrocyte hypertrophy by interacting with the myocyte-specific enhancer factor 2C transcription factor [28]. HDAC-5 has been demonstrated to interact with myocyte enhancer factor 2 (MEF2). MEF2 participates in diverse gene-regulatory programs during muscle- and neural-cell differentiation, cardiac morphogenesis, blood vessel formation and growth responsiveness. HDAC-5 knockout leads to inhibition of the target genes’ expression. These mice also develop cardiac hypertrophy. HDAC-6 plays a critical role in the cell response to misfolded protein stress. Knockdown of HDAC-6 induces Hsp-90 acetylation and reduction of the Hsp-90 chaperone activity [29]. HDAC-7 knockouts have defects in the maintenance of vascular integrity [30]. HDAC-7 inhibits the expression of the orphan nuclear receptor Nur77 and affects the apoptosis of T cells [31]. HDAC-9-knockout mice are sensitive to hypertrophic signals and develop cardiac hypertrophy with advanced age [32].

Cleavage factor Im (CFIm)-25, a component of mammalian CFIm, and poly(A) polymerase (PAP), a polyadenylating enzyme for the pre-mRNA, are acetylation targets. The residues acetyl ated in these proteins were mapped onto the regions required for interaction with each other. Whereas CBP acetylated these proteins, HDAC-1, HDAC-3, HDAC-10, SirT1 and SirT2 were involved in in vivo deacetylation. Shimazu and co-workers reported that HDACs regulate the 3′-end processing machinery and modulate the localization of PAP through the acetylation and deacetylation cycle [33]. These findings provide new insights into the role of protein acetylation in not only transcriptional initiation and elongation but also termination. There are still considerable gaps in our knowledge on the biological roles of the different HDACs.

Catalytic activities of HDACs are governed on multiple stages such as post-translational modifications, protein–protein interactions, availability of metabolic cofactors and subcellular localization (Table 2) [34]. Schuettengruber et al. reported that the murine HDAC-1 promoter is autoregulated by the HDAC inhibitor (HDACi) trichostatin A (TSA), which is involved in the SP1 binding sites and CCAAT box [35]. In the T cells, IL-2 is responsible for murine HDAC-1 induction. It has been reported that the HDAC-4 promoter is regulated by SP1/SP3 transcription factors [36]. Subcellular localization of HDAC-4, -5 and-7 is modulated by phosphorylation [36]. HDAC-4 phosphorylation and its nucleocytoplasmic shuttling depend upon CaMKIV [37]. Protein kinase D1 is important for phosphorylation of HDAC-7 and its nuclear export [37]. Myosin phosphatase dephosphorylates HDAC-7 and thus promotes its nuclear localization [38].

Table 2.

Nonhistone proteins known to be direct substrates for HDAC.

Substrate Category HDAC Function of acetylation Ref.
α-tubulin Cytoskeleton HDAC-6, SirT2 Modulates microtubule depolymerization [183] [14]
β-catenin Cell adhesion,
transcription regulator
HDAC-6 Regulates β-catenin function in a promoter-dependent manner [184]
TCF DNA-binding factor HDAC-3 Increases nuclear localization of high-mobility group (HMG) protein [185]
p53 Tumor suppressor HDAC-1, SirT2 Decreases DNA binding and transcriptional activity [51]
MyoD Muscle cell differentiation HDCA-1 Acetylation regulates transactivation and the conversion of naive
fibroblasts to muscle cells
[90]
E2F Transcription factor/cell cycle HDAC-1 Acetylation increases DNA-binding affinity [186]
Hsp90 Chaperone protein HDAC-6 Affects the maturation of glucocorticoid receptors [187]
HIF1α Transcription factor HDAC-7 Regulates degradation in the proteasome [123]
GATA Transcription factor HDAC-3, -4, -5 Affects erythroid differentiation [14]
YY1 Transcription factor HDAC-1, -2, -3 Acetylation promotes stable HDAC–YY1 interaction
Acetylation regulates YY1–DNA binding
[112]
NF-κB Transcription factor HDAC-3, SirT1 Promotes the p65–IκB interaction and inhibits
NF-κB-mediated transcription
[95]
Estrogen
receptor
Steroid hormone receptor HDAC-1 Increases cell proliferation [188]
pRb Tumor suppressor HDAC-1 Regulates the specific interaction with E2F-1 [189]

HDAC: Histone deacetylase; HIF1α: Hypoxia-inducible factor 1α; Hsp: Heat-shock protein; NF-κB: Nuclear factor-κB; TCF: T-cell factor; pRb: Retinoblastoma protein; YY1: Yin Yang 1.

Effects of nonhistone protein acetylation

Acetylation can affect a single factor in multiple ways. For example, acetylation of transcription factors can change their protein–protein interactions, protein turnover, protein localization and DNA-binding ability (Table 3) [39]. Also, depending on the protein and acetylation site, the same acetyl group may exert different, or even opposite, effects. For example, acetylation enhances sequence-specific DNA binding for nonhistone proteins such as p53, NF-κB, E2F, erythroid Kruppel-like factor (EKLF), p50 and PC4 [4044], whereas it reduces DNA binding of other factors such as FoxO1, HMGI (Y), p65 and so on [3,14,45]. The ability to activate or repress the DNA-binding ability often depends on the site of acetylation. If the acetylation sites lie in a DNA-binding domain, acetylation may repress the DNA-binding and if they are adjacent to a DNA-binding domain, it may activate DNA binding [3,14]. The following sections summarize the acetylation effects on several cell functions.

Table 3.

Biochemical and cellular activities regulated by acetylation.

Biological
implication
Nonhistone proteins affected by acetylation
Upregulation Downregulation
Protein stability p53, p73, Smad7, c-Myc, Runx3, AR, H2A.z, E2F1, NF-E4,
ER81, SREBP1a, HNF6, BACE1
GATA1, HIF-1α, pRb, SV40 T-Ag
DNA binding p53, SRY, STAT3, GATA transcription factors, E2F1, p50
(NF-κB), Erα, p65 (NF-κB), c-Myb, MyoD, HNF-4, AML1,
BETA2, NF-E2, KLF13, TAL1/SCL, TAF(I)68, AP endonuclease
YY1, HMG-A1, HMG-N2, p65 (NF-αB), DEK,
KLF13, Fen-1
Transcriptional
regulation
Transcriptional activation:
p53, HMG-A1, STAT3, AR, ERα (basal), GATA transcription
factors, EKLF, MyoD, E2F1, p65 (NF-αB), GR, p73, PGC1α,
MEF2D, GCMa, PLAG1, PLAGL2, Bcl-6, α-catenin, KLF5, Sp1,
BETA2, Cart1, RIP140, TAF(I)68
Transcriptional inactivation
ERα (ligand-bound), HIF-1α, STAT1, FoxO1,
FoxO4, RIP140
Protein–protein
interactions
STAT3, AR, EKLF, importin A, STAT1, TFIIB, α-tubulin,
actin, cortactin
p65(RelA), Ku70, heat-shock protein 90
DNA repair p53, E2F, TDF NAIL2, FEN1, Pol b
Protein localization PCAF, SRY, CtBP2, POP-1, HNF-4, PCNA subnuclear
WRN, PCNA
c-Abl, p300, PAP
mRNA stability p21, Brm Tyrosine hydrolase (Th), eNOS
Enzymatic activity p300, ATM PTEN, HDAC1, Mdm2, ACS, Neil2, Polα
Mitochondrial proteins ACS, Sod1/2 Profilin I, Thioredoxin
Viral proteins SV40 T-Ag E1A, HIV Tat, S-HDAg, L-HDAg

AR: Androgen receptor; EKLF: Erythroid Kruppel-like factor; ER: Estrogen receptor; GR: Glucocorticoid receptor; HMG: High-mobility group; NF-κB: Nuclear factor κB; SRY: Sex-determining region Y.

Signaling & transcription

p53

The p53 protein is a sequence-specific DNA-binding transcription factor known to maintain cellular homeostasis [46]. p53 can be acetylated by distinct acetyltransferases at multiple lysines: K120, K164, K320, K370, K372, K373, K381, K382 and K386 (Table 3) [47]. The resulting effects on p53 activity are still controversial [46]. Acetylation of K120 promoted by hMOF and Tip60 is able to mediate the expression of genes involved in DNA damage-induced apoptosis. Acetylation of K320 and poly ubiquitination of p53 apparently activate transcription. Additionally, p53 acetylated at K382 recruits CBP via its bromodomain to further stimulate transcription machinery, presumably binding through the increased DNA affinity to the target genes by acetylated p53 (Figure 1) [3]. Following the discovery of p53–HAT complexes, it was reported that p53 also interacts with HDAC-1, possibly through Sin3 or MTA2 proteins [48]. It was subsequently shown that HDAC-1 deacetylates p53 in vitro and in vivo [49]. Moreover, a pre acetylated p53 peptide was deacetylated by wild-type HDAC-1 but not by a deacetylase mutant. Furthermore, overexpression of HDAC-1 greatly reduced the in vivo acetylation level of p53. Finally, the activation potential of p53 on the BAX promoter, a natural p53-responsive system, is reduced in the presence of HDACs.

Figure 1. Acetylation regulates p53 function.

Figure 1

p53 loosely binds to the repressed, tightly packed chromatin. A cascade of events initiated with the phosphorylation of p53, followed by recruitment of acetyltransferases and methyltransferases takes place, leading to p53 acetylation. Acetylated p53 binds tightly to the acetylated and methylated chromatins. The overall effects are relaxation of chromatin and transcription activation. Deacetylation of p53 and nucleosomal histones by HDAC-1 or SirT2 reverses the process and leads to chromatin tightening and transcription repression.

Ac: Acetyl group; HAT: Histone acetyltransferase; HDAC: Histone deacetylase.

Interestingly, Mdm2 can promote p53 deacetylation by recruiting a separate complex containing HDAC-1 [50]. In addition to HDAC-1, a class III HDAC, SirT1, binds to and deacetylates p53 [51]. Studies indicate that SirT1 is involved in chromatin remodeling, gene silencing and the DNA-damage response. SirT1 can deacetylate nonhistone proteins, including various transcription factors involved in growth regulation, stress responses and endocrine signaling. SirT1-mediated deacetylation also suppresses DNA damage-induced, p53-dependent apoptosis [52]. In addition, Jin et al. reported that SirT2 interacts with 14–3-3 b/c proteins [53]; this interaction represents a novel negative regulatory mechanism for p53 besides the well-characterized Mdm2-mediated repression. Because lysine residues acetylated in p53 overlap with those that are ubiquitinated, acetylation serves to promote p53 stability. Unacetylated lysines are targets for ubiquitination catalyzed by Mdm2, which ultimately leads to the destruction of p53. Thus, p53 acetylation is critical for both the efficient recruitment of transcriptional complexes to promoter regions and the activation of p53 target genes in vivo [54]. On the other side, the deacetylation of p53 may provide a quick-acting mechanism to stop p53 function when the transcriptional activation of target genes is no longer needed.

The global transcriptional coactivator PC4 is known to be acetylated specifically by p300 in vitro and in vivo in humans [45]. Acetylation of PC4 enhances its DNA-binding ability. Phosphorylation of PC4 was found to negatively regulate the acetylation, presenting an intriguing case for the phosphorylation-mediated inhibition of acetylation [45]. In most cases, phosphorylation tends to exert positive effects on acetylation of the same proteins [55]. PC4 directly interacts with p53 in vitro and in vivo and the interaction promotes the sequence-specific DNA binding of p53 [56]. PC4 also induces the expression of p53-responsive genes and thereby enhances p53-dependent apoptosis. Since both p53 and PC4 are acetylated, it would be interesting to find out the role of reversible acetylation of these two proteins in the regulation of cellular homeostasis [57]. The deacetylation pathway of PC4 has not been understood, but initial data suggest that PC4 may be a target of HDAC-1 [57].

Han et al. reported that p300 interacts with SirT2 and acetylation of SirT2 by p300 relieves the inhibitory effect of SirT2 on the transcriptional activity of p53 [58]. These observations demon strate that p300 can inactivate SirT2 by acetylation and that p300 may regulate the activity of p53 indirectly through SirT2 in addition to its direct modification of p53. Deacetylation of p53 by SirT2 decreases the ability of p53 to transcriptionally activate the cell cycle inhibitor p21, which causes cells to re-enter the cell cycle following successful DNA repair.

Tubulin

Although acetylation of α-tubulin was found in mammalian cells two decades ago (Table 3), the acetyltransferase responsible remained unidentified for many years [59]. Stable microtubules contain high levels of acetyl ated α-tubulin. By contrast, dynamic microtubules, in which tubulin subunits are actively added or subtracted, such as at the leading edge of a migrating cell, are largely hypo-acetylated. HDAC-6 [60,61] and SirT2 [62] are able to deacetylate lysine 40 of α-tubulin and control the level of tubulin acetylation and assembly of the microtubule network. Deacetylation of α-tubulin by HDAC-6 and SirT2 enhances the stability of microtubules and consequently cell motility and other biological processes.

An interesting role of tubulin and tubulin-modifying deacetylases is their influence on the aggregation of misfolded proteins – the so-called aggresome formation, which serves to protect cells from stress and cell death [63]. HDAC-6 also deacetylates Hsp-90, pointing to a broader role of this enzyme in protein folding [64]. HDAC-6 contributes to the degradation of aggresomal proteins because it is able to bind to motor proteins (polyubiquitinated and dynein) as an adaptor protein to transport misfolded proteins along microtubules into aggresomes for lysosome-mediated degradation. Misfolded proteins can also be eliminated by the ubiquitin–proteasome pathway and the role of acetylation in this pathway is not clear. Small-molecule inhibitors of HDAC-6 inhibit the aggresome machinery and hence synergize with the proteasome inhibitor, which makes HDAC-6 an interesting target for new selective inhibitors. Inhibition of HDAC-6 results in a lower stability of the microtubules and consequently, increased cell stress and cell death. The inactivation of HDAC-6 reduces tumor formation in mice and HDAC-6-deficient cells are more resistant to oncogenic transformation [65,66]. Therefore, inhibition of HDAC-6 and SirT2 represents an interesting approach for the treatment of cancerous diseases.

