Literature
Home医源资料库在线期刊分子药理学杂志2006年第68卷第4期

Rational Development of Histone Deacetylase Inhibitors as Anticancer Agents: A Review

来源:分子药理学杂志
摘要:HistoneAcetylasesandDeacetylases:ClassificationandFunctionTheequilibriumofsteadystateacetylationistightlycontrolledbytheantagonisticeffectofbothHATsandHDACs,whichinturnregulatestranscriptionstatusofnotjusthistonesbutalsoofothersubstratessuchasp53(Grozing......

点击显示 收起

    Clinical Pharmacology Research Core, National Cancer Institute, National Institutes of Health, Bethesda, Maryland (M.R.A., A.S., W.D.F.)
    Department of Pharmaceutics, School of Pharmacy, Virginia Commonwealth University, Richmond, Virginia (M.R.A., J.V.)

    Abstract

    The epigenome is defined by DNA methylation patterns and the associated post-translational modifications of histones. This histone code determines the expression status of individual genes dependent upon their localization on the chromatin. The histone deacetylases (HDACs) play a major role in keeping the balance between the acetylated and deacetylated states of chromatin and eventually regulate gene expression. Recent developments in understanding the cancer cell cycle, specifically the interplay with chromatin control, are providing opportunities for developing mechanism-based therapeutic drugs. Inhibitors of HDACs are under considerable exploration, in part because of their potential roles in reversing the silenced genes in transformed tumor cells by modulating transcriptional processes. This review is an effort to summarize the nonclinical and clinical status of HDAC inhibitors currently under development in anticancer therapy.

    In eukaryotic cells, DNA has been conserved over evolution in a condensed and densely packed higher order structure called chromatin. Chromatin, present in the interphase nucleus, is composed of regular repeating units of nucleosomes, which represent the principal protein-nucleic acid relationship. The major components of chromatin are nucleic acids (DNA and RNA), which are negatively charged; associated proteins, including histones, that are positively charged at neutral pH; and nonhistone chromosomal proteins, which are acidic at neutral pH. Within the nucleus, chromatin can exists in two different forms: heterochromatin, which is highly compact and transcriptionally inactive, or euchromatin, which is loosely packed and accessible to RNA polymerases for involvement in transcriptional processes and gene expression. A nucleosome is a complex of 146 nucleotide base pairs of DNA wrapped around the core histone octamer that helps organize chromatin. The histone octamer is composed of two copies each of H2A, H2B, H3, and H4 proteins, which are very basic, mainly because of positively charged amino-terminal side chains rich in the amino acid lysine. Post-translational and other changes in chromatin, such as acetylation/deacetylation at lysine residues, methylation at lysine or arginine residues, phosphorylation at serine resides, ubiquitylation at lysines, and/or ADP ribosylation, are mediated by chemical modification of various sites on N-terminal tail (Marks et al., 2000, 2003, 2004).

    The structural modification of histones is regulated mainly by acetylation/deacetylation of the N-terminal tail and is crucial in modulating gene expression, because it affects the interaction of DNA with transcription-regulatory non-nucleosomal protein complexes. The balance between the acetylated/deacetylated states of histones is mediated by two different sets of enzymes: histone acetyltransferases (HATs) and histone deacetylases (HDACs). HATs preferentially acetylate specific lysine substrates among other nonhistone protein substrates and transcription factors, affecting DNA-binding properties and, in turn, altering gene transcription. HDACs restore the positive charge on lysine residues by removing acetyl groups and thus are involved primarily in the repression of gene transcription by compacting chromatin structure. Thus, open lysine residues attach firmly to the phosphate backbone of the DNA, preventing transcription. In this tight conformation, transcription factors, regulatory complexes, and RNA polymerases cannot bind to the DNA. Acetylation relaxes the DNA conformation, making it accessible to transcription machinery. High levels of acetylation of core histones are seen in chromatin-containing genes, which are highly transcribed genes; genes that are silent are associated with low levels of acetylation. Inappropriate silencing of critical genes can result in one or both hits of tumor suppressor gene inactivation in cancer; theoretically, therefore, the reactivation of affected TSGs could have an enormous therapeutic value in preventing and treating cancer (Thiagalingam et al., 2003).

    Histone Acetylases and Deacetylases: Classification and Function

    The equilibrium of steady state acetylation is tightly controlled by the antagonistic effect of both HATs and HDACs, which in turn regulates transcription status of not just histones but also of other substrates such as p53 (Grozinger and Schreiber, 2002). Several groups of proteins with HAT activity have been identified, including the GNAT (Gcn5-related N-acetyl transferase) family, the MYST (monocytic leukemia zinc finger protein) group, TIP60 (TAT-interactive protein), and the p300/CREB-binding protein family. HATs act as large multiprotein complexes containing other HATs, coactivators for transcription factors, and corepressors (Annunziato and Hansen, 2000; Chen et al., 2001; Gregory et al., 2001; Nakatani, 2001; Schreiber and Bernstein, 2002). HATs, which bind nonhistone protein substrates and transcription factors, have been called factor acetyltransferases. Acetylation of these transcription factors also affects their DNA binding properties and gene transcription (Roth et al., 2001; Gregoretti et al., 2004). HAT genes may be overexpressed, translocated, or mutated in both hematological and epithelial cancers (Mahlknecht and Hoelzer, 2000; Timmermann et al., 2001; Johnstone and Licht, 2003). Translocations of HATs, CREB-binding protein, and p300 acetyltransferases into genes have given rise to various hematological malignancies (Fenrick and Hiebert, 1998; Pandolfi, 2001).

    Mammalian HDACs are divided into three major groups or classes based on their structural homologies to the three distinct yeast HDACs: Rpd3 (class I), Hda1 (class II), and Sir2/Hst (class III). Class III HDACs consist of the large family of sirtuins [silent information regulators (SIRs)] that are evolutionarily distinct, with a unique enzymatic mechanism dependent on the cofactor NAD+, and are virtually unaffected by all HDAC inhibitors in current development (Imai et al., 2000; Gray and Ekstrom, 2001). Class I and II HDACs contain an active site zinc as a critical component of the enzymatic pocket, have been extensively described to have an association with cancers, and are believed to be comparably inhibited by all HDAC inhibitors in development. The Rpd3 homologous class I HDACs 1, 2, 3, and 8 are widely expressed in tissues and are primarily localized in the nucleus. Hda1 homologous class II HDACs 4, 5, 6, 7, 9a, 9b, and 10 are much larger in size, display limited tissue distribution, and can shuttle between the nucleus and cytoplasm, suggesting functions and cellular substrates different from Class I HDACs (Kao et al., 2001; Guardiola and Yao, 2002). HDACs 6 and 10 are unique in that they have two catalytic domains, whereas HDACs 4, 8, and 9 are expressed to greater extent in tumor tissues and have been shown to be specifically involved in differentiation (de Ruijter et al., 2003). There is some evidence that certain inhibitors display a variable degree of HDAC specificity; hence, it is imperative to identify differences in HDAC functions to better target and tailor specific drugs compounds (Jung, 2001; Grozinger and Schreiber, 2002; Miller et al., 2003; Heltweg et al., 2004). HDACs usually interacts with large protein complexes that down-regulate genes through association with corepressors [such as nuclear receptor corepressor (NcoR) and silencing mediator for retinoid and thyroid hormone receptor (SMRT)] transcription factors, estrogen receptors (ER), p53, cell cycle-specific regulators [such as retinoblastoma (Rb), E2F and other HDACs], as well as histones, but they can also bind to their receptor directly (Frye, 2000; Imai et al., 2000; Zhou et al., 2002).

    Class III HDACs (sirtuins, SIR T1, 2, 3, 4, 5, 6, and 7) are not inhibited by class I and II HDAC inhibitors; instead, they are inhibited by nicotinamide (vitamin B3). Nicotinamide inhibits a NAD-dependent p53 deacetylation induced by SIR2 and also enhances p53 acetylation levels in vivo (Luo et al., 2001). It has been shown recently that by restraining mammalian forkhead proteins, specifically foxo3a, SIRT1 also reduces apoptosis (Motta et al., 2004). The inhibition of forkhead activity by SIRT1 parallels the effect of this deacetylase on the tumor suppressor p53. These results have significant implications regarding an important role for Sirtuins in modulating the sensitivity of cells in p53-dependent apoptotic response and the possible effect in cancer therapy (Schwer et al., 2002; North et al., 2003).

    Chromatin Modification and Cancer. DNA gene expression is controlled by an assembly of nucleoproteins that includes histones and other architectural components of chromatin, nonhistone DNA-bound regulators, and additional chromatin-bound polypeptides. Changes in growth and differentiation leading to malignancy seem to occur by alterations in transcriptional control and gene silencing. It is becoming increasingly apparent that imbalances of both DNA methylation and histone acetylation may play an important role in cancer development and progression (Marks et al., 2001, 2004; Timmermann et al., 2001; Jones and Baylin, 2002). Unlike normal cells, in cancer, changes in genome expression are associated with the remodeling of long regions of regulatory DNA, including promoters, enhancers, locus control regions, and insulators, into specific chromatin architecture. These specific changes in the DNA architecture result in a general molecular signature for a type of cancer and complement its DNA methylation-based component. The changes in the infrastructure of chromatin over a target promoter are more profound than those observed by these enzymes acting independently (Wade, 2001; Davis and Brackmann, 2003). Apart from acetylation, histone tails undergo other modifications, including methylation, phosphorylation, ubiquitylation, and adenosine diphosphate ribosylation. These other areas of modifications ("histone code") have not yet been sufficiently explored to identify their roles in epigenetic modifications (Bhalla and List, 2004).

    Disruption of HAT and HDAC function is associated with the development of cancer, and malignant cells target chromatin-remodeling pathways as a means of disrupting transcriptional regulation (Mahlknecht and Hoelzer, 2000). Of the various hypotheses describing deregulation mechanisms, three have been put forth frequently: 1) disordered hyperacetylation could activate promoters that are normally repressed, leading to inappropriate expression of proteins, 2) abnormally decreased acetylation levels of promoter regions could repress the expression of genes necessary for a certain phenotype, and 3) mistargeted or aberrant recruitment of HAT/HDAC activity could act as a pathological trigger. Even though no direct alterations in HDAC genes have been demonstrated in cancer, the association of HDACs with various oncogenes and tumor suppressor genes is now well established, as is the potential for HDAC involvement in tumorigenesis (Kristeleit et al., 2004).

    Histone Deacetylase Inhibitors As Anticancer Agents. The discovery of recruitment of HDAC enzymes in cancer has provided a rationale for using inhibition of HDAC activity to release transcriptional repression as a viable option toward achieving eventual therapeutic benefit (Johnstone and Licht, 2003). Inhibition of HDAC function can release dysregulation of genes involved in cell cycle progression, differentiation, and apoptosis. HDAC inhibitors block deacetylation function, causing cell cycle arrest, differentiation, and/or apoptosis of many tumors (Pandolfi, 2001). Several HDAC inhibitors have exhibited potent antitumor activity in human xenograft models, suggesting their usefulness as novel cancer therapeutic agents. Several are currently in phase I/II clinical trials both in hematological malignancies and in solid tumors. Compared with agents used initially, some newer agents are effective at nanomolar concentrations and are relatively less toxic. A wide range of structures inhibits activity of class I/II HDAC enzymes; with a few exceptions, these can be divided into structural classes, including 1) carboxylates (short-chain fatty acids), 2) small-molecule hydroxamates, 3) electrophilic ketones (epoxides), 4) cyclic peptides, 5) benzamides, and 6) other hybrid compounds (Drummond et al., 2004). Table 1 describes the various compounds, their activities in cell lines and preclinical murine models, and their current clinical status.

    Aliases can be found in the text and in the abbreviations list.

    Comprehensive reviews on structure, medicinal chemistry, and structure-activity relationships of more than 80 different HDAC inhibitors and analogs have been previously published or reviewed (Curtin et al., 2002; Kouraklis and Theocharis, 2002; Remiszewski, 2002, 2003; Remiszewski et al., 2002; Arts et al., 2003; Bouchain and Delorme, 2003; Bouchain et al., 2003; Curtin and Glaser, 2003; Kim et al., 2003; Miller et al., 2003; Drummond et al., 2004; Kristeleit et al., 2004; Plumb et al., 2004). Despite the variety of structural distinctiveness, all of these HDAC inhibitors can be broadly characterized by a common pharmacophore that includes key elements of inhibitor-enzyme interactions (Miller et al., 2003). Most of these compounds were designed to have three basic components: a hydrophobic cap that blocks the entrance to active site, a polar site, and a hydroxamic acid type zinc-binding active site separated by a hydrophobic spacer that has optimal length spanning the hydrophobic pocket on the enzyme (Drummond et al., 2004).

    Short-Chain Fatty Acids

    Dimethyl sulfoxide was one of the first compounds identified as active in transformed cell differentiation. Because of this, several compounds were synthesized and screened for activity in differentiation, growth arrest, and/or apoptosis (Marks et al., 2001). Valproic acid is effective in vitro as an HDAC inhibitor at relatively high (millimolar) concentrations and has much weaker affinity. It has been shown to selectively induce proteasomal degeneration of HDAC2 and is antiangiogenic in vitro and in vivo (Kramer et al., 2003; Eyal et al., 2004; Michaelis et al., 2004). Valproic acid, a well established anticonvulsant for seizures and bipolar disorders, has been shown to have antigrowth activity of human endometrial cells and to inhibit proliferation and induce apoptosis in acute myeloid leukemia cells expressing ABCB1 (P-glycoprotein) and the multidrug resistance protein ABCC1 (MRP1) (Gurvich et al., 2004; Takai et al., 2004; Tang et al., 2004). Valproic acid has recently been shown to inhibit angiogenesis in vitro and in vivo and markedly effects genes relevant in proliferation and apoptosis (Michaelis et al., 2004; Thelen et al., 2004).

    Phenyl acetate (PA) can penetrate the central nervous system; when tested in solid tumors, it showed antitumor effects mediated by histone acetylation. PA is a metabolite of phenylbutyrate (PB) after B-oxidation in the liver and kidney (Piscitelli et al., 1995; Carducci et al., 1996). PB, a well studied member of the short-chain fatty acids, can arrest cells in G1-G0 by inducing p21WAF1 and other cdk-2-associated cell cycle proteins, alter levels of expression and activation of chemotaxis proteins, such as urokinase-plasminogen activator, induce apoptosis, inhibit telomerase, and increase major histocompatibility complex class I expression in various tumor models (Gore and Carducci, 2000). However, the short-chain fatty acids have a low potency because of their short side chains, limiting their contact with the catalytic pocket of HDACs (Johnstone, 2002). In human CCRF-CEM cells, acute T-lymphoblastic leukemia cells, butyrate and other HDAC inhibitors caused G2/M cell cycle arrest and apoptotic cell death (Bernhard et al., 1999). Butyrates induce histone acetylation and granulocyte maturation in acute myeloid leukemia (AML) and selectively inhibit growth in human prostate cancer and cervical carcinoma cells (Finzer et al., 2003; Gozzini et al., 2003; Kuefer et al., 2004). Butyrates have been under extensive clinical evaluation in both hematological malignancies and solid tumors. Butanoic acid or its prodrug pivaloyl oxymethyl butyrate (AN-9) is currently undergoing clinical trials after it showed 10-fold more potent activity than SB in leukemia tumor cell lines (Batova et al., 2002; Patnaik et al., 2002; Reid et al., 2004). The antineoplastic activity of AN-9 stems from rapid hydrolysis and release of butyrate, permitting efficient delivery to subcellular targets (Zimra et al., 1997, 2000). Despite the overall weak activity of short-chain fatty acids, several agents have been studied clinically because of their use for alternative medical conditions (Melchior et al., 1999; Gore et al., 2001, 2002).

    Hydroxamic Acids

    This is the broadest class of inhibitors with high affinity for HDAC that has been shown to inhibit both class I and II HDACs. Inhibitors containing hydroxamic acid (HA) residues bind with high affinity to the HDAC catalytic site, blocking the access of the substrate to the zinc ion (Finnin et al., 1999). The general structure of these substances consists of a hydrophobic linker that allows the hydroxamic acid moiety to chelate the cation at the bottom of the HDAC catalytic pocket while the bulky part of the molecule acts as a cap for the tube. Most of the chemicals in this group are very potent (functioning at nanomolar to micromolar concentrations in vitro) but are reversible inhibitors of class I/II HDACs.

    Trichostatin A (TSA) was one of the first HDAC inhibitors to be described and is widely used as a reference in research in this field (Yoshida et al., 1995, 2001). It was developed as an antifungal agent but is relatively unstable; its toxicity to patients and lack of specificity for certain HDACs motivated the search for other substances (Jung et al., 1999; Jung, 2001). The design of many synthetic drugs has been inspired by the TSA structure (the aromatic "cap", hydroxamic acid functionality, and hydrophobic linker between them). TSA blocks proliferation and triggers apoptosis in hepatocellular carcinoma cells, blocks cell cycle progression in HeLa cells, and differentiation in ovarian cancer cells by changing p21 tumor suppressor gene and DNA-binding Id1 protein (Hoshikawa et al., 1994; Herold et al., 2002; Strait et al., 2002). TSA has also been shown to suppress growth of pancreatic adenocarcinoma cells and ACHN renal cell carcinoma via cell cycle arrest in association with p27 or apoptosis (Donadelli et al., 2003; Park et al., 2003). TSA is more sensitive in estrogen receptor -positive breast cancer cells in inhibiting HDAC (Margueron et al., 2003).

    Simple hydroxamic acid derivatives such as suberoylanilide hydroxamic acid (SAHA) and pyroxamide have activity at submicromolar concentrations (Richon et al., 1998, 2001; Marks, 2004). SAHA is a second-generation polar-planar compound that induces growth arrest, differentiation, and/or apoptosis and is under clinical investigation in both hematological and nonhematological malignancies (Richon et al., 1996, 1998; Coffey et al., 2000; Munster et al., 2001). In studies with breast cancer cells, SAHA inhibited clonogenic growth and induced apoptosis, whereas in malignant human hemotopeoitic cells, SAHA induced marked toxicity but showed relatively minor maturation activity (Vrana et al., 1999; Huang and Pardee, 2000). SAHA also showed antiproliferative and pro-apoptotic actions in several mouse xenografts and cancer cells, including prostate, bladder carcinoma, and myeloma. SAHA also induced the CDK inhibitor p21WAF1/Cip1, and the inhibitory activity was independent of p53 status (Butler et al., 2000, 2002; Richon et al., 2000; Cohen et al., 2002; Gui et al., 2004). Pyroxamide is another compound in this class that induced terminal differentiation in murine erythroleukemic cells and caused growth inhibition in prostate carcinoma, bladder, and neuroblastoma cells via apoptosis (Butler et al., 2001; Kouraklis and Theocharis, 2002; Kutko et al., 2003). In experiments with SAHA and butyrates, a model was proposed in which induction of apoptosis in Bcr/Abl+ cells by HDAC inhibitors involves coordinate inactivation of the cytoprotective Raf/MEK/ERK pathway in conjunction with the reactive oxygen species-dependent activation of JNK (Yu et al., 2003c).

    Oxamflatin is another compound in the same class that induces transcriptional activation of junD, causing cell cycle arrest and morphological changes similar to those caused by TSA (Kim et al., 1999). Scriptaid was found to be one of the most potent analogs in a random search for substances that augment signal transduction pathways and, when screened in human and animal tumor cells, showed antiproliferative effects similar to those of SAHA (Su et al., 2000; Bouchain and Delorme, 2003). NVP-LAQ824, a cinnamic HA, has been shown to inhibit HDAC in vitro and cause transcriptional activation of p21 promoter in reporter gene assays at submicromolar concentrations in multiple myeloma (Catley et al., 2003). NVP-LAQ824, like most other HDAC inhibitors, was selective in its action because it required longer exposure and higher concentrations to retard growth of normal human fibroblasts (Atadja et al., 2004a). Another HA analog, suberic bishydroxamate, was shown to regulate expression of multiple apoptotic mediators and to induce mitochondria-dependent apoptosis in melanoma cells (Zhang et al., 2004b). PXD101 is a novel hydroxamate-type inhibitor of HDAC activity in nanomolar ranges in leukemia cells. It was shown to delay growth for xenografts of cisplatin-resistant ovarian tumor cells and had marked increase in acetylation of histone and showed good antitumor activity (Plumb et al., 2003).

    Newer compounds, such as cyclic HA peptides, structural combinations of a HA (such as TSA) and a cyclic tetrapeptide (such as trapoxin), inhibit isoform selective HDACs at nanomolar concentrations (Furumai et al., 2001; Nishino et al., 2003). One of the cyclic HA peptide derivatives inhibited growth in four of five human tumor lines implanted into nude mice and showed great promise as a therapeutic agent with higher selective inhibition of HDAC (Komatsu et al., 2001).

    Cyclic Peptides

    Cyclic peptides having epoxyketone (epoxides) may act by chemically modifying an active site nucleophile with the epoxy group and forming H-bonds with ketone. These chemicals are supposed to trap HDACs through the reaction of the epoxide moiety with the zinc cation or an amino acid (forming a covalent attachment) in the binding pocket. However, the lability of the epoxide functionality prevents significant in vivo activity, which makes them of little pharmacologic interest. The only HDAC inhibitors in this set of compounds are a number of natural products with significant in vitro activity, such as trapoxin (TPX) A and B, depudecin, and 2-amino-8-oxo-9,10-epoxydecanoic acid. TPX is a hybrid molecule containing cyclic peptide (acts as hydrophobic cap) and epoxyketone moiety that has shown irreversible inhibition of mammalian HDACs at nanomolar ranges (Kijima et al., 1993; Kosugi et al., 1999; Komatsu et al., 2001). Cyclic tetrapeptides such as apicidin, which has an ethyl ketone moiety, and FK228 (FR901228, also referred to as depsipeptide) inhibit HDACs at nanomolar concentrations. Apicidin is a fungal metabolite that is able to inhibit HDACs and proliferation of tumor cells via induction of p21WAF1/Cip1 and gelsolin (Han et al., 2000). It is postulated that apicidin interacts with the catalytic site and has been shown to inhibit cell proliferation in several human cancer cell lines because of its anti-invasive and antiangiogenic activity (Meinke et al., 2000; Singh et al., 2001, 2002; Hong et al., 2003; Kim et al., 2004). FK228 is a natural product derived from Chromobacterium violaceum that exhibits potent antitumor activity through currently unknown mechanism of action (Piekarz and Bates, 2004). One hypothesis proposes that the disulfide bridge is reduced inside the cell or organism and the 4-mercaptobut-1-enyl residue then fits inside the HDAC catalytic pocket, chelating Zn2+ in a manner similar to that of other inhibitors. In cultured cells, it is able to induce histone hyperacetylation and growth arrest at nanomolar concentrations. In human leukemia cells, FK228 had an IC50 at nanomolar concentrations and induced apoptosis in cells from patients with chronic lymphocytic leukemia (Byrd et al., 1999; Sasakawa et al., 2002, 2003a,b). In addition, depsipeptide has been shown to be antiangiogenic by modulating expression of c-myc and other regulatory genes (Kwon et al., 2002). FK228 is currently undergoing extensive evaluation in clinical trials (Kwon et al., 2002; Marshall et al., 2002; Sandor et al., 2002; Byrd et al., 2004).

    Benzamides

    The synthetic benzamide derivatives include a structurally diverse group of compounds such as MS-275 and CI-994. CI-994 has shown efficacy in solid tumors in murine models but did not inhibit HDAC directly. The mechanism of its action is unknown, but it seems to inhibit both histone deacetylation and cellular proliferation at the G1-S phase transition (LoRusso et al., 1996; Graziano et al., 1997; Prakash et al., 2001). MS-275 and some of its derivatives inhibit HDACs in vitro at micromolar concentrations, but the mechanism is not clearly understood. It is believed that the diaminophenyl group is very important for the inhibitory behavior; probably, both amino functionalities chelate the metallic ion in the catalytic site. MS-275eCassociated HDAC-inhibitory activity is accompanied by an increase in expression of cyclin-dependent kinase inhibitor p21WAF1/Cip1 and accumulation in G1 phase (Saito et al., 1999; Suzuki et al., 1999). MS-275 displays antiproliferative activity in several human cancer cell lines, including breast, colorectum, leukemia, lung, ovary, and pancreas. MS-275 suppressed growth of several pediatric cancer cell lines in a dose-dependent manner, as well as tumors transplanted in nude mice (Jaboin et al., 2002). MS-275 and CI-994 are undergoing clinical trials. There are reports of novel nonhydroxamate sulfonamide anilides similar in structure to MS-275 being synthesized that have shown lower toxicity and comparable antiproliferative activity (Fournel et al., 2002; Bouchain et al., 2003). Focus is on the development of novel compounds based on core structures of HA or benzamide platform, which may have a better HDAC inhibitory profile and lower toxicity compared with parent compounds.

    Mode of Action of HDAC Inhibitors in Cancer Cells. Even though a number of HDAC inhibitors have shown considerable promise in preclinical models, the mechanism of action has not been fully evaluated. HDAC inhibitors are effective in affecting cell cycle arrest, apoptosis, antiangiogenesis, and differentiation in cultured and transformed cells from both hematological (leukemias, lymphomas, and myelomas) and epithelial (breast, bladder, ovary, prostate, and lung) tumor sources. The change that occurs after treatment with HDAC inhibitors (growth arrest, terminal differentiation, or apoptosis) seems to be dependent upon the tumor cell rather than on the specific HDAC inhibitors used (Bhalla and List, 2004). The HDAC family is divided into Zn-dependent (class I and II) and Zn-independent/NAD-dependent (class III) enzymes. The Zn-dependent enzymes have been the focus of intense research, whereas class III has been recently implicated in acetylation and regulation of key cell cycle proteins such as p53 (Cheng et al., 2003; McLaughlin and La Thangue, 2004). It is interesting that a number of studies have showed that HDAC inhibitors are relatively nontoxic to normal cells or tissues but exhibit selective cytotoxicity against a wide range of cancer cells (Zhu et al., 2001a; Rosato and Grant, 2003). It has been postulated that defective cell cycle checkpoint regulation of neoplastic cells may render them susceptible to HDAC inhibition-induced apoptosis (Johnstone and Licht, 2003; Warrener et al., 2003).

