Abstract
The observation that Alzheimer’s disease (AD) patients with similar and even identical genetic backgrounds often present with heterogeneous pathologies has prompted the hypothesis that epigenetics may contribute to AD. While the study of epigenetics encompasses a variety of modifications including histone modifications and non-coding RNAs, much of the research on how epigenetics might impact AD pathology has been focused on DNA methylation. To this end, several studies have characterized DNA methylation alterations in various brain regions of individuals with AD, with conflicting results. This review examines the results of studies analyzing both global and gene-specific DNA methylation changes in AD and also assesses the results of studies analyzing DNA hydroxymethylation in patients with AD.
Keywords: DNA methylation, Alzheimer’s disease, epigenetics, dementia, DNA hydroxymethylation
Introduction to dementia and Alzheimer’s disease
Dementia, an age-related neurodegenerative disorder characterized by progressive cognitive decline, affects ∼35.6 million people worldwide and is becoming an increasingly relevant concern as the population ages [1, 2]. It is estimated that the prevalence of dementia is expected to double every 20 years such that 115.4 million individuals will have dementia by 2050 [2]. In terms of its economic burden, the total estimated cost of dementia worldwide was $604 billion U.S. dollars in 2010 [3]; in the United States alone, the annual societal cost of dementia was between $159 billion and $215 billion [1]. Moreover, predictive modeling suggests that there will be a cost increase of almost 80% per adult by 2040 [1]. In light of its substantial societal and economic burden, there has been a strong push within the research community to better understand the etiology of dementia.
Dementia is a broad term encompassing a wide span of neurodegenerative disorders including Alzheimer’s disease (AD), vascular dementia and dementia with Lewy bodies [4]. AD is the most common type of dementia, accounting for 50–80% of all dementia cases [5–7]. Currently, AD affects ∼5.4 million Americans, yet this number is expected to increase to 13.8 million by 2050 as the population ages [5]. As of 2013, AD was the sixth leading cause of death in the United States behind heart disease, cancer, lower respiratory disease, unintentional injury and stroke [8]. Strikingly, between 2000 and 2013, deaths from stroke and heart disease decreased by 23 and 14%, respectively, whereas deaths from AD increased by 71%, highlighting the necessity to better understand the etiology and pathogenesis of AD [5]. Because AD is the most common type of dementia, this review will focus specifically on AD.
AD is characterized by cerebral extracellular amyloid-β (Aβ) plaques and intracellular neurofibrillary tangles (NFTs) [9]. The sequential cleavage of amyloid precursor protein (APP) by the membrane-bound proteins β-secretase and γ-secretase generates amyloidogenic Aβ peptides, which accumulate within the extracellular space to form insoluble Aβ plaques [10, 11]. In addition, APP also can be cleaved by α-secretase and then again by γ-secretase, resulting in the production of soluble, nonamyloidogenic Aβ peptides. Besides Aβ plaques, intracellular NFTs are the other primary pathological hallmark of AD. These tangles are created by hyperphosphorylation of microtubule-associated protein tau (MAPT), and the number of NFTs in the neocortex is positively correlated with the severity of dementia [12]. The presence and severity of Aβ plaques and NFTs within the brain are often used for postmortem neuropathological assessment of individuals diagnosed with AD [13].
AD is classified as early-onset AD (EOAD) or late-onset AD (LOAD) [14, 15]. EOAD occurs in individuals younger than age 65 but may manifest as early as age 40 or 50 [16]. In addition, if EOAD coincides with genetic mutations in APP, presenilin-1 (PSEN1) or presenilin-2 (PSEN2), the latter two of which code for proteins that make up the main catalytic component of γ-secretase, the patient is diagnosed with familial AD. In this case, multiple generations within a family can be affected. EOAD is relatively rare and accounts for <10% of all AD cases [17]. LOAD, on the other hand, which occurs in individuals older than age 65, is more common yet has no clear genetic association or cause [18]. Despite its substantial socioeconomic burden, the etiology of LOAD has not been elucidated in part due to its complex pathogenesis and presentation.
Introduction to epigenetics
Recent research suggests that epigenetic factors often have an impact on chronic diseases including cancer and AD [19–22]. Although there is no single, universally accepted definition of epigenetics, many definitions touch upon the following components [23–25]: (i) heritability, defined as the ability of a dividing cell to pass epigenetic marks to its daughter cells; (ii) preservation of DNA sequence, in which the 5′ to 3′ order of nucleotides remains unchanged; (iii) transcriptional regulation, whereby epigenetic changes influence the transcription of genes and therefore the phenotype of the cell; and (iv) stability, defined as the maintenance of epigenetic modifications over time [26].
The number of studies examining epigenetic mechanisms has risen dramatically as big-data genomics has gained popularity. With this increased focus, the question of the stability of epigenetic modifications has become a highly debated topic. It was long thought that epigenetic marks were stable throughout a lifetime [27]. However, it is now known that epigenetic modifications are dynamic, and as a result, research has begun to focus on how environmental exposures can alter the epigenetic landscape [24, 28–31]. Although epigenetics is a broad term encompassing DNA methylation, histone modifications and non-coding RNA, this review will focus specifically on DNA methylation in AD; histone modifications and non-coding RNA have been reviewed in detail elsewhere [32–36].
DNA methylation
DNA methyltransferase (DNMT) enzymes, of which DNMT1, DNMT3A and DNMT3B are best characterized [37, 38], catalyze the transfer of a methyl group from S-adenosylmethionine to DNA. The product of this reaction, S-adenosylhomocysteine, is then converted back to S-adenosylmethionine through a series of reactions as part of the one-carbon metabolism cycle [39]. Importantly, an elevation in plasma homocysteine, an intermediate in this cycle, is associated with a greater risk for developing dementia and AD [40–42], thus implicating DNA methylation and alterations in one-carbon metabolism with AD pathogenesis [43–47].
In mammals, DNA methylation primarily occurs on cytosine residues of cytosine-guanine (CpG) dinucleotides and functions to modulate gene expression [22, 48]. Canonically, DNA methylation is associated with reduced gene expression [49], yet recent evidence suggests that the impact of DNA methylation on gene expression is dependent upon its context within the genome [50]. Though 5-methylcytosine (5mC) is the most abundant modified base in the mammalian genome, recent studies have identified additional modified bases such as N6-methyladenine (N6mA) and 5-formylcytosine (5fC). N6mA, though prevalent in prokaryotes, was thought to be absent in eukaryotes. In 2015, researchers identified N6mA in three eukaryotic species: green algae [51] (Chlamydomonas reinhardtii), nematodes [52] (Caenorhabditis elegans) and fruit flies [53] (Drosophila melanogaster), and in 2016, researchers characterized the existence of N6mA in mouse embryonic stem cells [54]. The results of these studies show that there is large variation in both the prevalence of this modification and its effect on gene expression across eukaryotic species. 5fC was originally thought to be an intermediate in enzyme-mediated DNA demethylation. Recent studies suggest, however, that 5fC can be a stable DNA modification that affects gene expression by altering the structure of the DNA double helix [55–57]. Despite the progress gained by these pioneering studies, knowledge of these modified bases is still incomplete, and more research is needed to further characterize their distribution and function.
For the purposes of this review, the term “global” DNA methylation describes the average percent methylation across the entire genome. The term “gene-specific” DNA methylation, on the other hand, refers to the average percent methylation within a specific gene. Global DNA methylation, although useful in that it provides an over-arching picture of methylation status in a sample, is sometimes misleading as the proportion of genes with significant alterations in DNA methylation to genes with insignificant DNA methylation differences is generally very small [58, 59]. Gene-specific DNA methylation, however, is able to detect these hidden significant DNA methylation differences. In studies on DNA methylation in AD, global DNA methylation is often assessed via antibody-based methods such as immunohistochemistry, while gene-specific DNA methylation is analyzed using array-based methods such as Illumina’s Infinium HumanMethylation450 BeadChip array. High-throughput techniques utilizing bisulfite conversion are used to measure both global and gene-specific DNA methylation.
DNA hydroxymethylation
The recent surge of interest in DNA hydroxymethylation began with the publication of two papers in 2009 characterizing this modification in separate systems: cerebellar purkinje neurons and mouse embryonic stem cells [60, 61]. In contrast with the relatively constant tissue distribution of 5mC, there is substantial variation in the tissue distribution of 5-hydroxymethylcytosine (5hmC). It is most abundant within the central nervous system, with the highest percentage of 5hmC found within the cerebral cortex, followed closely by the brainstem, spinal cord and cerebellum [62]. In addition to the central nervous system, 5hmC is also present at lower levels in the heart, kidney, liver, muscle and lung [62]. Ten-eleven translocation enzymes oxidize the 5-methyl group on methylated cytosines to a 5-hydroxymethyl group [63]. Despite a growing number of studies on the subject, the function of 5hmC is currently unknown. It is hypothesized to either be an intermediate in active DNA demethylation or be involved in gene regulation. More specifically, 5hmC within gene bodies is positively correlated with gene expression. Although it is still unclear what the mechanism behind this correlation is, it is thought to be due to the association of 5hmC with histone marks H3K4me1 and H3K27ac, both of which are markers of active promoters and enhancers [64–66].
