Next Article in Journal
Deletion of Cd44 Inhibits Metastasis Formation of Liver Cancer in Nf2-Mutant Mice
Next Article in Special Issue
Epigenetic Changes Associated with Osteosarcoma: A Comprehensive Review
Previous Article in Journal
Atherogenesis in Apoe−/− and Ldlr−/− Mice with a Genetically Resistant Background
Previous Article in Special Issue
Epigenetic Mechanisms of Aging and Aging-Associated Diseases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epigenetic Changes Associated with Different Types of Stressors and Suicide

1
College of Osteopathic Medicine, Rocky Vista University, Parker, CO 80112, USA
2
Molecular Biology, Department of Biomedical Sciences, Rocky Vista University, Parker, CO 80112, USA
3
College of Osteopathic Medicine, Des Moines University, Des Moines, IA 50312, USA
*
Author to whom correspondence should be addressed.
Cells 2023, 12(9), 1258; https://doi.org/10.3390/cells12091258
Submission received: 13 March 2023 / Revised: 21 April 2023 / Accepted: 24 April 2023 / Published: 26 April 2023

Abstract

:
Stress is associated with various epigenetic changes. Some stress-induced epigenetic changes are highly dynamic, whereas others are associated with lasting marks on the epigenome. In our study, a comprehensive narrative review of the literature was performed by investigating the epigenetic changes that occur with acute stress, chronic stress, early childhood stress, and traumatic stress exposures, along with examining those observed in post-mortem brains or blood samples of suicide completers and attempters. In addition, the transgenerational effects of these changes are reported. For all types of stress studies examined, the genes Nr3c1, OXTR, SLC6A4, and BDNF reproducibly showed epigenetic changes, with some modifications observed to be passed down to subsequent generations following stress exposures. The aforementioned genes are known to be involved in neuronal development and hormonal regulation and are all associated with susceptibility to mental health disorders including depression, anxiety, personality disorders, and PTSD (post-traumatic stress disorder). Further research is warranted in order to determine the scope of epigenetic actionable targets in individuals suffering from the long-lasting effects of stressful experiences.

1. Introduction

The field of epigenetics has gained greater appreciation in recent years. Determining the repertoire of epigenetic modifications that occur across the genome following stress can provide insight into not only the temporary effects of stress on the biology of an individual but can reveal valuable information about individualized future responses to stress. The term epigenetics is defined by Berger et al. as a “… stably heritable phenotype resulting from changes in a chromosome without alterations in the DNA sequence” [1]. It is well known that epigenetic changes are passed down somatically from cell to cell and in some cases transgenerationally, meaning the epigenetic change persists in subsequent generations, even in the absence of the stressor. Berger et al. further break down epigenetics into three categories which they label “Epigenators”, an environmental trigger that affects the cellular processes within the cell, “Initiators”, an intracellular signal that elicits an effect on the epigenome, and “Maintainers”, local histone variants and DNA methylation that maintain the chromatin state. The Epigenator is the upstream signal from the environment or another signaling pathway that activates the Initiator. An example of this is demonstrated by Cheesman and Weitzman in parasites from the apicomplexa phylum and their proteome acting as an Epigenator in their supposed pathogenesis [1,2]. The Initiator then interprets the signal for the precise location at which to act [1]. This location could include DNA, noncoding RNA, histones, and other structures involved in chromatin remodeling [1]. The Maintainer supports the “epigenetic chromatin state” at the specific location through DNA methylation and other specific epigenetic modifications [1].
Epigenetic modifications include DNA methylation, noncoding RNA such as microRNA, histone modification via methylation, acetylation, ubiquitination, and phosphorylation, and others including SUMOylating. DNA methylation works through enzymes known as DNA methyl transferases that transfer a methyl group to a cytosine-guanine dinucleotide (CpG) region in the DNA [3]. DNA methylation patterns correlate with gene expression levels [4], specifically a reduction in transcription. Histone modification can occur through either the methylation of an arginine or the acetylation of a lysine similar to DNA and can lead to chromosome condensing or opening [3]. Non-coding RNAs are RNA molecules that do not code for a protein but assist in other functions. An example of a non-coding RNA is microRNA (miRNA). miRNA is able to interact with genetic material such as mRNA in such a way as to control gene expression [5]. These epigenetic changes are subjected to a host of outside influences ranging from various types of stress, nutrition, lifestyle choices, and the environment [6,7]. The permanence of these markers is often a question of research and is still being explored.
Healthy physiological responses to stress are designed to maintain homeostasis and survivability [7,8]. Once stress is detected, the hypothalamic-pituitary-adrenal (HPA) axis is activated by the paraventricular nucleus of the hypothalamus releasing corticotropin-releasing hormone (CTRH), which activates the anterior pituitary to release adrenocorticotropic hormone (ACTH). ACTH exerts its effect on the adrenal gland by releasing glucocorticoids such as cortisol that have a body-wide effect on locations that have glucocorticoid receptors (Nr3c1) and a negative effect on the hypothalamus and anterior pituitary, releasing their respective stress hormones. When stress is encountered, it modifies the rhythmic pulsatile nature of the release of CTRH by increasing it [7]. Stress has also been shown to affect the immune system [9]. As reviewed by Dragos et al., acute stress has been shown to increase the resistance to infection whereas chronic stress impairs it [10]. Additionally, a study conducted on acute traumatic brain injury individuals showed an inverse correlation between circulating immune T cells and plasma cortisol levels in mice [11].
Stress is an inevitable part of life. In this review, we focus on five categories of different types of stress and their associated epigenetic changes. Depending on the type of stress experienced by an individual, stress can act as the epigenator in human cells, resulting in initiator-induced changes that can result in stable epigenetic changes through the activity of maintainers. Different types of stress epigenators include acute stress, chronic stress, early childhood stress, traumatic stress, and suicide. Acute stress is considered as an event that is brief in nature or mimics an acute event that causes stress on a subject. Chronic stress is defined as an event that occurs over a longer period of time or an experiment that mimics a chronic stressor on a subject. A stressful event occurring during adolescence is considered as early childhood stress but does not include in utero maternal stress. Traumatic stress is defined as an event that had some sort of terrifying, dangerous, and or shocking impact [12,13] leading to either acute stress disorder or post-traumatic stress disorder (PTSD) or is highly correlated with the development of these disorders following exposure [14]. Suicide is defined as either the thought of ending one’s life, attempting suicide, or completing the act.
This review incorporates the different aspects of stressors and suicide with their relation to epigenetics in one complete narrative review. We explore some of the epigenetic changes that are reported following different types of stress such as acute, chronic, early childhood, and traumatic, along with suicide contemplation/completion. We ask the question, what are the commonalities between the different epigenetic changes caused by the different forms of stress? Do different stressors affect the same genes or genomic regions? Are these epigenetic changes inherited in a transgenerational fashion? What are the implications of these findings with regard to molecular psychiatry and pharmacological treatments?

2. Acute Stress

Acute stress is defined as an event that inflicts stress on the subject for a short period of time before removing it. In the literature, to evaluate the effects of acute stress, researchers tested it in a multitude of ways ranging from acute exposure of benzo[a]pyrene and UV light, restraint tests, heat stress, trier social stress test (TSST), psychosocial stress, and swim tests. Studies assessing epigenetic changes associated with acute stress have been conducted in mice, chickens, rats, clams, and humans. In an experiment with blood clam gills, Guo et al. demonstrated that the acute exposure to benzo[a]pyrene, a compound found in tobacco smoke, coal, and tar, resulted in a decrease in global DNA methylation levels [15], suggesting brief exposure to chemical stressors such as benzo[a]pyrene could potentially lead to epigenetic changes in the model organism. In addition to these findings, the literature has recently demonstrated that immune cell portions (lymphocytes) can be a potential estimator for whole-genome DNA methylation [16]. In a study conducted on adult humans with an acute psychosocial stressor, Apsley et al. showed an increase in immune cells and a concordant increase in whole genome DNA methylation changes [16].
An important receptor in the stress response is the glucocorticoid receptor. Without it, the model organism (i.e., mice and humans) would not be able to elicit a full response to stress due to a lack of receptor binding. This could potentially lead to disorders of the brain including depression [17]. Various studies have looked at the methylation changes in different regions of the glucocorticoid receptor gene (Nr3c1). Using mice as a model system, Li et al., evaluated the 3′UTR region of Nr3c1 5-hydroxymethylcytosine (5-hmC) changes due to an acute restraint stress test [18,19]. They discovered an increase in 5-hmC in the 3′UTR region due to stress in the hippocampus of male mice. The second group of scientists, Rooij et al., looked at the methylation in the 1-C promoter region in the Nr3c1 gene of (N = 675) human participants due to three different psychosocial stressors such as a speech test, mirror-tracing tests, and a Stroop test [20]. They observed that lower stress reactivity measured through heart rate and cortisol response was associated with a lower level of methylation in the 1-C promoter region of fasting blood samples. However, when these results were adjusted for lifestyle variables such as sex, smoking, etc., the associations dissolved suggesting that lifestyle differences play a larger role. Interestingly, the lower methylation levels found in promoter region 1-C of the Nr3c1 gene were associated with a higher perception of stress and a decrease in perceived control and performance. The third group of scientists, Mifsud et al., focused on an area upstream of exon 2 in the Nr3c1 gene called “GR area 1” and a region around exon 17 called “GR area 2” in male Wistar rats’ hippocampi [21]. The CpG levels of “GR area 1” were not affected in dentate gyrus or Cornu Ammonis, but “GR area 2” showed a significant increase in methylation in dentate gyrus and a decrease in Cornu Ammonis due to a swim stress test. An examination of epigenetic changes along the Nr3c1 gene between human and animal studies showed variable region changes reflective of the different acute stressors applied.
Recent literature has assessed chromatin shape changes associated with acute stress, specifically histone methylation and acetylation. Hunter et al. looked at histone H3 lysine 3 (H3K9) modifications at retro-transposable element loci in the hippocampi of male Sprague Dawley rats [22]. They discovered that acute stress was associated with H3K9me3 and that the H3K9 methyl transferase, Suv39h2, is up-regulated in the hippocampus in acute restraint stress groups. In another study focusing on histone modification, Zheng et al. looked at Histone H3 lysine 27 (H3K27) methylation changes due to acute heat stress in (N = 192) Taiwan country chickens [23]. They discovered that chickens whose body temperature changed more than 6.5 °C (the susceptible group) had an increase in H3K27me3 compared to controls in addition to showing positive crosstalk with K36me and K37me in the tails of H3. When looking at the dentate gyrus and CA1 region of the hippocampus in adult male Sprague Dawley rats, Hunter et al.’s prior study examined histone (H3K3, H3K9, and H3K27) modifications [24]. The researchers determined that acute restraint stress was associated with an increase in H3K9me3 in both the dentate gyrus and CA1, whereas there was no effect on H3K4me3. They also observed a reduction in both H3K9me1 and H3K27me3 in the same regions. When comparing these results across the board, there is a consistent me3 in the H3 histone in animal models with no human study to compare to. In the last study focusing on histone modification, Ieraci et al. looked at histone H3 in the promoter region of the brain-derived neurotrophic factor (BDNF) of male mice after an acute restraint stress test [25]. Their results showed no change in the methylation or acetylation status of H3. Interestingly, they found a decrease in the mRNA levels of BDNF after the acute restraint stress test suggesting that there may be another epigenetic mechanism that regulates BDNF expression.
When turning our focus to miRNA (noncoding RNA) affecting gene expression under acute stress conditions, two literature articles were unearthed that report the involvement of specific miRNAs. The first article, which was previously discussed under Nr3c1 by Mifsud et al., also addressed miR-124a, a potential regulator of the Nr3c1 mRNA, and its association with the swim stress test [21]. They observed an increase in expression after a forced swim test in a time-dependent manner while also noticing a decrease in Nr3c1 mRNA levels in the dentate gyrus. Mifsud et al. hypothesize that this may offer a protective mechanism in response to acute stress in the dentate gyrus to repeated forms of stress. Mannironi et al. examined two different miRNAs, miR-135a and miR-124, in adult male mice amygdala after an acute restraint test in the context of acute stress [26]. They determined that in the acute stress response, these two miRNAs are down-regulated, which correlated with higher expression levels of another stress response receptor, the mineralocorticoid receptor, in the amygdala.
Additionally, various studies reported epigenetic changes to genes associated with the regulation of vitamin D levels, tumor suppressors, structural framework, immune response, and oxytocin regulation, associated with acute stress. These genes that encode for different types of proteins included CYP24A1, BRCA2, NOTCH2, FOXO3, GATA3, CSNK2A2, KRT17, CARD14, IRF8, BDNF, OXTR, and PRF1. In a study performed on (N = 32) healthy White females of different Fitzpatrick phototypes (a scale used to predict sunburn risk) by applying acute exposure of UV light (which is most likely different compared to psychological stress) to tissue from the sun-protected area of the lower back [27]. They observed CYP24A1, BRCA2, NOTCH2, FOXO3, and GATA3 to be methylated, KRT17 and CSNK2A2 to be hypomethylated, CARD14 to be demethylated, and IRF8 to be hypermethylated upon acute exposure to UV radiation. Unternaehrer et al. examined two regions in the oxytocin receptor gene (OXTR) and the exon V, Vh, and Vi in the BDNF of human blood samples that underwent the TSST [28]. They found no associated differences in BDNF methylation in the human blood samples but did find an increase in methylation from pre- to post-stress event in the OXTR exon III protein coding region (“OXTR1”) and a decrease in methylation from post-stress to follow up in the OXTR exon III noncoding/coding promoter region (“OXTR2”). These results remained significant even after the blood cell count was controlled. Another study that conducted the TSST but in chronic fatigue syndrome patients observed epigenetic changes in the promoter region of the perforin protein (PRF 1) that creates a channel in cell membranes during an immune response [29]. These epigenetic changes showed an increase in methylation in two PRF 1 CpG sites, -776 and -774, after the TSST in chronic fatigue syndrome blood sampling; however, there were no significant differences between the chronic and non-chronic fatigue syndrome groups.
Lastly, a group of researchers, Li et al., observed the down-regulation of gene expression in the hippocampus of a variety of different genes coding for phosphodiesterase/lipase, a bone morphogenetic protein antagonist, proteins involved with neuronal development, and secreted signaling proteins such as Enpp2, Sostdc1, Ulk4, and Wnt9a after an acute restraint test was performed on mice [19]. Enpp2 and Sostdc1 were hyper-differentially hydroxymethylated (DhMR) in the intron region and the upstream region of the transcription start site, respectively. Ulk4 and Wnt9a were hypo-DhMR in the upstream region of the transcription start site and intron/exon region, respectively. Li et al. also demonstrated up-regulation of genes involved in sequestering p53, transcription suppression and activation, the regulation of growth and apoptosis, cytoplasmic signaling molecules for insulin, structural proteins, and neuronal development such as Banp, Cbfa2t3, Gadd45b, Irs2, Klf15, Smtn, and Spns2 due to the same stress on mice [19]. All of these genes were observed to be hypo-DhMR. Banp and Gadd45b were observed to be hypo-DhMR downstream of the genes, whereas Cbfa2t3 was modified upstream of the transcription start site. The Irs2 and Spns2 hypo-DhMR region was reported in the exon and intron regions. Klf15 and Smtn were observed to be hypo-DhMR along the span of the gene sequence, including intronic and exonic regions. Overall, acute stress was shown to cause a variety of epigenetic changes in a diverse set of genes. The overall summary of these epigenetic changes associated with acute stress can be seen in Table 1.