Acetylated α-tubulin is present at the immune synapse, the junction between a T cell and antigen-presenting cell. The immune synapse also contains HDAC-6. By deacetylating tubulin, HDAC-6 aids in the remodeling of the synapse, which regulates the organization of adhesion and signaling molecules. The HDAC-6-mediated reorganization may attenuate the lymphocyte activation signaling cascade [67].

Nuclear receptor

Nuclear receptors (NRs) regulate gene expression by their association with HDAC complexes (Table 3). Hormones induce dramatic hyperacetylation at endogenous target genes by p300/CBP. This hyperacetylation is transient and coincides with the hormone-induced gene activation. The acetyltransferase, ACTR, can be acetylated by p300/CBP and PCAF [68]. The acetylation neutralizes the positive charges of the two lysine residues (629, 630) adjacent to the core LXLL motif of ACTR and disrupts the association of HAT–co-activator complexes with promoter-bound estrogen receptors. Thus, cofactor acetylation is a novel regulatory mechanism in hormonal signaling for transcription modulation.

The estrogen receptor (ER), peroxisome proliferator-activated receptor-γ (PPARγ) and androgen receptor (AR) are members of the NR superfamily. AR and ER are acetylated by HAT at a motif that is conserved across species and the NR family [3,8]. Acetylation of the NR appears to govern the ligand sensitivity and hormone antagonist responses. Mutagenesis of the six distinct AR-phosphorylation sites led to the identification of a single site capable of regulating HDAC responsiveness [69]. The gluco-corticoid receptor (GR) is a ligand-dependent transcriptional regulator that mediates a panoply of physiological, developmental and behavioral processes [70]. GR integrates, coordinates and responds to numerous cellular signals to accomplish its diverse functions. The GR is acetylated after ligand binding. Mutagenesis of the acetylated residues (K494/495) reduced GR acetylation and GR-mediated repression of NF-κB activity [71].

The short heterodimer partner (SHP) represses the actions of AR and ER. Unlike other NRs, SHP does not have a DNA-binding domain. Biological activities of SHP repression include competition with co-activators for the AF2-binding domains of NR, a dimerization partner [14]. HDAC-1, -3 and -6 each binds directly to SHP. Moreover, HDAC-1 and SHP can form complexes with both AR and ERα. However, it is not clear whether the HDAC recruited by SHP acts directly on the NR or on chromatin [72].

STAT proteins

Mammalian STAT proteins are a family of transcription factors consisting of seven members including STAT1–4, STAT5a, STAT5b and STAT6. It has been reported that the STAT proteins activate genes containing a γ-activating sequence (GAS) and interferon-stimulated response element (ISRE) in their promoters, thereby modulating cellular processes such as survival, apoptoapoptosis, cell proliferation and differentiation (Table 3) [3]. Outcomes of STAT-mediated reactions are tightly regulated by negative-feedback loops and tyrosine dephosphorylation [68]. STAT1 can be acetylated by CBP within its DNA-binding domain [5]. Overexpression of HDAC-1, -2 or -3 enhances STAT1-dependent gene expression upon cytokine stimulation. On the other hand, treatment with HDACi or specific siRNAs blocks the expression of interferon-responsive genes [73]. Acetylated STAT1 binds to the p65 subunit of NF-κB, which decreases p65 DNA binding and leads to downregulation of NF-κB target genes [5]. Tang reported that acetylated STAT2 acts as an adaptor for STAT1 and the STAT1/Ac-STAT2 heterodimer can interact with IFN-α receptor 2 or interferon-regulatory factor (IRF)-9 [46]. The expression of STAT1/STAT2-dependent genes is strongly suppressed upon HDAC inhibition [30,74]. In another study, Shankaranarayanan et al. correlated acetylation of STAT6 with the transcription of reticulocyte 15-lipoxygenase-1 [75]. p300/CBP-dependent acetylation of STAT3 may facilitate STAT3 dimerization, resulting in DNA binding and the transcriptional activation of STAT3 target genes cyclin D1, Bcl-xL and c-myc [75].

GATA factors

The transcriptional activator GATA1 plays an important role in hematopoietic cell differentiation. p300-mediated acetylation of GATA1 enhances its DNA-binding ability (Table 3). Since GATA1–DNA and Ac-GATA1–DNA complexes have differential mobility, presumably acetylation induces a conformational change of the protein [76]. The two highly conserved lysine-rich motifs within the central zinc finger domain of GATA1 are acetylated by p300/CBP. Mutation of these lysine-rich motifs in an erythroid cell lineage abrogated the cell differentiation, indicating that CBP-mediated acetylation of GATA1 is a key modification for its in vivo function.

GATA-2, another member of the zinc finger transcription factor family, expressed in hematopoietic stem cells and progenitors, is essential for cell survival. GATA-2 also exists as an acetylated protein in vivo. GATA-2 acetylation by p300 and GCN5 enhances its DNA-binding ability and transcriptional synergism with p300 [77]. GATA-2 directly associates with HDAC-3, but not HDAC-1 [78], and HDAC-3 can suppress the transcription potential of GATA-2. Deletion analyses showed that amino acids 270–393, encompassing two entire zinc fingers of GATA2, are required for binding to HDAC-3. Reciprocally, amino acids 132–180 of HDAC-3 are required for binding to GATA2. Both factors co-localized in the nucleus. GATA-2 also interacts with HDAC-5, but not with the other class II HDACs (HDAC-4 and -6). This indicates that a tissue-specific transcription factor can selectively interact with HDAC family members and such selectivity may define some tissue-specific function.

High-mobility group proteins

High-mobility group proteins bind preferentially to distorted DNA and provoke bending in linear DNA [79]. These proteins are divided into three subfamilies including HMG-A1/A2, HMG-B1/B2 and HMG-N1/N2 [80]. Among the HMG proteins, HMG-B1 and -B2 are the most extensively studied. Both the proteins are acetylated by CBP at lysine 2 [81], resulting in increased DNA-bending ability (Table 3). HMG-B1 is a chromatin component that, when leaked out by necrotic cells, triggers inflammation. In addition, HMG-B1 is also acetylated at lysine 11 in vivo [82]. HMG-B1 can be secreted by activated monocytes and macrophages and functions as a late mediator of inflammation. Bonaldi et al. showed that HMG-B1 shuttles actively between the nucleus and cytoplasm [83]. Monocytes and macrophages acetylate HMG-B1 extensively upon activation with lipopolysaccharide.

The main function of HMG-A1/A2 is the regulation of IFN-β production in response to viral infection upon acetylation by CBP and PCAF at lysines 65 and 71, respectively [84]. Acetylation by PCAF is connected with positive transcriptional regulation of the IFN-β gene promoter. The termination of the interferon response occurred following acetylation of lysine 65 by CBP owing to the decreased affinity of HMG-A1 for DNA [85]. An epigenetic alteration of HMG-N2 is related to tumorigenesis. In human adenocarcinoma HT29 cells, HDAC inhibitory butyrate induces HMG-N2 acetylation and downregulates HMG-N2 binding to chromatin, leading to modified gene expression [86]. Sex-determining region Y (SRY) is another HMG protein acetylated by p300. Acetylation increases SRY nuclear localization and DNA binding [86]. Conversely, deacetylation of SRY by HDAC-3 through the HMG box leads to a loss of nuclear localization [87].

E2F

Cell cycle regulator E2F (E2F-1, -2 and -3) forms a heterodimer with DP1 protein and is known to regulate S-phase-specific genes. PCAF acetylates one of the E2F family members, E2F1 at the DNA-binding domain, which increases the DNA-binding activity of the protein and thus E2F-mediated transcriptional activation (Table 3) [42]. Acetylation also stabilizes E2F1 protein. When complexed with retinoblastoma protein (Rb), E2F acts as a transcriptional repressor [88]. This interaction prevents E2F from getting acetylated. The Rb–HDAC complex deacetylates E2F1, indicating that HDACs and HATs act antagonistically on their non-histone substrates, as they do on histones. Interestingly, Rb itself is phosphorylated by G1-CDKs and acetylated by p300/CBP. While Rb acetylation does not affect its interaction with E2F, the modification increases its binding to MDM2. Thus, HAT- and HDAC-mediated Rb acetylation may define a new cell cycle control mechanism through protein–protein interactions [89].

MyoD

Myogenesis relies critically on the MyoD transcriptional factor. MyoD activity is regulated by acetylation at three evolutionarily conserved lysine residues (lysine 99, 102 and 104) [90]. Acetylation activates the transactivating ability of MyoD by inducing a conformational change that increases its affinity for DNA targets (Table 3). Although MyoD appears to interact with both the co-activators p300/CBP and PCAF, forming a multimeric protein complex associated with the promoter region, it can only be specifically acetylated by PCAF [91]. In myogenesis, MyoD binds the gene promoters, stimulating the expression of cyclin inhibitor p21 and certain muscle-specific genes. The acetylation of MyoD was found to enhance its DNA-binding activity and hence, transcription.

FoxO transcription factors

The mammalian family of FoxO proteins (FoxO1, 3, 4 and 6) belongs to the forkhead family of transcription factors. In the absence of insulin or growth factors, the FoxO proteins are located in the nucleus, triggering gene expression to regulate stress resistance, metabolism, cell cycle arrest and apoptosis. The p300/CBP- and PCAF-mediated acetylation diminishes the DNA-binding ability of FoxO proteins, in turn, reducing their activity [92]. In response to oxidative stress, SirT1 mediates deacetylation of FoxO1. SirT1-mediated FoxO1 deacetylation can increase lifespan and promote cellular survival [93]. Under conditions of caloric restriction, higher NAD+ levels could increase SirT1 activity towards FoxO1 and the resulting modulation of FoxO1 functions may contribute to SirT2-mediated lifespan extension.

NF-κB

To ensure a transient transcriptional response, cells have evolved mechanisms to regulate the proper termination of inducible transcription factor function. NF-κB, a heterodimer of p65 (RelA) and p50 proteins, is controlled by regulating its subcellular localization through its interaction with IκBα. Chen et al. showed that p300/CBP acetylates the RelA subunit at lysines 218, 221 and 310 [89]. Acetylation of K221 and K310 is required for the full activity of p65 [94]. SirT1-driven deacetylation of p65 K310 inhibits transcription of NF-κB target genes [95]. Likewise, HDAC-1 and HDAC-3 deacetylate p65 at either K221 or K310, resulting in the inhibition of NF-κB. Additionally, K122 and K123 acetylation reduces p65 DNA-binding affinity by increasing its IκBα interaction and nuclear export [96]. Recently, Buerki et al. identified a specific set of genes that are differentially regulated by TNF-α treatment when comparing wild-type and K314 and K315 mutant p65 using microarray ana lysis [97]. Acetylation of p65 thus regulates the NF-κB-dependent gene expression. Acetylation of p50 increases the NF-κB DNA binding towards target sequences, an action accompanied by increased p300 enrollment and activation of target-gene transcription [98].

Other nonhistone substrates

Many viral proteins are constitutive acetylation targets. Tat is a transactivator protein of HIV that plays an important role in HIV replication by binding to the trans-activator responsive region sequence of leader RNA [99]. Acetylation of Tat by p300 at K50 and K51 in its RNA-binding region decreases its affinity for the trans-activator responsive region sequences and leads to induction of transcription from the long terminal repeat by strengthening the elongation (Table 3) [100,101]. Adenoviral transforming protein E1A is acetylated by p300/CBP and PCAF at K239. Acetylated E1A can bind to the carboxyl-terminal binding protein (CtBP) and modulate global acetyltransferase activities, resulting in abnormal cell signaling and gene expression [102,103]. Moreover, acetylation impairs the capacity of E1A to bind importin-α3, resulting in cytosolic localization and thus, affects multiple cytoplasmic processes [104].

PTEN is an important phosphatase involved in cell signaling via phosphoinositols and the AKT/PI3 kinase pathway. Acetylation of PTEN by the HAT PCAF can stimulate its phosphatase activity; conversely, deacetylation of PTEN by SirT1 deacetylase and HDAC-1 can repress its activity [105,106].

Carboxyl-terminal binding protein participates in the regulation of cell differentiation, apoptosis, oncogenesis and development [107]. CtBP interacts with the p300 bromodomain and inhibits its transcriptional activity. It also interacts with nuclear hormone receptor co-repressor RIP140 and the acetylation of RIP140 inhibthe acetylation of RIP140 inhib the acetylation of RIP140 inhibits its interaction with CtBP [1]. Acetylation of adenoviral protein E1A inhibits its interaction with CtBP leading to alleviation of transcriptional repression mediated by CtBP [21]. The orphan NR SF-1 regulates the development and differentiation of steroidogenic tissues. Acetylation of SF1 by GCN5 regulates its transcriptional activity and stabilizes the protein (Table 3). Inhibition of deacetylation by TSA increases SF1-mediated transcriptional activation and nuclear export of SF1 protein [108].

Yin Yang 1 (YY1) is a nuclear protein with sequence-specific DNA-binding and transactivation/transrepression activities [109]. YY1 is able to interact with several HDACs and forms a multiprotein complex with co-activators and co-repressors [110]. The interaction between HDAC-1 and YY1 is phosphorylation dependent [111]. A discrete domain located within residues 261–333 of YY1 is necessary and sufficient for the recruitment of HDAC. Acetylation of 261–333 of YY1 suppresses its DNA-binding activity and hence, the transcriptional activity [112].

Erythroid Kruppel-like factor is an erythrocyte-specific transcription factor regulating normal hematopoiesis (Table 3). p300/CBP acetylates EKLF at residues 288 and 302, which reside in the transactivation and zinc finger domains, respectively [113], and activates adult β-globin gene expression. While acetylation does not affect the EKLF DNA-binding affinity, it enhances EKLF association with the SWI–SNF complex, resulting in an open chromatin structure at the β-globin promoter. Lysine 302 acetylation enables the interaction of EKLF with Sin3A and this complex represses transcription in a gene-specific manner [65,114].

mRNA stability

A HDACi-mediated decrease in endothelial nitric oxide synthase (eNOS) levels interferes with endothelial cell function [115]. The eNOS generates nitric oxide, a key second messenger in inflammatory diseases. Side effects of this free radical are cytotoxic through lipid, DNA and protein damage. The HDACi-mediated reduction of the half-life of eNOS mRNA is nevertheless sufficient to decrease eNOS protein levels. Furthermore, the HDACi TSA has been demonstrated to decrease the RNA stability of DNA methyltransferase-1 and -3B [116,117], which results in a significant reduction of de novo DNA methylation. HDACi can additionally decrease the expression of ERα [118] and tyrosine hydroxyl ase [119] by modulation of ERα mRNA stability. MicroRNAs (miRNAs) are noncoding RNAs that, through RNA interference, are involved in multiple cellular processes such as differentiation and apoptosis. Recent studies have demonstrated that deregulated miRNA expression contributes to the malignant phenotype. The mechanisms behind these effects on mRNA stability remain to be identified. Perhaps, dynamic protein acetylation affects mRNA turnover via RNase and/or mRNA stabilizing factors, which usually bind to the 3′-UTR of mRNA [119]. The HDACi LAQ-824 rapidly alters the levels of many miRNA species assessed. Interestingly, the study also observed that HDACi can modulate post-transcriptional processes [120]. The functional consequence of altered miRNA expression upon HDACi treatment is not yet understood in detail.