RNA modifications
Of the various RNA-related chemical modifications found in eukaryotic cells, most have been identified in rRNAs and tRNAs. They have been shown to play a role in altering the structure and function of mature RNAs to influence gene expression [67, 68]. Reversible mRNA modifications, on the other hand, have not been seriously studied until recently [69]. N6-methyladenosine (m6A), the most abundant of these modifications, was the first to be discovered [70–72]. Initial studies in mammalian cells using RNA immunoprecipitation followed by high-throughput sequencing demonstrate that m6A is highly enriched around stop codons and within long internal exons and may function to affect gene expression [73, 74]. Since the discovery of m6A, other dynamic RNA modifications such as pseudouridine and N1-methyladenosine have been identified, though more research is necessary to determine the mechanisms by which these modifications influence gene expression [75–79]. The information garnered thus far on RNA modifications is summarized in numerous review articles [67, 68].
Importantly, epigenetic mechanisms do not work in isolation [80, 81]. Rather, they often work together to bring about changes in gene expression and therefore phenotype. For example, methyl-CpG-binding domain proteins (MBD2) bind to methylated cytosines and recruit methyl-CpG binding protein 1 (MeCP1) complexes. Among other proteins, the MeCP1 complex includes a nucleosome remodeling protein with a histone deacetylase core [82]. Together, these proteins act to silence that particular region of the genome. Recent studies suggest that there are various types of epigenetic alterations associated with the pathophysiology of AD, including alterations in DNA methylation [17, 83, 84]. This review will focus on evidence concerning the association between DNA methylation alterations and AD.
Global DNA methylation alterations in AD
Due to the heterogeneous clinical presentation of AD in patients with very similar and sometimes identical genetic backgrounds, it has been hypothesized that epigenetics may contribute to the pathology of AD. The study of identical twins has classically been used to separate the effect of genetics and environment. In 2009, Mastroeni et al. [85] analyzed DNA methylation in a pair of monozygotic twins discordant for AD. Using immunohistochemical methods to detect 5mC, the authors found that global DNA methylation within the anterior temporal neocortex and the superior frontal gyrus was significantly decreased in the twin with AD compared with the neurologically normal, non-demented twin. Interestingly, although both twins were chemical engineers with similar levels of education, the AD twin worked extensively with pesticides whereas the non-demented twin did not, suggesting that work-related pesticide exposure could have contributed to the development of AD. Thus, the results of this study indicate that not only genetic mutations but also environmental influences can affect AD presentation.
Although the study of identical twins is often preferred to separate genetic and environmental factors, it is extremely rare. More commonly, research on DNA methylation alterations in AD involves unrelated participants discordant for AD. Both prior to and since Mastroeni’s study in 2009, other groups have analyzed global DNA methylation in various brain regions of unrelated individuals discordant for AD.
Temporal cortex
As a follow-up to their twin study, Mastroeni et al. [82] used immunohistochemistry to study DNA methylation in the entorhinal cortex of a sample population with and without AD. Like the conclusions drawn from their previous study, the authors found that there was a significant decrease in global DNA methylation in individuals with AD compared to individuals without AD. Importantly, immunoreactivity for 5mC was not significantly different in relatively AD-spared regions such as the cerebellum, suggesting that DNA methylation alterations in AD are region-specific. Although the results of this study in combination with the results from the 2009 twin study might lead readers to surmise that individuals with AD have lower levels of DNA methylation in the temporal cortex than individuals without AD, results from other studies using similar antibody-based methods do not support this conclusion. In contrast, Coppieters et al. [17] showed that there was an increase in the global DNA methylation of brain tissue samples derived from the middle temporal gyrus of subjects with AD compared with age-matched, cognitively normal controls. Lashley et al. [86] found that there was no significant difference in global DNA methylation in the entorhinal cortex of individuals with and without AD [86].
Phipps et al. [87], recognizing that perhaps DNA methylation could be cell-type specific, used immunohistochemistry to analyze 5mC and 5hmC in neuronal and glial cell types in the inferior temporal gyrus of human AD cases and age-matched controls. Results of their study suggest that extranuclear 5mC in neurofilament-labeled pyramidal neurons, which are particularly vulnerable to AD pathology [88–90], is significantly decreased in AD cases compared with controls. This same trend was seen in astrocytes. Interestingly, AD-spared calretinin interneurons and microglia did not have significant alterations in 5mC or 5hmC, which supports the hypothesis that DNA methylation alterations in AD are dependent upon the particular cell type studied.
Frontal cortex
The frontal cortex is a site of significant synapse loss in AD [91, 92], and as such, it is often analyzed in AD studies. As mentioned previously, Mastroeni’s twin study found decrements in DNA methylation in the frontal cortex of the twin with AD compared with the cognitively-normal twin [85]. Since then, however, other studies using similar antibody-based methods have found that individuals with AD tend to have a higher level of DNA methylation than individuals without AD. Coppieters et al. [93], for example, found that DNA methylation was higher in the middle frontal gyrus of individuals with AD than in the same region of age-matched controls. Similarly, Rao et al. [94] also found increased global DNA methylation in the frontal cortex of AD patients compared with cognitively-normal controls.
Hippocampus
One of the hallmark pathologies of AD is cerebral atrophy, and atrophy of the hippocampus, a region involved in memory formation, is particularly severe. Thus, researchers have long been analyzing the hippocampus as a brain region significantly impacted by AD. Like the results of studies on DNA methylation within the temporal cortex and frontal cortex, the results of studies analyzing DNA methylation alterations within the hippocampus are inconclusive. A study by Chouliaras et al. [95] found a significant decrease in hippocampal DNA methylation in AD cases compared to unrelated, age-matched controls. In addition, the authors found that glial cell DNA methylation was significantly different in the CA1 and CA3 subregions, whereas neuronal DNA methylation was significantly different only in the CA1 subregion, suggesting that cell type specific alterations in DNA methylation vary depending on hippocampal subregion. The same year that Chouliaras et al. published these results, Bradley-Whitman et al. [96] also published a study analyzing DNA methylation within the hippocampus. Interestingly, they found the opposite: individuals with AD had higher levels of DNA methylation in the hippocampus than individuals without AD.
Gene-specific DNA methylation alterations in AD
To date, there have been several studies analyzing gene-specific DNA methylation in brain tissue of patients with and without AD [86, 94, 97–100]. Much like the results from studies analyzing global DNA methylation, these studies have been inconclusive. However, a number of common DNA methylation alterations in specific genes have been observed across studies, providing relatively strong evidence that methylation within these genes may be altered in AD [98, 101, 102].
Genes related to AD pathology
Initially, studies of gene-specific DNA methylation differences in AD focused on genes associated with AD pathology such as APP, PSEN1, MAPT and apolipoprotein E (APOE). Barrachina et al. [103] analyzed DNA methylation of APP, PSEN1 and MAPT in the frontal cortex and hippocampus of non-demented control individuals and individuals in various stages of AD and found no significant differences in the DNA methylation in any of the three genes studied. Iwata et al. [104] used pyrosequencing to analyze DNA methylation of various CpG sites of AD-related genes in the inferior temporal lobe, the superior parietal lobe and the cerebellum in AD subjects and non-demented control subjects. They found significant differences in DNA methylation profiles of APP, MAPT and GSK3B, but not of PSEN1, BACE1 or APOE. Moreover, the authors demonstrated that these alterations in DNA methylation translated into changes in gene expression, which provides a potential mechanism by which DNA methylation can impact the AD phenotype in these subjects.
APOE is an apolipoprotein that associates with lipoproteins in the plasma as a component of systemic lipid metabolism. In addition, APOE is widely expressed in the central nervous system and functions as the primary cholesterol carrier necessary for the maintenance, growth and repair of neurons [105]. Importantly, APOE is reduced in the hippocampus of patients with AD and is thought to enhance proteolytic degradation of amyloidogenic Aβ [105–107]. Moreover, a specific allele of APOE, APOE-ε4, is significantly less effective at degrading Aβ than the other alleles [106]. APOE-ε4 is a risk factor for AD [108]; individuals homozygous for the APOE-ε4 allele are eight times as likely to develop AD as individuals without the APOE-ε4 allele [108]. Using MALDI-TOF mass spectrometry in postmortem brain tissue samples derived from the prefrontal cortex of individuals with AD and matched controls, Wang et al. [109] found that the promoter region of APOE, but not APP, was hypermethylated individuals with AD.