3. Chronic Stress

Chronic stress is defined as exposure to a reoccurring stressor over an extended period of time. In animal models, a chronic restraint test, a forced swim test in varying water temperatures over consecutive days, chronic water avoidance, exposure to other animals, and social defeat tests are common modalities of chronic stress tests. In human subjects, chronic stress effects are gleaned from participants being exposed to chronic job stress, living situations, elevated platforms, and chronic social defeat tests. Studies assessing epigenetic changes during chronic stress have been conducted and observed in mice, rats, and humans. In McEwen’s review on the system-wide effects of chronic stress, he highlights that chronic stress can have gross changes in the brain resulting in neuronal imbalance [30]. These large-scale modifications can also lead to behavioral changes [31]. A potential cause of this imbalance could be the influence of epigenetics when looking at the gene x environment model.
The Nr3c1 receptor plays a vital role in the stress response for both acute and chronic stress. In the literature, chronic stress has been shown to down-regulate the glucocorticoid receptor mRNA expression levels in the brain [32,33]. In particular, Louwies et al. investigated epigenetic changes in the Nr3c1 gene promoter region of exon 17 in (N = 36) female Fischer-344 rats [34]. They observed increases in DNA methylation in the central nucleus of the amygdala with 7 days of consecutive chronic water avoidance stress. Previously, they saw the down-regulation of the Nr3c1 receptor expression in the same tissue due to the same stressor [35]. Another group of scientists, Witzmann et al., also examined the promoter region of exon 17 but in male Sprague Dawley rats’ (N = 24 for chronic group) adrenal and pituitary glands [36]. They observed that individual site methylation did not change transcript levels after chronic restraint and concluded that there might be promoter-wide methylation changes. Desarnaud et al. applied a social defeat by exposing the model organism to another rat and inspected the promoter region of mice and found no increase in DNA methylation with the down-regulation of the Nr3c1 receptor in tissue samples from the hippocampus [37]. The inspection of a gene regulated by glucocorticoids, ZBTB16 (involved in cellular migration and proliferation), demonstrated hypermethylation in response to prolonged exposure to glucocorticoids in human fetal lung fibroblast cells [38]. These results demonstrate the differences in methylation results for tissue sampling in model organisms such as rats, mice, and human fibroblast cells.
A group of genes that code for specific proteins involved in the neuronal adhesion aspect of neuronal development, NCAM and CHL1, along with the serotonin receptor 5-hydroxytryptamine receptor 1A (5-HT1A), histone H3, and OXTR have been studied for changes of chronic stress. Desarnaud et al. also looked at the promoter regions for the genes, NCAM and CHL1, in mice that have been exposed to a rat or social defeat tests mimicking chronic stress from a social perspective [37]. This stress paradigm has been shown to induce social stress in mice via encountering an unfamiliar male in their home [39]. Interestingly, there was no DNA methylation change in the hippocampus between groups, suggesting that there was some other mechanism at play that down-regulated the adhesion molecules after chronic stress. When looking into the promoter site (-691 CpG) of 5-HT1A of male mice in the prefrontal cortex and midbrain, Le François et al. observed an increase in DNA methylation with an associated increase in 5-HT1A receptor after chronic unpredictable mild stress [40]. When chronic restraint stress was applied to adult male Sprague Dawley rats, Hunter et al. observed a mild increase in H3K4me3 and a reduction in H3K9me3 in the dentate gyrus [24]. An examination by Wiley et al. demonstrated that chronic stress by water avoidance altered proteins involved in tight junctions in the colon of Sprague Dawley rats via the up-regulation of IL-6 in association with H3K9 methylation and an increase in visceral hyperalgesia [41]. In an inspection of OXTR methylation patterns in relation to adult adversity (measured via Unmet Material Needs Scale and neighborhood crime) in (N = 100) African American women, Simons et al. observed increased methylation in the promoter region [42].
Surveying the effects of chronic stress on miRNAs surfaced two particular miRNAs, along with the gene responsible for the initiation of miRNA, Drosha. When (N = 35) adult male Long-Evans hooded rats underwent a 2-week restraint test, Babenko et al. discovered a down-regulation in miR-709 and up-regulation in miR-186 in the hippocampus and prefrontal cortex [43]. Computation analyses of targets for miR-186 were reported to be Gabra4, Creb3, Eps15, A2bp1, and MAP3k2, all important for various brain functions. For miR-709, the predicted targets were Creb5, Efnb3, Nav1, and Nab1 and were found to be important for diverse brain functions. Drosha, an RNase Type III protein, was found to have decreased methylation in intron 9 in the dentate gyrus of (N = 121) adult male mice after 14 days of chronic social defeat [44]. Curiously, Hing et al. also observed increased methylation in the intergenic region of chromosome X after chronic social defeat stress.
When looking at the effects of chronic stress on the heart, Zhang et al. examined different genes important for heart functioning [45]. In particular, they examined changes in desmin (forms the cytoskeleton in cardiac myocytes and aids in the mechanical strength of the heart), Tgfb1 (associations with dilation in the left ventricle and dysfunction in the systole), and genes involved in the ASPC pathway that give sympathetic responsiveness in the heart in (N = 56) male mice. They showed that four consecutive weeks of the chronic restraint test had an associated increase in DNA methylation upstream of the gene that codes for desmin, which correlated with a decrease in desmin mRNA. Similar results were found in Tgfb1 in the downstream region showing methylation and an associated decrease in mRNA levels. When examining the ASPC pathway genes involved in α1-adrenoceptor signaling such as Ppp2r2c (protein phosphatase 2, regulatory subunit B-γ), Ppp2r1a (protein phosphatase 2, regulatory subunit A), Prkca (protein Kinase C-α), and Adra1b (adrenergic receptor-α1B), Zhang et al. discovered that these genes showed alterations in DNA methylation that had associated changes of down-regulation in mRNA of Adra1b and Ppp2r2c while there was up-regulation of Ppp2rla and Prkca.
Lastly, many studies have shown chronic stress can result in epigenetic imprints on genes whose functions have been classically traced to the stress response, such as SLC6A4 or BDNF. Exploring the BDNF gene and its receptor, TrKB, Niknazar et al. observed an increase in the methylation of BDNF exon IV and TrKB in both male (N = 10) and female (N = 20) Wistar rats with an associated decrease in the expression of BDNF in the hippocampus after a forced swim test for 21 consecutive days [46]. Interestingly, females showed a higher methylation in BDNF when compared to male rats, potentially suggesting that female rats are more likely to be affected by chronic stress. Scientists assessing leukocytes in saliva observed lower DNA methylation in the BDNF CpG islands in exon 1 in the promoter region of Japanese workers (N = 774) with the highest job strain scores [47]. When chronic unpredicted mild stress was applied to rats, hippocampal BDNF (promoter region) demonstrated DNA hypermethylation in rats that also exhibited hyperhomocysteinemia [48]. This resulted in a reduction in the expression of the gene as well as cognitive decline. These results from BDNF demonstrate exon-specific variation in methylation in humans and animals with no supporting studies to validate findings. SLC6A4, a gene involved in the reuptake of serotonin, is another gene reported to have epigenetic changes as the result of high-stress environments, specifically in working female nurses (N = 49) [49]. Examining peripheral blood leukocytes revealed a decrease in methylation in the promoter region, potentially leading to an increase in the transcription of the reuptake receptor and less serotonin in the synaptic cleft. Alasaari et al. propose this as a potential coping mechanism for chronic stress. During an examination of chronic stress of living in disadvantaged neighborhoods, stress-related genes (CRF and SLC6A4) and inflammation-related genes (F8 and TLR1) have altered methylation profiles in blood cell samples of B cells, T cells, neutrophils, and natural killer cells (N = 1226) [4]. Smith et al. showed that living in a neighborhood with socioeconomic disadvantages had DNA methylation in the non-promoter regions of CRF, F8, and TLR1 and increased methylation in the shore/shelf site of SLC6A4 [4]. The overall summary of these epigenetic changes associated with chronic stress can be seen in Table 2.

4. Early Childhood Stress

Unique to a specific developmental stage during one’s lifetime, trauma experienced early in life is usually referred to as an ACE (adverse childhood experience), ELE (early life stress), or CM (childhood maltreatment) in the research literature. Examples of early trauma experienced in childhood include physical, sexual, or emotional abuse or neglect, violence within the home, or a hostile social environment, i.e., bullying [50]. A multitude of studies suggests that early traumatic experiences are associated with changes to the epigenome that are also linked to defects in brain developmental programs, psychiatric diseases, and an increased risk for drug abuse and suicide. In addition to psychological disorders, childhood trauma is also significantly correlated with negative physical health outcomes such as an increased risk for metabolic syndromes, chronic pain, and cancer [51]. It has been shown that the molecular changes caused by such incidents of negative childhood experiences are dependent upon the genetics of the individual, the type of stress, and the timing of the stressful event such as that earlier stressful experiences may be associated with longer-term effects [52] (Miguel et al). Collectively, both animal and human studies in response to early life stress report changes in differentially methylated genic and intergenic regions of the genome, altered levels of stress response miRNAs, and different global methylation patterns compared to those individuals who have not experienced childhood trauma.
When investigating possible epigenetic changes associated with childhood stress, the results are mixed. For example, numerous studies have replicated the finding that childhood adversity is associated with an increase in the methylation of the Nr3c1 promoter or the coding region of the gene in blood cells [53,54,55], but other studies show a decrease in the methylation of the Nr3c1 promoter or coding region in blood [56,57] or no change [58]. One recent study examined the epigenetic effects of daily hassles (DH) or daily stress on the Nr3c1 gene in adolescents. The results reported higher DNA methylation of Nr3c1 in adolescents with higher levels of DH, and this was associated with blunted HPA axis reactivity to psychosocial stress. This research group also found a correlation between higher DH and longer HPA axis stress recovery. Additionally, participants with higher Nr3c1 methylation had lower autonomic nervous system (ANS) adaptability to stress, as evidenced by lower parasympathetic withdrawal [59]. Another gene that has been shown in multiple studies to be epigenetically modified in response to childhood stress is the FKBP5 gene. One research group reported the allele-specific FKBP5 demethylation of intron 7 in blood cells [60], and another study showed decreased methylation [61] in leukocytes. On the contrary, a few other researchers show that there is no association between the methylation states of FKBP5 and stress [58,62]. Similar to the conflicting results reported for FKBP5, epigenetic changes in BDNF (brain-derived neurotrophic factor) with childhood stress show the hypermethylation of either the BDNF promoter/coding region [63], hypomethylation [64], hypomethylation [64], or no association of methylation changes with stress [65]. OXTR (oxytocin receptor) gene changes have also been reported, to varying degrees. Hypermethylation in CpG sites in blood [66,67] was found to be associated with childhood trajectories of anxiousness, although other studies fail to show a link between childhood trauma and epigenetic changes in the OXTR gene [68,69].
Of all published studies in the literature examining the association of childhood trauma and epigenetic modifications with SLC6A4 (serotonin transporter), the results suggest that there is an increase in methylation as a result of this type of stress in lymphoblasts [70,71], leukocytes [72], and blood [73]. Additionally, multiple studies show consistent changes in methylation globally across the genome [74] in the cingulate cortex and amygdala [60,75]. Other genes that show associations with childhood stress and epigenetic changes include KITLG [76], PRDM14 [77], and RAB14 [78]. Lastly, changes in miR-15a have been seen in blood cells from individuals who experienced the loss or separation of a parent, sexual abuse, or physical abuse [78].
Similar to results shown in humans, studies in rats reported early stress associated with increased methylation of BDNF [63,79]. Additional non-human studies looking at the epigenetic changes associated with early life stressors yielded hypomethylation in the Nr3c1 of male mice [80], and a reduction in AVP methylation in mice hippocampi following maternal separation stress [81]. An interesting study in mice showed sex-dependent epigenetic changes associated with early life stress. Early life adversities induced a depressive state and altered miR-34a levels in adulthood under acute stress, but solely in females. Specifically in the dorsal raphe nuclei, this miRNA is associated with prefrontal-accumbal serotonin release under acute stress exposure in females. Additional genes modified following early stress in animal models showed changes in MeCP2, CB1, and CRFR2 with maternal separation [82] and mi-RNA 133b when animal mothers were stressed during pregnancy [83]. Overall, studies do suggest that childhood stress does result in aberrations to the epigenome that persist throughout one’s lifetime and even through multiple generations (to be discussed later). The overall summary of these epigenetic changes associated with early childhood stress can be seen in Table 3.

5. Traumatic Stress

Traumatic stress is a form of stress that occurs in response to the rapid and often momentous loss of key valued resources [98]. Examples of trauma include natural disasters, war, sexual assault, tragic death, and other catastrophic events. Exposure to traumatic events has been recognized as part of the human experience and has the potential to impact subsequent development across the lifespan, although individual responses to trauma vary widely. Nevertheless, epigenetic changes have been reported in correlation with this type of stressful experience. Most studies report epigenetic changes for this type of stress in association with PTSD (post-traumatic stress disorder) or based on life experiences that are self-reported as traumatic events.
When comparing individuals diagnosed with PTSD to those non-PTSD control subjects, epigenetic changes abound. Hypomethylation is seen across the Nr3c1 promoter [74,77,80,93], UBE2L3 (Ubiquitin-conjugating enzyme) promoter [92], AHRR (aryl hydrocarbon receptor) [99], F2R, CNPY2, BAIAP2L1, and TBXAS1 [100]. On the contrary, some research efforts have shown the hypermethylation of various gene regions. Traumatic events have shown a significant association with increased methylation levels of CRHR1 [101], MANC1 [102], ADCYAP1R1 CpG island, and PACAP [103]. The gene FKBP5 has also been shown to be associated with PTSD. Yehuda et al. propose a mechanistic model of the relationship between GR and FKBP5 methylation in PTSD whereby GR responsiveness is increased resulting in a decrease in FKBP5 expression through a decrease in cortisol signaling. Additionally, epigenome-wide association studies or EWAS have revealed a number of genes epigenetically modified in correlation with PTSD severity. These genes include BRSK1, LCN8, NFG, DOCK2, ZFP57, and RNF39 [102,103,104].
Significant DNA methylation changes in additional genes have also been seen in post-deployment soldiers with PTSD. These genes include H19 and IL18 [105]. Post-deployment cases of PTSD showed a decrease in methylation levels along the H19 and IL18 genes, resulting in increased levels of both of these proteins. H19 codes for a long noncoding RNA thought to regulate body weight, cancer, inflammation, and aging [106]. Uddin et al. saw that two CpG island sites along the gene loci NRG1 and HGS had increases in DNA methylation in PTSD combat veterans [102]. NRG1 codes for Neuregulin-1 (NRG1), a component of the epidermal growth factor family, and induces the proliferation, differentiation, and survival of several cell types including epithelial cells, glial cells, neurons, and cardiomyocytes [107]. HGS is a gene that codes for hepatocyte growth factor-regulated tyrosine kinase substrate, which functions in regulating endosomal sorting and plays a critical role in the recycling and degradation of membrane receptors. The encoded protein sorts monoubiquitinated membrane proteins into the multivesicular body, targeting these proteins for lysosome-dependent degradation, and has been shown to play an important role in the central nervous system tissue [101]. Additionally, Montalvo-Ortiz et al. found that in 1135 male European–American U.S. veterans who participated in the National Health and Resilience in Veterans Study, CpG sites of genes involved in immune function, transcription regulation, axonal guidance, cell signaling, and protein binding were found to be differentially methylated. Among these, SENP7, which is involved in transcription regulation and has been linked to risk-taking behavior and alcohol consumption in genome-wide association studies, was replicated in an independent veteran cohort and was down-regulated in the medial orbitofrontal cortex of PTSD postmortem brain tissue. In a more recent study examining the blood cells of 290 trauma survivors, methylation analysis showed increases in the CpG site methylation of HPA-related genes, POMC and CRHBP, as predictors for chronic post-traumatic musculoskeletal pain. POMC is a gene that codes for proopiomelanocortin, a protein which is then cleaved to form functional peptides that play a role in the stress response, such as ACTH and melanocyte-simulating hormones [108].
Histone modification changes have also been seen in individuals with PTSD. A review by Zhang et al. concluded that histone modification that was enriched in the promoter regions of candidate genes such as the BDNF and Cdk5 could significantly increase the risk of PTSD [109]. Alterations in levels of histone acetylation and methylation in the hippocampus, amygdala, and prefrontal cortex are associated with PTSD and play key roles in the consolidation, reconsolidation, and extinction of fear memory in PTSD-like animals. It is worth noting that histone modifications of genes in the stress response are mainly involved in the regulation of the immune system, the serotonergic system, the neuropeptide Y-ergic system, and NMDA receptor-related pathways. In addition, histone modification can be regulated by a variety of enzymes, leading to the flexible regulation of PTSD, making drugs that target histone modification good choices for the clinical treatment of PTSD. The overall summary of these epigenetic changes associated with traumatic stress can be seen in Table 4.