Protein stability

Acetylation regulates protein stability in a refined mode and by surprisingly diverse mechanisms [39]. Acetylation of lysines can block ubiquitination at the same residue, thereby preventing proteasomal degradation. This was first suggested for tumor-suppressor protein p53, which is tightly controlled by the Mdm2 E3 ligase-driven proteasomal degradation pathway. Acetylation abrogates complex formation between p53 and Mdm2, whereas an unacetyl atable p53 mutant strongly interacts with Mdm2, resulting in p53 degradation [46]. In an overexpression system, Mdm2 formed a HDAC-1-containing complex binding to p53. Recruitment of HDAC-1 might thereby link two enzymatic activities promoting p53 degradation [50,52,121]. It is a little surprising that positive regulation of p53 levels by HDACi has not been reported. This could be owing to the fact that the HDACi-insensitive SirT1 probably represents the major p53 deacetylase [3,50]. Accordingly, SirT inhibitor treatment leads to p53 hyperacetylation. The combined effect of HDACs and SirTs on p53 stability remains to be investigated.

Nonproteasomal and proteasomal degradation may also be facilit ated by acetylation. For example, acetylation of non-proteasomal protein HNF-6 increases its half-life, whereas a mutant incapable of being acetylated is degraded non-proteasomally [122]. Similarly, SV40 large T-Ag stability is controlled by CBP, HDAC-1, HDAC-3 and SirT1. In this case, HDACi-initiated signaling enhances a proteasome-independent degradation of T-Ag [33].

Acetylation of HIF-1α, an important angiogenesis regulator, at K532 by ARD1 was reported to induce its degradation [123]. HIF-1α is steadily ubiquitinated by the E3 ligase pVHL and degraded by the proteasome under normoxic conditions [1]. HIF-1α acetylation facilitates interaction with pVHL and its degradation [124]. It has been reported that the metastasis-associated protein MTA1 forms a complex with HDAC-1 and is able to bind, deacetylate and stabilize HIF-1α [13]. However, other studies link an HDACi-mediated HIF-1α decrease to pVHL- and proteasome-independent degradation [124] or to the function of class II HDACs [125], probably suggesting multiple pathways regulating HIF-1α stability under different conditions [8].

TGF-β signaling is mediated through the Smad family of transcription factors. He et al. provided evidence that Smad7 is a potent in vivo inhibitor for signal transduction of the TGFβ super-family during development and maintenance of homeostasis of multiple epithelial tissues [126]. Acetylation of Smad7 by p300 on lysines 64 and 70 leads to increased protein stability by preventing the ubiquitination of overlapping lysines by the ubiquitin ligase Smurf1 [127]. Overexpression of HDAC-1 significantly increases the amount of ubiquitinated Smad7, leading to an overall decrease in the protein half-life [128]. HDAC-1 and -3 bind strongly to Smad7, HDAC-2, -5 and -6 bind weakly and HDAC-4 does not bind at all. Overexpression of HDAC-1 significantly increases the amount of ubiquitinated Smad7, leading to an overall decrease in the protein half-life [128].

HDAC & cancer

HDAC expression in cancer cells

Histone deacetylases are associated with a number of well-character ized cellular oncogenes and tumor-suppressor genes, leading to aberrant recruitment of HDAC activity, altered gene expression and the development of specific forms of leukemia and lymphoma [129]. The oncoprotein that is encoded by one of the translocation-generated fusion genes in acute promyelocytic leukemia, PML–retinoic-acid receptor (RAR)-α, represses transcription by associating with a co-repressor complex that contains HDAC activity. In non-Hodgkin’s lymphoma, the transcriptional repressor lymphoma-associated zinc finger-3/B cell lymphoma (LAZ3/BCL6) is overexpressed and associated with aberrant transcriptional repression through recruitment of HDAC, leading to lymphoid oncogenic transform ation. Acute myeloid leukemia (AML)-m2 subtype is associated with the t(8;21) chromosomal translation, which produces an AML1–ETO fusion protein, a potent dominant transcription repressor acting through its recruitment of HDAC activity.

The most extensively studied model of differentiation in which HDACs play an important role is that of myoblast differentiation, which has also led to a greater understanding of how class II HDACs (e.g., HDAC-4 and HDAC-5) are regulated. Class II HDACs are localized in either the cytoplasm or nucleus. Phosphorylation of HDAC-4 at the N-terminus by CaMK leads to sequestration of HDAC-4 by 14–3-3 proteins and its active transport out of the nucleus. Inactive HDAC-4 is a target of the Ras signal-transduction cascade, whereby phosphorylation (presumably at different residues) leads to either translocation into the nucleus or release from sequestration [130]. The hallmarks of the malignant phenotype include the loss of differentiated status and a decreased reliance on exogenous growth factors. Mutations resulting in constitutive activation of signal-transduction pathways, such as the Ras pathway, are among the most frequent genetic changes in cancer cells [130,131]. Therefore, a constitutively active form of Ras could lead to the nuclear localization of HDACs and consequently, alteration of gene transcription. The link between altered HDAC activity and tumorigenesis is probably best demonstrated in acute promyelotic leukemia. The retinoic-acid receptor transcription factors RARα and its heterodimerization partner RXR bind to retinoic acid response elements (RAREs) and, in the absence of retinoids, repress transcription through a complex involving SIN3/HDAC, N-CoR and SMRT.

Increased HDAC-1 levels have been detected in gastric cancers, esophageal squamous-cell carcinoma and hormone-refractory prostate cancer [132,133]. Increased HDAC-2 expression was found in colon cancer. Increased levels of class II HDAC enzymes (HDAC-6) have been linked to better survival in breast cancer, but reduced expression of class II HDAC enzymes HDAC-5 and HDAC-10 has been associated with poor prognosis in lung cancer patients [67,134]. The potential role of the sirtuins in regulating AR function has been investigated. SirT1 inhibited cellular proliferation in AR-expressing prostate cancer cell lines but not in cells that do not express AR [70], demonstrating that their interaction is physiologically relevant. Inhibition of SirT1 with antagonists increased androgen-regulated gene transcription. Importantly, inhibitors of endogenous SirT1 (Sirtinol, splitomycin and nicotinamide) induced endogenous AR gene expression. The repression of AR activity by SirT1 required the catalytic function of SirT1. The AR lysine residues that are acetylated by p300 serve as substrates for SirT1-mediated deacetylation [135]. The ability of SirT1 to deacetylate the AR and repress its activity might provide a novel and effective modality for cancer therapy.

HDAC involvement in apoptosis

Acetylation of p53 at distinct sites regulates different cellular activities performed by p53. Acetylation at position 373 in p53 by p300/CBP leads to cell apoptosis whereas acetylation at residue 320 by PCAF leads to cell cycle arrest [136]. Under stressed conditions, acetylation levels of p53 increase, leading to its activation. Acetylation of p53 is also controlled at the deacetylation level by HDAC-1, HDCA-3 and SirT1. Inhibitors of SirT1 and HDAC-2 enhance p53 acetylation and thereby induce p53-mediated apoptosis [137,138]. The p53 homolog p73 is also acetylated and activated in response to DNA damage and potentiates apoptosis [139].

The DNA end-joining protein Ku70 prevents apoptosis by sequestering a pro-apoptotic protein, Bax, from mitochondria [140]. However, the acetylation of Ku70 disrupts its interaction with Bax and elevates Bax-mediated apoptosis. STAT1 has been known to repress NF-κB-mediated cell signaling. Acetylated STAT1 interacts with NF-κB, thereby preventing its DNA-binding ability, nuclear localization and expression of antiapoptotic genes [5]. These examples clearly indicate that acetylation of nonhistone proteins may induce or inhibit apoptosis depending upon the protein and physiological status. Detailed information from this area of research would facilitate the design of novel therapeutic strategies capable of activating apoptosis in malignant cells.

HDAC & cell cycle regulation

Crucial stages of the cell cycle are generally controlled through transcriptional regulation of a subset of genes, which are in turn regulated by the acetylation/deacetylation of histone and non-histone proteins. One notable example is the regulation of C-myc gene expression and cell cycle progression. C-myc regulates the expression of several genes involved in growth promotion by associating with its DNA-binding partner Max. p300 associates with C-myc and promotes C-myc stabilization, independent of p300-mediated acetylation, whereas C-myc acetylation increases its turnover [141]. Another cell cycle regulatory protein, cyclin D1, plays key regulatory roles during G1 phase and is overexpressed in many cancers. Cyclin D1 interacts with PCAF and facilitates the association of ER and PCAF. Overexpression of PCAF results in cyclin D1-dependent regulation of ER activity [142]. Interestingly, cyclin D1 expression is downregulated by the HDAC-1 complex recruited to its promoter by SMAR1, a matrix attachment region-binding protein [143]. Indeed, one well-recognized effect of HDACi is its potent action on cell cycle regulation.

Acetylation by small inhibitors of HDAC can affect the cell cycle by direct or indirect alteration of p21 and pRb. p21 is a potent cyclin-dependent kinase (CDK) inhibitor and its expression is controlled by p53. Under DNA damage conditions, p53 induces p21 expression, which results in growth arrest and cell apoptosis [144]. It has been well observed that HDACi induced p21 expression and that the induction was mediated by Sp1/Sp3 and ATM [145]. On the other hand, HuR proteins bind to the 3′-UTR and stabilize p21 mRNA, resulting in increased p21 expression [146]. HDACi butyrate increased p21 mRNA stability depending on de novo protein synthesis in HepG2 cells [147]. Moreover, pRb is able to interact with mSin3 and HDACs and regulate cyclin A and E expression, which evokes apoptosis [148]. The underlying mechanisms for the interplay among HDACs, HuR, p21 and pRB are not clearly understood.

HDAC inhibitors

Interestingly, the discovery of small-molecule HDACi preceded the discovery of HDACs. Sodium butyrate was the first compound identified to induce acetylation of histone [149]. Later, TSA, a fungal antibiotic [150], valproic acid (VA), already used in the treatment of epileptic diseases [151] and several other compounds were identified as HDACi. To date, several classes of HDACi have been proven to have potent and specific anticancer activities in preclinical studies [152154]. Natural or synthetic small molecular compounds have been used as biological switches for probing into the functional mechanism of HDACs [155]. These compounds can be divided into six groups based on their structure, including hydroxamic acids (e.g., TSA) and suberoylanilide hydroxamic acids (SAHAs); cyclic tetrapeptides (e.g., trapoxin, apicidin and HC-toxin); depsipeptides (e.g., FK228); short-chain fatty acids (e.g., butyrate and VA); synthetic pyridyl carbamate (e.g., MS-275); synthetic benzamide derivatives (e.g., tacedinaline) and ketones (e.g., trifluoromethyl ketone and α-ketomides) (Box 1). Most of these upregulate the expression of tumor suppressors (such as p53 and p21) and block cell cycle progression [156].

Box 1. Histone deacetylase inhibitors.

Hydroxamic acid-derived compounds

  • Trichostatin A

  • Suberoylanilide hydroxamic acid

  • M-carboxycinnamic acid bishydroxamide

  • Azelaic bis-hydroxamic acid

  • NVP-LAQ824

  • LBH589

  • Oxamflatin

  • PXD101

  • Scriptaid

  • Pyroxamide Cyclic tetrapeptides

  • Depsipeptide (FK228, FR901228)

  • Apicidine

  • Trapoxin

  • HC-toxin

  • Chlamydocin

  • Depudesin

  • CHAPS Short-chain fatty acids

  • Valproic acid

  • Phenyl butyrate

  • Phenyl acetate

  • Sodium butyrate

  • AN-9 (Pivanexs)

  • Synthetic pyridyl carbamate derivative

  • MS-275 Synthetic benzamide derivatives

  • CI-994 (N-acetyldinaline)

  • Tacedinaline Ketones

  • Trifluoromethyl ketone

  • α-ketomides

Since epigenetic changes critically contribute to cancer onset and progression, HDACi were soon considered promising anti-cancer drugs [157]. Indeed, at the cellular level, HDACi can induce differentiation, programmed cell death, cell cycle arrest, senescence, apoptosis, reactive oxygen species (ROS) production and cell death [2,158]. For example, in preclinical studies, SAHA has been found to induce cell death by producing ROS [159]. Furthermore, Rosato et al. observed that HDACi MS-275 exerts dose-dependent growth arrest and differentiation at low drug concentrations and a marked induction of ROS, mitochondrial damage, caspase activation and apoptosis at higher concentrations [160]. HDACi are thought to interact with the catalytic domain of HDACs to block substrate recognition, resulting in restoration of the expression of relevant genes that are silenced in malignancies [161]. HDACi were found to reduce tumor invasiveness, angiogenesis and metastasis (Figure 2). An additional promising effect of HDACi for cancer therapy is their selective toxicity against tumor cells compared with normal cells [6,162]. However, HDACi appear to be equally competent at promoting the acetylation of nonhistone proteins [14]. While there have been attempts to reveal aberrant gene-expression patterns in tumors, less information is available for differences in the acetylation patterns between normal and cancer cells and the effects of HDACi.

Figure 2. Histone deacetylase recruitment and transcription repression.

Figure 2

Histone octamers are represented by circles and the DNA is shown in black lines. HDAC is recruited to gene promoters by nonhistone proteins including methylated DNA-binding protein MeCP2, ER, methyl transferases (DNMT) and transcription factors (E2F, Rb). HDAC removes the acetyl group from nonhistone as well as histone proteins, which leads to chromatin conformational change and transcription repression of tumor-suppression genes. The loss of tumor-suppressor gene products contributes to the block of cell differentiation, uncontrolled cell proliferation, interrupted apoptosis and, ultimately, tumor formation.

