Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2009 Sep 11.
Published in final edited form as: Crit Rev Eukaryot Gene Expr. 2008;18(3):189–206. doi: 10.1615/critreveukargeneexpr.v18.i3.10

The Epigenetics of Adult (Somatic) Stem Cells

Kenneth J Eilertsen 1,2,3,*, Z Elizabeth Floyd 2,4, Jeffrey M Gimble 2,3
PMCID: PMC2741686  NIHMSID: NIHMS138618  PMID: 18540823

Abstract

While genetic studies have provided a wealth of information about health and disease, there is a growing awareness that individual characteristics are also determined by factors other than genetic sequences. These “epigenetic” changes broadly encompass the influence of the environment on gene regulation and expression and in a more narrow sense, describe the mechanisms controlling DNA methylation, histone modification and genetic imprinting. In this review, we focus on the epigenetic mechanisms that regulate adult (somatic) stem cell differentiation, beginning with the metabolic pathways and factors regulating chromatin structure and DNA methylation and the molecular biological tools that are currently available to study these processes. The role of these epigenetic mechanisms in manipulating adult stem cells is followed by a discussion of the challenges and opportunities facing this emerging field.

I. INTRODUCTION: SCOPE AND INTENT OF REVIEW

Historically, discovery activities in the biological and medical sciences have placed equal emphasis on the “nature” and “nurture” components of individual phenomenon. However, with the advent of molecular biological methods, the majority of investigations have focused almost exclusively on the genetic basis of disease and health. Although these studies provided a wealth of information, the pendulum of science has begun to swing back and there is a resurging interest in the contribution of the nurturing environment to gene regulation and expression. Recently, the term “epigenetic” has been used to broadly encompass this growing field of investigation. In a stricter sense, the scope of epigenetics is restricted to mechanisms involving DNA methylation, histone acetylation, and genetic imprinting. We have set out to review recent studies emphasizing the role of epigenetic mechanisms in regulating the development and commitment of eukaryotic stem cells. We are by no means the first (or the last) to do so. Since 2000, the number of reviews published annually on “epigenetic” and “stem cell” has increased steadily from 1 (2000) to 46 (2006). We acknowledge these excellent summaries of the field; however, we have noted that the majority focus their attention on the contribution of epigenetic mechanisms to embryonic and cancer stem cells. In the current review, we will devote our attention to the epigenetics of somatic (adult) stem cells, which has received less attention. Emphasis is placed on the recent primary literature with speculation regarding future directions and opportunities facing this emerging field.

II. METABOLIC PATHWAYS AND MECHANISMS OF REGULATION

Cellular memory is regulated by epigenetic mechanisms that ensure heritable characteristics of cells, and functional differences between cell types, without changing DNA sequence. The epigenetic mechanisms alter chromatin (DNA and associated proteins) in ways that change the availability of genes to transcription factors required for their expression. These alterations are attributed to DNA methyltransferases, which add a methyl group to cytosine in the dinucleotide CpG and Polycomb group (PcG) and associated proteins, which modify histones.

The basic building block of chromatin is the nucleosome, which is based on an octamer of histone proteins. The octamer is formed by an H3-H4 tetramer surrounded by an H2A-H2B dimer. Histones have dual functions in the nucleus as important structural components and as regulators of gene expression. Histone tails protrude out of the nucleosome and undergo several posttranslational modifications including acetylation, methylation, phosphorylation, ubiquitylation, sumoylation, ADP ribosylation, glycosylation, biotinylation, and carbonylation.1 These modifications can exist in multiple combinations and together comprise what is being referred to as the “histone code.”2 Histone modifications can distinguish euchromatin from heterochromatin and influence associations of proteins and protein complexes that regulate gene transcription or repression by altering the availability of genes to transcription factors. Several families of proteins have been identified that modify histones.

Histone acetylation is typically associated with a transcriptionally permissive state and occurs on lysines −5, −9, and −13 of histone 2A, and lysines −5, −12, −15, and −20 of Histone 2B. Histone H3 is acetylated on lysines −9, −14, −18, −23, and −24, and histone H4 is acetylated on lysines −5, −8, −12, and −127. Histone acetyl transferases (HATs) add acetyl groups to histone tails and, to date, three superfamilies have been identified, namely, GNAT (Gcn5-related N-acetyltransferase), p300/CBP and MYST (MOZ, Ybf2/Sas3, and Sas2 and TIP60).3

Histone deacetylation is catalyzed by histone deacetylases (HDACs) and is associated with a closed chromatin structure and gene repression. Acetyl groups are removed by three classes of HDACs, namely, class I (HDAC1, HDAC2, HDAC3 and HDAC8), class II (HDAC4, HDAC5, HDAC6, HDAC7, HDAC9 and HDAC10), and class III consisting of the human sirtuin enzymes, of which there are seven (SIRT1–7). Class I HDACs have been reported to localize to the nucleus, whereas class II HDACs move between the nucleus and cytoplasm. SIRT proteins localize to various organelles depending on their function.4

Histone methyltransferases (HMTs) methylate histones at lysine residues and are associated with both gene activation and repression, depending on which lysine is methylated and whether it is mono-, di-, or trimethylated. These enzymes are characterized by a conserved SET [su(var)39, enhancer of zeste, trithorax] domain. Four subgroups of SET domain proteins have been identified based on homology, namely, SUV1, SUV2, SUV3, and RIZ.5

The complexity of the histone code is further illustrated by the effects of sumoylation and ubiquitylation on histone methylation and acetylation. Sumoylation of histones acts as a negative regulator of histone acetylation and ubiquitylation, resulting in repression of transcription.6 There is evidence that sumoylation also acts as a bridge between histone modification and DNA methylation. The methyl-CpG-binding protein MBD1 recognizes sites of DNA methylation and mediates chromatin structure changes leading to gene silencing.7,8 Sumoylation of MBD1 modulates methylation of histone H3,9 providing a direct link between sumoylation, histone methylation, and DNA methylation. Histones H2A and H2B were among the first proteins found to be reversibly modified by ubiquitin.10,11 Monoubiquitylation of H2B is a prerequisite for methylation of histone H312 and is found at transcriptionally active chromatin.13 However, both the addition and removal of ubiquitin on H2B regulates transcriptional activation,1416 suggesting that the dynamics of H2B ubiquitylation is the central feature of this modification. Ubiquitin modification of H2A leads to transcriptional repression and the E3 ligase responsible for the addition of ubiquitin to histone H2A contains proteins identified as Polycomb group proteins, which mediate gene silencing.17

Polycomb Group (PcG) proteins are a unique set of developmental regulators originally characterized in Drosophila as repressors of homeotic genes.1823 Homeotic genes, which regulate segment identity in the developing embryo, are initially repressed by transcriptional repressors that bind DNA. PcG proteins are subsequently recruited to the site and form Polycomb repressive complexes (PRCs), which further modify chromatin to maintain gene silencing.22,2426 For example, PRC2, a polycomb repressive complex required for embryonic development in mammals, catalyzes the methylation of histone H3 lysine-27 (H3K27), an activity required for PRC2-mediated gene silencing.2731 EZH2, a component of PRC2, is a histone methyltransferase that carries out the methylation of lysine 27 of histone H3.32,33 Methylation of H3K27 recruits additional repressive complexes such as PRC1, which adds an additional level of repression that is propagated.27,28,34,35

The DNA molecule can also be modified at the 5’ position of cytosine rings present in CG dinucleotide sequences by addition of a methyl group.36 Patterns of cytosine methylation are distinct for each cell type and confer cell type identity.37 DNA methylation patterns are known to be established during development and subsequently maintained by the maintenance DNA methyltransferases. For a long time it was held that the established methylation pattern was reliably maintained for the life of the organism and was also irreversible. Feng, Y.-Q., Desprat, R., Fu, H., Olivier, E., Lin, C.M., Lobell, A., Gowda, S.N., Aladjem, M.I., Bouhasira, E.E., 2006. DNA methylation supports intrinsic epigenetic memory in mammalian cells. PLOS Genetics 2, 0461-0470.[2006] have suggested that methylation supports intrinsic memory in mammalian cells and is the primary mark contributing to long-term gene silencing and repression of viruses, transposons, and transgenes. Recent data, however, has suggested that DNA methylation is a reversible signal and can change in response to environmental and other signals.38

DNA methylation patterns are closely linked to chromatin structure. Unmethylated DNA is typically associated with an active chromatin configuration, whereas methylated DNA is associated with inactive chromatin. The traditional view, or model, has maintained that cytosine methylation precedes chromatin structure and that DNA methylation attracts methylated cytosine binding proteins, which in turn recruit repressor complexes.39,40 The repressor complexes contain histone deacetylases, which further contribute to a repressive chromatin state. From this perspective, cytosine methylation is the primary epigenetic marker responsible for repressive chromatin structure. Recently, an alternative to this model has been described in which the state of or lysine 26 of histone H1, functions as part of Polycomb repressive complexes 2 and 3. The histone modifications catalyzed by EZH2 recruit comchromatin determines, or influences, DNA methylation or demethylation.37 Both scenarios may have implications for how DNA methylation patterns can be altered and how to approach demethylation in vitro to restore potential and improve reprogramming to a totipotent state after SCNT.

DNA methyltransferases (DNMTs) catalyze de novo and maintenance DNA methylation. These enzymes catalyze the transfer of a methyl group from the methyl donor S-adenosyl-methionine (SAM) onto the 5’ position of the cytosine ring found in CpG dinucleotides. Not all CpGs are methylated and patterns are tissue and time specific. Five enzymes have been identified, namely, DNMT1, DNMT2, DNMT3a, DNMT3b, and DNMT3L. DNMTs 2 and 3L lack enzymatic function due to loss of an amino-terminal regulatory domain in DNMT2 and the catalytic domain in DNMT3L.41 DNMT1 is regarded as a maintenance methyltransferase and recognizes heminiethylated DNA. DNMT1 localizes to the replication fork42 where it associates with PCNA (the replication processivity protein),43 and association with the hemimethylated sequence has been localized to the N-terminal domain.44 DNMTs 3a and 3b are classified as de novo methyltransferases; they bind to both hemimethylated and unmethylated CpG sites and add methyl groups to previously unmethylated cytosines. De novo methylation occurs extensively during early development and, once the methylation patterns are established, appear to be faithfully maintained for the life of the dividing cell. In contrast, if the cell were to exit the cell cycle, the methylation pattern would be maintained since demethylation is thermodynamically unfavorable.36 Recent data, however, have suggested that a clear distinction between maintenance and de novo methylation does not exist and that all three methyltransferases exist in a multiprotein complex45 and work together to maintain methylation patterns.46 DNTM3L has been reported to participate in de novo methylation of retrotransposons47 and establishing maternal imprints.48

Adding complexity to this system, Vire’ et al.49 have demonstrated that EZH2, a histone methyltransferase that methylates lysine 27 of histone H3 ponents of more elaborate repressive complexes resulting in gene silencing. Interestingly, Vire et al.49 have further shown that EZH2 interacts with DNMTs1, 3a, and 3b in vitro and in vivo, and that appropriate repression requires both EZH2 and DNMTs. Furthermore, appropriate recruitment of DNMTs absolutely requires EZH2. Moreover, reducing activity of DNMTs results in derepression of EZH2 target genes, and if EZH2 is down regulated, demethylation and derepression results. This suggests DNA methylation is dynamic and for some genes, silencing is driven by DNA binding and associated proteins.