6. Suicide

Suicide completers are defined as those individuals who end their life suddenly, passing away from non-natural causes. Suicide non-completers are those observed to have thoughts of suicide or attempted suicide but did not end their life. Research in the area of suicide completers presents limitations in small sample sizes, the freshness of tissue, and being confined to post-mortem studies. Additionally, mimicking suicide conditions in other model organisms such as rats offers challenges to individuals in this field for obvious reasons. With these limitations, researchers have turned their focus to human subjects that have attempted suicide and rely on samples of harvested brain tissue from brain banks or blood samples. Often the suicide completers have other mental disorders such as major depressive disorder (MDD) or bipolar disorder and age differences that must be factored in when examining the data.
One study examined epigenetic changes in suicide completers following a permutation test that controls for age since aging is associated with methylation changes [118]. Haghighi et al. demonstrated that the ventral prefrontal cortex, Brodmann area 47 (BA 47), of (N = 25) depressed-suicide completers had a substantial increase in DNA methylation in genes that involve embryonic and cellular development, cell cycle, cell death and survival, and behavior compared to (N = 28) control [118]. Examining the Brodmann area 10 of the prefrontal cortex in (N = 6) male suicide completers, Schneidera et al. observed lower global DNA methylation [119]. These results suggest region-specific epigenetic changes that occur in suicide brains post-mortem.
Suicidal individuals compared to non-suicidal individuals have shown epigenetic changes in the genes GRIK2, BEGAIN, BDNF, and TrkB, all of which are involved in neuron survival, cell-to-cell communication, and development [120,121,122,123,124]. Intron 13 of GRIK2 (glutamate ionotropic kainate receptor) and BEGAIN (brain-enriched granulated kinase-associated protein) were reported in the literature to be hypomethylated and hypermethylated, respectively, in MDD suicide patients’ (N = 76) cortical brain regions [120]. Looking at the hippocampus, Brodmann area 9, and the blood of suicide completers via hanging (N = 22), Ropret et al. found no methylation changes in the two brain regions and a decrease in methylation upstream of exon 1 in blood samples of the BDNF gene [121]. Interestingly, they also saw an increase in the transcription of BDNF (I-IX) in the Brodmann area 9 and not the hippocampus. When examining whole blood in a cohort of women with emotionally unstable personality disorder, Jamshidi et al. observed an increase in methylation in the promoter region of BDNF associated with suicide behaviors [125]. Turning our focus to the truncated version of the BDNF receptor (TrkB-T1), the 3′UTR region in BA 9 and 10 of (N = 11) suicide completers with low TrkB-T1 expression demonstrated epigenetic changes [122]. These epigenetic changes showed hypermethylation in the 3′UTR in suicide samples. Lastly, examining the BDNF-promoter region IV yielded hypermethylation of (N = 18) suicide individuals’ Wernicke areas with an associated down-regulation of BDNF [123,124]. Keller et al. also detected no correlation between TrkB or its truncated version TrkB-T1 and suicidal behaviors in the Wernicke area [123].
Another list of genes observed in the literature to have epigenetic changes in suicide brain samples includes Elovl5, ARHGEF38, PSORS1C3, OXTR, and CYP2D6. These genes are involved in the elongation of long-chain fatty acids, the regulation of catalytic activity, and a gene involved in psoriasis, respectively. In a blood sample study conducted on attempted suicide individuals with MDD (N = 22), the results showed Elovl5 (elongation of very long chain fatty acids protein 5) upstream regions proximal to the transcription start site underwent DNA methylation, and downstream regions from the transcription start site underwent lower CpG methylation [126]. An examination in the prefrontal cortex (BA 46) of ARHGEF38 (rho guanine nucleotide exchange factor 38) in (N = 23) suicide completers diagnosed with a form of bipolar disorder by the DSM-IV criteria demonstrated hypomethylation across four CpG sites [127]. Additionally, transcripts of ARHGEF38 were decreased in brain tissue. A relatively new psoriasis susceptibility gene, PSORS1C3 (psoriasis susceptibility 1 candidate 3), that lies close to genes involved in immune system regulation has surfaced in the literature with connections to MDD and suicide [128]. Murphy et al. saw hypomethylation in (N = 20) suicide completers with MDD in BA 11 and 25. An examination of the MT2 region of the OXTR (sites -901, -924, and -934) showed no correlation with suicidality assessed with the Columbia Suicide Severity Risk Scale Military Version in salivary samples of (N = 86) male Afghanistan and Iraq war veterans [129]. CYP2D6, a gene coding for the cytochrome P450 enzyme, demonstrated modifications of both hypo- and hypermethylation for males and females, respectively, and were correlated with suicidality, warranting further research [130].
Additionally, recent literature has shown a variety of different additional genes that have surfaced in relation to suicide. These genes and their protein products include ZNF714, NRIP3, Nr3c1, 5-HT1A, SKA2, MAOA, GABRA1, and CERC2. Observations of ZNF714 (zinc finger protein 714) and NRIP3 (nuclear receptor interacting protein 3) were found to be hypomethylated and mixed methylation, respectively, in (N = 9) male hanging suicides BA 9 [131]. Additionally, Kouter et al. also observed a higher expression of both genes when compared to controls in the same study. Analysis of neuropsychiatric genes and protein products Nr3c1, 5-HT1A, SKA2 (spindle- and kinetochore-associated complex subunit 2), MAOA (monoamine oxidase A), GABRA1 (gamma-aminobutyric acid type A receptor subunit alpha1), and NRIP3 showed epigenetic changes in (N = 25) male suicide completers [132]. Specifically, Nr3c1 in the prefrontal cortex and hippocampus demonstrated DNA methylation in the 5′UTR region for teenage suicide completers and an associated decrease in the expression of exon 1 [133]. A recent study by Kouter et al. found higher methylation in the 1B promoter region of Nr3c1 in the insula and blood, whereas lower levels of methylation in the hippocampus and mixed results in the BA 46 and amygdala were detected. SLC6A4_2 amplicon was hypomethylated in BA 46 while 5-HT1A was hypomethylated in blood but hypermethylated in the insula for suicide completers. SKA2 had mixed methylation data in all tissues examined. MAOA in BA 46 and insula were hypomethylated in amplicon MAOA_2. In the hippocampus and blood, GABRA1 had decreased methylation while the insula had increased methylation of suicide individuals. Lastly, the NRIP3 amplicon showed decreased methylation patterns in the hippocampus and insula [132]. Analysis of gene expression levels in suicide brain regions yielded a decrease in SLC6A4 and Nr3c1 gene products in the hippocampus, whereas 5-HT1A showed a borderline increase [132]. A meta-analysis conducted by Zhu et al. showed a correlation of hypermethylation in BDNF, SLC6A4, and Nr3c1 with a higher risk of depression [134]. An investigation of hypermethylation changes in the cerebellum conducted by Policicchio et al. showed one of the top-ranked suicide-associated genes as CERC2, a gene that has been shown to be involved in chromatin remodeling [135].
Lastly, new literature has surfaced reporting a new mouse model designed to study suicidality risk by increasing aggression in these mice through extended social isolation [136]. In particular, the researchers found that in mice subjected to social isolation for 4 weeks, there was an increase in aggression along with an increase in methylation and reduced expression of a gene known to be neuroprotective, PPAR-α [136]. The authors suggest that studying this mouse model may provide insight into mechanisms governing suicide risk associated with social isolation [136]. The overall summary of these epigenetic changes associated with suicide can be seen in Table 5.

7. Transgenerational Effects

Multiple studies have provided evidence that different types of stress exposure, and the associated epigenetic changes linked to maladaptive and poor mental health outcomes, can be passed down through generations [137,138,139]. Regarding chronic stress, Franklin et al. conducted a study with mice and showed that the effects of early chronic stress through changes in DNA methylation in the germline were transmitted through males and could affect the offspring in a sex-dependent manner [82]. In this study, mice were exposed to chronic and unpredictable maternal separation from days 1 to 14 following birth. The results showed the altered methylation of MeCP2, CB1, and the CRFR2 CpG island in the first-generation (F1) germline and second-generation (F2) brain, along with decreased mRNA expression in F2 brain [82]. These changes in methylation were associated with depressive-like behaviors and, additionally, altered behavioral responses to novel and aversive environments in adult mice. Another study showed evidence of intergenerational stress through epigenetic changes. Pregnant Wistar rats received restraint stress during the last week of gestation with male offspring sacrificed on 28 days and 60 days following birth. Prenatal stress induced changes in gpm6a (neuronal membrane glycoprotein) levels in hippocampal and prefrontal cortex tissues and at both ages analyzed, indicating the persistence of this change over time [83]. Another study revealed that chronic stress was found to be transmitted through male Long-Evans rats who were stressed for 27 consecutive days and then mated with control female rats. Specifically at day 21, the offspring were sacrificed and global DNA methylation levels in the hippocampus and frontal cortex were analyzed. Paternal stress prior to conception altered the behavior of all offspring and male offspring specifically, and there was a noted reduction in stress reactivity to novel environments. Paternal stress also altered DNA methylation patterns in the offspring on day 21. Global methylation was reduced in the frontal cortex of female offspring but increased in the hippocampus of both male and female offspring [140]. Additionally, another study in female Wistar rats who were gavaged with 5 mg/kg of fluoxetine (Anti-depressant) FLX during early pregnancy and the last day of lactation found brain methylation changes in male rats. On postnatal day 75, global DNA methylation levels showed an increase in the hypothalamus, cortex, or PAG. Furthermore, early exposure to FLX was also associated with a reduction in time mice spent in social interaction and a decrease in the plasma corticosterone level during restraint stress. Altogether, this study showed results suggesting that maternal exposure to FLX during gestation and lactation results in a long-lasting impact on the DNA methylation of the hippocampus and affects social interaction HPA axis activity during unique types of stress [90].
Multiple studies suggest a possible transgenerational phenomenon associated with traumatic stress. In one study, the sample population consisted of 24 mothers and newborns in the eastern Democratic Republic of Congo, a region with extreme conflict and violence against women. Maternal experiences of war trauma and chronic stress were associated with BDNF methylation in umbilical cord blood, placental tissue, and maternal venous blood. The majority of significant associations were observed in transcription factor binding regions of BDNF [91]. In a separate study, researchers investigated genome-wide sperm DNA methylation patterns in trauma-exposed Vietnam veterans. At the genome-wide level, they identified three CpG sites associated with PTSD in sperm including two intergenic and one CpG within the CCDC88C gene. Of those associated with PTSD, in sperm, 1868 CpGs were also associated with PTSD in peripheral blood including the RORA, CRHR1, and DOCK2 genes that have been previously implicated in PTSD. Additionally, 10 of these CpG sites were significantly associated with a reported history of a diagnosed mental health condition in children, with these same genes being reported to be resistant to demethylation, making them strong candidates for transgenerational inheritance. A different study found epigenetic effects based on the gene variant of interest. In mothers carrying the stress-sensitive T-allele for FKBP5, it was shown that maternal FKBP5 methylation negatively correlated with threat-based ACEs (adverse childhood events) and maternal PTSD symptoms during pregnancy but not deprivation-based ACEs. In infants homozygous for the C allele, infant FKBP5 methylation positively correlated with maternal threat-based ACEs and prenatal PTSD symptom severity, but not deprivation-based ACEs or adversity in adulthood, suggesting the type of ACE and allelic variant may affect the epigenetic change associated with symptoms of PTSD [81]. Lastly, Yehuda et al. found higher levels of methylation across the FKBP5 gene in Holocaust survivors compared with controls and lower levels of FKBP5 expression in offspring [112]. Altogether, these results suggest parental exposure to stress may be passed down to children, and the effects can persist over time.

8. Future Directions/Conclusions

In this review, we examined various epigenetic mechanisms seen with acute stress, chronic stress, early childhood stress, traumatic stress, and suicide. Upon examination of all these articles, we observed four particular genes that surfaced in all categories of stressors examined and suicide. These genes include Nr3c1, OXTR, SLC6A4, and BDNF, all of which demonstrated region- and site-specific methylation patterns in response to various stressors.
In acute stress, numerous studies demonstrated general methylation in the Nr3c1 gene in various regions and higher perceived stress reactivity, whereas the BDNF gene and promoter regional histone modification had no epigenetic changes [18,20,21,25,28]. Site-specific methylation patterns were observed in OXTR [28]. We were unable to find epigenetic alterations associated with SLC6A4 in the literature for acute stress, indicating that acute stress may not be strong enough to elicit an epigenetic alteration in this gene. Other epigenetic modifications that were identified in multiple articles included a histone modification in various subunits and miR-124 in response to acute stressors [21,22,23,24,26].
When examining chronic stress, the literature demonstrated increased methylated changes in exon 17 of Nr3c1 in the amygdala (a fear-processing region) and no changes in the hippocampus or the pituitary and adrenal glands [34,36,37]. Examination of the BDNF gene epigenetic changes in response to chronic stress yielded an increase in methylation in exon IV and the promoter region of rats and the opposite for exon 1 in humans [46,47,48]. Chronic stress yielded increased methylation in the promoter and shore/shelf site of OXTR and SLC6A4, respectively [4,42]. Other epigenetic modifications that were identified in multiple literature articles included a histone H3 modification increase in methylation in K4/K9 (dentate gyrus/colonic tissue), whereas K9 showed a decrease in methylation (dentate gyrus) in response to chronic stress [24,41]. Interestingly, chronic water avoidance stress was associated with an up-regulation of the acute phase reaction promoter of IL-6 [41].
In early childhood stress, the literature demonstrated mixed results for Nr3c1 methylation in the promoter and coding regions [53,54,55,56,57,58]. BDNF gene also demonstrated conflicting results with childhood stress [63,64,65]. An examination of OXTR showed an association with anxiousness and hypermethylation in CpG sites [66,67]. However, some studies fail to demonstrate an epigenetic link with childhood stress [68,69]. SLC6A4 has consistently been shown in the literature to have an increase in gene methylation in lymphoblasts, leukocytes, and blood with childhood stress [60,70,71,72,73,74,75].
Traumatic stress can cause epigenetic changes in the aforementioned genes as well. For Nr3c1, lower methylation in response to various traumatic stressors has been observed [70,73,76,86,88,105,128]. Opposite to Nr3c1, BDNF demonstrated hypermethylation in one combat veteran study in association with PTSD [114]. Exon region 3 of OXTR showed an increase in methylation for females with PTSD [109]. Lastly, the SLC6A4 promoter region was examined in individuals with PTSD, and no association was unearthed [110]. Interestingly, Koenen et al. discovered that individuals with methylation in the promoter region of SLC6A4 served as a protective effect against traumatic stress [117].
In suicide completers and non-completers, literature examining the insula and blood demonstrated Nr3c1 to be more methylated in the 1B promoter region, whereas the hippocampus had lower methylation levels [132]. BDNF showed a decrease in methylation in suicide completers in exon 1 and hypermethylation in promoter region IV in the blood and Wernicke area, respectively [121,123]. In addition to this, the promoter region in women with emotionally unstable personality disorders demonstrated hypermethylation in blood samples [125]. The OXTR MT2 region showed no methylation associations with suicidality indicating that the thoughts of suicide may not be strong enough to elicit an epigenetic alteration in this gene [129]. Lastly, SLC6A4 had hypermethylation in the amplicon region 2 with suicide completion for male human BA 46 [132].
Collectively, some of these studies have not been reproduced and specific results may only show up based on the model organism system, tissue sample location, or stressor. An overall summary of the specific epigenetic changes that occur in the BDNF, Nr3c1, OXTR, and SLC6A4 genes with various types of stress can be seen in Figure 1.
As outlined above, Nr3c1, OXTR, SLC6A4, and BDNF all vary in their associated epigenetic changes, depending on the types of associated stress exposures. Although further study is needed to investigate the full profile of epigenetic changes that occur following stress, current reported changes may provide insight into mechanisms governing development and susceptibility to stress-related pathologies such as PTSD or suicidality. Guide et al. demonstrated that allostatic load (i.e., cumulative stress) contributes to poor health outcomes including psychological dysfunction and depression [141]. Stress type, severity, and longevity likely contribute to a continuum between these epigenetic changes discussed. Specifically, research by Caradonna et al. demonstrated that even short-lived stress has a measurable impact on allostatic load [142]. Once we understand the complete profile of stress-induced epigenetic changes, preventative measures and actionable targets for treatment can be designed and implemented.
When determining the significance of discovered stress-induced epigenetic changes based on published results, it is important to note the reproducibility of the result (or lack thereof), variation in sampling location, gene regions analyzed/reported, and the model organism used. These variations in epigenetic outcomes of different types of stress can be seen in the Nr3c1 gene between rats and mice [34,36,37]. Additionally, there is potential for race-specific epigenomes, which have been shown to play a role in systemic diseases such as metabolic syndrome as outlined by Chitrala et al. [143]. Furthermore, there is a need for an increase in longitudinal human studies of both sexes and various races as these data are limited in the literature. Accessibility to more samples from humans and additional studies with large sample sizes will enable either validating or voiding animal data in terms of accuracy. Ultimately, this could offer better insight into the molecular underpinnings of the different types of stress and epigenetic changes, thus enabling the determination of accurate models that reflect the physiologic mechanism in humans.
If various forms of stress cause specific epigenetic changes and these changes are associated with poor psychiatric and physiological health outcomes, targeting these marks for reversal offers hope to improve patient outcomes. Ferioli et al. reviewed the role of exercise in providing beneficial impacts for patients battling various stress-related pathologies such as cancer and neurodegenerative conditions via epigenetic mechanisms [144]. Additionally, Ieraci et al. demonstrated that physical exercise mimicked the acute stress response in the hippocampus of mice with an increase in mRNA levels of BDNF while preventing its decrease seen shortly after an acute stressor via histone H3 acetylation in the promoter region of BDNF [25]. Additionally, yoga has been shown to demonstrate some changes in DNA methylation patterns and proteins involved in immunity [145]. Coping mechanisms such as social support, avoidance, and problem-solving may not offer a way to prevent or alter the course of methylation caused by stress overload [146]. A pilot study conducted on veterans with PTSD showed that some psychotherapies may offer an epigenetic fix for methylation changes in Nr3c1 and FKBP5 genes; however, the results were not statistically significant and need further validation [147]. The use of specific psychotherapy such as narrative exposure therapy in patients with PTSD has demonstrated protective epigenetic effects in Nr3c1 [148]. Additionally, a pilot study examining mechanisms of MDMA’s efficacy for treating PTSD found MDMA treatment-responsive patients showed more methylation change compared to placebo on one site of the NR3C1 gene [149]. The results from both narrative exposure therapy and MDMA interventions illustrate the need for personalized medicine approaches to treating patients with trauma-related disorders. Additionally, Venditti et al. reviewed some of the recent literature highlights of meditation offering a way to prevent or reverse the effects that occur due to stress from the environment [150]. However, the researchers discuss the ambiguous nature of the molecular basis of meditation and if it acts on the same epigenetic locations or through another mechanism. This illustrates a potential gap in the literature. Collectively, this non-exhaustive list of potential non-pharmacological “fixes” may offer ways to inhibit and even reverse epigenetic changes that arise throughout a person’s life experiences.
Epigenetics is still a relatively new area of research. The more we learn about stress-induced epigenetic changes, the better our chances are to develop preventative measures or treatments based on clinically actionable epigenetic-related targets so that we can treat and potentially break the devastating cycle of transgenerational stress.