ER: Estrogen receptor; HDAC: Histone deacetylase;

Rb: Retinoblastoma protein.

The fact that HATs and HDACs are deregulated in various cancers [69] suggests that anomalous acetylation occurs and may be corrected with HDACi treatment. Choudhary et al. used high-resolution mass spectrometry to identify 3600 lysine acetylation sites on 1750 proteins and quantified acetylation changes in response to the HDACi SAHA and MS-275 [163]. Acetylation of these sites preferentially targets large macro molecular complexes involved in diverse cellular processes, such as gene expression, RNA signaling, DNA-damage repair, cell cycle progression, cytoskeleton function, splicing, nuclear transport and actin nucleation, protein chaperones and ribosome formation. These findings suggest that the regulatory scope of lysine acetylation is broad and comparable with that of other major post-translational modifications.

Romidepsin and SAHA target many proteins whose structure and function are altered by acetylation, including chromatin-associated histones, nonhistone gene-transcription factors and proteins involved in regulation of cell proliferation, migration and death [164]. In clinical trials, both HDACi have shown significant anticancer activity against both hematologic and solid tumors at doses well tolerated by patients. SAHA and romidepsin have been approved by the US FDA for the treatment of CTCL. Other analogs of both drugs have shown unacceptable toxicity [164,165].

Histone deacetylase inhibitors appear to be highly effective in targeting endothelial cells. HDACi upregulate gene expression of p21WAF1/CIP1, which induces cell cycle arrest and down-regulates gene expression of survivin, an inhibitor of apoptosis, in proliferating endothelial cells [166]. TSA inhibits the VEGF-induced expression of VEGF receptors VEGFR1, VEGFR2 and neuropilin in a dose-dependent and reversible fashion. TSA and SAHA upregulate the expression of semaphorin III, a VEGF protein competitor, at both mRNA and protein levels [167]. The HDACi NVP-LAQ824 blocks mRNA and protein expression of the pro-angiogenic tyrosine kinase receptor Tie 2, as well as the Tie 2 ligand Ang-2, and results in reduced endothelial cell proliferation, tube formation and invasion into Matrigel plugs. In the mouse subcutaneous prostate and breast cancer models, a combination of NVP-LAQ824 with the VEGF receptor tyrosine kinase inhibitor PTK787 induced 80–85% inhibition of tumor growth without overt toxicity [166]. Butyrates upregulate cell adhesion molecules, including intercellular adhesion molecule-1 and E-selectin, on the surface of endothelial cells. Downregulation of eNOS in endothelial cells has also been shown to be critical for the antiangiogenic activity of the HDACi VA [168].

Isoform-specific inhibition of HDACs remains a challenging task [16]. TSA-induced acetylation of STAT3 was accompanied by enhanced nuclear localization of STAT3 [169]. Moreover, several genes under the control of STAT3 are indeed downregulated by HDACi and inhibition of this pro-oncogenic factor may contribute to the actions of HDACi [2]. Other pan-HDACi are SAHA, LAQ-824, MGCD0103 and LBH-589 [170]. One study indicated that tubacin can work as an HDAC-6-specific inhibitor [171]. VA has been demonstrated to be as a class I selective inhibitor [151], while MS-275 and depsipeptide are selective towards only a subset of class I HDACs [16]. In particular, HDAC-8 has been shown to play an important role in neuroblastoma pathogenesis [16]. Treatment with an undisclosed HDAC-8-selective small-molecule inhibitor induced differentiation of neuroblastoma cells in vitro [172]. Additional isoform-selective HDACi are under development. Such selective compounds will not only provide new investigative tools for molecular biology, but might also represent new candidate drugs for cancer treatment. Whether or not strictly isoform-specific HDACi would have additional therapeutic benefits is controversial [173].

Recently, Suzuki designed isoform-selective HDACi using the homology model of HDAC-6, one of the HDAC isoforms [174]. HDAC-6-insensitive inhibitor NCH-51, and HDAC-6-selective inhibitor NCT-10, were successfully identified as isoform-selective HDACi with potential therapeutic benefits. A novel butyrate-derived HDACi (OSU-HDAC-42) has recently been tested in the transgenic adenocarcinoma of the mouse prostate model. Interestingly, OSU-HDAC-42 achieved a remarkable suppression of prostate tumorigenesis, preventing the occurrence of macro scopic prostate tumors and poorly differentiated carcinomas [52,175]. An inhibition of HDAC-6 by such inhibiinhibitors results in a lower microtubule stability and consequently in increased cell stress and cell death.

Endometrial cancer is the most common malignancy of the female reproductive tract and treatment of advanced disease is a difficult challenge for clinicians. Treatment with single-agent paclitaxel results in median progression-free intervals of 7.3 months [66,176]. Dowdy et al. found that TSA and paclitaxel synergistically inhibited the proliferation of serous endometrial cancer cells [66]. Accompanying this effect was the dramatic activation of the apoptosis cascade. One of the most important mechanisms by which paclitaxel inhibits tumor growth is by microtubule stabilization. Importantly, they showed that, when added to paclitaxel, TSA causes a marked increase in micro tubule stabilization, indicating that these two agents work in a cooperative fashion in endometrial cancer cells. The effects of the TSA/paclitaxel combination on apoptosis and tubulin acetylation have subsequently been confirmed in a mouse xenograft model. There was a significant reduction in tumor weight in the mice treated with the drug combination compared with those treated with either agent alone. The ability of TSA to potentiate the anticancer effects achieved by paclitaxel may have important implications for the treatment of women with endometrial cancer and possibly for patients harboring other malignancies with limited sensitivity to paclitaxel. Chen et al. reported that the pretreatment of prostate cancer cells with HDACi (TSA, SAHA, MS-275 and OSU-HDAC-42) led to increased Ku70 acetylation and reduced Ku70 DNA-binding affinity without disrupting the Ku70–Ku80 heterodimer formation [177]. The impaired Ku70 function diminished the cellular capability to repair DNA double-strand breaks (DSBs) induced by bleomycin, doxorubicin and etoposide, thereby enhancing their cell-killing effect [177]. This sensitizing effect was most prominent when cells were treated with HDACi and DNA-damaging agents sequentially.

A recent review discussed the cancer chemopreventive properties of three reported dietary HDACi [178], namely butyrate, diallyl disulfide and sulforaphane (SFN). In general, these dietary agents are weak ligands and must be present in much higher concentrations than TSA or SAHA in order to inhibit HDAC activity. The diallyl disulfide compound, derived from garlic [179], was reported to inhibit HDAC activity following metabolic conversion to S-allylmercaptocysteine (SFN-Cys) [180]. SFN is found in a wide variety of cruciferous vegetables including broccoli, cabbages, watercress and Brussels sprouts. SFN and phenylhexyl isothiocyanate are among the synthetic isothiocyanates that have been shown to be HDACi and have antitumor activities in vitro and in vivo [169,181], whereas the SFN parent compound had no effect on HDAC unless it was incubated with cells to allow for metabolic conversion [182]. Interestingly, isothiocyanate was recently found to have dual epigenetic effects as both a HDACi and a demethylating agent [181].

Expert commentary

The growing number of nonhistone acetylation targets has opened new avenues for research. The regulation and function of acetylation in nonhistone compared with histone proteins are still understudied and many aspects remain to be investigated. To date the number of identified acetylated nonhistone proteins may be far below the actual number that exist in the in vivo acetylome. Switches between acetylation and alternative post-translational modifications at the same lysine residue appear to play a critical role. The functional consequences of acetylation can be as variable as their targets. It is difficult to predict the effect of acetylation of proteins at multiple sites without experimental testing. Animal as well as human data must be interpreted in the context of the counter balancing actions of HDAC and ACT. Acetylation appears to often occur in protein complexes containing multiple subunits and cofactors. The resulting substrate specificity and enzymatic activity are often difficult to reconstitute in vitro. Despite these technical challenges, research on acetylation of nonhistone substrates such as transcription factors and cellular and viral proteins has drawn broad attention. The dynamic nature and the diversified function of acetylation/deacetylation of these proteins have been increasingly appreciated. Specifically, more research efforts should be directed to elucidate the mechanisms by which acetylation is regulated and to determine how nonhistone acetylation affects mRNA stability and the localization, interaction, degradation and function of proteins.

Protein acetylation and gene expression regulate protein stability, which can influence cellular signaling, cell proliferation and apoptosis. Consequently, HDACs have been recognized as potential targets for developing effective anticancer therapeutics. HDACi have shown the potential for amelioration of disease states by promoting cell cycle arrest and apoptosis through acetylation. The first generation of HDACi is FDA approved and has been demonstrated to have a low toxicity profile in comparison with traditional cytotoxic agents. New HDACi, now in Phase I and II clinical trials, appear to be even better tolerated. Acetylation of histone and nonhistone proteins can interfere with other post-transcriptional modifications, such as ubiquitination, and prevent or induce the proteasomal degradation of (proto-) oncoproteins, such as HIF-1α and c-Myc. Moreover, HDACi-induced abrogation of chaperone function correlates with enhanced degradation of client proteins, such as FLT3, Bcr–Abl, mutated p53 or ERα. Analyzing how HDACi influence the fine-tuned abundance of enzymes of the proteasomal pathway and induce the turnover of cancer-relevant proteins, such as AML1–ETO, PML–RARα, HDAC-2, or cyclin D1, provides further insights into the actions of these compounds. Given the pleiotropic effects HDACi have on multiple pathways, such data will aid in understanding how HDACi specifically affect different cell types. Moreover, data from in vitro experiments, animal models and clinical studies on HDACi are often explained solely based on their effects on histones and transcription control. In the future, studies will fill this gap of knowledge and the effects of nonhistone acetylation will be taken into account in the development of new HDACi.

Five-year view

Acetylation of nonhistone proteins regulates a wide variety of cellular events. In addition, it is highly likely that many of these reactions require formation of a complex by multiple proteins. Despite the wide array of possible outcomes, the fast technical advancement has made it possible to dissect the individual effects of the acetylation status of different sites through empirical testing. More information will become available through the establishment of mRNA/protein expression and signaling databases. Given the deregulation of HDAC and extensive involvement of aberrant protein acetylation in tumorigenesis, epigenetic research will become an ever increasing presence in the field of cancer research and therapeutics. This will facilitate comparison between normal and diseased subjects, not just in cancer research, but in the study of medical disorders as well.

Technical innovation will lead to further streamlining and/or automation of studies on protein acetylation. Recombinant proteins, new antibodies and molecular labeling techniques, such as mass spectrometry, can be applied for in vitro experiments as well as in vivo using engineered animal models. Although acetylation lysine-specific antibodies have been used for the collection of evidence of acetylation, such antibodies may not be as specific for acetylation as advertised. Undesirable false-positive or -negative results may result from the nonspecificity of these antibodies. This will make it necessary to use other means, such as in vivo labeling or mass spectrometry, to confirm the antibody-based findings. These developments will significantly speed up the process of determining the impact of nonhistone proteins on cell signaling, transcription, mRNA/protein stability and the cell cycle and apoptosis.

It is particularly exciting that, owing to its dynamic nature, protein acetylation is a reversible process. Epigenetic inhibitors could possibly correct aberrant acetylation patterns and ameliorate the diseased state. The growing number of identified acetylation target proteins beyond chromatin will provide a large pool of potential targets for therapeutic interference. Development of drugs with specificity to subclasses of tumor-selective transferases will also occur in the near future. More research will uncover important differences in splice variants, which will also be exploited by drug development. Preliminary data support synergy between these inhibitors and traditional cytotoxic agents. Future drug regimens will include combinations of these agents with improved efficacy and lower toxicity.

Key issues.

  • Nonhistone proteins undergo aberrant acetylation that may modulate a wide variety of cellular events that are involved in critical biological processes such as gene expression, cell signaling, mRNA stability, protein folding, cytoskeleton assembly, enzymatic activity, protein localization, protein stability and protein–protein interactions.

  • Hundreds of nonhistone proteins have been identified as acetylation targets. Effects of acetylation/deacetylation on their capacity for DNA binding, protein–protein interaction and cellular signaling have been characterized. 2D gel electrophoresis and LC/MS-MS ana lysis have been used to identify acetylated proteins.

  • p53 acetylation favors DNA binding and transcriptional activation of tumor-suppressor genes. Similar mechanisms have been suggested for other transcription factors, GATA factors, MyoD, E2F1 and many other proteins.

  • The recruitment of histone acetyltransferases (HATs) and histone deacetylases (HDACs) to the transcriptional machinery is a key element in the dynamic regulation of genes controlling cellular proliferation and differentiation. Histone acetyltransferase and HDAC often modify protein acetylation status in multiprotein complexes.

  • Gene silencing can be reversed using small-molecule inhibitors of HDAC, some of which are either US FDA approved or currently in Phase I and II clinical trials. Several tested HDAC inhibitors (HDACi) have shown tolerable side effect profiles.

  • HDACi exhibit potent inhibitory and stimulatory effects on the cell cycle and cell apoptosis, respectively. Preclinical studies suggest that HDACi may synergize with other cytogenic agents to alter chromatin structure, gene-expression patterns and tubulin acetylation and inhibit tumor growth in animal models.

Acknowledgments

Shi-Wen Jiang is the Distinguished Cancer Scholar supported by the Georgia Cancer Coalition (GCC). He also serves as a Mayo Collaborative Researcher and receives traveling support from the Department of Obstetrics and Gynecology, Mayo Clinic. This work is partially funded by a grant from the National Institute of Health (NIH R01 HD 41577, Shi-Wen Jiang), NIH/NCI MD Anderson Uterine Cancer SPORE (Jinping Li, Shi-Wen Jiang), Gynecologic Cancer Foundation (GCF; Sean C Dowdy, Shi-Wen Jiang), the Fraternal Order of Eagles Cancer Research Award (Sean C Dowdy, Shi-Wen Jiang), Merck Pharmaceutical Research Funding for Vorinostat Study (Sean C Dowdy, Shi-Wen Jiang), the NIH Clinical Investigator Loan Repayment Program (Sean C Dowdy, Shi-Wen Jiang), the NIH K12 training program (Sean C Dowdy, Shi-Wen Jiang), the Seeding Grant from Mercer University (Jinping Li) and research supplements from the Mercer University School of Medicine (Shi-Wen Jiang).