II. TOOLS FOR STUDY

Sophisticated molecular biological methods account for many of the advances relating to epigenetics (Table 1). Until recently, the detection of methylated cytosine nucleotides relied on the use of the methylation inhibitory compound 5 aza-cytadine and methylation-sensitive restriction endonucleases. To perform these latter assays, genomic DNA was digested in parallel with enzymes that could cut a specific 4–8 bp fragment either independent or dependent on the phosphorylation status of a single CpG. Subsequent Southern blot analysis examined the digestion pattern within a single gene of interest. Limitations facing this method included its inability to evaluate all possible CpG sequences and its low throughput. Direct bisulfite sequencing is capable of examining the methylation status of all CpG elements within any gene; however, it is not an efficient high-throughput method. Genome microarrays have been developed that allow the evaluation of CpG methylation status across thousands of regulatory elements in a high-throughput manner. These tools are more widely available through commercial vendors with proven quality assurance and control.

TABLE 1.

Epigenetic Methods

DNA methylation—bisulfite sequencing, restriction enzyme isoforms, genome microarrays
Histone acetylation/methylation/phosphorylation/ubiquitylation/sumoylation immunoblot, immunoprecipitation, pharmacologic agents, targeted gene expression
Chromatin conformation—chromatin immunoprecipitation (ChIP) assay, DNase hypersensitivity assay, in vivo footprinting
Circular chromosome conformation capture (4-C)
Immunohistochemical—FISH
Global expression—transcriptomics, proteomics, metabolomics

Histone protein modification (acetylation, methylation, phosphorylation, ubiquitylation, and sumoylation) has become a science in and of itself. There is a growing interest in the development of pharmacologic agents modulating histone acetylase and deacetylase activities. In addition to well-studied histone deacetylase inhibitors such as the trichostatin A and valproic acid, studies are now able to use new compounds such as MS-275,50 pivaloyloxymethyl butyrate,51,52 and resveretrol, a polyphenolic compound found at high concentrations in functional foods. Antibody reagents are now commercially available to epitopes associated with modified histone residues. The tools have been used for immunoblots and immunoprecipitations of total cell protein extracts. Such studies can identify histone associated proteins that change as a function of epigenetic inputs.

Antibodies to modified histone and other proteins have been used in chromatin immunoprecipitation (ChIP) assays to monitor the epigenetic modulation of gene regulatory regions. The traditional ChIP assay relies on PCR methods to detect the cross-linking of a protein of interest to a specific genomic DNA regulatory element following DNA fragmentation and antibody immunoprecipitation. Alternatively, this ChIP assay can be used in tandem with a genomic microarray (ChIP-on-chip assay) to evaluate targets in a high-throughput manner. Recently, the ChIP-on-chip method has been used productively to evaluate the repressor function of Polycomb proteins on developmental genes in embryonic stem cell. The circular-chromosome-conformation-capture (4-C) method provides a novel variation on the technology underlying the ChIP assay. Chromatin proteins are cross-linked to DNA and digested to completion with a restriction endonuclease. The isolated DNA fragments are incubated with ligase for an extended period (one week) to promote circularization of protein–cross-linked DNA fragments, followed by amplification with primers for a gene of interest. The resulting PCR fragments of varying size are then subcloned into a library, sequenced, and the novel genes identified. This approach has been used successfully to document the interchromosomal and intrachromosomal interaction of the maternally inherited H19 imprinting control region.

Additional molecular tools can be applied. Fluorescent in situ hybridization (FISH) analysis has been used to colocalize the expression of modified histone proteins to chromosomal translocation sites associated with imprinted genes or sites of X inactivation.53 Multiple studies have applied global transcriptomic methods to epigenetic questions using stem cell cultures and whole animal models. There are likely to be similar studies appearing using both proteomic and metabolomic methods.

III. EPIGENETIC MECHANISMS AND MANIPULATION OF ADULT (SOMATIC) STEM

Although the role of epigenetic mechanisms regulating adult stem cell differentiation has garnered greater attention recently, the literature spans a number of decades. Studies exploring the role of DNA methylation in fibroblast cell lines (3T3, CH310T1/2)54 paved the way for the discovery of the “master regulatory” factor MyoD55,56 that revolutionized the transcriptional field. To systematically address recent developments, this section will explore findings related to specific adult stem cell lineages.

A. Hematopoietic Stem Cells (HSCs)

The HSCs are the best characterized adult stem cells with respect to their epigenetic regulation, consistent with the fact that HSCs have been used clinically for over four decades. Interest in epigenetic modification of HSC differentiation can be traced to studies documenting the ability of 5’ aza-cytadine57 and butyrate to promote demethylation of fetal globin promoters. Consequently, these agents have been used to treat patients suffering from hemoglobinopathies such as beta thalassemia and sickle cell disease.57,58 Ex vivo, the histone deacetylase inhibitors (trichostatin A, sodium buyrate, valproic acid) have been found to reactivate fetal hemoglobin expression in HSC cultures5961 The phenomenon can be potentiated by the addition of c-kit ligand in the presence of sodium butyrate.59 The addition of granulocyte-macrophage colony stimulating factor (GM-CSF) alone likewise induced fetal hemoglobin switching, although to a lesser degree,59 consistent with the observation that GM-CSF induces methylation of the p15 gene in an HDAC- and DNMT-dependent manner.62

In vitro, HSCs require both hematopoietic cytokines and adhesion molecules, provided in vivo by bone marrow–derived mesenchymal stem cells (MSCs), to proliferate and differentiate. The addition of DNA methylation modifiers (5 aza-cytadine) or HDAC inhibitors (trichostatin A) directly to HSCs promoted their expansion in culture6366 however, this has been associated with apoptosis.67 In contrast, when these agents were added to cocultures of HSCs and MSCs together, the CD34+ HSC population was enhanced, suggesting that MSCs may provide survival factors that balance the apoptotic effects of the epigenetic agents.67 A number of epigenetic genes have been implicated in the process of HSC renewal and differentiation. These include Hmgb3,68 polycomb group genes,69,70 and Xist.71

B. Mesodermal Lineages

Taylor and Jones54 first demonstrated that murine embryonic fibroblasts treated with 5 aza-cytadine were induced to undergo differentiation along the myocyte, adipocyte, and osteoblast pathways. Induction by 5 aza-cytadine has been used to promote the differentiation of ASCs and MSCs along the myocyte and cardiomyocyte pathways.7276 Investigators have continued to pursue the role of CpG methylation in the context of adipogenesis using primary adipose-derived stem cells (ASCs) as well as MSCs.7779 Analysis of the CpG methylation pattern of promoter elements in ASCs finds that while adipocyte-associated genes (lipoprotein lipase, leptin, peroxisome proliferator activated receptor gamma) were hypomethylated,79 those of endothelial-associated genes (CD31, CD 144) remained stably methylated,77 suggesting that ASC differentiation along the endothelial pathway is limited. Comparisons between ASC and MSC cell lines have indicated that CpG methylation regulates adipogenic events.78

The HDAC inhibitors have been found to promote myocyte growth in vitro and in vivo. When trichostatin A was given to wild-type mice, their myocytes’ size increased in a follistatin-dependent manner.80 Similar findings have been reported in dystrophic mice, where HDAC inhibitors enhanced myofiber size and muscle function.81 This suggests a potential therapeutic role for HDAC inhibitors targeting adult stem cells or lineage-specific progenitors. In contrast, the HDAC inhibitors valproic acid and trichostatin A blocked adipogenesis in murine cell lines (3T3-L1) and primary human preadipocytes in a CAAT/enhancer-binding protein alpha–dependent manner.82 Whether such epigenetic-related actions are associated with chromatin organization, as suggested by studies on adipocyterelated liposarcomas tumors, remains to be determined.83

In general, adipogenesis and osteogenesis display an inverse relationship;84 agents that promote differentiation along one pathway often do so at the expense of the other.85 Consistent with this is the observation that the HDAC inhibitors valproic acid and trichostatin A promote osteogenesis by human MSCs in a bone morphogenetic protein–dependent manner in vitro.86,87 These findings suggest that HDAC inhibitors could enhance regeneration in bone, but further in vitro and in vivo studies will be necessary.86 Although the mechanisms accounting for osteogenic promotion are unknown, several possibilities merit further consideration. One is the role of the polycomb genes, since their epigenetic regulation has been associated with skeleton formation in vivo.88 A second is the role of histone arginine methylation, which has recently been identified as an epigenetic mechanism at work in murine embryos.89 A third is the role of posttranscriptional regulators, such as microRNAs and their interactive RNA-binding proteins.90,91 Animals deficient in the expression of the Sam68 RNA-binding protein display a phenotype characterized by decreased bone marrow adipogenesis and protection against bone loss associated with aging.91 In vitro, fibroblasts with reduced expression of Sma68 increased their expression of the osteoblast-associated osteocalcin gene.91 These latter findings are consistent with the inverse relationship hypothesis regulating adipogenesis and osteogenesis in the bone marrow microenvironment and all merit further investigation with respect to the epigenetic regulation of adult stem cell fate.

C. Endodermal and Ectodermal Lineages

Bone marrow–derived MSCs have been reported to express markers consistent with beta-islet cell differentiation (insulin, glucagon, PDX-1) following in vitro exposure to the HDAC inhibitor trichostatin A in the presence of glucagon.92 Likewise, the presence of trichostatin A in combination with hepatocyte growth factor, fibroblast growth factor-4, and other agents, was found to induce hepatocyte biochemical markers in human MSCs.93 Although these studies suggest that epigenetic mechanisms can be used to maximize the pluripotentiality of bone marrow–derived adult stem cells, such findings need to be reproduced and further validated using in vivo models. Others have used trichostatin A to manipulate the differentiation of liver-derived stellate cells, preventing myofibrotic biomarker expression.94,95 Again, although this raises the possibility of exploiting epigenetic regulators for the treatment of fibrotic disease in the liver, in vivo studies are needed.

In the central nervous system, the histone deacetylase inhibitor valproic acid has been found to induce differentiation neural stem cells derived from the hippocampus.96 Neuronal differentiation is accompanied by expression of the transcription factor Neuro D and inhibition of both the astrocyte and oligodendrocyte pathways.96 Similar findings have been made with other HDAC inhibitors in vitro and in vivo, suggesting that these agents may have potential utility for the treatment of spinal muscular atrophy.97,98 In studies of both MSCs and ASCs in vitro, this same agent has been used in combination with other factors to induce neuronal biomarker expression.99,100 It is hypothesized that micro-RNAs, which modulate posttranscriptional events, may be downstream effectors of both HATs and HDACs.90 The tissue-specific expression of these short RNAs may coordinate lineage commitment by adult stem cells in the brain and elsewhere in an epigenetic manner.90

IV. LESSONS FROM ASSISTED REPRODUCTION: IMPACT OF CULTURE ON THE EPIGENOME OF STEM CELL-BASED THERAPEUTICS

Similar to stem cells planned for therapeutic purposes, the methods of assisted reproduction technology (ART) require culture. Embryo culture is an environment that can be regarded as less than optimal. As a result, perturbations can be introduced that may be detrimental to differentiation and development. In light of growing concerns about epigenetic disturbances associated with ART, there is clearly a need to understand adult stem cell epigenetic events and the impact of culture on these processes. What follows is a brief review of observed consequences associated with embryo culture.

A. Methylation Reprogramming Defects

In mouse, the maternal and paternal genomes are demethylated after fertilization by different mechanisms. Drastic differences in methylation reprogramming and development have been observed in embryos fertilized in vitro and cultured. These differences are likely due to an environment (e.g., culture) that is suboptimal that impacts epigenetic reprogramming. It has been demonstrated in mice and ruminants that isolation, manipulation, and culture of gametes and early embryos can disrupt methylation, resulting in phenotypic defects.101,102 Disruptions in establishing and maintaining appropriate methylation patterns may also contribute to issues associated with assisted reproductive technology (ART). Children conceived with ART have been reported to have unusually high incidences of rare imprinting-related diseases, including Beckwith-Wiedemann and Angelman syndromes. These diseases are associated with abnormal hypomethylation of typically methylated maternal alleles.103,104

Similarly, somatic cell nuclear transfer (SCNT, or cloning) disrupts epigenetic reprogramming events and the consequences are more frequent and amplified. Methylcytosine labeling of cloned bovine embryos has demonstrated incomplete or delayed demethylation of the donor genome,48,105 likely resulting in a mixture of normally and abnormally methylated sequences. Several reports have described consequences of disrupted epigenetic reprogramming associated with SCNT including aberrant gene expression in preimplantation embryos,106 placentas, and livers of neonatal cloned mice,107 improper reactivation of genes required for cell lineage differentiation,108 and aberrant expression of imprinted genes in cloned mouse blastocysts.109 Thus, the developmental failures and longer-term defects associated with cloning are associated with improper epigenetic reprogramming. The contribution of culture conditions to aberrant reprogramming in clones has not been fully assessed.

B. Embryo Culture Results in Shifts in Energy Metabolism

In vivo, the early cleavage-stage, relatively metabolically quiescent embryo uses pyruvate/lactate, but not glucose, as a sole energy source. In contrast, blastocysts are highly metabolically active and use glucose as a sole energy source.110 Culture results in a stress response that is manifested by changes in gene expression (e.g., expression of heat-shock genes), apoptosis, and metabolism. For example, mouse blastocysts that develop in vivo convert 40–50% of the glucose to lactate, whereas blastocysts that develop in vitro from the morula stage convert ~ 100% of the glucose to lactate. Early cleavage-stage embryos also exhibit an increase in glycolysis in response to culture that is associated with reduced implantation and development following embryo transfer. Inclusion of amino acids in the medium prevents this metabolic change and restores the embryo’s ability to implant and develop following transfer.111

The changes in metabolism that occur as a response to culture conditions may be coupled either directly or indirectly to changes in gene expression. Both Whitten’s medium and KSOM support development of one-cell mouse embryos to the blastocyst stage. The major difference between these media is that KSOM has a reduced sodium chloride concentration relative to that in Whitten’s medium. When comparing gene expression in cultured embryos to in vivo–derived embryos, it has been shown that the level of expression of transcripts encoding several growth factors and their receptors is significantly reduced in embryos cultured in Whitten’s medium, whereas no significant differences were observed when the embryos were cultured in KSOM that includes amino acids.112 Of greater interest is that although culture of embryos to the blastocyst stage in KSOM maintains the appropriate maternal monoallelic expression of the imprinted H19 gene, H19 expression is essentially biallelic following culture in Whitten’s medium. 113 Moreover, the loss of imprinting was associated with the loss of methylation on the paternal, repressed allele in the differentially methylated region (DMR);113 DNA methylation of the paternal allele in this region is strongly linked with the repression of the paternal H19 allele.114

The loss of imprinting that occurs during culture under certain conditions, as well as global changes in gene expression, could have long-term consequences on development. In fact, a growing body of evidence documents that culture of bovine and ovine embryos to the blastocyst stage prior to embryo transfer results in higher incidences of fetal and perinatal loss.115117 In the bovine, these abnormalities are not observed if the embryos are first transferred to the oviducts of sheep and then transferred at the blastocyst stage to the reproductive tracts of recipient heifers.115 Thus, the abnormalities are likely attributable to embryo culture. Interestingly, large offspring derived following culture of sheep preimplantation embryos exhibit lower levels of expression of maternally expressed imprinted Igf2r gene.102 This demethylation provides a molecular basis for the reduced Igf2r expression, since, in contrast to other imprinted genes, methylation of the Igf2r gene is associated with expression via a complicated mechanism that entails expression of an antisense transcript.118,119 It should be noted that in mice, reduced Igf2r expression also leads to increased size.120 Thus, embryo culture can result in perturbations that manifest in the offspring and it is likely that epigenetic changes underlie long-term effects. Steele et al.121 have recently surveyed components of common embryo culture media and have learned that several components, including those related to methyl metabolism, are present in nophysiological concentrations. For example, vitamin B12 was either absent or present in up to 3000-fold to 5000-fold and vitamin B6 is present up to 160-fold to 500-fold the concentration reported in maternal serum or human follicular fluid. Simlarly, folate and methionine were either absent or present in supraphysiological concentrations. Based on what is being realized with respect to ART, it is conceivable that suboptimal culture conditions, such as nonphysiological levels of methyl cycle–related components in culture media, could introduce novel epigenetic perturbations in adult stem cells during processes of culture and propagation.

V. POSSIBLE IN VITRO STRATEGIES TO MINIMIZE SPONTANEOUS DIFFERENTIATION IN CULTURE

A. DNA Methyltransferase Inhibitors

The classic DNA methyltransferase inhibitor is 5-azacytidine, a derivative of the nucleoside cytidine. The inhibitor was discovered more than 40 years ago122 and its demethylating activity was discovered subsequently due to its ability to influence cellular differentiation in vitro.123 5-azacytidine is a nucleoside inhibitor that can be incorporated into DNA and can be methylated by DNA methyltransferases. The DNA methyltransferase, however, becomes covalently trapped because the intermediate cannot be resolved, inactivating the enzyme. As the enzyme is depleted, genomic DNA is demethylated as a result of continued DNA replication. To become incorporated into DNA, 5-azacytidine must be modified to a deoxyribonucleoside triphosphate prior to incorporation into DNA. Prior to modification, 5-azacytidine can become incorporated into RNA, resulting in a variety of consequences including cytotoxicity. An analogue of 5-azacytidine, 5-aza-2’deoxycytidine (decitabine), does not require modification and is thought to be less cytotoxic, but still has severe cytotoxic affects.124 For example, decitabine has been used in clinical trials and has shown promise for treatment of myeloid malignancies,125,126 but also has toxic effects including myelosuppression and neutropenic fever.124 At least three reports have investigated the treatment of donor cells with 5-aza-2’deoxycytidine prior to nuclear transfer.127129 Treatment at high levels (0.04 uM or greater) resulted in demethylation to levels of in vitro embryos, yet blastocyst development rates were reduced or unchanged compared with controls. It is unclear whether the decrease in blastocyst development rates were due to cytotoxicity or demethylation of the genome.

Several nonnucleoside compounds also inhibit DNA methyltransferase activity, none of which have been examined in nuclear transfer models. Two compounds have been reasonably characterized, whereas compounds representing three subclasses are less understood. The two characterized compounds include (−)-epigallocatechin-3-gallate (EGCG), the key polyphenol in green tea,130 and RG108, a molecule identified in an in silico screening assay.131 EGCG inhibits DNA methyltransferase activity in protein extracts and human cancer cell lines,132 and is thought to block the active site of DNMT1. Degradation of EGCG results in production of hydrogen peroxide, a strong oxidizing agent133 that may result in cytoxicity. In contrast, RG108 appears to have low toxicity and has been demonstrated to inhibit the catalytic activity of recombinant DNA methyltransferases. The inhibitory activity appears to be direct and specific for DNA methyltransferases.131 The additional classes of nonnucleoside inhibitors include the following: (i) 4-aminobenzoic acid derivatives, exemplified by Procaine,134,135 which is thought to bind to CpG-rich sequences, preventing access of DNMT to DNA, but high concentrations are thought to be required for effectiveness based on studies in cancer cell lines;136 (ii) The psammaplins, which inhibit DNMT as well as histone deacetylase activity; and (iii) oligonucleotides.

B. Limiting the Methyl Group Supply

The availability of nutrients appears to play an important role in regulating DNA methylation. A growing body of evidence suggests environmental factors such as diet can influence DNA methylation. For example, prenatal feeding of a methyl-supplemented diet can increase the level of DNA methylation and phenotypic expression of genes in offspring. The coat color in mice is determined by agouti gene expression. This is determined by the DNA methylation status of the long terminal repeat of the agouti gene in the hair follicle. If this region is hypermethylated, the mouse is agouti in color whereas if the region is hypomethylated the mouse is yellow. When pregnant female mice were fed a methyl-supplemented diet enriched in zinc, methionine, betaine, choline, folate, and vitamin B12, there was an alteration in the methylation status of the agouti long terminal repeat and none of the pups had a yellow coat.137 Thus, in utero exposure to nutrients can lead to epigenetic modifications of the genome in the offspring.

Folate, a water-soluble B vitamin, plays a significant role in DNA methylation status. The biochemical function of folate is to mediate the transfer of one-carbon moieties138140 and thus has a central role in one-carbon metabolism. Animal studies have further shown that folate deficiency causes genomic and gene-specific hypomethylation in rat liver, and the degree appears to depend on the severity and duration of folate depletion.141,142 DNA hypomethylation also has been identified in lymphocytes of humans on low dietary folate and can be reversed by folate repletion.143

Factors involved in one-carbon metabolism also likely play an important role in methylation status because they influence the supply of methyl groups and, therefore, the biochemical pathways of methylation processes. Methyl groups supplied from the “diet” (or culture environment) in the form of choline and methionine, or from the folate-dependent one-carbon pool, must be activated to S-adenosylmethionine (SAM) to serve as substrates in transmethylation reactions. Because S-adenosylhomocysteine (SAH) is a product of transmethylation reactions and a potent inhibitor of SAM-dependent methyltransferases, the ratio of SAM/SAH is an important index of transmethylation potential.144,145 When the ratio is high, methylation potential is high; when the ratio is low, methylation potential is low. It is well known that folate depletion alone is a sufficient perturbing force to diminish SAM pools. This leads to an increase in cellular levels of SAH because the equilibrium of the SAH-homocysteine interconversion actually favors SAH synthesis. Therefore, when homocysteine metabolism is inhibited (as in folate deficiency), cellular SAH levels will be increased. Increased SAH inhibits methyltransferase activity and consequently DNA methylation reactions.

The cytosolic enzyme glycine-N-methyltransferase (GNMT) functions to optimize transmethylation reactions by regulating the SAM:SAH ratio. When methyl groups are abundant and SAM levels are elevated, GNMT disposes of the excess methyl groups by forming sarcosine from glycine. SAM also reduces the supply of methyl groups originating from the one-carbon pool by inhibiting 5,10-methylenetetrahydrofolate reductase (MTHFR),146,147 the enzyme responsible for the synthesis of 5-methyltetrahydrofolate (5-methyl-THF), the folate coenzyme that donates its methyl group to homocysteine to form methionine. Because 5-methyl-THF also binds to GNMT and inhibits its activity,148,149 a decrease in 5-methyl-THF levels due to inhibition of MTHFR by SAM results in an increase in the activity of GNMT. Factors that activate GNMT lead to down regulation of methyltransferases, including DNA methyltransferases. To date, there have been no reports testing the hypothesis that activation of GNMT in cells in culture result in reduced DNA methylation; however, a few reports have described dietary manipulation of GNMT in rodents.150152 For example, diets supplemented with vitamin A and its derivatives (e.g., all-trans retinoic acid, ATRA) result in increased GNMT activity and hypomethylated DNA in rat hepatocytes, while decreases in dietary folate, choline, betaine, and vitamins B6 and B12 also result in decreased DNA methylation in specific tissues. Thus, alterations in the availability of single-carbon metabolism components (e.g., methyl donors) and adjusting the nutrient availability of methyl donors represent possible in vitro approaches to demethylate DNA and restore cell differentiation potential.

VI. FUTURE DIRECTIONS

Much remains to be done in the field of adult stem cell epigenetics despite the fact that the epigenetics of embryonic and cancer stem cells have received greater attention to date. Indeed, it is likely that adult stem cells exhibit unique epigenetic features. The following questions merit further investigation:

  1. Can culture conditions be used to direct or reprogram the epigenome in adult stem cells to improve their utility for tissue regeneration and repair?

  2. What other epigenetic targets, in addition to histones, chromatin proteins, and CpG islands, should be evaluated in adult stem cells?

  3. Can diet be used to positively modify the epigenome of adult stem cells in the living organism? If so, can this approach be used effectively as a preventive medical therapy?

  4. What elements of the adult stem cell epigenome are reversible? Can exposure to an environmental toxin, such as cigarette smoke, be reversed in adult stem cells prior to transformation into cancer stem cells? If so, how?

These are just some of the more obvious directions for research into the epigenetics of adult stem cells. New technologies and discoveries will point investigators to further, more probing questions in the future.

ACKNOWLEDGMENTS

This work was partially supported by the Pennington Biomedical Research Foundation and a CNRU Center Grant No. 1P30 DK072476 entitled “Nutritional Programming: Environmental and Molecular Interactions” sponsored by NIDDK.

REFERENCES

  • 1.Strahl BD, Allis CD. The language of covalent histone modifications. Nature. 2000;403(6765):4–5. doi: 10.1038/47412. [DOI] [PubMed] [Google Scholar]
  • 2.Jenuwein T, Allis CD. Translating the histone code. Science. 2001 Aug 10;293(5532):1074–1080. doi: 10.1126/science.1063127. [DOI] [PubMed] [Google Scholar]
  • 3.Gibbons RJ. Histone modifying and chromatin remodelling enzymes in cancer and dysplastic syndromes. Hum Mol Genet. 2005;14(Spec No 1):R85–R92. doi: 10.1093/hmg/ddi106. [DOI] [PubMed] [Google Scholar]
  • 4.Michishita E, Park JY, Burneskis JM, Barrett JC, Horikawa I. Evolutionarily conserved and nonconserved cellular localizations and functions of human SIRT proteins. Mol Biol Cell. 2005;16:4623–4635. doi: 10.1091/mbc.E05-01-0033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Schneider R, Bannister AJ, Kouzarides T. Unsafe SETs: histone lysine methyltransferases and cancer. Trends Biochem Sci. 2002;27:396–402. doi: 10.1016/s0968-0004(02)02141-2. [DOI] [PubMed] [Google Scholar]
  • 6.Nathan D, Ingvarsdottir K, Sterner DE, Bylebyl GR, Dokmanovic M, Dorsey JA, Whelan KA, Krsmanovic M, Lane WS, Meluh PB, Johnson ES, Berger SL. Histone sumoylation is a negative regulator in Saccharomyces cerevisiae and shows dynamic interplay with positive-acting histone modifications. Genes Dev. 2006;20:966–976. doi: 10.1101/gad.1404206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Fujita N, Watanabe S, Ichimura T, Tsuruzoe S, Shinkai Y, Tachibana M, Chiba T, Nakao M. Methyl-CpG binding domain 1 (MBD1) interacts with the Suv39hl-HPl heterochromatic complex for DNA methylation-based transcriptional repression. J Biol Chem. 2003;278:24132–24138. doi: 10.1074/jbc.M302283200. [DOI] [PubMed] [Google Scholar]
  • 8.Sarraf SA, Stancheva I. Methyl-CpG binding protein MBD1 couples histone H3 methylation at lysine 9 by SETDB1 to DNA replication and chromatin assembly. Mol Cell. 2004;15(4):595–605. doi: 10.1016/j.molcel.2004.06.043. [DOI] [PubMed] [Google Scholar]
  • 9.Uchimura Y, Ichimura T, Uwada J, Tachibana T, Sugahara S, Nakao M, Saitoh H. Involvement of SUMO modification in MBD1- and MCAF1-mediated heterochromatin formation. J Biol Chem. 2006;281:23180–23190. doi: 10.1074/jbc.M602280200. [DOI] [PubMed] [Google Scholar]
  • 10.Goldknopf IL, French MF, Musso R, Busch H. Presence of protein A24 in rat liver nucleosomes. Proc Natl Acad Sci USA. 1977;74:5492–5495. doi: 10.1073/pnas.74.12.5492. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.West MH, Bonner WM. Histone 2B can be modified by the attachment of ubiquitin. Nucleic Acids Res. 1980;8(20):4671–4680. doi: 10.1093/nar/8.20.4671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Sun ZW, Allis CD. Ubiquitination of histone H2B regulates H3 methylation and gene silencing in yeast. Nature. 2002;418(6893):104–108. doi: 10.1038/nature00883. [DOI] [PubMed] [Google Scholar]
  • 13.Nickel BE, Allis CD, Davie JR. Ubiquitinated histone H2B is preferentially located in transcriptionally active chromatin. Biochemistry. 1989;28(3):958–963. doi: 10.1021/bi00429a006. [DOI] [PubMed] [Google Scholar]
  • 14.Daniel JA, Torok MS, Sun ZW, Schieltz D, Allis CD, Yates JR, 3rd, Grant PA. Deubiquitination of histone H2B by a yeast acetyltransferase complex regulates transcription. J Biol Chem. 2004;279:1867–1871. doi: 10.1074/jbc.C300494200. [DOI] [PubMed] [Google Scholar]
  • 15.Henry KW, Wyce A, Lo WS, Duggan LJ, Emre NC, Kao CF, Pillus L, Shilatifard A, Osley MA, Berger SL. Transcriptional activation via sequential histone H2B ubiquitylation and deubiquitylation, mediated by SAGA-associated Ubp8. Genes Dev. 2003;17:2648–2663. doi: 10.1101/gad.1144003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Sridhar VV, Kapoor A, Zhang K, Zhu J, Zhou T, Hasegawa PM, Bressan RA, Zhu JK. Control of DNA methylation and heterochromatic silencing by histone H2B deubiquitination. Nature. 2007;447:735–738. doi: 10.1038/nature05864. [DOI] [PubMed] [Google Scholar]
  • 17.Wang H, Wang L, Erdjument-Bromage H, Vidal M, Tempst P, Jones RS, Zhang Y. Role of histone H2A ubiquitination in Polycomb silencing. Nature. 2004;431:873–878. doi: 10.1038/nature02985. [DOI] [PubMed] [Google Scholar]
  • 18.Denell RE, Frederick RD. Homoeosis in Drosophila: a description of the Polycomb lethal syndrome. Dev Biol. 1983;97:34–47. doi: 10.1016/0012-1606(83)90061-1. [DOI] [PubMed] [Google Scholar]
  • 19.Kennison JA. Introduction to Trx-G and Pc-G genes. Methods Enzymol. 2004;377:61–70. doi: 10.1016/S0076-6879(03)77003-7. [DOI] [PubMed] [Google Scholar]
  • 20.Lewis EB. A gene complex controlling segmentation in Drosophila. Nature. 1978;276(5688):565–570. doi: 10.1038/276565a0. [DOI] [PubMed] [Google Scholar]
  • 21.Orlando V, Paro R. Chromatin multiprotein complexes involved in the maintenance of transcription patterns. Curr Opin Genet Dev. 1995;5:174–179. doi: 10.1016/0959-437x(95)80005-0. [DOI] [PubMed] [Google Scholar]
  • 22.Pirrotta V. Polycombing the genome: PcG, trxG, and chromatin silencing. Cell. 1998;93:333–336. doi: 10.1016/s0092-8674(00)81162-9. [DOI] [PubMed] [Google Scholar]
  • 23.Simon J, Chiang A, Bender W. Ten different Polycomb group genes are required for spatial control of the abdA and AbdB homeotic products. Development. 1992;114:493–505. doi: 10.1242/dev.114.2.493. [DOI] [PubMed] [Google Scholar]
  • 24.Levine SS, King IF, Kingston RE. Division of labor in polycomb group repression. Trends Biochem Sci. 2004;29:478–485. doi: 10.1016/j.tibs.2004.07.007. [DOI] [PubMed] [Google Scholar]
  • 25.Lund AH, van Lohuizen M. Polycomb complexes and silencing mechanisms. Curr Opin Cell Biol. 2004;16:239–246. doi: 10.1016/j.ceb.2004.03.010. [DOI] [PubMed] [Google Scholar]
  • 26.Ringrose L, Paro R. Epigenetic regulation of cellular memory by the Polycomb and Trithorax group proteins. Annu Rev Genet. 2004;38:413–443. doi: 10.1146/annurev.genet.38.072902.091907. [DOI] [PubMed] [Google Scholar]
  • 27.Cao R, Wang L, Wang H, Xia L, Erdjument-Bromage H, Tempst P, Jones RS, Zhang Y. Role of histone H3 lysine 27 methylation in Polycomb-group silencing. Science. 2002;298:1039–1043. doi: 10.1126/science.1076997. [DOI] [PubMed] [Google Scholar]
  • 28.Czermin B, Melfi R, McCabe D, Seitz V, Imhof A, Pirrotta V. Drosophila enhancer of Zeste/ESC complexes have a histone H3 methyltransferase activity that marks chromosomal Polycomb sites. Cell. 2002;111(2):185–196. doi: 10.1016/s0092-8674(02)00975-3. [DOI] [PubMed] [Google Scholar]
  • 29.Kirmizis A, Bartley SM, Kuzmichev A, Margueron R, Reinberg D, Green R, Farnham PJ. Silencing of human polycomb target genes is associated with methylation of histone H3 Lys 27. Genes Dev. 2004;18:1592–1605. doi: 10.1101/gad.1200204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Kuzmichev A, Nishioka K, Erdjument-Bromage H, Tempst P, Reinberg D. Histone methyltransferase activity associated with a human multiprotein complex containing the Enhancer of Zeste protein. Genes Dev. 2002;16(22):2893–2905. doi: 10.1101/gad.1035902. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Muller J, Hart CM, Francis NJ, Vargas ML, Sengupta A, Wild B, Miller EL, O'Connor MB, Kingston RE, Simon JA. Histone methyltransferase activity of a Drosophila Polycomb group repressor complex. Cell. 2002;111:197–208. doi: 10.1016/s0092-8674(02)00976-5. [DOI] [PubMed] [Google Scholar]
  • 32.Cao R, Zhang Y. SUZ12 is required for both the histone methyltransferase activity and the silencing function of the EED-EZH2 complex. Mol Cell. 2004;15(1):57–67. doi: 10.1016/j.molcel.2004.06.020. [DOI] [PubMed] [Google Scholar]
  • 33.Pasini D, Bracken AP, Jensen MR, Lazzerini Denchi E, Helin K. Suzl2 is essential for mouse development and for EZH2 histone methyltransferase activity. Embo J. 2004;23(20):4061–4071. doi: 10.1038/sj.emboj.7600402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Francis NJ, Saurin AJ, Shao Z, Kingston RE. Reconstitution of a functional core polycomb repressive complex. Mol Cell. 2001;8(3):545–556. doi: 10.1016/s1097-2765(01)00316-1. [DOI] [PubMed] [Google Scholar]
  • 35.Shao Z, Raible F, Mollaaghababa R, Guyon JR, Wu CT, Bender W, Kingston RE. Stabilization of chromatin structure by PRC1, a Polycomb complex. Cell. 1999;98:37–46. doi: 10.1016/S0092-8674(00)80604-2. [DOI] [PubMed] [Google Scholar]
  • 36.Razin A, Riggs AD. DNA methylation and gene function. Science. 1980;210(4470):604–610. doi: 10.1126/science.6254144. [DOI] [PubMed] [Google Scholar]
  • 37.Szyf M. DNA methylation and demethylation as targets for anticancer therapy. Biochemistry (Mosc) 2005;70:533–549. doi: 10.1007/s10541-005-0147-7. [DOI] [PubMed] [Google Scholar]
  • 38.Ramchandani S, Bhattacharya SK, Cervoni N, Szyf M. DNA methylation is a reversible biological signal. Proc Natl Acad Sci U S A. 1999;96(11):6107–6112. doi: 10.1073/pnas.96.11.6107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Jones PL, Veenstra GJ, Wade PA, Vermaak D, Kass SU, Landsberger N, Strouboulis J, Wolffe AP. Methylated DNA and MeCP2 recruit histone deacetylase to repress transcription. Nat Genet. 1998;19:187–191. doi: 10.1038/561. [DOI] [PubMed] [Google Scholar]
  • 40.Nan X, Ng HH, Johnson CA, Laherty CD, Turner BM, Eisenman RN, Bird A. Transcriptional repression by the methyl-CpG-binding protein MeCP2 involves a histone deacetylase complex. Nature. 1998;393:386–389. doi: 10.1038/30764. [DOI] [PubMed] [Google Scholar]
  • 41.Goll MG, Bestor TH. Eukaryotic cytosine methyltransferases. Annu Rev Biochem. 2005;74:481–514. doi: 10.1146/annurev.biochem.74.010904.153721. [DOI] [PubMed] [Google Scholar]
  • 42.Leonhardt H, Page AW, Weier HU, Bestor TH. A targeting sequence directs DNA methyltransferase to sites of DNA replication in mammalian nuclei. Cell. 1992;71(5):865–873. doi: 10.1016/0092-8674(92)90561-p. [DOI] [PubMed] [Google Scholar]
  • 43.Chuang LS, Ian HI, Koh TW, Ng HH, Xu G, Li BF. Human DNA-(cytosine-5) methyltransferase-PCNA complex as a target for p21WAFl. Science. 1997 doi: 10.1126/science.277.5334.1996. 1996-2000277(5334) [DOI] [PubMed] [Google Scholar]
  • 44.Araujo FD, Croteau S, Slack AD, Milutinovic S, Bigey P, Price GB, Zannis-Hajopoulos M, Szyf M. The DNMT1 target recognition domain resides in the N terminus. J Biol Chem. 2001;276:6930–6936. doi: 10.1074/jbc.M009037200. [DOI] [PubMed] [Google Scholar]
  • 45.Kim GD, Ni J, Kelesoglu N, Roberts RJ, Pradhan S. Co-operation and communication between the human maintenance and de novo DNA (cytosine-5) methyltransferases. Embo J. 2002;21(15):4183–4195. doi: 10.1093/emboj/cdf401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Liang G, Chan MF, Tomigahara Y, Tsai YC, Gonzales FA, Li E, Laird PW, Jones PA. Cooperativity between DNA methyltransferases in the maintenance methylation of repetitive elements. Mol Cell Biol. 2002;22:480–491. doi: 10.1128/MCB.22.2.480-491.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Bourc'his D, Bestor TH. Meiotic catastrophe and retrotransposon reactivation in male germ cells lacking Dnmt3L. Nature. 2004;431(7004):96–99. doi: 10.1038/nature02886. [DOI] [PubMed] [Google Scholar]
  • 48.Bourc'his D, Xu GL, Lin CS, Bollman B, Bestor TH. Dnmt3L and the establishment of maternal genomic imprints. Science. 2001;294(5551):2536–2539. doi: 10.1126/science.1065848. [DOI] [PubMed] [Google Scholar]
  • 49.Vire E, Brenner C, Deplus R, Blanchon L, Fraga M, Didelot C, Morey L, Van Eynde A, Bernard D, Vanderwinden JM, Bollen M, Esteller M, Di Croce L, de Launoit Y, Fuks F. The Polycomb group protein EZH2 directly controls DNA methylation. Nature. 2006;439:871–874. doi: 10.1038/nature04431. [DOI] [PubMed] [Google Scholar]
  • 50.Acharya MR, Karp JE, Sausville EA, Hwang K, Ryan Q, Gojo I, Venitz J, Figg WD, Sparreboom A. Factors affecting the pharmacokinetic profile of MS-275, a novel histone deacetylase inhibitor, in patients with cancer. Invest New Drugs. 2006;24:367–375. doi: 10.1007/s10637-005-5707-6. [DOI] [PubMed] [Google Scholar]
  • 51.Rabizadeh E, Merkin V, Belyaeva I, Shaklai M, Zimra Y. Pivanex, a histone deacetylase inhibitor, induces changes in BCR-ABL expression and when combined with STI571, acts synergistically in a chronic myelocytic leukemia cell line. Leuk Res. 2007;31(8):1115–1123. doi: 10.1016/j.leukres.2006.12.015. [DOI] [PubMed] [Google Scholar]
  • 52.Vinodhkumar R, Song YS, Ravikumar V, Ramakrishnan G, Devaki T. Depsipeptide a histone deacetlyase inhibitor down regulates levels of matrix metalloproteinases 2 and 9 mRNA and protein expressions in lung cancer cells (A549) Chem Biol Interact. 2007;165(3):220–229. doi: 10.1016/j.cbi.2006.12.012. [DOI] [PubMed] [Google Scholar]
  • 53.Popova BC, Tada T, Takagi N, Brockdorff N, Nesterova TB. Attenuated spread of X-inactivation in an X; autosome translocation. Proc Natl Acad Sci USA. 2006;103(20):7706–7711. doi: 10.1073/pnas.0602021103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Taylor SM, Jones PA. Multiple new phenotypes induced in 10T1/2 and 3T3 cells treated with 5-azacytidine. Cell. 1979;17(4):771–779. doi: 10.1016/0092-8674(79)90317-9. [DOI] [PubMed] [Google Scholar]
  • 55.Davis RL, Weintraub H, Lassar AB. Expression of a single transfected cDNA converts fibroblasts to myoblasts. Cell. 1987;51(6):987–1000. doi: 10.1016/0092-8674(87)90585-x. [DOI] [PubMed] [Google Scholar]
  • 56.Lassar AB, Paterson BM, Weintraub H. Transfection of a DNA locus that mediates the conversion of 10T1/2 fibroblasts to myoblasts. Cell. 1986;47(5):649–656. doi: 10.1016/0092-8674(86)90507-6. [DOI] [PubMed] [Google Scholar]
  • 57.Ley TJ, DeSimone J, Anagnou NP, Keller GH, Humphries RK, Turner PH, Young NS, Keller P, Nienhuis AW. 5-azacytidine selectively increases gamma-globin synthesis in a patient with beta+ thalassemia. N Engl J Med. 1982;307:1469–1475. doi: 10.1056/NEJM198212093072401. [DOI] [PubMed] [Google Scholar]
  • 58.Perrine SP, Ginder GD, Faller DV, Dover GH, Ikuta T, Witkowska HE, Cai SP, Vichinsky EP, Olivieri NF. A short-term trial of butyrate to stimulate fetal-globin-gene expression in the beta-globin disorders. N Engl J Med. 1993;328:81–86. doi: 10.1056/NEJM199301143280202. [DOI] [PubMed] [Google Scholar]
  • 59.Gabbianelli M, Testa U, Massa A, Pelosi E, Sposi NM, Riccioni R, Luchetti L, Peschle C. Hemoglobin switching in unicellular erythroid culture of sibling erythroid burst-forming units: kit ligand induces a dose-dependent fetal hemoglobin reactivation potentiated by sodium butyrate. Blood. 2000;95:3555–3561. [PubMed] [Google Scholar]
  • 60.Marianna P, Kollia P, Akel S, Papassotiriou Y, Stamoulakatou A, Loukopoulos D. Valproic acid, trichostatin and their combination with hemin preferentially enhance gamma-globin gene expression in human erythroid liquid cultures. Haematologica. 2001;86:700–705. [PubMed] [Google Scholar]
  • 61.Swank RA, Skarpidi E, Papayannopoulou T, Stamatoyannopoulos G. The histone deacetylase inhibitor, trichostatin A, reactivates the developmentally silenced gamma globin expression in somatic cell hybrids and induces gamma gene expression in adult BFUe cultures. Blood Cells Mol Dis. 2003;30(3):254–257. doi: 10.1016/s1079-9796(03)00024-x. [DOI] [PubMed] [Google Scholar]
  • 62.Zhao XY, Sakashita K, Kamijo T, Hidaka E, Sugane K, Kubota T, Koike K. Granulocyte-macrophage colony-stimulating factor induces de novo methylation of the pi5 CpG island in hematopoietic cells. Cytokine. 2005;31:203–212. doi: 10.1016/j.cyto.2005.04.010. [DOI] [PubMed] [Google Scholar]
  • 63.Araki H, Yoshinaga K, Boccuni P, Zhao Y, Hoffman R, Mahmud N. Chromatin-modifying agents permit human hematopoietic stem cells to undergo multiple cell divisions while retaining their repopulating potential. Blood. 2007;109:3570–3578. doi: 10.1182/blood-2006-07-035287. [DOI] [PubMed] [Google Scholar]
  • 64.Araki H, Yoshinaga K, Boccuni P, Zhao Y, Hoffman R, Mahmud N. Chromatin-modifying agents permit human hematopoietic stem cells to undergo multiple cell divisions while retaining their repopulating potential. Blood. 2007;109(8):3570–3578. doi: 10.1182/blood-2006-07-035287. [DOI] [PubMed] [Google Scholar]
  • 65.Milhem M, Mahmud N, Lavelle D, Araki H, DeSimone J, Saunthararajah Y, Hoffman R. Modification of hematopoietic stem cell fate by 5aza 2'deoxycytidine and trichostatin A. Blood. 2004;103:4102–4110. doi: 10.1182/blood-2003-07-2431. [DOI] [PubMed] [Google Scholar]
  • 66.Young JC, Wu S, Hansteen G, Du C, Sambucetti L, Remiszewski S, O'Farrell AM, Hill B, Lavau C, Murray LJ. Inhibitors of histone deacetylases promote hematopoietic stem cell self-renewal. Cytotherapy. 2004;6:328–336. doi: 10.1080/14653240410004899. [DOI] [PubMed] [Google Scholar]
  • 67.Koh SH, Choi HS, Park ES, Kang HJ, Ahn HS, Shin HY. Co-culture of human CD34+ cells with mesenchymal stem cells increases the survival of CD34+ cells against the 5-aza-deoxycytidine- or trichostatin A-induced cell death. Biochem Biophys Res Commun. 2005;329(3):1039–1045. doi: 10.1016/j.bbrc.2005.02.077. [DOI] [PubMed] [Google Scholar]
  • 68.Nemeth MJ, Kirby MR, Bodine DM. Hmgb3 regulates the balance between hematopoietic stem cell self-renewal and differentiation. Proc Natl Acad Sci USA. 2006;103(37):13783–13788. doi: 10.1073/pnas.0604006103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Kajiume T, Ninomiya Y, Ishihara H, Kanno R, Kanno M. Polycomb group gene mel-18 modulates the self-renewal activity and cell cycle status of hematopoietic stem cells. Exp Hematol. 2004;32(6):571–578. doi: 10.1016/j.exphem.2004.03.001. [DOI] [PubMed] [Google Scholar]
  • 70.Lessard J, Baban S, Sauvageau G. Stage-specific expression of polycomb group genes in human bone marrow cells. Blood. 1998;91(4):1216–1224. [PubMed] [Google Scholar]
  • 71.Savarese F, Flahndorfer K, Jaenisch R, Busslinger M, Wutz A. Hematopoietic precursor cells transiently reestablish permissiveness for X inactivation. Mol Cell Biol. 2006;26(19):7167–7177. doi: 10.1128/MCB.00810-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Balana B, Nicoletti C, Zahanich I, Graf EM, Christ T, Boxberger S, Ravens U. 5-Azacytidine induces changes in electrophysiological properties of human mesenchymal stem cells. Cell Res. 2006;16:949–960. doi: 10.1038/sj.cr.7310116. [DOI] [PubMed] [Google Scholar]
  • 73.Fukuda K. Use of adult marrow mesenchymal stem cells for regeneration of cardiomyocytes. Bone Marrow Transplant. 2003;32 Suppl 1:S25–S27. doi: 10.1038/sj.bmt.1703940. [DOI] [PubMed] [Google Scholar]
  • 74.Makino S, Fukuda K, Miyoshi S, Konishi F, Kodama H, Pan J, Sano M, Takahashi T, Hori S, Abe H, Hata J, Umezawa A, Ogawa S. Cardiomyocytes can be generated from marrow stromal cells in vitro. J Clin Invest. 1999;103:697–705. doi: 10.1172/JCI5298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Planat-Benard V, Menard C, Andre M, Puceat M, Perez A, Garcia-Verdugo JM, Penicaud L, Casteilla L. Spontaneous cardiomyocyte differentiation from adipose tissue stroma cells. Circ Res. 2004;94:223–229. doi: 10.1161/01.RES.0000109792.43271.47. [DOI] [PubMed] [Google Scholar]
  • 76.Rangappa S, Fen C, Lee EH, Bongso A, Sim EK. Transformation of adult mesenchymal stem cells isolated from the fatty tissue into cardiomyocytes. Ann Thorac Surg. 2003;75(3):775–779. doi: 10.1016/s0003-4975(02)04568-x. [DOI] [PubMed] [Google Scholar]
  • 77.Boquest AC, Noer A, Sorensen AL, Vekterud K, Collas P. CpG methylation profiles of endothelial cell-specific gene promoter regions in adipose tissue stem cells suggest limited differentiation potential toward the endothelial cell lineage. Stem Cells. 2007;25(4):852–861. doi: 10.1634/stemcells.2006-0428. [DOI] [PubMed] [Google Scholar]
  • 78.Kang MI, Kim HS, Jung YC, Kim YH, Hong SJ, Kim MK, Baek KH, Kim CC, Rhyu MG. Transitional CpG methylation between promoters and retroelements of tissue-specific genes during human mesenchymal cell differentiation. J Cell Biochem. 2007 doi: 10.1002/jcb.21291. [DOI] [PubMed] [Google Scholar]
  • 79.Noer A, Sorensen AL, Boquest AC, Collas P. Stable CpG hypomethylation of adipogenic promoters in freshly isolated, cultured, and differentiated mesenchymal stem cells from adipose tissue. Mol Biol Cell. 2006;17:3543–3556. doi: 10.1091/mbc.E06-04-0322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Iezzi S, Di Padova M, Serra C, Caretti G, Simone C, Maklan E, Minetti G, Zhao P, Hoffman EP, Puri PL, Sartorelli V. Deacetylase inhibitors increase muscle cell size by promoting myoblast recruitment and fusion through induction of follistatin. Dev Cell. 2004;6:673–684. doi: 10.1016/s1534-5807(04)00107-8. [DOI] [PubMed] [Google Scholar]
  • 81.Minetti GC, Colussi C, Adami R, Serra C, Mozzetta C, Parente V, Fortuni S, Straino S, Sampaolesi M, Di Padova M, Illi B, Gallinari P, Steinkuhler C, Capogrossi MC, Sartorelli V, Bottinelli R, Gaetano C, Puri PL. Functional and morphological recovery of dystrophic muscles in mice treated with deacetylase inhibitors. Nat Med. 2006;12:1147–1150. doi: 10.1038/nm1479. [DOI] [PubMed] [Google Scholar]
  • 82.Lagace DC, Nachtigal MW. Inhibition of histone deacetylase activity by valproic acid blocks adipogenesis. J Biol Chem. 2004;279(18):18851–18860. doi: 10.1074/jbc.M312795200. [DOI] [PubMed] [Google Scholar]
  • 83.Kuroda M, Tanabe H, Yoshida K, Oikawa K, Saito A, Kiyuna T, Mizusawa H, Mukai K. Alteration of chromosome positioning during adipocyte differentiation. J Cell Sci. 2004;l17:5897–5903. doi: 10.1242/jcs.01508. [DOI] [PubMed] [Google Scholar]
  • 84.Beresford JN, Bennett JH, Devlin C, Leboy PS, Owen ME. Evidence for an inverse relationship between the differentiation of adipocytic and osteogenic cells in rat marrow stromal cell cultures. J Cell Sci. 1992;102(Pt 2):341–351. doi: 10.1242/jcs.102.2.341. [DOI] [PubMed] [Google Scholar]
  • 85.Gimble JM, Zvonic S, Floyd ZE, Kassem M, Nuttall ME. Playing with bone and fat. J Cell Biochem. 2006;98(2):251–266. doi: 10.1002/jcb.20777. [DOI] [PubMed] [Google Scholar]
  • 86.Boer JD, Licht R, Bongers M, Klundert TV, Arends R, Blitterswijk CV. Inhibition of Histone Acetylation as a Tool in Bone Tissue Engineering. Tissue Eng. 2006;12(10):2927–2937. doi: 10.1089/ten.2006.12.2927. [DOI] [PubMed] [Google Scholar]
  • 87.Cho HH, Park HT, Kim YJ, Bae YC, Suh KT, Jung JS. Induction of osteogenic differentiation of human mesenchymal stem cells by histone deacetylase inhibitors. J Cell Biochem. 2005;96(3):533–542. doi: 10.1002/jcb.20544. [DOI] [PubMed] [Google Scholar]
  • 88.Thomas D, Kansara M. Epigenetic modifications in osteogenic differentiation and transformation. J Cell Biochem. 2006;98(4):757–769. doi: 10.1002/jcb.20850. [DOI] [PubMed] [Google Scholar]
  • 89.Torres-Padilla ME, Parfitt DE, Kouzarides T, Zernicka-Goetz M. Histone arginine methylation regulates pluripotency in the early mouse embryo. Nature. 2007;445(7124):214–218. doi: 10.1038/nature05458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Cheng LC, Tavazoie M, Doetsch F. Stem cells: from epigenetics to microRNAs. Neuron. 2005;46(3):363–367. doi: 10.1016/j.neuron.2005.04.027. [DOI] [PubMed] [Google Scholar]
  • 91.Richard S, Torabi N, Franco GV, Tremblay GA, Chen T, Vogel G, Morel M, Cleroux P, Forget-Richard A, Komarova S, Tremblay ML, Li W, Li A, Gao YJ, Henderson JE. Ablation of the Sam68 RNA binding protein protects mice from age-related bone loss. PLoS Genet 1. 2005:e74. doi: 10.1371/journal.pgen.0010074. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Tayaramma T, Ma B, Rohde M, Mayer H. Chromatin-remodeling factors allow differentiation of bone marrow cells into insulin-producing cells. Stem Cells. 2006;24:2858–2867. doi: 10.1634/stemcells.2006-0109. [DOI] [PubMed] [Google Scholar]
  • 93.Snykers S, Vanhaecke T, De Becker A, Papeleu P, Vinken M, Van Riet I, Rogiers V. Chromatin remodelling agent trichostatin A: a key-factor in the hepatic differentiation of human mesenchymal stem cells derived of adult bone marrow. BMC Dev Biol. 2007;7:24. doi: 10.1186/1471-213X-7-24. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Niki T, Rombouts K, De Bleser P, De Smet K, Rogiers V, Schuppan D, Yoshida M, Gabbiani G, Geerts A. A histone deacetylase inhibitor, trichostatin A, suppresses myofibroblastic differentiation of rat hepatic stellate cells in primary culture. Hepatology. 1999;29:858–867. doi: 10.1002/hep.510290328. [DOI] [PubMed] [Google Scholar]
  • 95.Rombouts K, Niki T, Wielant A, Hellemans K, Geerts A, Trichostatin A. lead compound for development of antifibrogenic drugs. Acta Gastroenterol Belg. 2001;64(3):239–246. [PubMed] [Google Scholar]
  • 96.Hsieh J, Nakashima K, Kuwabara T, Mejia E, Gage FH. Histone deacetylase inhibition-mediated neuronal differentiation of multipotent adult neural progenitor cells. Proc Natl Acad Sci USA. 2004;101(47):16659–16664. doi: 10.1073/pnas.0407643101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Hahnen E, Eyupoglu IY, Brichta L, Haastert K, Trankle C, Siebzehnrubl FA, Riessland M, Holker I, Claus P, Romstock J, Buslei R, Wirth B, Blumcke I. In vitro and ex vivo evaluation of second-generation histone deacetylase inhibitors for the treatment of spinal muscular atrophy. J Neurochem. 2006;98:193–202. doi: 10.1111/j.1471-4159.2006.03868.x. [DOI] [PubMed] [Google Scholar]
  • 98.Siebzehnrubl FA, Buslei R, Eyupoglu IY, Seufert S, Hahnen E, Blumcke I. Histone deacetylase inhibitors increase neuronal differentiation in adult forebrain precursor cells. Exp Brain Res. 2007;176(4):672–678. doi: 10.1007/s00221-006-0831-x. [DOI] [PubMed] [Google Scholar]
  • 99.Safford KM, Safford SD, Gimble JM, Shetty AK, Rice HE. Characterization of neuronal/glial differentiation of murine adipose-derived adult stromal cells. Exp Neurol. 2004;187(2):319–328. doi: 10.1016/j.expneurol.2004.01.027. [DOI] [PubMed] [Google Scholar]
  • 100.Woodbury D, Schwarz EJ, Prockop DJ, Black IB. Adult rat and human bone marrow stromal cells differentiate into neurons. J Neurosci Res. 2000;61(4):364–370. doi: 10.1002/1097-4547(20000815)61:4<364::AID-JNR2>3.0.CO;2-C. [DOI] [PubMed] [Google Scholar]
  • 101.Khosla S, Dean W, Brown D, Reik W, Feil R. Culture of preimplantation mouse embryos affects fetal development and the expression of imprinted genes. Biol Reprod. 2001;64:918–926. doi: 10.1095/biolreprod64.3.918. [DOI] [PubMed] [Google Scholar]
  • 102.Young LE, Fernandes K, McEvoy TG, Butterwith SC, Gutierrez CG, Carolan C, Broadbent PJ, Robinson JJ, Wilmut I, Sinclair KD. Epigenetic change in IGF2R is associated with fetal overgrowth after sheep embryo culture. Nat Genet. 2001;27:153–154. doi: 10.1038/84769. [DOI] [PubMed] [Google Scholar]
  • 103.Ludwig M, Katalinic A, Gross S, Sutcliffe A, Varon R, Horsthemke B. Increased prevalence of imprinting defects in patients with Angelman syndrome born to subfertile couples. J Med Genet. 2005;42:289–291. doi: 10.1136/jmg.2004.026930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Maher ER, Brueton LA, Bowdin SC, Luharia A, Cooper W, Cole TR, Macdonald F, Sampson JR, Barratt CL, Reik W, Hawkins MM. Beckwith-Wiedemann syndrome and assisted reproduction technology (ART) J Med Genet. 2003;40:62–64. doi: 10.1136/jmg.40.1.62. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Dean W, Santos F, Stojkovic M, Zakhartchenko V, Walter J, Wolf E, Reik W. Conservation of methylation reprogramming in mammalian development: aberrant reprogramming in cloned embryos. Proc Natl Acad Sci USA. 2001;98:13734–13738. doi: 10.1073/pnas.241522698. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Pfister-Genskow M, Myers C, Childs LA, Lacson JC, Patterson T, Betthauser JM, Goueleke PJ, Koppang RW, Lange G, Fisher P, Watt SR, Forsberg EJ, Zheng Y, Leno GH, Schultz RM, Liu B, Chetia C, Yang X, Hoeschele I, Eilertsen KJ. Identification of differentially expressed genes in individual bovine preimplantation embryos produced by nuclear transfer: improper reprogramming of genes required for development. Biol Reprod. 2005;72:546–555. doi: 10.1095/biolreprod.104.031799. [DOI] [PubMed] [Google Scholar]
  • 107.Humphries A, Klein D, Baler R, Carter DA. cDNA array analysis of pineal gene expression reveals circadian rhythmicity of the dominant negative helix-loop-helix protein-encoding gene, Id-1. J Neuroendocrinol. 2002;14:101–108. doi: 10.1046/j.0007-1331.2001.00738.x. [DOI] [PubMed] [Google Scholar]
  • 108.Bortvin A, Eggan K, Skaletsky H, Akutsu H, Berry DL, Yanagimachi R, Page DC, Jaenisch R. Incomplete reactivation of Oct4- related genes in mouse embryos cloned from somatic nuclei. Development. 2003;130:1673–1680. doi: 10.1242/dev.00366. [DOI] [PubMed] [Google Scholar]
  • 109.Mann MR, Chung YG, Nolen LD, Verona RI, Latham KE, Bartolomei MS. Disruption of imprinted gene methylation and expression in cloned preimplantation stage mouse embryos. Biol Reprod. 2003;69(3):902–914. doi: 10.1095/biolreprod.103.017293. [DOI] [PubMed] [Google Scholar]
  • 110.Gardner DK. Changes in requirements and utilization of nutrients during mammalian preimplantation embryo development and their significance in embryo culture. Theriogenology. 1998;49:83–102. doi: 10.1016/s0093-691x(97)00404-4. [DOI] [PubMed] [Google Scholar]
  • 111.Lane M, Gardner DK. Amino acids and vitamins prevent culture-induced metabolic perturbations and associated loss of viability of mouse blastocysts. Hum Reprod. 1998;13:991–997. doi: 10.1093/humrep/13.4.991. [DOI] [PubMed] [Google Scholar]
  • 112.Ho Y, Wigglesworth K, Eppig JJ, Schultz RM. Preimplantation development of mouse embryos in KSOM: augmentation by amino acids and analysis of gene expression. Mol Reprod Dev. 1995;41(2):232–238. doi: 10.1002/mrd.1080410214. [DOI] [PubMed] [Google Scholar]
  • 113.Doherty AS, Mann MR, Tremblay KD, Bartolomei MS, Schultz RM. Differential effects of culture on imprinted H19 expression in the preimplantation mouse embryo. Biol Reprod. 2000;62:1526–1536. doi: 10.1095/biolreprod62.6.1526. [DOI] [PubMed] [Google Scholar]
  • 114.Thorvaldsen JL, Duran KL, Bartolomei MS. Deletion of the H19 differentially methylated domain results in loss of imprinted expression of H19 and Igf2. Genes Dev. 1998;12(23):3693–3702. doi: 10.1101/gad.12.23.3693. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Behboodi E, Anderson GB, Bondurant RH, Cargill SL, Kreuscher BR, Medrano JF, Murray JD. Birth of large calves that developed from in vitro-derived bovine embryos. Theriogenology. 1995;44:227–232. doi: 10.1016/0093-691x(95)00172-5. [DOI] [PubMed] [Google Scholar]
  • 116.Kruip TAM, den Dass JHG. In vitro produced and cloned embryos: Effects on pregnancy, parturition and offspring. Theriogenology. 1997;47(1):43–52. [Google Scholar]
  • 117.Thompson JG, Gardner DK, Pugh PA, McMillan WH, Tervit HR. Lamb birth weight is affected by culture system utilized during in vitro pre-elongation development of ovine embryos. Biol Reprod. 1995;53(6):1385–1391. doi: 10.1095/biolreprod53.6.1385. [DOI] [PubMed] [Google Scholar]
  • 118.Lyle R, Watanabe D, te Vruchte D, Lerchner W, Smrzka OW, Wutz A, Schageman J, Hahner L, Davies C, Barlow DP. The imprinted antisense RNA at the Igf2r locus overlaps but does not imprint Mas1. Nat Genet. 2000;25:19–21. doi: 10.1038/75546. [DOI] [PubMed] [Google Scholar]
  • 119.Reik W, Walter J. Evolution of imprinting mechanisms: the battle of the sexes begins in the zygote. Nat Genet. 2001;27:255–256. doi: 10.1038/85804. [DOI] [PubMed] [Google Scholar]
  • 120.Wang ZQ, Fung MR, Barlow DP, Wagner EF. Regulation of embryonic growth and lysosomal targeting by the imprinted Igf2/Mpr gene. Nature. 1994;372(6505):464–467. doi: 10.1038/372464a0. [DOI] [PubMed] [Google Scholar]
  • 121.Steele SR, Mullenix PS, Martin MJ, Place RJ. The effect of combat rations on bowel habits in a combat environment. Am J Surg. 2005;189:518–521. doi: 10.1016/j.amjsurg.2005.01.026. discussion 21. [DOI] [PubMed] [Google Scholar]
  • 122.Sorm F, Piskala A, Cihak A, Vesely J. 5-Azacytidine, a new, highly effective cancerostatic. Experientia. 1964;20(4):202–203. doi: 10.1007/BF02135399. [DOI] [PubMed] [Google Scholar]
  • 123.Jones PA, Taylor SM. Cellular differentiation, cytidine analogs and DNA methylation. Cell. 1980;20(1):85–93. doi: 10.1016/0092-8674(80)90237-8. [DOI] [PubMed] [Google Scholar]
  • 124.Issa JP, Gharibyan V, Cortes J, Jelinek J, Morris G, Verstovsek S, Talpaz M, Garcia-Manero G, Kantarjian HM. Phase II study of low-dose decitabine in patients with chronic myelogenous leukemia resistant to imatinib mesylate. J Clin Oncol. 2005;23:3948–3956. doi: 10.1200/JCO.2005.11.981. [DOI] [PubMed] [Google Scholar]
  • 125.Leone G, Voso MT, Teofili L, Lubbert M. Inhibitors of DNA methylation in the treatment of hematological malignancies and MDS. Clin Immunol. 2003;109(l):89–102. doi: 10.1016/s1521-6616(03)00207-9. [DOI] [PubMed] [Google Scholar]
  • 126.Santini V, Kantarjian HM, Issa JP. Changes in DNA methylation in neoplasia: pathophysiology and therapeutic implications. Ann Intern Med. 2001;134(7):573–586. doi: 10.7326/0003-4819-134-7-200104030-00011. [DOI] [PubMed] [Google Scholar]
  • 127.Enright BP, Kubota C, Yang X, Tian XC. Epigenetic characteristics and development of embryos cloned from donor cells treated by trichostatin A or 5-aza-2'-deoxycytidine. Biol Reprod. 2003;69(3):896–901. doi: 10.1095/biolreprod.103.017954. [DOI] [PubMed] [Google Scholar]
  • 128.Enright BP, Sung LY, Chang CC, Yang X, Tian XC. Methylation and acetylation characteristics of cloned bovine embryos from donor cells treated with 5-aza-2'-deoxycytidine. Biol Reprod. 2005;72:944–948. doi: 10.1095/biolreprod.104.033225. [DOI] [PubMed] [Google Scholar]
  • 129.Jones KL, Hill J, Shin TY, Lui L, Westhusin M. DNA hypomethylation of karyoplasts for bovine nuclear transplantation. Mol Reprod Dev. 2001;60(2):208–213. doi: 10.1002/mrd.1079. [DOI] [PubMed] [Google Scholar]
  • 130.Moyers SB, Kumar NB. Green tea polyphenols and cancer chemoprevention: multiple mechanisms and endpoints for phase II trials. Nutr Rev. 2004;62:204–211. doi: 10.1111/j.1753-4887.2004.tb00041.x. [DOI] [PubMed] [Google Scholar]
  • 131.Brueckner B, Boy RG, Siedlecki P, Musch T, Kliem HC, Zielenkiewicz P, Suhai S, Wiessler M, Lyko F. Epigenetic reactivation of tumor suppressor genes by a novel small-molecule inhibitor of human DNA methyltransferases. Cancer Res. 2005;65:6305–6311. doi: 10.1158/0008-5472.CAN-04-2957. [DOI] [PubMed] [Google Scholar]
  • 132.Fang MZ, Wang Y, Ai N, Hou Z, Sun Y, Lu H, Welsh W, Yang CS. Tea polyphenol (-)-epigallocatechin-3-gallate inhibits DNA methyltransferase and reactivates methylation-silenced genes in cancer cell lines. Cancer Res. 2003;63:7563–7570. [PubMed] [Google Scholar]
  • 133.Nakagawa H, Hasumi K, Woo JT, Nagai K, Wachi M. Generation of hydrogen peroxide primarily contributes to the induction of Fe(II)-dependent apoptosis in Jurkat cells by (-)-epigallocatechin gallate. Carcinogenesis. 2004;25:1567–1574. doi: 10.1093/carcin/bgh168. [DOI] [PubMed] [Google Scholar]
  • 134.Lin X, Asgari K, Putzi MJ, Gage WR, Yu X, Cornblatt BS, Kumar A, Piantadosi S, DeWeese TL, De Marzo AM, Nelson WG. Reversal of GSTP1 CpG island hypermethylation and reactivation of pi-class glutathione S-transferase (GSTP1) expression in human prostate cancer cells by treatment with procainamide. Cancer Res. 2001;61:8611–8616. [PubMed] [Google Scholar]
  • 135.Villar-Garea A, Fraga MF, Espada J, Esteller M. Procaine is a DNA-demethylating agent with growth-inhibitory effects in human cancer cells. Cancer Res. 2003;63(16):4984–4989. [PubMed] [Google Scholar]
  • 136.Nieto M, Samper E, Fraga MF, Gonzalez de Buitrago G, Esteller M, Serrano M. The absence of p53 is critical for the induction of apoptosis by 5-aza-2'-deoxycytidine. Oncogene. 2004;23(3):735–743. doi: 10.1038/sj.onc.1207175. [DOI] [PubMed] [Google Scholar]
  • 137.Cooney CA, Dave AA, Wolff GL. Maternal methyl supplements in mice affect epigenetic variation and DNA methylation of offspring. J Nutr. 2002;132(8 Suppl):2393S–2400S. doi: 10.1093/jn/132.8.2393S. [DOI] [PubMed] [Google Scholar]
  • 138.Choi SW, Mason JB. Folate status: effects on pathways of colorectal carcinogenesis. J Nutr. 2002;132(8 Suppl):2413S–2418S. doi: 10.1093/jn/132.8.2413S. [DOI] [PubMed] [Google Scholar]
  • 139.Kim YI. Folate and carcinogenesis: evidence, mechanisms, and implications. J Nutr Biochem. 1999;10:66–88. doi: 10.1016/s0955-2863(98)00074-6. [DOI] [PubMed] [Google Scholar]
  • 140.Lamprecht SA, Lipkin M. Chemoprevention of colon cancer by calcium, vitamin D and folate: molecular mechanisms. Nat Rev Cancer. 2003;3(8):601–614. doi: 10.1038/nrc1144. [DOI] [PubMed] [Google Scholar]
  • 141.Balaghi M, Wagner C. DNA methylation in folate deficiency: use of CpG methylase. Biochem Biophys Res Commun. 1993;193(3):1184–1190. doi: 10.1006/bbrc.1993.1750. [DOI] [PubMed] [Google Scholar]
  • 142.Kim YI, Pogribny IP, Basnakian AG, Miller JW, Selhub J, James SJ, Mason JB. Folate deficiency in rats induces DNA strand breaks and hypomethylation within the p53 tumor suppressor gene. Am J Clin Nutr. 1997;65:46–52. doi: 10.1093/ajcn/65.1.46. [DOI] [PubMed] [Google Scholar]
  • 143.Ingrosso D, Cimmino A, Perna AF, Masella L, De Santo NG, De Bonis ML, Vacca M, D'Esposito M, D'Urso M, Galletti P, Zappia V. Folate treatment and unbalanced methylation and changes of allelic expression induced by hyperhomocysteinaemia in patients with uraemia. Lancet. 2003;361:1693–1699. doi: 10.1016/S0140-6736(03)13372-7. [DOI] [PubMed] [Google Scholar]
  • 144.Cantoni GL, Chiang PK. The role of S-adenosylhomocysteine and S-adenosylhomocysteine hydrolase in the control of biological methylations. In: Cavallini D, Gaull GE, Zappia V, editors. Natural sulfur compounds. New York: Plenum Press; 1980. pp. 67–80. [Google Scholar]
  • 145.Kerr SJ. Competing methyltransferase systems. J Biol Chem. 1972;247(13):4248–4252. [PubMed] [Google Scholar]
  • 146.Wagner C, Briggs WT, Cook RJ. Inhibition of glycine N-methyltransferase activity by folate derivatives: implications for regulation of methyl group metabolism. Biochem Biophys Res Commun. 1985;127(3):746–752. doi: 10.1016/s0006-291x(85)80006-1. [DOI] [PubMed] [Google Scholar]
  • 147.Wagner C, Decha-Umphai W, Corbin J. Phosphorylation modulates the activity of glycine N-methyltransferase, a folate binding protein. In vitro phosphorylation is inhibited by the natural folate ligand. J Biol Chem. 1989;264(16):9638–9642. [PubMed] [Google Scholar]
  • 148.Jencks DA, Mathews RG. Allosteric inhibition of methylenetetrahydrofolate reductase by adenosylmethionine. Effects of adenosylmethionine and NADPH on the equilibrium between active and inactive forms of the enzyme and on the kinetics of approach to equilibrium. J Biol Chem. 1987;262(6):2485–2493. [PubMed] [Google Scholar]
  • 149.Kutabach C, Stokstad ELR. Feedback inhibition of methyleneterahydrofolate reductase. Biochem Biophys Acta. 1967;250:459–477. doi: 10.1016/0005-2744(71)90247-6. [DOI] [PubMed] [Google Scholar]
  • 150.McMullen MH, Rowling MJ, Ozias MK, Schalinske KL. Activation and induction of glycine N-methyltransferase by retinoids are tissue and gender-specific. Arch Biochem Biophys. 2002;401(1):73–80. doi: 10.1016/S0003-9861(02)00030-9. [DOI] [PubMed] [Google Scholar]
  • 151.Rowling MJ, McMullen MH, Schalinske KL. Vitamin A and its derivatives induce hepatic glycine N-methyltransferase and hypomethylation of DNA in rats. J Nutr. 2002;132:365–369. doi: 10.1093/jn/132.3.365. [DOI] [PubMed] [Google Scholar]
  • 152.Rowling MJ, Schalinske KL. Retinoid compounds activate and induce hepatic glycine N-methyltransferase in rats. J Nutr. 2001;131:1914–1917. doi: 10.1093/jn/131.7.1914. [DOI] [PubMed] [Google Scholar]

RESOURCES