Author Contributions

Conceptualization, G.D. and R.R.; methodology, G.D. and R.R.; validation, G.D. and R.R.; formal analysis, G.D. and R.R.; investigation, G.D. and R.R.; resources, G.D. and R.R.; data curation, G.D. and R.R.; writing—original draft preparation, G.D. and R.R.; writing—review and editing, G.D., R.R. and C.D.; abstract, introduction, acute, chronic, suicide, G.D.; abstract, early childhood, traumatic stress, and transgenerational effect, R.R.; conclusion, G.D., R.R. and C.D.; supervision, R.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Pubmed-NIH and other various widely accessible and trusted search engines.

Acknowledgments

We thank the Physician Scientist track at RVU for enabling work like this to be available to medical students. A special thank you to Caitlyn Dee for helping be a reference of aid in organizing and constructing the figures.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Berger, S.L.; Kouzarides, T.; Shiekhattar, R.; Shilatifard, A. An operational definition of epigenetics. Genes Dev. 2009, 23, 781–783. [Google Scholar] [CrossRef]
  2. Cheeseman, K.; Weitzman, J.B. Host-parasite interactions: An intimate epigenetic relationship. Cell Microbiol. 2015, 17, 1121–1132. [Google Scholar] [CrossRef] [PubMed]
  3. Peixoto, P.; Cartron, P.F.; Serandour, A.A.; Hervouet, E. From 1957 to Nowadays: A Brief History of Epigenetics. Int. J. Mol. Sci. 2020, 21, 7571. [Google Scholar] [CrossRef]
  4. Smith, J.A.; Zhao, W.; Wang, X.; Ratliff, S.M.; Mukherjee, B.; Kardia, S.L.R.; Liu, Y.; Roux, A.V.D.; Needham, B.L. Neighborhood characteristics influence DNA methylation of genes involved in stress response and inflammation: The Multi-Ethnic Study of Atherosclerosis. Epigenetics 2017, 12, 662–673. [Google Scholar] [CrossRef] [PubMed]
  5. Mattick, J.S.; Makunin, I.V. Non-coding RNA. Hum. Mol. Genet. 2006, 15 Spec No 1, R17–R29. [Google Scholar] [CrossRef]
  6. Alegría-Torres, J.A.; Baccarelli, A.; Bollati, V. Epigenetics and lifestyle. Epigenomics 2011, 3, 267–277. [Google Scholar] [CrossRef] [PubMed]
  7. Gatta, E.; Saudagar, V.; Auta, J.; Grayson, D.R.; Guidotti, A. Epigenetic landscape of stress surfeit disorders: Key role for DNA methylation dynamics. Int. Rev. Neurobiol. 2021, 156, 127–183. [Google Scholar] [CrossRef]
  8. Yaribeygi, H.; Panahi, Y.; Sahraei, H.; Johnston, T.P.; Sahebkar, A. The impact of stress on body function: A review. Excli. J. 2017, 16, 1057–1072. [Google Scholar] [CrossRef]
  9. Biesmans, S.; Acton, P.D.; Cotto, C.; Langlois, X.; Ver Donck, L.; Bouwknecht, J.A.; Aelvoet, S.A.; Hellings, N.; Meert, T.F.; Nuydens, R. Effect of stress and peripheral immune activation on astrocyte activation in transgenic bioluminescent Gfap-luc mice. Glia 2015, 63, 1126–1137. [Google Scholar] [CrossRef]
  10. Dragoş, D.; Tănăsescu, M.D. The effect of stress on the defense systems. J. Med. Life 2010, 3, 10–18. [Google Scholar]
  11. Dong, T.; Zhi, L.; Bhayana, B.; Wu, M.X. Cortisol-induced immune suppression by a blockade of lymphocyte egress in traumatic brain injury. J. Neuroinflamm. 2016, 13, 197. [Google Scholar] [CrossRef]
  12. Howie, H.; Rijal, C.M.; Ressler, K.J. A review of epigenetic contributions 
to post-traumatic stress disorder. Dialogues Clin. Neurosci. 2019, 21, 417–428. [Google Scholar] [CrossRef]
  13. Pai, A.; Suris, A.M.; North, C.S. Posttraumatic Stress Disorder in the DSM-5: Controversy, Change, and Conceptual Considerations. Behav. Sci. 2017, 7, 7. [Google Scholar] [CrossRef]
  14. American Psychiatric Association. Diagnostic and Statistical Manual of Mental Disorders: DSM-5TM, 5th ed.; American Psychiatric Publishing, a Division of American Psychiatric Association: Washington, DC, USA, 2013. [Google Scholar]
  15. Guo, B.; Feng, D.; Xu, Z.; Qi, P.; Yan, X. Acute benzo[a]pyrene exposure induced oxidative stress, neurotoxicity and epigenetic change in blood clam Tegillarca granosa. Sci. Rep. 2021, 11, 18744. [Google Scholar] [CrossRef]
  16. Apsley, A.T.; Etzel, L.; Hastings, W.J.; Heim, C.C.; Noll, J.G.; O’donnell, K.J.; Schreier, H.M.C.; Shenk, C.E.; Ye, Q.; Shalev, I. Investigating the effects of maltreatment and acute stress on the concordance of blood and DNA methylation methods of estimating immune cell proportions. Clin. Epigenetics 2023, 15, 1–13. [Google Scholar] [CrossRef]
  17. Holsboer, F. The corticosteroid receptor hypothesis of depression. Neuropsychopharmacology 2000, 23, 477–501. [Google Scholar] [CrossRef]
  18. Li, S.; Papale, L.A.; Kintner, D.B.; Sabat, G.; Barrett-Wilt, G.A.; Cengiz, P.; Alisch, R.S. Hippocampal increase of 5-hmC in the glucocorticoid receptor gene following acute stress. Behav. Brain Res. 2015, 286, 236–240. [Google Scholar] [CrossRef]
  19. Li, S.; Papale, L.A.; Zhang, Q.; Madrid, A.; Chen, L.; Chopra, P.; Keleş, S.; Jin, P.; Alisch, R.S. Genome-wide alterations in hippocampal 5-hydroxymethylcytosine links plasticity genes to acute stress. Neurobiol. Dis. 2016, 86, 99–108. [Google Scholar] [CrossRef]
  20. de Rooij, S.R.; Costello, P.M.; Veenendaal, M.V.E.; Lillycrop, K.A.; Gluckman, P.D.; Hanson, M.A.; Painter, R.C.; Roseboom, T.J. Associations between DNA methylation of a glucocorticoid receptor promoter and acute stress responses in a large healthy adult population are largely explained by lifestyle and educational differences. Psychoneuroendocrinology 2012, 37, 782–788. [Google Scholar] [CrossRef]
  21. Mifsud, K.R.; Saunderson, E.A.; Spiers, H.; Carter, S.D.; Trollope, A.F.; Mill, J.; Reul, J.M. Rapid Down-Regulation of Glucocorticoid Receptor Gene Expression in the Dentate Gyrus after Acute Stress in vivo: Role of DNA Methylation and MicroRNA Activity. Neuroendocrinology 2017, 104, 157–169. [Google Scholar] [CrossRef]
  22. Hunter, R.G.; Murakami, G.; Dewell, S.; Seligsohn, M.; Baker, M.E.; Datson, N.A.; McEwen, B.S.; Pfaff, D.W. Acute stress and hippocampal histone H3 lysine 9 trimethylation, a retrotransposon silencing response. Proc. Natl. Acad. Sci. USA 2012, 109, 17657–17662. [Google Scholar] [CrossRef]
  23. Zheng, H.T.; Zhuang, Z.X.; Chen, C.J.; Liao, H.Y.; Chen, H.L.; Hsueh, H.C.; Chen, C.F.; Chen, S.E.; Huang, S.Y. Effects of acute heat stress on protein expression and histone modification in the adrenal gland of male layer-type country chickens. Sci. Rep. 2021, 11, 6499. [Google Scholar] [CrossRef]
  24. Hunter, R.G.; McCarthy, K.J.; Milne, T.A.; Pfaff, D.W.; McEwen, B.S. Regulation of hippocampal H3 histone methylation by acute and chronic stress. Proc. Natl. Acad. Sci. USA 2009, 106, 20912–20917. [Google Scholar] [CrossRef]
  25. Ieraci, A.; Mallei, A.; Musazzi, L.; Popoli, M. Physical exercise and acute restraint stress differentially modulate hippocampal brain-derived neurotrophic factor transcripts and epigenetic mechanisms in mice. Hippocampus 2015, 25, 1380–1392. [Google Scholar] [CrossRef]
  26. Mannironi, C.; Camon, J.; De Vito, F.; Biundo, A.; De Stefano, M.E.; Persiconi, I.; Bozzoni, I.; Fragapane, P.; Mele, A.; Presutti, C. Acute stress alters amygdala microRNA miR-135a and miR-124 expression: Inferences for corticosteroid dependent stress response. PLoS ONE 2013, 8, e73385. [Google Scholar] [CrossRef]
  27. Holzscheck, N.; Söhle, J.; Schläger, T.; Falckenhayn, C.; Grönniger, E.; Kolbe, L.; Wenck, H.; Terstegen, L.; Kaderali, L.; Winnefeld, M.; et al. Concomitant DNA methylation and transcriptome signatures define epidermal responses to acute solar UV radiation. Sci. Rep. 2020, 10, 12918. [Google Scholar] [CrossRef]
  28. Unternaehrer, E.; Luers, P.; Mill, J.; Dempster, E.; Meyer, A.H.; Staehli, S.; Lieb, R.; Hellhammer, D.H.; Meinlschmidt, G. Dynamic changes in DNA methylation of stress-associated genes (OXTR, BDNF ) after acute psychosocial stress. Transl. Psychiatry 2012, 2, e150. [Google Scholar] [CrossRef]
  29. Falkenberg, V.R.; Whistler, T.; Murray, J.R.; Unger, E.R.; Rajeevan, M.S. Acute psychosocial stress-mediated changes in the expression and methylation of perforin in chronic fatigue syndrome. Genet. Epigenet. 2013, 5, 1–9. [Google Scholar] [CrossRef] [PubMed]
  30. McEwen, B.S. Neurobiological and Systemic Effects of Chronic Stress. Chronic Stress 2017, 1, 2470547017692328. [Google Scholar] [CrossRef]
  31. Sanacora, G.; Yan, Z.; Popoli, M. The stressed synapse 2.0: Pathophysiological mechanisms in stress-related neuropsychiatric disorders. Nat. Rev. Neurosci. 2022, 23, 86–103. [Google Scholar] [CrossRef]
  32. Dickens, M.; Romero, L.M.; Cyr, N.E.; Dunn, I.C.; Meddle, S.L. Chronic stress alters glucocorticoid receptor and mineralocorticoid receptor mRNA expression in the European starling (Sturnus vulgaris) brain. J. Neuroendocrinol. 2009, 21, 832–840. [Google Scholar] [CrossRef]
  33. Kitraki, E.; Karandrea, D.; Kittas, C. Long-lasting effects of stress on glucocorticoid receptor gene expression in the rat brain. Neuroendocrinology 1999, 69, 331–338. [Google Scholar] [CrossRef]
  34. Louwies, T.; Greenwood-Van Meerveld, B. Chronic stress increases DNA methylation of the GR promoter in the central nucleus of the amygdala of female rats. Neurogastroenterol. Motil. 2022, 34, e14377. [Google Scholar] [CrossRef]
  35. Louwies, T.; Orock, A.; Greenwood-Van Meerveld, B. Stress-induced visceral pain in female rats is associated with epigenetic remodeling in the central nucleus of the amygdala. Neurobiol. Stress 2021, 15, 100386. [Google Scholar] [CrossRef]
  36. Witzmann, S.R.; Turner, J.D.; Mériaux, S.B.; Meijer, O.C.; Muller, C.P. Epigenetic regulation of the glucocorticoid receptor promoter 1(7) in adult rats. Epigenetics 2012, 7, 1290–1301. [Google Scholar] [CrossRef]
  37. Desarnaud, F.; Jakovcevski, M.; Morellini, F.; Schachner, M. Stress downregulates hippocampal expression of the adhesion molecules NCAM and CHL1 in mice by mechanisms independent of DNA methylation of their promoters. Cell Adh. Migr. 2008, 2, 38–44. [Google Scholar] [CrossRef]
  38. Leung, C.S.; Kosyk, O.; Welter, E.M.; Dietrich, N.; Archer, T.K.; Zannas, A.S. Chronic stress-driven glucocorticoid receptor activation programs key cell phenotypes and functional epigenomic patterns in human fibroblasts. iScience 2022, 25, 104960. [Google Scholar] [CrossRef]
  39. Bartolomucci, A.; Palanza, P.; Sacerdote, P.; Panerai, A.E.; Sgoifo, A.; Dantzer, R.; Parmigiani, S. Social factors and individual vulnerability to chronic stress exposure. Neurosci. Biobehav. Rev. 2005, 29, 67–81. [Google Scholar] [CrossRef]
  40. Le François, B.; Soo, J.; Millar, A.M.; Daigle, M.; Le Guisquet, A.M.; Leman, S.; Minier, F.; Belzung, C.; Albert, P.R. Chronic mild stress and antidepressant treatment alter 5-HT1A receptor expression by modifying DNA methylation of a conserved Sp4 site. Neurobiol. Dis. 2015, 82, 332–341. [Google Scholar] [CrossRef]
  41. Wiley, J.W.; Zong, Y.; Zheng, G.; Zhu, S.; Hong, S. Histone H3K9 methylation regulates chronic stress and IL-6-induced colon epithelial permeability and visceral pain. Neurogastroenterol. Motil. 2020, 32, e13941. [Google Scholar] [CrossRef]
  42. Simons, R.L.; Lei, M.K.; Beach, S.R.H.; Cutrona, C.E.; Philibert, R.A. Methylation of the oxytocin receptor gene mediates the effect of adversity on negative schemas and depression. Dev. Psychopathol. 2017, 29, 725–736. [Google Scholar] [CrossRef] [PubMed]
  43. Babenko, O.; Golubov, A.; Ilnytskyy, Y.; Kovalchuk, I.; Metz, G.A. Genomic and epigenomic responses to chronic stress involve miRNA-mediated programming. PLoS ONE 2012, 7, e29441. [Google Scholar] [CrossRef]
  44. Hing, B.; Braun, P.; Cordner, Z.A.; Ewald, E.R.; Moody, L.; McKane, M.; Willour, V.L.; Tamashiro, K.L.; Potash, J.B. Chronic social stress induces DNA methylation changes at an evolutionary conserved intergenic region in chromosome X. Epigenetics 2018, 13, 627–641. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, P.; Li, T.; Liu, Y.Q.; Zhang, H.; Xue, S.M.; Li, G.; Cheng, H.M.; Cao, J.M. Contribution of DNA methylation in chronic stress-induced cardiac remodeling and arrhythmias in mice. FASEB J. 2019, 33, 12240–12252. [Google Scholar] [CrossRef] [PubMed]
  46. Niknazar, S.; Nahavandi, A.; Peyvandi, A.A.; Peyvandi, H.; Akhtari, A.S.; Karimi, M. Comparison of the Adulthood Chronic Stress Effect on Hippocampal BDNF Signaling in Male and Female Rats. Mol. Neurobiol. 2016, 53, 4026–4033. [Google Scholar] [CrossRef] [PubMed]
  47. Song, Y.; Miyaki, K.; Suzuki, T.; Sasaki, Y.; Tsutsumi, A.; Kawakami, N.; Shimazu, A.; Takahashi, M.; Inoue, A.; Kan, C.; et al. Altered DNA methylation status of human brain derived neurotrophis factor gene could be useful as biomarker of depression. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2014, 165b, 357–364. [Google Scholar] [CrossRef]
  48. Wang, S.-D.; Wang, X.; Zhao, Y.; Xue, B.-H.; Wang, X.-T.; Chen, Y.-X.; Zhang, Z.-Q.; Tian, Y.-R.; Xie, F.; Qian, L.-J. Homocysteine-Induced Disturbances in DNA Methylation Contribute to Development of Stress-Associated Cognitive Decline in Rats. Neurosci. Bull. 2022, 38, 887–900. [Google Scholar] [CrossRef]
  49. Alasaari, J.S.; Lagus, M.; Ollila, H.M.; Toivola, A.; Kivimäki, M.; Vahtera, J.; Kronholm, E.; Härmä, M.; Puttonen, S.; Paunio, T. Environmental stress affects DNA methylation of a CpG rich promoter region of serotonin transporter gene in a nurse cohort. PLoS ONE 2012, 7, e45813. [Google Scholar] [CrossRef]
  50. Austin, A. Association of Adverse Childhood Experiences with Life Course Health and Development. N. C. Med. J. 2018, 79, 99–103. [Google Scholar] [CrossRef]
  51. Thumfart, K.M.; Jawaid, A.; Bright, K.; Flachsmann, M.; Mansuy, I.M. Epigenetics of childhood trauma: Long term sequelae and potential for treatment. Neurosci. Biobehav. Rev. 2022, 132, 1049–1066. [Google Scholar] [CrossRef]
  52. Miguel, P.M.; Pereira, L.O.; Silveira, P.P.; Meaney, M.J. Early environmental influences on the development of children’s brain structure and function. Dev. Med. Child Neurol. 2019, 61, 1127–1133. [Google Scholar] [CrossRef] [PubMed]
  53. Shields, A.E.; Wise, L.A.; Ruiz-Narvaez, E.A.; Seddighzadeh, B.; Byun, H.M.; Cozier, Y.C.; Rosenberg, L.; Palmer, J.R.; Baccarelli, A.A. Childhood abuse, promoter methylation of leukocyte NR3C1 and the potential modifying effect of emotional support. Epigenomics 2016, 8, 1507–1517. [Google Scholar] [CrossRef] [PubMed]
  54. McGowan, P.O.; Sasaki, A.; D’Alessio, A.C.; Dymov, S.; Labonté, B.; Szyf, M.; Turecki, G.; Meaney, M.J. Epigenetic regulation of the glucocorticoid receptor in human brain associates with childhood abuse. Nat. Neurosci. 2009, 12, 342–348. [Google Scholar] [CrossRef] [PubMed]
  55. Tyrka, A.R.; Price, L.H.; Marsit, C.; Walters, O.C.; Carpenter, L.L. Childhood adversity and epigenetic modulation of the leukocyte glucocorticoid receptor: Preliminary findings in healthy adults. PLoS ONE 2012, 7, e30148. [Google Scholar] [CrossRef]
  56. Misiak, B.; Karpiński, P.; Szmida, E.; Grąźlewski, T.; Jabłoński, M.; Cyranka, K.; Rymaszewska, J.; Piotrowski, P.; Kotowicz, K.; Frydecka, D. Adverse Childhood Experiences and Methylation of the FKBP5 Gene in Patients with Psychotic Disorders. J. Clin. Med. 2020, 9, 3792. [Google Scholar] [CrossRef]
  57. Vangeel, E.B.; Kempke, S.; Bakusic, J.; Godderis, L.; Luyten, P.; Van Heddegem, L.; Compernolle, V.; Persoons, P.; Lambrechts, D.; Izzi, B.; et al. Glucocorticoid receptor DNA methylation and childhood trauma in chronic fatigue syndrome patients. J. Psychosom. Res. 2018, 104, 55–60. [Google Scholar] [CrossRef]
  58. Marzi, S.J.; Sugden, K.; Arseneault, L.; Belsky, D.W.; Burrage, J.; Corcoran, D.L.; Danese, A.; Fisher, H.L.; Hannon, E.; Moffitt, T.E.; et al. Analysis of DNA Methylation in Young People: Limited Evidence for an Association between Victimization Stress and Epigenetic Variation in Blood. Am. J. Psychiatry 2018, 175, 517–529. [Google Scholar] [CrossRef]
  59. Chubar, V.; Vaessen, T.; Noortgate, W.V.D.; Lutin, E.; Bosmans, G.; Bekaert, B.; Van Leeuwen, K.; Calders, F.; Weyn, S.; Bijttebier, P.; et al. Mild daily stress, in interaction with NR3C1 DNA methylation levels, is linked to alterations in the HPA axis and ANS response to acute stress in early adolescents. Psychoneuroendocrinology 2023, 150, 106045. [Google Scholar] [CrossRef]
  60. Mehta, D.; Klengel, T.; Conneely, K.N.; Smith, A.K.; Altmann, A.; Pace, T.W.; Rex-Haffner, M.; Loeschner, A.; Gonik, M.; Mercer, K.B.; et al. Childhood maltreatment is associated with distinct genomic and epigenetic profiles in posttraumatic stress disorder. Proc. Natl. Acad. Sci. USA 2013, 110, 8302–8307. [Google Scholar] [CrossRef]
  61. Ramo-Fernández, L.; Boeck, C.; Koenig, A.M.; Schury, K.; Binder, E.B.; Gündel, H.; Fegert, J.M.; Karabatsiakis, A.; Kolassa, I.T. The effects of childhood maltreatment on epigenetic regulation of stress-response associated genes: An intergenerational approach. Sci. Rep. 2019, 9, 983. [Google Scholar] [CrossRef]
  62. Alexander, N.; Kirschbaum, C.; Stalder, T.; Muehlhan, M.; Vogel, S. No association between FKBP5 gene methylation and acute and long-term cortisol output. Transl. Psychiatry 2020, 10, 175. [Google Scholar] [CrossRef] [PubMed]
  63. Roth, T.L.; Lubin, F.D.; Funk, A.J.; Sweatt, J.D. Lasting epigenetic influence of early-life adversity on the BDNF gene. Biol. Psychiatry 2009, 65, 760–769. [Google Scholar] [CrossRef]
  64. Wang, Q.; Shelton, R.C.; Dwivedi, Y. Interaction between early-life stress and FKBP5 gene variants in major depressive disorder and post-traumatic stress disorder: A systematic review and meta-analysis. J. Affect. Disord. 2018, 225, 422–428. [Google Scholar] [CrossRef]
  65. Hossack, M.R.; Reid, M.W.; Aden, J.K.; Gibbons, T.; Noe, J.C.; Willis, A.M. Adverse Childhood Experience, Genes, and PTSD Risk in Soldiers: A Methylation Study. Mil. Med. 2020, 185, 377–384. [Google Scholar] [CrossRef]
  66. Smearman, E.L.; Almli, L.M.; Conneely, K.N.; Brody, G.H.; Sales, J.M.; Bradley, B.; Ressler, K.J.; Smith, A.K. Oxytocin Receptor Genetic and Epigenetic Variations: Association with Child Abuse and Adult Psychiatric Symptoms. Child Dev. 2016, 87, 122–134. [Google Scholar] [CrossRef]
  67. Gouin, J.P.; Zhou, Q.Q.; Booij, L.; Boivin, M.; Côté, S.M.; Hébert, M.; Ouellet-Morin, I.; Szyf, M.; Tremblay, R.E.; Turecki, G.; et al. Associations among oxytocin receptor gene (OXTR) DNA methylation in adulthood, exposure to early life adversity, and childhood trajectories of anxiousness. Sci. Rep. 2017, 7, 7446. [Google Scholar] [CrossRef]
  68. Womersley, J.S.; Hemmings, S.M.J.; Ziegler, C.; Gutridge, A.; Ahmed-Leitao, F.; Rosenstein, D.; Domschke, K.; Seedat, S. Childhood emotional neglect and oxytocin receptor variants: Association with limbic brain volumes. World J. Biol. Psychiatry 2020, 21, 513–528. [Google Scholar] [CrossRef]
  69. Parianen Lesemann, F.H.; Spencer, H.; Montoya, E.R.; Kraaijenvanger, E.J.; He, Y.; Branje, S.; Boks, M.P.; Bos, P.A. Methylation of oxytocin related genes and early life trauma together shape the N170 response to human faces. Eur. Neuropsychopharmacol. 2020, 39, 19–28. [Google Scholar] [CrossRef] [PubMed]
  70. Beach, S.R.H.; Brody, G.H.; Todorov, A.A.; Gunter, T.D.; Philibert, R.A. Methylation at SLC6A4 is linked to family history of child abuse: An examination of the Iowa Adoptee sample. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2010, 153b, 710–713. [Google Scholar] [CrossRef]
  71. Vijayendran, M.; Beach, S.R.; Plume, J.M.; Brody, G.H.; Philibert, R.A. Effects of genotype and child abuse on DNA methylation and gene expression at the serotonin transporter. Front. Psychiatry 2012, 3, 55. [Google Scholar] [CrossRef] [PubMed]
  72. Kang, H.J.; Kim, J.M.; Stewart, R.; Kim, S.Y.; Bae, K.Y.; Kim, S.W.; Shin, I.S.; Shin, M.G.; Yoon, J.S. Association of SLC6A4 methylation with early adversity, characteristics and outcomes in depression. Prog. Neuropsychopharmacol. Biol. Psychiatry 2013, 44, 23–28. [Google Scholar] [CrossRef]
  73. Booij, L.; Szyf, M.; Carballedo, A.; Frey, E.M.; Morris, D.; Dymov, S.; Vaisheva, F.; Ly, V.; Fahey, C.; Meaney, J.; et al. DNA methylation of the serotonin transporter gene in peripheral cells and stress-related changes in hippocampal volume: A study in depressed patients and healthy controls. PLoS ONE 2015, 10, e0119061. [Google Scholar] [CrossRef]
  74. Labonté, B.; Suderman, M.; Maussion, G.; Navaro, L.; Yerko, V.; Mahar, I.; Bureau, A.; Mechawar, N.; Szyf, M.; Meaney, M.J.; et al. Genome-wide epigenetic regulation by early-life trauma. Arch. Gen. Psychiatry 2012, 69, 722–731. [Google Scholar] [CrossRef] [PubMed]
  75. Lutz, P.E.; Tanti, A.; Gasecka, A.; Barnett-Burns, S.; Kim, J.J.; Zhou, Y.; Chen, G.G.; Wakid, M.; Shaw, M.; Almeida, D.; et al. Association of a History of Child Abuse with Impaired Myelination in the Anterior Cingulate Cortex: Convergent Epigenetic, Transcriptional, and Morphological Evidence. Am. J. Psychiatry 2017, 174, 1185–1194. [Google Scholar] [CrossRef]
  76. Houtepen, L.C.; Vinkers, C.H.; Carrillo-Roa, T.; Hiemstra, M.; van Lier, P.A.; Meeus, W.; Branje, S.; Heim, C.M.; Nemeroff, C.B.; Mill, J.; et al. Genome-wide DNA methylation levels and altered cortisol stress reactivity following childhood trauma in humans. Nat. Commun. 2016, 7, 10967. [Google Scholar] [CrossRef]
  77. Nätt, D.; Johansson, I.; Faresjö, T.; Ludvigsson, J.; Thorsell, A. High cortisol in 5-year-old children causes loss of DNA methylation in SINE retrotransposons: A possible role for ZNF263 in stress-related diseases. Clin. Epigenetics 2015, 7, 91. [Google Scholar] [CrossRef]
  78. Volk, N.; Pape, J.C.; Engel, M.; Zannas, A.S.; Cattane, N.; Cattaneo, A.; Binder, E.B.; Chen, A. Amygdalar MicroRNA-15a is Essential for Coping with Chronic Stress. Cell Rep. 2016, 17, 1882–1891. [Google Scholar] [CrossRef]
  79. Boersma, G.J.; Lee, R.S.; Cordner, Z.A.; Ewald, E.R.; Purcell, R.H.; Moghadam, A.A.; Tamashiro, K.L. Prenatal stress decreases Bdnf expression and increases methylation of Bdnf exon IV in rats. Epigenetics 2014, 9, 437–447. [Google Scholar] [CrossRef]
  80. Bockmühl, Y.; Patchev, A.V.; Madejska, A.; Hoffmann, A.; Sousa, J.C.; Sousa, N.; Holsboer, F.; Almeida, O.F.; Spengler, D. Methylation at the CpG island shore region upregulates Nr3c1 promoter activity after early-life stress. Epigenetics 2015, 10, 247–257. [Google Scholar] [CrossRef]
  81. Kember, R.L.; Dempster, E.L.; Lee, T.H.; Schalkwyk, L.C.; Mill, J.; Fernandes, C. Maternal separation is associated with strain-specific responses to stress and epigenetic alterations to Nr3c1, Avp, and Nr4a1 in mouse. Brain Behav. 2012, 2, 455–467. [Google Scholar] [CrossRef]
  82. Franklin, T.B.; Russig, H.; Weiss, I.C.; Gräff, J.; Linder, N.; Michalon, A.; Vizi, S.; Mansuy, I.M. Epigenetic transmission of the impact of early stress across generations. Biol. Psychiatry 2010, 68, 408–415. [Google Scholar] [CrossRef] [PubMed]
  83. Monteleone, M.C.; Adrover, E.; Pallarés, M.E.; Antonelli, M.C.; Frasch, A.C.; Brocco, M.A. Prenatal stress changes the glycoprotein GPM6A gene expression and induces epigenetic changes in rat offspring brain. Epigenetics 2014, 9, 152–160. [Google Scholar] [CrossRef]
  84. Fransquet, P.D.; Hjort, L.; Rushiti, F.; Wang, S.J.; Krasniqi, S.P.; Çarkaxhiu, S.I.; Arifaj, D.; Xhemaili, V.D.; Salihu, M.; Leku, N.A.; et al. DNA methylation in blood cells is associated with cortisol levels in offspring of mothers who had prenatal post-traumatic stress disorder. Stress Health 2022, 38, 755–766. [Google Scholar] [CrossRef]
  85. D’Addario, S.L.; Municchi, D.; Mancini, C.; Ielpo, D.; Babicola, L.; Di Segni, M.; Iacono, L.L.; Ferlazzo, F.; Cifani, C.; Andolina, D.; et al. The long-lasting effects of early life adversities are sex dependent: The signature of miR-34a. J. Affect. Disord. 2022. [Google Scholar] [CrossRef]
  86. Grasso, D.J.; Drury, S.; Briggs-Gowan, M.; Johnson, A.; Ford, J.; Lapidus, G.; Scranton, V.; Abreu, C.; Covault, J. Adverse childhood experiences, posttraumatic stress, and FKBP5 methylation patterns in postpartum women and their newborn infants. Psychoneuroendocrinology 2020, 114, 104604. [Google Scholar] [CrossRef] [PubMed]
  87. Mulder, R.H.; Walton, E.; Neumann, A.; Houtepen, L.C.; Felix, J.F.; Bakermans-Kranenburg, M.J.; Suderman, M.; Tiemeier, H.; van IJzendoorn, M.H.; Relton, C.L.; et al. Epigenomics of being bullied: Changes in DNA methylation following bullying exposure. Epigenetics 2020, 15, 750–764. [Google Scholar] [CrossRef] [PubMed]
  88. van der Knaap, L.J.; Riese, H.; Hudziak, J.J.; Verbiest, M.M.; Verhulst, F.C.; Oldehinkel, A.J.; van Oort, F.V. Glucocorticoid receptor gene (NR3C1) methylation following stressful events between birth and adolescence. The TRAILS study. Transl. Psychiatry 2014, 4, e381. [Google Scholar] [CrossRef]
  89. Papale, L.A.; Madrid, A.; Li, S.; Alisch, R.S. Early-life stress links 5-hydroxymethylcytosine to anxiety-related behaviors. Epigenetics 2017, 12, 264–276. [Google Scholar] [CrossRef]
  90. Silva, A.S.; Toffoli, L.V.; Estrada, V.B.; Veríssimo, L.F.; Francis-Oliveira, J.; Moreira, E.G.; Gomes, M.V.; Pelosi, G.G. Maternal exposure to fluoxetine during gestation and lactation induces long lasting changes in the DNA methylation profile of offspring’s brain and affects the social interaction of rat. Brain Res. Bull. 2018, 142, 409–413. [Google Scholar] [CrossRef]
  91. Kertes, D.A.; Bhatt, S.S.; Kamin, H.S.; Hughes, D.A.; Rodney, N.C.; Mulligan, C.J. BNDF methylation in mothers and newborns is associated with maternal exposure to war trauma. Clin. Epigenetics 2017, 9, 68. [Google Scholar] [CrossRef]
  92. Martin, C.A.; Vorn, R.; Schrieber, M.; Lai, C.; Yun, S.; Kim, H.S.; Gill, J. Identification of DNA Methylation Changes That Predict Onset of Post-traumatic Stress Disorder and Depression Following Physical Trauma. Front. Neurosci. 2021, 15, 738347. [Google Scholar] [CrossRef] [PubMed]
  93. Misiak, B.; Samochowiec, J.; Konopka, A.; Gawrońska-Szklarz, B.; Beszłej, J.A.; Szmida, E.; Karpiński, P. Clinical Correlates of the NR3C1 Gene Methylation at Various Stages of Psychosis. Int. J. Neuropsychopharmacol. 2021, 24, 322–332. [Google Scholar] [CrossRef]
  94. Misiak, B.; Szmida, E.; Karpiński, P.; Loska, O.; Sąsiadek, M.M.; Frydecka, D. Lower LINE-1 methylation in first-episode schizophrenia patients with the history of childhood trauma. Epigenomics 2015, 7, 1275–1285. [Google Scholar] [CrossRef] [PubMed]
  95. Zhou, A.; Ancelin, M.-L.; Ritchie, K.; Ryan, J. Childhood adverse events and BDNF promoter methylation in later-life. Front. Psychiatry 2023, 14. [Google Scholar] [CrossRef] [PubMed]
  96. Bustamante, D.; Amstadter, A.B.; Pritikin, J.N.; Brick, T.R.; Neale, M.C. Associations between Traumatic Stress, Brain Volumes and Post-traumatic Stress Disorder Symptoms in Children: Data from the ABCD Study. Behav. Genet. 2022, 52, 75–91. [Google Scholar] [CrossRef] [PubMed]
  97. Klinger-König, J.; Hertel, J.; Van der Auwera, S.; Frenzel, S.; Pfeiffer, L.; Waldenberger, M.; Golchert, J.; Teumer, A.; Nauck, M.; Homuth, G.; et al. Methylation of the FKBP5 gene in association with FKBP5 genotypes, childhood maltreatment and depression. Neuropsychopharmacology 2019, 44, 930–938. [Google Scholar] [CrossRef]
  98. Gerhart, J.I.; Canetti, D.; Hobfoll, S.E. Traumatic stress in overview: Definition, context, scope, and long-term outcomes. In Traumatic Stress and Long-Term Recovery; Springer: Berlin/Heidelberg, Germany, 2015; pp. 3–24. [Google Scholar]
  99. Logue, M.W.; Miller, M.W.; Wolf, E.J.; Huber, B.R.; Morrison, F.G.; Zhou, Z.; Zheng, Y.; Smith, A.K.; Daskalakis, N.P.; Ratanatharathorn, A.; et al. An epigenome-wide association study of posttraumatic stress disorder in US veterans implicates several new DNA methylation loci. Clin. Epigenetics 2020, 12, 46. [Google Scholar] [CrossRef]
  100. Katrinli, S.; Maihofer, A.X.; Wani, A.H.; Pfeiffer, J.R.; Ketema, E.; Ratanatharathorn, A.; Baker, D.G.; Boks, M.P.; Geuze, E.; Kessler, R.C.; et al. Epigenome-wide meta-analysis of PTSD symptom severity in three military cohorts implicates DNA methylation changes in genes involved in immune system and oxidative stress. Mol. Psychiatry 2022, 27, 1720–1728. [Google Scholar] [CrossRef]
  101. Pape, J.C.; Carrillo-Roa, T.; Rothbaum, B.O.; Nemeroff, C.B.; Czamara, D.; Zannas, A.S.; Iosifescu, D.; Mathew, S.J.; Neylan, T.C.; Mayberg, H.S.; et al. DNA methylation levels are associated with CRF(1) receptor antagonist treatment outcome in women with post-traumatic stress disorder. Clin. Epigenetics 2018, 10, 136. [Google Scholar] [CrossRef]
  102. Uddin, M.; Galea, S.; Chang, S.C.; Aiello, A.E.; Wildman, D.E.; de los Santos, R.; Koenen, K.C. Gene expression and methylation signatures of MAN2C1 are associated with PTSD. Dis. Markers 2011, 30, 111–121. [Google Scholar] [CrossRef]
  103. Hammamieh, R.; Chakraborty, N.; Gautam, A.; Muhie, S.; Yang, R.; Donohue, D.; Kumar, R.; Daigle, B.J., Jr.; Zhang, Y.; Amara, D.A.; et al. Whole-genome DNA methylation status associated with clinical PTSD measures of OIF/OEF veterans. Transl. Psychiatry 2017, 7, e1169. [Google Scholar] [CrossRef] [PubMed]
  104. Rutten, B.P.F.; Vermetten, E.; Vinkers, C.H.; Ursini, G.; Daskalakis, N.P.; Pishva, E.; de Nijs, L.; Houtepen, L.C.; Eijssen, L.; Jaffe, A.E.; et al. Longitudinal analyses of the DNA methylome in deployed military servicemen identify susceptibility loci for post-traumatic stress disorder. Mol. Psychiatry 2018, 23, 1145–1156. [Google Scholar] [CrossRef]
  105. Rusiecki, J.A.; Byrne, C.; Galdzicki, Z.; Srikantan, V.; Chen, L.; Poulin, M.; Yan, L.; Baccarelli, A. PTSD and DNA Methylation in Select Immune Function Gene Promoter Regions: A Repeated Measures Case-Control Study of U.S. Military Service Members. Front. Psychiatry 2013, 4, 56. [Google Scholar] [CrossRef]
  106. Wang, B.; Suen, C.W.; Ma, H.; Wang, Y.; Kong, L.; Qin, D.; Lee, Y.W.W.; Li, G. The Roles of H19 in Regulating Inflammation and Aging. Front. Immunol. 2020, 11, 579687. [Google Scholar] [CrossRef] [PubMed]
  107. Lemmens, K.; Doggen, K.; De Keulenaer, G.W. Role of neuregulin-1/ErbB signaling in cardiovascular physiology and disease: Implications for therapy of heart failure. Circulation 2007, 116, 954–960. [Google Scholar] [CrossRef]
  108. Branham, E.M.; McLean, S.A.; Deliwala, I.; Mauck, M.C.; Zhao, Y.; McKibben, L.A.; Lee, A.; Spencer, A.B.; Zannas, A.S.; Lechner, M.; et al. CpG Methylation Levels in HPA Axis Genes Predict Chronic Pain Outcomes Following Trauma Exposure. J. Pain 2023. [Google Scholar] [CrossRef]
  109. Zhang, Y.; Zhao, G.; Han, Y.; Zhang, J.; Cao, C.; Wang, L.; Zhang, K. The mechanisms of histone modification in post-traumatic stress disorder. Adv. Psychol. Sci. 2022, 30, 98. [Google Scholar] [CrossRef]
  110. Wolf, E.J.; Maniates, H.; Nugent, N.; Maihofer, A.X.; Armstrong, D.; Ratanatharathorn, A.; Ashley-Koch, A.E.; Garrett, M.; Kimbrel, N.A.; Lori, A.; et al. Traumatic stress and accelerated DNA methylation age: A meta-analysis. Psychoneuroendocrinology 2018, 92, 123–134. [Google Scholar] [CrossRef]
  111. Snijders, C.; Maihofer, A.X.; Ratanatharathorn, A.; Baker, D.G.; Boks, M.P.; Geuze, E.; Jain, S.; Kessler, R.C.; Pishva, E.; Risbrough, V.B.; et al. Longitudinal epigenome-wide association studies of three male military cohorts reveal multiple CpG sites associated with post-traumatic stress disorder. Clin. Epigenetics 2020, 12, 11. [Google Scholar] [CrossRef]
  112. Yehuda, R.; Flory, J.D.; Bierer, L.M.; Henn-Haase, C.; Lehrner, A.; Desarnaud, F.; Makotkine, I.; Daskalakis, N.P.; Marmar, C.R.; Meaney, M.J. Lower methylation of glucocorticoid receptor gene promoter 1F in peripheral blood of veterans with posttraumatic stress disorder. Biol. Psychiatry 2015, 77, 356–364. [Google Scholar] [CrossRef]
  113. Bam, M.; Yang, X.; Busbee, B.P.; Aiello, A.E.; Uddin, M.; Ginsberg, J.P.; Galea, S.; Nagarkatti, P.S.; Nagarkatti, M. Increased H3K4me3 methylation and decreased miR-7113-5p expression lead to enhanced Wnt/β-catenin signaling in immune cells from PTSD patients leading to inflammatory phenotype. Mol. Med. 2020, 26, 110. [Google Scholar] [CrossRef]
  114. Kim, T.Y.; Kim, S.J.; Chung, H.G.; Choi, J.H.; Kim, S.H.; Kang, J.I. Epigenetic alterations of the BDNF gene in combat-related post-traumatic stress disorder. Acta Psychiatr. Scand. 2017, 135, 170–179. [Google Scholar] [CrossRef]
  115. Ressler, K.J.; Mercer, K.B.; Bradley, B.; Jovanovic, T.; Mahan, A.; Kerley, K.; Norrholm, S.D.; Kilaru, V.; Smith, A.K.; Myers, A.J.; et al. Post-traumatic stress disorder is associated with PACAP and the PAC1 receptor. Nature 2011, 470, 492–497. [Google Scholar] [CrossRef] [PubMed]
  116. Nawijn, L.; Krzyzewska, I.M.; van Zuiden, M.; Henneman, P.; Koch, S.B.J.; Mul, A.N.; Frijling, J.L.; Veltman, D.J.; Mannens, M.; Olff, M. Oxytocin receptor gene methylation in male and female PTSD patients and trauma-exposed controls. Eur. Neuropsychopharmacol. 2019, 29, 147–155. [Google Scholar] [CrossRef] [PubMed]
  117. Koenen, K.C.; Uddin, M.; Chang, S.C.; Aiello, A.E.; Wildman, D.E.; Goldmann, E.; Galea, S. SLC6A4 methylation modifies the effect of the number of traumatic events on risk for posttraumatic stress disorder. Depress. Anxiety 2011, 28, 639–647. [Google Scholar] [CrossRef]
  118. Haghighi, F.; Xin, Y.; Chanrion, B.; O’Donnell, A.H.; Ge, Y.; Dwork, A.J.; Arango, V.; Mann, J.J. Increased DNA methylation in the suicide brain. Dialogues Clin. Neurosci. 2014, 16, 430–438. [Google Scholar] [CrossRef] [PubMed]
  119. Schneider, E.; El Hajj, N.; Müller, F.; Navarro, B.; Haaf, T. Epigenetic Dysregulation in the Prefrontal Cortex of Suicide Completers. Cytogenet. Genome Res. 2015, 146, 19–27. [Google Scholar] [CrossRef]
  120. Nagy, C.; Suderman, M.; Yang, J.; Szyf, M.; Mechawar, N.; Ernst, C.; Turecki, G. Astrocytic abnormalities and global DNA methylation patterns in depression and suicide. Mol. Psychiatry 2015, 20, 320–328. [Google Scholar] [CrossRef]
  121. Ropret, S.; Kouter, K.; Zupanc, T.; Videtic Paska, A. BDNF methylation and mRNA expression in brain and blood of completed suicides in Slovenia. World J. Psychiatry 2021, 11, 1301–1313. [Google Scholar] [CrossRef]
  122. Maussion, G.; Yang, J.; Suderman, M.; Diallo, A.; Nagy, C.; Arnovitz, M.; Mechawar, N.; Turecki, G. Functional DNA methylation in a transcript specific 3′UTR region of TrkB associates with suicide. Epigenetics 2014, 9, 1061–1070. [Google Scholar] [CrossRef]
  123. Keller, S.; Sarchiapone, M.; Zarrilli, F.; Tomaiuolo, R.; Carli, V.; Angrisano, T.; Videtic, A.; Amato, F.; Pero, R.; di Giannantonio, M.; et al. TrkB gene expression and DNA methylation state in Wernicke area does not associate with suicidal behavior. J. Affect. Disord. 2011, 135, 400–404. [Google Scholar] [CrossRef] [PubMed]
  124. Keller, S.; Sarchiapone, M.; Zarrilli, F.; Videtic, A.; Ferraro, A.; Carli, V.; Sacchetti, S.; Lembo, F.; Angiolillo, A.; Jovanovic, N.; et al. Increased BDNF promoter methylation in the Wernicke area of suicide subjects. Arch. Gen. Psychiatry 2010, 67, 258–267. [Google Scholar] [CrossRef]
  125. Jamshidi, E.; Boström, A.E.D.; Wilczek, A.; Nilsonne, Å.; Åsberg, M.; Jokinen, J. Increased Methylation of Brain-Derived Neurotrophic Factor (BDNF) Is Related to Emotionally Unstable Personality Disorder and Severity of Suicide Attempt in Women. Cells 2023, 12, 350. [Google Scholar] [CrossRef]
  126. Haghighi, F.; Galfalvy, H.; Echen, S.; Ehuang, Y.-Y.; Cooper, T.B.; Burke, A.K.; Oquendo, M.A.; Emann, J.J.; Esublette, M.E. DNA Methylation Perturbations in Genes Involved in Polyunsaturated Fatty Acid Biosynthesis Associated with Depression and Suicide Risk. Front. Neurol. 2015, 6, 92. [Google Scholar] [CrossRef] [PubMed]
  127. Gaine, M.E.; Seifuddin, F.; Sabunciyan, S.; Lee, R.S.; Benke, K.S.; Monson, E.T.; Zandi, P.P.; Potash, J.B.; Willour, V.L. Differentially methylated regions in bipolar disorder and suicide. Am. J. Med. Genet. B Neuropsychiatr. Genet. 2019, 180, 496–507. [Google Scholar] [CrossRef] [PubMed]
  128. Murphy, T.M.; Crawford, B.; Dempster, E.L.; Hannon, E.; Burrage, J.; Turecki, G.; Kaminsky, Z.; Mill, J. Methylomic profiling of cortex samples from completed suicide cases implicates a role for PSORS1C3 in major depression and suicide. Transl. Psychiatry 2017, 7, e989. [Google Scholar] [CrossRef]
  129. Warrener, C.D.; Valentin, E.M.; Gallin, C.; Richey, L.; Ross, D.B.; Hood, C.J.; Lori, A.; Cubells, J.; Rauch, S.A.M.; Rilling, J.K. The role of oxytocin signaling in depression and suicidality in returning war veterans. Psychoneuroendocrinology 2021, 126, 105085. [Google Scholar] [CrossRef]
  130. Boström, A.E.D.; Jamshidi, E.; Manu, D.-M.; Kular, L.; Schiöth, H.B.; Åsberg, M.; Jokinen, J. Epigenetic changes in the CYP2D6 gene are related to severity of suicide attempt: A cross-sectional study of suicide attempters. J. Psychiatr. Res. 2023, 160, 217–224. [Google Scholar] [CrossRef]
  131. Kouter, K.; Zupanc, T.; Videtič Paska, A. Genome-wide DNA methylation in suicide victims revealing impact on gene expression. J. Affect. Disord. 2019, 253, 419–425. [Google Scholar] [CrossRef]
  132. Kouter, K.; Zupanc, T.; Videtič Paska, A. Targeted sequencing approach: Comprehensive analysis of DNA methylation and gene expression across blood and brain regions in suicide victims. World J. Biol. Psychiatry 2022, 24, 12–23. [Google Scholar] [CrossRef]
  133. Rizavi, H.S.; Khan, O.S.; Zhang, H.; Bhaumik, R.; Grayson, D.R.; Pandey, G.N. Methylation and expression of glucocorticoid receptor exon-1 variants and FKBP5 in teenage suicide-completers. Transl. Psychiatry 2023, 13, 1–9. [Google Scholar] [CrossRef]
  134. Zhu, J.-H.; Bo, H.-H.; Liu, B.-P.; Jia, C.-X. The associations between DNA methylation and depression: A systematic review and meta-analysis. J. Affect. Disord. 2023, 327, 439–450. [Google Scholar] [CrossRef]
  135. Policicchio, S.; Washer, S.; Viana, J.; Iatrou, A.; Burrage, J.; Hannon, E.; Turecki, G.; Kaminsky, Z.; Mill, J.; Dempster, E.L.; et al. Genome-wide DNA methylation meta-analysis in the brains of suicide completers. Transl. Psychiatry 2020, 10, 1–13. [Google Scholar] [CrossRef] [PubMed]
  136. Matrisciano, F.; Pinna, G. PPAR-α Hypermethylation in the Hippocampus of Mice Exposed to Social Isolation Stress is Associated with Enhanced Neuroinflammation and Aggressive Behavior. Int. J. Mol. Sci. 2021, 22, 10678. [Google Scholar] [CrossRef] [PubMed]
  137. Nestler, E.J. Transgenerational Epigenetic Contributions to Stress Responses: Fact or Fiction? PLoS Biol. 2016, 14, e1002426. [Google Scholar] [CrossRef]
  138. Skinner, M.K. Environmental stress and epigenetic transgenerational inheritance. BMC Med. 2014, 12, 153. [Google Scholar] [CrossRef] [PubMed]
  139. Švorcová, J. Transgenerational Epigenetic Inheritance of Traumatic Experience in Mammals. Genes 2023, 14, 120. [Google Scholar] [CrossRef]
  140. Mychasiuk, R.; Harker, A.; Ilnytskyy, S.; Gibb, R. Paternal stress prior to conception alters DNA methylation and behaviour of developing rat offspring. Neuroscience 2013, 241, 100–105. [Google Scholar] [CrossRef]
  141. Guidi, J.; Lucente, M.; Sonino, N.; Fava, G.A. Allostatic Load and Its Impact on Health: A Systematic Review. Psychother. Psychosom. 2021, 90, 11–27. [Google Scholar] [CrossRef] [PubMed]
  142. Caradonna, S.G.; Paul, M.R.; Marrocco, J. An allostatic epigenetic memory on chromatin footprints after double-hit acute stress. Neurobiol. Stress 2022, 20, 100475. [Google Scholar] [CrossRef]
  143. Chitrala, K.N.; Hernandez, D.G.; Nalls, M.A.; Mode, N.A.; Zonderman, A.B.; Ezike, N.; Evans, M.K. Race-specific alterations in DNA methylation among middle-aged African Americans and Whites with metabolic syndrome. Epigenetics 2020, 15, 462–482. [Google Scholar] [CrossRef] [PubMed]
  144. Ferioli, M.; Zauli, G.; Maiorano, P.; Milani, D.; Mirandola, P.; Neri, L.M. Role of physical exercise in the regulation of epigenetic mechanisms in inflammation, cancer, neurodegenerative diseases, and aging process. J. Cell Physiol. 2019, 234, 14852–14864. [Google Scholar] [CrossRef]
  145. Harkess, K.N.; Ryan, J.; Delfabbro, P.H.; Cohen-Woods, S. Preliminary indications of the effect of a brief yoga intervention on markers of inflammation and DNA methylation in chronically stressed women. Transl. Psychiatry 2016, 6, e965. [Google Scholar] [CrossRef]
  146. Brown, K.M.; Hui, Q.; Huang, Y.; Taylor, J.Y.; Prescott, L.; de Mendoza, V.B.; Crusto, C.; Sun, Y.V. Association between Stress and Coping with DNA Methylation of Blood Pressure-Related Genes Among African American Women. Chronic Stress 2019, 3, 2470547019879088. [Google Scholar] [CrossRef] [PubMed]
  147. Yehuda, R.; Daskalakis, N.P.; Desarnaud, F.; Makotkine, I.; Lehrner, A.L.; Koch, E.; Flory, J.D.; Buxbaum, J.D.; Meaney, M.J.; Bierer, L.M. Epigenetic Biomarkers as Predictors and Correlates of Symptom Improvement following Psychotherapy in Combat Veterans with PTSD. Front. Psychiatry 2013, 4, 118. [Google Scholar] [CrossRef] [PubMed]
  148. Wilker, S.; Vukojevic, V.; Schneider, A.; Pfeiffer, A.; Inerle, S.; Pauly, M.; Elbert, T.; Papassotiropoulos, A.; de Quervain, D.; Kolassa, I.-T. Epigenetics of traumatic stress: The association of NR3C1 methylation and posttraumatic stress disorder symptom changes in response to narrative exposure therapy. Transl. Psychiatry 2023, 13, 1–7. [Google Scholar] [CrossRef]
  149. Lewis, C.R.; Tafur, J.; Spencer, S.; Green, J.M.; Harrison, C.; Kelmendi, B.; Rabin, D.M.; Yehuda, R.; Yazar-Klosinski, B.; Cahn, B.R. Pilot study suggests DNA methylation of the glucocorticoid receptor gene (NR3C1) is associated with MDMA-assisted therapy treatment response for severe PTSD. Front. Psychiatry 2023, 14, 101. [Google Scholar] [CrossRef]
  150. Venditti, S.; Verdone, L.; Reale, A.; Vetriani, V.; Caserta, M.; Zampieri, M. Molecules of Silence: Effects of Meditation on Gene Expression and Epigenetics. Front. Psychol. 2020, 11, 1767. [Google Scholar] [CrossRef]
Figure 1. Stress-induced epigenetic changes associated with Nr3c1, BDNF, OXTR, and FKBP5. Stressors are color-coded with their respective epigenetic changes in the various tissue samples analyzed. Blue represents acute stress, red represents chronic stress, brown represents early childhood stress, green represents traumatic stress, and teal represents suicide. The simplistic gene diagram includes a promoter (blacked-out box), followed by an exon, (empty white box), and lastly by the intron (a squiggly line). Various abbreviations used include: H: humans, R: rats, M: mice, I: insula, Hip: hippocampus, Hi: hypothalamus, PC: prefrontal cortex, W: Wernicke’s area, A: amygdala, arrow up: increase in methylation, arrow down: decrease in methylation, tilde symbol: variable epigenetic changes, and numerical number corresponds to the source. The location of the arrow is not exact but will indicate whether the epigenetic change occurs in the promoter, intron, or exon region as shown in the key. An up arrow indicates increase in methylation and a down arrow indicates a decrease in methylation. If no methylation changes were observed the area was left blank. Created with BioRender.com (accessed on 12 March 2023).
Figure 1. Stress-induced epigenetic changes associated with Nr3c1, BDNF, OXTR, and FKBP5. Stressors are color-coded with their respective epigenetic changes in the various tissue samples analyzed. Blue represents acute stress, red represents chronic stress, brown represents early childhood stress, green represents traumatic stress, and teal represents suicide. The simplistic gene diagram includes a promoter (blacked-out box), followed by an exon, (empty white box), and lastly by the intron (a squiggly line). Various abbreviations used include: H: humans, R: rats, M: mice, I: insula, Hip: hippocampus, Hi: hypothalamus, PC: prefrontal cortex, W: Wernicke’s area, A: amygdala, arrow up: increase in methylation, arrow down: decrease in methylation, tilde symbol: variable epigenetic changes, and numerical number corresponds to the source. The location of the arrow is not exact but will indicate whether the epigenetic change occurs in the promoter, intron, or exon region as shown in the key. An up arrow indicates increase in methylation and a down arrow indicates a decrease in methylation. If no methylation changes were observed the area was left blank. Created with BioRender.com (accessed on 12 March 2023).
Cells 12 01258 g001
Table 1. Summary of epigenetic changes to genes associated with acute stress.
Table 1. Summary of epigenetic changes to genes associated with acute stress.
GeneEpigenetic ChangesStress TypeMethodModel OrganismTissue Location
[16]Whole-genome DNA methylationIncrease in methylation in concordance with increase in lymphocytesPsychosocial stressIllumina Infinium EPIC arrayHumanWhole blood samples
[18]3′UTR Nr3c1Increase in 5-hmCAcute restraint stressImmunohistochemistry, tandem mass-spec, TAB-seq analysisMale miceHippocampus
[20]Promoter region 1-C of Nr3c1Lower levels of methylation with lower stress reactivity, but failed to survive test corrections Psychosocial stressMethylation-sensitive polymerase chain reaction (PCR)HumanFasting blood samples
[21]Upstream of exon 2 in Nr3c1/Region around exon I7 Nr3c1No effect on CpG levels in both tissues/increase in methylation in dentate gyrus, decrease in Cornu Ammonis Swim stress testEZ 96-DNA methylation kitMale Wistar ratDentate gyrus and Cornu Ammonis
[22]Histone H3K9 at retro-transposable element lociH3K9me3Acute restraint stressChIP sequencing against H3K9me3Male Sprague Dawley ratsHippocampus
[23]Histone H3K7H3K27me3Acute heat stressNano-UPLC-ESI-Q-TOF-MS/MSMale Taiwan country chickensAdrenal gland
[24]Histone H3K9/H3K27Increase in H3K9me3 and decrease in H3K9me1 in both tissues/reduction in H3K27 me3 in both tissuesAcute restraint stressImmunohistochemistry analysisAdult male Sprague Dawley ratsDentate gyrus and CA1
[25]Histone H3 in promoter region of BDNFNo methylation or acetylation changeAcute restraint stressChIP assayMale miceHippocampus
[21]MiR-124aIncreased expressionForced swim testMicroRNA analysis targeted against miR-124aMale Wistar ratsDentate gyrus
[26]MiR-135a/miR-124Down-regulatedAcute restraint stress Microarray analysis and reverse transcription real-time PCRAdult male miceAmygdala
[27]CYP24A1, BRCA2, NOTCH2, FOXO3, GATA3MethylatedAcute exposure to UV lightInfinity methylation nEPIC array Female Caucasian humansSkin from lower back
[27]KRT17, CSNK2A2HypomethylatedAcute exposure to UV lightInfinity methylation nEPIC arrayFemale Caucasian humansSkin from lower back
[27]CARD14/IRF8Demethylated/hypermethylatedAcute exposure to UV lightInfinity methylation nEPIC arrayFemale Caucasian humansSkin from lower back
[28]OXTR1”/“OXTR2”/BDNF exon V, Vh, and ViIncrease in methylation/decrease in methylation/no methylation differencesTSSTSequenom EpiTYPERHuman Blood samples
[29]PRF1 CpG sites -776 and -774Increase in methylationTSSTQuantitative methylation analysis by Bisulfite-pyrosequencing Chronic fatigue syndrome humansBlood samples
[19]Enpp2/Sostdc1Hyper-DhMRAcute restraint stressGenome-wide map of 5-hmC via NEBNext ChIP-Seq and RNAseqMiceHippocampus
[19]Ulk4/Wnt9aHypo-DhMRAcute restraint stressGenome-wide map of 5-hmC via NEBNext ChIP-Seq and RNAseqMiceHippocampus
[19]Banp/Gadd45b/Cbfa2t3/Irs2/Klf15/Smtn/Spns2Hypo-DhMRAcute restraint stressGenome-wide map of 5-hmC via NEBNext ChIP-Seq and RNAseqMiceHippocampus
Table 2. Summary of epigenetic changes to genes associated with chronic stress.
Table 2. Summary of epigenetic changes to genes associated with chronic stress.
GeneEpigenetic ChangesStress TypeMethodModel OrganismTissue Location
[34]Nr3c1 promoter region of exon 17Increase in DNA methylationChronic water avoidance stress test (7 days)Sodium Bisulfite sequencing, pyrosequencingFemale Fischer-344 ratsCentral nucleus of the amygdala
[36]Nr3c1 promoter region of exon 17No individual site methylation changeChronic restraint test (1 h) Bisulfite sequencingMale Sprague Dawley ratsAdrenal and pituitary gland
[37]Nr3c1 promoter regionNo increase in DNA methylationSocial defeat test by rat exposureSodium Bisulfite sequencingMiceHippocampus
[37]NCAM/CHL1No DNA methylation changesSocial defeat test by rat exposureSodium Bisulfite sequencingMiceHippocampus
[38]ZBTB16HypermethylationProlonged exposure to glucocorticoids GenFind V3 DNA extraction and purification kit, Illumina Infinium HumanMethylationEPIC BeadChipHumansFetal lung fibroblast cells
[40]5-HT1A promoter (-691 CpG)Increase in DNA methylationChronic unpredictable mild stressBisulfite treated, PCR, followed by random selection of amplifications and DNA methylation sequencingMale micePrefrontal cortex and midbrain
[24]Histone H3H3K4me3 mild increase, H3K9me3 decreaseChronic restraint test (7 days)Immunohistochemistry analysisAdult male Sprague Dawley ratsDentate gyrus
[41]Histone H3H3K9 methylation increase, IL-6 up-regulationChronic water avoidance stress (10 consecutive days)Chip, quantitative PCR, Western blot, and immunofluorescenceYoung-adult male Sprague Dawley ratsColonic
[42]OXTR promoter regionIncreased DNA methylationAdult adversity (measured via Unmet Material Needs Scale and neighborhood crime)Illumina 450K Human Methylation BeadChipAfrican American Women Blood
[43]MiR-709/miR-186Down-regulation/up-regulationChronic restraint test (2 weeks)miRNA microarray analysis using µParaflo ® BiochipAdult male Long-Evans hooded ratsHippocampus and prefrontal cortex
[44]Drosha intron 9/chromosome X intergenic regionDecrease methylation/increased methylationChronic social defeat (14 days)Methyl-Seq followed by Bisulfite-pyrosequencing Adult male miceDentate gyrus
[45]Desmin upstream region/Tgfb1 downstream regionIncrease in DNA methylation/increase in DNA methylationChronic restraint test (4 weeks)Reduced representation Bisulfite sequencing and methylation-specific PCRMale miceHeart
[45]Ppp2r2c/Ppp2r1a/Prkca/Adra1bAlterations in DNA methylationChronic restraint test (4 weeks)Reduced representation Bisulfite sequencing and methylation-specific PCRMale miceHeart
[46]BDNF exon IV/TrkBIncreased methylation/increased methylation Forced swim test (21 days)One-way ANOVA, real-time PCR, and Western blottingMale and female Wistar ratsHippocampus
[47]BDNF CpG islands in exon 1 promoter regionLower DNA methylationChronic high job stressIllumina Infinium HumanMethylation 450 BeadChip arrayJapanese workersLeukocytes in saliva
[48]BDNF promoter regionDNA hypermethylation in rats with hyperhomocysteinemiaChronic unpredicted mild stressMethyl-Specific PCR, Sequenom Mass Spectrometry, and enzyme-linked immunosorbent assay RatsHippocampus
[49]SLC6A4 promoter regionDecrease in methylationChronic job stressMethylation 450K BeadChip and Bisulfite sequencingFemale nursesPeripheral blood leukocytes
[4]Non-promoter CRF/shore shelf site of SLC6A4DNA methylation/increased methylationChronic stress from living in disadvantaged neighborhoodsIllumina HumanMethylation450 BeadChipHumansB cells, T cells, Neutrophils, and natural killer cells found in blood
[4]Non-promoter F8/non-promoter TLR1DNA methylationChronic stress from living in disadvantaged neighborhoodsIllumina HumanMethylation450 BeadChipHumansB cells, T cells, Neutrophils, and natural killer cells found in blood
Table 3. Summary of epigenetic changes to genes associated with early childhood stress.
Table 3. Summary of epigenetic changes to genes associated with early childhood stress.
GeneEpigenetic ChangesStress TypeMethodModel OrganismTissue Location
[84]FRMD4A, CCDC174, FBXL2, CHD4Methylation status associated with cortisol levelsTorture and/or sexual assaultEpigenome-wide methylation via illuminates Infinium HumanMethylationEPIC BeadChipHuman mothersBlood
[82]MeCP2, CB1, CRFR2Methylation of the CpG island surrounding the transcription initiation site of MeCP2 and CB1 genes was increased/CRFR2 gene methylation in a stretch of the CpG island located 5′ of the transcription initiation site was decreased Chronic and unpredictable maternal separation from postnatal days 1–14Bisulfite-converted DNA followed by pyrosequencing analysis MiceGermline