Footnotes

Financial & competing interests disclosure The authors have no other relevant affiliations or financial involvements with any organization or entity with a financial interest in or financial conflict with the subject manner or materials discussed in the manuscript apart from those disclosed.

No writing assistance was utilized in the production of this manuscript.

References

Papers of special note have been highlighted as:

• of interest

•• of considerable interest

  • 1.Jiang SW, Li J, Podratz K, Dowdy S. Application of DNA methylation biomarkers for endometrial cancer management. Expert Rev. Mol. Diagn. 2008;8(5):607–616. doi: 10.1586/14737159.8.5.607. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Buchwald M, Kramer OH, Heinzel T. HDACi – targets beyond chromatin. Cancer Lett. 2009;280(2):160–167. doi: 10.1016/j.canlet.2009.02.028. •• Describes histone deacetylase inhibitor (HDACi)-mediated depletion of oncoproteins via modulation of the abundance of enzymes of the ubiquitin–proteasome pathway. It highlights the cytokine- and HDACi-stimulated acetylation-dependent cross-talks of STAT/NF-κB signaling pathways. It also describes how the reagents could be beneficial for the treatment and prevention of human ailments, such as cancer and unbalanced immune functions.
  • 3.Spange S, Wagner T, Heinzel T, Krämer OH. Acetylation of non-histone proteins modulates cellular signalling at multiple levels. Int. J. Biochem. Cell Biol. 2009;41(1):185–198. doi: 10.1016/j.biocel.2008.08.027. •• Detailed review of the post-translational acetylation of nonhistone proteins as well as the complex and dynamic nature of the cellular acetylome. Acetylation alterations are frequently involved in physiological and pathophysiological pathways.
  • 4.Yang XJ, Seto E. Lysine acetylation: codified crosstalk with other posttranslational modifications. Mol. Cell. 2008;31(4):449–461. doi: 10.1016/j.molcel.2008.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Krämer OH, Baus D, Knauer SK, et al. Acetylation of Stat1 modulates NF-κB activity. Genes Dev. 2006;20(4):473–485. doi: 10.1101/gad.364306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Xu WS, Parmigiani RB, Marks PA. Histone deacetylase inhibitors: molecular mechanisms of action. Oncogene. 2007;26(37):5541–5552. doi: 10.1038/sj.onc.1210620. [DOI] [PubMed] [Google Scholar]
  • 7.Botrugno OA, Santoro F, Minucci S. Histone deacetylase inhibitors as a new weapon in the arsenal of differentiation therapies of cancer. Cancer Lett. 2009;280(2):134–144. doi: 10.1016/j.canlet.2009.02.027. [DOI] [PubMed] [Google Scholar]
  • 8.Bilton R, Trottier E, Pouyssegur J, Brahimi-Horn MC. ARDent about acetylation and deacetylation in hypoxia signalling. Trends Cell Biol. 2006;16(12):616–621. doi: 10.1016/j.tcb.2006.10.002. [DOI] [PubMed] [Google Scholar]
  • 9.Sakuma T, Uzawa K, Onda T, et al. Aberrant expression of histone deacetylase 6 in oral squamous cell carcinoma. Int. J. Oncol. 2006;29(1):117–124. [PubMed] [Google Scholar]
  • 10.Polevoda B, Sherman F. Nα -terminal acetylation of eukaryotic proteins. J. Biol. Chem. 2000;275(47):36479–36482. doi: 10.1074/jbc.R000023200. [DOI] [PubMed] [Google Scholar]
  • 11.Bergel M, Herrera JE, Thatcher BJ, et al. Acetylation of novel sites in the nucleosomal binding domain of chromosomal protein HMG-14 by p300 alters its interaction with nucleosomes. J. Biol. Chem. 2000;275:11514–11520. doi: 10.1074/jbc.275.15.11514. [DOI] [PubMed] [Google Scholar]
  • 12.Herrera J, Sakaguchi K, Bergel M, Trieschmann L, Nakatani Y, Bustin M. Specific acetylation of chromosomal protein HMG-17 by P/CAF alters its interaction with nucleosomes. Mol. Cell Biol. 1999;19:3466–3473. doi: 10.1128/mcb.19.5.3466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Allfrey VG, Faulkner R, Mirsky AE. Acetylation and methylation of histones and their possible role in the regulation of RNA synthesis. Proc. Natl Acad. Sci. USA. 1964;51:786–794. doi: 10.1073/pnas.51.5.786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Glozak MA, Sengupta N, Zhang X, Seto E. Acetylation and deacetylation of non-histone proteins. Gene. 2005;363:15–23. doi: 10.1016/j.gene.2005.09.010. •• Broad review covering the nonhistone targets of histone acetyltransferases (HATs) and HDACs and the consequences of their modifications. The authors note that the acetylation of transcription factors, cellular proteins and viral proteins may be closely associated with tumorigenesis.
  • 15.Yoo CB, Jones PA. Epigenetic therapy of cancer: past, present and future. Nat. Rev. Drug Discov. 2006;5(1):37–50. doi: 10.1038/nrd1930. [DOI] [PubMed] [Google Scholar]
  • 16.Hahn CK, Ross KN, Warrington IM, et al. Expression-based screening identifies the combination of histone deacetylase inhibitors and retinoids for neuroblastoma differentiation. Proc. Natl Acad. Sci. USA. 2008;105(28):9751–9756. doi: 10.1073/pnas.0710413105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Gregoretti IV, Lee YM, Goodson HV. Molecular evolution of the histone deacetylase family: functional implications of phylogenetic analysis. J. Mol. Biol. 2004;338(1):17–31. doi: 10.1016/j.jmb.2004.02.006. [DOI] [PubMed] [Google Scholar]
  • 18.Yamamoto H, Schoonjans K, Auwerx J. Sirtuin functions in health and disease. Mol. Endocrinol. 2007;21(8):1745–1755. doi: 10.1210/me.2007-0079. [DOI] [PubMed] [Google Scholar]
  • 19.Yang XJ, Gregoire S. Class II histone deacetylases: from sequence to function, regulation, and clinical implication. Mol. Cell Biol. 2005;25(8):2873–2884. doi: 10.1128/MCB.25.8.2873-2884.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Glass CK, Rosenfeld MG. The coregulator exchange in transcriptional functions of nuclear receptors. Genes Dev. 2000;14(2):121–141. [PubMed] [Google Scholar]
  • 21.de Ruijter AJ, van Gennip AH, Caron HN, Kemp S, van Kuilenburg AB. Histone deacetylases (HDACs): characterization of the classical HDAC family. Biochem. J. 2003;370(3):737–749. doi: 10.1042/BJ20021321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Marks PA. Discovery and development of SAHA as an anticancer agent. Oncogene. 2007;26(9):1351–1356. doi: 10.1038/sj.onc.1210204. [DOI] [PubMed] [Google Scholar]
  • 23.Bandyopadhyay D, Mishra A, Medrano EE. Overexpression of histone deacetylase 1 confers resistance to sodium butyrate-mediated apoptosis in melanoma cells through a p53-mediated pathway. Cancer Res. 2004;64(21):7706–7710. doi: 10.1158/0008-5472.CAN-03-3897. [DOI] [PubMed] [Google Scholar]
  • 24.Johnson CA, White DA, Lavender JS, O’Neill LP, Turner BM. Human class I histone deacetylase complexes show enhanced catalytic activity in the presence of ATP and co-immunoprecipitate with the ATP-dependent chaperone protein Hsp70. J. Biol. Chem. 2002;277(11):9590–9597. doi: 10.1074/jbc.M107942200. [DOI] [PubMed] [Google Scholar]
  • 25.Tran AD, Marmo TP, Salam AA, et al. HDAC6 deacetylation of tubulin modulates dynamics of cellular adhesions. J. Cell Sci. 2007;120(8):1469–1479. doi: 10.1242/jcs.03431. [DOI] [PubMed] [Google Scholar]
  • 26.Lagger G, O’Carroll D, Rembold M, et al. Essential function of histone deacetylase 1 in proliferation control and CDK inhibitor repression. EMBO J. 2002;21(11):2672–2681. doi: 10.1093/emboj/21.11.2672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Vega RB, Matsuda JK, Oh AC, et al. Histone deacetylase 4 controls chondrocyte hypertrophy during skeletogenesis. Cell. 2004;119(4):555–566. doi: 10.1016/j.cell.2004.10.024. [DOI] [PubMed] [Google Scholar]
  • 28.Arnold MA, Kim Y, Czubryt MP, et al. MEF2C transcription factor controls chondrocyte hypertrophy and bone development. Dev. Cell. 2007;12(3):377–389. doi: 10.1016/j.devcel.2007.02.004. [DOI] [PubMed] [Google Scholar]
  • 29.Scroggins BT, Robzyk K, Wang D, et al. An acetylation site in the middle domain of Hsp90 regulates chaperone function. Mol. Cell. 2007;25(1):151–159. doi: 10.1016/j.molcel.2006.12.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Chang S, Young BD, Li S, Qi X, Richardson JA, Olson EN. Histone deacetylase 7 maintains vascular integrity by repressing matrix metalloproteinase 10. Cell. 2006;126(2):321–334. doi: 10.1016/j.cell.2006.05.040. [DOI] [PubMed] [Google Scholar]
  • 31.Dequiedt F, et al. HDAC7, a thymus-specific class II histone deacetylase, regulates Nur77 transcription and TCR-mediated apoptosis. Immunity. 2003;18(5):687–698. doi: 10.1016/s1074-7613(03)00109-2. [DOI] [PubMed] [Google Scholar]
  • 32.Zhang Q, Vo N, Goodman RH. Acetylation of adenovirus E1A regulates binding of the transcriptional corepressor CtBP. Proc. Natl Acad. Sci. USA. 2000;97(26):14323–14328. doi: 10.1073/pnas.011283598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Shimazu T, Komatsu Y, Nakayama KI, Fukazawa H, Horinouchi S, Yoshida M. Regulation of SV40 large T-antigen stability by reversible acetylation. Oncogene. 2006;25(56):7391–7400. doi: 10.1038/sj.onc.1209731. [DOI] [PubMed] [Google Scholar]
  • 34.Sengupta N, Seto E. Regulation of histone deacetylase activities. J. Cell Biochem. 2004;93(1):57–67. doi: 10.1002/jcb.20179. [DOI] [PubMed] [Google Scholar]
  • 35.Schuettengruber B, Simboeck E, Khier H, Seiser C. Autoregulation of mouse histone deacetylase 1 expression. Mol. Cell Biol. 2003;23(19):6993–7004. doi: 10.1128/MCB.23.19.6993-7004.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Grozinger CM, Schreiber SL. Regulation of histone deacetylase 4 and 5 and transcriptional activity by 14–13-3-dependent cellular localization. Proc. Natl Acad. Sci. USA. 2000;97(14):7835–7840. doi: 10.1073/pnas.140199597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Zhao X, Ito A, Kane CD, et al. The modular nature of histone deacetylase HDAC4 confers phosphorylation-dependent intracellular trafficking. J. Biol. Chem. 2001;276(37):35042–35048. doi: 10.1074/jbc.M105086200. [DOI] [PubMed] [Google Scholar]
  • 38.Parra M, Mahmoudi T, Verdinn E. Myosin phosphatase dephosphorylates HDAC7, controls its nucleocytoplasmic shuttling, and inhibits apoptosis in thymocytes. Genes Dev. 2007;21(6):638–643. doi: 10.1101/gad.1513107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Sadoul K, Boyault C, Pabion M, Khochbin S. Regulation of protein turnover by acetyltransferases and deacetylases. Biochimie. 2008;90(2):306–312. doi: 10.1016/j.biochi.2007.06.009. [DOI] [PubMed] [Google Scholar]
  • 40.Bode AM, Dong Z. Post-translational modification of p53 in tumorigenesis. Nat. Rev. Cancer. 2004;4(10):793–805. doi: 10.1038/nrc1455. [DOI] [PubMed] [Google Scholar]
  • 41.Martínez-Balbás MA, Bauer UM, Nielsen SJ, Brehm A, Kouzarides T. Regulation of E2F1 activity by acetylation. EMBO J. 2000;19(4):662–671. doi: 10.1093/emboj/19.4.662. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Chen X, Bieker JJ. Stage-specific repression by the EKLF transcriptional activator. Mol. Cell Biol. 2004;24(23):10416–10424. doi: 10.1128/MCB.24.23.10416-10424.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Chen LF, Greene WC. Shaping the nuclear action of NF-κB. Nat. Rev. Mol. Cell Biol. 2004;5(5):392–401. doi: 10.1038/nrm1368. [DOI] [PubMed] [Google Scholar]
  • 44.Kumar BRP, Swaminathan V, Banerjee S, Kundu TK. p300-mediated acetylation of human transcriptional coactivator PC4 is inhibited by phosphorylation. J. Biol. Chem. 2001;276(20):16804–16809. doi: 10.1074/jbc.M100934200. [DOI] [PubMed] [Google Scholar]
  • 45.Matsuzaki H, Daitoku H, Hatta M, Aoyama H, Yoshimochi K, Fukamizu A. Acetylation of FoxO1 alters its DNA-binding ability and sensitivity to phosphorylation. Proc. Natl Acad. Sci. USA. 2005;102(32):11278–11283. doi: 10.1073/pnas.0502738102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Zhao Y, Lu S, Wu L, et al. Acetylation of p53 at lysine 373/382 by the histone deacetylase inhibitor depsipeptide induces expression of p21(Waf1/Cip1) Mol. Cell Biol. 2006;26(7):2782–2790. doi: 10.1128/MCB.26.7.2782-2790.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Tang Y, Zhao W, Chen Y, Zhao Y, Gu W. Acetylation is indispensable for p53 activation. Cell. 2008;133(4):612–626. doi: 10.1016/j.cell.2008.03.025. • Highlights the p53 activation by acetylation in various malignancies, including gastric, ovarian, colon, oral, breast and pancreatic cancers. Deacetylation of p53 takes place in separate pathways involving HDAC-1 or SirT2, leading to transcriptional repression.
