The emerging role of DNA methylation in ... - Wiley Online Library

44 downloads 11873 Views 237KB Size Report
tural adaptation of chromosomal regions so as to register, ..... hydrofolate reductase; NSD1, nuclear receptor binding SET domain protein 1; REST, RE1-silencing ..... Fischle W. (2008) Talk is cheap – cross-talk in establishment, mainte- nance ...
Epilepsia, 53(Suppl. 9):11–20, 2012 doi: 10.1111/epi.12031

MOLECULAR PLASTICITY

The emerging role of DNA methylation in epileptogenesis Katja Kobow and Ingmar Blu¨mcke Department of Neuropathology, University Hospital Erlangen, Erlangen, Germany

function, and aging. Aberrant DNA methylation patterns, either localized to specific gene regions or scattered throughout the genome, are associated with many neurologic disorders. Herein, we discuss the emerging role of DNA methylation in epileptogenesis and the perspectives arising from epigenetic medicine as new therapeutic strategy in difficult-to-treat epilepsies. KEY WORDS: Brain, Epilepsy, Seizures, Epigenetics, Transmethylation, Epigenetic medicine.

SUMMARY DNA methylation is a covalent chromatin modification, characterized by the biochemical addition of a methyl group (-CH3) to cytosine nucleotides via a DNA methyltransferase enzyme. 5¢-Methylcytosine (5-mC), frequently called the fifth base, has been implicated in genome stability, silencing of transposable elements, and repression of gene expression. Through the latter, DNA methylation dynamics broadly influence brain development,

it also applies to numerous posttranslational histone tail modifications (e.g., acetylation, methylation, phosphorylation, ubiquitination, or sumoylation), incorporation of histone variants (e.g., macroH2A, H2A.X, H2A.Z, H3.3, and CENP-A), nucleosome and higher-order chromatin remodeling processes, nuclear localization, and action of noncoding RNAs (ncRNAs). Great advances in epigenetic research, mainly in the field of developmental biology and cancer, changed the prevalent view that chromatin was largely inert packing material for eukaryotic DNA. Instead, epigenetic signals are fundamental for a cell’s ability to ‘‘remember’’ past events, such as environmental stimuli, developmental cues, or internal events (Bonasio et al., 2010; Feil & Fraga, 2012). In other words, epigenetic mechanisms build the molecular memory of a cell, thereby affecting cellular plasticity and defining biologic phenotypes, also in the adult mammalian brain (Guo et al., 2011a; Miller-Delaney et al., 2012).

Epigenetics Somatic cells from multicellular organisms all share the same DNA. Nevertheless, we distinguish highly organized tissues composed of several hundred specific cell types, all differing in morphology and function. It requires tightly controlled gene expression patterns—at the least with respect to timing, space and quantity—to ensure celltype specific functionality. Epigenetic regulation is a key mechanism to define the versatile activity states of a gene, which certainly exceed a simple ‘‘on’’ and ‘‘off.’’ A set of chromatin-modifying actions leads to a change in genetic activity, without affecting DNA sequence itself. There is an ongoing debate of whether epigenetics summarizes only ‘‘mitotically and/or meiotically heritable changes to the chromatin template’’ and respective gene function (Russo, 1996). If heritability is a hallmark, so far only DNA methylation and the Trithorax/Polycomb system seem to be eligible. In contrast, Adrian Bird more recently suggested that epigenetic mechanisms include ‘‘any structural adaptation of chromosomal regions so as to register, signal or perpetuate altered activity states’’ (Bird, 2007). This unifying definition applies to DNA methylation, but

DNA Methylation DNA methylation is so far the best studied epigenetic mechanism and is present in all higher eukaryotes including humans (Dulac, 2010). It was long viewed as the most stable and long-lasting chromatin modification, but more recent studies have revealed the capability of DNA methylation being dynamic not only during development and cellular differentiation, but also in adult and postmitotic

Address correspondence to Ingmar Blmcke, Department of Neuropathology, University Hospital Erlangen, Schwabachanlage 6, DE 91054 Erlangen, Germany. E-mail: [email protected] Wiley Periodicals, Inc. ª 2012 International League Against Epilepsy

11

12 K. Kobow and I. Blu¨mcke cells (Reik, 2007; Wu & Zhang, 2010; Guo et al., 2011a). In mammals, methylation is confined mainly to CpG dinucleotides, but has also repeatedly been reported to occur at non-CpG sites (Clark et al., 1995; Grandjean et al., 2007; Sulewska et al., 2007; Barres et al., 2009). The distribution of DNA methylation along the genome shows enrichment at noncoding regions (e.g., centromeric heterochromatin) and interspersed highly repetitive elements (endogenous retroviruses and transposons), suggesting that host defense and genome stabilization are primary functions of methylation (Bestor, 2000). However, regions of high CG density, referred to as CG islands, are also frequently found in the 5¢ end of genes (60–70%) embedded in promoter regions and are sites of transcription initiation (Goll & Bestor, 2005). The great majority of CG islands are unmethylated at all stages of development and in all normal, nondiseased tissue types (Bird, 2002; Edwards et al., 2010). Chromatin, containing CG islands, is generally heavily acetylated, is most frequently marked by lysine 4 trimethylation of histone H3 (H3K4me3), lacks linker histone H1, and is relatively nucleosome deficient, which allows interaction of transcription factors with the promoter (Deaton & Bird, 2011). Of interest, mammalian transcription factor binding sites are more CG rich than the bulk genome and many contain CpG within their recognition sequence. DNA methylation is fundamental during embryonic development, where massive changes in genomic DNA methylation occur during gametogenesis and early embryogenesis (Reik & Walter, 2001; Reik et al., 2001; Jaenisch & Bird, 2003). Beside general stabilization of the genome DNA methylation is associated with parent-of-origin imprinting (Edwards & Ferguson-Smith, 2007), X chromosome inactivation in females (Minkovsky et al., 2012), the process of aging (Madrigano et al., 2012; Numata et al., 2012), lineage commitment, for example, during neurogenesis (Ma et al., 2010), and in neural plasticity (Levenson et al., 2006; Miller & Sweatt, 2007; Nelson et al., 2008; Feng et al., 2010; Guo et al., 2011a).

DNA Methylating Enzymes Cytosine methylation is mediated by DNA methyltransferases (DNMTs) (Bestor, 2000; Goll & Bestor, 2005). DNMT1 shows strong preference for hemimethylated DNA and is thought to be a maintenance enzyme, preserving DNA methylation marks during mitosis following semiconservative DNA replication. In contrast, DNMT3A and 3B show activity toward unmethylated DNA, and establish DNA methylation at novel sites, particularly during gametogenesis and embryogenesis. However, overlap in specificity and function between the three isoforms cannot be excluded. A recent study provided first evidence for redundant function of at least Dnmt1 and Dnmt3a using conditional mutant mice targeting Dnmt1, Epilepsia, 53(Suppl. 9):11–20, 2012 doi: 10.1111/epi.12031