In addition to the genes classically associated with AD, numerous genome-wide association studies have identified genetic variants associated with increased AD susceptibility, including but not limited to sortilin-related receptor, low-density lipoprotein receptor class A repeat-containing protein (SORL1), ATP-binding cassette, sub-family A, member 7 (ABCA7), and bridging integrator 1 (BIN1) [110–113]. Yu et al. [113] analyzed DNA methylation in the dorsolateral prefrontal cortex of subjects with and without AD at 28 gene loci associated with AD pathology. Of the 28 gene loci, DNA methylation of five genes (SORL1, ABCA7, HLA-DRB5, SLC2A4 and BIN1) was significantly associated with pathological AD. SORL1 regulates recycling of APP back into the cell, and in its absence, APP is instead directed towards endosomal secretase cleavage pathways that generate Aβ [114, 115]. ABCA7 is primarily known for its role in lipid efflux from cells to lipoprotein particles, but it has also been shown to regulate APP processing, secretion and clearance [116–118]. Although there is relatively little known about the role of BIN1 in AD pathogenesis, it is hypothesized to primarily affect tau pathology, APP endocytosis and intracellular trafficking, neuroinflammation and calcium transients [119]. Results from studies such as those mentioned here suggest that alterations in DNA methylation of AD-related genes are significantly associated with AD pathology.
Genes related to neuroplasticity, memory formation and neuroinflammation
Multiple genes involved in neuroplasticity and memory formation are differentially methylated between individuals with AD and individuals without AD. Sanchez-Mut et al. [84] used a customized Illumina VeraCode GoldenGate DNA Methylation array to study DNA methylation of genes related to sensory perception, cognition, neuroplasticity, brain physiology and neurological diseases in the frontal cortex of two transgenic mouse models of AD (APP/PSEN1 and 3xTg-AD). The authors validated the results of the array by pyrosequencing and found significant DNA hypermethylation in various genes, including thromboxane A2 receptor (Tbxa2r) and sorbin and SH3 domain containing 3 (Sorbs3). Tbxa2r encodes a G-protein coupled receptor that modulates cAMP response-element binding protein (CREB), which in turn is a transcription factor involved in neuronal plasticity, long-term memory formation and neuroprotection [120]. The protein product of Sorbs3 is a scaffold protein that, among other functions, modulates cell survival by regulating epidermis growth factor-induced activation of c-jun N-terminal kinase (JNK) [121, 122]. The results of Sanchez-Mut’s study were similar to those of an earlier study that found hypermethylation within SORBS3 in the temporal cortex of AD patients [31].
Mendioroz et al. [123] analyzed gene expression and methylation of CREB-regulated transcription factor 1 (CRTC1), a coactivator of the CREB-dependent gene transcription pathway involved in synaptic plasticity and long-term memory formation that is highly expressed in hippocampal neurons. CRTC1 was significantly hypomethylated in the hippocampus of individuals with AD compared with controls, and furthermore, methylation within CRTC1 was inversely correlated with Aβ and phosphorylated tau expression. Also, the expression of CRTC1 mRNA was significantly lower in AD cases compared with controls, and downstream target genes of CRTC1 were down regulated, demonstrating that DNA hypomethylation within CRTC1 translated functionally into decreased transcriptional expression and downstream effects.
Neuroinflammation is commonly associated with AD [124–126]. There is evidence that expression levels of IL-1β and IL-6 peak in the early stages of AD but return to more normal levels in the later stages [127]. Since both IL-1β and IL-6 are modulated by DNA methylation in various chronic diseases [128, 129], Nicolia et al. [130] sought to determine whether or not DNA methylation in the promoter region of these genes contributes to the fluctuation in expression levels seen across AD stages. After analyzing DNA methylation in the frontal cortex of patients with AD and middle-aged controls using techniques based on bisulfite modification, the authors found evidence of hypomethylation in the IL-1β promoter in the early stages of AD that returned to middle-aged control levels in later stages. In contrast, they found that IL-6 methylation decreased with AD progression.
DNA methylation in peripheral blood
Aberrant methylation in the peripheral blood is correlated with AD disease status [131, 132], so some argue that peripheral blood is a good indicator for AD-associated methylation changes when brain tissue is not available. However, evidence concerning DNA methylation differences in peripheral blood is inconclusive [133, 134]. As an example, brain-derived neurotrophic factor (BDNF), a protein involved in neuronal survival, differentiation and plasticity, is often reduced in AD and is commonly analyzed in AD studies [135–137]. Carboni et al. [138] found no significant differences in DNA methylation in the promoter region of BDNF in peripheral blood. On the other hand, Chang et al. [139] showed an elevation of promoter BDNF DNA methylation in peripheral blood samples of AD patients compared with gender- and age-matched controls. Similarly, Nagata et al. [140] found that DNA methylation of the BDNF promoter in peripheral blood mononuclear cells of patients with AD was significantly higher than that of age-matched cognitively normal controls. Importantly, percent methylation of certain CpG sites within the BDNF promotor correlated negatively and significantly with neuropsychological test scores, suggesting that BDNF promoter methylation is associated with clinical manifestations of AD.
DNA hydroxymethylation in AD
Alteration of global 5hmC has been shown in a variety of neurological disorders such as Rett syndrome, autism spectrum disorder and Huntington’s disease [141–143]. It is not surprising, then, that differences in 5hmC occur in AD as well, although the differences appear to be region-specific. For instance, Condliffe et al. [144] showed that there is a significant decrease in 5hmC in the entorhinal cortex and cerebellum of individuals with AD compared with their age-matched controls. Bradley-Whitman et al. [96], who studied hydroxymethylation within the hippocampus and cerebellum, demonstrated that there is a significant increase in 5hmC in the hippocampus, but not in the cerebellum. In contrast to these results, Chouliaras et al. [95] used immunohistochemistry to show that there was a 20.2% reduction in 5hmC immunoreactivity in the hippocampus of AD patients when compared with non-demented, age-matched controls. Similar results were obtained when analyzing monozygotic twins discordant for AD; 5hmC immunoreactivity was 31.4% lower in the CA1 region of the hippocampus of the AD twin compared with the non-demented twin. Coppieters et al. [93] found that there was a significant increase in 5hmC in the mid-frontal gyrus and mid-temporal gyrus of patients with AD and further showed that 5hmC was relatively low in astrocytes and microglia, but high in neurons.
Evidence suggests that neurogenesis, which is associated with alterations in hydroxymethylation, is altered in AD [145]. The subventricular zone (SVZ) and the subgranular zone are the primary sites of neurogenesis in the aging brain [146]. It has been hypothesized that neurogenesis decreases in the early stages of AD but increases during later stages as part of a compensatory mechanism [145, 147, 148]. Mastroeni et al. [149] found an increase of 5hmC in the SVZ of AD patients compared with non-diseased controls in vivo (brain tissue) and in vitro (primary SVZ cultures) using antibody-based methodologies. They concluded that the compensatory increase in cell proliferation in AD is linked to an increase in 5hmC.
Due to technological advancements in the area of 5hmC high-throughput sequencing, studies have recently begun to analyze gene-specific alterations in 5hmC in AD patients. Bernstein et al. [150] analyzed genome-wide 5hmC in the prefrontal cortex of AD cases and then correlated this with transcriptional changes using RNA-sequencing. They identified 325 genes containing differentially hydroxymethylated loci, of which 140 displayed concomitant changes in gene expression.
The type of starting material used may impact the distribution of 5hmC. A recent study by Shu et al. [65] using adult neural stem cells shows that treatment with Aβ decreased global 5hmC, suggesting that Aβ itself may have a direct impact on methylation. Moreover, when comparing 5hmC levels in HEK293ft cells, adult neural stem cells and neuronal tissue, they demonstrated that 5hmC is significantly higher in tissues than in cells. This suggests that the type of starting material itself has an impact on DNA hydroxymethylation. In addition, the authors characterized changes in DNA hydroxymethylation associated with the interaction between aging and AD by studying 5hmC levels in various brain regions of wildtype mice and a transgenic AD mouse model at 12 weeks of age and at 67 weeks of age. Although there were no significant differences in DNA hydroxymethylation in either the cortex or cerebellum with aging or with AD, there was a significant difference between transgenic mice and wildtype mice in the hippocampus at 67 weeks. It would appear, then, that brain region is important when analyzing DNA hydroxymethylation levels. Finally, the authors characterized the location of differentially methylated 5hmC regions within the genome. These regions were primarily located in gene bodies associated with genes involved in neuronal development and function. More studies like the one described above will further support a complete understanding of 5hmC distribution and characterization.
Conclusions
Despite the surge of interest in DNA methylation and AD, evidence is still inconclusive as to whether or not DNA methylation or DNA hydroxymethylation is altered, particularly on a global scale. Conflicting results have been reported for both types of modifications. Additional studies measuring DNA methylation and DNA hydroxymethylation both accurately and reproducibly are needed before any conclusions can be reached. Numerous limitations exist that impede progress in studying DNA methylation in AD. For example, most studies on this topic predominately use immunohistochemical approaches and/or microarray-based methods, such as the Illumina Infinium arrays, on small sample sizes. Genomic coverage in these arrays is extremely limited; Illumina’s Infinium HumanMethylation450 BeadChip array, for example, interrogates only 1.5% of total CpG sites in the human genome. Moreover, CpG sites included in the array are located primarily in promoter regions, so relevant DNA methylation in gene bodies is often missed. Whole-genome bisulfite sequencing (WGBS), on the other hand, is able to analyze the entirety of a genome at single-nucleotide resolution and is thus considered the ideal methodology for DNA methylation studies. WGBS introduces a whole host of other challenges, however. Currently, WGBS is generally too expensive to run the large-scale studies needed to account for interindividual variation seen in DNA methylation in AD. Furthermore, there is no standardized method for analyzing WGBS data. For these reasons, studying DNA methylation in AD has been difficult, with studies often giving conflicting results.
Inconsistencies in the conclusions of studies reported here could be due to any number of reasons. For instance, most studies on global methylation in AD are done in humans, which is important for three reasons. First, it is extremely difficult to collect the brain from a human subject exactly at the time of death, and there is concern that the sample degrades during the time between death and sample collection (postmortem interval). Studies have shown, however, that DNA methylation is well preserved in the postmortem brain for up to 48 h or even 72 h [103, 151]. As the postmortem intervals recorded in most studies are well below 48 h, sample degradation due to a long delay between death and tissue processing does not account for the variable results described here. Second, there is a naturally large interindividual variation across human subjects, which can veil anything but the most significant of differences. To further complicate the matter, most human subject studies concerning DNA methylation in AD use a relatively small sample size. The inherently large interindividual variation across humans coupled with the use of small sample sizes may bias the results of studies. Third, AD is a complex, multifactorial disease that often presents with heterogeneous pathologies. Hence, DNA methylation in AD may depend largely on AD stage and pathological presentation. Thus, conclusive results from human subject studies may be difficult to acquire, particularly in studies with small sample sizes.
In addition, the inconsistencies seen when analyzing DNA methylation in AD could be due to a failure to recognize and appreciate the complexity of the brain as an organ; the brain is comprised of multiple regions and subregions, each with its own function and unique DNA methylation profile [152]. In AD, pathologies are often dependent upon brain region (e.g. the cerebellum is relatively spared in AD pathogenesis). Thus, it would be tempting to conclude that variation seen across studies could be due to a lack of consistency in DNA methylation alterations in AD across brain regions. As demonstrated in the studies presented in this review, however, there are differences even between studies analyzing the same brain region. Importantly, not only does the brain consist of numerous distinct regions, each of these regions is also composed of numerous different cell types. Cells are specialized with their own unique gene expression and methylation signatures. Therefore, it is possible that significant differences between methylation profiles could be lost when cell types are analyzed together.
Finally, methodological differences could contribute to the inconsistencies seen across studies. Earlier studies on DNA methylation in AD use immunohistochemistry, a semi-quantitative method that does not provide an accurate, quantitative measurement of methylation or hydroxymethylation. In addition, immunohistochemistry relies heavily on the particular antibody probe being used; concerns about antibody specificity must always be taken into account when comparing studies using this method. More recent studies use methods based on bisulfite conversion, in which unmethylated cytosines are converted to uracils. Pyrosequencing and WGBS, both of which are based on the process of bisulfite conversion, are considered gold standards for quantifying DNA methylation due to their ability to measure DNA methylation at single-nucleotide resolution. Despite this advantage, however, bisulfite conversion does not distinguish between 5mC and 5hmC. This is a significant problem for studying DNA methylation in AD, as 5hmC is particularly abundant in the brain. In addition, because these bisulfite-based methods are relatively new, standardized and user-friendly software for analyzing bisulfite sequencing data has only recently been developed. Thus, post-collection analysis of data is different across studies, which could contribute to the inconsistencies reported here. For more information regarding the advantages and disadvantages of DNA methylation analysis techniques, the reader is referred to excellent reviews by Laird [153] and Kurdyukov et al. [154].
It is apparent that more research is needed to provide conclusive results. As the cost of WGBS decreases and the bioinformatics software associated with WGBS analysis becomes more user-friendly and widely available, it should become easier to carry out large-scale studies on DNA methylation in the brain and its association with AD. With further advancements in cell-sorting and DNA methylation technology using small quantities of sample, we should begin to see an increase in studies looking at DNA methylation in specific cell types. In addition, studies should focus on the interaction between the methylome, transcriptome and the plethora of histone marks and non-coding RNA in AD, as epigenetic modifications often work together to achieve a particular phenotype [155]. Finally, the field should consider: (i) using a combination of standard techniques such as immunohistochemistry coupled with bisulfite sequencing and (ii) comparing the results of genome-wide studies to focus on common genes to improve comparability across studies. In essence, although results are inconclusive, it is evident that the environment, through its effect on DNA methylation, has an impact on AD pathology. Thus, studies of epigenetic alterations in AD will continue to be a topic of immense interest within the research field.
Funding
This study was supported by the Richard A. and Nora Eccles Harrison Endowed Chair in Diabetes Research Fund (to J.C.R.) and the National Institutes of Health through the following grants: the National Institute on Aging Grant AG039094 (to J.C.R.) and the National Institute of Diabetes and Digestive and Kidney Diseases Grants U24 DK092993-05S1 (to J.C.R.), U24 DK092993 (to Kent Lloyd), R01DK104770-01A1 (to V.M.) and R03DK099427-02 (to V.M.). J.C.R. was partially funded through Grant U24 DK092993, which was awarded to Kent Lloyd. Kent Lloyd did not contribute to this manuscript.
Conflict of interest statement. None declared.
References
- 1. Hurd MD, Martorell P, Delavande A, Mullen KJ, Langa Km.. Monetary costs of dementia in the United States. N Engl J Med 2013;368: 1326–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2. Prince M, Bryce R, Albanese E, Wimo A, Ribeiro W, Ferri CP.. The global prevalence of dementia: a systematic review and metaanalysis. Alzheimers Dement 2013;9: 63–75.e62. [DOI] [PubMed] [Google Scholar]
- 3. Wimo A, Jönsson L, Bond J, Prince M, Winblad B; Alzheimer Disease International. The worldwide economic impact of dementia 2010. Alzheimer's Dementia 2013;9: 1–11.e13. [DOI] [PubMed] [Google Scholar]
- 4. Morris J. Classification of dementia and Alzheimer's disease. Acta Neurologica Scandinavica 1996;94: 41–50. [DOI] [PubMed] [Google Scholar]
- 5. Association AS. Alzheimer's disease facts and figures. Alzheimer's Dementia 2016;12: 459–509. [DOI] [PubMed] [Google Scholar]
- 6. Barnes DE, Yaffe K.. The projected effect of risk factor reduction on Alzheimer's disease prevalence. Lancet Neurol 2011;10: 819–28. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7. Vos SJ, Verhey F, Frölich L, Kornhuber J, Wiltfang J, Maier W, Peters O, Rüther E, Nobili F, Morbelli S.. Prevalence and prognosis of Alzheimer’s disease at the mild cognitive impairment stage. Brain 2015;awv029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8. Murphy S, Xu J, Kochanek K, Bastian B.. Deaths: final data for 2013. Natl Vital Stat Rep 2016;64: 1–119. [PubMed] [Google Scholar]
- 9. Esch FS, Keim PS, Beattie EC, Blacher RW, Culwell AR, Oltersdorf T, McClure D, Ward PJ.. Cleavage of amyloid beta peptide during constitutive processing of its precursor. Science 1990;248: 1122–4. [DOI] [PubMed] [Google Scholar]
- 10. Cordy JM, Hooper NM, Turner AJ.. The involvement of lipid rafts in Alzheimer's disease (review). Mol Membrane Biol 2006;23: 111–22. [DOI] [PubMed] [Google Scholar]
- 11. Ehehalt R, Keller P, Haass C, Thiele C, Simons K.. Amyloidogenic processing of the Alzheimer beta-amyloid precursor protein depends on lipid rafts. J Cell Biol 2003;160: 113–123. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Arriagada PV, Growdon JH, Hedley-Whyte ET, Hyman BT.. Neurofibrillary tangles but not senile plaques parallel duration and severity of Alzheimer's disease. Neurology 1992;42: 631. [DOI] [PubMed] [Google Scholar]
- 13. Khachaturian ZS. Diagnosis of Alzheimer’s disease. Arch Neurol 1985;42: 1097–105. [DOI] [PubMed] [Google Scholar]
- 14. Lippa CF, Saunders AM, Smith TW, Swearer JM, Drachman DA, Ghetti B, Nee L, Pulaski-Salo D, Dickson D, Robitaille Y.. Familial and sporadic Alzheimer's disease neuropathology cannot exclude a final common pathway. Neurology 1996;46: 406–12. [DOI] [PubMed] [Google Scholar]
- 15. Rossor M, Iversen L, Reynolds G, Mountjoy C, Roth M.. Neurochemical characteristics of early and late onset types of Alzheimer's disease. Br Med J 1984;288: 961–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Rossor MN, Fox NC, Mummery CJ, Schott JM, Warren JD.. The diagnosis of young-onset dementia. Lancet Neurol 2010;9: 793–806. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Coppieters N, Dragunow M.. Epigenetics in Alzheimer's disease: a focus on DNA modifications. Current Pharmaceutical Design 2011;17:3398–3412. [DOI] [PubMed] [Google Scholar]
- 18. Wang W-X, Rajeev BW, Stromberg AJ, Ren N, Tang G, Huang Q, Rigoutsos I, Nelson PT.. The expression of microRNA miR-107 decreases early in Alzheimer's disease and may accelerate disease progression through regulation of β-site amyloid precursor protein-cleaving enzyme 1. J Neurosci 2008;28: 1213–23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Egger G, Liang G, Aparicio A, Jones PA.. Epigenetics in human disease and prospects for epigenetic therapy. Nature 2004;429: 457–63. [DOI] [PubMed] [Google Scholar]
- 20. Portela A, Esteller M.. Epigenetic modifications and human disease. Nat Biotechnol 2010;28: 1057–68. [DOI] [PubMed] [Google Scholar]
- 21. Feinberg AP. Phenotypic plasticity and the epigenetics of human disease. Nature 2007;447: 433–40. [DOI] [PubMed] [Google Scholar]
- 22. Mastroeni D, Grover A, Delvaux E, Whiteside C, Coleman PD, Rogers J.. Epigenetic mechanisms in Alzheimer's disease. Neurobiol Aging 2011;32: 1161–80. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Bird A. Perceptions of epigenetics. Nature 2007;447: 396–8. [DOI] [PubMed] [Google Scholar]
- 24. Richards EJ. Inherited epigenetic variation—revisiting soft inheritance. Nat Rev Genet 2006;7: 395–401. [DOI] [PubMed] [Google Scholar]
- 25. Goldberg AD, Allis CD, Bernstein E.. Epigenetics: a landscape takes shape. Cell 2007;128: 635–8. [DOI] [PubMed] [Google Scholar]
- 26. Irier HA, Jin P.. Dynamics of DNA methylation in aging and Alzheimer's disease. DNA Cell Biol 2012;31: S-42–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Skinner MK. Environmental epigenetic transgenerational inheritance and somatic epigenetic mitotic stability. Epigenetics 2011;6: 838–42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Fraga MF, Ballestar E, Paz MF, Ropero S, Setien F, Ballestar ML, Heine-Suñer D, Cigudosa JC, Urioste M, Benitez J.. Epigenetic differences arise during the lifetime of monozygotic twins. Proc Natl Acad Sci U S A 2005;102: 10604–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Poulsen P, Esteller M, Vaag A, Fraga MF.. The epigenetic basis of twin discordance in age-related diseases. Pediat Res 2007;61: 38R–42R. [DOI] [PubMed] [Google Scholar]
- 30. Guo JU, Ma DK, Mo H, Ball MP, Jang M-H, Bonaguidi MA, Balazer JA, Eaves HL, Xie B, Ford E.. Neuronal activity modifies the DNA methylation landscape in the adult brain. Nat Neurosci 2011;14: 1345–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Siegmund KD, Connor CM, Campan M, Long TI, Weisenberger DJ, Biniszkiewicz D, Jaenisch R, Laird PW, Akbarian S.. DNA methylation in the human cerebral cortex is dynamically regulated throughout the life span and involves differentiated neurons. PLoS One 2007;2: e895. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32. Bannister AJ, Kouzarides T.. Regulation of chromatin by histone modifications. Cell Res 2011;21: 381–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Castel SE, Martienssen RA.. RNA interference in the nucleus: roles for small RNAs in transcription, epigenetics and beyond. Nat Rev Genet 2013;14: 100–12. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Esteller M. Non-coding RNAs in human disease. Nat Rev Genet 2011;12: 861–74. [DOI] [PubMed] [Google Scholar]
- 35. Kouzarides T. Chromatin modifications and their function. Cell 2007;128: 693–705. [DOI] [PubMed] [Google Scholar]
- 36. Mercer TR, Dinger ME, Mattick JS.. Long non-coding RNAs: insights into functions. Nat Rev Genet 2009;10: 155–9. [DOI] [PubMed] [Google Scholar]
- 37. Bestor TH. The DNA methyltransferases of mammals. Hum Mol Genet 2000;9: 2395–402. [DOI] [PubMed] [Google Scholar]
- 38. Handel AE, Ebers GC, Ramagopalan SV.. Epigenetics: molecular mechanisms and implications for disease. Trends Mol Med 2010;16: 7–16. [DOI] [PubMed] [Google Scholar]
- 39. Selhub J. Folate, vitamin B12 and vitamin B6 and one carbon metabolism. J Nutr Health Aging 2001;6: 39–42. [PubMed] [Google Scholar]
- 40. Clarke R, Smith AD, Jobst KA, Refsum H, Sutton L, Ueland PM.. Folate, vitamin B12, and serum total homocysteine levels in confirmed Alzheimer disease. Arch Neurol 1998;55: 1449–55. [DOI] [PubMed] [Google Scholar]
- 41. Leblhuber F, Walli J, Artner-Dworzak E, Vrecko K, Widner B, Reibnegger G, Fuchs D.. Hyperhomocysteinemia in dementia. J Neural Transm 2000;107: 1469–74. [DOI] [PubMed] [Google Scholar]
- 42. Seshadri S, Beiser A, Selhub J, Jacques PF, Rosenberg IH, D'Agostino RB, Wilson PW, Wolf PA.. Plasma homocysteine as a risk factor for dementia and Alzheimer's disease. N Engl J Med 2002;346: 476–83. [DOI] [PubMed] [Google Scholar]
- 43. Breteler MM. Vascular risk factors for Alzheimer’s disease: an epidemiologic perspective. Neurobiol Aging 2000;21: 153–60. [DOI] [PubMed] [Google Scholar]
- 44. Kivipelto M, Helkala E-L, Laakso MP, Hänninen T, Hallikainen M, Alhainen K, Soininen H, Tuomilehto J, Nissinen A.. Midlife vascular risk factors and Alzheimer's disease in later life: longitudinal, population based study. BMJ 2001;322: 1447–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Kivipelto M, Ngandu T, Fratiglioni L, Viitanen M, Kåreholt I, Winblad B, Helkala E-L, Tuomilehto J, Soininen H, Nissinen A.. Obesity and vascular risk factors at midlife and the risk of dementia and Alzheimer disease. Arch Neurol 2005;62: 1556–60. [DOI] [PubMed] [Google Scholar]
- 46. Luchsinger J, Reitz C, Honig LS, Tang M-X, Shea S, Mayeux R.. Aggregation of vascular risk factors and risk of incident Alzheimer disease. Neurology 2005;65: 545–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. Yoshitake T, Kiyohara Y, Kato I, Ohmura T, Iwamoto H, Nakayama K, Ohmori S, Nomiyama K, Kawano H, Ueda K.. Incidence and risk factors of vascular dementia and Alzheimer's disease in a defined elderly Japanese population The Hisayama Study. Neurology 1995;45: 1161–8. [DOI] [PubMed] [Google Scholar]
- 48. Suzuki MM, Bird A.. DNA methylation landscapes: provocative insights from epigenomics. Nat Rev Genet 2008;9: 465–76. [DOI] [PubMed] [Google Scholar]
- 49. Bird A. DNA methylation patterns and epigenetic memory. Genes Dev 2002;16: 6–21. [DOI] [PubMed] [Google Scholar]
- 50. Jones PA. Functions of DNA methylation: islands, start sites, gene bodies and beyond. Nat Rev Genet 2012;13: 484–92. [DOI] [PubMed] [Google Scholar]
- 51. Fu Y, Luo G-Z, Chen K, Deng X, Yu M, Han D, Hao Z, Liu J, Lu X, Doré LC.. N6-methyldeoxyadenosine marks active transcription start sites in Chlamydomonas. Cell 2015;161: 879–92. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Greer EL, Blanco MA, Gu L, Sendinc E, Liu J, Aristizábal-Corrales D, Hsu C-H, Aravind L, He C, Shi Y.. DNA methylation on N 6-adenine in C. elegans. Cell 2015;161: 868–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. Zhang G, Huang H, Liu D, Cheng Y, Liu X, Zhang W, Yin R, Zhang D, Zhang P, Liu J.. N6-methyladenine DNA modification in Drosophila. Cell 2015;161: 893–906. [DOI] [PubMed] [Google Scholar]
- 54. Wu TP, Wang T, Seetin MG, Lai Y, Zhu S, Lin K, Liu Y, Byrum SD, Mackintosh SG, Zhong M, et al. DNA methylation on N6-adenine in mammalian embryonic stem cells. Nature 2016 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Bachman M, Uribe-Lewis S, Yang X, Burgess HE, Iurlaro M, Reik W, Murrell A, Balasubramanian S.. 5-Formylcytosine can be a stable DNA modification in mammals. Nat Chem Biol 2015;11: 555–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Bachman M, Uribe-Lewis S, Yang X, Williams M, Murrell A, Balasubramanian S.. 5-Hydroxymethylcytosine is a predominantly stable DNA modification. Nat Chem 2014;6: 1049–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57. Raiber E-A, Murat P, Chirgadze DY, Beraldi D, Luisi BF, Balasubramanian S.. 5-Formylcytosine alters the structure of the DNA double helix. Nat Struct Mol Biol 2015;22: 44–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Bell JT, Tsai P-C, Yang T-P, Pidsley R, Nisbet J, Glass D, Mangino M, Zhai G, Zhang F, Valdes A.. Epigenome-wide scans identify differentially methylated regions for age and age-related phenotypes in a healthy ageing population. PLoS Genet 2012;8: e1002629. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Song F, Smith JF, Kimura MT, Morrow AD, Matsuyama T, Nagase H, Held WA.. Association of tissue-specific differentially methylated regions (TDMs) with differential gene expression. Proc Natl Acad Sci USA 2005;102: 3336–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Kriaucionis S, Heintz N.. The nuclear DNA base 5-hydroxymethylcytosine is present in Purkinje neurons and the brain. Science 2009;324: 929–30. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Tahiliani M, Koh KP, Shen Y, Pastor WA, Bandukwala H, Brudno Y, Agarwal S, Iyer LM, Liu DR, Aravind L.. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 2009;324: 930–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Globisch D, Münzel M, Müller M, Michalakis S, Wagner M, Koch S, Brückl T, Biel M, Carell T.. Tissue distribution of 5-hydroxymethylcytosine and search for active demethylation intermediates. PLoS One 2010;5: e15367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Münzel M, Globisch D, Brückl T, Wagner M, Welzmiller V, Michalakis S, Müller M, Biel M, Carell T. Quantification of the sixth DNA base hydroxymethylcytosine in the brain. Angewandte Chemie International Edition 2010;49: 5375–7. [DOI] [PubMed] [Google Scholar]
- 64. Creyghton MP, Cheng AW, Welstead GG, Kooistra T, Carey BW, Steine EJ, Hanna J, Lodato MA, Frampton GM, Sharp PA.. Histone H3K27ac separates active from poised enhancers and predicts developmental state. Proc Natl Acad Sci USA 2010;107: 21931–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Shu L, Sun W, Li L, Xu Z, Lin L, Xie P, Shen H, Huang L, Xu Q, Jin P, et al. Genome-wide alteration of 5-hydroxymenthylcytosine in a mouse model of Alzheimer's disease. BMC Genomics 2016;17: 381. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Szulwach KE, Li X, Li Y, Song CX, Han JW, Kim S, Namburi S, Hermetz K, Kim JJ, Rudd MK, et al. Integrating 5-hydroxymethylcytosine into the epigenomic landscape of human embryonic stem cells. PLoS Genet 2011;7: e1002154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. He C. Grand challenge commentary: RNA epigenetics?. Nat Chem Biol 2010;6: 863.. [DOI] [PubMed] [Google Scholar]
- 68. Motorin Y, Helm M.. RNA nucleotide methylation. Wiley Interdisc Rev: RNA 2011;2: 611–31. [DOI] [PubMed] [Google Scholar]
- 69. Liu J, Jia G.. Methylation modifications in eukaryotic messenger RNA. J Genet Genom 2014;41: 21–33. [DOI] [PubMed] [Google Scholar]
- 70. Fu Y, Dominissini D, Rechavi G, He C.. Gene expression regulation mediated through reversible m6A RNA methylation. Nat Rev Genet 2014;15: 293–306. [DOI] [PubMed] [Google Scholar]
- 71. Liu N, Pan T.. RNA epigenetics. Transl Res 2015;165: 28–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Liu N, Pan T.. N6-methyladenosine-encoded epitranscriptomics. Nat Struct Mol Biol 2016;23: 98–102. [DOI] [PubMed] [Google Scholar]
- 73. Dominissini D, Moshitch-Moshkovitz S, Schwartz S, Salmon-Divon M, Ungar L, Osenberg S, Cesarkas K, Jacob-Hirsch J, Amariglio N, Kupiec M.. Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq. Nature 2012;485: 201–6. [DOI] [PubMed] [Google Scholar]
- 74. Meyer KD, Saletore Y, Zumbo P, Elemento O, Mason CE, Jaffrey SR.. Comprehensive analysis of mRNA methylation reveals enrichment in 3′ UTRs and near stop codons. Cell 2012;149: 1635–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75. Ge J, Yu Y-T.. RNA pseudouridylation: new insights into an old modification. Trends Biochem Sci 2013;38: 210–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Fernández IS, Ng CL, Kelley AC, Wu G, Yu Y-T, Ramakrishnan V.. Unusual base pairing during the decoding of a stop codon by the ribosome. Nature 2013;500: 107–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Karijolich J, Yu Y-T.. Converting nonsense codons into sense codons by targeted pseudouridylation. Nature 2011;474: 395–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Dominissini D, Nachtergaele S, Moshitch-Moshkovitz S, Peer E, Kol N, Ben-Haim MS, Dai Q, Di Segni A, Salmon-Divon M, Clark WC.. The dynamic N1-methyladenosine methylome in eukaryotic messenger RNA. Nature 2016;530: 441–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Li X, Xiong X, Wang K, Wang L, Shu X, Ma S, Yi C.. Transcriptome-wide mapping reveals reversible and dynamic N1-methyladenosine methylome. Nat Chem Biol 2016 [DOI] [PubMed] [Google Scholar]
- 80. Gupta RA, Shah N, Wang KC, Kim J, Horlings HM, Wong DJ, Tsai M-C, Hung T, Argani P, Rinn JL.. Long non-coding RNA HOTAIR reprograms chromatin state to promote cancer metastasis. Nature 2010;464: 1071–6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81. Hashimoto H, Vertino PM, Cheng X.. Molecular coupling of DNA methylation and histone methylation. Epigenomics.2010;2: 657–69 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82. Mastroeni D, Grover A, Delvaux E, Whiteside C, Coleman PD, Rogers J.. Epigenetic changes in Alzheimer's disease: decrements in DNA methylation. Neurobiol Aging 2010;31: 2025–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83. Lahiri DK, Maloney B, Zawia NH.. The LEARn model: an epigenetic explanation for idiopathic neurobiological diseases. Mol Psychiatry 2009;14: 992–1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Sanchez-Mut JV, Aso E, Panayotis N, Lott I, Dierssen M, Rabano A, Urdinguio RG, Fernandez AF, Astudillo A, Martin-Subero JI.. DNA methylation map of mouse and human brain identifies target genes in Alzheimer’s disease. Brain 2013;awt237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Mastroeni D, McKee A, Grover A, Rogers J, Coleman PD.. Epigenetic differences in cortical neurons from a pair of monozygotic twins discordant for Alzheimer's disease. PLoS One 2009;4: e6617.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Lashley T, Gami P, Valizadeh N, Li A, Revesz T, Balazs R.. Alterations in global DNA methylation and hydroxymethylation are not detected in Alzheimer's disease. Neuropathol Appl Neurobiol 2015;41: 497–506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Phipps AJ, Vickers JC, Taberlay PC, Woodhouse A.. Neurofilament-labeled pyramidal neurons and astrocytes are deficient in DNA methylation marks in Alzheimer's disease. Neurobiol Aging 2016;45: 30–42. [DOI] [PubMed] [Google Scholar]
- 88. Hof PR, Morrison JH.. Quantitative analysis of a vulnerable subset of pyramidal neurons in Alzheimer's disease: II. Primary and secondary visual cortex. J Compar Neurol 1990;301: 55–64. [DOI] [PubMed] [Google Scholar]
- 89. Thangavel R, Sahu Sk, Van Hoesen Gw, Zaheer A.. Loss of nonphosphorylated neurofilament immunoreactivity in temporal cortical areas in Alzheimer's disease. Neuroscience 2009;160: 427–33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90. Vickers J, Riederer B, Marugg R, Buee-Scherrer V, Buee L, Delacourte A, Morrison J.. Alterations in neurofilament protein immunoreactivity in human hippocampal neurons related to normal aging and Alzheimer's disease. Neuroscience 1994;62: 1–13. [DOI] [PubMed] [Google Scholar]
- 91. DeKosky St, Scheff Sw.. Synapse loss in frontal cortex biopsies in Alzheimer's disease: correlation with cognitive severity. Ann Neurol 1990;27: 457–64. [DOI] [PubMed] [Google Scholar]
- 92. Davies C, Mann D, Sumpter P, Yates P.. A quantitative morphometric analysis of the neuronal and synaptic content of the frontal and temporal cortex in patients with Alzheimer's disease. J Neurol Sci 1987;78: 151–64. [DOI] [PubMed] [Google Scholar]
- 93. Coppieters N, Dieriks BV, Lill C, Faull RL, Curtis MA, Dragunow M.. Global changes in DNA methylation and hydroxymethylation in Alzheimer's disease human brain. Neurobiol Aging 2014;35: 1334–44. [DOI] [PubMed] [Google Scholar]
- 94. Rao J, Keleshian V, Klein S, Rapoport S.. Epigenetic modifications in frontal cortex from Alzheimer's disease and bipolar disorder patients. Transl Psychiatry 2012;2: e132.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95. Chouliaras L, Mastroeni D, Delvaux E, Grover A, Kenis G, Hof PR, Steinbusch HW, Coleman PD, Rutten BP, van den Hove DL.. Consistent decrease in global DNA methylation and hydroxymethylation in the hippocampus of Alzheimer's disease patients. Neurobiol Aging 2013;34: 2091–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96. Bradley-Whitman MA, Lovell MA.. Epigenetic changes in the progression of Alzheimer's disease. Mech Ageing Dev 2013;134: 486–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97. Bakulski KM, Dolinoy DC, Sartor MA, Paulson HL, Konen JR, Lieberman AP, Albin RL, Hu H, Rozek LS.. Genome-wide DNA methylation differences between late-onset Alzheimer's disease and cognitively normal controls in human frontal cortex. J Alzheimer's Dis: JAD 2012;29: 571–88. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. De Jager PL, Srivastava G, Lunnon K, Burgess J, Schalkwyk LC, Yu L, Eaton ML, Keenan BT, Ernst J, McCabe C.. Alzheimer's disease: early alterations in brain DNA methylation at ANK1, BIN1, RHBDF2 and other loci. Nat Neurosci 2014;17: 1156–63. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99. Humphries CE, Kohli MA, Nathanson L, Whitehead P, Beecham G, Martin E, Mash DC, Pericak-Vance MA, Gilbert J.. Integrated whole transcriptome and DNA methylation analysis identifies gene networks specific to late-onset Alzheimer's disease. J Alzheimer's Dis 2015;44: 977–87. [DOI] [PubMed] [Google Scholar]
- 100. Watson CT, Roussos P, Garg P, Ho DJ, Azam N, Katsel PL, Haroutunian V, Sharp AJ.. Genome-wide DNA methylation profiling in the superior temporal gyrus reveals epigenetic signatures associated with Alzheimer’s disease. Genome Med 2016;8: 1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Lord J, Cruchaga C.. The epigenetic landscape of Alzheimer's disease. Nat Neurosci 2014;17: 1138–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102. Lunnon K, Smith R, Hannon E, De Jager PL, Srivastava G, Volta M, Troakes C, Al-Sarraj S, Burrage J, Macdonald R.. Methylomic profiling implicates cortical deregulation of ANK1 in Alzheimer's disease. Nat Neurosci 2014;17: 1164–70. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103. Barrachina M, Ferrer I.. DNA methylation of Alzheimer disease and tauopathy-related genes in postmortem brain. J Neuropathol Exp Neurol 2009;68: 880–91. [DOI] [PubMed] [Google Scholar]
- 104. Iwata A, Nagata K, Hatsuta H, Takuma H, Bundo M, Iwamoto K, Tamaoka A, Murayama S, Saido T, Tsuji S.. Altered CpG methylation in sporadic Alzheimer's disease is associated with APP and MAPT dysregulation. Hum Mol Genet 2013;ddt451. [DOI] [PubMed] [Google Scholar]
- 105. Poirier J, Bertrand P, Kogan S, Gauthier S, Davignon J, Bouthillier D.. Apolipoprotein E polymorphism and Alzheimer's disease. Lancet 1993;342: 697–9. [DOI] [PubMed] [Google Scholar]
- 106. Jiang Q, Lee CD, Mandrekar S, Wilkinson B, Cramer P, Zelcer N, Mann K, Lamb B, Willson TM, Collins JL.. ApoE promotes the proteolytic degradation of Aβ. Neuron 2008;58: 681–93. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Wisniewski T, Golabek A, Matsubara E, Ghiso J, Frangione B.. Apolipoprotein E: binding to soluble Alzheimer′ s β-amyloid. Biochem Biophys Res Commun 1993;192: 359–65. [DOI] [PubMed] [Google Scholar]
- 108. Corder E, Saunders A, Strittmatter W, Schmechel D, Gaskell P, Small G, Roses A, Haines J, Pericak-Vance MA.. Gene dose of apolipoprotein E type 4 allele and the risk of Alzheimer’s disease in late onset families. Science 1993;261: 921–3. [DOI] [PubMed] [Google Scholar]
- 109. Wang S-C, Oelze B, Schumacher A.. Age-specific epigenetic drift in late-onset Alzheimer's disease. PLoS One 2008;3: e2698.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110. Hollingworth P, Harold D, Sims R, Gerrish A, Lambert J-C, Carrasquillo MM, Abraham R, Hamshere ML, Pahwa JS, Moskvina V.. Common variants at ABCA7, MS4A6A/MS4A4E, EPHA1, CD33 and CD2AP are associated with Alzheimer's disease. Nat Genet 2011;43: 429–35. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111. Humphries C, Kohli MA, Whitehead P, Mash DC, Pericak-Vance MA, Gilbert J.. Alzheimer disease (AD) specific transcription, DNA methylation and splicing in twenty AD associated loci. Mol Cell Neurosci 2015;67: 37–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112. Lambert J-C, Ibrahim-Verbaas CA, Harold D, Naj AC, Sims R, Bellenguez C, Jun G, DeStefano AL, Bis JC, Beecham GW.. Meta-analysis of 74,046 individuals identifies 11 new susceptibility loci for Alzheimer's disease. Nat Genet 2013;45: 1452–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113. Yu L, Chibnik LB, Srivastava GP, Pochet N, Yang J, Xu J, Kozubek J, Obholzer N, Leurgans SE, Schneider JA.. Association of brain DNA methylation in SORL1, ABCA7, HLA-DRB5, SLC24A4, and BIN1 with pathological diagnosis of Alzheimer disease. JAMA Neurol 2015;72: 15–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114. LaFerla FM, Green KN, Oddo S.. Intracellular amyloid-β in Alzheimer's disease. Nat Rev Neurosci 2007;8: 499–509. [DOI] [PubMed] [Google Scholar]
- 115. Rogaeva E, Meng Y, Lee JH, Gu Y, Kawarai T, Zou F, Katayama T, Baldwin CT, Cheng R, Hasegawa H.. The neuronal sortilin-related receptor SORL1 is genetically associated with Alzheimer disease. Nat Genet 2007;39: 168–77. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116. Li H, Karl T, Garner B.. Understanding the function of ABCA7 in Alzheimer's disease. Biochem Soc Trans 2015;43: 920–3. [DOI] [PubMed] [Google Scholar]
- 117. Satoh K, Abe-Dohmae S, Yokoyama S, George-Hyslop PS, Fraser P.. ATP-binding cassette transporter A7 (ABCA7) effects on amyloid processing and relevance to Alzheimer's disease. Alzheimer's Dementia 2012;8: P473.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118. Satoh K, Abe-Dohmae S, Yokoyama S, St George-Hyslop P, Fraser PE.. ATP-binding cassette transporter A7 (ABCA7) loss of function alters Alzheimer amyloid processing. J Biol Chem 2015;290: 24152–65. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119. Tan M-S, Yu J-T, Tan L.. Bridging integrator 1 (BIN1): form, function, and Alzheimer's disease. Trends Mol Med 2013;19: 594–603. [DOI] [PubMed] [Google Scholar]
- 120. Lonze BE, Ginty DD.. Function and regulation of CREB family transcription factors in the nervous system. Neuron 2002;35: 605–23. [DOI] [PubMed] [Google Scholar]
- 121. Akamatsu M, Aota S-I, Suwa A, Ueda K, Amachi T, Yamada KM, Akiyama SK, Kioka N.. Vinexin forms a signaling complex with Sos and modulates epidermal growth factor-induced c-Jun N-terminal kinase/stress-activated protein kinase activities. J Biol Chem 1999;274: 35933–7. [DOI] [PubMed] [Google Scholar]