  • 48.Murphy M, Ahn J, Walker KK, et al. Transcriptional repression by wild-type p53 utilizes histone deacetylases, mediated by interaction with mSin3a. Genes Dev. 1999;13(19):2490–2501. doi: 10.1101/gad.13.19.2490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Luo J, Su F, Chen D, Shiloh A, Gu W. Deacetylation of p53 modulates its effect on cell growth and apoptosis. Nature. 2000;408(6810):377–381. doi: 10.1038/35042612. [DOI] [PubMed] [Google Scholar]
  • 50.Ito A, Kawaguchi Y, Lai CH, et al. MDM2-HDAC1-mediated deacetylation of p53 is required for its degradation. EMBO J. 2002;21(22):6236–6245. doi: 10.1093/emboj/cdf616. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Luo J, Nikolaev AY, Imai S, et al. Negative control of p53 by Sir2a promotes cell survival under stress. Cell. 2001;107:137–148. doi: 10.1016/s0092-8674(01)00524-4. [DOI] [PubMed] [Google Scholar]
  • 52.Vaquero A, Lee MD Scher, Erdjument-Bromage H, Tempst P, Reinberg D. Human SirT1 interacts with histone H1 and promotes formation of facultative heterochromatin. Mol. Cell. 2004;16(1):93–105. doi: 10.1016/j.molcel.2004.08.031. [DOI] [PubMed] [Google Scholar]
  • 53.Jin YH, Kim YJ, Kim DW, Baek KH, Kang BY, Yeo CY. SirT2 interacts with 14–13-3 β/γ and down-regulates the activity of p53. Biochem. Biophys. Res. Commun. 2008;368(3):690–695. doi: 10.1016/j.bbrc.2008.01.114. [DOI] [PubMed] [Google Scholar]
  • 54.Barlev NA, Liu L, Chehab NH, et al. Acetylation of p53 activates transcription through recruitment of coactivators/ histone acetyltransferases. Mol. Cell. 2001;8(6):1243–1254. doi: 10.1016/s1097-2765(01)00414-2. [DOI] [PubMed] [Google Scholar]
  • 55.Warnock LJ, Raines SA, Mee TR, Milner J. Role of phosphorylation in p53 acetylation and PAb421 epitope recognition in baculoviral and mammalian expressed proteins. FEBS J. 2005;272(7):1669–1675. doi: 10.1111/j.1742-4658.2005.04589.x. [DOI] [PubMed] [Google Scholar]
  • 56.Banerjee S, Kumar BR, Kundu TK. General transcriptional coactivator PC4 activates p53 function. Mol. Cell Biol. 2004;24(5):2052–2062. doi: 10.1128/MCB.24.5.2052-2062.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Batta K, Das C, Gadad S, Shandilya J, Kundu TK. Reversible acetylation of non histone proteins: role in cellular function and disease. Subcell. Biochem. 2007;4:193–212. [PubMed] [Google Scholar]
  • 58.Han Y, Jin YH, Kim YJ, et al. Acetylation of SirT2 by p300 attenuates its deacetylase activity. Biochem. Biophys. Res. Commun. 2008;375(4):576–580. doi: 10.1016/j.bbrc.2008.08.042. • Direct cross-talk between SirT2 and p300 was investigated. Experimental evidence supports a model in which the deacetylase activity of Sir2 is negatively regulated through acetylation by p300. The authors point out that the relief of the repression of p53 transcriptional activity by SirT2 could be used by the cells to efficiently regulate gene transcription.
  • 59.Schulze E, Asai DJ, Bulinski JC, Kirschner M. Posttranslational modification and microtubule stability. J. Cell Biol. 1987;105(5):2167–2177. doi: 10.1083/jcb.105.5.2167. • Mechanistic investigation on how acetylation and deacetylation modification affects the assembly of the microtubule network.
  • 60.Hubbert C, Guardiola A, Shao R, et al. HDAC6 is a microtubule-associated deacetylase. Nature. 2002;417:455–458. doi: 10.1038/417455a. [DOI] [PubMed] [Google Scholar]
  • 61.Matsuyama A, Shimazu T, Sumida Y, et al. In vivo destabilization of dynamic microtubules by HDAC6-mediated deacetylation. EMBO J. 2002;21:6820–6831. doi: 10.1093/emboj/cdf682. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.North BJ, Marshall BL, Borra MT, Denu JM, Verdin E. The human Sir2 ortholog, SirT2, is an NAD+-dependent tubulin deacetylase. Mol. Cell. 2003;11(2):437–444. doi: 10.1016/s1097-2765(03)00038-8. [DOI] [PubMed] [Google Scholar]
  • 63.Boyault C, Zhang Y, Fritah S, et al. HDAC6 controls major cell response pathways to cytotoxic accumulation of protein aggregates. Genes Dev. 2007;21(17):2172–2181. doi: 10.1101/gad.436407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Marrocco DL, Tilley WD, Bianco-Miotto T, et al. Suberoylanilide hydroxamic acid (vorinostat) represses androgen receptor expression and acts synergistically with an androgen receptor antagonist to inhibit prostate cancer cell proliferation. Mol. Cancer Ther. 2007;6(1):51–60. doi: 10.1158/1535-7163.MCT-06-0144. [DOI] [PubMed] [Google Scholar]
  • 65.Marrocco DL, Tilley WD, Bianco-Miotto T, et al. Histone deacetylase inhibitors and paclitaxel cause synergistic effects on apoptosis and microtubule stabilization in papillary serous endometrial cancer cells. Mol. Cancer Ther. 2006;5(11):2767–2776. doi: 10.1158/1535-7163.MCT-06-0209. [DOI] [PubMed] [Google Scholar]
  • 66.Lee YS, Lim KH, Guo X, et al. The cytoplasmic deacetylase HDAC6 is required for efficient oncogenic tumorigenesis. Cancer Res. 2008;68(18):7561–7569. doi: 10.1158/0008-5472.CAN-08-0188. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Serrador JM, Cabrero JR, Sancho D, Mittelbrunn M, Urzainqui A, Sanchez-Madrid F. HDAC6 deacetylase activity links the tubulin cytoskeleton with immune synapse organization. Immunity. 2004;20(4):417–428. doi: 10.1016/s1074-7613(04)00078-0. [DOI] [PubMed] [Google Scholar]
  • 68.Das C, Kundu TK. Transcriptional regulation by the acetylation of nonhistone proteins in humans – a new target for therapeutics. IUBMB Life. 2005;57(3):137–149. doi: 10.1080/15216540500090629. •• Comprehensive review on acetylation of nonhistone proteins and their roles in the regulation of transcription. Authors also illustrate the pathways through which dysfunction of HATs and HDACs may lead to several diseases. Therapeutic application and clinical significance of inhibitors of these enzymes are discussed in detail.
  • 69.Fu M, Rao M, Wu K, et al. The androgen receptor acetylation site regulates cAMP and AKT but not ERK-induced activity. J. Biol. Chem. 2004;279(28):29436–29449. doi: 10.1074/jbc.M313466200. [DOI] [PubMed] [Google Scholar]
  • 70.Tronche F, Kellendonk C, Reichardt HM, Schutz G. Genetic dissection of glucocorticoid receptor function in mice. Curr. Opin. Genet. Dev. 1998;8(5):532–538. doi: 10.1016/s0959-437x(98)80007-5. [DOI] [PubMed] [Google Scholar]
  • 71.Lin HY, Hopkins R, Cao HJ, et al. Acetylation of nuclear hormone receptor superfamily members: thyroid hormone causes acetylation of its own receptor by a mitogen-activated protein kinase-dependent mechanism. Steroids. 2005;70(5–7):444–449. doi: 10.1016/j.steroids.2005.03.001. [DOI] [PubMed] [Google Scholar]
  • 72.Gobinet J, Carascossa S, Cavailles V, Vignon F, Nicolas JC, Jalaguier S. SHP represses transcriptional activity via recruitment of histone deacetylases. Biochemistry. 2005;44(16):6312–6320. doi: 10.1021/bi047308d. [DOI] [PubMed] [Google Scholar]
  • 73.Klampfer L, Huang J, Swaby LA, Augenlicht L. Requirement of histone deacetylase activity for signaling by STAT1. J. Biol. Chem. 2004;279(29):30358–30368. doi: 10.1074/jbc.M401359200. [DOI] [PubMed] [Google Scholar]
  • 74.Chang HM, Paulson M, Holko M, et al. Induction of interferon-stimulated gene expression and antiviral responses require protein deacetylase activity. Proc. Natl Acad. Sci. USA. 2004;101(26):9578–9583. doi: 10.1073/pnas.0400567101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Shankaranarayanan P, Chaitidis P, Kühn H, et al. Acetylation by histone acetyltransferase CREB-binding protein/ p300 of STAT6 is required for transcriptional activation of the 15-lipoxygenase-1 gene. J. Biol. Chem. 2001;276(46):42753–42760. doi: 10.1074/jbc.M102626200. [DOI] [PubMed] [Google Scholar]
  • 76.Yuan ZL, Guan YJ, Chatterjee D, Chin YE. STAT3 dimerization regulated by reversible acetylation of a single lysine residue. Science. 2005;307(5707):269–273. doi: 10.1126/science.1105166. [DOI] [PubMed] [Google Scholar]
  • 77.Hayakawa F, Towatari M, Ozawa Y, Tomita A, Privalsky ML, Saito H. Functional regulation of GATA-2 by acetylation. J. Leukoc. Biol. 2004;75(3):529–540. doi: 10.1189/jlb.0603389. [DOI] [PubMed] [Google Scholar]
  • 78.Osawa M, Takayanagi S, Kato Y, et al. Histone deacetylase 3 associates with and represses the transcription factor GATA-2. Blood. 2001;98(7):2116–2123. doi: 10.1182/blood.v98.7.2116. [DOI] [PubMed] [Google Scholar]
  • 79.Thomas JO, Travers AA. HMG1 and 2, and related ‘architectural’ DNA-binding proteins. Trends Biochem. Sci. 2001;26(3):167–174. doi: 10.1016/s0968-0004(01)01801-1. [DOI] [PubMed] [Google Scholar]
  • 80.Bustin M, Reeves R. High-mobility-group chromosomal proteins: architectural components that facilitate chromatin function. Prog. Nucleic Acid Res. Mol. Biol. USA. 1996;54:35–100. doi: 10.1016/s0079-6603(08)60360-8. [DOI] [PubMed] [Google Scholar]
  • 81.Pasheva E, Sarov M, Bidjekov K, et al. In vitro acetylation of HMGB-1 and -2 proteins by CBP: the role of the acidic tail. Biochemistry. 2004;43(10):2935–2940. doi: 10.1021/bi035615y. [DOI] [PubMed] [Google Scholar]
  • 82.Sterner R, Vidali G, Allfrey VG. Studies of acetylation and deacetylation in high mobility group proteins. Identification of the sites of acetylation in HMG-1. J. Biol. Chem. 1979;254(22):11577–11583. [PubMed] [Google Scholar]
  • 83.Bonaldi T, Fabio, Scaffidi P, et al. Monocytic cells hyperacetylate chromatin protein HMGB1 to redirect it towards secretion. EMBO J. 2003;22(20):5551–5560. doi: 10.1093/emboj/cdg516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Munshi N, Agalioti T, Lomvardas S, Merika M, Chen G, Thanos D. Coordination of a transcriptional switch by HMGI(Y) acetylation. Science. 2001;293(5532):1133–1136. doi: 10.1126/science.293.5532.1133. [DOI] [PubMed] [Google Scholar]
  • 85.Munshi N, Merika M, Yie J, Senger K, Chen G, Thanos D. Acetylation of HMG I(Y) by CBP turns off IFN β expression by disrupting the enhanceosome. Mol. Cell. 1998;2(4):457–467. doi: 10.1016/s1097-2765(00)80145-8. [DOI] [PubMed] [Google Scholar]
  • 86.Luhrs H, Hock R, Schauber J, et al. Modulation of HMG-N2 binding to chromatin by butyrate-induced acetylation in human colon adenocarcinoma cells. Int. J. Cancer. 2002;97(5):567–573. doi: 10.1002/ijc.10098. [DOI] [PubMed] [Google Scholar]
  • 87.Thevenet L, Mejean C, Moniot B, et al. Regulation of human SRY subcellular distribution by its acetylation/deacetylation. EMBO J. 2004;23(16):3336–3345. doi: 10.1038/sj.emboj.7600352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Weintraub SJ, Chow KN, Luo RX, Zhang SH, He S, Dean DC. Mechanism of active transcriptional repression by the retinoblastoma protein. Nature. 1995;375(6534):812–815. doi: 10.1038/375812a0. [DOI] [PubMed] [Google Scholar]
  • 89.Chan HM, Krstic-Demonacos M, Smith L, Demonacos C, La Thangue NB. Acetylation control of the retinoblastoma tumour-suppressor protein. Nat. Cell Biol. 2001;3(7):667–674. doi: 10.1038/35083062. [DOI] [PubMed] [Google Scholar]
  • 90.Polesskaya A, Duquet A, Naguibneva I, et al. CREB-binding protein/p300 activates MyoD by acetylation. J. Biol. Chem. 2000;275(44):34359–34364. doi: 10.1074/jbc.M003815200. [DOI] [PubMed] [Google Scholar]
  • 91.Sartorelli V, Puri PL, Hamamori Y, et al. Acetylation of MyoD directed by PCAF is necessary for the execution of the muscle program. Mol. Cell. 1999;4(5):725–734. doi: 10.1016/s1097-2765(00)80383-4. [DOI] [PubMed] [Google Scholar]
  • 92.Calnan DR, Brunet A. The FoxO code. Oncogene. 2008;27(16):2276–2288. doi: 10.1038/onc.2008.21. [DOI] [PubMed] [Google Scholar]
  • 93.van der Horst A, Tertoolen LG, de Vries-Smits LM, Frye RA, Medema RH, Burgering BM. FoxO4 is acetylated upon peroxide stress and deacetylated by the longevity protein hSir2(SIRT1) J. Biol. Chem. 2004;279(28):28873–28879. doi: 10.1074/jbc.M401138200. [DOI] [PubMed] [Google Scholar]