Dnmt3a, or both, in postnatal postmitotic neurons (Feng et al., 2010). Furthermore, homozygous deletions of Dnmt1, Dnmt3a, or Dnmt3b in mice are not viable, underscoring the overall physiologic importance of DNA methylation (Li et al., 1992; Okano et al., 1999). Although DNA methylation can be established and maintained, both passive and active DNA demethylation may also occur (Bhutani et al., 2011). Passive DNA demethylation may result from ‘‘dilution’’ during mitosis, with the DNA methylation signature not being transmitted from mother to daughter strand. In addition, various mechanisms have been proposed for active, cell cycle–independent DNA demethylation including deamination of 5-mC to thymidine (T) followed by base excision repair (Wu & Zhang, 2010). De novo methylating enzymes Dnmt3a and 3b have been shown to possess deaminase activity in vitro and may therefore participate in both methylation and active demethylation, thereby facilitating rapid transcriptional cycling (Metivier et al., 2008). However, conditional double knockout mice lacking Dnmt1 and Dnmt3a in postmitotic neurons show active DNA demethylation (Feng et al., 2010). Other studies have suggested that growth arrest and DNA-damage-inducible 45 (Gadd45), a base excision repair protein, is involved in active DNA demethylation, because overexpression of Gadd45 fosters loci-specific and global demethylation, whereas knockdown results in DNA hypermethylation (Ma et al., 2009). Another possible DNA demethylation mechanism exclusively reported in neurons includes conversion of 5-mC to 5-hydroxy-methylcytosine (5-hmC) by tet methylcytosine dioxygenase 1 (Tet1) and subsequent replacement of 5-hmC through base excision (Wu & Zhang, 2010; Bhutani et al., 2011; Guo et al., 2011b).

Transmethylation, Metabolism, and Disease S-Adenosyl-L-methionine (SAM) is the principal biologic methyl group donor. It is synthesized in the cytosol of every cell and mainly utilized in transmethylation reactions. Thereby, its methyl group is donated to a large variety of acceptor substrates including DNA, histones, and other proteins as well as phospholipids (Fig. 1A). Given the critical role of methylation in determining various cellular processes—for example, gene expression, protein function, and membrane fluidity—it is not surprising that abnormalities in SAM metabolism have been well recognized in various diseases including neurologic and psychiatric disorders (Mulder et al., 2005; Papakostas, 2009; Popp et al., 2009). Despite the great variety of substrates (e.g., DNA, histone and nonhistone proteins, and lipids), transmethylation reactions have a common end product, namely S-adenosylhomocysteine (SAH). The metabolite SAH is also a potent competitive inhibitor of transmethylation.

13 DNA Methylation and Epileptogenesis

ƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒƒ

Both increase in SAH levels and decrease in SAM or SAH/SAM ratio interfere with effective transmethylation. It is, therefore, essential to continuously remove SAH (Lu, 2000). The reaction that converts SAH to homocysteine and adenosine is reversible and catalyzed by SAH hydrolase (SAHH; Fig. 1A). Thermodynamics favor the synthesis of SAH. Physiologically the equilibrium of this reaction is shifted toward hydrolysis by rapid and constant removal of adenosine and homocysteine. Adenosine is an endogenous inhibitory neuromodulator with anticonvulsive and neuroprotective properties. Postnatally, adenosine kinase (ADK) is the primary metabolizing enzyme of adenosine contributing to the maintenance of SAM-dependent transmethylation (Fig. 1A). Mouse models of epileptogenesis suggest initial downregulation of Adk and elevation of ambient adenosine as a protective response following initial precipitating injury. In chronic disease, astrogliosis and its associated overexpression of Adk reduce adenosine levels (Fig. 1B; Boison, 2008). Adk overexpressing transgenic mice display increased sensitivity to brain injury and seizures (Loscher & Brandt, 2010). Concordant with these findings, astrocytic ADK expression was increased also in the HC and temporal cortex of patients with temporal lobe epilepsy (Aronica et al., 2011). In the brain, homocysteine is metabolized to regenerate methionine in a reaction that is catalyzed by methionine synthase and requires normal levels of folate and vitamin

Figure 1. Transmethylation in physiologic and epileptogenic condition. (A) Summary of the transmethylation pathway and interplay with folate and adenosine metabolism under physiologic conditions. SAM is the primary methyl group donor in all transmethylation reactions irrespective of the substrate to be methylated (e.g., DNA, proteins including histones, and lipids) and is converted to SAH. SAH is further hydrolyzed to adenosine and homocysteine. Although adenosine may function as a neuromodulator or be converted to AMP, homocysteine is remethylated to methionine in a folate-dependent manner and further converted to SAM. (B) Summary of expected molecular changes in the transmethylation pathway (marked in violet) as previously reported in experimental and/or human epilepsy. Increased expression of enzymes or accumulation of metabolites is indicated by arrow up. Reduced gene expression or decrease in metabolites is indicated by arrow down. Both increased expression of de novo and maintenance Dnmts (Zhu et al., 2011) and aberrant DNA methylation have been described in hippocampi of epileptic animals and patients (Kobow et al., 2009; Miller-Delaney et al., 2012). A constant shift in the equilibrium of the SAHH reaction towards hydrolysis of SAH may result from efficient adenosine removal and has been suggested to effectively promote transmethylation (Gouder et al., 2004). Intriguingly, increased Adk expression is frequently observed in chronic epileptic tissue, accounting for continuous removal of adenosine and shifting chemical equilibrium toward AMP synthesis (Boison, 2008). Disturbances in folate metabolism (e.g., low-folate diet or MTHFR mutations/reduced enzymatic activity) have been associated with increased CSF levels in homocysteine, a proconvulsant, and seizures (Goyette et al., 1995; Ono et al., 2000; Chen et al., 2001; Baldelli et al., 2010). AMP, adenosine monophosphate; ADK, adenosine kinase; DNMTs, DNA methyltransferases; MAT, methionine adenosyltransferase; MTHFR, methylene tetrahydrofolate reductase; 5-MTHF, 5¢ methylene tetrahydrofolate; SAHH, S-adenosylhomocysteine hydrolase; THF, tetrahydrofolate. Epilepsia ILAE

B12. Homocysteine is a potent N-methyl-D-aspartate (NMDA) receptor agonist and exerts proconvulsant and neurotoxic activity via increased glutamatergic neurotransmission (Lipton et al., 1997; Mattson & Shea, 2003). Disturbances in both folate and homocysteine metabolism might severely affect transmethylation and associate with seizures (Fig. 1B). Epilepsia, 53(Suppl. 9):11–20, 2012 doi: 10.1111/epi.12031

14 K. Kobow and I. Blu¨mcke Methylenetetrahydrofolate reductase (MTHFR) is a key metabolic enzyme involved in the conversion of 5,10-methylenetetrahydrofolate (5,10-MTHF) to 5methyltetrahydrofolate (5-MTHF), the methyl donor for homocysteine remethylation to methionine. Mice heterozygous or homozygous for a null allele of Mthfr display significantly decreased SAM or increased SAH levels or both, with global DNA hypomethylation in various tissues including brain (Chen et al., 2001). Mutations in the MTHFR gene accompanied with severe MTHFR deficiency have been described in patients with homocystinuria, an inborn error of metabolism (Table 1). Intriguingly, these patients present with developmental delay, psychiatric disturbances, and seizures or abnormal electroencephalography (EEG) pattern among other features (Haan et al., 1985; Goyette et al., 1995; Tonetti et al., 2000). Of interest, there is a considerable phenotypic overlap of severe MTHFR deficiency with Angelman syndrome and Rett syndrome, both of which are autism spectrum disorders with high incidence of seizures. Furthermore, the 677C>T MTHFR variant, which causes mild reduction in MTHFR activity, is a genetic risk factor for seizures or epilepsy (Ono et al., 2000; Baldelli et al., 2010). Cerebral folate deficiency (CFD) syndrome is a pediatric neurologic disorder characterized by low cerebrospinal fluid (CSF) levels of 5-MTHF. Intriguingly, patients with CFD syndrome

present with psychomotor retardation, cerebellar ataxia, and seizures (Ramaekers et al., 2005).

DNA Methylation in Epilepsy and Other Neurologic Disorders Epigenetic mechanisms including DNA methylation regulate not only cell fate determination and maturation, but also play a role in the induction of activity-dependent synaptic plasticity, memory formation, and cognition (Levenson & Sweatt, 2005; Levenson et al., 2006; Fischer et al., 2007; Qureshi & Mehler, 2010). The importance of epigenetic mechanisms is emphasized by the growing number of neurologic diseases that associate with alterations in, for example, DNA methylation patterns, including autism spectrum disorders (James et al., 2004; Feinberg, 2007), schizophrenia (Grayson et al., 2005), Alzheimer’s disease (Scarpa et al., 2003; Graff et al., 2012), brain tumors (Hegi et al., 2005), and spinal muscular atrophy (Hauke et al., 2009). These epigenetic lesions include changes in localized or global density of DNA methylation and are likely to be acquired by environmental or intrinsic factors that promote the pathogenic condition (Fig. 2) (Feil & Fraga, 2012). Of interest, a number of autism spectrum disorders with high incidence of a seizure

Table 1. Epigenetics in human diseases with associated seizure phenotype Disorder Prader-Willi syndrome (PWS) Angelman syndrome (AS) Rett syndrome (RTT) Rubinstein-Taybi syndrome (RSTS) Sotos syndrome Progressive myoclonus epilepsy (PME) a-Thalassemia and X-linked mental retardation (ATRX) Fragile X syndrome (FXS) MTHFR deficiency

Genomic region

Gene involved

15q11-q13

snoRNAs?

15q11-q13

UBE3A

Xq27

MeCP2

16p13

CBP/p300

5q35

NSD1

12q12

RILP1

Xq13

ATR-X

Xq27

FMR1

1p36

MTHFR

Comments Deletion, uniparental disomy, imprinting defect Deletion, uniparental disomy, imprinting defect, point mutation Loss of function, duplication, epigenetic dysregulation Loss of function, duplication Loss of function, epigenetic dysregulation Loss of function, point mutation Loss of function, duplication (transgenic mice), DNA hyper-/hypomethylation Expansion of CCG, DNA hypermethylation DNA hypomethylation

Function/pathway Unknown Ubiquitin-E3-ligase, histone ubiquitination Binds methylated CpGs HAT activity Nuclear receptor, HMT activity Nuclear receptor, necessary for nuclear localization of REST Member of SWI/SNF chromatin-remodeling enzymes RNA-binding protein Folate metabolism and transmethylation

Reference Ledbetter et al. (1981), Nicholls et al. (1989) Kishino et al. (1997)

Amir et al. (1999), McGraw et al. (2011) Petrij et al. (1995) Turkmen et al. (2003) Bassuk et al. (2008)

Gibbons et al. (2000)

Santoro et al. (2012) Goyette et al. (1995)

ATRX, a-Thalassemia/mental retardation syndrome X-linked; CBP, CREB-binding protein; FMR1, fragile X mental retardation 1; HAT, histone acetyl-transferase; HMT, histone methyltransferase; MeCP2, methyl CpG-binding protein 2; MTHFR, methylenetetrahydrofolate reductase; NSD1, nuclear receptor binding SET domain protein 1; REST, RE1-silencing transcription factor; RILP1, REST-interacting LIM domain protein 1; snoRNA, small nucleolar RNA; SWI/SNF, SWItch/Sucrose NonFermentable (yeast); UBE3A, ubiquitin protein ligase E3A. Epilepsia, 53(Suppl. 9):11–20, 2012 doi: 10.1111/epi.12031

15 DNA Methylation and Epileptogenesis

Figure 2. Epigenetic drift and epileptogenesis (modified from Feil & Fraga, 2012). Integration of the progression model of epileptogenesis (initial precipitating injury, latency period, development of chronic recurrent seizures, that is, epilepsy) and epigenetic drift as hypothesized pathomechanisms in epilepsy development and chronification. Intrinsic and environmental factors contribute to epigenetic alterations over time. Initial precipitating events as much as seizures by themselves may alter chromatin function and thereby induce the epileptogenic phenotype including chronic recurrent seizures, structural lesion, drug resistance, and cognitive impairment. Both active demethylation of previously silenced gene promoters and accumulation of de novo DNA methylation at formerly unmethylated gene loci may occur and account for deregulation of multiple candidate genes. However, DNA methylation is not a binary code with a simple on/off mode. Rather subtle, accumulating changes over time may lead to different adjustments in gene expression patterns at a specific locus and contribute to the chronification of the disease state. Noteworthy, not all loci are equally prone to epigenetic drift. FS, febrile seizures; CpG, cytosine guanine dinucleotides; epigenetic drift, localized or global changes in DNA methylation and other chromatin marks over time (i.e., in postmitotic neurons and/or proliferating adult neural stem cells); epigenetic maintenance, chromatin state remains unchanged over time. Epilepsia ILAE phenotype result from chromatin effector mutations (e.g., mutation of methyl-CpG-binding protein 2, MeCP2, in Rett syndrome), DNA regulatory sequence mutations (e.g., imprinting center deletions in Angelman and PraderWilli syndrome), or triplet repeat expansion (e.g., CGG expansion in fragile X syndrome), and thus are of primary genetic origin. However, all of these primary genetic mutations are associated with secondary epimutations that severely alter neuronal gene function (Urdinguio et al., 2009; Qureshi & Mehler, 2010) (Table 1). A growing number of studies have shown the impact of aberrant epigenetic changes in experimental epilepsy including alterations in H3 and H4 histone tail modifications (Huang et al., 2002; Tsankova et al., 2004; Jia et al., 2006; Sng et al., 2006), increased phosphorylation of histone variant H2A.X (Crowe et al., 2011), increased binding of neuron restrictive silencing factor (NRSF, a major

transcriptional repressor in neuronal gene regulation that recruits DNMTs) (Garriga-Canut et al., 2006; McClelland et al., 2011), aberrant microRNA expression (Aronica et al., 2010; Jimenez-Mateos et al., 2012), and altered DNA methylation patterns (Miller-Delaney et al., 2012). Martinowich et al. were the first to describe dynamic changes in Bdnf promoter methylation patterns, concomitant reciprocal association of MeCP2 and cAMP response element binding protein (CREB) to the promoter, as well as altered Bdnf expression upon potassium chloride-driven depolarization of mouse cortical postmitotic neurons. The most remarkable finding was, however, that neuronal activity induced active demethylation of the Bdnf IV promoter, and the decrease in DNA methylation within the regulatory region of the Bdnf gene occurred within days (Martinowich et al., 2003). Two other studies demonstrated decrease of spontaneous excitatory neurotransmission and Epilepsia, 53(Suppl. 9):11–20, 2012 doi: 10.1111/epi.12031

16 K. Kobow and I. Blu¨mcke network activity following 5-Aza- or zebularine-mediated inhibition of Dnmts in both hippocampal slices from young mice (Levenson et al., 2006) and hippocampal neurons obtained from postnatal mice (Nelson et al., 2008). Our group was first to show aberrant DNA promoter methylation in human TLE-HS (Kobow et al., 2009), and Zhu et al. (2011) reported increased Dnmt gene expression in the temporal neocortex of patients with TLE. More recently, Miller-Delaney et al. (2012) described differential DNA methylation patterns in defining status epilepticus (SE) and epileptic tolerance.

The Methylation Hypothesis in Epilepsy Stimulus-induced methylation changes have been identified in postnatal brain, supporting the notion that active DNA methylation modifications contribute to the molecular memory of postmitotic neurons. This may occur in physiologic conditions, but may as well contribute to the pathogenesis and disease progression as recently suggested for epilepsy, a major neurologic disease affecting about 1–2% of the world’s population (Kobow & Blumcke, 2011). Chronic, intractable TLEs are of particular interest due to their progressive etiopathology and frequently evolving drug-resistance (Loscher et al., 2011). The genesis and development of TLE primarily relate to the hippocampus. Surgical resection has proved successful and beneficial to control seizures (Wiebe et al., 2001). Novel strategies targeting the molecular pathomechanism will be mandatory to improve treatment. MTLE-HS as well as other epilepsies are often characterized by an initial precipitating injury, may it be trauma, inflammation, or febrile convulsions, followed by a clinically silent latency phase and later development of spontaneous recurrent seizures (Blumcke et al., 2002; Pitkanen & Lukasiuk, 2011). With growing evidence for aberrant epigenetic chromatin modifications in experimental and human TLE, we hypothesized that initial precipitating injuries as much as seizures by themselves are a potent inducer of epigenetic alterations and thereby aggravate the epileptogenic condition resulting in structural brain lesion, drug resistance, and cognitive dysfunction (Kobow & Blumcke, 2011). It is the dynamics of epigenetic mechanisms that provide a likely explanation for some features of complex diseases, for example, epilepsy, including late onset, parent-of-origin effects, discordance of monozygotic twins, and fluctuation of symptoms (Petronis, 2001). Indeed, TLE is the most common epilepsy syndrome in adults, and disease progression with development of drug resistance is very frequent in these patients. Surprisingly, cure for some TLE patients has been described after treatment with a high fat, low carbohydrate ketogenic diet, leading to the question ‘‘what happened to the diseaserelated candidate genes’’? (Martinez et al., 2007). There is Epilepsia, 53(Suppl. 9):11–20, 2012 doi: 10.1111/epi.12031

also a puzzling interrelation between hormones and epilepsy in humans with hormones affecting seizure susceptibility and frequency (Veliskova & Desantis, 2012). Epidemiologic studies provide evidence for an increased risk of seizures in the offspring of mothers with epilepsy (Ottman et al., 1988). Discordance for monozygotic twins reaches about 30% in idiopathic epilepsies (Briellmann et al., 2001; Petronis, 2001). Progressive gain of DNA methylation has been associated with aging and age-related complex diseases. In contrast, inadequate epigenetic maintenance and loss of DNA methylation may also contribute to human pathologies, including neurodegenerative and psychiatric diseases (Pogribny & Beland, 2009). However, we need to keep in mind that DNA methylation is not required for gene silencing in the first instance, but may contribute to the stabilization, or locking, of an inactive gene state and thus serve as a reinforcing signal for preexisting, less-stable epigenetic signatures (Bonasio et al., 2010). It has been suggested that transiently expressed or activated factors in response to environmental stimuli or developmental cues establish epigenetic chromatin states. The transient effects mediated by transcription factors (TFs) may be maintained even when the TF itself disappears. The Polycomb and Trithorax groups of proteins are likely to be involved in the maintenance of epigenetic states, as many members of both groups possess chromatin-modifying activities (e.g., acetyltransferase, deacetylase, or methyltransferase activity). Chromatin-modifying enzymes may be recruited to specific genomic targets by physical interaction, either with TFs or with ncRNAs. The latter are heavily transcribed from noncoding intergenic and intragenic regions, frequently possess regulatory functions, and are well suited to bridge chromatin modifiers with the genome in a sequence specific fashion. Given the fact that most of our genome is noncoding in terms of not protein-coding, the number, functional variability, and regulatory impact of ncRNAs is probably underestimated (Mehler & Mattick, 2007). Irrespective of the initial events in the establishment of an epigenetic state, the activation or silencing of a gene locus can be reinforced by the following: (1) feedback loops, for example, where enzymes that modify histones specifically interact with proteins that bind to these histone marks; and (2) cross-talk between different epigenetic marks and mechanisms, for example, histone modifications and DNA methylation (Fischle, 2008). There is evidence that de novo DNA methyltransferases and associated factors preferentially bind to unmethylated H3K4, whereas di-/trimethylation at lysine 9 as well as trimethylation at lysine 27 seem to trigger DNA methylation (Goll & Bestor, 2005; Vire et al., 2006; Ooi et al., 2007). This interplay suggests that, when present, DNA methylation may serve as a reinforcing signal for preexisting, less stable epigenetic signatures such as histone

17 DNA Methylation and Epileptogenesis modifications (for detailed review please see Bonasio et al., 2010). By targeting transcriptionally silent chromatin, DNA methylation contributes to the stability of gene silencing (Bird, 2002) and in the context of experimental and human TLE is thought to represent a late stage in the process of misregulation of multiple genes accounting for the molecular and cellular changes associated with HS pathology.

Epigenetic Medicine Growing evidence for epigenetic pathomechanisms opens new avenues also for treatment. The reversible nature of epigenetic marks highlights DNA methyltransferases (DNMTs), histone acetyltransferases, deacetylases (HATs and HDACs), and histone methyltransferases and demethylases (HKTs and HDMTs) as interesting targets for novel drugs (Kelly et al., 2010) (Table 2). Two DNMT inhibitors (DNMTi) were already approved by the U.S. Food and Drug Administration (FDA) for the treatment of malignancies, i.e., 5-azacytidine (Vidaza, Celgene, Summit, NJ, U.S.A.; approved 2004) and 5-aza-2¢-deoxycytidine (Dacogen, Astex Pharmaceuticals, Dublin, CA, U.S.A.; approved 2006). Both are cytosine analogs that incorporate into DNA during the S phase of the cell cycle, and thus primarily target rapidly proliferating cells and induce global demethylation. Intriguingly, both inhibitors are covalently bound to DNMTs, targeting these enzymes for proteasomal degradation and thereby reducing DNMT levels (Eglen & Reisine, 2011). Despite their proven efficacy in some types of cancer, DNMTi are yet limited by either target-related or side-effect toxicity. Zebularine is another promising nucleoside analog mimicking reactive intermediates of DNA methylation (Champion et al., 2010). It is the first orally active DNMTi, more stable than the parent compound (i.e., 5-aza-2¢-deoxycytidine), and with minimal cytotoxicity

both in vivo and in vitro. Unfortunately, zebularine was also reported to have poor bioavailability due to rapid metabolization (Klecker et al., 2006). To avoid toxicity and stability problems associated with nucleoside inhibitors, recent efforts have focused on the development of nonnucleoside DNMTi. These compounds inhibit DNA methylation by binding directly to DNMTs without being incorporated into DNA and, thus, show higher specificity. RG108 is a small compound that selectively blocks the enzymatic activity of DNMT1 (Brueckner et al., 2005). Of interest, repositioned drugs approved for other indications such as the antihypertensive drug hydralazine, or the local anesthetic procaine, or even the antiarrhythmic drug procainamide, have also been identified as nonnucleoside DNMTi (SeguraPacheco et al., 2003; Candelaria et al., 2007; Castellano et al., 2008). These compounds may either target DNMTs directly or inhibit signaling pathways associated with DNA methylation. Another novel approach addresses the physiologic RNA interference (RNAi) pathway to silence DNMTs. MG98 is an antisense oligonucleotide specifically binding to the 3¢ untranslated region (3¢UTR) of DNMT1 mRNA and targeting mRNA degradation. A clinical phase I trial in patients with tumor has proved its safety and tolerability as well as early evidence of antitumor activity (Plummer et al., 2009). Physiologic microRNAs targeting DNMT gene expression may also represent interesting targets for pharmacologic intervention, for example, miR-29 and miR-143, but have not yet been addressed in preclinical or clinical trials.

Summary We hypothesize that aberrant DNA promoter methylation in concert with other epigenetic chromatin modifications provides a key mechanism for the misregulation of multiple genes all being affected in a certain neuropathologic

Table 2. Epigenetic compounds targeting DNA methylation DNMT inhibitor 5-Azacytidine (Vidaza) 5-Aza-20-deoxycytidine (Dacogen, Decitabine) Zebularine RG108 Hydralazine Procainamide Procaine MG98 miR-143, miR-29

Specificity

Developmental stage

Ribonucleoside analog, competitive inhibitor for all DNMTs, incorporates into DNA and RNA, inhibits DNA methylation and may as well inhibit RNA translation Deoxyribonucleoside analog, competitive inhibitor for all DNMTs, incorporates into DNA but not RNA, inhibits DNA methylation Deoxyribonucleoside analog Nonnucleoside small molecule, L-tryptophan derivative, competitive inhibitor for DNMT1 Antihypertensive drug, specifically inhibits DNMT1 activity, also potent mitogen-activated protein kinase inhibitor, targets DNMT gene expression Antiarrhythmic drug, specifically inhibits DNMT1 activity Local anesthetic, possible inhibition of signaling pathways associated with DNA methylation Antisense oligonucleotide directed at the 3¢ UTR of DNMT1 mRNA, selectively reduces DNMT1 gene expression Physiologic microRNAs, miR-143 targets DNMT3A; miR-29 targets DNMT3A and 3B

Approved 2004 Approved 2006 Phase I Preclinical Phase II Preclinical Preclinical Phase I

DNMT, DNA methyltransferase; miR, microRNA; 3¢ UTR, 3¢ untranslated region. Epilepsia, 53(Suppl. 9):11–20, 2012 doi: 10.1111/epi.12031

18 K. Kobow and I. Blu¨mcke condition, that is, mesial TLE with HS, and their collective contribution to the associated symptomatology. The ‘‘methylation hypothesis’’ is compatible with current pathogenic concepts of mesial TLE with HS, where spontaneous recurrent seizures usually develop after an early initial precipitating injury (status epilepticus, brain trauma) followed by a clinically silent latency period, in which many structural and molecular reorganization processes establish (Blumcke et al., 2002). The presumable intimate relationship between neuronal hyperactivity and functional gene regulation by means of promoter methylation and other epigenetic mechanisms needs further analysis, as understanding of these mechanisms may open new therapeutic perspectives in difficult-to-treat epilepsies.

Acknowledgments Our work is supported by the German research council (DFG) within the EpiGENet programme of the newly established European EuroEPINOMICS initiative.

Disclosure The authors declare no conflicts of interest. We confirm that we have read the Journal’s position on issues involved in ethical publication and affirm that this report is consistent with those guidelines.

References Amir RE, Van den Veyver IB, Wan M, Tran CQ, Francke U, Zoghbi HY. (1999) Rett syndrome is caused by mutations in X-linked MECP2, encoding methyl-CpG-binding protein 2. Nat Genet 23:185–188. Aronica E, Fluiter K, Iyer A, Zurolo E, Vreijling J, van Vliet EA, Baayen JC, Gorter JA. (2010) Expression pattern of miR-146a, an inflammation-associated microRNA, in experimental and human temporal lobe epilepsy. Eur J Neurosci 31:1100–1107. Aronica E, Zurolo E, Iyer A, de Groot M, Anink J, Carbonell C, van Vliet EA, Baayen JC, Boison D, Gorter JA. (2011) Upregulation of adenosine kinase in astrocytes in experimental and human temporal lobe epilepsy. Epilepsia 52:1645–1655. Baldelli E, Leo G, Andreoli N, Fuxe K, Biagini G, Agnati LF. (2010) Homocysteine potentiates seizures and cell loss induced by pilocarpine treatment. Neuromolecular Med 12:248–259. Barres R, Osler ME, Yan J, Rune A, Fritz T, Caidahl K, Krook A, Zierath JR. (2009) Non-CpG methylation of the PGC-1alpha promoter through DNMT3B controls mitochondrial density. Cell Metab 10:189–198. Bassuk AG, Wallace RH, Buhr A, Buller AR, Afawi Z, Shimojo M, Miyata S, Chen S, Gonzalez-Alegre P, Griesbach HL, Wu S, Nashelsky M, Vladar EK, Antic D, Ferguson PJ, Cirak S, Voit T, Scott MP, Axelrod JD, Gurnett C, Daoud AS, Kivity S, Neufeld MY, Mazarib A, Straussberg R, Walid S, Korczyn AD, Slusarski DC, Berkovic SF, El-Shanti HI. (2008) A homozygous mutation in human PRICKLE1 causes an autosomal-recessive progressive myoclonus epilepsyataxia syndrome. Am J Hum Genet 83:572–581. Bestor TH. (2000) The DNA methyltransferases of mammals. Hum Mol Genet 9:2395–2402. Bhutani N, Burns DM, Blau HM. (2011) DNA demethylation dynamics. Cell 146:866–872. Bird A. (2002) DNA methylation patterns and epigenetic memory. Genes Dev 16:6–21. Bird A. (2007) Perceptions of epigenetics. Nature 447:396–398. Blumcke I, Thom M, Wiestler OD. (2002) Ammon’s horn sclerosis: a maldevelopmental disorder associated with temporal lobe epilepsy. Brain Pathol 12:199–211. Epilepsia, 53(Suppl. 9):11–20, 2012 doi: 10.1111/epi.12031

Boison D. (2008) The adenosine kinase hypothesis of epileptogenesis. Prog Neurobiol 84:249–262. Bonasio R, Tu S, Reinberg D. (2010) Molecular signals of epigenetic states. Science 330:612–616. Briellmann RS, Jackson GD, Torn-Broers Y, Berkovic SF. (2001) Causes of epilepsies: insights from discordant monozygous twins. Ann Neurol 49:45–52. Brueckner B, Garcia Boy R, Siedlecki P, Musch T, Kliem HC, Zielenkiewicz P, Suhai S, Wiessler M, Lyko F. (2005) Epigenetic reactivation of tumor suppressor genes by a novel small-molecule inhibitor of human DNA methyltransferases. Cancer Res 65:6305–6311. Candelaria M, Gallardo-Rincon D, Arce C, Cetina L, Aguilar-Ponce JL, Arrieta O, Gonzalez-Fierro A, Chavez-Blanco A, de la CruzHernandez E, Camargo MF, Trejo-Becerril C, Perez-Cardenas E, Perez-Plasencia C, Taja-Chayeb L, Wegman-Ostrosky T, RevillaVazquez A, Duenas-Gonzalez A. (2007) A phase II study of epigenetic therapy with hydralazine and magnesium valproate to overcome chemotherapy resistance in refractory solid tumors. Ann Oncol 18: 1529–1538. Castellano S, Kuck D, Sala M, Novellino E, Lyko F, Sbardella G. (2008) Constrained analogues of procaine as novel small molecule inhibitors of DNA methyltransferase-1. J Med Chem 51:2321–2325. Champion C, Guianvarc’h D, Senamaud-Beaufort C, Jurkowska RZ, Jeltsch A, Ponger L, Arimondo PB, Guieysse-Peugeot AL. (2010) Mechanistic insights on the inhibition of c5 DNA methyltransferases by zebularine. PLoS ONE 5:e12388. Chen Z, Karaplis AC, Ackerman SL, Pogribny IP, Melnyk S, LussierCacan S, Chen MF, Pai A, John SW, Smith RS, Bottiglieri T, Bagley P, Selhub J, Rudnicki MA, James SJ, Rozen R. (2001) Mice deficient in methylenetetrahydrofolate reductase exhibit hyperhomocysteinemia and decreased methylation capacity, with neuropathology and aortic lipid deposition. Hum Mol Genet 10:433–443. Clark SJ, Harrison J, Frommer M. (1995) CpNpG methylation in mammalian cells. Nat Genet 10:20–27. Crowe SL, Tsukerman S, Gale K, Jorgensen TJ, Kondratyev AD. (2011) Phosphorylation of histone H2A.X as an early marker of neuronal endangerment following seizures in the adult rat brain. J Neurosci 31:7648–7656. Deaton AM, Bird A. (2011) CpG islands and the regulation of transcription. Genes Dev 25:1010–1022. Dulac C. (2010) Brain function and chromatin plasticity. Nature 465:728–735. Edwards CA, Ferguson-Smith AC. (2007) Mechanisms regulating imprinted genes in clusters. Curr Opin Cell Biol 19:281–289. Edwards JR, O’Donnell AH, Rollins RA, Peckham HE, Lee C, Milekic MH, Chanrion B, Fu Y, Su T, Hibshoosh H, Gingrich JA, Haghighi F, Nutter R, Bestor TH. (2010) Chromatin and sequence features that define the fine and gross structure of genomic methylation patterns. Genome Res 20:972–980. Eglen RM, Reisine T. (2011) Screening for compounds that modulate epigenetic regulation of the transcriptome: an overview. J Biomol Screen 16:1137–1152. Feil R, Fraga MF. (2012) Epigenetics and the environment: emerging patterns and implications. Nat Rev Genet 13:97–109. Feinberg AP. (2007) Phenotypic plasticity and the epigenetics of human disease. Nature 447:433–440. Feng J, Zhou Y, Campbell SL, Le T, Li E, Sweatt JD, Silva AJ, Fan G. (2010) Dnmt1 and Dnmt3a maintain DNA methylation and regulate synaptic function in adult forebrain neurons. Nat Neurosci 13:423– 430. Fischer A, Sananbenesi F, Wang X, Dobbin M, Tsai LH. (2007) Recovery of learning and memory is associated with chromatin remodelling. Nature 447:178–182. Fischle W. (2008) Talk is cheap – cross-talk in establishment, maintenance, and readout of chromatin modifications. Genes Dev 22:3375– 3382. Garriga-Canut M, Schoenike B, Qazi R, Bergendahl K, Daley TJ, Pfender RM, Morrison JF, Ockuly J, Stafstrom C, Sutula T, Roopra A. (2006) 2-Deoxy-D-glucose reduces epilepsy progression by NRSFCtBP-dependent metabolic regulation of chromatin structure. Nat Neurosci 9:1382–1387. Gibbons RJ, McDowell TL, Raman S, O’Rourke DM, Garrick D, Ayyub H, Higgs DR. (2000) Mutations in ATRX, encoding a SWI/SNF-like

19 DNA Methylation and Epileptogenesis protein, cause diverse changes in the pattern of DNA methylation. Nat Genet 24:368–371. Goll MG, Bestor TH. (2005) Eukaryotic cytosine methyltransferases. Annu Rev Biochem 74:481–514. Gouder N, Scheurer L, Fritschy JM, Boison D. (2004) Overexpression of adenosine kinase in epileptic hippocampus contributes to epileptogenesis. J Neurosci 24:692–701. Goyette P, Frosst P, Rosenblatt DS, Rozen R. (1995) Seven novel mutations in the methylenetetrahydrofolate reductase gene and genotype/ phenotype correlations in severe methylenetetrahydrofolate reductase deficiency. Am J Hum Genet 56:1052–1059. Graff J, Rei D, Guan JS, Wang WY, Seo J, Hennig KM, Nieland TJ, Fass DM, Kao PF, Kahn M, Su SC, Samiei A, Joseph N, Haggarty SJ, Delalle I, Tsai LH. (2012) An epigenetic blockade of cognitive functions in the neurodegenerating brain. Nature 483:222–226. Grandjean V, Yaman R, Cuzin F, Rassoulzadegan M. (2007) Inheritance of an epigenetic mark: the CpG DNA methyltransferase 1 is required for de novo establishment of a complex pattern of non-CpG methylation. PLoS ONE 2:e1136. Grayson DR, Jia X, Chen Y, Sharma RP, Mitchell CP, Guidotti A, Costa E. (2005) Reelin promoter hypermethylation in schizophrenia. Proc Natl Acad Sci U S A 102:9341–9346. Guo JU, Ma DK, Mo H, Ball MP, Jang MH, Bonaguidi MA, Balazer JA, Eaves HL, Xie B, Ford E, Zhang K, Ming GL, Gao Y, Song H. (2011a) Neuronal activity modifies the DNA methylation landscape in the adult brain. Nat Neurosci 14:1345–1351. Guo JU, Su Y, Zhong C, Ming GL, Song H. (2011b) Hydroxylation of 5-methylcytosine by TET1 promotes active DNA demethylation in the adult brain. Cell 145:423–434. Haan EA, Rogers JG, Lewis GP, Rowe PB. (1985) 5,10-Methylenetetrahydrofolate reductase deficiency. Clinical and biochemical features of a further case. J Inherit Metab Dis 8:53–57. Hauke J, Riessland M, Lunke S, Eyupoglu IY, Blumcke I, El-Osta A, Wirth B, Hahnen E. (2009) Survival motor neuron gene 2 silencing by DNA methylation correlates with spinal muscular atrophy disease severity and can be bypassed by histone deacetylase inhibition. Hum Mol Genet 18:304–317. Hegi ME, Diserens AC, Gorlia T, Hamou MF, de Tribolet N, Weller M, Kros JM, Hainfellner JA, Mason W, Mariani L, Bromberg JE, Hau P, Mirimanoff RO, Cairncross JG, Janzer RC, Stupp R. (2005) MGMT gene silencing and benefit from temozolomide in glioblastoma. N Engl J Med 352:997–1003. Huang Y, Doherty JJ, Dingledine R. (2002) Altered histone acetylation at glutamate receptor 2 and brain-derived neurotrophic factor genes is an early event triggered by status epilepticus. J Neurosci 22:8422– 8428. Jaenisch R, Bird A. (2003) Epigenetic regulation of gene expression: how the genome integrates intrinsic and environmental signals. Nat Genet 33(Suppl.):245–254. James SJ, Cutler P, Melnyk S, Jernigan S, Janak L, Gaylor DW, Neubrander JA. (2004) Metabolic biomarkers of increased oxidative stress and impaired methylation capacity in children with autism. Am J Clin Nutr 80:1611–1617. Jia YH, Zhu X, Li SY, Ni JH, Jia HT. (2006) Kainate exposure suppresses activation of GluR2 subunit promoter in primary cultured cerebral cortical neurons through induction of RE1-silencing transcription factor. Neurosci Lett 403:103–108. Jimenez-Mateos EM, Engel T, Merino-Serrais P, McKiernan RC, Tanaka K, Mouri G, Sano T, O’Tuathaigh C, Waddington JL, Prenter S, Delanty N, Farrell MA, O’Brien DF, Conroy RM, Stallings RL, Defelipe J, Henshall DC. (2012) Silencing microRNA-134 produces neuroprotective and prolonged seizure-suppressive effects. Nat Med 18:1087–1094. Kelly TK, De Carvalho DD, Jones PA. (2010) Epigenetic modifications as therapeutic targets. Nat Biotechnol 28:1069–1078. Kishino T, Lalande M, Wagstaff J. (1997) UBE3A/E6-AP mutations cause Angelman syndrome. Nat Genet 15:70–73. Klecker RW, Cysyk RL, Collins JM. (2006) Zebularine metabolism by aldehyde oxidase in hepatic cytosol from humans, monkeys, dogs, rats, and mice: influence of sex and inhibitors. Bioorg Med Chem 14:62–66. Kobow K, Blumcke I. (2011) The methylation hypothesis: do epigenetic chromatin modifications play a role in epileptogenesis? Epilepsia 52(Suppl. 4):15–19.

Kobow K, Jeske I, Hildebrandt M, Hauke J, Hahnen E, Buslei R, Buchfelder M, Weigel D, Stefan H, Kasper B, Pauli E, Blumcke I. (2009) Increased reelin promoter methylation is associated with granule cell dispersion in human temporal lobe epilepsy. J Neuropathol Exp Neurol 68:356–364. Ledbetter DH, Riccardi VM, Airhart SD, Strobel RJ, Keenan BS, Crawford JD. (1981) Deletions of chromosome 15 as a cause of the Prader–Willi syndrome. N Engl J Med 304:325–329. Levenson JM, Sweatt JD. (2005) Epigenetic mechanisms in memory formation. Nat Rev Neurosci 6:108–118. Levenson JM, Roth TL, Lubin FD, Miller CA, Huang IC, Desai P, Malone LM, Sweatt JD. (2006) Evidence that DNA (cytosine-5) methyltransferase regulates synaptic plasticity in the hippocampus. J Biol Chem 281:15763–15773. Li E, Bestor TH, Jaenisch R. (1992) Targeted mutation of the DNA methyltransferase gene results in embryonic lethality. Cell 69:915–926. Lipton SA, Kim WK, Choi YB, Kumar S, D’Emilia DM, Rayudu PV, Arnelle DR, Stamler JS. (1997) Neurotoxicity associated with dual actions of homocysteine at the N-methyl-D-aspartate receptor. Proc Natl Acad Sci U S A 94:5923–5928. Loscher W, Brandt C. (2010) Prevention or modification of epileptogenesis after brain insults: experimental approaches and translational research. Pharmacol Rev 62:668–700. Loscher W, Luna-Tortos C, Romermann K, Fedrowitz M. (2011) Do ATP-binding cassette transporters cause pharmacoresistance in epilepsy? Problems and approaches in determining which antiepileptic drugs are affected. Curr Pharm Des 17:2808–2828. Lu SC. (2000) S-Adenosylmethionine. Int J Biochem Cell Biol 32: 391–395. Ma DK, Guo JU, Ming GL, Song H. (2009) DNA excision repair proteins and Gadd45 as molecular players for active DNA demethylation. Cell Cycle 8:1526–1531. Ma DK, Marchetto MC, Guo JU, Ming GL, Gage FH, Song H. (2010) Epigenetic choreographers of neurogenesis in the adult mammalian brain. Nat Neurosci 13:1338–1344. Madrigano J, Baccarelli A, Mittleman MA, Sparrow D, Vokonas PS, Tarantini L, Schwartz J. (2012) Aging and epigenetics: longitudinal changes in gene-specific DNA methylation. Epigenetics 7, 63–70. Martinez CC, Pyzik PL, Kossoff EH. (2007) Discontinuing the ketogenic diet in seizure-free children: recurrence and risk factors. Epilepsia 48:187–190. Martinowich K, Hattori D, Wu H, Fouse S, He F, Hu Y, Fan G, Sun YE. (2003) DNA methylation-related chromatin remodeling in activitydependent BDNF gene regulation. Science 302:890–893. Mattson MP, Shea TB. (2003) Folate and homocysteine metabolism in neural plasticity and neurodegenerative disorders. Trends Neurosci 26:137–146. McClelland S, Flynn C, Dub C, Richichi C, Zha Q, Ghestem A, Esclapez M, Bernard C, Baram TZ. (2011) Neuron-restrictive silencer factor-mediated hyperpolarization-activated cyclic nucleotide gated channelopathy in experimental temporal lobe epilepsy. Ann Neurol 70(3):454–464. McGraw CM, Samaco RC, Zoghbi HY. (2011) Adult neural function requires MeCP2. Science 333:186. Mehler MF, Mattick JS. (2007) Noncoding RNAs and RNA editing in brain development, functional diversification, and neurological disease. Physiol Rev 87:799–823. Metivier R, Gallais R, Tiffoche C, Le Peron C, Jurkowska RZ, Carmouche RP, Ibberson D, Barath P, Demay F, Reid G, Benes V, Jeltsch A, Gannon F, Salbert G. (2008) Cyclical DNA methylation of a transcriptionally active promoter. Nature 452:45–50. Miller CA, Sweatt JD. (2007) Covalent modification of DNA regulates memory formation. Neuron 53:857–869. Miller-Delaney SF, Das S, Sano T, Jimenez-Mateos EM, Bryan K, Buckley PG, Stallings RL, Henshall DC. (2012) Differential DNA methylation patterns define status epilepticus and epileptic tolerance. J Neurosci 32:1577–1588. Minkovsky A, Patel S, Plath K. (2012) Concise review: pluripotency and the transcriptional inactivation of the female Mammalian X chromosome. Stem Cells 30:48–54. Mulder C, Schoonenboom NS, Jansen EE, Verhoeven NM, van Kamp GJ, Jakobs C, Scheltens P. (2005) The transmethylation cycle in the brain of Alzheimer patients. Neurosci Lett 386:69–71. Epilepsia, 53(Suppl. 9):11–20, 2012 doi: 10.1111/epi.12031

20 K. Kobow and I. Blu¨mcke Nelson ED, Kavalali ET, Monteggia LM. (2008) Activity-dependent suppression of miniature neurotransmission through the regulation of DNA methylation. J Neurosci 28:395–406. Nicholls RD, Knoll JH, Butler MG, Karam S, Lalande M. (1989) Genetic imprinting suggested by maternal heterodisomy in nondeletion Prader–Willi syndrome. Nature 342:281–285. Numata S, Ye T, Hyde TM, Guitart-Navarro X, Tao R, Wininger M, Colantuoni C, Weinberger DR, Kleinman JE, Lipska BK. (2012) DNA methylation signatures in development and aging of the human prefrontal cortex. Am J Hum Genet 90:260–272. Okano M, Bell DW, Haber DA, Li E. (1999) DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell 99:247–257. Ono H, Sakamoto A, Mizoguchi N, Sakura N. (2000) Methylenetetrahydrofolate reductase 677C > T mutation and epilepsy. J Inherit Metab Dis 23:525–526. Ooi SK, Qiu C, Bernstein E, Li K, Jia D, Yang Z, Erdjument-Bromage H, Tempst P, Lin SP, Allis CD, Cheng X, Bestor TH. (2007) DNMT3L connects unmethylated lysine 4 of histone H3 to de novo methylation of DNA. Nature 448:714–717. Ottman R, Annegers JF, Hauser WA, Kurland LT. (1988) Higher risk of seizures in offspring of mothers than of fathers with epilepsy. Am J Hum Genet 43:257–264. Papakostas GI. (2009) Evidence for S-adenosyl-L-methionine (SAM-e) for the treatment of major depressive disorder. J Clin Psychiatry 70(Suppl. 5):18–22. Petrij F, Giles RH, Dauwerse HG, Saris JJ, Hennekam RC, Masuno M, Tommerup N, van Ommen GJ, Goodman RH, Peters DJ, Breuning MH. (1995) Rubinstein-Taybi syndrome caused by mutations in the transcriptional co-activator CBP. Nature 376:348–351. Petronis A. (2001) Human morbid genetics revisited: relevance of epigenetics. Trends Genet 17:142–146. Pitkanen A, Lukasiuk K. (2011) Mechanisms of epileptogenesis and potential treatment targets. Lancet Neurol 10:173–186. Plummer R, Vidal L, Griffin M, Lesley M, de Bono J, Coulthard S, Sludden J, Siu LL, Chen EX, Oza AM, Reid GK, McLeod AR, Besterman JM, Lee C, Judson I, Calvert H, Boddy AV. (2009) Phase I study of MG98, an oligonucleotide antisense inhibitor of human DNA methyltransferase 1, given as a 7-day infusion in patients with advanced solid tumors. Clin Cancer Res 15:3177–3183. Pogribny IP, Beland FA. (2009) DNA hypomethylation in the origin and pathogenesis of human diseases. Cell Mol Life Sci 66:2249–2261. Popp J, Lewczuk P, Linnebank M, Cvetanovska G, Smulders Y, Kolsch H, Frommann I, Kornhuber J, Maier W, Jessen F. (2009) Homocysteine metabolism and cerebrospinal fluid markers for Alzheimer’s disease. J Alzheimers Dis 18:819–828. Qureshi IA, Mehler MF. (2010) Epigenetic mechanisms underlying human epileptic disorders and the process of epileptogenesis. Neurobiol Dis 39:53–60. Ramaekers VT, Rothenberg SP, Sequeira JM, Opladen T, Blau N, Quadros EV, Selhub J. (2005) Autoantibodies to folate receptors in the cerebral folate deficiency syndrome. N Engl J Med 352:1985– 1991. Reik W. (2007) Stability and flexibility of epigenetic gene regulation in mammalian development. Nature 447:425–432.

Epilepsia, 53(Suppl. 9):11–20, 2012 doi: 10.1111/epi.12031

Reik W, Walter J. (2001) Genomic imprinting: parental influence on the genome. Nat Rev Genet 2:21–32. Reik W, Dean W, Walter J. (2001) Epigenetic reprogramming in mammalian development. Science 293:1089–1093. Russo VEA, Martienssen RA, Riggs AD (eds) (1996) Introduction. In Russo VEA (Ed) Epigenetic mechanisms of gene regulation, Cold Spring Harbor Monograph, Vol. 32. Cold Spring Harbor Laboratory Press, Woodbury, pp. 1–4. Santoro MR, Bray SM, Warren ST. (2012) Molecular mechanisms of fragile X syndrome: a twenty-year perspective. Annu Rev Pathol 7:219–245. Scarpa S, Fuso A, D’Anselmi F, Cavallaro RA. (2003) Presenilin 1 gene silencing by S-adenosylmethionine: a treatment for Alzheimer disease? FEBS Lett 541:145–148. Segura-Pacheco B, Trejo-Becerril C, Perez-Cardenas E, Taja-Chayeb L, Mariscal I, Chavez A, Acuna C, Salazar AM, Lizano M, DuenasGonzalez A. (2003) Reactivation of tumor suppressor genes by the cardiovascular drugs hydralazine and procainamide and their potential use in cancer therapy. Clin Cancer Res 9:1596–1603. Sng JC, Taniura H, Yoneda Y. (2006) Histone modifications in kainateinduced status epilepticus. Eur J Neurosci 23:1269–1282. Sulewska A, Niklinska W, Kozlowski M, Minarowski L, Naumnik W, Niklinski J, Dabrowska K, Chyczewski L. (2007) DNA methylation in states of cell physiology and pathology. Folia Histochem Cytobiol 45:149–158. Tonetti C, Burtscher A, Bories D, Tulliez M, Zittoun J. (2000) Methylenetetrahydrofolate reductase deficiency in four siblings: a clinical, biochemical, and molecular study of the family. Am J Med Genet 91:363–367. Tsankova NM, Kumar A, Nestler EJ. (2004) Histone modifications at gene promoter regions in rat hippocampus after acute and chronic electroconvulsive seizures. J Neurosci 24:5603–5610. Turkmen S, Gillessen-Kaesbach G, Meinecke P, Albrecht B, Neumann LM, Hesse V, Palanduz S, Balg S, Majewski F, Fuchs S, Zschieschang P, Greiwe M, Mennicke K, Kreuz FR, Dehmel HJ, Rodeck B, Kunze J, Tinschert S, Mundlos S, Horn D. (2003) Mutations in NSD1 are responsible for Sotos syndrome, but are not a frequent finding in other overgrowth phenotypes. Eur J Hum Genet 11:858–865. Urdinguio RG, Sanchez-Mut JV, Esteller M. (2009) Epigenetic mechanisms in neurological diseases: genes, syndromes, and therapies. Lancet Neurol 8:1056–1072. Veliskova J, Desantis KA. (2012) Sex and hormonal influences on seizures and epilepsy. Horm Behav Apr 4 [Epub ahead of print]. 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. (2006) The Polycomb group protein EZH2 directly controls DNA methylation. Nature 439:871–874. Wiebe S, Blume WT, Girvin JP, Eliasziw M. (2001) A randomized, controlled trial of surgery for temporal-lobe epilepsy. N Engl J Med 345:311–318. Wu SC, Zhang Y. (2010) Active DNA demethylation: many roads lead to Rome. Nat Rev Mol Cell Biol 11:607–620. Zhu Q, Wang L, Zhang Y, Zhao FH, Luo J, Xiao Z, Chen GJ, Wang XF. (2011) Increased Expression of DNA methyltransferase 1 and 3a in Human Temporal Lobe Epilepsy. J Mol Neurosci 46:420–426.