- 122. Ito H, Usuda N, Atsuzawa K, Iwamoto I, Sudo K, Katoh‐Semba R, Mizutani K, Morishita R, Deguchi T, Nozawa Y.. Phosphorylation by extracellular signal‐regulated kinase of a multidomain adaptor protein, vinexin, at synapses. J Neurochem 2007;100: 545–54. [DOI] [PubMed] [Google Scholar]
- 123. Mendioroz M, Celarain N, Altuna M, Sanchez-Ruiz de Gordoa J, Zelaya MV, Roldan M, Rubio I, Larumbe R, Erro ME, Mendez I, et al. CRTC1 gene is differentially methylated in the human hippocampus in Alzheimer's disease. Alzheimers Res Ther 2016;8: 15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124. Akiyama H, Barger S, Barnum S, Bradt B, Bauer J, Cole GM, Cooper NR, Eikelenboom P, Emmerling M, Fiebich BL.. Inflammation and Alzheimer’s disease. Neurobiol Aging 2000;21: 383–421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125. Lee Y-J, Han SB, Nam S-Y, Oh K-W, Hong JT.. Inflammation and Alzheimer’s disease. Arch Pharmacal Res 2010;33: 1539–56. [DOI] [PubMed] [Google Scholar]
- 126. Heneka MT, O’Banion MK, Terwel D, Kummer MP.. Neuroinflammatory processes in Alzheimer’s disease. J Neural Transm 2010;117: 919–47. [DOI] [PubMed] [Google Scholar]
- 127. López-González I, Schlüter A, Aso E, Garcia-Esparcia P, Ansoleaga B, Llorens F, Carmona M, Moreno J, Fuso A, Portero-Otin M.. Neuroinflammatory signals in alzheimer disease and APP/PS1 transgenic mice. J Neuropathol Exp Neurol 2015;74: 319–44. [DOI] [PubMed] [Google Scholar]
- 128. Tekpli X, Landvik NE, Anmarkud KH, Skaug V, Haugen A, Zienolddiny S.. DNA methylation at promoter regions of interleukin 1B, interleukin 6, and interleukin 8 in non-small cell lung cancer. Cancer Immunol, Immunother 2013;62: 337–45. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129. Hashimoto K, Oreffo RO, Gibson MB, Goldring MB, Roach HI.. DNA demethylation at specific CpG sites in the IL1B promoter in response to inflammatory cytokines in human articular chondrocytes. Arthritis Rheum 2009;60: 3303–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130. Nicolia V, Cavallaro RA, López-González I, Maccarrone M, Scarpa S, Ferrer I, Fuso A.. DNA methylation profiles of selected pro-inflammatory cytokines in Alzheimer disease. J Neuropathol Exp Neurol 2017;nlw099. [DOI] [PubMed] [Google Scholar]
- 131. Di Francesco A, Arosio B, Falconi A, Di Bonaventura MVM, Karimi M, Mari D, Casati M, Maccarrone M, D’Addario C.. Global changes in DNA methylation in Alzheimer’s disease peripheral blood mononuclear cells. Brain Behav Immun 2015;45: 139–44. [DOI] [PubMed] [Google Scholar]
- 132. Bollati V, Galimberti D, Pergoli L, Dalla Valle E, Barretta F, Cortini F, Scarpini E, Bertazzi P, Baccarelli A.. DNA methylation in repetitive elements and Alzheimer disease. Brain Behav Immun 2011;25: 1078–83. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133. Ji H, Wang Y, Liu G, Xu X, Dai D, Chen Z, Zhou D, Zhou X, Han L, Li Y.. OPRK1 promoter hypermethylation increases the risk of Alzheimer’s disease. Neurosci Lett 2015;606: 24–9. [DOI] [PubMed] [Google Scholar]
- 134. Tannorella P, Stoccoro A, Tognoni G, Petrozzi L, Salluzzo MG, Ragalmuto A, Siciliano G, Haslberger A, Bosco Pand Bonuccelli U.. Methylation analysis of multiple genes in blood DNA of Alzheimer’s disease and healthy individuals. Neurosci Lett 2015;600: 143–7. [DOI] [PubMed] [Google Scholar]
- 135. Connor B, Young D, Yan Q, Faull R, Synek Band Dragunow M.. Brain-derived neurotrophic factor is reduced in Alzheimer's disease. Mol Brain Res 1997;49: 71–81. [DOI] [PubMed] [Google Scholar]
- 136. Narisawa-Saito M, Wakabayashi K, Tsuji S, Takahashi H, Nawa H.. Regional specificity of alterations in NGF, BDNF and NT-3 levels in Alzheimer's disease. Neuroreport 1996;7: 2925–8. [DOI] [PubMed] [Google Scholar]
- 137. Tanila H. The role of BDNF in Alzheimer's disease. Neurobiol Dis 2016; [DOI] [PubMed] [Google Scholar]
- 138. Carboni L, Lattanzio F, Candeletti S, Porcellini E, Raschi E, Licastro F, Romualdi P.. Peripheral leukocyte expression of the potential biomarker proteins Bdnf, Sirt1, and Psen1 is not regulated by promoter methylation in Alzheimer's disease patients. Neurosci Lett 2015;605: 44–8. [DOI] [PubMed] [Google Scholar]
- 139. Chang L, Wang Y, Ji H, Dai D, Xu X, Jiang D, Hong Q, Ye H, Zhang X, Zhou X.. Elevation of peripheral BDNF promoter methylation links to the risk of Alzheimer's disease. PLoS One 2014;9: e110773. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140. Nagata T, Kobayashi N, Ishii J, Shinagawa S, Nakayama R, Shibata N, Kuerban B, Ohnuma T, Kondo K, Arai H.. Association between DNA methylation of the BDNF promoter region and clinical presentation in Alzheimer's disease. Dement Geriatr Cogn Disord Extra 2015;5: 64–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141. Szulwach KE, Li X, Li Y, Song CX, Wu H, Dai Q, Irier H, Upadhyay AK, Gearing M, Levey AI, et al. 5-hmC-mediated epigenetic dynamics during postnatal neurodevelopment and aging. Nat Neurosci 2011;14: 1607–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142. Wang F, Yang Y, Lin X, Wang JQ, Wu YS, Xie W, Wang D, Zhu S, Liao YQ, Sun Q, et al. Genome-wide loss of 5-hmC is a novel epigenetic feature of Huntington's disease. Hum Mol Genet 2013;22: 3641–53. [DOI] [PubMed] [Google Scholar]
- 143. Wang T, Pan Q, Lin L, Szulwach KE, Song CX, He C, Wu H, Warren ST, Jin P, Duan R, et al. Genome-wide DNA hydroxymethylation changes are associated with neurodevelopmental genes in the developing human cerebellum. Hum Mol Genet 2012;21: 5500–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144. Condliffe D, Wong A, Troakes C, Proitsi P, Patel Y, Chouliaras L, Fernandes C, Cooper J, Lovestone S, Schalkwyk L, et al. Cross-region reduction in 5-hydroxymethylcytosine in Alzheimer's disease brain. Neurobiol Aging 2014;35: 1850–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145. Mu Y, Gage FH.. Adult hippocampal neurogenesis and its role in Alzheimer's disease. Mol Neurodegener 2011;6: 85.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146. Fitzsimons CP, van Bodegraven E, Schouten M, Lardenoije R, Kompotis K, Kenis G, van den Hurk M, Boks MP, Biojone C, Joca S.. Epigenetic regulation of adult neural stem cells: implications for Alzheimer’s disease. Mol Neurodegener 2014;9: 25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147. Jin K, Peel AL, Mao XO, Xie L, Cottrell BA, Henshall DC, Greenberg DA.. Increased hippocampal neurogenesis in Alzheimer's disease. Proc Natl Acad Sci USA 2004;101: 343–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148. Perry EK, Johnson M, Ekonomou A, Perry RH, Ballard C, Attems J.. Neurogenic abnormalities in Alzheimer's disease differ between stages of neurogenesis and are partly related to cholinergic pathology. Neurobiol Dis 2012;47: 155–62. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149. Mastroeni D, Chouliaras L, Van den Hove DL, Nolz J, Rutten BP, Delvaux E, Coleman PD.. Increased 5-hydroxymethylation levels in the sub ventricular zone of the Alzheimer's brain. Neuroepigenetics 2016;6: 26–31. [Google Scholar]
- 150. Bernstein AI, Lin Y, Street RC, Lin L, Dai Q, Yu L, Bao H, Gearing M, Lah JJ, Nelson PT.. 5-Hydroxymethylation-associated epigenetic modifiers of Alzheimer’s disease modulate Tau-induced neurotoxicity. Hum Mol Genet 2016;ddw109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151. Rhein M, Hagemeier L, Klintschar M, Muschler M, Bleich S, Frieling H.. DNA methylation results depend on DNA integrity—role of post mortem interval. Front Genet 2015;6: 182. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152. Ladd-Acosta C, Pevsner J, Sabunciyan S, Yolken RH, Webster MJ, Dinkins T, Callinan PA, Fan J-B, Potash JB, Feinberg AP.. DNA methylation signatures within the human brain. Am J Hum Genet 2007;81: 1304–15. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153. Laird PW. Principles and challenges of genome-wide DNA methylation analysis. Nat Rev Genet 2010;11: 191–203. [DOI] [PubMed] [Google Scholar]
- 154. Kurdyukov S, Bullock M.. DNA methylation analysis: choosing the right method. Biology 2016;5: 3.. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155. Gjoneska E, Pfenning AR, Mathys H, Quon G, Kundaje A, Tsai L-H, Kellis M.. Conserved epigenomic signals in mice and humans reveal immune basis of Alzheimer/'s disease. Nature 2015;518: 365–9. [DOI] [PMC free article] [PubMed] [Google Scholar]