  • 94.Chen LF, Mu Y, Greene WC. Acetylation of RelA at discrete sites regulates distinct nuclear functions of NF-κB. EMBO J. 2002;21(23):6539–6548. doi: 10.1093/emboj/cdf660. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Yeung F, Hoberg JE, Ramsey CS, et al. Modulation of NF-κB-dependent transcription and cell survival by the SIRT1 deacetylase. EMBO J. 2004;23(12):2369–2380. doi: 10.1038/sj.emboj.7600244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Kiernan R, Bres V, Ng RW, et al. Post-activation turn-off of NF-κB-dependent transcription is regulated by acetylation of p65. J. Biol. Chem. 2003;278(4):2758–2766. doi: 10.1074/jbc.M209572200. [DOI] [PubMed] [Google Scholar]
  • 97.Buerki C, Rothgiesser KM, Valovka T, et al. Functional relevance of novel p300-mediated lysine 314 and 315 acetylation of RelA/p65. Nucleic Acids Res. 2008;36(5):1665–1680. doi: 10.1093/nar/gkn003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Deng WG, Zhu Y, Wu KK. Up-regulation of p300 binding and p50 acetylation in tumor necrosis factor-α-induced cyclooxygenase-2 promoter activation. J. Biol. Chem. 2003;278(7):4770–4777. doi: 10.1074/jbc.M209286200. [DOI] [PubMed] [Google Scholar]
  • 99.Berkhout B, Klaver B, Das AT. Forced evolution of a regulatory RNA helix in the HIV-1 genome. Nucleic Acids Res. 1997;25(5):940–947. doi: 10.1093/nar/25.5.940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Ott M, Schnolzer M, Garnica J, et al. Acetylation of the HIV-1 Tat protein by p300 is important for its transcriptional activity. Curr. Biol. 1999;9(24):1489–1492. doi: 10.1016/s0960-9822(00)80120-7. [DOI] [PubMed] [Google Scholar]
  • 101.Kiernan RE, Vanhulle C, Schiltz L, et al. HIV-1 tat transcriptional activity is regulated by acetylation. EMBO J. 1999;18(21):6106–6118. doi: 10.1093/emboj/18.21.6106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Deng Q, Li Y, Tedesco D, et al. The ability of E1A to rescue ras-induced premature senescence and confer transformation relies on inactivation of both p300/CBP and Rb family proteins. Cancer Res. 2005;65(18):8298–8307. doi: 10.1158/0008-5472.CAN-05-0054. [DOI] [PubMed] [Google Scholar]
  • 103.Molloy D, Mapp KL, Webster R, Gallimore PH, Grand RJ. Acetylation at a lysine residue adjacent to the CtBP binding motif within adenovirus 12 E1A causes structural disruption and limited reduction of CtBP binding. Virology. 2006;355(2):115–126. doi: 10.1016/j.virol.2006.05.004. [DOI] [PubMed] [Google Scholar]
  • 104.Madison DL, Yaciuk P, Kwok RP, Lundblad JR. Acetylation of the adenovirus-transforming protein E1A determines nuclear localization by disrupting association with importin-α. J. Biol. Chem. 2002;277(41):38755–38763. doi: 10.1074/jbc.M207512200. •• Describes post-translational acetylation of multiple transcription factors and other cellular proteins. These proteins play a significant role in regulating the chromatin conformation, transcription dynamics and, thereby, gene transcription regulation. Strategies for designing a new generation of drugs specifically targeting acetylated proteins and their potential clinical application are discussed.
  • 105.Ikenoue T, Inoki K, Zhao B, Guan KL. PTEN acetylation modulates its interaction with PDZ domain. Cancer Res. 2008;68(17):6908–6912. doi: 10.1158/0008-5472.CAN-08-1107. [DOI] [PubMed] [Google Scholar]
  • 106.Yao XH, Nyomba BL. Hepatic insulin resistance induced by prenatal alcohol exposure is associated with reduced PTEN and TRB3 acetylation in adult rat offspring. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2008;294(6):R1797–R1806. doi: 10.1152/ajpregu.00804.2007. [DOI] [PubMed] [Google Scholar]
  • 107.Corda D, Colanzi A, Luini A. The multiple activities of CtBP/BARS proteins: the Golgi view. Trends Cell Biol. 2006;16(3):167–173. doi: 10.1016/j.tcb.2006.01.007. [DOI] [PubMed] [Google Scholar]
  • 108.Jacob AL, Lund J, Martinez P, Hedin L. Acetylation of steroidogenic factor 1 protein regulates its transcriptional activity and recruits the coactivator GCN5. J. Biol. Chem. 2001;276(40):37659–37664. doi: 10.1074/jbc.M104427200. [DOI] [PubMed] [Google Scholar]
  • 109.Thomas MJ, Seto E. Unlocking the mechanisms of transcription factor YY1: are chromatin modifying enzymes the key? Gene. 1999;236(2):197–208. doi: 10.1016/s0378-1119(99)00261-9. [DOI] [PubMed] [Google Scholar]
  • 110.Yang WM, Yao YL, Sun JM, Davie JR, Seto E. Isolation and characterization of cDNAs corresponding to an additional member of the human histone deacetylase gene family. J. Biol. Chem. 1997;272(44):28001–28007. doi: 10.1074/jbc.272.44.28001. [DOI] [PubMed] [Google Scholar]
  • 111.Galasinski SC, Resing KA, Goodrich JA, Ahn NG. Phosphatase inhibition leads to histone deacetylases 1 and 2 phosphorylation and disruption of corepressor interactions. J. Biol. Chem. 2002;277(22):19618–19626. doi: 10.1074/jbc.M201174200. [DOI] [PubMed] [Google Scholar]
  • 112.Yao YL, Yang WM, Seto E. Regulation of transcription factor YY1 by acetylation and deacetylation. Mol. Cell Biol. 2001;21(17):5979–5991. doi: 10.1128/MCB.21.17.5979-5991.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Zhang W, Kadam S, Emerson BM, Bieker JJ. Site-specific acetylation by p300 or CREB binding protein regulates erythroid Kruppel-like factor transcriptional activity via its interaction with the SWI–SNF complex. Mol. Cell Biol. 2001;21(7):2413–2422. doi: 10.1128/MCB.21.7.2413-2422.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Chen X, Bieker JJ. Unanticipated repression function linked to erythroid Kruppel-like factor. Mol. Cell Biol. 2001;21(9):3118–3125. doi: 10.1128/MCB.21.9.3118-3125.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Rössig L, Li H, Fisslthaler B, et al. Inhibitors of histone deacetylation downregulate the expression of endothelial nitric oxide synthase and compromise endothelial cell function in vasorelaxation and angiogenesis. Circ. Res. 2002;91(9):837–844. doi: 10.1161/01.res.0000037983.07158.b1. [DOI] [PubMed] [Google Scholar]
  • 116.Januchowski R, Dabrowski M, Ofori H, Jagodzinski PP. Trichostatin A downregulate DNA methyltransferase 1 in Jurkat T cells. Cancer Lett. 2007;246(1–2):313–317. doi: 10.1016/j.canlet.2006.03.010. [DOI] [PubMed] [Google Scholar]
  • 117.Xiong Y, Dowdy SC, Podratz KC, et al. Histone deacetylase inhibitors decrease DNA methyltransferase-3B messenger RNA stability and down-regulate de novo DNA methyltransferase activity in human endometrial cells. Cancer Res. 2005;65(7):2684–2689. doi: 10.1158/0008-5472.CAN-04-2843. [DOI] [PubMed] [Google Scholar]
  • 118.De los Santos M, Martinez-Iglesias O, Aranda A. Anti-estrogenic actions of histone deacetylase inhibitors in MCF-7 breast cancer cells. Endocr. Relat. Cancer. 2007;14(4):1021–1028. doi: 10.1677/ERC-07-0144. [DOI] [PubMed] [Google Scholar]
  • 119.Arányi T, Sarkis C, Berrard S, et al. Sodium butyrate modifies the stabilizing complexes of tyrosine hydroxylase mRNA. Biochem. Biophys. Res. Commun. 2007;359(1):15–19. doi: 10.1016/j.bbrc.2007.05.025. [DOI] [PubMed] [Google Scholar]
  • 120.Rausa FM, III, Hughes DE, Costa RH. Stability of the hepatocyte nuclear factor 6 transcription factor requires acetylation by the CREB-binding protein coactivator. J. Biol. Chem. 2004;279(41):43070–43076. doi: 10.1074/jbc.M407472200. [DOI] [PubMed] [Google Scholar]
  • 121.Juan LJ, Shia WJ, Chen MH, et al. Histone deacetylases specifically down-regulate p53-dependent gene activation. J. Biol. Chem. 2000;275(27):20436–20443. doi: 10.1074/jbc.M000202200. [DOI] [PubMed] [Google Scholar]
  • 122.Sheng W, Yan H, Rausa FM, Costa RH, Liao X. Structure of the hepatocyte nuclear factor 6α and its interaction with DNA. J. Biol. Chem. 2004;279(32):33928–33936. doi: 10.1074/jbc.M403805200. [DOI] [PubMed] [Google Scholar]
  • 123.Jeong JW, Bae MK, Ahn MY, et al. Regulation and destabilization of HIF-1α by ARD1-mediated acetylation. Cell. 2002;111(5):709–720. doi: 10.1016/s0092-8674(02)01085-1. [DOI] [PubMed] [Google Scholar]
  • 124.Kong X, Lin Z, Liang D, Fath D, Sang N, Caro J. Histone deacetylase inhibitors induce VHL and ubiquitin-independent proteasomal degradation of hypoxia-inducible factor 1α. Mol. Cell Biol. 2006;26(6):2019–2028. doi: 10.1128/MCB.26.6.2019-2028.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Qian DZ, Kachhap SK, Collis SJ, et al. Class II histone deacetylases are associated with VHL-independent regulation of hypoxia-inducible factor 1 α. Cancer Res. 2006;66(17):8814–8821. doi: 10.1158/0008-5472.CAN-05-4598. [DOI] [PubMed] [Google Scholar]
  • 126.He W, Li AG, Wang D, Han S, Zheng B, Goumans M-J. Overexpression of Smad7 results in severe pathological alterations in multiple epithelial tissues. EMBO J. 2002;21(11):2580–2590. doi: 10.1093/emboj/21.11.2580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Grönroos E, Hellman U, Heldin CH, Ericsson J. Control of Smad7 stability by competition between acetylation and ubiquitination. Mol. Cell. 2002;10(3):483–493. doi: 10.1016/s1097-2765(02)00639-1. [DOI] [PubMed] [Google Scholar]
  • 128.Simonsson M, Heldin CH, Ericsson J, Gronroos E. The balance between acetylation and deacetylation controls Smad7 stability. J. Biol. Chem. 2005;280(23):21797–21803. doi: 10.1074/jbc.M503134200. [DOI] [PubMed] [Google Scholar]
  • 129.Tang X, Gao JS, Guan YJ, et al. Acetylation-dependent signal transduction for type I interferon receptor. Cell. 2007;131:93–105. doi: 10.1016/j.cell.2007.07.034. [DOI] [PubMed] [Google Scholar]
  • 130.Wade PA. Transcriptional control at regulatory checkpoints by histone deacetylases: molecular connections between cancer and chromatin. Hum. Mol. Genet. 2001;10(7):693–698. doi: 10.1093/hmg/10.7.693. [DOI] [PubMed] [Google Scholar]
  • 131.Gabrielli BG, Johnstone RW, Saunders NA. Identifying molecular targets mediating the anticancer activity of histone deacetylase inhibitors: a work in progress. Curr. Cancer Drug Targets. 2002;2(4):337–353. doi: 10.2174/1568009023333818. [DOI] [PubMed] [Google Scholar]
  • 132.Choi CH, Burton ZF, Usheva A. Auto-acetylation of transcription factors as a control mechanism in gene expression. Cell Cycle. 2004;3(2):114–115. doi: 10.4161/cc.3.2.651. [DOI] [PubMed] [Google Scholar]
  • 133.Halkidou K, Gnanapragasam VJ, Meht PB, et al. Upregulation and nuclear recruitment of HDAC1 in hormone refractory prostate cancer. Prostate. 2004;59(2):177–189. doi: 10.1002/pros.20022. [DOI] [PubMed] [Google Scholar]
  • 134.Osada H, Tatematsu Y, Saito H, Yatabe Y, Mitsudomi T, Takahashi T. Reduced expression of class II histone deacetylase genes is associated with poor prognosis in lung cancer patients. Int. J. Cancer. 2004;112(1):26–32. doi: 10.1002/ijc.20395. [DOI] [PubMed] [Google Scholar]
  • 135.Saunders LR, Verdin E. Sirtuins: critical regulators at the crossroads between cancer and aging. Oncogene. 2007;26(37):5489–5504. doi: 10.1038/sj.onc.1210616. [DOI] [PubMed] [Google Scholar]
  • 136.Knights CD, Catania J, Di Giovanni S, et al. Distinct p53 acetylation cassettes differentially influence gene-expression patterns and cell fate. J. Cell Biol. 2006;173(4):533–544. doi: 10.1083/jcb.200512059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Huang BH, Laban M, Leung CH, et al. Inhibition of histone deacetylase 2 increases apoptosis and p21Cip1/WAF1 expression, independent of histone deacetylase 1. Cell Death Differ. 2005;12(4):395–404. doi: 10.1038/sj.cdd.4401567. [DOI] [PubMed] [Google Scholar]
  • 138.Olaharski AJ, Rine J, Marshall BL, et al. The flavoring agent dihydrocoumarin reverses epigenetic silencing and inhibits sirtuin deacetylases. PLoS Genet. 2005;1(6):E77. doi: 10.1371/journal.pgen.0010077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Costanzo A, Merlo P, Pediconi N, et al. DNA damage-dependent acetylation of p73 dictates the selective activation of apoptotic target genes. Mol. Cell. 2002;9(1):175–186. doi: 10.1016/s1097-2765(02)00431-8. [DOI] [PubMed] [Google Scholar]
  • 140.Cohen HY, Lavu S, Bitterman KJ, et al. Acetylation of the C terminus of Ku70 by CBP and PCAF controls Bax-mediated apoptosis. Mol. Cell. 2004;13(5):627–638. doi: 10.1016/s1097-2765(04)00094-2. [DOI] [PubMed] [Google Scholar]