    As noted above, histone acetylation is known to precede gene transcription; among the genes that are consistently up-regulated because their promoters are associated with acetylated histones is the cell cycle gene CDKN1A, which encodes cyclin-dependent kinase inhibitor p21WAF1. Cyclin-dependent kinase inhibitor WAF1 inhibits cell-cycle progression by blocking cyclin-dependent kinase activity and the arrest of the cell cycle in G1 stage. Most HDAC inhibitors (i.e., butyrates, TSA, depsipetide, oxamflatin, MS-275, and SAHA) induce expression of p21 (Archer et al., 1998; Saito et al., 1999; Sowa et al., 1999a,b; Vrana et al., 1999; Chai et al., 2000; Han et al., 2000; Huang et al., 2000; Sandor et al., 2000; Siavoshian et al., 2000; Lavelle et al., 2001; Blagosklonny et al., 2002; Wang et al., 2002). Some cDNA microarray studies have shown that treatment with TSA or SAHA alters the expression of a selective subset of approximately 2% of cellular genes that are either up- or down-regulated (Mitsiades et al., 2003; Chiba et al., 2004a,b). The genes that are usually affected by these inhibitors are CDKN1A and CDKN2A; the latter encodes genes of cell cycle regulation, such as p16, cyclin E, and thioredoxin binding protein 2 (Huang and Pardee, 2000; Kim et al., 2000). Thus, gene promoters have specific sites, such as SP1, that bind HDAC-containing transcription complexes and repress gene transcription (Li and Wu, 2004; Yokota et al., 2004). Inhibition of HDACs will activate these silenced genes, contributing to growth arrest, differentiation, and/or apoptosis of transformed cells. Treatment with HDAC inhibitors triggers both the intrinsic and sensitizes tumor cells to the death ligands that initiate the extrinsic pathway of apoptosis (Bhalla and List, 2004). Several HDAC inhibitors, including SB, SAHA, and MS-275, induce mitochondrial permeability transition, in which pro-apoptotic molecules, such as cytochrome c, are released into the cytosol, resulting in eventual activation of caspase-dependent apoptotic cascades (both receptor- and mitochondria-mediated) (Rosato et al., 2001; Aron et al., 2003; Nguyen et al., 2003; Guo et al., 2004). Up-regulation and induction of a conformational change of the pro-apoptotic proteins are some of the HDAC inhibitor-induced upstream events that may trigger the mitochondrial pathway of apoptosis, as described for MS-275 and SB or, as proposed in case of SAHA, may not require key caspases such as caspase-8 and caspase-3 (Ruefli et al., 2001; Lucas et al., 2004). Reactive oxygen species have recently been identified as a major cell death mechanism of several HDAC inhibitors (Ruefli et al., 2001; Rosato et al., 2003). There is some evidence that HDAC inhibitors may induce acetylation of nonhistone proteins, such as the 90-kDa heat shock protein (hsp90). Depsipeptide, SAHA, and LAQ824 lower the threshold for apoptosis by inducing the acetylation hsp90 and thus affect oncoproteins such as Bcr-Abl and FLT-3 (Yao et al., 2003; Atadja et al., 2004b). This eventually results in the inhibition of its chaperone association with important pro-survival client proteins such as Erk, Akt, and c-Raf (Yu et al., 2003c). SAHA and oxamflatin were also shown to kill both ABCB1-positive and -negative cells, whereas FK228 was shown to be substrate for ABCB1 (Peart et al., 2003). These data may provide insight into defining rational approaches to chemotherapy in which the genetic profile of the tumor is matched with a functional profile to promote favorable clinical response.

    Induction of the cell cycle inhibitor plays an important role in the induction of differentiation by HDAC inhibitors. SAHA and sodium butyrate were shown to induce differentiation of leukemia and breast cancer cells (Gore et al., 2001, 2002). Induction of the expression of other molecules involved in differentiation (such as gelsolin, an actin-binding protein involved in cell morphology) and structural changes were observed during treatment with HDAC inhibitors (Hoshikawa et al., 1994; Mielnicki et al., 1999; Han et al., 2000; Kamitani et al., 2002). In addition to pro-apoptotic and cytostatic activities, another mode of tumor regression after treatment with HDAC inhibitors may be by indirect inhibition of angiogenesis. In in vitro models, depsipeptide potently blocked the hypoxia-stimulated proliferation, invasion, migration, adhesion, and tube formation of bovine aortic endothelial cells (Kwon et al., 2002). Effective concentrations were comparable with cytotoxic concentrations, and there was an indication of possible modulation of gene transcription as evidenced by the expression of angiogenic-inhibiting factors such as von Hippel Lindau and neurofibromin 2 and the suppression of angiogenic-stimulating factors such as vascular endothelial growth factor (Mie Lee et al., 2003; Sasakawa et al., 2003b). Other HDAC inhibitors, such as apicidin, TSA, butyrate, and newer analog LAQ824 were all shown to inhibit angiogenesis through vascular endothelial growth factor inhibition (Williams, 2001; Sawa et al., 2002; Kim et al., 2004; Michaelis et al., 2004; Qian et al., 2004). Such insights into the mechanisms by which HDAC inhibitors interfere with cancer cell growth and survival have prompted the search for combination strategies to optimize therapy.

    Combination Therapy of HDAC Inhibitors with Other Drugs. Silencing of genes that affect growth and differentiation has been shown to occur by aberrant DNA methylation in the promoter region and by changes in chromatin structure that involve histone deacetylation (Baylin et al., 2001; Herman and Baylin, 2003). DNA methylation and histone deacetylation seem to act as synergistic layers for the transcriptional silencing of genes in cancer (Cameron et al., 1999; Zhu et al., 2001; Zhu and Otterson, 2003). Such findings have great implication in development of combination therapies.

    Epigenetic mechanisms, such as DNA methylation and histone deacetylation, may also play a role in loss of ER expression in ER-negative human breast cancer cells. Previous studies showed that pharmacologic inhibition of these mechanisms using the DNA methyltransferase inhibitor 5-aza-2'deoxycitidine (AZA) and TSA resulted in expression of functional ER mRNA and protein (Yang et al., 2001). Scriptaid, a novel TPX-HA analog, inhibits tumor growth in vitro and in vivo and, in conjunction with AZA, acts to reexpress functional ER (Keen et al., 2003). In another study, TSA was shown to sensitize ER-negative antihormone-unresponsive breast cancer cells to tamoxifen treatment by up-regulating tamoxifen's activity (Jang et al., 2004). The in vitro antineoplastic activity of AZA, in combination with TSA or depsipeptide, on the human myeloid leukemic cell lines produced a greater inhibition of growth and DNA synthesis and a greater loss of clonogenicity than that caused by either agent alone (Shaker et al., 2003). Similar results were noted with PB and AZA combination in lymphoid leukemic cells (Lemaire et al., 2004). Another study found that when AZA was combined with PB, murine lung tumor development was significantly reduced >50%, whereas no effect was observed with PB alone (Belinsky et al., 2003).

    Chromatin DNA is tightly packed; hence, accessibility to the drug target may reduce the efficiency of these anticancer drugs. When six cancer cell lines were pretreated with TSA or SAHA followed by exposure to anticancer drugs that target chromatin DNA, such as etoposide (VP-16), camptothecin, cisplatin, doxorubicin, 5-fluorouracil, cyclophosphamide, or ellipcitine, there was >10-fold sensitization of cells for VP-16. The data suggested that loosening up the chromatin structure by histone acetylation can increase efficiency of several anticancer agents (Kim et al., 2003). SAHA significantly potentiated the DNA damage by topoisomerase II inhibitors; however, synergy was dependent on the sequence of drug administration and expression of target. Pre-exposure of cells to SAHA for 48 h was synergistic, whereas shorter periods of exposure abrogated synergy, and pretreatment with topoisomerase II inhibitor showed antagonistic effects (Marchion et al., 2004).

    Inhibition of cell survival signals and proliferation by using inhibitors of tyrosine kinase activity, in combination with HDAC inhibitors is another mechanism to induce differentiation and/or apoptosis (Nimmanapalli and Bhalla, 2002). The cytotoxic effects that occurred after the introduction of SAHA with imatinib mesylate showed accumulation of acetylated histones H3 and H4 and induction of p21 and p27; after SAHA treatment, there was a decline in the mRNA and protein levels of Bcr-Abl, resulting in G1 arrest and apoptosis of leukemic cells. Cotreatment with imatinib mesylate and SAHA caused significantly more down-regulation of tyrosine kinase activity of Bcr-Abl and apoptosis of these cells compared with treatment with SAHA alone. These findings suggested that cotreatment with SAHA and imatinib mesylate or arsenic trioxide are cytotoxic to Bcr-AbleCpositive acute leukemia cells, and these agents may be promising therapy against imatinib mesylate-refractory Bcr-Abl positive acute leukemia (Nimmanapalli et al., 2003a; Yu et al., 2003a). Similar results were achieved on combined exposure of Bcr/Abl-positive human myeloid leukemia cells to imatinib (STI571) and SAHA, leading to diverse perturbations in signaling and cell cycle-regulatory proteins associated with a marked increase in mitochondrial damage and cell death (Yu et al., 2003a). SAHA and PB were also shown to synergistically induce apoptosis in human leukemic cells when cotreated with the hsp90 antagonist 17-allylamino-17-demethoxygeldenamycin (Rahmani et al., 2003). Similar cumulative inhibitory effects were noted on combined treatment of SB and flavopiridol, in which interruption of HDAC-mediated p21WAF1/Cip1 induction by flavopiridol potentiated apoptosis (Rosato et al., 2002, 2004). The same investigators showed recently that MS-275 acts synergistically with fludarabine to increase the apoptotic activity in leukemia cells (Maggio et al., 2004). Moreover, the proteasome inhibitor bortezomib interacts synergistically with SB or SAHA to cause oxidative injury and apoptosis in Bcr/Abl-positive multiple myeloma and leukemia cells that are sensitive and resistant to imatinib (Yu et al., 2003b; Pei et al., 2004).

    LAQ824 both lowers expression and promotes proteasomal degradation of Bcr-Abl and induces apoptosis of imatinib-sensitive or refractory chronic myelogenous leukemia-blast crisis cells (Nimmanapalli et al., 2003b). Recent studies show that LAQ824 can also promote degradation of mutant FLT-3 and induce apoptosis of AML cells carrying the mutated FLT-3. The addition of the Flt-3 kinase inhibitor PKC412 had a synergistic effect on apoptosis in AML cells with mutant FLT-3 (Bali et al., 2004). The combination of SAHA or LAQ824 with various cytotoxic agents such as taxotere, trastuzumab, gemcitabine, and epothilone B enhanced the cytotoxic effects in breast cancer cells, whereas the combination of 5-fluorouracil and other chemotherapy agents with PB also enhanced the cytotoxic effects in colorectal cancer cells (Huang and Waxman, 1998; Huang et al., 2000; Fuino et al., 2003). In two separate studies, SAHA also potentiated sensitizing melanoma cells to tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)-induced apoptosis by simultaneous activation of intrinsic and extrinsic pathways (Rosato et al., 2003; Zhang et al., 2003). In another study, VA was shown to increase cellular sensitivity to estrogens, progestins, and other hormone nuclear ligands by functioning as an activator of p42/p44 mitogen-activated protein kinase (Jansen et al., 2004). TSA up-regulated RECK glycoprotein, which negatively regulates matrix metalloproteinases (MMPs) and inhibits tumor metastasis and angiogenesis by specifically inhibiting MMP-2 (Liu et al., 2003). Radiotherapy is an effective treatment for several cancers but causes cutaneous radiation syndrome. PB, TSA, and VA were shown to decrease skin fibrosis and tumorigenesis by suppressing aberrant expression of transforming growth factor- and tumor necrosis factor- (Chung et al., 2004). In human gastric and colorectal cancer cells, depsipeptide, MS-275, and m-carboxycinnamic acid bishydroxamide augmented radiation-induced cell death (Zhang et al., 2004c). Moreover, HDAC inhibitors have shown synergism when combined with all-trans-retinoic acid to overcome the block in differentiation due to specific translocations associated with acute promyelocytic leukemia (Coffey et al., 2000, 2001; He et al., 2001).

    Histone Deacetylase Inhibitors in Clinical Trials. Based on promising preclinical data, several HDAC inhibitors are currently being investigated in early phase trials in humans, both as single agents and in combination with known cytotoxic compounds. HDAC inhibitors such as PA, PB, VA, AN-9, SAHA, LAQ824, pyroxamide, FK228, MS-275, and CI-994 are evaluated in patients with various metastatic or refractory solid tumors in advanced stages and those with hematological malignancies such as AML, acute lymphocytic leukemia, chronic myeloid leukemia, chronic lymphoid leukemia, or lymphomas. Details about phase of development, major toxicities, pharmacokinetics, and preliminary data on clinical response of various HDAC inhibitors used as single agent or given in combination with cytotoxic agents that are undergoing clinical development are summarized in Tables 2 and 3, respectively.

    Future Directions

    The concept of mechanism-based therapeutic development of novel anticancer agents is now being recognized, because better targeting of abnormalities has been shown to offer new directions. The HDAC inhibitors in clinical trials have shown encouraging antitumor effects and well tolerated safety profiles. There may be significant repercussions in success or failure of an anticancer agent when targeting a specific subtype of HDAC without having a broader understanding of mechanism of action and the differential role each enzyme plays in chromatin remodeling in cancer cells. Although none of these agents in clinical trials were developed to be selective inhibitors of individual HDAC subtype, they do show some target selectivity (McLaughlin and La Thangue, 2004). For example, MS-275 showed selective inhibition of HDAC1 and HDAC3 but was inactive against HDAC8 (Hu et al., 2003). Likewise, FK228 has activity against class I (HDAC1 and HDAC2) enzymes but not against class II (HDAC4 and HDAC6) (Furumai et al., 2002). It remains a challenge to develop specific inhibitors of class I HDACs that are primarily located within the nucleus and class II HDACs that are known to shuttle between nucleus and cytoplasm (Kao et al., 2001; Johnstone, 2002). Recent findings using siRNA techniques to understand HDAC isotypes as potential targets suggested that class I HDAC enzymes may be more relevant targets for intervention in oncology (Curtin and Glaser, 2003). In any case, chromatin-modifying enzymes have provided an increasingly validated therapeutic target, and there is now compelling evidence that these compounds exhibit efficacy in human diseases.

    Phase I and II clinical trials with HDAC inhibitors have been completed, and others are being initiated. Most of these have been able to identify suitable doses for treatment with relatively less toxicity and reasonable efficacy in various cancers. Remission seemed to be transient in some of the patient trials, suggesting a need for determination of dosing parameters (Bhalla and List, 2004). Based on preliminary clinical data and the apparent cytostatic mechanism of action, most HDAC inhibitors, with the possible exception of FK228 in the treatment of renal cell carcinoma, seem to be more suited to combination treatment with existing chemotherapy regimens and to being used in other mechanism-based agents. Nonetheless, various questions still remain to be answered: 1) what role do altered HAT or HDAC activities have in conjunction with tumorigenesis Is it a direct effect or is an epigenetic adaptive phenomenon 2) Why are tumor cells more sensitive to HDAC inhibitors than normal cells, and is there a possibility that there may be increased HAT/HDAC activity in tumors 3) Is modification of histone(s) the only mechanism leading to antineoplastic effects, or are there targets responsible that are as-yet undefined and 4) What is the target specificity of HDAC inhibitors (Piekarz and Bates, 2004). Unraveling specific roles of HDAC isozymes during human tumorigenesis will provide further incentive for the development of more specific HDAC inhibitors, potentially those enhancing clinical activity as well as decreasing nonspecific toxicities. In addition, optimizing potential interactions with other rationally designed and integrated therapeutic agents remains a promising premise for exploration. In addition, there is a general current lack of knowledge on the pharmacokinetics and biodistribution of various HDAC inhibitors studied clinically. Current evidence suggests that novel formulations and drug delivery strategies that allow better targeting may significantly enhance the therapeutic potential of HDAC inhibitors (Drummond et al., 2004).

    Conclusion

    A wealth of recent data has become available suggesting that histone modification is a promising therapeutic strategy affecting many of the hallmark traits of cancer (Hanahan and Weinberg, 2000). Drugs such as HDAC inhibitors that have pleiotropic actions in modulating multiple genes, pathways, and biological features of malignancy might prove to be suited for dealing with combinatorial oncogenic abnormalities seen with most cancer types (Kristeleit et al., 2004). In addition to applications in oncology, manipulation of histones involved in other diseases, such as Huntington's disease and hepatic fibrosis, may be avenues for further explorations in other therapeutic areas (Penner et al., 1987; Ferrante et al., 2003; Hockly et al., 2003). Although the clinical development of novel HDAC inhibitors seems certain, their actual value will greatly depend on identification of molecular and cellular predictors and elucidation of their mechanism of action as anticancer agents.

    Acknowledgements

    We thank Richard Piekarz for critical review of the manuscript and for his helpful suggestions.

    doi:10.1124/mol.105.014167.

    References

    Annunziato AT and Hansen JC (2000) Role of histone acetylation in the assembly and modulation of chromatin structures. Gene Expr 9: 37eC61.

    Archer SY, Meng S, Shei A, and Hodin RA (1998) p21(WAF1) is required for butyrate-mediated growth inhibition of human colon cancer cells. Proc Natl Acad Sci USA 95: 6791eC6796.

    Aron JL, Parthun MR, Marcucci G, Kitada S, Mone AP, Davis ME, Shen T, Murphy T, Wickham J, Kanakry C, et al. (2003) Depsipeptide (FR901228) induces histone acetylation and inhibition of histone deacetylase in chronic lymphocytic leukemia cells concurrent with activation of caspase 8-mediated apoptosis and down-regulation of c-FLIP protein. Blood 102: 652eC658.

    Arts J, de Schepper S, and Van Emelen K (2003) Histone deacetylase inhibitors: from chromatin remodeling to experimental cancer therapeutics. Curr Med Chem 10: 2343eC2350.

    Atadja P, Gao L, Kwon P, Trogani N, Walker H, Hsu M, Yeleswarapu L, Chandramouli N, Perez L, Versace R, et al. (2004a) Selective growth inhibition of tumor cells by a novel histone deacetylase inhibitor, NVP-LAQ824. Cancer Res 64: 689eC695.

    Atadja P, Hsu M, Kwon P, Trogani N, Bhalla K, and Remiszewski S (2004b) Molecular and cellular basis for the anti-proliferative effects of the HDAC inhibitor LAQ824. Novartis Found Symp 259: 249eC266; discussion 266eC8, 285eC8.

    Atmaca AM, Heinzel T, Gttlicher M, Neumann A, Al-Batran S-E, Martin E, Bartsch I, Knuth A, and Jaeger E (2004) A dose-escalating phase I study with valproic acid (VPA) in patients (pts) with advanced cancer. J Clin Oncol 22 (14 Suppl):Abstract 3169.

    Bali P, George P, Cohen P, Tao J, Guo F, Sigua C, Vishvanath A, Scuto A, Annavarapu S, Fiskus W, et al. (2004) Superior activity of the combination of histone deacetylase inhibitor LAQ824 and the FLT-3 kinase inhibitor PKC412 against human acute myelogenous leukemia cells with mutant FLT-3. Clin Cancer Res 10: 4991eC4997.

    Batova A, Shao LE, Diccianni MB, Yu AL, Tanaka T, Rephaeli A, Nudelman A, and Yu J (2002) The histone deacetylase inhibitor AN-9 has selective toxicity to acute leukemia and drug-resistant primary leukemia and cancer cell lines. Blood 100: 3319eC3324.

    Baylin SB, Esteller M, Rountree MR, Bachman KE, Schuebel K, and Herman JG (2001) Aberrant patterns of DNA methylation, chromatin formation and gene expression in cancer. Hum Mol Genet 10: 687eC692.

    Beck J, Fischer T, Rowinsky E, Huber C, Mita M, Atadja P, Peng B, Kwong C, Dugan M, and Patnaik A (2004) Phase I pharmacokinetic (PK) and pharmacodynamic (PD) study of LBH589A: a novel histone deacetylase inhibitor. J Clin Oncol 22 (14 Suppl):Abstract 3025.

    Belinsky SA, Klinge DM, Stidley CA, Issa JP, Herman JG, March TH, and Baylin SB (2003) Inhibition of DNA methylation and histone deacetylation prevents murine lung cancer. Cancer Res 63: 7089eC7093.

    Bernhard D, Ausserlechner MJ, Tonko M, Loffler M, Hartmann BL, Csordas A, and Kofler R (1999) Apoptosis induced by the histone deacetylase inhibitor sodium butyrate in human leukemic lymphoblasts. FASEB J 13: 1991eC2001.

    Bhalla K and List A (2004) Histone deacetylase inhibitors in myelodysplastic syndrome. Best Pract Res Clin Haematol 17: 595eC611.

    Blagosklonny MV, Robey R, Sackett DL, Du L, Traganos F, Darzynkiewicz Z, Fojo T, and Bates SE (2002) Histone deacetylase inhibitors all induce p21 but differentially cause tubulin acetylation, mitotic arrest and cytotoxicity. Mol Cancer Ther 1: 937eC941.

    Blumenschein G, Lu C, Kies M, Glisson B, Papadimitrakopoulou V, Zinner R, Kim E, Gillenwater A, Chiao J, and W Hong (2004) Phase II clinical trial of suberoylanilide hydroxamic acid (SAHA) in patients (pts) with recurrent and/or metastatic head and neck cancer(SCCHN). J Clin Oncol 22 (14 Suppl):Abstract 5578.

    Bouchain G and Delorme D (2003) Novel hydroxamate and anilide derivatives as potent histone deacetylase inhibitors: synthesis and antiproliferative evaluation. Curr Med Chem 10: 2359eC2372.

    Bouchain G, Leit S, Frechette S, Khalil EA, Lavoie R, Moradei O, Woo SH, Fournel M, Yan PT, Kalita A, et al. (2003) Development of potential antitumor agents. Synthesis and biological evaluation of a new set of sulfonamide derivatives as histone deacetylase inhibitors. J Med Chem 46: 820eC830.

    Butler LM, Agus DB, Scher HI, Higgins B, Rose A, Cordon-Cardo C, Thaler HT, Rifkind RA, Marks PA, and Richon VM (2000) Suberoylanilide hydroxamic acid, an inhibitor of histone deacetylase, suppresses the growth of prostate cancer cells in vitro and in vivo. Cancer Res 60: 5165eC5170.

    Butler LM, Webb Y, Agus DB, Higgins B, Tolentino TR, Kutko MC, LaQuaglia MP, Drobnjak M, Cordon-Cardo C, Scher HI, et al. (2001) Inhibition of transformed cell growth and induction of cellular differentiation by pyroxamide, an inhibitor of histone deacetylase. Clin Cancer Res 7: 962eC970.

    Butler LM, Zhou X, Xu WS, Scher HI, Rifkind RA, Marks PA, and Richon VM (2002) The histone deacetylase inhibitor SAHA arrests cancer cell growth, up-regulates thioredoxin-binding protein-2 and down-regulates thioredoxin. Proc Natl Acad Sci USA 99: 11700eC11705.

    Byrd JC, Marcucci G, Parthun MR, Xiao JJ, Klisovic RB, Moran M, Lin TS, Liu S, Sklenar AR, Davis ME, et al. (2004) A phase I and pharmacodynamic study of depsipeptide (FK228) in chronic lymphocytic leukemia and acute myeloid leukemia. Blood 105: 959eC967.

    Byrd JC, Shinn C, Ravi R, Willis CR, Waselenko JK, Flinn IW, Dawson NA, and Grever MR (1999) Depsipeptide (FR901228): a novel therapeutic agent with selective, in vitro activity against human B-cell chronic lymphocytic leukemia cells. Blood 94: 1401eC1408.

    Cameron EE, Bachman KE, Myohanen S, Herman JG, and Baylin SB (1999) Synergy of demethylation and histone deacetylase inhibition in the re-expression of genes silenced in cancer. Nat Genet 21: 103eC107.

    Carducci MA, Gilbert J, Bowling MK, Noe D, Eisenberger MA, Sinibaldi V, Zabelina Y, Chen TL, Grochow LB, and Donehower RC (2001) A phase I clinical and pharmacological evaluation of sodium phenylbutyrate on an 120-h infusion schedule. Clin Cancer Res 7: 3047eC3055.

    Carducci MA, Nelson JB, Chan-Tack KM, Ayyagari SR, Sweatt WH, Campbell PA, Nelson WG, and Simons JW (1996) Phenylbutyrate induces apoptosis in human prostate cancer and is more potent than phenylacetate. Clin Cancer Res 2: 379eC387.

    Catley L, Weisberg E, Tai YT, Atadja P, Remiszewski S, Hideshima T, Mitsiades N, Shringarpure R, LeBlanc R, Chauhan D, et al. (2003) NVP-LAQ824 is a potent novel histone deacetylase inhibitor with significant activity against multiple myeloma. Blood 102: 2615eC2622.

    Chai F, Evdokiou A, Young GP, and Zalewski PD (2000) Involvement of p21(Waf1/ Cip1) and its cleavage by DEVD-caspase during apoptosis of colorectal cancer cells induced by butyrate. Carcinogenesis 21: 7eC14.

    Chang SM, Kuhn JG, Ian Robins H, Clifford Schold S, Spence AM, Berger MS, Mehta MP, Pollack I, Gilbert M, and Prados MD (2003) A study of a different dose-intense infusion schedule of phenylacetate in patients with recurrent primary brain tumors consortium report. Investig New Drugs 21: 429eC433.

    Chang SM, Kuhn JG, Robins HI, Schold SC, Spence AM, Berger MS, Mehta MP, Bozik ME, Pollack I, Schiff D, et al. (1999) Phase II study of phenylacetate in patients with recurrent malignant glioma: a North American Brain Tumor Consortium report. J Clin Oncol 17: 984eC990.

    Chen H, Tini M, and Evans RM (2001) HATs on and beyond chromatin. Curr Opin Cell Biol 13: 218eC224.

    Cheng HL, Mostoslavsky R, Saito S, Manis JP, Gu Y, Patel P, Bronson R, Appella E, Alt FW, and Chua KF (2003) Developmental defects and p53 hyperacetylation in Sir2 homolog (SIRT1)-deficient mice. Proc Natl Acad Sci USA 100: 10794eC10799.

    Chiba T, Yokosuka O, Arai M, Tada M, Fukai K, Imazeki F, Kato M, Seki N, and Saisho H (2004a) Identification of genes up-regulated by histone deacetylase inhibition with cDNA microarray and exploration of epigenetic alterations on hepatoma cells. J Hepatol 41: 436eC445.

    Chiba T, Yokosuka O, Fukai K, Kojima H, Tada M, Arai M, Imazeki F, and Saisho H (2004b) Cell growth inhibition and gene expression induced by the histone deacetylase inhibitor, trichostatin A, on human hepatoma cells. Oncology 66: 481eC491.

    Chung YL, Wang AJ, and Yao LF (2004) Antitumor histone deacetylase inhibitors suppress cutaneous radiation syndrome: implications for increasing therapeutic gain in cancer radiotherapy. Mol Cancer Ther 3: 317eC325.

    Coffey DC, Kutko MC, Glick RD, Butler LM, Heller G, Rifkind RA, Marks PA, Richon VM, and La Quaglia MP (2001) The histone deacetylase inhibitor, CBHA, inhibits growth of human neuroblastoma xenografts in vivo, alone and synergistically with all-trans retinoic acid. Cancer Res 61: 3591eC3594.

    Coffey DC, Kutko MC, Glick RD, Swendeman SL, Butler L, Rifkind R, Marks PA, Richon VM, and LaQuaglia MP (2000) Histone deacetylase inhibitors and retinoic acids inhibit growth of human neuroblastoma in vitro. Med Pediatr Oncol 35: 577eC581.

    Cohen LA, Marks PA, Rifkind RA, Amin S, Desai D, Pittman B, and Richon VM (2002) Suberoylanilide hydroxamic acid (SAHA), a histone deacetylase inhibitor, suppresses the growth of carcinogen-induced mammary tumors. Anticancer Res 22: 1497eC1504.

    Curtin M and Glaser K (2003) Histone deacetylase inhibitors: the Abbott experience. Curr Med Chem 10: 2373eC2392.

    Curtin ML, Garland RB, Heyman HR, Frey RR, Michaelides MR, Li J, Pease LJ, Glaser KB, Marcotte PA, and Davidsen SK (2002) Succinimide hydroxamic acids as potent inhibitors of histone deacetylase (HDAC). Bioorg Med Chem Lett 12: 2919eC2923.

    Davis PK and Brackmann RK (2003) Chromatin remodeling and cancer. Cancer Biol Ther 2: 22eC29.

    de Ruijter AJ, van Gennip AH, Caron HN, Kemp S, and van Kuilenburg AB (2003) Histone deacetylases (HDACs): characterization of the classical HDAC family. Biochem J 370: 737eC749.

    Donadelli M, Costanzo C, Faggioli L, Scupoli MT, Moore PS, Bassi C, Scarpa A, and Palmieri M (2003) Trichostatin A, an inhibitor of histone deacetylases, strongly suppresses growth of pancreatic adenocarcinoma cells. Mol Carcinog 38: 59eC69.

    Drummond DC, Noble CO, Kirpotin DB, Guo Z, Scott GK, and Benz CC (2004) Clinical development of histone deacetylase inhibitors as anticancer agents. Annu Rev Pharmacol Toxicol 45: 495eC528.

    Eyal S, Yagen B, Sobol E, Altschuler Y, Shmuel M, and Bialer M (2004) The activity of antiepileptic drugs as histone deacetylase inhibitors. Epilepsia 45: 737eC744.

    Fenrick R and Hiebert SW (1998) Role of histone deacetylases in acute leukemia. J Cell Biochem Suppl 30eC31: 194eC202.

    Ferrante RJ, Kubilus JK, Lee J, Ryu H, Beesen A, Zucker B, Smith K, Kowall NW, Ratan RR, Luthi-Carter R, et al. (2003) Histone deacetylase inhibition by sodium butyrate chemotherapy ameliorates the neurodegenerative phenotype in Huntington's disease mice. J Neurosci 23: 9418eC9427.

    Finnin MS, Donigian JR, Cohen A, Richon VM, Rifkind RA, Marks PA, Breslow R, and Pavletich NP (1999) Structures of a histone deacetylase homologue bound to the TSA and SAHA inhibitors. Nature (Lond) 401: 188eC193.

    Finzer P, Stohr M, Seibert N, and Rosl F (2003) Phenylbutyrate inhibits growth of cervical carcinoma cells independent of HPV type and copy number. J Cancer Res Clin Oncol 129: 107eC113.

    Fournel M, Trachy-Bourget MC, Yan PT, Kalita A, Bonfils C, Beaulieu C, Frechette S, Leit S, Abou-Khalil E, Woo SH, et al. (2002) Sulfonamide anilides, a novel class of histone deacetylase inhibitors, are antiproliferative against human tumors. Cancer Res 62: 4325eC4330.

    Frye RA (2000) Phylogenetic classification of prokaryotic and eukaryotic Sir2-like proteins. Biochem Biophys Res Commun 273: 793eC798.

    Fuino L, Bali P, Wittmann S, Donapaty S, Guo F, Yamaguchi H, Wang HG, Atadja P, and Bhalla K (2003) Histone deacetylase inhibitor LAQ824 down-regulates Her-2 and sensitizes human breast cancer cells to trastuzumab, taxotere, gemcitabine and epothilone B. Mol Cancer Ther 2: 971eC984.

    Furumai R, Komatsu Y, Nishino N, Khochbin S, Yoshida M, and Horinouchi S (2001) Potent histone deacetylase inhibitors built from trichostatin A and cyclic tetrapeptide antibiotics including trapoxin. Proc Natl Acad Sci USA 98: 87eC92.

    Furumai R, Matsuyama A, Kobashi N, Lee KH, Nishiyama M, Nakajima H, Tanaka A, Komatsu Y, Nishino N, Yoshida M, et al. (2002) FK228 (depsipeptide) as a natural prodrug that inhibits class I histone deacetylases. Cancer Res 62: 4916eC4921.

    Garcia-Manero G, Issa J-P, Cortes J, Koller C, O'Brien S, Estey E, Canalli AA, Chiao J, Richon V, and Kantarjian H (2004) Phase I study of oral suberoylanilide hydroxamic acid (SAHA), a histone deacetylase inhibitor, in patients (pts) with advanced leukemias or myelodysplastic syndromes (MDS). J Clin Oncol 22 (14 Suppl):Abstract 3027.

    Gilbert J, Baker SD, Bowling MK, Grochow L, Figg WD, Zabelina Y, Donehower RC, and Carducci MA (2001) A phase I dose escalation and bioavailability study of oral sodium phenylbutyrate in patients with refractory solid tumor malignancies. Clin Cancer Res 7: 2292eC2300.

    Gore L, Holden SN, Basche M, Raj SKS, Arnold I, O'Bryant C, Witta S, Rohde B, McCoy C, and Eckhardt SG (2004) Updated results from a phase I trial of the histone deacetylase (HDAC) inhibitor MS-275 in patients with refractory solid tumors. J Clin Oncol 22 (14 Suppl):Abstract 3026.

    Gore SD and Carducci MA (2000) Modifying histones to tame cancer: clinical development of sodium phenylbutyrate and other histone deacetylase inhibitors. Expert Opin Investig Drugs 9: 2923eC2934.

    Gore SD, Weng LJ, Figg WD, Zhai S, Donehower RC, Dover G, Grever MR, Griffin C, Grochow LB, Hawkins A, et al. (2002) Impact of prolonged infusions of the putative differentiating agent sodium phenylbutyrate on myelodysplastic syndromes and acute myeloid leukemia. Clin Cancer Res 8: 963eC970.

    Gore SD, Weng LJ, Zhai S, Figg WD, Donehower RC, Dover GJ, Grever M, Griffin CA, Grochow LB, Rowinsky EK, et al. (2001) Impact of the putative differentiating agent sodium phenylbutyrate on myelodysplastic syndromes and acute myeloid leukemia. Clin Cancer Res 7: 2330eC2339.

    Gozzini A, Rovida E, Dello Sbarba P, Galimberti S, Santini V, and Galimbert S (2003) Butyrates, as a single drug, induce histone acetylation and granulocytic maturation: possible selectivity on core binding factor-acute myeloid leukemia blasts. Cancer Res 63: 8955eC8961.

    Gray SG and Ekstrom TJ (2001) The human histone deacetylase family. Exp Cell Res 262: 75eC83.

    Graziano MJ, Pilcher GD, Walsh KM, Kasali OB, and Radulovic L (1997) Preclinical toxicity of a new oral anticancer drug, CI-994 (acetyldinaline), in rats and dogs. Investig New Drugs 15: 295eC310.

    Gregoretti IV, Lee YM, and Goodson HV (2004) Molecular evolution of the histone deacetylase family: functional implications of phylogenetic analysis. J Mol Biol 338: 17eC31.

    Gregory PD, Wagner K, and Horz W (2001) Histone acetylation and chromatin remodeling. Exp Cell Res 265: 195eC202.

    Grozinger CM and Schreiber SL (2002) Deacetylase enzymes: biological functions and the use of small-molecule inhibitors. Chem Biol 9: 3eC16.

    Guardiola AR and Yao TP (2002) Molecular cloning and characterization of a novel histone deacetylase HDAC10. J Biol Chem 277: 3350eC3356.

    Gui CY, Ngo L, Xu WS, Richon VM, and Marks PA (2004) Histone deacetylase (HDAC) inhibitor activation of p21WAF1 involves changes in promoter-associated proteins, including HDAC1. Proc Natl Acad Sci USA 101: 1241eC1246.

    Guo F, Sigua C, Tao J, Bali P, George P, Li Y, Wittmann S, Moscinski L, Atadja P, and Bhalla K (2004) Cotreatment with histone deacetylase inhibitor LAQ824 enhances Apo-2L/tumor necrosis factor-related apoptosis inducing ligand-induced death inducing signaling complex activity and apoptosis of human acute leukemia cells. Cancer Res 64: 2580eC2589.

    Gurvich N, Tsygankova OM, Meinkoth JL, and Klein PS (2004) Histone deacetylase is a target of valproic acid-mediated cellular differentiation. Cancer Res 64: 1079eC1086.

    Han JW, Ahn SH, Park SH, Wang SY, Bae GU, Seo DW, Kwon HK, Hong S, Lee HY, Lee YW, et al. (2000) Apicidin, a histone deacetylase inhibitor, inhibits proliferation of tumor cells via induction of p21WAF1/Cip1 and gelsolin. Cancer Res 60: 6068eC6074.

    Hanahan D and Weinberg RA (2000) The hallmarks of cancer. Cell 100: 57eC70.

    He LZ, Tolentino T, Grayson P, Zhong S, Warrell RP Jr, Rifkind RA, Marks PA, Richon VM, and Pandolfi PP (2001) Histone deacetylase inhibitors induce remission in transgenic models of therapy-resistant acute promyelocytic leukemia. J Clin Investig 108: 1321eC1330.

    Heltweg B, Dequiedt F, Marshall BL, Brauch C, Yoshida M, Nishino N, Verdin E, and Jung M (2004) Subtype selective substrates for histone deacetylases. J Med Chem 47: 5235eC5243.

    Herman JG and Baylin SB (2003) Gene silencing in cancer in association with promoter hypermethylation. N Engl J Med 349: 2042eC2054.

    Herold C, Ganslmayer M, Ocker M, Hermann M, Geerts A, Hahn EG, and Schuppan D (2002) The histone-deacetylase inhibitor trichostatin A blocks proliferation and triggers apoptotic programs in hepatoma cells. J Hepatol 36: 233eC240.

    Hockly E, Richon VM, Woodman B, Smith DL, Zhou X, Rosa E, Sathasivam K, Ghazi-Noori S, Mahal A, Lowden PA, et al. (2003) Suberoylanilide hydroxamic acid, a histone deacetylase inhibitor, ameliorates motor deficits in a mouse model of Huntington's disease. Proc Natl Acad Sci USA 100: 2041eC2046.

    Hong J, Ishihara K, Yamaki K, Hiraizumi K, Ohno T, Ahn JW, Zee O, and Ohuchi K (2003) Apicidin, a histone deacetylase inhibitor, induces differentiation of HL-60 cells. Cancer Lett 189: 197eC206.

    Hoshikawa Y, Kwon HJ, Yoshida M, Horinouchi S, and Beppu T (1994) Trichostatin A induces morphological changes and gelsolin expression by inhibiting histone deacetylase in human carcinoma cell lines. Exp Cell Res 214: 189eC197.

    Hu E, Dul E, Sung CM, Chen Z, Kirkpatrick R, Zhang GF, Johanson K, Liu R, Lago A, Hofmann G, et al. (2003) Identification of novel isoform-selective inhibitors within class I histone deacetylases. J Pharmacol Exp Ther 307: 720eC728.

    Huang L and Pardee AB (2000) Suberoylanilide hydroxamic acid as a potential therapeutic agent for human breast cancer treatment. Mol Med 6: 849eC866.

    Huang L, Sowa Y, Sakai T, and Pardee AB (2000) Activation of the p21WAF1/CIP1 promoter independent of p53 by the histone deacetylase inhibitor suberoylanilide hydroxamic acid (SAHA) through the Sp1 sites. Oncogene 19: 5712eC5719.

    Huang Y, Horvath CM, and Waxman S (2000) Regrowth of 5-fluorouracil-treated human colon cancer cells is prevented by the combination of interferon gamma, indomethacin and phenylbutyrate. Cancer Res 60: 3200eC3206.

    Huang Y and Waxman S (1998) Enhanced growth inhibition and differentiation of fluorodeoxyuridine-treated human colon carcinoma cells by phenylbutyrate. Clin Cancer Res 4: 2503eC2509.

    Imai S, Armstrong CM, Kaeberlein M, and Guarente L (2000) Transcriptional silencing and longevity protein Sir2 is an NAD-dependent histone deacetylase. Nature (Lond) 403: 795eC800.

    Jaboin J, Wild J, Hamidi H, Khanna C, Kim CJ, Robey R, Bates SE, and Thiele CJ (2002) MS-27eC275, an inhibitor of histone deacetylase, has marked in vitro and in vivo antitumor activity against pediatric solid tumors. Cancer Res 62: 6108eC6115.

    Jang ER, Lim SJ, Lee ES, Jeong G, Kim TY, Bang YJ, and Lee JS (2004) The histone deacetylase inhibitor trichostatin A sensitizes estrogen receptor alpha-negative breast cancer cells to tamoxifen. Oncogene 23: 1724eC1736.

    Jansen MS, Nagel SC, Miranda PJ, Lobenhofer EK, Afshari CA, and McDonnell DP (2004) Short-chain fatty acids enhance nuclear receptor activity through mitogen-activated protein kinase activation and histone deacetylase inhibition. Proc Natl Acad Sci USA 101: 7199eC7204.

    Johnstone RW (2002) Histone-deacetylase inhibitors: novel drugs for the treatment of cancer. Nat Rev Drug Discov 1: 287eC299.

    Johnstone RW and Licht JD (2003) Histone deacetylase inhibitors in cancer therapy: is transcription the primary target Cancer Cell 4: 13eC18.

    Jones PA and Baylin SB (2002) The fundamental role of epigenetic events in cancer. Nat Rev Genet 3: 415eC428.

    Jung M (2001) Inhibitors of histone deacetylase as new anticancer agents. Curr Med Chem 8: 1505eC1511.

    Jung M, Brosch G, Kolle D, Scherf H, Gerhauser C, and Loidl P (1999) Amide analogues of trichostatin A as inhibitors of histone deacetylase and inducers of terminal cell differentiation. J Med Chem 42: 4669eC4679.

    Kamitani H, Taniura S, Watanabe K, Sakamoto M, Watanabe T, and Eling T (2002) Histone acetylation may suppress human glioma cell proliferation when p21 WAF/Cip1 and gelsolin are induced. Neuro-oncol 4: 95eC101.

    Kao HY, Verdel A, Tsai CC, Simon C, Juguilon H, and Khochbin S (2001) Mechanism for nucleocytoplasmic shuttling of histone deacetylase 7. J Biol Chem 276: 47496eC47507.

    Keen JC, Yan L, Mack KM, Pettit C, Smith D, Sharma D, and Davidson NE (2003) A novel histone deacetylase inhibitor, scriptaid, enhances expression of functional estrogen receptor alpha (ER) in ER negative human breast cancer cells in combination with 5-aza 2'-deoxycytidine. Breast Cancer Res Treat 81: 177eC186.

    Kelly WK, Richon VM, O'Connor O, Curley T, MacGregor-Curtelli B, Tong W, Klang M, Schwartz L, Richardson S, Rosa E, et al. (2002) A phase I clinical trial of an oral formulation of the histone deacetylase inhibitor of suberoylanilide hydroxamic acid (SAHA) (Abstract 286). 14th EORTC-NCI-AACR Symposium on Molecular Targets and Cancer Therapeutics; 2002 Nov 19eC22; Frankfurt, Germany. Eur J Cancer 38 (Suppl 7): S1eCS188.

    Kelly WK, Richon VM, O'Connor O, Curley T, MacGregor-Curtelli B, Tong W, Klang M, Schwartz L, Richardson S, Rosa E, et al. (2003) Phase I clinical trial of histone deacetylase inhibitor: suberoylanilide hydroxamic acid administered intravenously. Clin Cancer Res 9: 3578eC3588.

    Kijima M, Yoshida M, Sugita K, Horinouchi S, and Beppu T (1993) Trapoxin, an antitumor cyclic tetrapeptide, is an irreversible inhibitor of mammalian histone deacetylase. J Biol Chem 268: 22429eC22435.

    Kim DH, Kim M, and Kwon HJ (2003) Histone deacetylase in carcinogenesis and its inhibitors as anti-cancer agents. J Biochem Mol Biol 36: 110eC119.

    Kim MS, Blake M, Baek JH, Kohlhagen G, Pommier Y, and Carrier F (2003) Inhibition of histone deacetylase increases cytotoxicity to anticancer drugs targeting DNA. Cancer Res 63: 7291eC7300.

    Kim SH, Ahn S, Han JW, Lee HW, Lee HY, Lee YW, Kim MR, Kim KW, Kim WB, and Hong S (2004) Apicidin is a histone deacetylase inhibitor with anti-invasive and anti-angiogenic potentials. Biochem Biophys Res Commun 315: 964eC970.

    Kim YB, Ki SW, Yoshida M, and Horinouchi S (2000) Mechanism of cell cycle arrest caused by histone deacetylase inhibitors in human carcinoma cells. J Antibiot (Tokyo) 53: 1191eC1200.

    Kim YB, Lee KH, Sugita K, Yoshida M, and Horinouchi S (1999) Oxamflatin is a novel antitumor compound that inhibits mammalian histone deacetylase. Oncogene 18: 2461eC2470.

    Komatsu Y, Tomizaki KY, Tsukamoto M, Kato T, Nishino N, Sato S, Yamori T, Tsuruo T, Furumai R, Yoshida M, et al. (2001) Cyclic hydroxamic-acid-containing peptide 31, a potent synthetic histone deacetylase inhibitor with antitumor activity. Cancer Res 61: 4459eC4466.

    Kosugi H, Towatari M, Hatano S, Kitamura K, Kiyoi H, Kinoshita T, Tanimoto M, Murate T, Kawashima K, Saito H, 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: 1316eC1324.

    Kouraklis G and Theocharis S (2002) Histone deacetylase inhibitors and anticancer therapy. Curr Med Chem Anti-Canc Agents 2: 477eC484.

    Kramer OH, Zhu P, Ostendorff HP, Golebiewski M, Tiefenbach J, Peters MA, Brill B, Groner B, Bach I, Heinzel T, et al. (2003) The histone deacetylase inhibitor valproic acid selectively induces proteasomal degradation of HDAC2. EMBO (Eur Mol Biol Organ) J 22: 3411eC3420.

    Kristeleit R, Stimson L, Workman P, and Aherne W (2004) Histone modification enzymes: novel targets for cancer drugs. Expert Opin Emerg Drugs 9: 135eC154.

    Kuefer R, Hofer MD, Altug V, Zorn C, Genze F, Kunzi-Rapp K, Hautmann RE, and Gschwend JE (2004) Sodium butyrate and tributyrin induce in vivo growth inhibition and apoptosis in human prostate cancer. Br J Cancer 90: 535eC541.

    Kutko MC, Glick RD, Butler LM, Coffey DC, Rifkind RA, Marks PA, Richon VM, and LaQuaglia MP (2003) Histone deacetylase inhibitors induce growth suppression and cell death in human rhabdomyosarcoma in vitro. Clin Cancer Res 9: 5749eC5755.

    Kwon HJ, Kim MS, Kim MJ, Nakajima H, and Kim KW (2002) Histone deacetylase inhibitor FK228 inhibits tumor angiogenesis. Int J Cancer 97: 290eC296.

    Lavelle D, Chen YH, Hankewych M, and DeSimone J (2001) Histone deacetylase inhibitors increase p21(WAF1) and induce apoptosis of human myeloma cell lines independent of decreased IL-6 receptor expression. Am J Hematol 68: 170eC178.

    Lemaire M, Momparler LF, Farinha NJ, Bernstein M, and Momparler RL (2004) Enhancement of antineoplastic action of 5-aza-2'-deoxycytidine by phenylbutyrate on L1210 leukemic cells. Leuk Lymphoma 45: 147eC154.

    Li H and Wu X (2004) Histone deacetylase inhibitor, Trichostatin A, activates p21(WAF1/CIP1) expression through downregulation of c-myc and release of the repression of c-myc from the promoter in human cervical cancer cells. Biochem Biophys Res Commun 324: 860eC867.

    Liu LT, Chang HC, Chiang LC, and Hung WC (2003) Histone deacetylase inhibitor up-regulates RECK to inhibit MMP-2 activation and cancer cell invasion. Cancer Res 63: 3069eC3072.

    LoRusso PM, Demchik L, Foster B, Knight J, Bissery MC, Polin LM, Leopold WR, 3rd and Corbett TH (1996) Preclinical antitumor activity of CI-994. Investig New Drugs 14: 349eC356.

    Lucas DM, Davis ME, Parthun MR, Mone AP, Kitada S, Cunningham KD, Flax EL, Wickham J, Reed JC, Byrd JC, et al. (2004) The histone deacetylase inhibitor MS-275 induces caspase-dependent apoptosis in B-cell chronic lymphocytic leukemia cells. Leukemia 18: 1207eC1214.

    Luo J, Nikolaev AY, Imai S, Chen D, Su F, Shiloh A, Guarente L, and Gu W (2001) Negative control of p53 by Sir2alpha promotes cell survival under stress. Cell 107: 137eC148.

    Maggio SC, Rosato RR, Kramer LB, Dai Y, Rahmani M, Paik DS, Czarnik AC, Payne SG, Spiegel S, and Grant S (2004) The histone deacetylase inhibitor MS-275 interacts synergistically with fludarabine to induce apoptosis in human leukemia cells. Cancer Res 64: 2590eC2600.

    Mahlknecht U and Hoelzer D (2000) Histone acetylation modifiers in the pathogenesis of malignant disease. Mol Med 6: 623eC644.

    Marchion DC, Bicaku E, Daud AI, Richon V, Sullivan DM, and Munster PN (2004) Sequence-specific potentiation of topoisomerase II inhibitors by the histone deacetylase inhibitor suberoylanilide hydroxamic acid. J Cell Biochem 92: 223eC237.

    Margueron R, Licznar A, Lazennec G, Vignon F, and Cavailles V (2003) Oestrogen receptor alpha increases p21(WAF1/CIP1) gene expression and the antiproliferative activity of histone deacetylase inhibitors in human breast cancer cells. J Endocrinol 179: 41eC53.

    Marks PA (2004) The mechanism of the anti-tumor activity of the histone deacetylase inhibitor, suberoylanilide hydroxamic acid (SAHA). Cell Cycle 3: 534eC535.

    Marks PA, Miller T, and Richon VM (2003) Histone deacetylases. Curr Opin Pharmacol 3: 344eC351.

    Marks PA, Richon VM, Miller T, and Kelly WK (2004) Histone deacetylase inhibitors. Adv Cancer Res 91: 137eC168.

    Marks PA, Richon VM, and Rifkind RA (2000) Histone deacetylase inhibitors: inducers of differentiation or apoptosis of transformed cells. J Natl Cancer Inst 92: 1210eC1216.

    Marks P, Rifkind RA, Richon VM, Breslow R, Miller T, and Kelly WK (2001) Histone deacetylases and cancer: causes and therapies. Nat Rev Cancer 1: 194eC202.

    Marshall JL, Rizvi N, Kauh J, Dahut W, Figuera M, Kang MH, Figg WD, Wainer I, Chaissang C, Li MZ, et al. (2002) A phase I trial of depsipeptide (FR901228) in patients with advanced cancer. J Exp Ther Oncol 2: 325eC332.

    McLaughlin F and La Thangue NB (2004) Histone deacetylase inhibitors open new doors in cancer therapy. Biochem Pharmacol 68: 1139eC1144.

    Meinke PT, Colletti SL, Doss G, Myers RW, Gurnett AM, Dulski PM, Darkin-Rattray SJ, Allocco JJ, Galuska S, Schmatz DM, et al. (2000) Synthesis of apicidin-derived quinolone derivatives: parasite-selective histone deacetylase inhibitors and antiproliferative agents. J Med Chem 43: 4919eC4922.

    Melchior SW, Brown LG, Figg WD, Quinn JE, Santucci RA, Brunner J, Thuroff JW, Lange PH, and Vessella RL (1999) Effects of phenylbutyrate on proliferation and apoptosis in human prostate cancer cells in vitro and in vivo. Int J Oncol 14: 501eC508.

    Michaelis M, Michaelis UR, Fleming I, Suhan T, Cinatl J, Blaheta RA, Hoffmann K, Kotchetkov R, Busse R, Nau H, et al. (2004) Valproic acid inhibits angiogenesis in vitro and in vivo. Mol Pharmacol 65: 520eC527.

    Mie Lee Y, Kim SH, Kim HS, Jin Son M, Nakajima H, Jeong Kwon H, and Kim KW (2003) Inhibition of hypoxia-induced angiogenesis by FK228, a specific histone deacetylase inhibitor, via suppression of HIF-1alpha activity. Biochem Biophys Res Commun 300: 241eC246.

    Mielnicki LM, Ying AM, Head KL, Asch HL, and Asch BB (1999) Epigenetic regulation of gelsolin expression in human breast cancer cells. Exp Cell Res 249: 161eC176.

    Miller TA, Witter DJ, and Belvedere S (2003) Histone deacetylase inhibitors. J Med Chem 46: 5097eC5116.

    Mitsiades N, Mitsiades CS, Richardson PG, McMullan C, Poulaki V, Fanourakis G, Schlossman R, Chauhan D, Munshi NC, Hideshima T, et al. (2003) Molecular sequelae of histone deacetylase inhibition in human malignant B cells. Blood 101: 4055eC4062.

    Motta MC, Divecha N, Lemieux M, Kamel C, Chen D, Gu W, Bultsma Y, McBurney M, and Guarente L (2004) Mammalian SIRT1 represses forkhead transcription factors. Cell 116: 551eC563.

    Munster PN, Troso-Sandoval T, Rosen N, Rifkind R, Marks PA, and Richon VM (2001) The histone deacetylase inhibitor suberoylanilide hydroxamic acid induces differentiation of human breast cancer cells. Cancer Res 61: 8492eC8497.

    Nakatani Y (2001) Histone acetylaseseCversatile players. Genes Cells 6: 79eC86.

    Nemunaitis JJ, Orr D, Eager R, Cunningham CC, Williams A, Mennel R, Grove W, and Olson S (2003) Phase I study of oral CI-994 in combination with gemcitabine in treatment of patients with advanced cancer. Cancer J 9: 58eC66.

    Nguyen DM, Schrump WD, Tsai WS, Chen A, Stewart JHT, Steiner F, and Schrump DS (2003) Enhancement of depsipeptide-mediated apoptosis of lung or esophageal cancer cells by flavopiridol: activation of the mitochondria-dependent death-signaling pathway. J Thorac Cardiovasc Surg 125: 1132eC1142.

    Nimmanapalli R and Bhalla K (2002) Mechanisms of resistance to imatinib mesylate in Bcr-Abl-positive leukemias. Curr Opin Oncol 14: 616eC620.

    Nimmanapalli R, Fuino L, Bali P, Gasparetto M, Glozak M, Tao J, Moscinski L, Smith C, Wu J, Jove R, et al. (2003a) Histone deacetylase inhibitor LAQ824 both lowers expression and promotes proteasomal degradation of Bcr-Abl and induces apoptosis of imatinib mesylate-sensitive or -refractory chronic myelogenous leukemia-blast crisis cells. Cancer Res 63: 5126eC5135.

    Nimmanapalli R, Fuino L, Stobaugh C, Richon V, and Bhalla K (2003b) Cotreatment with the histone deacetylase inhibitor suberoylanilide hydroxamic acid (SAHA) enhances imatinib-induced apoptosis of Bcr-Abl-positive human acute leukemia cells. Blood 101: 3236eC3239.

    Nishino N, Jose B, Okamura S, Ebisusaki S, Kato T, Sumida Y, and Yoshida M (2003) Cyclic tetrapeptides bearing a sulfhydryl group potently inhibit histone deacetylases. Org Lett 5: 5079eC5082.

    North BJ, Marshall BL, Borra MT, Denu JM, and Verdin E (2003) The human Sir2 ortholog, SIRT2, is an NAD+-dependent tubulin deacetylase. Mol Cell 11: 437eC444.

    Olivares J, Williams A, Olson S, Pauer LR, Grove W, and Nemunaitis J (2001) Phase I pharmacokinetic (PK) study of CI-994 in combination with carboplatin (C) and paclitaxel (T) in patients (pts) with advanced solid tumors (Abstract 346). Proceedings of the 37th Annual Meeting of the American Society for Clinical Oncology; 2001 May 12eC15; San Francisco, California. American Society for Clinical Oncology, Alexandria, VA.

    O'Shaughnessy J, Flaherty L, Fiorica J, and Grove W (1999) Phase II trial of CI-994 in patients (pts) with metastatic renal cell carcinoma (RCC). (Abstract 1346). Proceedings of the 35th Annual Meeting of the American Society for Clinical Oncology; 1999 May 15eC18; Atlanta, Georgia. American Society for Clinical Oncology, Alexandria, VA.

    Ottmann OG, Deangelo DJ, Stone RM, Pfeifer H, Lowenberg B, Atadja P, Peng B, Scott JW, Dugan M, and Sonneveld P (2004) A phase I, pharmacokinetic (PK) and pharmacodynamic (PD) study of a novel histone deacetylase inhibitor LAQ824 in patients with hematologic malignancies. J Clin Oncol 22 (14 Suppl):Abstract 3024.

    Pandolfi PP (2001) Histone deacetylases and transcriptional therapy with their inhibitors. Cancer Chemother Pharmacol 48 (Suppl 1): S17eCS19.

    Park WH, Jung CW, Park JO, Kim K, Kim WS, Im YH, Lee MH, Kang WK, and Park K (2003) Trichostatin inhibits the growth of ACHN renal cell carcinoma cells via cell cycle arrest in association with p27, or apoptosis. Int J Oncol 22: 1129eC1134.

    Patnaik A, Rowinsky EK, Villalona MA, Hammond LA, Britten CD, Siu LL, Goetz A, Felton SA, Burton S, Valone FH, et al. (2002) A phase I study of pivaloyloxymethyl butyrate, a prodrug of the differentiating agent butyric acid, in patients with advanced solid malignancies. Clin Cancer Res 8: 2142eC2148.

    Peart MJ, Tainton KM, Ruefli AA, Dear AE, Sedelies KA, O'Reilly LA, Waterhouse NJ, Trapani JA, and Johnstone RW (2003) Novel mechanisms of apoptosis induced by histone deacetylase inhibitors. Cancer Res 63: 4460eC4471.

    Pei XY, Dai Y, and Grant S (2004) Synergistic induction of oxidative injury and apoptosis in human multiple myeloma cells by the proteasome inhibitor bortezomib and histone deacetylase inhibitors. Clin Cancer Res 10: 3839eC3852.

    Penner E, Muller S, Zimmermann D, and Van Regenmortel MH (1987) High prevalence of antibodies to histones among patients with primary biliary cirrhosis. Clin Exp Immunol 70: 47eC52.

    Piekarz R and Bates S (2004) A review of depsipeptide and other histone deacetylase inhibitors in clinical trials. Curr Pharm Des 10: 2289eC2298.

    Piekarz RL, Robey R, Sandor V, Bakke S, Wilson WH, Dahmoush L, Kingma DM, Turner ML, Altemus R, and Bates SE (2001) Inhibitor of histone deacetylation, depsipeptide (FR901228), in the treatment of peripheral and cutaneous T-cell lymphoma: a case report. Blood 98: 2865eC2868.

    Piscitelli SC, Thibault A, Figg WD, Tompkins A, Headlee D, Lieberman R, Samid D, and Myers CE (1995) Disposition of phenylbutyrate and its metabolites, phenylacetate and phenylacetylglutamine. J Clin Pharmacol 35: 368eC373.

    Plumb JA, Finn PW, Williams RJ, Bandara MJ, Romero MR, Watkins CJ, La Thangue NB, and Brown R (2003) Pharmacodynamic response and inhibition of growth of human tumor xenografts by the novel histone deacetylase inhibitor PXD101. Mol Cancer Ther 2: 721eC728.

    Plumb JA, Steele N, Finn PW, and Brown R (2004) Epigenetic approaches to cancer therapy. Biochem Soc Trans 32: 1095eC1097.

    Prakash S, Foster BJ, Meyer M, Wozniak A, Heilbrun LK, Flaherty L, Zalupski M, Radulovic L, Valdivieso M, and LoRusso PM (2001) Chronic oral administration of CI-994: a phase 1 study. Investig New Drugs 19: 1eC11.

    Qian DZ, Wang X, Kachhap SK, Kato Y, Wei Y, Zhang L, Atadja P and Pili R (2004) The histone deacetylase inhibitor NVP-LAQ824 inhibits angiogenesis and has a greater antitumor effect in combination with the vascular endothelial growth factor receptor tyrosine kinase inhibitor PTK787/ZK222584. Cancer Res 64: 6626eC6634.

    Rahmani M, Yu C, Dai Y, Reese E, Ahmed W, Dent P, and Grant S (2003) Coadministration of the heat shock protein 90 antagonist 17-allylamino-17-demethoxygeldanamycin with suberoylanilide hydroxamic acid or sodium butyrate synergistically induces apoptosis in human leukemia cells. Cancer Res 63: 8420eC8427.

    Reid T, Valone F, Lipera W, Irwin D, Paroly W, Natale R, Sreedharan S, Keer H, Lum B, Scappaticci F, et al. (2004) Phase II trial of the histone deacetylase inhibitor pivaloyloxymethyl butyrate (Pivanex, AN-9) in advanced non-small cell lung cancer. Lung Cancer 45: 381eC386.

    Reid T, Weeks A, Vakil T, Cosgriff T, Harper F, Valone F, Magnuson A, and Bhatnagar A (2004) Dose escalation study of pivanex (a histone deacetylase inhibitor) in combination with docetaxel for advanced non-small cell lung cancer. J Clin Oncol 22 (14 Suppl):Abstract 7279.

    Remiszewski SW (2002) Recent advances in the discovery of small molecule histone deacetylase inhibitors. Curr Opin Drug Discov Dev 5: 487eC499.

    Remiszewski SW (2003) The discovery of NVP-LAQ824: from concept to clinic. Curr Med Chem 10: 2393eC2402.

    Remiszewski SW, Sambucetti LC, Atadja P, Bair KW, Cornell WD, Green MA, Howell KL, Jung M, Kwon P, Trogani N, et al. (2002) Inhibitors of human histone deacetylase: synthesis and enzyme and cellular activity of straight chain hydroxamates. J Med Chem 45: 753eC757.

    Richon VM, Emiliani S, Verdin E, Webb Y, Breslow R, Rifkind RA, and Marks PA (1998) A class of hybrid polar inducers of transformed cell differentiation inhibits histone deacetylases. Proc Natl Acad Sci USA 95: 3003eC3007.

    Richon VM, Sandhoff TW, Rifkind RA, and Marks PA (2000) Histone deacetylase inhibitor selectively induces p21WAF1 expression and gene-associated histone acetylation. Proc Natl Acad Sci USA 97: 10014eC10019.

    Richon VM, Webb Y, Merger R, Sheppard T, Jursic B, Ngo L, Civoli F, Breslow R, Rifkind RA, and Marks PA (1996) Second generation hybrid polar compounds are potent inducers of transformed cell differentiation. Proc Natl Acad Sci USA 93: 5705eC5708.

    Richon VM, Zhou X, Rifkind RA, and Marks PA (2001) Histone deacetylase inhibitors: development of suberoylanilide hydroxamic acid (SAHA) for the treatment of cancers. Blood Cells Mol Dis 27: 260eC264.

    Rosato RR, Almenara JA, Cartee L, Betts V, Chellappan SP, and Grant S (2002) The cyclin-dependent kinase inhibitor flavopiridol disrupts sodium butyrate-induced p21WAF1/CIP1 expression and maturation while reciprocally potentiating apoptosis in human leukemia cells. Mol Cancer Ther 1: 253eC266.

    Rosato RR, Almenara JA, Dai Y, and Grant S (2003) Simultaneous activation of the intrinsic and extrinsic pathways by histone deacetylase (HDAC) inhibitors and tumor necrosis factor-related apoptosis-inducing ligand (TRAIL) synergistically induces mitochondrial damage and apoptosis in human leukemia cells. Mol Cancer Ther 2: 1273eC1284.

    Rosato RR, Almenara JA, and Grant S (2003) The histone deacetylase inhibitor MS-275 promotes differentiation or apoptosis in human leukemia cells through a process regulated by generation of reactive oxygen species and induction of p21CIP1/WAF1 1. Cancer Res 63: 3637eC3645.

    Rosato RR, Almenara JA, Yu C, and Grant S (2004) Evidence of a functional role for p21WAF1/CIP1 down-regulation in synergistic antileukemic interactions between the histone deacetylase inhibitor sodium butyrate and flavopiridol. Mol Pharmacol 65: 571eC581.

    Rosato RR and Grant S (2003) Histone deacetylase inhibitors in cancer therapy. Cancer Biol Ther 2: 30eC37.

    Rosato RR, Wang Z, Gopalkrishnan RV, Fisher PB, and Grant S (2001) Evidence of a functional role for the cyclin-dependent kinase-inhibitor p21WAF1/CIP1/MDA6 in promoting differentiation and preventing mitochondrial dysfunction and apoptosis induced by sodium butyrate in human myelomonocytic leukemia cells (U937). Int J Oncol 19: 181eC191.

    Roth SY, Denu JM, and Allis CD (2001) Histone acetyltransferases. Annu Rev Biochem 70: 81eC120.

    Rowinsky EK, Pacey S, Patnaik A, O'Donnell A, Mita MM, Atadja P, Peng B, Dugan M, Scott JW, and De Bono JS (2004) A phase I, pharmacokinetic (PK) and pharmacodynamic (PD) study of a novel histone deacetylase (HDAC) inhibitor LAQ824 in patients with advanced solid tumors. J Clin Oncol 22 (14 Suppl): Abstract 3022.

    Ruefli AA, Ausserlechner MJ, Bernhard D, Sutton VR, Tainton KM, Kofler R, Smyth MJ, and Johnstone RW (2001) The histone deacetylase inhibitor and chemotherapeutic agent suberoylanilide hydroxamic acid (SAHA) induces a cell-death pathway characterized by cleavage of Bid and production of reactive oxygen species. Proc Natl Acad Sci USA 98: 10833eC10838.

    Ryan QC, Headlee D, Sparreboom A, Figg W, Zhai S, Trepel J, Murgo A, Elsayed Y, Karp J, and Sausville E (2003) A phase I trial of an oral histone deacetylase inhibitor, MS-275, in advanced solid tumor and lymphoma patients. Proc Am Soc Clin Oncol 22: 200 (Abstract 802).

    Saito A, Yamashita T, Mariko Y, Nosaka Y, Tsuchiya K, Ando T, Suzuki T, Tsuruo T, and Nakanishi O (1999) A synthetic inhibitor of histone deacetylase, MS-27eC275, with marked in vivo antitumor activity against human tumors. Proc Natl Acad Sci USA 96: 4592eC4597.

    Sandor V, Bakke S, Robey RW, Kang MH, Blagosklonny MV, Bender J, Brooks R, Piekarz RL, Tucker E, Figg WD, et al. (2002) Phase I trial of the histone deacetylase inhibitor, depsipeptide (FR901228, NSC 630176), in patients with refractory neoplasms. Clin Cancer Res 8: 718eC728.

    Sandor V, Senderowicz A, Mertins S, Sackett D, Sausville E, Blagosklonny MV, and Bates SE (2000) P21-dependent g(1)arrest with downregulation of cyclin D1 and upregulation of cyclin E by the histone deacetylase inhibitor FR901228. Br J Cancer 83: 817eC825.

    Sasakawa Y, Naoe Y, Inoue T, Sasakawa T, Matsuo M, Manda T, and Mutoh S (2002) Effects of FK228, a novel histone deacetylase inhibitor, on human lymphoma U-937 cells in vitro and in vivo. Biochem Pharmacol 64: 1079eC1090.

    Sasakawa Y, Naoe Y, Inoue T, Sasakawa T, Matsuo M, Manda T, and Mutoh S (2003a) Effects of FK228, a novel histone deacetylase inhibitor, on tumor growth and expression of p21 and c-myc genes in vivo. Cancer Lett 195: 161eC168.

    Sasakawa Y, Naoe Y, Noto T, Inoue T, Sasakawa T, Matsuo M, Manda T, and Mutoh S (2003b) Antitumor efficacy of FK228, a novel histone deacetylase inhibitor, depends on the effect on expression of angiogenesis factors. Biochem Pharmacol 66: 897eC906.

    Sawa H, Murakami H, Ohshima Y, Murakami M, Yamazaki I, Tamura Y, Mima T, Satone A, Ide W, Hashimoto I, et al. (2002) Histone deacetylase inhibitors such as sodium butyrate and trichostatin A inhibit vascular endothelial growth factor (VEGF) secretion from human glioblastoma cells. Brain Tumor Pathol 19: 77eC81.

    Schreiber SL and Bernstein BE (2002) Signaling network model of chromatin. Cell 111: 771eC778.

    Schwer B, North BJ, Frye RA, Ott M, and Verdin E (2002) The human silent information regulator (Sir)2 homologue hSIRT3 is a mitochondrial nicotinamide adenine dinucleotide-dependent deacetylase. J Cell Biol 158: 647eC657.

    Shaker S, Bernstein M, Momparler LF, and Momparler RL (2003) Preclinical evaluation of antineoplastic activity of inhibitors of DNA methylation (5-aza-2'-deoxycytidine) and histone deacetylation (trichostatin A, depsipeptide) in combination against myeloid leukemic cells. Leuk Res 27: 437eC444.

    Siavoshian S, Segain JP, Kornprobst M, Bonnet C, Cherbut C, Galmiche JP, and Blottiere HM (2000) Butyrate and trichostatin A effects on the proliferation/ differentiation of human intestinal epithelial cells: induction of cyclin D3 and p21 expression. Gut 46: 507eC514.

    Singh SB, Zink DL, Liesch JM, Dombrowski AW, Darkin-Rattray SJ, Schmatz DM, and Goetz MA (2001) Structure, histone deacetylase and antiprotozoal activities of apicidins B and C, congeners of apicidin with proline and valine substitutions. Org Lett 3: 2815eC2818.

    Singh SB, Zink DL, Liesch JM, Mosley RT, Dombrowski AW, Bills GF, Darkin-Rattray SJ, Schmatz DM, and Goetz MA (2002) Structure and chemistry of apicidins, a class of novel cyclic tetrapeptides without a terminal alpha-keto epoxide as inhibitors of histone deacetylase with potent antiprotozoal activities. J Org Chem 67: 815eC825.

    Sowa Y, Orita T, Hiranabe-Minamikawa S, Nakano K, Mizuno T, Nomura H, and Sakai T (1999) Histone deacetylase inhibitor activates the p21/WAF1/Cip1 gene promoter through the Sp1 sites. Ann NY Acad Sci 886: 195eC199.

    Sowa Y, Orita T, Minamikawa-Hiranabe S, Mizuno T, Nomura H, and Sakai T (1999) Sp3, but not Sp1, mediates the transcriptional activation of the p21/WAF1/Cip1 gene promoter by histone deacetylase inhibitor. Cancer Res 59: 4266eC4270.

    Strait KA, Dabbas B, Hammond EH, Warnick CT, Iistrup SJ, and Ford CD (2002) Cell cycle blockade and differentiation of ovarian cancer cells by the histone deacetylase inhibitor trichostatin A are associated with changes in p21, Rb and Id proteins. Mol Cancer Ther 1: 1181eC1190.

    Su GH, Sohn TA, Ryu B, and Kern SE (2000) A novel histone deacetylase inhibitor identified by high-throughput transcriptional screening of a compound library. Cancer Res 60: 3137eC3142.

    Suzuki T, Ando T, Tsuchiya K, Fukazawa N, Saito A, Mariko Y, Yamashita T, and Nakanishi O (1999) Synthesis and histone deacetylase inhibitory activity of new benzamide derivatives. J Med Chem 42: 3001eC3003.

    Takai N, Desmond JC, Kumagai T, Gui D, Said JW, Whittaker S, Miyakawa I, and Koeffler HP (2004) Histone deacetylase inhibitors have a profound antigrowth activity in endometrial cancer cells. Clin Cancer Res 10: 1141eC1149.

    Tang R, Faussat AM, Majdak P, Perrot JY, Chaoui D, Legrand O, and Marie JP (2004) Valproic acid inhibits proliferation and induces apoptosis in acute myeloid leukemia cells expressing P-gp and MRP1. Leukemia 18: 1246eC1251.

    Thelen P, Schweyer S, Hemmerlein B, Wuttke W, Seseke F, and Ringert RH (2004) Expressional changes after histone deacetylase inhibition by valproic acid in LNCaP human prostate cancer cells. Int J Oncol 24: 25eC31.

    Thiagalingam S, Cheng KH, Lee HJ, Mineva N, Thiagalingam A, and Ponte JF (2003) Histone deacetylases: unique players in shaping the epigenetic histone code. Ann NY Acad Sci 983: 84eC100.

    Thibault A, Cooper MR, Figg WD, Venzon DJ, Sartor AO, Tompkins AC, Weinberger MS, Headlee DJ, McCall NA, Samid D, and et al. (1994) A phase I and pharmacokinetic study of intravenous phenylacetate in patients with cancer. Cancer Res 54: 1690eC1694.

    Thibault A, Samid D, Cooper MR, Figg WD, Tompkins AC, Patronas N, Headlee DJ, Kohler DR, Venzon DJ, and Myers CE (1995) Phase I study of phenylacetate administered twice daily to patients with cancer. Cancer 75: 2932eC2938.

    Timmermann S, Lehrmann H, Polesskaya A, and Harel-Bellan A (2001) Histone acetylation and disease. Cell Mol Life Sci 58: 728eC736.

    Undevia SD, Kindler HL, Janisch L, Olson SC, Schilsky RL, Vogelzang NJ, Kimmel KA, Macek TA, and Ratain MJ (2004) A phase I study of the oral combination of CI-994, a putative histone deacetylase inhibitor and capecitabine. Ann Oncol 15: 1705eC1711.

    Vrana JA, Decker RH, Johnson CR, Wang Z, Jarvis WD, Richon VM, Ehinger M, Fisher PB, and Grant S (1999) Induction of apoptosis in U937 human leukemia cells by suberoylanilide hydroxamic acid (SAHA) proceeds through pathways that are regulated by Bcl-2/Bcl-XL, c-Jun and p21CIP1, but independent of p53. Oncogene 18: 7016eC7025.

    Wade PA (2001) Transcriptional control at regulatory checkpoints by histone deacetylases: molecular connections between cancer and chromatin. Hum Mol Genet 10: 693eC698.

    Wang ZM, Hu J, Zhou D, Xu ZY, Panasci LC, and Chen ZP (2002) Trichostatin A inhibits proliferation and induces expression of p21WAF and p27 in human brain tumor cell lines. Ai Zheng 21: 1100eC1105.

    Warrener R, Beamish H, Burgess A, Waterhouse NJ, Giles N, Fairlie D, and Gabrielli B (2003) Tumor cell-selective cytotoxicity by targeting cell cycle checkpoints. FASEB J 17: 1550eC1552.

    Williams RJ (2001) Trichostatin A, an inhibitor of histone deacetylase, inhibits hypoxia-induced angiogenesis. Expert Opin Investig Drugs 10: 1571eC1573.

    Wozniak A, O'Shaughnessy, Fiorica J, and Grove W (1999) Phase II trial of CI-994 in patients (pts) with advanced nonsmall cell lung cancer (NSCLC) (Abstract 1878). Proceedings of the 35th Annual Meeting of the American Society for Clinical Oncology; 1999 May 15eC18; Atlanta, Georgia. American Society for Clinical Oncology, Alexandria, VA.

    Yang X, Phillips DL, Ferguson AT, Nelson WG, Herman JG, and Davidson NE (2001) Synergistic activation of functional estrogen receptor (ER)-alpha by DNA methyltransferase and histone deacetylase inhibition in human ER-alpha-negative breast cancer cells. Cancer Res 61: 7025eC7029.

    Yao Q, Nishiuchi R, Li Q, Kumar AR, Hudson WA, and Kersey JH (2003) FLT3 expressing leukemias are selectively sensitive to inhibitors of the molecular chaperone heat shock protein 90 through destabilization of signal transduction-associated kinases. Clin Cancer Res 9: 4483eC4493.

    Yokota T, Matsuzaki Y, Miyazawa K, Zindy F, Roussel MF, and Sakai T (2004) Histone deacetylase inhibitors activate INK4d gene through Sp1 site in its promoter. Oncogene 23: 5340eC5349.

    Yoshida M, Furumai R, Nishiyama M, Komatsu Y, Nishino N, and Horinouchi S (2001). Histone deacetylase as a new target for cancer chemotherapy. Cancer Chemother Pharmacol 48 (Suppl 1): S20eCS26.

    Yoshida M, Horinouchi S, and Beppu T (1995) Trichostatin A and trapoxin: novel chemical probes for the role of histone acetylation in chromatin structure and function. Bioessays 17: 423eC430.

    Yu C, Rahmani M, Almenara J, Subler M, Krystal G, Conrad D, Varticovski L, Dent P, and Grant S (2003) Histone deacetylase inhibitors promote STI571-mediated apoptosis in STI571-sensitive and -resistant Bcr/Abl+ human myeloid leukemia cells. Cancer Res 63: 2118eC2126.

    Yu C, Rahmani M, Conrad D, Subler M, Dent P, and Grant S (2003) The proteasome inhibitor bortezomib interacts synergistically with histone deacetylase inhibitors to induce apoptosis in Bcr/Abl+ cells sensitive and resistant to STI571. Blood 102: 3765eC3774.

    Yu C, Subler M, Rahmani M, Reese E, Krystal G, Conrad D, Dent P, and Grant S (2003) Induction of apoptosis in BCR/ABL+ cells by histone deacetylase inhibitors involves reciprocal effects on the RAF/MEK/ERK and JNK pathways. Cancer Biol Ther 2: 544eC551.

    Zalupski M, O'Shaughnessy J, Vukelja S, Shields A, Diener K, and Grove W (2000) Phase II trial of II-994 in patients (PTS) with advanced pancreatic cancer (APC). (Abstract 1115). Proceedings of the 36th Annual Meeting of the American Society for Clinical Oncology; 2000 May 20eC23; New Orleans, Louisiana. American Society for Clinical Oncology, Alexandria, VA.

    Zhang XD, Gillespie SK, Borrow JM, and Hersey P (2003) The histone deacetylase inhibitor suberic bishydroxamate: a potential sensitizer of melanoma to TNF-related apoptosis-inducing ligand (TRAIL) induced apoptosis. Biochem Pharmacol 66: 1537eC1545.

    Zhang XD, Gillespie SK, Borrow JM, and Hersey P (2004a) The histone deacetylase inhibitor suberic bishydroxamate regulates the expression of multiple apoptotic mediators and induces mitochondria-dependent apoptosis of melanoma cells. Mol Cancer Ther 3: 425eC435.

    Zhang Y, Adachi M, Zhao X, Kawamura R, and Imai K (2004b) Histone deacetylase inhibitors FK228, N-(2-aminophenyl)-4-[N-(pyridin-3-yl-methoxycarbonyl)aminomethyl]benzamide and m-carboxycinnamic acid bis-hydroxamide augment radiation-induced cell death in gastrointestinal adenocarcinoma cells. Int J Cancer 110: 301eC308.

    Zhou Y, Santoro R, and Grummt I (2002) The chromatin remodeling complex NoRC targets HDAC1 to the ribosomal gene promoter and represses RNA polymerase I transcription. EMBO (Eur Mol Biol Organ) J 21: 4632eC4640.

    Zhu WG, Dai Z, Ding H, Srinivasan K, Hall J, Duan W, Villalona-Calero MA, Plass C, and Otterson GA (2001a) Increased expression of unmethylated CDKN2D by 5-aza-2'-deoxycytidine in human lung cancer cells. Oncogene 20: 7787eC7796.

    Zhu WG, Lakshmanan RR, Beal MD, and Otterson GA (2001b) DNA methyltransferase inhibition enhances apoptosis induced by histone deacetylase inhibitors. Cancer Res 61: 1327eC1333.

    Zhu WG and Otterson GA (2003) The interaction of histone deacetylase inhibitors and DNA methyltransferase inhibitors in the treatment of human cancer cells. Curr Med Chem Anti-Canc Agents 3: 187eC199.

    Zimra Y, Nudelman A, Zhuk R, Rabizadeh E, Shaklai M, Aviram A, and Rephaeli A (2000) Uptake of pivaloyloxymethyl butyrate into leukemic cells and its intracellular esterase-catalyzed hydrolysis. J Cancer Res Clin Oncol 126: 693eC698.

    Zimra Y, Wasserman L, Maron L, Shaklai M, Nudelman A, and Rephaeli A (1997) Butyric acid and pivaloyloxymethyl butyrate, AN-9, a novel butyric acid derivative, induce apoptosis in HL-60 cells. J Cancer Res Clin Oncol 123: 152eC160.

作者: Milin R. Acharya, Alex Sparreboom, Jegen Venitz, a 2007-5-15
医学百科App—中西医基础知识学习工具
  • 相关内容
  • 近期更新
  • 热文榜
  • 医学百科App—健康测试工具