Skip to main content
  • Review Article
  • Published:

Histone Acetylation Modifiers in the Pathogenesis of Malignant Disease

Abstract

Chromatin structure is gaining increasing attention as a potential target in the treatment of cancer. Relaxation of the chromatin fiber facilitates transcription and is regulated by two competing enzymatic activities, histone acetyltransferases (HATs) and histone deacetylases (HDACs), which modify the acetylation state of histone proteins and other promoter-bound transcription factors. While HATs, which are frequently part of multisubunit coactivator complexes, lead to the relaxation of chromatin structure and transcriptional activation, HDACs tend to associate with multisubunit corepressor complexes, which result in chromatin condensation and transcriptional repression of specific target genes. HATs and HDACs are known to be involved both in the pathogenesis as well as in the suppression of cancer. Some of the genes encoding these enzymes have been shown to be rearranged in the context of chromosomal translocations in human acute leukemias and solid tumors, where fusions of regulatory and coding regions of a variety of transcription factor genes result in completely new gene products that may interfere with regulatory cascades controlling cell growth and differentiation. On the other hand, some histone acetylation-modifying enzymes have been located within chromosomal regions that are particularly prone to chromosomal breaks. In these cases gains and losses of chromosomal material may affect the availability of functionally active HATs and HDACs, which in turn disturbs the tightly controlled equilibrium of histone acetylation. We review herein the recent achievements, which further help to elucidate the biological role of histone acetylation modifying enzymes and their potential impact on our current understanding of the molecular changes involved in the development of solid tumors and leukemias.

DNA in chromatin is organized in arrays of nucleosomes, where two copies of each histone protein—H2A, H2B, H3, and H4—are assembled into an octamer that has approximately 146 base pairs of DNA wrapped around it in 1.8 turns to form a nucleosome. The nucleosome is an invariant component of euchromatin and heterochromatin in the interphase nucleus, and of mitotic chromosomes. This highly conserved nucleoprotein complex occurs fundamentally every 200 ± 40 bp throughout all eukaryotic genomes (1). During mitosis, the tightly packed metaphase chromosomes need to be accurately distributed between two daughter cells, while the DNA has to be accessible to various enzymatic machineries during interphase, when DNA is replicated, specific parts are transcribed, and mutated DNA segments are repaired. Under these circumstances, the nucleosomal architecture represents a major structural obstacle that limits the access of factors to nucleosome-bound DNA (2). The interaction of DNA with histone proteins is highly complex and may—at least in part—be explained by electrostatic interactions between negatively charged phosphate groups in the DNA backbone and positively charged amino acids in the histone proteins (35). A number of post-translational modifications of the histone components of chromatin, including acetylation, phosphorylation, ubiquitination, methylation, and ADP-ribosylation, which altogether affect transcriptional regulation, have been described (68). However, our focus in this review is on the role of histone modification through acetylation in the pathogenesis of cancer.

First observations linking transcriptional activity with histone acetylation and deacetylation of the ε-amino groups of conserved lysine residues, which are present in the amino terminal tails of all four core histones (H2A, H2B, H3, and H4), were made more than three decades ago (9). These observations have been reinforced by studies that demonstrated transcriptionally active euchromatin domains to be highly acetylated and/or hypomethylated (912), while densely methylated inactive DNA has been associated with hypoacetylated histone proteins (9,13,14). Notably, most DNA in mammals is methylated at CpG dinucleotides, with the exception of promoter elements, which contain undermethylated CpG islands (15). Methyl-CpG binding protein 2 (MeCP2) is a protein that recognizes methylated DNA and interacts with histone deacetylases, which are part of the mSIN3A/histone deacetylases (HDAC) multi-subunit repressor complex. This suggests that MeCP2 mediates silencing of methylated DNA through deacetylation (1618) (Fig. 1).

Fig 1
figure 1

In a Giemsa-stained chromosome (left) the unstained regions (white) represent regions, which are thought to be less compact and transcriptionally active (euchromatin), and heavily stained regions (black) are considered transcriptionally inactive and highly condensed (heterochromatin). These observations have been reinforced by studies that demonstrated transcriptionally active euchromatin domains to be highly acetylated when compared to transcriptionally silenced chromosomal regions, which are hypoacetylated (9–12, 228–230). Electrostatic interactions between DNA (negatively charged) and N-terminal histone protein tails (which positively charged when deacetylated and loose their positive charge through acetylation) may at least in part serve as an explanation for the highly complex mechanisms that are involved in chromatin compaction. MeCP2, which recognizes methylated DNA, recruits HDACs, which are part of multisubunit repressor complexes (e.g., SIN3a or N-COR/SMRT-SIN3a) and mediates silencing of methylated DNA through deacetylation (1517). Under such conditions chromosomal DNA is inaccessible to DNA-binding factors, which are necessary for transcription, repair, replication, etc. Conversely, demethylation and/or hyperacetylation are associated with transcriptional activation and the unfolding of chromatin, thereby allowing access to transcription factors and other regulators (231233).

It took more than three decades to test the validity of the hypothesis that linked transcriptional activity with the post-translational modification of histone proteins, following the identification of the regulators of histone acetylation, histone acetyltransferases, and histone deacetylases (19). These enzymes allow reversible modification of histone proteins through the addition or removal of acetyl groups, which alter the strength of the bonding between histones and DNA, thereby modifying the regulation of biological processes such as DNA replication and repair, gene expression, chromatin assembly, condensation, and cell division (see also 20,21 for reviews). In addition to the effect of histone acetyltransferases (HATs) and HDACs on the charge of the histone octamer, these enzymes may also directly alter the activity of basal and sequence-specific transcription factors as well as other cellular regulators (cell-cycle regulators, signaling cascades, etc.) (Fig. 2) (5,22,23).

Fig. 2
figure 2

Histone acetylation levels in cells result from a dynamic equilibrium between the competing enzymatic activities of HATs and HDACs. Changes in histone acetylation levels have been reported to affect transcriptional regulation, signal transduction cascades, cell survival, differentiation, and the activities of target proteins (e.g., transcription factors, cell-cycle regulators, etc.). While HDACs, which may be recruited to specific promotors by transcription factor-bound multisubunit repressor complexes (e.g., N-COR/SMRT-SIN3A) have mainly been associated with transcriptional repression (16,74,75,80–83,158,193,194,234), HATs may be part of enhancer complexes and have predominantly been associated with transcriptional activation (19,41,151,235,236). In addition, HAT modifiers have been implicated in the regulation of diverse signaling cascades (122,237,238), in the alteration of protein conformation and protein activities (e.g., hormone receptors, transcription factors, DNA-associated regulators) (20,239241), and in the regulation of cell-cycle events, whereas HDAC has been observed to result in a lengthening of G2 and M phases (20,242). Inhibition of HDAC arrested the cell cycle in Gl and G2 (20,212). HATs, on the other hand, have been connected mainly with cell-cycle progression (235,243246).

Histone Modification and Transcriptional Control

The work of many investigators during the last few years has contributed to almost explosive advances in our understanding of the molecular details of transcriptional regulation and chromatin modification within the context of the highly complex interplay of protein-DNA binding factors and protein-protein interactions. It is now becoming increasingly obvious that most enzymes that regulate the acetylation state of histone proteins and other promoter-bound transcription factors (i.e., HATs and HDACs) exert their enzymatic activities as members of large multisubunit protein complexes. A deregulation of the tightly controlled equilibrium of acetylation and deacetylation plays a causative role in the generation as well as in the suppression of several types of cancer (20,2427). Depending on the specific target promotors, hyperacetylation and deacetylation may exert contradictive effects on gene expression (28) and suppress tumorigenesis in some cases, while they facilitate cancer development in others. This could be either (1) a consequence of chromosomal translocations, where histone acetylation modifiers may be fused to or recruited by a newly generated transcription factor hybrid protein, which alters the expression of specific target genes, or (2) an effect of overall changes in the concentrations of functionally available histone acetylation modifiers.

Histone Acetyltransferases

The first HAT gene to be cloned was HAT1, a yeast histone acetyltransferase, initially identified as a temperature-sensitive mutant lacking the capability to acetylate specific lysine residues of the Histone H4 peptide (2931). This was followed by the discovery of a homolog of the yeast HAT GCN5 (general control of amino acid synthesis) in Tetrahymena, which was identified by virtue of its HAT enzymatic activity (19). GCN5 is known to be the catalytic unit of both the yeast ADA (Ada2-Ada3-Gcn5) and SAGA (SPT-ADA-GCN5-acetyltransferase) coactivator complexes, which exert HAT activity (3234) and of the human SAGA-homolog STAGA (SPT3-TAFII31-GCN5 acetyltransferase) (35). Since then, a number of enzymes with HAT activity have been identified in humans, including CREB binding protein (CBP)/p300 (3638); p300/CBP associated factor (p/CAF) (36,38,39); the p160 family of proteins (NCOA1-3) (4043); the MYST family, which includes the human proteins monocytic leukemia zinc finger (MOZ), monocytic leukemia zinc finger protein-related factor (MORF), Tat interacting protein 60 (Tip60), and histone acetyltransferase binding to ORC (HBO1) (4451); hTFIIIC90 (52,53); and TAFII250 (54). For the HAT proteins MOF (45), HAT-A4 (55), Esa1 (45), NuA3/NuA4 (56), and Elp3 (57), a human homolog has not been identified to date.

So far, only one report for a HAT knock-out is available, where p300 nullizygous mice were found to die early after gestation, exhibiting defects in neurulation, cell proliferation, and heart development and where heterozygous mice also revealed considerable embryonic lethality. In the same study, cells derived from p300-deficient embryos displayed specific transcriptional defects and proliferated poorly. Mice that were double heterozygous for p300 and CBP were consistently associated with embryonic death (58). Taken together, HATs can be subdivided into two broad categories, type A and type B, by virtue of their subcellular localization. While type A HATs, which are located in the nucleus, essentially are believed to acetylate chromosomal histones, thereby playing important roles in the regulation of gene expression, type B HATs are found in the cytoplasm, where they acetylate cytoplasmic histones prior to chromatin assembly (for review see 20,21,27,59,60) (Table 1).

Table 1 Human histone acetyltransferases

Histone Deacetylases

First links between the modification of histone acetylation in conjunction with transcriptional activity were observed in the early 1960s (9). In the 1970s, an inhibition of histone deacetylase activity was shown to result in the accumulation of acetylated histones in vivo (61). Finally, the biochemical fractionation of yeast extracts led to the discovery of two distinct yeast histone deacetylating activities, HDA, the catalytic subunit of a 350-kDa histone deacetylase complex, which contains the histone deacetylase 1 (HDA1) protein and HDB, the catalytic subunit of a 600-kDa histone deacetylase complex, which contains reduced potassium dependency 3 (RPD3). HDA1 and RPD3 share a significant degree of sequence homology at the protein level. Both proteins act mainly as negative regulators of transcription. They may however, counteract repression at telomeric loci, where the general hyperacetylation of histones is associated with gene activation. Whether this observation, that HDAC enzymes repress genes in some parts of the genome while they activate transcription in other parts, reflects indirect mechanisms (e.g., reduction of the expression of genes, which encode other repressor proteins) or direct, gene-specific effects, remains still to be elucidated (28,62,63). Functionally, HDA1 and RPD3 have nonoverlapping effects on the modulation of lifespan: while the deletion of RPD3 was observed to extend life in yeast, this was not the case for HDA1, unless it was combined with the deletion of additional genes (e.g., SIR3). The simultaneous deletion of both HDA1 and RPD3 has been shown to decrease lifespan and because the expression of both enzymes declines with age, this could provide a possible explanation for the increase in mortality during senescence (64). While HDA activity in yeast is strongly inhibited by Zn2+, spermine, and spermidine (65), RPD3 mutants are highly sensitive to cycloheximide (63). Null mutants of both HDA1 and RPD3 are viable and result in a general increase of histone acetylation and gene expression, except for genes located in telomeric regions, where histone acetylation has been associated with silencing (28).