  • 141.Faiola F, Liu X, Lo S, et al. Dual regulation of c-Myc by p300 via acetylation-dependent control of Myc protein turnover and coactivation of Myc-induced transcription. Mol. Cell Biol. 2005;25(23):10220–10234. doi: 10.1128/MCB.25.23.10220-10234.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.McMahon C, Suthiphongchai T, DiRenzo J, Ewen ME. P/CAF associates with cyclin D1 and potentiates its activation of the estrogen receptor. Proc. Natl Acad. Sci. USA. 1999;96(10):5382–5387. doi: 10.1073/pnas.96.10.5382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Rampalli S, Pavithra L, Bhatt A, Kundu TK, Chattopadhyay S. Tumor suppressor SMAR1 mediates cyclin D1 repression by recruitment of the SIN3/ histone deacetylase 1 complex. Mol. Cell Biol. 2005;25(19):8415–8429. doi: 10.1128/MCB.25.19.8415-8429.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Ocker M, Schneider-Stock R. Histone deacetylase inhibitors: signalling towards p21cip1/waf1. Int. J. Biochem. Cell Biol. 2007;39(7–8):1367–1374. doi: 10.1016/j.biocel.2007.03.001. [DOI] [PubMed] [Google Scholar]
  • 145.Ju R, Muller MT. Histone deacetylase inhibitors activate p21(WAF1) expression via ATM. Cancer Res. 2003;63(11):2891–2897. [PubMed] [Google Scholar]
  • 146.Yang XJ. The diverse superfamily of lysine acetyltransferases and their roles in leukemia and other diseases. Nucleic Acids Res. 2004;32(3):959–976. doi: 10.1093/nar/gkh252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Hirsch CL, Bonham K. Histone deacetylase inhibitors regulate p21WAF1 gene expression at the post-transcriptional level in HepG2 cells. FEBS Lett. 2004;570(1–3):37–40. doi: 10.1016/j.febslet.2004.06.018. [DOI] [PubMed] [Google Scholar]
  • 148.Zhang HS, Gavin M, Dahiya A, et al. Exit from G1 and S phase of the cell cycle is regulated by repressor complexes containing HDAC-Rb-hSWI/SNF and Rb-hSWI/SNF. Cell. 2000;101(1):79–89. doi: 10.1016/S0092-8674(00)80625-X. [DOI] [PubMed] [Google Scholar]
  • 149.Riggs MG, Whittaker RG, Neumann JR, Ingram VM. N-Butyrate causes histone modification in HeLa and Friend erythroleukaemia cells. Nature. 1977;268(5619):462–464. doi: 10.1038/268462a0. [DOI] [PubMed] [Google Scholar]
  • 150.Yoshida M, Kijima M, Akita M, Beppu T. Potent and specific inhibition of mammalian histone deacetylase both in vivo and in vitro by trichostatin A. J. Biol. Chem. 1990;265(28):17174–17179. [PubMed] [Google Scholar]
  • 151.Göttlicher M, Minucci S, Zhu P, Krämer OH, Schimpf A, Giavara S, et al. Valproic acid defines a novel class of HDAC inhibitors inducing differentiation of transformed cells. EMBO J. 2001;20(24):6969–6978. doi: 10.1093/emboj/20.24.6969. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Ma X, Ezzeldin HH, Diasio RB. Histone deacetylase inhibitors: current status and overview of recent clinical trials. Drugs. 2009;69(14):1911–1934. doi: 10.2165/11315680-000000000-00000. [DOI] [PubMed] [Google Scholar]
  • 153.Steele NL, Plumb JA, Vidal L, et al. A Phase I pharmacokinetic and pharmacodynamic study of the histone deacetylase inhibitor belinostat in patients with advanced solid tumors. Clin. Cancer Res. 2008;14(3):804–810. doi: 10.1158/1078-0432.CCR-07-1786. [DOI] [PubMed] [Google Scholar]
  • 154.Undevia SD, Kindler HL, Janisch L, et al. A Phase I study of the oral combination of CI-994, a putative histone deacetylase inhibitor, and capecitabine. Ann. Oncol. 2004;15(11):1705–1711. doi: 10.1093/annonc/mdh438. [DOI] [PubMed] [Google Scholar]
  • 155.Acharya MR, Sparreboom A, Venitz J, Figg WD. Rational development of histone deacetylase inhibitors as anticancer agents: a review. Mol. Pharmacol. 2005;68(4):917–932. doi: 10.1124/mol.105.014167. [DOI] [PubMed] [Google Scholar]
  • 156.Schneider-Stock R, Ocker M. Epigenetic therapy in cancer: molecular background and clinical development of histone deacetylase and DNA methyltransferase inhibitors. IDrugs. 2007;10(8):557–561. [PubMed] [Google Scholar]
  • 157.Kramer OH, Göttlicher M, Heinzel T. Histone deacetylase as a therapeutic target. Trends Endocrinol. Metab. 2001;12(7):294–300. doi: 10.1016/s1043-2760(01)00438-6. [DOI] [PubMed] [Google Scholar]
  • 158.Marks PA, Xu WS. Histone deacetylase inhibitors: potential in cancer therapy. J. Cellular Biochem. 2009;107(4):600–608. doi: 10.1002/jcb.22185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Ungerstedt JS, Sowa Y, Xu WS, et al. Role of thioredoxin in the response of normal and transformed cells to histone deacetylase inhibitors. Proc. Natl Acad. Sci. USA. 2005;102:673–678. doi: 10.1073/pnas.0408732102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Rosato RR, Almenara JA, Grant S. The histone deacetylase inhibitor MS-275 promotes differentiation or apoptosis in human leukemia cells through a process regulated by generation of reactive oxygen species and induction of p21CIP1/WAF1 1. Cancer Res. 2003;63(13):637–645. [PubMed] [Google Scholar]
  • 161.Finnin MS, Donigian JR, Cohen A, et al. Structures of a histone deacetylase homologue bound to the TSA and SAHA inhibitors. Nature. 1999;401(6749):188–193. doi: 10.1038/43710. [DOI] [PubMed] [Google Scholar]
  • 162.Minucci S, Pelicci PG. Histone deacetylase inhibitors and the promise of epigenetic (and more) treatments for cancer. Nat. Rev. Cancer. 2006;6(1):38–51. doi: 10.1038/nrc1779. [DOI] [PubMed] [Google Scholar]
  • 163.Choudhary C, Kumar C, Gnad F, et al. Lysine acetylation targets protein complexes and co-regulates major cellular functions. Science. 2009;325(5942):834–840. doi: 10.1126/science.1175371. [DOI] [PubMed] [Google Scholar]
  • 164.Marks PA, Breslow R. Dimethyl sulfoxide to vorinostat: development of this histone deacetylase inhibitor as an anticancer drug. Nat. Biotechnol. 2007;25(1):84–90. doi: 10.1038/nbt1272. [DOI] [PubMed] [Google Scholar]
  • 165.Piekarz RL, Frye R, Turner M, et al. Phase II multi-institutional trial of the histone deacetylase inhibitor romidepsin as monotherapy for patients with cutaneous T-cell lymphoma. J. Clin. Oncol. 2009;27(32):5410–7. doi: 10.1200/JCO.2008.21.6150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Qian DZ, Wang X, Kachhap SK, et al. The histone deacetylase inhibitor NVP-LAQ824 inhibits angiogenesis and has a greater antitumor effect in combination with the vascular endothelial growth factor receptor tyrosine kinase inhibitor PTK787/ZK222584. Cancer Res. 2004;64(18):6626–6634. doi: 10.1158/0008-5472.CAN-04-0540. [DOI] [PubMed] [Google Scholar]
  • 167.Deroanne CF, Bonjean K, Servotte S, et al. Histone deacetylases inhibitors as anti-angiogenic agents altering vascular endothelial growth factor signaling. Oncogene. 2002;21(3):427–436. doi: 10.1038/sj.onc.1205108. • Provides an overview of the HDAC superfamily, highlights the positive results of HDACi in cancer clinical trials and comments on the prospects for the next generation of such inhibitors.
  • 168.Michaelis M, Michaelis UR, Fleming I, et al. Valproic acid inhibits angiogenesis in vitro and in vivo. Mol. Pharmacol. 2004;65(3):520–527. doi: 10.1124/mol.65.3.520. [DOI] [PubMed] [Google Scholar]
  • 169.Wang LG, Liu XM, Fang Y, et al. De-repression of the p21 promoter in prostate cancer cells by an isothiocyanate via inhibition of HDACs and c-Myc. Int. J. Oncol. 2008;33(2):375–380. [PubMed] [Google Scholar]
  • 170.Le Tourneau C, Siu LL. Promising antitumor activity with MGCD0103, a novel isotype-selective histone deacetylase inhibitor. Expert Opin. Investig. Drugs. 2008;17(8):1247–1254. doi: 10.1517/13543784.17.8.1247. [DOI] [PubMed] [Google Scholar]
  • 171.Haggarty SJ, Koeller KM, Wong JC, Grozinger CM, Schreiber SL. Domain-selective small-molecule inhibitor of histone deacetylase 6 (HDAC6)-mediated tubulin deacetylation. Proc. Natl Acad. Sci. USA. 2003;100(8):4389–4394. doi: 10.1073/pnas.0430973100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Oehme I, Deubzer HE, Wegener D, et al. Histone deacetylase 8 in neuroblastoma tumorigenesis. Clin. Cancer Res. 2009;15(1):91–99. doi: 10.1158/1078-0432.CCR-08-0684. [DOI] [PubMed] [Google Scholar]
  • 173.Karagiannis TC, El-Osta A. Will broad-spectrum histone deacetylase inhibitors be superseded by more specific compounds? Leukemia. 2007;21(1):61–65. doi: 10.1038/sj.leu.2404464. [DOI] [PubMed] [Google Scholar]
  • 174.Suzuki T. Explorative study on isoform-selective histone deacetylase inhibitors. Chem. Pharm. Bull. (Tokyo) 2009;57(9):897–906. doi: 10.1248/cpb.57.897. [DOI] [PubMed] [Google Scholar]
  • 175.Sargeant AM, Rengel RC, Kulp SK, et al. OSU-HDAC42, a histone deacetylase inhibitor, blocks prostate tumor progression in the transgenic adenocarcinoma of the mouse prostate model. Cancer Res. 2008;68(10):3999–4009. doi: 10.1158/0008-5472.CAN-08-0203. [DOI] [PubMed] [Google Scholar]
  • 176.Ramondetta L, Burke TW, Levenback C, et al. Treatment of uterine papillary serous carcinoma with paclitaxel. Gynecol. Oncol. 2001;82(1):156–161. doi: 10.1006/gyno.2001.6211. [DOI] [PubMed] [Google Scholar]
  • 177.Chen C-S, Wang Y-C, Yang H-C, et al. Histone deacetylase inhibitors sensitize prostate cancer cells to agents that produce DNA double-strand breaks by targeting Ku70 acetylation. Cancer Res. 2007;67(11):5318–5327. doi: 10.1158/0008-5472.CAN-06-3996. [DOI] [PubMed] [Google Scholar]
  • 178.Myzak MC, Dashwood RH. Chemoprotection by sulforaphane: keep one eye beyond Keap1. Cancer Lett. 2006;233(2):208–218. doi: 10.1016/j.canlet.2005.02.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179.Druesne N, Pagniez A, Mayeur C, et al. Diallyl disulfide (DADS) increases histone acetylation and p21(waf1/cip1) expression in human colon tumor cell lines. Carcinogenesis. 2004;25(7):1227–1236. doi: 10.1093/carcin/bgh123. [DOI] [PubMed] [Google Scholar]
  • 180.Guyonnet D, Bergès R, Siess MH, et al. Post-initiation modulating effects of allyl sulfides in rat hepatocarcinogenesis. Food Chem. Toxicol. 2004;42(9):1479–1485. doi: 10.1016/j.fct.2004.04.010. [DOI] [PubMed] [Google Scholar]
  • 181.Lu Q, Lin X, Feng J, et al. Phenylhexyl isothiocyanate has dual function as histone deacetylase inhibitor and hypomethylating agent and can inhibit myeloma cell growth by targeting critical pathways. J. Hematol. Oncol. 2008;1:6. doi: 10.1186/1756-8722-1-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Myzak MC, Karplus PA, Chung F-L, Dashwood RH. A novel mechanism of chemoprotection by sulforaphane: inhibition of histone deacetylase. Cancer Res. 2004;64(16):5767–5774. doi: 10.1158/0008-5472.CAN-04-1326. [DOI] [PubMed] [Google Scholar]
  • 183.Gu W, Roeder RG. Activation of p53 sequence-specific DNA binding by acetylation of the p53C-terminal domain. Cell. 1997;90:595–606. doi: 10.1016/s0092-8674(00)80521-8. [DOI] [PubMed] [Google Scholar]
  • 184.Ma H, Nguyen C, Lee KS, Kahn M. Differential roles for the coactivators CBP and p300 on TCF/β-catenin-mediated survivin gene expression. Oncogene. 2005;24(22):3619–3631. doi: 10.1038/sj.onc.1208433. [DOI] [PubMed] [Google Scholar]
  • 185.Gay F, Calvo D, Lo MC, et al. Acetylation regulates subcellular localization of the Wnt signaling nuclear effector POP-1. Genes Dev. 2003;17(6):717–722. doi: 10.1101/gad.1042403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Marzio G, Wagener C, Gutierrez MI, Cartwright P, Helin K, Giacca M. E2F family members are differentially regulated by reversible acetylation. J. Biol. Chem. 2000;275(15):10887–10892. doi: 10.1074/jbc.275.15.10887. [DOI] [PubMed] [Google Scholar]
  • 187.Kovacs JJ, Cohen TJ, Yao TP. Chaperoning steroid hormone signaling via reversible acetylation. Nucl. Recept. Signal. 2005;3:E004. doi: 10.1621/nrs.03004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Kawai H, Yamada Y, Tatsuka M, Niwa O, Yamamoto K, Suzuki F. Down-regulation of nuclear factor κB is required for p53-dependent apoptosis in x-ray-irradiated mouse lymphoma cells and thymocytes. Cancer Res. 1999;59:6038–6041. [PubMed] [Google Scholar]
  • 189.Markham D, Munro S, Soloway J, O’Connor DP, La Thangue NB. DNA-damage-responsive acetylation of pRb regulates binding to E2F-1. EMBO Rep. 2006;7(2):192–198. doi: 10.1038/sj.embor.7400591. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES