Review Targeting DNA methylation for epigenetic therapy

Xiaojing Yang1, Fides Lay1,2, Han Han1,3 and Peter A. Jones1

1 Department of Urology, Norris Comprehensive Cancer Center, Keck School of Medicine, University of Southern California, Los Angeles, CA, USA 2 Program in Genetics, Molecular and Cellular Biology, Norris Comprehensive Cancer Center, Keck School of Medicine, University of Southern California, Los Angeles, CA, USA 3 Department of Pharmacology and Pharmaceutical Sciences, Norris Comprehensive Cancer Center, Keck School of Medicine, University of Southern California, Los Angeles, CA, USA

Patterns of DNA methylation are established during and also CpHpG nucleotide sequences in embryonic stem embryonic development and faithfully copied through cells (ES) [2]. In somatic cells, most CpG dinucleotides are somatic cell divisions. Based on current understanding methylated except those located in CpG islands [3],which of DNA methylation and other interrelated epigenetic are defined as regions of DNA >500 bp with a GC content modifications, a comprehensive view of the ‘epigenetic 55%, and a ratio of observed and expected CpG >0.65 [4]. landscape’ and cancer epigenome is evolving. The can- Nearly 60% of mammalian gene promoters are located in cer methylome is highly disrupted, making DNA meth- CpG islands [5].DNAmethylationoftheseCpG-rich ylation an excellent target for anticancer therapies. promoters silences gene expression by changing the ac- During the last few decades, an increasing number of cessibility of DNA to transcription factors (TFs) or by drugs targeting DNA methylation have been developed recruiting additional silencing-associated proteins [6]. to increase efficacy and stability and to decrease toxicity. Nevertheless, there is also growing evidence suggesting The earliest and the most successful epigenetic drug to that even in CpG-poor regions such as the Oct4 promoter date, 5-Azacytidine, is currently recommended as the DNA methylation still has a role in gene regulation [7,8]. first-line treatment of high-risk myelodysplastic syn- Patterns of DNA methylation are established and main- dromes (MDS). Encouraging results from clinical trials tained by DNA methyltransferases (DNMTs). During early have prompted further efforts to elucidate epigenetic embryogenesis, de novo DNA methylation is mediated by alterations in cancer, and to subsequently develop new DNMT3A and DNMT3B associated with DNMT3L, which epigenetic therapies. This review delineates the latest lacks a methyltransferase domain [9–11]. To maintain pat- cancer epigenetic models, the recent discovery of hypo- terns of DNA methylation in daughter cells, ubiquitin-like, methylation agents as well as their application in the containing PHD and RING finger domains 1 (UHRF1) clinic. recognizes hemi-methylated DNA and directs DNMT1 to methylate the appropriate cytosine in newly synthesized The epigenetic landscape DNA strands during successive replications [12,13]. Recent Primary heritable genetic information is dictated by the studies have proposed an updated model, suggesting that DNA sequence packed into chromatin. The basic repeating DNMT3A/B are also required for the maintenance of pat- unit of chromatin is a nuclesome consisting of 146–147 terns of DNA methylation in somatic cells, particularly of base pairs (bp) of DNA wrapped around a histone octamer. repeat regions and imprinted genes (Figure 1a) [6,14]. DNA and histone proteins can undergo various post- In contrast to DNA methylation, the patterns of histone synthetic modifications. These modifications, along with modifications are more labile. The N-termini of histones histone variants and nucleosome remodelers, can strongly can undergo various post-synthetic modifications, includ- influence chromatin structure and transcriptional regula- ing a diverse combination of methylation, acetylation, tion without changing the underlying DNA sequence, a ubiquitylation, phosphorylation and sumoylation. In re- process called ‘epigenetic regulation’. Epigenetic regula- cent years, the development of modification-specific anti- tion can be separated into three inter-related layers: bodies in chromatin immunoprecipitation (ChIP), coupled nucleosome positioning, histone modifications, and DNA with microarray (ChIP-chip) or direct sequencing technol- methylation [1]. ogy (ChIP-seq), has revolutionized assessment of the global DNA methylation is probably the most extensively distribution of histone modifications and their roles in studied epigenetic mark. It plays an important part in gene regulation [15–17]. For instance, H3K4me3 has been genomic imprinting, in X-chromosome inactivation, and in found to occur near the transcription start sites (TSS) of the silencing of retrotransposon, repetitive elements and genes or miRNA, and is positively correlated with tran- tissue-specific genes. In mammalian cells, a methyl group scription level [18–21]. Similarly, H3K9/14 acetylation is is covalently added to cytosine in the context of cytosine- highly associated with gene transcriptional activation phosphate-guanine (CpG) dinucleotides in somatic cells [22]. By contrast, H3K9me and H3K27me are enriched in the vicinity of suppressed or silenced gene promoters Corresponding author: Jones, P.A. ([email protected]). (Figure 1a) [15]. The enzymes that catalyze these

536 0165-6147/$ – see front matter ß 2010 Elsevier Ltd. All rights reserved. doi:10.1016/j.tips.2010.08.001 Trends in Pharmacological Sciences, November 2010, Vol. 31, No. 11 [(Figure_1)TD$IG]Review Trends in Pharmacological Sciences Vol.31 No.11

Figure 1. Epigenetic regulation in normal and cancer cells. Shown are schematic promoters. Arrows represent transcription start sites (TSS); filled circles represent methylated CpG dinucleotide, and empty circles represent unmethylated CpG dinucleotide. In normal cells (a), genes such as MLH1, CDKN2A are, in general, unmethylated and packaged with active modified histone proteins containing active marks (e.g. H3K4me3) as well as histone variants (e.g. H2A.Z). These epigenetic modifications constitute an ‘open’ chromatin structure which, with a nucleosome-depleted region (NDR), favor transcription. In other genomic regions, such as in the repetitive elements, the CpG sites are methylated and thereby maintain a closed chromatin structure. In cancer cells (b), epigenetic modifications are disrupted. Besides cancer-specific hypomethylation (e.g. in repetitive sequences), there are two inter-related epigenetic mechanisms to repress gene expression. Some genes (e.g. FBXO32) could be recognized by polycomb proteins such as EZH2 which catalyses H3K27 methylation, and are consequently repressed. By contrast, CpG sites within gene promoters could undergo de novo methylation by DNMT3A/B, which are compartmentalized to these regions to complete the methylation process. The methylated CpG sites attract methyl-binding proteins such as MBD, which is coupled with HDAC proteins to remove histone acetylation as well as histone methyltransferase to methylate H3K9. Associated with all the repressive factors shown, nucleosomes cover the promoter region, generating a tightly closed chromatin status to shut down gene expression. modifications and the crosstalk between these complexes cing the suppressive nature of promoters (Figure 1) [36– have been summarized in several important review arti- 37]. The interplay between DNA methylation and histone cles [23–26]. modifications can be further observed in ES cells. Some Other crucial components of the epigenetic landscape studies have proposed that histone modifications and as- are histone variants which partially determine nucleosome sociated proteins (e.g., EZH2 or G9a) take part in establish- stability and ultimately affect gene expression. H2A.Z is ing de novo patterns of DNA methylation during early commonly localized in promoter regions. The differences in development [38–41]. the amino-acid sequence of H2A.Z compared with H2A might affect the interactions of the protein with the H3/ Cancer H4 tetramer, thereby altering nucleosome stability [27– It has been documented that epigenetic alterations are 28]. In mammalian cells, the role of H2A.Z in gene regula- involved in the initiation and progression of cancer in tion is incompletely understood, but, there is increasing addition to abnormal genetic events. Early studies that evidence suggesting that H2A.Z assists in the transcrip- measured the global content of 5-methylcytosine of tumors tion initiation of a special group of genes such as endoplas- showed that hypomethylation was a common feature of mic reticulum (ER) target genes [29–32]. carcinogenesis, leading to abnormal chromosomal instabil- The interplay between histone modifications, histone ity and transcriptional regulation [42–44]. However, most variants, and DNA methylation constitutes a global regu- cancer epigenetic studies subsequently concentrated on lation network. Extensive studies have elucidated the focal CpG island hypermethylation in cancer, and revealed biological and functional interactions among these three many tumor suppressor genes, cellular functional genes, epigenetic layers. For example, DNA methylation anti- and miRNAs silenced by promoter DNA methylation [45– correlates with the active mark H3K4me3, H3ac or 48]. Recent genome-wide studies have demonstrated dis- H2A.Z [33–35]. Meanwhile, highly methylated promoter tinct patterns of DNA methylation in cancerous tissues regions are bound by methylcytosine binding proteins such compared with their normal counterparts [49–52]. as MeCP2, which subsequently recruits histone deacety- The detailed mechanisms by which these discrete lase (HDAC) to deacetylate histones, thus further reinfor- regions undergo hyper- or hypomethylation are unclear.

537 Review Trends in Pharmacological Sciences Vol.31 No.11

Early evidence suggested that elevated DNA methyltrans- interaction between DNA and DNMTs through the nitro- ferase levels might trigger hypermethylation of tumor sup- gen in the 5 position of the modified pyrimidine, and traps pressor gene promoters, which would consequentially result DNMTs for proteosomal degradation [74–75] (Figure 3b). in proliferation of cancer cells [53]. In addition to this The depletion of DNMTs results in the passive loss of ‘selection’ model, an alternative mechanism has been pro- cytosine methylation in the daughter cells after replica- posed that takes advantage of the current genome-wide tion. This is associated with reduced H3K9me3, increased epigenetic studies in stem cells. Investigators have sug- H3ac and H3K4me3 modifications around gene promoter gested that the establishment of aberrant epigenetic profiles regions, as well as the formation of a nucleosome-deficient in cancer undergoes a process similar to epigenetic repro- region [66,76]. A recent genome-wide study of alterations gramming during development [54]. During cancer initia- in the epigenetic landscape after 5-Aza-CR treatment fur- tion, the promoters of genes (which are repressed by histone ther validated this concept [77]. The demethylation func- H3K27me3 in normal differentiated cells) might become tion of 5-Aza-CR and 5-Aza-CdR is most evident at low methylated and thereby set up for long-term silencing. This drug concentration because the drugs exhibit greater cy- ‘epigenetic switch’ could be regulated by the cooperation of totoxicity, interfere with DNA synthesis, and cause DNA polycomb proteins and DNMTs [55–57] (Figure 1b). damage at higher concentrations [78]. The S-phase is In cooperation with DNA methylation, other epigenetic required for the selective and effective incorporation of mechanisms also exhibit abnormal regulation in cancer. these two drugs into the DNA of rapidly proliferating cells, For example, histone deacetylases (HDACs) are often thereby limiting unwanted hypomethylation in cells un- found to be overexpressed in various types of cancer, dergoing the normal cell cycle. resulting in histone deacetylation around the TSS region Besides inhibiting DNMTs, 5-Aza-CR incorporates into and the formation of a more compact structure to silence RNA and interrupts normal cellular processes by inducing genes [58–59]. In addition, H3K4me is also selectively ribosomal disassembly and preventing the translation of demethylated by the histone lysine demethylase (LSD1), oncogenic proteins [73,79]. The ability of 5-Aza-CR to be which is upregulated in cancer, making it a potential drug incorporated into DNA and RNA increases its side effects target [60–61]. In some loci, polycomb-group (PcG) proteins in vitro and in vivo because it can function in resting and associated with H3K27me work independently of DNA dividing cells [80]. However, 5-Aza-CR and 5-Aza-CdR are methylation to aberrantly repress genes in cancer cells readily hydrolyzed in aqueous solution and subject to [62–65]. Nucleosome occupancy is also switched from an deamination by cytidine deaminase. The instabilities of ‘open’ to a ‘covered’ status in gene regulation elements in these compounds inevitably present a challenge to their neoplastic cells [66–67] (Figure 1b). clinical applications. To improve the stability and efficacy of 5-Azanucleo- In vitro study of DNMT inhibitors (DNMTi) sides, several other cytidine analogs have been developed Epigenetic modifications play a crucial part in regulating (Figure 2a). For example, zebularine (a cytidine analog normal cells, but these processes are disrupted during that lacks an amino group in the 4 position of the pyrimi- tumorigenesis. The relatively reversible character of epi- dine ring) can inhibit DNMTs and cytidine deaminase after genetic alterations (in contrast to genetic changes) has oral administration [81–83]. Studies have shown that inspired the development of therapeutic strategies target- zebularine induces hypomethylation in breast cancer cell ing various epigenetic components. Among them, DNA lines and reactivates silenced tumor suppressor genes [84– methylation and its associated enzymes have been well- 85]. The inefficient metabolic activation of this compound studied. The understanding of their fundamental mecha- has, however, delayed its clinical use as a single agent. As nism of action and correlation with other epigenetic mod- an inhibitor of cytidine deaminase, its co-administration ifications makes them attractive drug targets. has increased the efficacy of 5-Aza-CdR [82]. The cytidine analog 5-Fluoro-20-deoxycytidine (5-F-CdR) had also been Nucleoside analogs demonstrated to have hypomethylating activity in mouse 5-Azacytidine (5-Aza-CR) and 5-aza-20-deoxycitidine (5- cells as well as in human breast and lung carcinoma cells Aza-CdR) are the two most potent DNMTi. They have [68,86]. Clinical studies further showed that co-adminis- been approved by the Food and Drug Administration tration of 5-F-CdR with the cytidine deaminase inhibitor (FDA) in the USA for the treatment of myeloid malignan- tetrahydrouridine (THU) improved the stability of 5-F- cies (Figure 2a). They were first synthesized as cytotoxic CdR [87]. The therapeutic potential of another stable agents. In the 1980s, these compounds were found to have analog, dihydro-5-azacytidine (DHAC), was also assessed hypomethylating activity after incorporation into the DNA for the treatment of malignant mesothelioma, but results of actively replicating tumor cells [68–70]. on this compound’s clinical efficacy have been inconsistent Upon transport into cells by the human concentrative [88]. nucleoside transporter 1 (hCNT1), 5-Aza-CR and 5-Aza- The effort to improve the stability of DNMTi includes CdR are phosphorylated by different kinases, converting the development of prodrugs of the nucleoside analogs. A them to their active triphosphate forms, 5-Aza-CTP and 5- preclinical study showed that NPEOC-DAC (a prodrug of Aza-dCTP, respectively [71–72] (Figure 3a). 5-Aza-CR can 5-Aza-CdR containing a 2-( p-nitrophenyl)ethoxycarbonyl be incorporated into RNA and DNA after the reduction of 5- (NPEOC) group at the 4 position of the pyrimidine ring) Aza-CDP by ribonucleotide reductase, whereas 5-Aza-CdR can be incorporated into DNA and inhibit DNMTs after its incorporates into DNA after its phosphorylation to 5-Aza- activation by human carboxylesterase 1 in a liver cancer dCTP [73]. The incorporated 5-azanucleoside disrupts the cell line. The NPEOC moiety protects 5-Aza-CdR from

538 [(Figure_2)TD$IG]Review Trends in Pharmacological Sciences Vol.31 No.11

Figure 2. Chemical structure of DNMTi:

A. Nucleoside analogs B. Non-nucleoside analogs. deamination, but the compound itself is less potent if Non-nucleoside analogs administered at the same concentration as 5-Aza-CdR. Unlike cytidine analogs, non-nucleoside DNMTis (Figure Moreover, the activity of NPEOC-DAC is dependent upon 2b) do not require incorporation into DNA, and thus might carboxylesterases, which are not expressed in all tissues. exhibit less cytotoxicity. Some of the compounds assessed Further studies are required to explore the use of this for their potential to induce hypomethylation in solid compound in combination therapy [89]. Alternatively, tumors are hydralazine and procainamide, the widely used S110 (a dinucleotide containing the 5-azacytosine ring) vasodilator and antiarrhythmic agents, respectively [91]. has also been shown to improve the efficacy of 5-Aza- Hydralazine has been reported to block the activity of CdR by protecting it from deamination. The compound DNMTs by the interaction of its nitrogen atoms with the is well tolerated and can reduce the level of DNA methyla- Lys-162 and Arg-240 residues of the enzyme, whereas tion in the CDKN2A promoter region in xenografts [90]. procainamide acts similarly as a competitive inhibitor by

539 [(Figure_3)TD$IG]Review Trends in Pharmacological Sciences Vol.31 No.11

Figure 3. Metabolic pathway and mechanism of action of 5-Azanucleosides. (a) Human concentrated nucleoside transporter 1 (hCNT1) facilitates the entry of 5-Aza-CR and 5-Aza-CdR into cells, where they are phosphorylated by uridine-cytidine kinase and deoxycytidine kinase, respectively. 5-Aza-CMP and 5-Aza-dCMP are subsequently phosphorylated into their active triphosphate forms. 5 Aza-dCR can be incorporated solely into DNA, whereas 5-Aza-CR can be incorporated into RNA as well as DNA after the reduction of its 5-Aza-CDP form to the 5-Aza-dCDP form. Incorporation of 5-Azanucleosides into DNA induces hypomethylation of the daughter DNA strands, whereas incorporation of 5-Aza-CR into RNA disrupts crucial cellular process such as protein translation and causes ribosomal disassembly. 5-Azanucleosides, however, are also very unstable and can be deaminated to inactive uridine. (b). Schematic working model of 5-Azanucleosides. During DNA replication, 5-Azanucleosides are incorporated into DNA and trap DNMTs, which are subsequently targeted for proteosomal degradation. DNA-containing azanucleosides are hemi-methylated after the first round of DNA replication and become fully demethylated after several rounds of replication. Using hypomethylation agents, the silenced epigenetic modifications could be switched to an active status (e.g. H3K4me3 and AcH3). preferentially binding to DNMT1 [92–93]. These com- successful results encouraged the use of DNMTi to develop pounds, however, have limited DNA hypomethylation ac- novel clinical regimens. We have summarized several tivity in living cells [94]. The small molecule RG108 also recent important clinical trials in Table 1. shows the potential to reactivate tumor suppressor genes in human colon cancer cells [95–96]. Recently, the lipophil- Single-agent therapy ic, quinoline-based compound SGI-1027 was demonstrated 5-Azacytidine (Vidaza, Azacitidine) In early clinical trials, to be a novel DNMTi in vitro. In RKO cells, SGI-1027 5-Aza-CR was administered at maximum tolerated doses causes the degradation of DNMT1 and the demethylation (MTD) to patients with osteogenic sarcoma or other cancer- of the CDKN2A gene promoter as well as reactivating related , but showed unfavorable toxicity [100– silenced genes [97]. 101]. Increasing knowledge of the mechanism of action An alternative strategy to inhibit DNMT1 includes the of 5-Aza-CR indicated that lower dosages of 5-Aza-CR use of short-chain oligodeoxynucleotides and microRNAs. could act as DNMTi with minimal effects on DNA MG98 is a 20-bp antisense oligonucleotide that specifically synthesis [102]. 5-Aza-CR was approved by the FDA for binds to the 30 UTR of human DNMT1 mRNA to prevent its the treatment of MDS based on the positive results from translation. Despite promising results in preclinical stud- the GALGB9221 clinical trial [103]. A reported 60% of 5- ies, the clinical use of MG98 has not been validated [98]. Aza-CR-treated patients exhibited various levels of MicroRNA miR29a, which targets DNMT3A/B directly and response, whereas only 5% of patients in the supportive DNMT1 indirectly in a similar way to MG98, can reduce group showed hematological improvement (HI). 5-Aza-CR global DNA methylation and reactivate CDKN2B [99]. also benefited patients by delaying progression time to acute myeloid leukemia (AML), improving quality of life, Clinical data on DNMTi and prolonging overall survival in Refractory Anemia with Accumulated in vitro evidence has demonstrated that Excess Blasts (RAEB) or Refractory Anemia with Excess of DNMTi can reverse abnormal DNA methylation, indirectly Blasts in Transformation (RAEB-T) subgroups [104–105]. affect histone modifications, and eventually restore normal To further validate the efficiencies of 5-Aza-CR in MDS gene expression profiles in various cancer cell lines. These patients, the European AZA-001 trial was conducted for

540 Table 1. Selected Recent Clinical Trials with DNMT Inhibitors Review Drug(s) Indication(s) Author (Year) Total Results Comments Ref. Patients Azacitidine MDS Silverman 191 7% CR, 16% PR, Low dose treatment improved quality of life and survival rate; delayed [102, (2002/2006) 37% HI progression time to AML(CALGB9221). 104] MDS Fenaux (2009) 358 17% CR, 12% PR, 42% SD Improved OS compared to CCR in high risk MDS patients (AZA-001). [105] MDS Lyons (2009) 151 44%, 45%, 56% HI RBC transfusion independence achieved in 3 alternative dosing regimens. [109] MDS Martin (2009) 25 27% ORR Response rates are comparable to CALGB 9221 and AZA-001 trials but shorter OS in this 5 [110] days treatment regimen. MDS/AML Muller-Thomas 32 15.6% CR, 25% HI, Survival advantage in responder group of high-risk MDS [111] (2009) 34.4% SD MDS Musto (2010) 74 10.8% CR, 9.5% PR, Survival benefits in responder group. [108] 20.3% HI Decitabine MDS Kantarjian (2006) 170 9% CR, 13% HI, RBC transfusion independence and delayed progression time toward AML (D-0007). [112] 17% ORR MDS Kantarjian (2007) 95 34% CR Better response rate at low dose. [114] MDS Steensma (2009) 99 17% CR, 18% HI, Alternative dosing and significant difference in ORR compared to single-center trial by [115] 32% ORR Kantarjian et al, (ADOPT). Refractory Stewart (2009) 31 Relative reduction of tumor size but no correlation between PBMC DNA methylation and [117] Solid Tumor tumor DNA methylation. MG98 MDS/AML Klisovic (2009) 23 26% SD No clinical benefits observed in administered doses and no correlation [119] between DNMT1 downregulation with SD. Advanced Solid Plummer (2009) 33 Suppression of DNMT1 expression in PBMC in all doses with mild toxicity observed. [121] Malignancies Azacitidine in Combination Azacitidine + Valproic Acid + AML/MDS Soriano (2007) 53 22% CR, 42% ORR The increase of histone acetylation level is not correlated with VPA dose level. [122] all-trans-Retinoic Acid Azacitidine + Thalidomide MDS/AML Raza (2008) 40 42% HI, 14% SD High expression of cellular proliferation genes was observed in non-responders. [135] Azacitidine + Valproic Acid Advanced cancer Braiteh (2008) 55 25%SD Safe dosage of azacitidine up to 75mg/m2 in advanced malignancies. [124] Azacitidine + Refractory Lin (2009) 27 No clear benefits seen in the three dosing regimens tested. [126] Sodium Phenylbutyrate Solid Tumor Azacitidine + Valproic Acid MDS Voso (2009) 62 30.7% CR/PR, 15.4% Increasing therapeutic level of VPA promoted efficacy of azacitidine [123] HI, 38.5% SD Azacitidine + Entinostat MDS/AML Fandy (2009) 30 42%ORR Lack of association between reversal of DNA methylation and clinical response. [125] Azacitidine + Allogeneic SCT MDS/AML Jabbour 17 High response rate in patients with recurrent after HSCT . [136] (2009) Azacitidine + Lenalidomide MDS Sekeres (2010) 18 44% CR, 17% HI, Higher CR is associated with normal cytogenetics and lower methylation levels [129] Sciences Pharmacological in Trends 67% ORR Azacitidine + Cytarabine MDS/AML Borthakur (2010) 34 Limited clinical activity in patients with relapsed, refractory AML. [137] Azacitidine + Etanercept MDS/CMML Scott (2010) 32 72% ORR Increased response rate and duration of response observed. [138] Decitabine in Combination Decitabine + Valproic Acid Advanced Garcia-Manero 54 19% CR Lower pretreatment CDKN2B methylation level correlates with [127] leukemia (2006) higher response rate. Decitabine + Carboplatin Solid Tumor Appleton (2007) 33 A correlation between different doses of decitabine with PBC DNA demethylation. [132] Decitabine + Valproic Acid AML Blum (2007) 25 44% ORR VPA increased treatment-related toxicity; Re-expression of ER associated with clinical [128] response. Decitabine + AML Chowdhury 12 42% CR Treatment facilitated subsequent HSCT. [131] Gemtuzumab ozogamicin (2009) Decitabine + Allogeneic SCT MDS De Padua Silva (2009) 17 No increased toxicity due to decitabine. [139] Decitabine + Imatinib CML Oki (2007) 28 32% CR, 4% PR, Higher activity in patients without BCR-ABL mutations. [130] Mesylate 7% HI o.1No.11 Vol.31 Decitabine + IL-2 Metastatic Gollob (2006) 21 Reexpreesion of immunomodulatory genes. [133] Melanoma; Renal Carcinoma

541 Abbreviations: CR, complete remission; PR, partial response; HI, hematologic improvement; ORR, overall response rate; SD, stable disease. CCR, conventional care regimens; RBC, red blood cells; HSCT, hematopoietic stem cells transplantation; PBMC, peripheral blood mononuclear cell; PBC, peripheral-blood cells. Review Trends in Pharmacological Sciences Vol.31 No.11 intermediate- and high-risk MDS patients. In that study, trial reported an ORR of 32%, suggesting that this 5-day patients treated with 5-Aza-CR showed a significant im- schedule was as effective as the approved inpatient provement in median overall survival (OS) than patients regimen [116]. 5-Aza-CdR is currently being investigated who received conventional care regimens (CCR): 24.5 in other cancer types to assess the best conditions for months vs 15.0 months. It was the first time that 5-Aza- administration [117–119]. CR treatment was demonstrated to prolong OS in high-risk MDS patients [106]. Detailed analyses of AZA-001 indicat- MG98 Several phase I/II clinical trials have been ed that low-dose cytarabine (one of CCR for patients with conducted to find out the tumor types sensitive to MG98 higher-risk MDS) less efficient and more toxic as compared and an appropriate working dosage. Some of the patients with 5-Aza-CR [107]. Fenaux et al. also demonstrated that already showed decreases in DNMT1 levels, but a 5-Aza-CR benefited older AML patients by prolonging the consistent correlation of DNMT1 level and dosage has OS from 16.0 months to 24.5 months while reducing the not been observed [120–121]. New evidence indicated prevalence of side effects [107]. Based on the outcome of that 7-day continuous dosing of MG98 was well the AZA-001 trial, the National Comprehensive Cancer tolerated for patients with advanced solid tumors, but Network (NCCN) recommended 5-Aza-CR as the preferred an objective clinical response was not reported [122]. therapy for patients with high-risk MDS. In addition to high-risk MDS, 5-Aza-CR can also be used as a potentially Combination therapy effective treatment for patients with low-risk MDS [108– HDAC inhibitors (HDACi) are a family of epigenetic drugs 109]. that increase acetylation of histone proteins and cyto- However, the current FDA-approved, 7-day course of 5- plasmic proteins such as p53. As discussed above, revers- Aza-CR treatment requires weekend treatment and is ing abnormal silencing of epigenetic genes is coordinated inconvenient for patients and care providers. To overcome by increasing promoter histone acetylation levels and DNA this problem, Lyons et al. [110] designed three alternative demethylation. Based on these observations, numerous regimens which avoided weekend treatment. Patients who clinical trials have been conducted to study the synergistic received any one of the three regimens showed similar effects of DNMTi and HDACi (Table 1). hematological improvement as the previously approved 7- Several research teams have combined 5-Aza-CR with day 5-Aza-CR regimen, as well as a higher transfusion- the HDACi valproic acid (VPA) to treat MDS and AML. independent rate. More patients who needed transfusion of Soriano et al. [123] conducted a phase-I/II clinical trial in red blood cells (RBCs) at baseline became independent of which patients were treated with 5-Aza-CR and VPA every RBC transfusion. Another phase-II trial administered an day for 7 days. An ORR of 42% indicated that this combi- alternative 5-day 5-Aza-CR intravenous schedule and nation strategy was clinically effective. Another phase-II reported a 27% partial response (PR) + complete remission study sequentially administered 5-Aza-CR after VPA and (CR) rate that was comparable with the 7-day subcutane- observed a CR+PR of 30.7%. As the plasma concentration ous regimen [111]. However, further studies are required of VPA increased, (>50 mg/mL vs <50 mg/mL), a better to illuminate the survival benefit of these modified regi- median survival rate was observed (18.7 vs 10 months), mens. Although one group reported that a limited number indicating HDACi have the potential to increase 5-Aza-CR of treatment cycles could achieve a overall response rate efficacy [124]. Combination therapy of 5-Aza-CR + VPA has (ORR) of 50% according to new criteria set by the Interna- also been investigated in breast cancer, colon cancer and tional Working Group, most clinical trials indicate that other advanced cancers [125]. Other HDACi have also been prolonged exposure to 5-Aza-CR will benefit patients [112]. combined with 5-Aza-CR. For example, 46% of patients with MDS or AML who received 5-Aza-CR supplemented 5-aza-2’-deoxycitidine (Decitabine) 5-Aza-CdR was also with entinostat (another potential HDACi) showed prom- approved by the FDA for MDS therapy, but there is no clear ising outcomes. Among them, 3 patients had CR, 4 patients evidence indicating 5-Aza-CdR improves OS. In the USA- had PR and 7 patients had hematologic improvement registered trial (D-0007), patients treated with 5-Aza-CdR [126]. In a phase-I study, researchers co-administered 5- had a 17% ORR, which was significantly higher than that Aza-CR and the first-generation HDACi sodium phenylbu- in the best supportive care (BSC) group (0%). When tyrate to patients with refractory solid tumors. The three- comparing the OS between the 5-Aza-CdR and control dose schedule administered in that study showed mild arms, a statistical improvement (14.0 vs 14.9 months) toxicity, albeit with few benefits for the patients [127]. was not observed, even though clinical benefits (e.g. Similar to 5-Aza-CR, 5-Aza-CdR has also been tested for independence from RBC transfusion or elongation of its therapeutic efficacy in combination with HDACi. 5-Aza- median time to AML progression) were seen after 5-Aza- CdR + VPA achieved a 22% objective response in patients CdR treatment [113]. Wijermans et al. presented a similar with leukemia although, in this case, the higher level of negative result of the OS advantage of 5-Aza-CdR for older VPA did not correlate with clinical activity [128]. A later patients with MDS or Chronic Myelomonocytic Leukemia study confirmed this conclusion, demonstrating that 5- (CMML) [114]. To increase the CR rate of patients with Aza-CdR itself had promising clinical activity in elderly these diseases taking 5-Aza-CdR in the outpatient setting, patients with AML. Adding 25 mg/kg/d of VPA neither several clinical trials explored alternative schedules. enhanced the efficacy of 5-Aza-CdR nor did it increase Kantarjian et al [115] showed that a 5-day intravenous the re-expression of CDKN2B [129]. schedule with the highest dose-intensity yielded the 5-Aza-CR and 5-Aza-CdR have also been combined with highest CR rate (39%). In a follow-up study, the ADOPT other conventional therapies. For example, a recent study

542 Review Trends in Pharmacological Sciences Vol.31 No.11 combining 5-Aza-CR with the FDA-approved MDS chemo- respond to hypomethylation therapy best are those who therapy drug lenalidomide revealed an impressive ORR of carry markers such as lower methylation levels in the 67% [130]. 5-Aza-CdR has been combined with carboplatin, CDKN2B promoter or higher levels of miR29 [128,143]. imatinib, or with gemtuzumab ozogamicin (monoclonal Ultimately, efforts to create a comprehensive picture of the antibody) for the treatment of various cancer types [131– epigenetic landscape should help in the understanding of 133]. The encouraging results from these clinical trials cancer pathology and the molecular basis of tumorigenesis. could broaden the application of DNMTi to various tumor This could lay the foundation for future drug development types, but further randomized clinical trials with control to improve the survival and quality of life of cancer arms are needed to demonstrate the feasibility of these patients. approaches. Acknowledgments Validating epigenetic targets in clinical trials This work was supported by a R01 grant from the National Institutes of The ability of 5-Aza-CR and 5-Aza-CdR to reduce global Health (‘Mechanisms of de novo methylation in cancer’ (CA082422-12)) DNA methylation in vivo has been validated in multiple and a grant from the organization ‘Stand Up To Cancer’. clinical trials in which they have been used as single agents or combined with other anticancer therapies [118,123,125– References 1 Jones, P.A. and Baylin, S.B. (2007) The epigenomics of cancer. Cell 126,129,134–140]. The reversal of DNA methylation 128, 683–692 around individual gene promoters such as CDKN2B and 2 Lister, R. et al. (2009) Human DNA methylomes at base resolution MAGE1A has also been confirmed in various tumor types show widespread epigenomic differences. Nature 462, 315–322 [126,128,133]. However, most of these studies did not show 3 Suzuki, M.M. and Bird, A. (2008) DNA methylation landscapes: a clear association between the level of induced demethyl- provocative insights from epigenomics. Nat. Rev. Genet. 9, 465–476 4 Takai, D. and Jones, P.A. (2002) Comprehensive analysis of CpG ation and clinical response. Two recent studies reported islands in human chromosomes 21 and 22. Proc. Natl. Acad. Sci. that responders to 5-Aza-CR have more significant U.S.A. 99, 3740–3745 decreases in global methylation level and methylated pro- 5 Wang, Y. and Leung, F.C. (2004) An evaluation of new criteria for CpG moters compared with non-responders as detected by the islands in the human genome as gene markers. Bioinformatics 20, TM 1170–1177 Infinium methylation assay [111,130]. 6 Jones, P.A. and Liang, G. (2009) Rethinking how DNA methylation patterns are maintained. Nat. Rev. Genet. 10, 805–811 Conclusion and future directions 7 Hattori, N. et al. (2004) Epigenetic control of mouse Oct-4 gene This review summaries the recent advances and future expression in embryonic stem cells and trophoblast stem cells. J. prospects for epigenetic therapy through DNA methyla- Biol. Chem. 279, 17063–17069 8 Irizarry, R.A. et al. (2009) The human colon cancer methylome shows tion. After decades of development, DNMTi are now been similar hypo- and hypermethylation at conserved tissue-specific CpG used in cancer therapies. Progress made in epigenetic island shores. Nat. Genet. 41, 178–186 research will further aid translation from the ‘bench to 9 Okano, M. et al. (1999) DNA methyltransferases Dnmt3a and Dnmt3b the bedside’. are essential for de novo methylation and mammalian development. A densely methylated region is strongly associated with Cell 99, 247–257 10 Bourc’his, D. et al. (2001) Dnmt3L and the establishment of maternal other silencing mechanisms (including chromatin pro- genomic imprints. Science 294, 2536–2539 teins), and consequently the use of a single hypomethylat- 11 Jia, D. et al. (2007) Structure of Dnmt3a bound to Dnmt3L suggests a ing agent might not be sufficient to reverse epigenetic model for de novo DNA methylation. Nature 449, 248–251 aberrations in cancer. This concept has guided the clinical 12 Arita, K. et al. (2008) Recognition of hemi-methylated DNA by the application of DNMTi and HDACi combination therapy. SRA protein UHRF1 by a base-flipping mechanism. Nature 455, 818– 821 Further systematic exploration of the overall epigenetic 13 Bostick, M. et al. (2007) UHRF1 plays a role in maintaining DNA changes after drug-induced DNA demethylation might methylation in mammalian cells. Science 317, 1760–1764 lead to the discovery of other drug targets with the goal 14 Jeong, S. et al. (2009) Selective anchoring of DNA methyltransferases of minimizing toxicity and maximizing response. Such 3A and 3B to nucleosomes containing methylated DNA. Mol. Cell Biol. 29, 5366–5376 examples might be inhibitors to LSD1 (which removes 15 Barski, A. et al. (2007) High-resolution profiling of histone the active H3K4me mark) or inhibitors of EZH2 (which methylations in the human genome. Cell 129, 823–837 plays an important part in gene suppression in cancer) 16 Mikkelsen, T.S. et al. (2007) Genome-wide maps of chromatin state in [141,142]. In addition to targeting DNMT1, new drugs pluripotent and lineage-committed cells. Nature 448, 553–560 could be designed to inhibit DNA methyltransferases 3A 17 Schones, D.E. and Zhao, K. (2008) Genome-wide approaches to studying chromatin modifications. Nat. Rev. Genet. 9, 179–191 and 3B with the accumulated understanding of their func- 18 Guenther, M.G. et al. (2007) A chromatin landmark and transcription tions in maintaining methylation patterns in somatic cells. initiation at most promoters in human cells. Cell 130, 77–88 The elucidation of differences between the responses of 5- 19 Ozsolak, F. et al. (2008) Chromatin structure analyses identify Aza-CR and 5-Aza-CdR might help in understanding the miRNA promoters. Genes Dev. 22, 3172–3183 different outcomes in clinical trials, and prompt trials to 20 Heintzman, N.D. et al. (2007) Distinct and predictive chromatin signatures of transcriptional promoters and enhancers in the compare them head-to-head [80]. human genome. Nat. Genet. 39, 311–318 Analyzing available clinical data might also provide 21 Liang, G. et al. (2004) Distinct localization of histone H3 acetylation better prognostic factors for patient survival or predict and H3-K4 methylation to the transcription start sites in the human markers for patient sensitivity to DNMTi. It has been genome. Proc. Natl. Acad. Sci. U.S.A. 101, 7357–7362 suggested that there is no positive correlation between a 22 Wang, Z. et al. (2008) Combinatorial patterns of histone acetylations and methylations in the human genome. Nat. Genet. 40, 897–903 good response to DNMT treatment and baseline methyla- 23 Schuettengruber, B. et al. (2007) Genome regulation by polycomb and tion in the patient [126]. In fact, patients who seem to trithorax proteins. Cell 128, 735–745

543 Review Trends in Pharmacological Sciences Vol.31 No.11

24 Kouzarides, T. (2007) Chromatin modifications and their function. 53 Kautiainen, T.L. and Jones, P.A. (1986) DNA methyltransferase Cell 128, 693–705 levels in tumorigenic and nontumorigenic cells in culture. J. Biol. 25 Berger, S.L. (2007) The complex language of chromatin regulation Chem. 261, 1594–1598 during transcription. Nature 447, 407–412 54 Cedar, H. and Bergman, Y. (2009) Linking DNA methylation and 26 Li, B. et al. (2007) The role of chromatin during transcription. Cell 128, histone modification: patterns and paradigms. Nat. Rev. Genet. 10, 707–719 295–304 27 Zlatanova, J. and Thakar, A. (2008) H2A.Z: view from the top. 55 Ohm, J.E. et al. (2007) A stem cell-like chromatin pattern may Structure 16, 166–179 predispose tumor suppressor genes to DNA hypermethylation and 28 Jin, C. et al. (2009) H3.3/H2A.Z double variant-containing heritable silencing. Nat. Genet. 39, 237–242 nucleosomes mark ‘nucleosome-free regions’ of active promoters 56 Schlesinger, Y. et al. (2007) Polycomb-mediated methylation on Lys27 and other regulatory regions. Nat. Genet. 41, 941–945 of histone H3 pre-marks genes for de novo methylation in cancer. Nat. 29 Gevry, N. et al. (2009) Histone H2A.Z is essential for estrogen receptor Genet. 39, 232–236 signaling. Genes Dev. 23, 1522–1533 57 Gal-Yam, E.N. et al. (2008) Frequent switching of Polycomb 30 Creyghton, M.P. et al. (2008) H2AZ is enriched at polycomb complex repressive marks and DNA hypermethylation in the PC3 prostate target genes in ES cells and is necessary for lineage commitment. Cell cancer cell line. Proc. Natl. Acad. Sci. U.S.A. 105, 12979–12984 135, 649–661 58 Halkidou, K. et al. (2004) Upregulation and nuclear recruitment of 31 Sutcliffe, E.L. et al. (2009) Dynamic histone variant exchange HDAC1 in hormone refractory prostate cancer. Prostate 59, 177–189 accompanies gene induction in T cells. Mol. Cell Biol. 29, 1972– 59 Song, J. et al. (2005) Increased expression of histone deacetylase 2 is 1986 found in human gastric cancer. APMIS 113, 264–268 32 Marques, M. et al. (2010) Reconciling the positive and negative roles of 60 Schulte, J.H. et al. (2009) Lysine-specific demethylase 1 is strongly histone H2A.Z in gene transcription. Epigenetics 5, 267–272 expressed in poorly differentiated neuroblastoma: implications for 33 Zilberman, D. et al. (2008) Histone H2A.Z and DNA methylation are therapy. Cancer Res. 69, 2065–2071 mutually antagonistic chromatin marks. Nature 456, 125–129 61 Lim, S. et al. (2010) Lysine-specific demethylase 1 (LSD1) is highly 34 Okitsu, C.Y. and Hsieh, C.L. (2007) DNA methylation dictates histone expressed in ER-negative breast cancers and a biomarker predicting H3K4 methylation. Mol. Cell Biol. 27, 2746–2757 aggressive biology. Carcinogenesis 31, 512–520 35 Zemach, A. et al. (2010) Genome-wide evolutionary analysis of 62 Kondo, Y. et al. (2008) Gene silencing in cancer by histone H3 lysine 27 eukaryotic DNA methylation. Science 328, 916–919 trimethylation independent of promoter DNA methylation. Nat. 36 Nan, X. et al. (1998) Transcriptional repression by the methyl-CpG- Genet. 40, 741–750 binding protein MeCP2 involves a histone deacetylase complex. 63 Varambally, S. et al. (2002) The polycomb group protein EZH2 is Nature 393, 386–389 involved in progression of prostate cancer. Nature 419, 624–629 37 Jones, P.L. et al. (1998) Methylated DNA and MeCP2 recruit histone 64 Varambally, S. et al. (2008) Genomic loss of microRNA-101 leads to deacetylase to repress transcription. Nat. Genet. 19, 187–191 overexpression of histone methyltransferase EZH2 in cancer. Science 38 Ooi, S.K. et al. (2007) DNMT3L connects unmethylated lysine 4 of 322, 1695–1699 histone H3 to de novo methylation of DNA. Nature 448, 714–717 65 Friedman, J.M. et al. (2009) The putative tumor suppressor 39 Vire, E. et al. (2006) The Polycomb group protein EZH2 directly microRNA-101 modulates the cancer epigenome by repressing the controls DNA methylation. Nature 439, 871–874 polycomb group protein EZH2. Cancer Res. 69, 2623–2629 40 Tachibana, M. et al. (2008) G9a/GLP complexes independently 66 Lin, J.C. et al. (2007) Role of nucleosomal occupancy in the epigenetic mediate H3K9 and DNA methylation to silence transcription. silencing of the MLH1 CpG island. Cancer Cell 12, 432–444 EMBO J. 27, 2681–2690 67 He, H.H. et al. (2010) Nucleosome dynamics define transcriptional 41 Dong, K.B. et al. (2008) DNA methylation in ES cells requires the enhancers. Nat. Genet. 42, 343–347 lysine methyltransferase G9a but not its catalytic activity. EMBO J. 68 Jones, P.A. and Taylor, S.M. (1980) Cellular differentiation, cytidine 27, 2691–2701 analogs and DNA methylation. Cell 20, 85–93 42 Eden, A. et al. (2003) Chromosomal instability and tumors promoted 69 Taylor, S.M. and Jones, P.A. (1979) Multiple new phenotypes induced by DNA hypomethylation. Science 300, 455 in 10T1/2 and 3T3 cells treated with 5-azacytidine. Cell 17, 771–779 43 Pogribny, I.P. and Beland, F.A. (2009) DNA hypomethylation in the 70 Vesely´, J. and Cˇ iha?k, A. (1978) 5-Azacytidine: mechanism of action and origin and pathogenesis of human diseases. Cell Mol. Life Sci. 66, biological effects in mammalian cells. Pharmac. Ther. A. 2, 813–840 2249–2261 71 Rius, M. et al. (2009) Human concentrative nucleoside transporter 1- 44 Wu, H. et al. (2005) Hypomethylation-linked activation of PAX2 mediated uptake of 5-azacytidine enhances DNA demethylation. Mol. mediates tamoxifen-stimulated endometrial carcinogenesis. Nature Cancer Ther. 8, 225–231 438, 981–987 72 Issa, J.P. and Kantarjian, H.M. (2009) Targeting DNA methylation. 45 Esteller, M. (2007) Cancer epigenomics: DNA methylomes and Clin. Cancer Res. 15, 3938–3946 histone-modification maps. Nat. Rev. Genet. 8, 286–298 73 Stresemann, C. and Lyko, F. (2008) Modes of action of the DNA 46 Lujambio, A. et al. (2008) A microRNA DNA methylation signature for methyltransferase inhibitors azacytidine and decitabine. Int. J. human cancer metastasis. Proc. Natl. Acad. Sci. U.S.A. 105, 13556– Cancer 123, 8–13 13561 74 Ghoshal, K. et al. (2005) 5-Aza-deoxycytidine induces selective 47 Toyota, M. et al. (2008) Epigenetic silencing of microRNA-34b/c and B- degradation of DNA methyltransferase 1 by a proteasomal cell translocation gene 4 is associated with CpG island methylation in pathway that requires the KEN box, bromo-adjacent homology colorectal cancer. Cancer Res. 68, 4123–4132 domain, and nuclear localization signal. Mol. Cell Biol. 25, 4727–4741 48 Saito, Y. et al. (2006) Specific activation of microRNA-127 75 Kuo, H.K. et al. (2007) 5-Azacytidine induced methyltransferase-DNA with downregulation of the proto-oncogene BCL6 by chromatin- adducts block DNA replication in vivo. Cancer Res. 67, 8248–8254 modifying drugs in human cancer cells. Cancer Cell 9, 435– 76 Nguyen, C.T. et al. (2002) Histone H3-lysine 9 methylation is 443 associated with aberrant gene silencing in cancer cells and is 49 Ordway, J.M. et al. (2007) Identification of novel high-frequency DNA rapidly reversed by 5-aza-2’-deoxycytidine. Cancer Res. 62, 6456–6461 methylation changes in breast cancer. PLoS One 2, e1314 77 Komashko, V.M. and Farnham, P.J. (2010) 5-azacytidine treatment 50 Rauch, T.A. et al. (2008) High-resolution mapping of DNA reorganizes genomic histone modification patterns. Epigenetics 5, hypermethylation and hypomethylation in lung cancer. Proc. Natl. 229–240 Acad. Sci. U.S.A. 105, 252–257 78 Qin, T. et al. (2009) Mechanisms of resistance to 5-aza-2’- 51 Figueroa, M.E. et al. (2009) MDS and secondary AML display unique deoxycytidine in human cancer cell lines. Blood 113, 659–667 patterns and abundance of aberrant DNA methylation. Blood 114, 79 Li, L.H. et al. (1970) Cytotoxicity and mode of action of 5-azacytidine 3448–3458 on L1210 leukemia. Cancer Res. 30, 2760–2769 52 Noushmehr, H. et al. (2010) Identification of a CpG island methylator 80 Hollenbach, P.W. et al. (2010) A comparison of azacitidine and phenotype that defines a distinct subgroup of glioma. Cancer Cell 17, decitabine activities in acute myeloid leukemia cell lines. PLoS 510–522 One 5, e9001

544 Review Trends in Pharmacological Sciences Vol.31 No.11

81 Yoo, C.B. et al. (2008) Long-term epigenetic therapy with oral 104 Kornblith, A.B. et al. (2002) Impact of azacytidine on the quality of life zebularine has minimal side effects and prevents intestinal tumors of patients with myelodysplastic syndrome treated in a randomized in mice. Cancer Prev. Res. (Phila PA) 1, 233–240 phase III trial: a Cancer and Leukemia Group B study. J. Clin. Oncol. 82 Lemaire, M. et al. (2009) Inhibition of cytidine deaminase by 20, 2441–2452 zebularine enhances the antineoplastic action of 5-aza-2’- 105 Silverman,L.R. etal. (2006) Furtheranalysisoftrialswith azacitidinein deoxycytidine. Cancer Chemother. Pharmacol. 63, 411–416 patients with myelodysplastic syndrome: studies 8421, 8921, and 9221 83 Yoo, C.B. et al. (2004) Zebularine: a new drug for epigenetic therapy. by the Cancer and Leukemia Group B. J. Clin. Oncol. 24, 3895–3903 Biochem. Soc. Trans. 32, 910–912 106 Fenaux, P. et al. (2009) Efficacy of azacitidine compared with that of 84 Billam, M. et al. (2010) Effects of a novel DNA methyltransferase conventional care regimens in the treatment of higher-risk inhibitor zebularine on human breast cancer cells. Breast Cancer Res. myelodysplastic syndromes: a randomised, open-label, phase III Treat. 120, 581–592 study. Lancet Oncol. 10, 223–232 85 Flotho, C. et al. (2009) The DNA methyltransferase inhibitors 107 Fenaux, P. et al. (2010) Prolonged survival with improved tolerability azacitidine, decitabine and zebularine exert differential effects on in higher-risk myelodysplastic syndromes: azacitidine compared with cancer gene expression in acute myeloid leukemia cells. Leukemia 23, low dose ara-C. Br J Haematol 149, 244–249 1019–1028 108 Santini, V. (2009) Azacitidine in lower-risk myelodysplastic 86 Beumer, J.H. et al. (2006) Pharmacokinetics, metabolism, and oral syndromes. Leuk. Res. 33 (suppl 2), S22–S26 bioavailability of the DNA methyltransferase inhibitor 5-fluoro-2’- 109 Musto, P. et al. (2010) Azacitidine for the treatment of lower risk deoxycytidine in mice. Clin. Cancer Res. 12, 7483–7491 myelodysplastic syndromes: a retrospective study of 74 patients 87 Beumer, J.H. et al. (2008) Concentrations of the DNA enrolled in an Italian named patient program. Cancer 116, 1485–1494 methyltransferase inhibitor 5-fluoro-2’-deoxycytidine (FdCyd) and 110 Lyons, R.M. et al. (2009) Hematologic response to three alternative its cytotoxic metabolites in plasma of patients treated with FdCyd dosing schedules of azacitidine in patients with myelodysplastic and tetrahydrouridine (THU). Cancer Chemother. Pharmacol. 62, syndromes. J. Clin. Oncol. 27, 1850–1856 363–368 111 Martin, M.G. et al. (2009) A phase II study of 5-day intravenous 88 Kratzke, R.A. et al. (2008) Response to the methylation inhibitor azacitidine in patients with myelodysplastic syndromes. Am. J. dihydro-5-azacytidine in mesothelioma is not associated with Hematol. 84, 560–564 methylation of p16INK4a: results of cancer and leukemia group B 112 Muller-Thomas, C. et al. (2009) A limited number of 5-azacitidine 159904. J. Thorac. Oncol. 3, 417–421 cycles can be effective treatment in MDS. Ann. Hematol. 88, 213–219 89 Byun, H.M. et al. (2008) 2’-Deoxy-N4-[2-(4-nitrophenyl)ethoxycarbonyl]- 113 Kantarjian, H. et al. (2006) Decitabine improves patient outcomes in 5-azacytidine: a novel inhibitor of DNA methyltransferase that myelodysplastic syndromes: results of a phase III randomized study. requires activation by human carboxylesterase 1. Cancer Lett. 266, Cancer 106, 1794–1803 238–248 114 WijerMans P, et al. (2008) Low dose decitabine versus best supportive 90 Chuang, J.C. et al. (2010) S110, a 5-Aza-20-deoxycytidine-containing care in elderly patients with intermediate or high risk MDS not dinucleotide, is an effective DNA methylation inhibitor in vivo and eligible for intensive chemotherapy: final results of the randomized can reduce tumor growth. Mol Cancer Ther 9, 1443–1450 phase III study (06011) of the EORTC Leukemia and German MDS 91 Segura-Pacheco, B. et al. (2006) Global DNA hypermethylation- Study Groups. Blood, abstract 226 associated cancer chemotherapy resistance and its reversion with 115 Kantarjian, H. et al. (2007) Results of a randomized study of 3 the demethylating agent hydralazine. J. Transl. Med. 4, 32 schedules of low-dose decitabine in higher-risk myelodysplastic 92 Singh, N. et al. (2009) Molecular modeling and molecular dynamics syndrome and chronic myelomonocytic leukemia. Blood 109, 52–57 studies of hydralazine with human DNA methyltransferase 1. Chem. 116 Steensma, D.P. et al. (2009) Multicenter study of decitabine Med. Chem. 4, 792–799 administered daily for 5 days every 4 weeks to adults with 93 Song, Y. and Zhang, C. (2009) Hydralazine inhibits human cervical myelodysplastic syndromes: the alternative dosing for outpatient cancer cell growth in vitro in association with APC demethylation and treatment (ADOPT) trial. J. Clin. Oncol. 27, 3842–3848 re-expression. Cancer Chemother. Pharmacol. 63, 605–613 117 Steensma, D.P. (2009) Decitabine treatment of patients with higher- 94 Chuang, J.C. et al. (2005) Comparison of biological effects of non- risk myelodysplastic syndromes. Leuk. Res. 33 (suppl 2), S12–S17 nucleoside DNA methylation inhibitors versus 5-aza-2’- 118 Stewart, D.J. et al. (2009) Decitabine effect on tumor global DNA deoxycytidine. Mol. Cancer Ther. 4, 1515–1520 methylation and other parameters in a phase I trial in refractory solid 95 Suzuki, T. et al. (2010) Design, synthesis, inhibitory activity, and tumors and lymphomas. Clin. Cancer Res. 15, 3881–3888 binding mode study of novel DNA methyltransferase 1 inhibitors. 119 Schrump, D.S. et al. (2006) Phase I study of decitabine-mediated gene Bioorg. Med. Chem. Lett. 20, 1124–1127 expression in patients with cancers involving the lungs, esophagus, or 96 Stresemann, C. et al. (2006) Functional diversity of DNA pleura. Clin. Cancer Res. 12, 5777–5785 methyltransferase inhibitors in human cancer cell lines. Cancer 120 Klisovic, R.B. et al. (2008) A phase I biological study of MG98, an Res. 66, 2794–2800 oligodeoxynucleotide antisense to DNA methyltransferase 1, in 97 Datta, J. et al. (2009) A new class of quinoline-based DNA patients with high-risk myelodysplasia and acute myeloid hypomethylating agents reactivates tumor suppressor genes by leukemia. Clin. Cancer Res. 14, 2444–2449 blocking DNA methyltransferase 1 activity and inducing its 121 Winquist, E. et al. (2006) Phase II trial of DNA methyltransferase 1 degradation. Cancer Res. 69, 4277–4285 inhibition with the antisense oligonucleotide MG98 in patients with 98 Amato, R.J. (2007) Inhibition of DNA methylation by antisense metastatic renal carcinoma: a National Cancer Institute of Canada oligonucleotide MG98 as cancer therapy. Clin. Genitourin. Cancer Clinical Trials Group investigational new drug study. Invest. New 5, 422–426 Drugs 24, 159–167 99 Garzon, R. et al. (2009) MicroRNA-29b induces global DNA 122 Plummer, R. et al. (2009) Phase I study of MG98, an oligonucleotide hypomethylation and tumor suppressor gene reexpression in acute antisense inhibitor of human DNA methyltransferase 1, given as a 7- myeloid leukemia by targeting directly DNMT3A and 3B and day infusion in patients with advanced solid tumors. Clin. Cancer Res. indirectly DNMT1. Blood 113, 6411–6418 15, 3177–3183 100 Velez-Garcia, E. et al. (1977) Twice weekly 5-azacytidine infusion in 123 Soriano, A.O. et al. (2007) Safety and clinical activity of the dissmeinated metastatic cancer: a phase II study. Cancer Treat. Rep. combination of 5-azacytidine, valproic acid, and all-trans retinoic 61, 1675–1677 acid in acute myeloid leukemia and myelodysplastic syndrome. 101 Srinivasan, U. et al. (1982) Phase II study of 5-azacytidine in Blood 110, 2302–2308 sarcomas of bone. Am. J. Clin. Oncol. 5, 411–415 124 Voso, M.T. et al. (2009) Valproic acid at therapeutic plasma levels may 102 Yoo, C.B. and Jones, P.A. (2006) Epigenetic therapy of cancer: past, increase 5-azacytidine efficacy in higher risk myelodysplastic present and future. Nat. Rev. Drug Discov. 5, 37–50 syndromes. Clin. Cancer Res. 15, 5002–5007 103 Silverman, L.R. et al. (2002) Randomized controlled trial of 125 Braiteh, F. et al. (2008) Phase I study of epigenetic modulation with 5- azacitidine in patients with the myelodysplastic syndrome: a study azacytidine and valproic acid in patients with advanced cancers. Clin. of the cancer and leukemia group B. J. Clin. Oncol. 20, 2429–2440 Cancer Res. 14, 6296–6301

545 Review Trends in Pharmacological Sciences Vol.31 No.11

126 Fandy, T.E. et al. (2009) Early epigenetic changes and DNA damage patients with melanoma or renal cell carcinoma. Clin. Cancer Res. do not predict clinical response in an overlapping schedule of 5- 12, 4619–4627 azacytidine and entinostat in patients with myeloid malignancies. 135 Yang, A.S. et al. (2006) DNA methylation changes after 5-aza-2’- Blood 114, 2764–2773 deoxycytidine therapy in patients with leukemia. Cancer Res. 66, 127 Lin, J. et al. (2009) A phase I dose-finding study of 5-azacytidine in 5495–5503 combination with sodium phenylbutyrate in patients with refractory 136 Raza, A. et al. (2008) Combination of 5-azacytidine and thalidomide solid tumors. Clin. Cancer Res. 15, 6241–6249 for the treatment of myelodysplastic syndromes and acute myeloid 128 Garcia-Manero, G. et al. (2006) Phase 1/2 study of the combination of leukemia. Cancer 113, 1596–1604 5-aza-2’-deoxycytidine with valproic acid in patients with leukemia. 137 Jabbour, E. et al. (2009) Low-dose azacitidine after allogeneic stem Blood 108, 3271–3279 cell transplantation for acute leukemia. Cancer 115, 1899–1905 129 Blum, W. et al. (2007) Phase I study of decitabine alone or in 138 Borthakur, G. et al. (2010) Report of a phase 1/2 study of a combination with valproic acid in acute myeloid leukemia. J. Clin. combination of azacitidine and cytarabine in acute myelogenous Oncol. 25, 3884–3891 leukemia and high-risk myelodysplastic syndromes. Leuk. 130 Sekeres, M.A. et al. (2010) Phase I combination trial of lenalidomide Lymphoma 51, 73–78 and azacitidine in patients with higher-risk myelodysplastic 139 Scott, B.L. et al. (2010) Prolonged responses in patients with MDS and syndromes. J. Clin. Oncol. 28, 2253–2258 CMML treated with azacitidine and etanercept. Brit. J. Haematol. 131 Oki, Y. et al. (2007) Phase II study of low-dose decitabine in 148, 944–947 combination with imatinib mesylate in patients with accelerated or 140 De Padua Silva, L. et al. (2009) Feasibility of allo-SCT after myeloid blastic phase of chronic myelogenous leukemia. Cancer 109, hypomethylating therapy with decitabine for myelodysplastic 899–906 syndrome. Bone Marrow Transplant. 43, 839–843 132 Chowdhury, S. et al. (2009) Decitabine combined with fractionated 141 Huang, Y. et al. (2009) Polyamine analogues targeting epigenetic gene gemtuzumab ozogamicin therapy in patients with relapsed or regulation. Essays Biochem. 46, 95–110 refractory acute myeloid leukemia. Am. J. Hematol. 84, 599–600 142 Tan, J. et al. (2007) Pharmacologic disruption of Polycomb-repressive 133 Appleton, K. et al. (2007) Phase I and pharmacodynamic trial of the complex 2-mediated gene repression selectively induces apoptosis in DNA methyltransferase inhibitor decitabine and carboplatin in solid cancer cells. Genes Dev. 21, 1050–1063 tumors. J. Clin. Oncol. 25, 4603–4609 143 Blum, W. et al. (2010) Clinical response and miR-29b predictive 134 Gollob, J.A. et al. (2006) Phase I trial of sequential low-dose 5-aza-2’- significance in older AML patients treated with a 10-day schedule deoxycytidine plus high-dose intravenous bolus interleukin-2 in of decitabine. Proc. Natl. Acad. Sci. U.S.A. 107, 7473–7478

546