Using trapoxin, a potent histone deacetylase inhibitor, as a bait in an affinity matrix, HDAC1, the first human RPD3 ortholog, was purified (66). HDAC2, a closely related protein, was found when screening studies for corepressors that interact with YY1, a transcription repressor/activator, were performed (67). Moreover, HOS1, HOS2, and HOS3, three yeast histone deacetylases, which are homologous to both HDA1 and RPD3, were identified (28). In most cases, RPD3 is physically associated with SIN3 (also referred to as RPD1) and exerts its transcriptional repression function within a 2-MDa corepressor complex, which is distinct from the 600-kDa HDB complex described above (63,6871). This corepressor complex is recruited to promoters by sequence-specific repressor proteins, including the mammalian heterodimeric repressors Mad/Max and Mxi/Max, which are repressors in large part because of their ability to recruit the RPD3/SIN3 complex to DNA-bound regulators of transcription (66,7279), while other repressors (e.g., unliganded nuclear receptors) recruit the RPD3/SIN3 complex via SMRT (silencing mediator of retinoic acid and thyroid hormone receptors) or nuclear corepressor (N-COR) (7882) and regulators like UME6 recognize URS1 and bind SIN3, which in turn interacts with RPD3 (75,83). RPD3, which contains the catalytic deacetylase subunit of the SIN3/RPD3 complex, is clearly required for repression by SIN3 (83). However, although it is likely that the SIN3/RPD3 complex performs multiple functions, some of which may play a more prominent role in the repression of transcription, it remains to be elucidated whether histone deacetylation per se is the primary mechanism of transcriptional repression (Table 2).

Table 2 Human histone deacetylases

So far, seven human histone deacetylase proteins, all of which share a highly conserved catalytic domain, have been identified, of which HDAC1, HDAC2, and HDAC3 are orthologs of yeast RPD3 (66,67,8486), while HDAC4, HDAC5, HDAC6, and HDAC7 are yeast HDA1 orthologs (87,88). All human RPD3 orthologs that have been reported so far repress transcription when targeted to DNA via a DNA-binding domain. They all bind transcription factor YY1, which can act both as an activator and a repressor of transcription (89). Accordingly, the inhibition of HDACs by trichostatin or trapoxin is associated with the activation or repression of specific gene products (90). While mammalian HDAC1 and HDAC2 have been shown to interact with mSIN3 and the N-COR or SMRT corepressor complexes, which may associate additional proteins (e.g., SAP18, SAP30, RbAp48, or RbAp46), HDAC3 does not appear to be part of such multiprotein complexes (74,75,80–82,91). Unlike the other deacetylases, HDAC4, which belongs to the HDA1 family of HDACs, has been reported to shuttle between the nucleus and the cytoplasm in a process involving active nuclear export (87,92). Unfortunately, human orthologs of the yeast histone deacetylases HOS1 (28), HOS2 (28,93) and HOS3 (28,94) have not been reported so far.

Histone Acetylation, Solid Tumors, and Leukemias: When HATs Are the Key Players

More recently, an increasing number of disease processes have been observed to involve abnormalities of the tightly regulated interplay of acetylating and deacetylating cellular events, which are maintained by the enzymatic activities of HATs and HDACs (95100). A decrease in the amount of functionally available HDACs (e.g., if HDAC loci are part of chromosomal deletions) or an increase of functionally active HAT enzymatic activities (e.g., if chromosomal segments, which encode HAT proteins, are amplified) may shift the equilibrium of histone acetylation toward acetylation, which in turn has an effect on conformation and activity of associated transcription factors (e.g., GATA-1, TFIIEβ, TFIIF, EKLF, and p53) (23) and subsequently on gene expression, interaction targets, and activities of downstream signaling pathways (28,101104). If, by contrast, HATs are fused to a transcription factor in the context of a chromosomal translocation (46–48, 105–110), which creates a novel chimeric protein, hyperacetylation may be confined to the target promoters of that specific transcription factor and, because transcription factors have many domains of protein-protein interaction, the transcription factor/HAT-fusion may allow acetylation of associated regulators and result in the transcriptional activation of a restricted number of genes. As a consequence, chromosomal regions that were silenced under normal conditions may now be derepressed and change the entire pattern of gene expression within affected cells: genes that were previously silenced may now be activated or even overexpressed, whereas other genes, which were previously expressed, may now secondarily be repressed (5,28,111). HAT enzymes have been found amplified, translocated, overexpressed, deleted, or point mutated in several types of cancers:

  • ■ An overexpression of HAT enzymatic activities has, for example, been described for the steroid receptor coactivator-1 (SRC1) homolog AIB1, which is involved in the pathogenesis of both breast and ovarian cancers. The associated HAT defect that has been described for this particular type of cancer is a more than 20-fold amplification of the chromosomal region, which contains the AIB1 gene, and has been determined by fluorescence in situ hybridization (FISH) (104,112).

  • ■ Changes in the availability of functionally available HATs may also directly affect-functions and activities of nonhistone proteins, in view of the fact that some histone acetyltransferases (p300, P/CAF, and TAF-250) have been described to directly acetylate general transcription factors (TFIIF and TFIIEβ) (113), sequence-specific transcription factors (e.g., GATA-1 or EKLF) (114, 115), tumor suppressors (e.g., p53 or NFκB) (116118), architectural chromatin proteins (e.g., HMGI(Y)) (119), and DNA repair complexes, where it seems to increase the DNA-binding capacity of the protein (115,116).

  • ■ Both p300 and CBP possess HAT activity, which is partly intrinsic and partly the result of association with other proteins (e.g., p/CAF) (3638). The CBP/p300 HAT-protein complex integrates many signaling pathways (e.g., the TGF-β signaling pathway) (120122) and is able to interact with a series of transcription factors (e.g., CREB, c-Jun, JunB, c-Fos, Myb, MyoD, YY1, nuclear receptors and basal components of the transcriptional apparatus, etc.) (123) and to participate in the direct or indirect stimulation of transcription through scaffolding different classes of transcriptional regulators onto specific chromatin domains (124,125). p300 and CBP have been envisioned as negative regulators of cell growth; mutations or translocations of the p300 or of the CBP genes have been found to be associated with several solid tumors (e.g., point mutations of p300, which may be found in colorectal and gastric carcinomas, are usually located within the cysteine histidine-rich regions of the protein, known to play important roles in the biological activities of p300 (121,126,127); loss of heterozygosity for p300 in 80% of glioblastomas (128) and acute leukemias (e.g., in the M4/M5 subtype of acute myeloid leukemia [AML] where CBP is found translocated and fused to the putative acetyltransferase MOZ [t(8;16)(p11;p13)], which is a human homolog of the yeast SAS genes [SAS: something about silencing]) (48,109,129). Remarkably, p300 mutations are located within the cysteine histidine-rich regions, which have been observed to play an important role in the biological activities of p300 (130). In addition, it has been reported that the oncogenic viral proteins E1A and SV40 are able to antagonize CBP-dependent transcription, thereby promoting cellular proliferation (38,131133).

Leukemia-associated chromosome 8 inversions of the genotype inv(8)(p11;q13) characteristically fuse MOZ to TIF2/NCOA2/GRIP1 (transcriptional mediator/intermediary factor 2) (134,135), a NR (nuclear hormone receptor) coactivator that itself binds CBP/p300 (136,137). The phenotype of the resulting MOZ-TIF2 fusion is therefore highly similar to the MOZ-CBP fusion.

In other leukemias, particularly therapy-related AML, myelodysplastic syndrome, and chronic myelomonocytic leukemia, CBP may be fused to MLL (mixed lineage leukemia), a gene that has been associated with the myelodysplastic syndrome [t(11;16)(q23;p13)](48,105108). In both translocations (MLL-CBP and MOZ-CBP) the HAT domain of CBP remains intact within the fusion protein and because both MOZ and MLL have been implicated in the modification of chromatin structure, it is likely that the molecular mechanism through which the fusion proteins perturb growth is by dysregulating gene expression patterns (130).

In a separate subset of AMLs, which has been reported to be associated with therapy-induced leukemia, the p300 gene was found rearranged and fused in frame with the MLL gene [t(11;22)(q23;q13)] (106,138). This suggests that alterations of CBP function may occur in the later stages of leukemogenesis, possibly as a way to eliminate cell-cycle checkpoints and apoptotic responses (130). Notably, the MLL fusion proteins described herein lack the carboxy terminal SET (suppressor of variegation) domain, which is a hall-mark of many chromatinassociated proteins (139,140). The MLL SET domain interacts with the human SWI/SNF (switch defective/sucrose nonfermenting) chromatin-remodeling complex, a powerful transcriptional activator that belongs to a family of DNA-stimulated ATPases that can either disrupt the structure of nucleosome core particles or influence the mobility and spacing of nucleosome arrays (141,142). Therefore, a fusion of MLL and CBP results in the dysregulation of transcription by failing to recruit SWI/SNF. Conversely, the leukemia-associated MLL-AF9 and MLL-ENL translocations fuse MLL to genes, which encode transcriptional activators expressing carboxy terminal domains, which are highly homologous to SET and may function as distinct targets for the SWI/SNF complex (143146). Somatic mutations within the SWI/SNF complex have been identified in several aggressive pediatric malignant rhabdoid tumors (147).

Humans lacking one functional allele of the CBP gene, having point mutations or microdeletions within the 16p13.3 chromosomal region that contains CBP, develop a condition, which has been described as “Rubinstein-Taybi syndrome” (autosomal dominant). Individuals exhibiting this condition have a particular propensity for malignancy, skeletal abnormalities, and growth retardation (128,148). Concordantly, CBP-heterozygous mice reveal skeletal abnormalities corresponding to the changes that are seen in Rubinstein-Taybi syndrome (149). Interestingly, in spite of apparently overlapping functions between CBP and p300, patients with Rubinstein-Taybi syndrome have an intact p300 allele, which is potentially unable to sufficiently substitute for CBP (130,150).

Unfortunately, it is not presently clear exactly how p300 and CBP are involved in the development of cancer. What is known, however, is that p300 and CBP, which have well-documented activity as transcriptional activators, are important key players in cell-cycle control, within apoptotic pathways, in the promotion of differentiation, and in p53 signaling and activation (116,124,151,152). In a simplified scenario the anti-oncogenic properties of p300 and CBP seem to go hand in hand with the anti-oncogenic activities that have been proposed for p53; since p300 forms a complex with p53 that exerts its anticancerogenic activity by negatively regulating cell growth. Several promoter/enhancer elements, such as the AP1 and c-Fos elements, function in a p300-dependent manner (153), which has been correlated with a promotion of G1-S transition, resulting thereby in cellular proliferation and potentially transformation (154). Although the formation of p53–p300 may in part be responsible for the recruitment of p300 onto some promoters, it may inhibit the transactivating effects of p300 on others (e.g., on promoters containing the DNA binding sites for the transcription factor AP1, where increased levels of p300 are able to overcome p53-mediated inhibition of AP1). Most interestingly, p53 binds to p300 in a region that is required for its intrinsic HAT activity (37). However, this region is distinct form the domains that bind to c-Jun (155), P/CAF (38), and TBP (156), all of which are important modulators of transcription. It has therefore been suggested that p53 might function through direct protein-protein interactions via p300 and possibly also through other p300-associated factors (124).

It has been observed that factors like the nuclear hormone receptors (e.g., RAR/RXR) mediate transcriptional repression by recruiting HDAC complexes in their unliganded form (8082), while they exhibit ligand-inducible transcriptional activator functions through the recruitment of HAT-coactivator complexes (p300/CAF) when hormone is bound (Fig. 3) (4,5,40,150,157). Similarly, E2F and Rb, for example, form a repressor complex, which recruits HDAC1 and HDAC2 (158,159) and subsequently represses the cyclin E promoter. Frequently, phosphorylation is a key event that induces a conformational change within a transcription factor or other regulatory elements (e.g., NFκB, the IFNβ enhancer complex or Rb), which then readily recruits a HAT-coactivator complex and stimulates transcription (160162).

Fig. 3
figure 3

Some factors that associate with histone modifying-enzymes (nuclear hormone receptors, transcription factors, and other regulators) have the capacity to mediate both transcriptional activation and repression, depending on whether they associate with HATs or HDACs.

Histone Acetylation, Solid Tumors, and Leukemias: When HDACs Are the Key Players

Leukemias are generally associated with characteristic chromosomal translocations, which may result either in the generation of a chimeric protein with novel functional properties or in the aberrant expression of a regulatory element. Usually, chromosomal translocations affect only one allele of a gene. Therefore, to cause a phenotypic effect, the activity of the newly generated fusion protein needs to be dominant over that of the wild-type protein. Frequently, the translocations found in leukemias target regulatory transcription factors that control cellular proliferation, survival, and differentiation and may obviate the need for multistep mutation pathways because they are observed for protooncogenes and tumor suppressor genes in solid tumors (163,164). In accordance with the disease-linked HAT defects that have been described above, HDAC defects may very similarly be associated with a tumorigenic phenotype of affected cells. When HDACs are excessively available or if the amount of functionally available HATs is decreased, the balance of histone acetylation will be shifted toward deacetylation, resulting subsequently in a dysregulation of gene expression. Several transcriptional repressors (e.g., Mad and members of the nuclear receptor superfamily), trans-cription factors, and cellular regulators have been described to associate with HDAC activities (5).

  • ■ The analysis of the transforming chimeric proteins PML-RARα [t(15;17)(q22;q21)] (163,165167) and PLZF-RARα [t(11;17) (q23;q21)] (167170), which are found in different acute promyelocytic leukemias (APL), has shown a clear connection between the action of histone deacetylases and the development of cancer. In these instances the promyelocytic leukemia gene (PML) [t(15;17) (q22;q21)] or the promyelocytic leukemia zinc finger gene (PLZF) [t(11;17)(q23;q21)] are fused to the retinoic acid receptor-alpha (RARα) and are no longer responsive to physiological levels of retinoic acid (167,171,172). These chromosomal changes result in a block of cellular differentiation (i.e., in the clonal expansion of cells arrested in the promyelocyte stage of development) yielding the clinical picture of an acute leukemia (78,171,172). While patients featuring the PML-RARα [t(15;17)(q22;q21)] translocation, readily differentiate upon treatment with all-trans-retinoic acid (ATRA) (78,171), patients having the PLZF-RARα [t(11;17)(q23;q21)] type of translocation do not respond adequately to treatment with ATRA (78,171). On the molecular level, RARα represses target genes by tethering corepressors such as N-COR and SMRT to promoter DNA (78,173). These corepressors are part of one or more large complexes that also contain mSIN3A and HDAC proteins (5). In cells that express PML-RARα, retinoic acids lead to the dissociation of the SMRT-mSIN3A-HDAC1 and N-COR-mSIN3A-HDAC1,2 complexes from RARα (97,98,172). By contrast, cells that express PLZF-RARα have two N-COR binding sites, one in the RARα region (which is responsive to retinoic acids) and one in the PLZF amino terminal region (which is nonresponsive to retinoic acids) of the fusion protein (97). Because PLZF binds N-COR and SMRT independently from RARα, the HDAC corepressor complex is readily released from RARα upon treatment with retinoic acids, but not from PLZF (97). As a consequence, transcriptional repression is preserved. However, the sensitivity of PLZF to ATRA may be restored by the treatment with an HDAC inhibitor (e.g., trichostatin A, an antibiotic). The PLZF-bound HDAC corepressor complex is then readily released, allowing these leukemic cells to differentiate (97,98,172,174, 175).

  • ■ An additional example of a translocation found in AML and that has been shown to involve HDACs is the translocation t(8;21) (q22;q22), which results in a fusion of AML-1 and ETO (176178) and accounts for approximately 10–12% of AMLs (164). Similarly, AML-1 may be fused to MTG16 (myeloid tumor gene 16) in the context of a translocation t(16;21)(q24;q22) (179181), or to EVI1, a transcriptional repressor, in association with a translocation t(3;21)(q26;q22) (177,182, 183). In all these translocations of AML-1, the Runt homology domain, which is the region of AML-1 that interacts with both DNA and the core binding factor CBFβ, is preserved (178,184,185). Other translocations that are frequently seen are an inversion of chromosome 16 [inv(16)] in AML, where the CBFβ gene, which forms a transcription factor complex with AML-1, is fused to the smooth muscle myosin heavy chain gene MYH11 (164) and the translocation t(12;21) (p12;q22), which is found in 15–35% of pediatric B-lineage ALLs and where a TEL-AML1 gene fusion yields a novel chimeric protein (186,187). The ability of CBFβ to associate with AML-1, hereby increasing the affinity of AML-1 for its DNA-binding site, is retained even when chromosome 16 is inverted (184,185). Because all the translocations mentioned in this paragraph [t(8;21) (q22;q22), t(3;21) (q26;q22), t(12;21)(p12; q22) and inv(16)] interfere with the transcriptional regulation of AML-1 responsive genes (177,178,188191), evidence emerges that transcriptional repression of AML-1 target genes is critical in the pathogenesis of AMLs. In the case of AML-1/ETO, overexpression of AML-1 and anti AML-1/ETO antisense oligonucleotides can induce differentiation in cells containing this fusion protein (185,191). In analogy to the observations that have been made for PML-RARα and PLZF-RARα, ETO has been found to interact with N-COR and mSIN3A, thereby recruiting histone deacetylases to repress transcription. This effect can be inhibited by histone deacetylase inhibitors (95,192). Additionally, it has been found that critical domains that mediate the interaction of ETO with N-COR both when AML-1 is fused to ETO [t(8;21)(q22;q22)] and when it is fused to MTG16 [t(16;21)(q24;q22)] are highly conserved, suggesting that t(16; 21) (q24;q22) equally represses transcription through the recruitment of HDACs. Besides CBFβ, AML-1 associates with other transcription factors and regulators and activates transcription if bound to HATs (e.g., CBP and p300) (191) or represses transcription when it interacts with mSIN3 (192) (Fig. 3).

  • ■ Other than the indirect recruitment of HDAC with the assistance of protein complexes, which include mSIN3A, N-COR, and SMRT or Rb, several transcription factors (e.g., YY1) and regulators are able to recruit HDAC enzymes directly (reviewed in 21), thereby interfering with the generation of a functional initiation complex. These HDAC-associated effects on transcriptional regulation may be abrogated by HDAC-inhibitors (trapoxin, trichostatin A, etc.). The retinoblastoma protein (RB), which is important in the induction of cell-cycle arrest under unfavorable growth conditions, mediates E2F-bound promoter repression through its interaction with HDAC1. This binding of HDAC to RB has been highlighted by the observations that tumor-specific mutations found in RB disrupt its association with HDAC (159,193,194) and that viral oncoproteins (e.g., HPV16 E7 or the SV40 T-antigen) are able to displace HDAC from RB (193,194). These findings suggest a fundamental role for histone modification in the suppression of cancer.

  • ■ So far, the chromosomal localizations of the human RPD3 orthologs HDAC1 3 (102,195) and the HDA1 orthologs HDAC4 7 have been identified more or less precisely (87) (Table 2) and, interestingly, HDAC1 3 and HDAC7 localize to chromosomal sites, which are particularly fragile and frequently altered through mutations, translocations, and deletions, particularly in myeloproliferative disorders and solid tumors (101–103, 196). This may potentially result in a shift of the acetylation equilibrium toward acetylation.

It had long been epidemiologically postulated that a diet high in fiber was associated with a low incidence of colon cancer (197200) until the Nurses’ Health Study at Harvard, which was conducted on 88,757 middle-aged women, proved that the “protective effect” of dietary fiber against colorectal cancer or adenoma was not significant (198). Even though there is considerable and to some extent inconclusive literature on dietary fiber in connection with colon cancer, it appears that butyrate, a fiber fermentation product, could in fact have a protective effect against colon cancer (201204). More recently, several investigators demonstrated that butyrate administration effectively reduced incidence and size of colonic tumors (205,206), their likelihood to metastasize (207), and that it shifted their histological phenotype to one that appeared less aggressive (206). Even though the molecular mechanisms by which butyrate mediates its protective effects are still very unclear, butyrate has been shown to induce both histone and nonhistone hyperacetylation through a noncompetitive and nonspecific inhibition of HDACs via a serine-threonine protein phosphatase of the PP1 type (208210). In addition to its capability to induce differentiation, butyrate has been found to cause a G1 cell-cycle arrest (211,212), which is mediated through induction of the G1 cell-cycle inhibitor p21 gene (213), thereby requiring an inhibition of HDAC1. The fact that p21 is deleted in the human colon carcinoma cell line HCT1116 further supports that p21 is essentially involved in butyrate-mediated cellular growth arrest (214). Because butyrate is rapidly metabolized and it has not been possible to maintain adequate butyrate concentrations in patients, butyrate homologs and alternative substrates [e.g., trichostatins and trapoxins (215,216), depudecin (217), oxamflatin (218), benzamide derivatives (219)], which appear more promising than butyrate itself (e.g., phenylbutyrate and tributyrin) are currently under study (220224).

Conclusion and Future Directions

The equilibrium of reversible histone acetylation is maintained by the activity of two families of enzymes, HATs and HDACs, which have been found to participate in the regulation of cellular proliferation and differentiation as co-factors of several mammalian transcriptional complexes. More recently, increasing evidence suggests a close connection between imbalanced histone acetylation and carcinogenesis. This goes hand in hand with the widely accepted concepts that reversible modification of chromatin influences its transcriptional competence and that promoters may be targeted specifically by either activating or repressing complexes. Histone acetylation modifiers have been found to be engaged and mutated in several types of cancer. Although genes encoding HAT enzymes have been found to be preferentially translocated, amplified, overexpressed, or point mutated, HDACs have repeatedly been identified to mediate the function of oncogenic translocation products, thereby accounting for at least 30% of AMLs, 25% of childhood B-ALLs, and more than 99% of APLs (26). They have also been found to associate with tumor suppressor proteins, which themselves are frequently mutated (e.g., RB). Many questions persist regarding the molecular mechanisms that involve histone-modifying enzymes. The further characterization of HATs and HDACs will therefore not only continue to unravel the role that these enzymes play in transcription, it will also help to identify the molecular mechanisms that promote leukemogenesis. Additionally, it looks like acetylation is not just limited to histones. It could therefore, in analogy to phosphorylation, be a process that influences the function of many proteins and cellular processes. The identification of proteins that interact with histone-modifying enzymes and target genes, which are misregulated as a consequence of mistargeted, or defective histone-modifying enzymes may help to hasten the development of less toxic, more refined and specific forms of pharmacological interventions for some forms of cancer and leukemias. First experiences with histone deacetylase inhibitors as “differentiation therapy” reagents have shown promising results for several types of leukemias (96,172,221224) and solid tumors (220,225227) with few, if any, significant side effects, indicating that such treatment could have great therapeutic advantages when compared to conventional chemotherapeutic agents. Even though almost explosive advances in the understanding of the molecular details of transcriptional regulation and chromatin modification through acetylation have been reported in the last few years, many questions remain. The identification of novel HAT/HDAC interaction targets, the analysis of HAT/HDAC levels in primary cells, and the response and tolerability of histone deacetylase inhibitors in patients may help to answer and generate new questions, such as how far a therapeutic modulation of intracellular acetylation levels could complement or replace existing chemotherapeutic strategies in the treatment of cancer.

References

  1. Luger K, Mader AW, Richmond RK, et al. (1997) Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 389: 251–260.

    Article  CAS  PubMed  Google Scholar 

  2. Loidl P. (1994) Histone acetylation: facts and questions. Chromosoma 103: 441–449.

    Article  PubMed  CAS  Google Scholar 

  3. Jeppesen P. (1997) Histone acetylation: a possible mechanism for the inheritance of cell memory at mitosis. Bioessays 19: 67–74.

    Article  PubMed  CAS  Google Scholar 

  4. Wade PA, Pruss D, Wolffe AP. (1997) Histone acetylation: chromatin in action. Trends Biochem. Sci. 22: 128–132.

    Article  PubMed  CAS  Google Scholar 

  5. Struhl K. (1998) Histone acetylation and transcriptional regulatory mechanisms. Genes Dev. 12: 599–606.

    Article  PubMed  CAS  Google Scholar 

  6. Spencer VA, Davie JR. (1999) Role of covalent modifications of histones in regulating gene expression. Gene 240: 1–12.

    Article  PubMed  CAS  Google Scholar 

  7. Bradbury EM. (1992) Reversible histone modifications and the chromosome cell cycle. Bioessays 14: 9–16.

    Article  PubMed  CAS  Google Scholar 

  8. Barratt MJ, Hazzalin CA, Cano E, et al. (1994) Mitogen-stimulated phosphorylation of histone H3 is targeted to a small hyperacetylation-sensitive fraction. Proc. Natl. Acad. Sci. U.S.A. 91: 4781–4785.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Allfrey VG, Faulkner R, Mirsky AE. (1964) Acetylation and methylation of histones and their possible role in the regulation of RNA synthesis. Proc. Natl. Acad. Sci. U.S.A. 51: 786–794.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Cameron EE, Bachman KE, Myohanen S, et al. (1999) Synergy of demethylation and histone deacetylase inhibition in the re-expression of genes silenced in cancer. Nat. Genet. 21: 103–107.

    Article  CAS  PubMed  Google Scholar 

  11. Hebbes TR, Thorne AW, Crane-Robinson C. (1988) A direct link between core histone acetylation and transcriptionally active chromatin. Embo J. 7: 1395–1402.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. Grunstein M. (1997) Histone acetylation in chromatin structure and transcription. Nature 389: 349–352.

    Article  CAS  PubMed  Google Scholar 

  13. Braunstein M, Rose AB, Holmes SG, et al. (1993) Transcriptional silencing in yeast is associated with reduced nucleosome acetylation. Genes Dev. 7: 592–604.

    Article  CAS  PubMed  Google Scholar 

  14. Johnson CA, O’Neill LP, Mitchell A, et al. (1998) Distinctive patterns of histone H4 acetylation are associated with defined sequence elements within both heterochromatic and euchromatic regions of the human genome. Nucleic Acids Res. 26: 994–1001.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  15. Ng HH, Bird A. (1999) DNA methylation and chromatin modification. Curr. Opin. Genet. Dev. 9: 158–163.

    Article  CAS  PubMed  Google Scholar 

  16. Nan X, Ng HH, Johnson CA, et al. (1998) Transcriptional repression by the methyl-CpG-binding protein MeCP2 involves a histone deacetylase complex. Nature 393: 386–389.

    Article  CAS  PubMed  Google Scholar 

  17. Jones PL, Veenstra GJ, Wade PA, et al. (1998) Methylated DNA and MeCP2 recruit histone deacetylase to repress transcription. Nat. Genet. 19: 187–191.

    Article  CAS  PubMed  Google Scholar 

  18. Wong CW, Privalsky ML. (1998) Transcriptional repression by the SMRT-mSin3 corepressor: multiple interactions, multiple mechanisms, and a potential role for TFIIB. Mol. Cell. Biol. 18: 5500–5510.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  19. Brownell JE, Zhou J, Ranalli T, et al. (1996) Tetrahymena histone acetyltransferase A: a homolog to yeast Gcn5p linking histone acetylation to gene activation. Cell 84: 843–851.

    Article  CAS  PubMed  Google Scholar 

  20. Kouzarides T. (1999) Histone acetylases and deacetylases in cell proliferation. Curr. Opin. Genet. Dev. 9: 40–48.

    Article  PubMed  CAS  Google Scholar 

  21. Davie JR, Chadee DN. (1998) Regulation and regulatory parameters of histone modifications. J. Cell. Biochem. Suppl. 31: 203–213.

    Article  Google Scholar 

  22. Muscat GE, Burke LJ, Downes M. (1998) The corepressor N-CoR and its variants RIP13a and RIP13Delta1 directly interact with the basal transcription factors TFIIB, TAFII32 and TAFII70. Nucleic Acids Res. 26: 2899–2907.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  23. Boyes J, Byfield P, Nakatani Y, et al. (1998) Regulation of activity of the transcription factor GATA-1 by acetylation. Nature 396: 594–598.

    Article  PubMed  CAS  Google Scholar 

  24. Minucci S, Pelicci PG. (1999) Retinoid receptors in health and disease: co-regulators and the chromatin connection. Semin. Cell Dev. Biol. 10: 215–225.

    Article  PubMed  CAS  Google Scholar 

  25. Behre G, Zhang P, Zhang DE, et al. (1999) Analysis of the modulation of transcriptional activity in myelopoiesis and leukemogenesis. Methods 17: 231–237.

    Article  PubMed  CAS  Google Scholar 

  26. Fenrick R, Hiebert SW. (1998) Role of histone deacetylases in acute leukemia. J. Cell. Biochem. Suppl. 31: 194–202.

    Article  Google Scholar 

  27. Magnaghi-Jaulin L, Ait-Si-Ali S, Harel-Bellan A. (1999) Histone acetylation in signal transduction by growth regulatory signals. Semin. Cell Dev. Biol. 10: 197–203.

    Article  CAS  PubMed  Google Scholar 

  28. Rundlett SE, Carmen AA, Kobayashi R, et al. (1996) HDA1 and RPD3 are members of distinct yeast histone deacetylase complexes that regulate silencing and transcription. Proc. Natl. Acad. Sci. U.S.A. 93: 14503–14508.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  29. Kleff S, Andrulis ED, Anderson CW, et al. (1995) Identification of a gene encoding a yeast histone H4 acetyltransferase. J. Biol. Chem. 270: 24674–24677.

    Article  CAS  PubMed  Google Scholar 

  30. Parthun MR, Widom J, Gottschling DE. (1996) The major cytoplasmic histone acetyltransferase in yeast: links to chromatin replication and histone metabolism. Cell 87: 85–94.

    Article  PubMed  CAS  Google Scholar 

  31. Chang L, Loranger SS, Mizzen C, et al. (1997) Histones in transit: cytosolic histone complexes and diacetylation of H4 during nucleosome assembly in human cells. Biochemistry 36: 469–480.

    Article  PubMed  CAS  Google Scholar 

  32. Kuo MH, Brownell JE, Sobel RE, et al. (1996) Transcription-linked acetylation by Gcn5p of histones H3 and H4 at specific lysines. Nature 383: 269–272.

    Article  PubMed  CAS  Google Scholar 

  33. Grant PA, Duggan L, Cote J, et al. (1997) Yeast Gcn5 functions in two multisubunit complexes to acetylate nucleosomal histones: characterization of an Ada complex and the SAGA (Spt/Ada) complex. Genes Dev. 11: 1640–1650.

    Article  PubMed  CAS  Google Scholar 

  34. Eberharter A, Sterner DE, Schieltz D, et al. (1999) The ADA complex is a distinct histone acetyltransferase complex in Saccharomyces cerevisiae. Mol. Cell. Biol. 19: 6621–6631.

    Article  PubMed  CAS  Google Scholar 

  35. Martinez E, Kundu TK, Fu J, et al. (1998) A human SPT3-TAFII31-GCN5-L acetylase complex distinct from transcription factor IID. J. Biol. Chem. 273: 23781–23785.

    Article  PubMed  CAS  Google Scholar 

  36. Bannister AJ, Kouzarides T. (1996) The CBP co-activator is a histone acetyltransferase. Nature 384: 641–643.

    Article  PubMed  CAS  Google Scholar 

  37. Ogryzko VV, Schiltz RL, Russanova V, et al. (1996) The transcriptional coactivators p300 and CBP are histone acetyltransferases. Cell 87: 953–959.

    Article  PubMed  CAS  Google Scholar 

  38. Yang XJ, Ogryzko VV, Nishikawa J, et al. (1996) A p300/CBP-associated factor that competes with the adenoviral oncoprotein E1A. Nature 382: 319–324.

    Article  PubMed  CAS  Google Scholar 

  39. Blanco JC, Minucci S, Lu J, et al. (1998) The histone acetylase PCAF is a nuclear receptor coactivator. Genes Dev. 12: 1638–1651.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Glass CK, Rose DW, Rosenfeld MG. (1997) Nuclear receptor coactivators. Curr. Opin. Cell Biol. 9: 222–232.

    Article  PubMed  CAS  Google Scholar 

  41. Chen H, Lin RJ, Schiltz RL, et al. (1997) Nuclear receptor coactivator ACTR is a novel histone acetyltransferase and forms a multimeric activation complex with P/CAF and CBP/p300. Cell 90: 569–580.

    Article  PubMed  CAS  Google Scholar 

  42. Spencer TE, Jenster G, Burcin MM, et al. (1997) Steroid receptor coactivator-1 is a histone acetyltransferase. Nature 389: 194–198.

    Article  PubMed  CAS  Google Scholar 

  43. Onate SA, Tsai SY, Tsai MJ, et al. (1995) Sequence and characterization of a coactivator for the steroid hormone receptor superfamily. Science 270: 1354–1357.

    Article  PubMed  CAS  Google Scholar 

  44. Champagne N, Bertos NR, Pelletier N, et al. (1999) Identification of a human histone acetyltransferase related to monocytic leukemia zinc finger protein. J. Biol. Chem. 274: 28528–28536.

    Article  PubMed  CAS  Google Scholar 

  45. Smith ER, Eisen A, Gu W, et al. (1998) ESA1 is a histone acetyltransferase that is essential for growth in yeast. Proc. Natl. Acad. Sci. U.S.A. 95: 3561–3565.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  46. Kamine J, Elangovan B, Subramanian T, et al. (1996) Identification of a cellular protein that specifically interacts with the essential cysteine region of the HIV-1 Tat transactivator. Virology 216: 357–366.

    Article  PubMed  CAS  Google Scholar 

  47. Reifsnyder C, Lowell J, Clarke A, et al. (1996) Yeast SAS silencing genes and human genes associated with AML and HIV-1 Tat interactions are homologous with acetyltransferases. Nat. Genet. 14: 42–49.

    Article  PubMed  CAS  Google Scholar 

  48. Borrow J, Stanton VP, Jr., Andresen JM, et al. (1996) The translocation t(8;16)(p11;p13) of acute myeloid leukaemia fuses a putative acetyltransferase to the CREB-binding protein. Nat. Genet. 14: 33–41.

    Article  PubMed  CAS  Google Scholar 

  49. Hilfiker A, Hilfiker-Kleiner D, Pannuti A, et al. (1997) mof, a putative acetyl transferase gene related to the Tip60 and Moz human genes and to the Sas genes of yeast, is required for dosage compensation in Drosophila. Embo J. 16: 2054–2060.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Iizuka M, Stillman B. (1999) Histone acetyltransferase HBO1 interacts with the ORC1 subunit of the human initiator protein. J. Biol. Chem. 274: 23027–23034.

    Article  PubMed  CAS  Google Scholar 

  51. Yamamoto T, Horikoshi M. (1997) Novel substrate specificity of the histone acetyltransferase activity of HIV-1-Tat interactive protein Tip60. J. Biol. Chem. 272: 30595–30598.

    Article  PubMed  CAS  Google Scholar 

  52. Hsieh YJ, Kundu TK, Wang Z, et al. (1999) The TFIIIC90 subunit of TFIIIC interacts with multiple components of the RNA polymerase III machinery and contains a histone-specific acetyltransferase activity. Mol. Cell. Biol. 19: 7697–7704.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  53. Kundu TK, Wang Z, Roeder RG. (1999) Human TFIIIC relieves chromatin-mediated repression of RNA polymerase III transcription and contains an intrinsic histone acetyltransferase activity. Mol. Cell. Biol. 19: 1605–1615.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  54. Mizzen CA, Yang XJ, Kokubo T, et al. (1996) The TAF(II)250 subunit of TFIID has histone acetyltransferase activity. Cell 87: 1261–1270.

    Article  PubMed  CAS  Google Scholar 

  55. Ruiz-Garcia AB, Sendra R, Galiana M, et al. (1998) HAT1 and HAT2 proteins are components of a yeast nuclear histone acetyltransferase enzyme specific for free histone H4. J. Biol. Chem. 273: 12599–12605.

    Article  PubMed  CAS  Google Scholar 

  56. Ohba R, Steger DJ, Brownell JE, et al. (1999) A novel H2A/H4 nucleosomal histone acetyltransferase in Tetrahymena thermophila. Mol. Cell. Biol. 19: 2061–2068.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  57. Wittschieben BO, Otero G, de Bizemont T, et al. (1999) A novel histone acetyltransferase is an integral subunit of elongating RNA polymerase II holoenzyme. Mol. Cell 4: 123–128.

    Article  PubMed  CAS  Google Scholar 

  58. Yao TP, Oh SP, Fuchs M, et al. (1998) Gene dosage-dependent embryonic development and proliferation defects in mice lacking the transcriptional integrator p300. Cell 93: 361–372.

    Article  PubMed  CAS  Google Scholar 

  59. Workman JL, Kingston RE. (1998) Alteration of nucleosome structure as a mechanism of transcriptional regulation. Annu. Rev. Biochem. 67: 545–579.

    Article  PubMed  CAS  Google Scholar 

  60. Grant PA, Berger SL. (1999) Histone acetyltransferase complexes. Semin. Cell Dev. Biol. 10: 169–177.

    Article  PubMed  CAS  Google Scholar 

  61. Boffa LC, Vidali G, Mann RS, et al. (1978) Suppression of histone deacetylation in vivo and in vitro by sodium butyrate. J. Biol. Chem. 253: 3364–3366.

    PubMed  CAS  Google Scholar 

  62. Carmen AA, Rundlett SE, Grunstein M. (1996) HDA1 and HDA3 are components of a yeast histone deacetylase (HDA) complex. J. Biol. Chem. 271: 15837–15844.

    Article  PubMed  CAS  Google Scholar 

  63. Vidal M, Gaber RF. (1991) RPD3 encodes a second factor required to achieve maximum positive and negative transcriptional states in Saccharomyces cerevisiae. Mol. Cell. Biol. 11: 6317–6327.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  64. Kim S, Benguria A, Lai CY, et al. (1999) Modulation of life-span by histone deacetylase genes in saccharomyces cerevisiae. Mol. Biol. Cell 10: 3125–3136.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  65. Vu QA, Zhang DE, Chroneos ZC, et al. (1987) Polyamines inhibit the yeast histone deacetylase. FEBS Lett. 220: 79–83.

    Article  PubMed  CAS  Google Scholar 

  66. Taunton J, Hassig CA, Schreiber SL. (1996) A mammalian histone deacetylase related to the yeast transcriptional regulator Rpd3p. Science 272: 408–411.

    Article  PubMed  CAS  Google Scholar 

  67. Yang WM, Inouye C, Zeng Y, et al. (1996) Transcriptional repression by YY1 is mediated by interaction with a mammalian homolog of the yeast global regulator RPD3. Proc. Natl. Acad. Sci. U.S.A. 93: 12845–12850.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  68. McKenzie EA, Kent NA, Dowell SJ, et al. (1993) The centromere and promoter factor, 1, CPF1, of Saccharomyces cerevisiae modulates gene activity through a family of factors including SPT21, RPD1 (SIN3), RPD3 and CCR4. Mol. Gen. Genet. 240: 374–386.

    PubMed  CAS  Google Scholar 

  69. Vidal M, Strich R, Esposito RE, et al. (1991) RPD1 (SIN3/UME4) is required for maximal activation and repression of diverse yeast genes. Mol. Cell. Biol. 11: 6306–6316.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  70. Pazin MJ, Kadonaga JT. (1997) What’s up and down with histone deacetylation and transcription? Cell 89: 325–328.

    Article  PubMed  CAS  Google Scholar 

  71. Wolffe AP. (1997) Transcriptional control. Sinful repression. Nature 387: 16–17.

    Article  PubMed  CAS  Google Scholar 

  72. Ayer DE, Lawrence QA, Eisenman RN. (1995) Mad-Max transcriptional repression is mediated by ternary complex formation with mammalian homologs of yeast repressor Sin3. Cell 80: 767–776.

    Article  PubMed  CAS  Google Scholar 

  73. Schreiber-Agus N, Chin L, Chen K, et al. (1995) An amino-terminal domain of Mxi1 mediates anti-Myc oncogenic activity and interacts with a homolog of the yeast transcriptional repressor SIN3. Cell 80: 777–786.

    Article  PubMed  CAS  Google Scholar 

  74. Hassig CA, Fleischer TC, Billin AN, et al. (1997) Histone deacetylase activity is required for full transcriptional repression by mSin3A. Cell 89: 341–347.

    Article  PubMed  CAS  Google Scholar 

  75. Laherty CD, Yang WM, Sun JM, et al. (1997) Histone deacetylases associated with the mSin3 corepressor mediate mad transcriptional repression. Cell 89: 349–356.

    Article  PubMed  CAS  Google Scholar 

  76. Amati B, Dalton S, Brooks MW, et al. (1992) Transcriptional activation by the human c-Myc oncoprotein in yeast requires interaction with Max. Nature 359: 423–426.

    Article  PubMed  CAS  Google Scholar 

  77. Kretzner L, Blackwood EM, Eisenman RN. (1992) Myc and Max proteins possess distinct transcriptional activities. Nature 359: 426–429.

    Article  PubMed  CAS  Google Scholar 

  78. Chen JD, Evans RM. (1995) A transcriptional co-repressor that interacts with nuclear hormone receptors. Nature 377: 454–457.

    Article  PubMed  CAS  Google Scholar 

  79. Horlein AJ, Naar AM, Heinzel T, et al. (1995) Ligand-independent repression by the thyroid hormone receptor mediated by a nuclear receptor co-repressor. Nature 377: 397–404.

    Article  PubMed  CAS  Google Scholar 

  80. Alland L, Muhle R, Hou H, Jr., et al. (1997) Role for N-CoR and histone deacetylase in Sin3-mediated transcriptional repression. Nature 387: 49–55.

    Article  PubMed  CAS  Google Scholar 

  81. Heinzel T, Lavinsky RM, Mullen TM, et al. (1997) A complex containing N-CoR, mSin3 and histone deacetylase mediates transcriptional repression. Nature 387: 43–48.

    Article  PubMed  CAS  Google Scholar 

  82. Nagy L, Kao HY, Chakravarti D, et al. (1997) Nuclear receptor repression mediated by a complex containing SMRT, mSin3A, and histone deacetylase. Cell 89: 373–380.

    Article  PubMed  CAS  Google Scholar 

  83. Kadosh D, Struhl K. (1997) Repression by Ume6 involves recruitment of a complex containing Sin3 corepressor and Rpd3 histone deacetylase to target promoters. Cell 89: 365–371.

    Article  PubMed  CAS  Google Scholar 

  84. Yang WM, Yao YL, Sun JM, et al. (1997) Isolation and characterization of cDNAs corresponding to an additional member of the human histone deacetylase gene family. J. Biol. Chem. 272: 28001–28007.

    Article  PubMed  CAS  Google Scholar 

  85. Emiliani S, Fischle W, Van Lint C, et al. (1998) Characterization of a human RPD3 ortholog, HDAC3. Proc. Natl. Acad. Sci. U.S.A. 95: 2795–2800.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  86. Dangond F, Hafler DA, Tong JK, et al. (1998) Differential display cloning of a novel human histone deacetylase (HDAC3) cDNA from PHA-activated immune cells. Biochem. Biophys. Res. Commun. 242: 648–652.

    Article  CAS  Google Scholar 

  87. Wang AH, Bertos NR, Vezmar M, et al. (1999) HDAC4, a human histone deacetylase related to yeast HDA1, is a transcriptional corepressor. Mol. Cell. Biol. 19: 7816–7827.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  88. Grozinger CM, Hassig CA, Schreiber SL. (1999) Three proteins define a class of human histone deacetylases related to yeast Hda1p. Proc. Natl. Acad. Sci. U.S.A. 96: 4868–4873.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  89. Shi Y, Seto E, Chang LS, et al. (1991) Transcriptional repression by YY1, a human GLI-Kruppel-related protein, and relief of repression by adenovirus E1A protein. Cell 67: 377–388.

    Article  PubMed  CAS  Google Scholar 

  90. Van Lint C, Emiliani S, Ott M, et al. (1996) Transcriptional activation and chromatin remodeling of the HIV-1 promoter in response to histone acetylation. Embo J. 15: 1112–1120.

    Article  PubMed  PubMed Central  Google Scholar 

  91. Zhang Y, Iratni R, Erdjument-Bromage H, et al. (1997) Histone deacetylases and SAP18, a novel polypeptide, are components of a human Sin3 complex. Cell 89: 357–364.

    Article  PubMed  CAS  Google Scholar 

  92. Miska EA, Karlsson C, Langley E, et al. (1999) HDAC4 deacetylase associates with and represses the MEF2 transcription factor. Embo J. 18: 5099–5107.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  93. Aiba H, Kawaura R, Yamamoto E, et al. (1998) Isolation and characterization of high-osmolarity-sensitive mutants of fission yeast. J. Bacteriol. 180: 5038–5043.

    PubMed  PubMed Central  CAS  Google Scholar 

  94. Carmen AA, Griffin PR, Calaycay JR, et al. (1999) Yeast HOS3 forms a novel trichostatin A-insensitive homodimer with intrinsic histone deacetylase activity. Proc. Natl. Acad. Sci. U.S.A. 96: 12356–12361.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  95. Wang J, Hoshino T, Redner RL, et al. (1998) ETO, fusion partner in t(8;21) acute myeloid leukemia, represses transcription by interaction with the human N-CoR/mSin3/HDAC1 complex. Proc. Natl. Acad. Sci. U.S.A. 95: 10860–10865.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  96. Wang J, Saunthararajah Y, Redner RL, et al. (1999) Inhibitors of histone deacetylase relieve ETO-mediated repression and induce differentiation of AML1-ETO leukemia cells. Cancer Res. 59: 2766–2769.

    PubMed  CAS  Google Scholar 

  97. Grignani F, De Matteis S, Nervi C, et al. (1998) Fusion proteins of the retinoic acid receptor-alpha recruit histone deacetylase in promyelocytic leukaemia. Nature 391: 815–818.

    Article  PubMed  CAS  Google Scholar 

  98. Lin RJ, Nagy L, Inoue S, et al. (1998) Role of the histone deacetylase complex in acute promyelocytic leukaemia. Nature 391: 811–814.

    Article  PubMed  CAS  Google Scholar 

  99. Kosugi H, Towatari M, Hatano S, et al. (1999) Histone deacetylase inhibitors are the potent inducer/enhancer of differentiation in acute myeloid leukemia: a new approach to anti-leukemia therapy. Leukemia 13: 1316–1324.

    Article  PubMed  CAS  Google Scholar 

  100. Horiuchi K, Fujimoto D, Fukushima M, et al. (1981) Increased histone acetylation and deacetylation in rat ascites hepatoma cells. Cancer Res. 41: 1488–1491.

    PubMed  CAS  Google Scholar 

  101. Le Beau MM, Espinosa Rd, Neuman WL, et al. (1993) Cytogenetic and molecular delineation of the smallest commonly deleted region of chromosome 5 in malignant myeloid diseases. Proc. Natl. Acad. Sci. U.S.A. 90: 5484–5488.

    Article  PubMed  PubMed Central  Google Scholar 

  102. Randhawa GS, Bell DW, Testa JR, et al. (1998) Identification and mapping of human histone acetylation modifier gene homologues. Genomics 51: 262–269.

    Article  PubMed  CAS  Google Scholar 

  103. Mahlknecht U, Bucala R, Hoelzer D, et al. (1999) High resolution physical mapping of human HDAC3, a potential tumor suppressor gene in the 5q31 region. Cytogenet. Cell Genet. 86: 237–239.

    Article  PubMed  CAS  Google Scholar 

  104. Anzick SL, Kononen J, Walker RL, et al. (1997) AIB1, a steroid receptor coactivator amplified in breast and ovarian cancer. Science 277: 965–968.

    Article  PubMed  CAS  Google Scholar 

  105. Rowley JD, Reshmi S, Sobulo O, et al. (1997) All patients with the T(11;16) (q23;p13.3) that involves MLL and CBP have treatment-related hematologic disorders. Blood 90: 535–541.

    PubMed  CAS  Google Scholar 

  106. Taki T, Sako M, Tsuchida M, et al. (1997) The t(11;16)(q23;p13) translocation in myelodysplastic syndrome fuses the MLL gene to the CBP gene. Blood 89: 3945–3950.

    PubMed  CAS  Google Scholar 

  107. Sobulo OM, Borrow J, Tomek R, et al. (1997) MLL is fused to CBP, a histone acetyltransferase, in therapy-related acute myeloid leukemia with a t(11;16)(q23;p13.3). Proc. Natl. Acad. Sci. U.S.A. 94: 8732–8737.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  108. Satake N, Ishida Y, Otoh Y, et al. (1997) Novel MLL-CBP fusion transcript in therapy-related chronic myelomonocytic leukemia with a t(11;16)(q23;p13) chromosome translocation. Genes Chromosomes Cancer 20: 60–63.

    Article  PubMed  CAS  Google Scholar 

  109. Quesnel B, Kantarjian H, Bjergaard JP, et al. (1993) Therapy-related acute myeloid leukemia with t(8;21), inv(16), and t(8;16): a report on 25 cases and review of the literature. J. Clin. Oncol. 11: 2370–2379.

    Article  PubMed  CAS  Google Scholar 

  110. Dutnall RN, Tafrov ST, Sternglanz R, et al. (1998) Structure of the histone acetyltransferase Hat1: a paradigm for the GCN5-related N-acetyltransferase superfamily. Cell 94: 427–438.

    Article  CAS  PubMed  Google Scholar 

  111. Mahlknecht U, Ottmann OG, Hoelzer D. (2000) When the band begins to play: Histone acetylation caught in the crossfire of gene control. Mol. Carcinog. (in press).

  112. Lee SK, Anzick SL, Choi JE, et al. (1999) A nuclear factor, ASC-2, as a cancer-amplified transcriptional coactivator essential for ligand-dependent transactivation by nuclear receptors in vivo. J. Biol. Chem. 274: 34283–34293.

    Article  CAS  PubMed  Google Scholar 

  113. Imhof A, Yang XJ, Ogryzko VV, et al. (1997) Acetylation of general transcription factors by histone acetyltransferases. Curr. Biol. 7: 689–692.

    Article  CAS  PubMed  Google Scholar 

  114. Boyes J, Byfield P, Nakatani Y, et al. (1998) Regulation of activity of the transcription factor GATA-1 by acetylation. Nature 396: 594–598.

    Article  PubMed  CAS  Google Scholar 

  115. Zhang W, Bieker JJ. (1998) Acetylation and modulation of erythroid Kruppel-like factor (EKLF) activity by interaction with histone acetyltransferases. Proc. Natl. Acad. Sci. U.S.A. 95: 9855–9860.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  116. Gu W, Roeder RG. (1997) Activation of p53 sequence-specific DNA binding by acetylation of the p53 C-terminal domain. Cell 90: 595–606.

    Article  CAS  PubMed  Google Scholar 

  117. Sakaguchi K, Herrera JE, Saito S, et al. (1998) DNA damage activates p53 through a phosphorylation-acetylation cascade. Genes Dev. 12: 2831–2841.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  118. Perkins ND, Felzien LK, Betts JC, et al. (1997) Regulation of NF-kappaB by cyclin-dependent kinases associated with the p300 coactivator. Science 275: 523–527.

    Article  CAS  PubMed  Google Scholar 

  119. Munshi N, Merika M, Yie J, et al. (1998) Acetylation of HMG I(Y) by CBP turns off IFN beta expression by disrupting the enhanceosome. Mol. Cell 2: 457–467.

    Article  CAS  PubMed  Google Scholar 

  120. Feng XH, Zhang Y, Wu RY, et al. (1998) The tumor suppressor Smad4/DPC4 and transcriptional adaptor CBP/p300 are coactivators for smad3 in TGF-beta-induced transcriptional activation. Genes Dev. 12: 2153–2163.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  121. Janknecht R, Wells NJ, Hunter T. (1998) TGF-beta-stimulated cooperation of smad proteins with the coactivators CBP/p300. Genes Dev. 12: 2114–2119.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  122. Luo K, Stroschein SL, Wang W, et al. (1999) The Ski oncoprotein interacts with the Smad proteins to repress TGFbeta signaling. Genes Dev. 13: 2196–2206.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  123. Janknecht R, Hunter T. (1996) Transcription. A growing coactivator network. Nature 383: 22–23.

    Article  CAS  PubMed  Google Scholar 

  124. Avantaggiati ML, Ogryzko V, Gardner K, et al. (1997) Recruitment of p300/CBP in p53-dependent signal pathways. Cell 89: 1175–1184.

    Article  CAS  PubMed  Google Scholar 

  125. Martinez-Balbas MA, Bannister AJ, Martin K, et al. (1998) The acetyltransferase activity of CBP stimulates transcription. Embo J. 17: 2886–2893.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  126. Zhou S, Buckhaults P, Zawel L, et al. (1998) Targeted deletion of Smad4 shows it is required for transforming growth factor beta and activin signaling in colorectal cancer cells. Proc. Natl. Acad. Sci. U.S.A. 95: 2412–2416.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  127. Muraoka M, Konishi M, Kikuchi-Yanoshita R, et al. (1996) p300 gene alterations in colorectal and gastric carcinomas. Oncogene 12: 1565–1569.

    CAS  PubMed  Google Scholar 

  128. Giles RH, Peters DJ, Breuning MH. (1998) Conjunction dysfunction: CBP/p300 in human disease. Trends Genet. 14: 178–183.

    Article  CAS  PubMed  Google Scholar 

  129. Aguiar RC, Chase A, Coulthard S, et al. (1997) Abnormalities of chromosome band 8p11 in leukemia: two clinical syndromes can be distinguished on the basis of MOZ involvement. Blood 90: 3130–3135.

    CAS  PubMed  Google Scholar 

  130. Giordano A, Avantaggiati ML. (1999) p300 and CBP: partners for life and death. J. Cell Physiol. 181: 218–230.

    Article  CAS  PubMed  Google Scholar 

  131. Arany Z, Newsome D, Oldread E, et al. (1995) A family of transcriptional adaptor proteins targeted by the E1A oncoprotein. Nature 374: 81–84.

    Article  CAS  PubMed  Google Scholar 

  132. Lundblad JR, Kwok RP, Laurance ME, et al. (1995) Adenoviral E1A-associated protein p300 as a functional homologue of the transcriptional co-activator CBP. Nature 374: 85–88.

    Article  CAS  PubMed  Google Scholar 

  133. Eckner R, Ludlow JW, Lill NL, et al. (1996) Association of p300 and CBP with simian virus 40 large T antigen. Mol. Cell. Biol. 16: 3454–3464.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  134. Carapeti M, Aguiar RC, Goldman JM, et al. (1998) A novel fusion between MOZ and the nuclear receptor coactivator TIF2 in acute myeloid leukemia. Blood 91: 3127–3133.

    CAS  PubMed  Google Scholar 

  135. Liang J, Prouty L, Williams BJ, et al. (1998) Acute mixed lineage leukemia with an inv(8)(p11q13) resulting in fusion of the genes for MOZ and TIF2. Blood 92: 2118–2122.

    CAS  PubMed  Google Scholar 

  136. Voegel JJ, Heine MJ, Zechel C, et al. (1996) TIF2, a 160 kDa transcriptional mediator for the ligand-dependent activation function AF-2 of nuclear receptors. Embo J. 15: 3667–3675.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  137. Voegel JJ, Heine MJ, Tini M, et al. (1998) The coactivator TIF2 contains three nuclear receptor-binding motifs and mediates transactivation through CBP binding-dependent and -independent pathways. Embo J. 17: 507–519.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  138. Ida K, Kitabayashi I, Taki T, et al. (1997) Adenoviral E1A-associated protein p300 is involved in acute myeloid leukemia with t(11;22) (q23;q13). Blood 90: 4699–4704.

    CAS  PubMed  Google Scholar 

  139. Stassen MJ, Bailey D, Nelson S, et al. (1995) The Drosophila trithorax proteins contain a novel variant of the nuclear receptor type DNA binding domain and an ancient conserved motif found in other chromosomal proteins. Mech. Dev. 52: 209–223.

    Article  CAS  PubMed  Google Scholar 

  140. Nislow C, Ray E, Pillus L. (1997) SET1, A Yeast Member of the Trithorax Family, Functions in Transcriptional Silencing and Diverse Cellular Processes. Mol. Biol. Cell 8: 2421–2436.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  141. Rozenblatt-Rosen O, Rozovskaia T, Burakov D, et al. (1998) The C-terminal SET domains of ALL-1 and TRITHORAX interact with the INI1 and SNR1 proteins, components of the SWI/SNF complex. Proc. Natl. Acad. Sci. U.S.A. 95: 4152–4157.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  142. Pollard KJ, Peterson CL. (1998) Chromatin remodeling: a marriage between two families? Bioessays 20: 771–780.

    Article  CAS  PubMed  Google Scholar 

  143. Welch MD, Drubin DG. (1994) A nuclear protein with sequence similarity to proteins implicated in human acute leukemias is important for cellular morphogenesis and actin cytoskeletal function in Saccharomyces cerevisiae. Mol. Biol. Cell 5: 617–632.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  144. Jacobson S, Pillus L. (1999) Modifying chromatin and concepts of cancer. Curr. Opin. Genet. Dev. 9: 175–184.

    Article  CAS  PubMed  Google Scholar 

  145. Rubnitz JE, Morrissey J, Savage PA, et al. (1994) ENL, the gene fused with HRX in t(11;19) leukemias, encodes a nuclear protein with transcriptional activation potential in lymphoid and myeloid cells. Blood 84: 1747–1752.

    CAS  PubMed  Google Scholar 

  146. Prasad R, Yano T, Sorio C, et al. (1995) Domains with transcriptional regulatory activity within the ALL1 and AF4 proteins involved in acute leukemia. Proc. Natl. Acad. Sci. U.S.A. 92: 12160–12164.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  147. Versteege I, Sevenet N, Lange J, et al. (1998) Truncating mutations of hSNF5/INI1 in aggressive paediatric cancer. Nature 394: 203–206.

    Article  CAS  PubMed  Google Scholar 

  148. Petrij F, Giles RH, Dauwerse HG, et al. (1995) Rubinstein-Taybi syndrome caused by mutations in the transcriptional co-activator CBP. Nature 376: 348–351.

    Article  CAS  PubMed  Google Scholar 

  149. Tanaka Y, Naruse I, Maekawa T, et al. (1997) Abnormal skeletal patterning in embryos lacking a single Cbp allele: a partial similarity with Rubinstein-Taybi syndrome. Proc. Natl. Acad. Sci. U.S.A. 94: 10215–10220.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  150. Kamei Y, Xu L, Heinzel T, et al. (1996) A CBP integrator complex mediates transcriptional activation and AP-1 inhibition by nuclear receptors. Cell 85: 403–414.

    Article  CAS  PubMed  Google Scholar 

  151. Gu W, Shi XL, Roeder RG. (1997) Synergistic activation of transcription by CBP and p53. Nature 387: 819–823.

    Article  CAS  PubMed  Google Scholar 

  152. Lill NL, Grossman SR, Ginsberg D, et al. (1997) Binding and modulation of p53 by p300/CBP coactivators. Nature 387: 823–827.

    Article  CAS  PubMed  Google Scholar 

  153. Arias J, Alberts AS, Brindle P, et al. (1994) Activation of cAMP and mitogen responsive genes relies on a common nuclear factor. Nature 370: 226–229.

    Article  CAS  PubMed  Google Scholar 

  154. Angel P, Karin M. (1991) The role of Jun, Fos and the AP-1 complex in cell-proliferation and transformation. Biochim. Biophys. Acta. 10: 2–3.

    Google Scholar 

  155. Lee JS, See RH, Deng T, et al. (1996) Adenovirus E1A downregulates cJun- and JunB-mediated transcription by targeting their coactivator p300. Mol. Cell. Biol. 16: 4312–4326.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  156. Yuan W, Condorelli G, Caruso M, et al. (1996) Human p300 protein is a coactivator for the transcription factor MyoD. J. Biol. Chem. 271: 9009–9013.

    Article  CAS  PubMed  Google Scholar 

  157. Chakravarti D, LaMorte VJ, Nelson MC, et al. (1996) Role of CBP/P300 in nuclear receptor signalling. Nature 383: 99–103.

    Article  CAS  PubMed  Google Scholar 

  158. DePinho RA. (1998) Transcriptional repression. The cancer-chromatin connection. Nature 391: 535–536.

    Article  CAS  Google Scholar 

  159. Luo RX, Postigo AA, Dean DC. (1998) Rb interacts with histone deacetylase to repress transcription. Cell 92: 463–473.

    Article  CAS  PubMed  Google Scholar 

  160. McMahon SB, Wood MA, Cole MD. (2000) The essential cofactor TRRAP recruits the histone acetyltransferase hGCN5 to c-Myc. Mol. Cell. Biol. 20: 556–562.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  161. Zhong H, Voll RE, Ghosh S. (1998) Phosphorylation of NF-kappa B p65 by PKA stimulates transcriptional activity by promoting a novel bivalent interaction with the coactivator CBP/p300. Mol. Cell 1: 661–671.

    Article  CAS  PubMed  Google Scholar 

  162. Merika M, Williams AJ, Chen G, et al. (1998) Recruitment of CBP/p300 by the IFN beta enhanceosome is required for synergistic activation of transcription. Mol. Cell 1: 277–287.

    Article  CAS  PubMed  Google Scholar 

  163. Ruddon RW. (1995). Genetic alterations in cancer cells. In: Cancer Biology. Oxford University Press, New York, pp 91–95.

    Google Scholar 

  164. Look AT. (1997) Oncogenic transcription factors in the human acute leukemias. Science 278: 1059–1064.

    Article  CAS  PubMed  Google Scholar 

  165. de The H, Chomienne C, Lanotte M, et al. (1990) The t(15;17) translocation of acute promyelocytic leukaemia fuses the retinoic acid receptor alpha gene to a novel transcribed locus. Nature 347: 558–561.

    Article  PubMed  Google Scholar 

  166. Borrow J, Goddard AD, Sheer D, et al. (1990) Molecular analysis of acute promyelocytic leukemia breakpoint cluster region on chromosome 17. Science 249: 1577–1580.

    Article  CAS  PubMed  Google Scholar 

  167. Kalantry S, Delva L, Gaboli M, et al. (1997) Gene rearrangements in the molecular pathogenesis of acute promyelocytic leukemia. J. Cell Physiol. 173: 288–296.

    Article  CAS  PubMed  Google Scholar 

  168. Najfeld V, Scalise A, Troy K. (1989) A new variant translocation 11;17 in a patient with acute promyelocytic leukemia together with t(7;12). Cancer Genet. Cytogenet. 43: 103–108.

    Article  CAS  PubMed  Google Scholar 

  169. Licht JD, Chomienne C, Goy A, et al. (1995) Clinical and molecular characterization of a rare syndrome of acute promyelocytic leukemia associated with translocation (11;17). Blood 85: 1083–1094.

    CAS  PubMed  Google Scholar 

  170. Li JY, English MA, Ball HJ, et al. (1997) Sequence-specific DNA binding and transcriptional regulation by the promyelocytic leukemia zinc finger protein. J. Biol. Chem. 272: 22447–22455.

    Article  CAS  PubMed  Google Scholar 

  171. Grimwade D, Solomon E. (1997) Characterisation of the PML/RAR alpha rearrangement associated with t(15;17) acute promyelocytic leukaemia. Curr. Top. Microbiol. Immunol. 220: 81–112.

    CAS  PubMed  Google Scholar 

  172. He LZ, Guidez F, Tribioli C, et al. (1998) Distinct interactions of PML-RARalpha and PLZF-RARalpha with corepressors determine differential responses to RA in APL. Nat. Genet. 18: 126–135.

    Article  CAS  PubMed  Google Scholar 

  173. Kurokawa R, Soderstrom M, Horlein A, et al. (1995) Polarity-specific activities of retinoic acid receptors determined by a co-repressor. Nature 377: 451–454.

    Article  CAS  PubMed  Google Scholar 

  174. Nakajima H, Kim YB, Terano H, et al. (1998) FR901228, a potent antitumor antibiotic, is a novel histone deacetylase inhibitor. Exp. Cell Res. 241: 126–133.

    Article  CAS  PubMed  Google Scholar 

  175. Richon VM, Emiliani S, Verdin E, et al. (1998) A class of hybrid polar inducers of transformed cell differentiation inhibits histone deacetylases. Proc. Natl. Acad. Sci. U.S.A. 95: 3003–3007.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  176. Ohki M. (1993) Molecular basis of the t(8;21) translocation in acute myeloid leukaemia. Semin. Cancer Biol. 4: 369–375.

    PubMed  CAS  Google Scholar 

  177. Nucifora G, Rowley JD. (1995) AML1 and the 8;21 and 3;21 translocations in acute and chronic myeloid leukemia. Blood 86: 1–14.

    PubMed  CAS  Google Scholar 

  178. Lenny N, Westendorf JJ, Hiebert SW. (1997) Transcriptional regulation during myelopoiesis. Mol. Biol. Rep. 24: 157–168.

    Article  PubMed  CAS  Google Scholar 

  179. Berger R, Le Coniat M, Romana SP, et al. (1996) Secondary acute myeloblastic leukemia with t(16;21) (q24;q22). involving the AML1 gene. Hematol. Cell Ther. 38: 183–186.

    Article  PubMed  CAS  Google Scholar 

  180. Shimada M, Ohtsuka E, Shimizu T, et al. (1997) A recurrent translocation, t(16;21) (q24;q22), associated with acute myelogenous leukemia: identification by fluorescence in situ hybridization. Cancer Genet. Cytogenet. 96: 102–105.

    Article  PubMed  CAS  Google Scholar 

  181. Gamou T, Kitamura E, Hosoda F, et al. (1998) The partner gene of AML1 in t(16;21) myeloid malignancies is a novel member of the MTG8(ETO) family. Blood 91: 4028–4037.

    PubMed  CAS  Google Scholar 

  182. Horsman DE, Gascoyne RD, Barnett MJ. (1995) Acute leukemia with structural rearrangements of chromosome 3. Leuk. Lymphoma 16: 369–377.

    Article  PubMed  CAS  Google Scholar 

  183. Secker-Walker LM, Mehta A, Bain B. (1995) Abnormalities of 3q21 and 3q26 in myeloid malignancy: a United Kingdom Cancer Cytogenetic Group study. Br. J. Haematol. 91: 490–501.

    Article  PubMed  CAS  Google Scholar 

  184. Speck NA, Terryl S. (1995) A new transcription factor family associated with human leukemias. Crit. Rev. Eukaryot. Gene Expr. 5: 337–364.

    Article  PubMed  CAS  Google Scholar 

  185. Meyers S, Hiebert SW. (1995) Indirect and direct disruption of transcriptional regulation in cancer: E2F and AML-1. Crit. Rev. Eukaryot. Gene Expr. 5: 365–383.

    Article  PubMed  CAS  Google Scholar 

  186. Romana SP, Poirel H, Leconiat M, et al. (1995) High frequency of t(12;21) in childhood B-lineage acute lymphoblastic leukemia. Blood 86: 4263–4269.

    PubMed  CAS  Google Scholar 

  187. Raynaud S, Cave H, Baens M, et al. (1996) The 12;21 translocation involving TEL and deletion of the other TEL allele: two frequently associated alterations found in childhood acute lymphoblastic leukemia. Blood 87: 2891–2899.

    PubMed  CAS  Google Scholar 

  188. Westendorf JJ, Yamamoto CM, Lenny N, et al. (1998) The t(8;21) fusion product, AML-1-ETO, associates with C/EBP-alpha, inhibits C/EBP-alpha-dependent transcription, and blocks granulocytic differentiation. Mol. Cell. Biol. 18: 322–333.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  189. Tanaka T, Tanaka K, Ogawa S, et al. (1995) An acute myeloid leukemia gene, AML1, regulates hemopoietic myeloid cell differentiation and transcriptional activation antagonistically by two alternative spliced forms. Embo J. 14: 341–350.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  190. Niitsu N, Yamamoto-Yamaguchi Y, Miyoshi H, et al. (1997) AML1a but not AML1b inhibits erythroid differentiation induced by sodium butyrate and enhances the megakaryocytic differentiation of K562 leukemia cells. Cell Growth Differ. 8: 319–326.

    PubMed  CAS  Google Scholar 

  191. Kitabayashi I, Yokoyama A, Shimizu K, et al. (1998) Interaction and functional cooperation of the leukemia-associated factors AML1 and p300 in myeloid cell differentiation. Embo J. 17: 2994–3004.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  192. Lutterbach B, Westendorf JJ, Linggi B, et al. (1998) ETO, a target of t(8;21) in acute leukemia, interacts with the N-CoR and mSin3 corepressors. Mol. Cell. Biol. 18: 7176–7184.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  193. Brehm A, Miska EA, McCance DJ, et al. (1998) Retinoblastoma protein recruits histone deacetylase to repress transcription. Nature 391: 597–601.

    Article  PubMed  CAS  Google Scholar 

  194. Magnaghi-Jaulin L, Groisman R, Naguibneva I, et al. (1998) Retinoblastoma protein represses transcription by recruiting a histone deacetylase. Nature 391: 601–605.

    Article  PubMed  CAS  Google Scholar 

  195. Mahlknecht U, Emiliani S, Najfeld V, et al. (1999) Genomic organization and chromosomal localization of the human histone deacetylase 3 gene. Genomics 56: 197–202.

    Article  PubMed  CAS  Google Scholar 

  196. Wlodarska I, De Wolf-Peeters C, Michaux L, et al. (1995) A new t(2;5) translocation in a null cell type CD30 positive anaplastic large cell lymphoma case. Leukemia 9: 1685–1688.

    PubMed  CAS  Google Scholar 

  197. Burkitt DP. (1971) Epidemiology of cancer of the colon and rectum. Cancer 28: 3–13.

    Article  PubMed  CAS  Google Scholar 

  198. Fuchs CS, Giovannucci EL, Colditz GA, et al. (1999) Dietary fiber and the risk of colorectal cancer and adenoma in women. N. Engl. J. Med. 340: 169–176.

    Article  PubMed  CAS  Google Scholar 

  199. Wargovich MJ, Levin B. (1996) Grist for the mill: role of cereal fiber and calcium in prevention of colon cancer. J. Natl. Cancer Inst. 88: 67–69.

    Article  PubMed  CAS  Google Scholar 

  200. Zoran DL, Turner ND, Taddeo SS, et al. (1997) Wheat bran diet reduces tumor incidence in a rat model of colon cancer independent of effects on distal luminal butyrate concentrations. J. Nutr. 127: 2217–2225.

    Article  PubMed  CAS  Google Scholar 

  201. Folino M, McIntyre A, Young GP. (1995) Dietary fibers differ in their effects on large bowel epithelial proliferation and fecal fermentation-dependent events in rats. J. Nutr. 125: 1521–1528.

    PubMed  CAS  Google Scholar 

  202. DeCosse JJ, Miller HH, Lesser ML. (1989) Effect of wheat fiber and vitamins C and E on rectal polyps in patients with familial adenomatous polyposis. J. Natl. Cancer Inst. 81: 1290–1297.

    Article  PubMed  CAS  Google Scholar 

  203. Weaver GA, Krause JA, Miller TL, et al. (1988) Short chain fatty acid distributions of enema samples from a sigmoidoscopy population: an association of high acetate and low butyrate ratios with adenomatous polyps and colon cancer. Gut 29: 1539–1543.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  204. Clausen MR, Bonnen H, Mortensen PB. (1991) Colonic fermentation of dietary fibre to short chain fatty acids in patients with adenomatous polyps and colonic cancer. Gut 32: 923–928.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  205. D’Argenio G, Cosenza V, Delle Cave M, et al. (1996) Butyrate enemas in experimental colitis and protection against large bowel cancer in a rat model. Gastroenterology 110: 1727–1734.

    Article  PubMed  Google Scholar 

  206. Medina V, Afonso JJ, Alvarez-Arguelles H, et al. (1998) Sodium butyrate inhibits carcinoma development in a 1,2-dimethylhydrazine-induced rat colon cancer. JPEN J. Parenter. Enteral Nutr. 22: 14–17.

    Article  PubMed  CAS  Google Scholar 

  207. Velazquez OC, Jabbar A, DeMatteo RP, et al. (1996) Butyrate inhibits seeding and growth of colorectal metastases to the liver in mice. Surgery 120: 440–447.

    Article  PubMed  CAS  Google Scholar 

  208. Sealy L, Chalkley R. (1978) The effect of sodium butyrate on histone modification. Cell 14: 115–121.

    Article  PubMed  CAS  Google Scholar 

  209. Boffa LC, Lupton JR, Mariani MR, et al. (1992) Modulation of colonic epithelial cell proliferation, histone acetylation, and luminal short chain fatty acids by variation of dietary fiber (wheat bran) in rats. Cancer Res. 52: 5906–5912.

    PubMed  CAS  Google Scholar 

  210. Cuisset L, Tichonicky L, Delpech M. (1998) A protein phosphatase is involved in the inhibition of histone deacetylation by sodium butyrate. Biochem. Biophys. Res. Commun. 246: 760–764.

    Article  CAS  PubMed  Google Scholar 

  211. Barnard JA, Warwick G. (1993) Butyrate rapidly induces growth inhibition and differentiation in HT-29 cells. Cell Growth Differ. 4: 495–501.

    CAS  PubMed  Google Scholar 

  212. Yoshida M, Beppu T. (1988) Reversible arrest of proliferation of rat 3Y1 fibroblasts in both the G1 and G2 phases by trichostatin A. Exp. Cell Res. 177: 122–131.

    Article  CAS  PubMed  Google Scholar 

  213. Archer SY, Meng S, Shei A, et al. (1998) p21(Waf1) is required for butyrate-mediated growth inhibition of human colon cancer cells. Proc. Natl. Acad. Sci. U. S. A. 95: 6791–6796.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  214. Archer SY, Hodin RA. (1999) Histone acetylation and cancer. Curr. Opin. Genet. Dev. 9: 171–174.

    Article  CAS  PubMed  Google Scholar 

  215. Kijima M, Yoshida M, Sugita K, et al. (1993) Trapoxin, an antitumor cyclic tetrapeptide, is an irreversible inhibitor of mammalian histone deacetylase. J. Biol. Chem. 268: 22429–22435.

    CAS  PubMed  Google Scholar 

  216. Yoshida M, Kijima M, Akita M, et al. (1990) Potent and specific inhibition of mammalian histone deacetylase both in vivo and in vitro by trichostatin A. J. Biol. Chem. 265: 17174–17179.

    CAS  PubMed  Google Scholar 

  217. Kwon HJ, Owa T, Hassig CA, et al. (1998) Depudecin induces morphological reversion of transformed fibroblasts via the inhibition of histone deacetylase. Proc. Natl. Acad. Sci. U.S.A. 95: 3356–3361.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  218. Kim YB, Lee KH, Sugita K, et al. (1999) Oxamflatin is a novel antitumor compound that inhibits mammalian histone deacetylase. Oncogene 18: 2461–2470.

    Article  CAS  PubMed  Google Scholar 

  219. Saito A, Yamashita T, Mariko Y, et al. (1999) A synthetic inhibitor of histone deacetylase, MS-27–275, with marked in vivo antitumor activity against human tumors. Proc. Natl. Acad. Sci. U.S.A. 96: 4592–4597.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  220. Samid D, Hudgins WR, Shack S, et al. (1997) Phenylacetate and phenylbutyrate as novel, nontoxic differentiation inducers. Adv. Exp. Med. Biol. 400A: 501–505.

    Article  PubMed  CAS  Google Scholar 

  221. Warrell RP, Jr., He LZ, Richon V, et al. (1998) Therapeutic targeting of transcription in acute promyelocytic leukemia by use of an inhibitor of histone deacetylase. J. Natl. Cancer Inst. 90: 1621–1625.

    Article  PubMed  CAS  Google Scholar 

  222. Lea MA, Tulsyan N. (1995) Discordant effects of butyrate analogues on erythroleukemia cell proliferation, differentiation and histone deace-tylase. Anticancer Res. 15: 879–883.

    PubMed  CAS  Google Scholar 

  223. Lea MA, Randolph VM, Hodge SK. (1999) Induction of histone acetylation and growth regulation in eryrthroleukemia cells by 4-phenylbutyrate and structural analogs. Anticancer Res. 19: 1971–1976.

    PubMed  CAS  Google Scholar 

  224. Saunders N, Dicker A, Popa C, et al. (1999) Histone deacetylase inhibitors as potential anti-skin cancer agents. Cancer Res. 59: 399–404.

    PubMed  CAS  Google Scholar 

  225. Engelhard HH, Duncan HA, Dal Canto M. (1997) Molecular characterization of glioblastoma cell differentiation. Neurosurgery 41: 886–896.

    Article  PubMed  CAS  Google Scholar 

  226. Huang H, Reed CP, Zhang JS, et al. (1999) Carboxypeptidase A3 (CPA3): a novel gene highly induced by histone deacetylase inhibitors during differentiation of prostate epithelial cancer cells. Cancer Res. 59: 2981–2988.

    PubMed  CAS  Google Scholar 

  227. Ellerhorst J, Nguyen T, Cooper DN, et al. (1999) Induction of differentiation and apoptosis in the prostate cancer cell line LNCaP by sodium butyrate and galectin-1. Int. J. Oncol. 14: 225–232.

    PubMed  CAS  Google Scholar 

  228. Keohane AM, LP On, Belyaev ND, et al. (1996) X-Inactivation and histone H4 acetylation in embryonic stem cells. Dev. Biol. 180: 618–630.

    Article  PubMed  CAS  Google Scholar 

  229. Braunstein M, Sobel RE, Allis CD, et al. (1996) Efficient transcriptional silencing in Saccharomyces cerevisiae requires a heterochromatin histone acetylation pattern. Mol. Cell Biol. 16: 4349–4356.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  230. Moens PB. (1995) Histones H1 and H4 of surface-spread meiotic chromosomes. Chromosoma 104: 169–174.

    Article  CAS  PubMed  Google Scholar 

  231. Lee DY, Hayes JJ, Pruss D, et al. (1993) A positive role for histone acetylation in transcription factor access to nucleosomal DNA. Cell 72: 73–84.

    Article  CAS  PubMed  Google Scholar 

  232. Mutskov V, Gerber D, Angelov D, et al. (1998) Persistent interactions of core histone tails with nucleosomal DNA following acetylation and transcription factor binding. Mol. Cell Biol. 18: 6293–6304.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  233. Vettese-Dadey M, Grant PA, Hebbes TR, et al. (1996) Acetylation of histone H4 plays a primary role in enhancing transcription factor binding to nucleosomal DNA in vitro. Embo J. 15: 2508–2518.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  234. Rundlett SE, Carmen AA, Suka N, et al. (1998) Transcriptional repression by UME6 involves deacetylation of lysine 5 of histone H4 by RPD3. Nature 392: 831–835.

    Article  CAS  PubMed  Google Scholar 

  235. Ait-Si-Ali S, Ramirez S, Barre FX, et al. (1998) Histone acetyltransferase activity of CBP is controlled by cycle-dependent kinases and oncoprotein E1A. Nature 396: 184–186.

    Article  CAS  PubMed  Google Scholar 

  236. Cosma MP, Tanaka T, Nasmyth K. (1999) Ordered recruitment of transcription and chromatin remodeling factors to a cell cycle- and developmentally regulated promoter. Cell 97: 299–311.

    Article  CAS  PubMed  Google Scholar 

  237. Kao HY, Downes M, Ordentlich P, et al. (2000) Isolation of a novel histone deacetylase reveals that class I and class II deacetylases promote SMRT-mediated repression. Genes Dev. 14: 55–66.

    PubMed  PubMed Central  CAS  Google Scholar 

  238. Akiyoshi S, Inoue H, Hanai J, et al. (1999) c-Ski acts as a Transcriptional Co-repressor in Transforming Growth Factor-beta Signaling through interaction with Smads. J. Biol. Chem. 274: 35269–35277.

    Article  CAS  PubMed  Google Scholar 

  239. Torchia J, Glass C, Rosenfeld MG. (1998) Co-activators and co-repressors in the integration of transcriptional responses. Curr. Opin. Cell Biol. 10: 373–383.

    Article  CAS  PubMed  Google Scholar 

  240. Davie JR, Spencer VA. (1999) Control of histone modifications. J. Cell Biochem. 33: 141–148.

    Article  Google Scholar 

  241. Bauer A, Mikulits W, Lagger G, et al. (1998) The thyroid hormone receptor functions as a ligand-operated developmental switch between proliferation and differentiation of erythroid progenitors. Embo J. 17: 4291–4303.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  242. Bartl S, Taplick J, Lagger G, et al. (1997) Identification of mouse histone deacetylase 1 as a growth factor-inducible gene. Mol. Cell Biol. 17: 5033–5043.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  243. Allard S, Utley RT, Savard J, et al. (1999) NuA4, an essential transcription adaptor/histone H4 acetyltransferase complex containing esa1p and the ATM-related cofactor tra1p. Embo J. 18: 5108–5119.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  244. Clarke AS, Lowell JE, Jacobson SJ, et al. (1999) Esa1p is an essential histone acetyltransferase required for cell cycle progression. Mol. Cell Biol. 19: 2515–2526.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  245. Zhang W, Bone JR, Edmondson DG, et al. (1998) Essential and redundant functions of histone acetylation revealed by mutation of target lysines and loss of the Gcn5p acetyltransferase. Embo J. 17: 3155–3167.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  246. Dunphy EL, Johnson T, Auerbach SS, et al. (2000) Requirement for TAF(II)250 Acetyltransferase Activity in Cell Cycle Progression. Mol. Cell Biol. 20: 1134–1139.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  247. Kasten MM, Dorland S, Stillman DJ. (1997) A large protein complex containing the yeast Sin3p and Rpd3p transcriptional regulators. Mol. Cell Biol. 17: 4852–4858.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  248. Kim S, Benguria A, Lai CY, et al. (1999) Modulation of life-span by histone deacetylase genes in Saccharomyces cerevisiae. Mol. Biol. Cell 10: 3125–3136.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

Download references

Acknowledgments

We wish to apologize to those investigators whose relevant work was not discussed or cited directly in this manuscript due to space limitations. This work was partly supported by the German National Science Foundation (Deutsche Forschungsgemeinschaft, MA 2057/1-1) and institutional funds from the University of Frankfurt Medical Center.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ulrich Mahlknecht MD, PhD.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Mahlknecht, U., Hoelzer, D. Histone Acetylation Modifiers in the Pathogenesis of Malignant Disease. Mol Med 6, 623–644 (2000). https://doi.org/10.1007/BF03402044

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/BF03402044

Keywords