[76]KITLG (cg27512205)MethylationChildhood traumaGenome-wide analysisHealthy humansBlood and buccal cells
[77]PRDM14HypomethylatedChildhood chronic stress assessed via hair cortisolHair cortisol levels in relationship to whole-genome DNA-methylation sequencing5-year-old childrenBlood and hair cortisol
[83]gpm6a intron 1/miRNA-133bNo overall differences in methylation, CpG 6 methylation, and CpG 7 not methylated/overexpression of miRNA suppressed gpm6a mRNARestraint test on mothers in last week of gestationBisulfite conversion and quantitative PCRMale offspring mice (postnatal days 28 and 60)Hippocampus and prefrontal cortex
[85]miRNA-34aIncreasedMaternal separation followed by repeated cross-fosteringQuantitative real-time PCRFemale miceMedial prefrontal cortex and dorsal raphe nuclei
[80]Nr3c1 CpG island shore regionHypomethylationMaternal separationBisulfite sequencingMale mice (postnatal day 0)Hypothalamic neurons
[86]FKBP5Methylation with CC geneAdverse childhood experiences Bisulfite pyrosequencing and DNA methylation analysisPostpartum women and babies (within 24 h)Saliva
[58]Nr3c1, FKBP5, BDNF, AVP, CRHR1, SLC6A4No robust epigenetic variationPhysical, emotional, or sexual abuse/neglect/intimate-partner violence/crime/bullying/cyber-victimizationBisulfite sequencing with Illumina Infinium HumanMethylation450 BeadChipHuman twin cohort (assessed at ages 5, 7, 10, 12, and 18)Peripheral blood
[87]RAB14Decrease in methylationChild bullyingEpigenome-wide methylation, Bisulfite conversion followed with Illumina Infinium HumanMethylation450 BeadChipChildrenBlood
[88]Nr3c1 amplicon 1Higher methylationPerinatal stress/stressful life events/traumatic youth experiencesBisulfite converted DNA, PCR, reverse transcription, cleavage of RNA product followed with mass spectrometry Children (mean age 16.1)Blood
[89] Nr3c2, Nrxn1, Nfia, and Clip1Variable methylation in adult female mice who had experienced early life stressEarly life stress before weaning (postnatal days 12 to 18)Chemical labeling-based 5hmC enrichmentFemale mice 3-month-oldHypothalamus
[79]BDNF exon IVIncrease methylationPrenatal stressBisulfite pyrosequencing genomic DNAMale offspring of pregnant Sprague Dawley ratsAmygdala and hippocampus
[90] Global DNA methylationIncrease in DNA methylation in hippocampus Medication used for antidepressants Imprint methylated DNA quantification KitFemale Wistar ratsHippocampus, cortex, hypothalamus, and periaqueductal gray matter
[91]BDNFIncrease in DNA methylation with trauma exposureWar traumaHumanMethylation450 BeadChipMothers and newborns in the eastern democratic republic of Congo Umbilical cord blood, placental tissue, maternal venous blood
[53]Nr3c1Increase in DNA methylationEarly childhood abuseBi-sulfite sequencingWomen from Black Women’s Health Societyleukocytes
[54] Nr3c1Increase in promoter methylationEarly childhood abuseBi-sulfite sequencingPost-mortem suicide victims of childhood abuseHippocampus sections post-mortem
[92] Nr3c1Increase in DNA methylationEarly childhood physical abuseBi-sulfite sequencingSubjects with Borderline Personality DisorderPeripheral Blood Cells
[55]Nr3c1Increase in promoter methylationLack of adequate nurturing, as measured by parental loss, childhood maltreatment, and parental careBi-sulfite sequencingHealthy adultsLeukocytes
[93]Nr3c1Lower Nr3c1 methylationSchizophrenia patients with ACESPyrosequencingHumanLeukocytes
[59]Nr3c1 CpG sitesIncreased methylationTSST for an acute stressPyrosequencingHumanSaliva
[60]FKBP5Demethylation in functional glucocorticoid response elements of FKBP5 geneChildhood traumaBisulfite pyrosequencingHumanWhole blood cells
[94]LINE1Lower LINE1 methylationSchizophrenia patients with ACESPyrosequencingHumanWhole blood leukocytes
[61]FKBP5Decreased levels of DNA methylationmother-infant dyads shortly after parturitionMass array spectrometryHumanImmune cells from blood
[56]FKBP5ACES associated with lower methylation levels at CpG site Patients with psychotic disorders (with and without ACES)PyrosequencingHumanPeripheral Blood leukocytes
[63]BDNFHypermethylation of promoterMaltreatment compared to positive caregivingMethylation-specific real-time PCR (MSP) or direct bisulfite DNA sequencing PCR
(BSP) on bisulfite-modified DNA (Chemicon or Qiagen), or via methylated DNA
immunoprecipitation using an antibody against 5-methylcytosine
RatsPrefrontal cortex and
hippocampus
[95]BDNF promoter I and IVNo long-term epigenetic changes Early life abuse, experiencing a war or natural disaster, and povertySequenom MassArrayHumansBlood and buccal tissue
[66]OXTRHypermethylation of CpG sites predictive for pathology393 African American adults with or without childhood abuseBisulfite sequencingHumanWhole blood
[67]OXTRELA exposure was associated with one significant CpG site in the first intron among females, but not among males46 adults (23 males/23 females) with varying degrees of childhood adverse eventsPyrosequencingHumanWhole blood
[70]SLC64ASignificant effect of Child Abuse on overall methylation levels192 (96 males, 96 females from Iowa Adoption Study) with varying levels of childhood abuse experiencesBisulfite conversionHumanlymphoblasts
[71]SLC64AExamination of these four CpG residues indicated that methylation of cg22584138 was influenced by both genotype and sex abuse, whereas methylation of cg05016953 was influenced only by sex abuse history158 female subjects in the Iowa Adoption Studies Illumina HumanMethylation450 BeadChip Humanlymphoblasts
[72]SLC64AHigher promoter methylation status was significantly associated with childhood adversities108 patients with major depressive disordersBisulfite sequencingHumanLeukocytes
[73]SLC64AChildhood trauma, being male, and smaller hippocampal volume were independently associated with greater peripheral serotonin transporter methylationThirty-three adults with Major Depressive Disorder (MDD) (23 females) and 36 matched healthy controls (21 females) were included in the studyPyrosequencingHumansWhole Blood
[74]Whole genome362 differentially methylated promoters in individuals with a history of abuse compared with controls. Among these promoters, 248 showed hypermethylation and 114 demonstrated hypomethylation, highest methylation difference in ALS2 gene41 French-Canadian men (25 with a history of severe childhood abuse and 16 control subjects)Methylated DNA
immunoprecipitation (meDIP) method
HumansCingulate cortex
[75]Whole genomeA history of child abuse was associated with cell type-specific changes in DNA methylation of oligodendrocyte genes and a global impairment of the myelin-related transcriptional programPostmortem brain samples from human subjects (N = 78) and from a rodent model of the impact of early life environment (N = 24) were analyzedBisulfite sequencingHumansAmygdala/Post-mortem brains
[60]Whole genomeGenome-wide methylation evidence of distinct biological modifications in PTSD in the presence or absence of exposure to childhood abuse396 with 169 trauma-exposed individualsHumanMethylation 450k BeadChipHumansWhole Blood
[76]KIT4GNine DMRs replicated across cohorts, respectively associated with the ACE scoreTwo cohorts (mothers from the Avon Longitudinal Study of Parents and Children, ALSPAC, n  =  780 and women from the MRC National Survey of Health and Development, NSHD, n  =  552HumanMethylation450 BeadChipHumansPeripheral Blood
[87]RAB14One site, cg17312179, showed small changes in DNA methylation associated with bullying exposure and RAB14 methylation levels decreased for exposed but increased for nonexposedPopulation-based Generation R Study and Avon Longitudinal Study of Parents and Children (combined n = 1352)Bisulfite sequencingHumansPeripheral Blood
[78]mIR-15aIncreased levels of childhood stressChildhood adversityAffymetrix miRNA 2.0 arrayHumansBlood cells
[81]AVPDecrease methylationMaternal separationBisulfite sequencingMicehippocampus
[96]Nr3c1CM was associated with an increase in DNA methylation in an EGR1 transcription factor binding siteChildhood maltreatment (CM)Pyrosequencing147 adult participants from the Detroit Neighborhood Health StudyWhole blood
[97]FKBP5No change in methylation for childhood abuseChildhood abusePyrosequencing3965 subjects of the Study of
Health in Pomerania
Whole blood
[57]Nr3c1Associations between DNA methylation and severity of fatigue as well as with childhood emotional abuse in CFS patients, although these findings were not significant after correction for multiple testingChildhood traumaBisulfite sequencing80 female CFS (chronic fatigue syndrome) patients and 91 female controlsPeripheral blood
[69]OXTRMethylation of both OXTR and OXT genes shaped the directionality of adversity effectsMemories of Childhood TraumaBisulfite sequencing81 womenBlood
[65]BDNF, Nr3c1, and MAN2C1Increased methylation with adverse childhood experiencesAdverse childhood experience [ACE] scorePyrosequencing70 active military members with and without PTSDPeripheral blood
Table 4. Summary of epigenetic changes to genes associated with traumatic stress.
Table 4. Summary of epigenetic changes to genes associated with traumatic stress.
GeneEpigenetic ChangesStress TypeMethodModel OrganismTissue Location
[100]F2R, CNPY2, BAIAP2L1, TBXAS1All sites showed lower DNA methylationPTSD severityMethylationEPIC BeadChipPTSD military personnelBlood
[92]UBE2L3Differentials methylated and promoter was hypomethylated in PTSD and MDD patientsTraumatic eventBisulfite sequencingPTSD with MDD, MDD alone, and no PTSD or MDDPeripheral blood
[101]CRHR1Increase in methylation from baseline in rs110402 GG allele individuals after treatment Traumatic eventIllumina 450K array and Bisulfite conversionWomen with PTSD Blood
[110]Epigenetic ageTraumatic stress associated with advanced epigenetic ageTraumatic eventIllumina Infinium HumanMethylation BeadChipHumansPeripheral blood
[102]MAN2C1Higher methylation had increased risk of PTSDTraumatic eventBisulfite conversion followed by Infinium HumanMethylation 27K BeadChipHumans from Detroit Blood
[104]ZFP57, RNF39, HIST1H2APS2Decrease in DNA methylation with increased PTSD symptomsCombat trauma leading to PTSDIllumina HumanMethylation 450 BeadChipDutch military cohort and male US marine cohortBlood
[74]Nr3c1 promoter regions 1B and 1CLower methylation levelsPTSD from trauma Sequence EpiTYPER, Sodium Bisulfite conversion Humans with PTSDT-lymphocyte isolated from blood
[99]GOS2 cg19534438/AHRR cg05575921Methylated at locus/decreased methylation at locusPTSD from traumaEWAS using Illumina EPIC methylation BeadChipHumans with PTSDBlood samples and prefrontal cortex
[105]H19 and IL18Those who did not develop PTSD had
reduced %5-mC levels of H19 and IL18 after deployment
PTSD from combat traumaBisulfite treatment followed by pyrosequencingPost-deployment military individuals with PTSDBlood serum
[111]HEXDC rs4789774/MAD1L1 Development of combat-related PTSD is associated with distinct methylation patterns in HLA region, HEXDC, and MAD1L1PTSD from combat traumaHumanMethylation450 BeadChipThree cohorts of male military membersBlood
[91] CRH, CRHBP, Nr3c1, and FKBP5Methylation changes in offspring associated with war exposure in mothersWar traumaHumanMethylation450 BeadChipMothers and newborns in the eastern Democratic Republic of Congo Umbilical cord blood, placental tissue, maternal venous blood
[4]AHRRLower DNA methylation in PTSDTraumatic eventIllumina HumanMethylation450 BeadChipMilitary and civilian cohorts Blood
[60] SENP7Six genome-wide significant (GWS) CpG sites associated with past-month PTSD and three CpGs with lifetime PTSDPTSD from the Vietnam WarBisulfite converted analysis 1135 male European–American U.S. veterans who participated in the National Health and Resilience in Veterans Study (NHRVS)Semen and blood
[112]Nr3c1 1F promoter Lower methylation in veterans with PTSDPTSD from combat trauma Bisulfite mapping and colonial sequencingMale combat veteransPeripheral blood mononuclear cells
[113]H3K4me3 histone modification, WNT 10BIncrease in H3K4me3 around WNT 10B promoter in patients with PTSDPTSD from various traumatic stressorsRNA-seq, ChIP-seq, and microarrayPatients with PTSDPeripheral blood mononuclear cells
[102]NRG1, HGSTwo CpG sites significantly associated with current PTSD in NRG1 (cg23637605) and in HGS (cg19577098)PTSD from combat trauma Methylation microarrayPatients with PTSDWhole blood cells
[103]5600 CpG islandsMajority of CpG islands were hypermethylated in PTSD casesPTSD from combat traumaAgilent whole genome methylation array/targeted bisulfite sequencingOperation Enduring Freedom/Iraqi Freedom Combat veteransPeripheral whole blood cells
[114]BDNFSubjects with PTSD showed a higher DNA methylation of four CpG sites at the BDNF promoter compared with those without PTSDCombat veterans with PTSDPyrosequencingCombat veteransPeripheral blood cells
[115]ADCYAP1Methylation of ADCYAP1R1 is associated with PTSDPatients with PTSDHumanMethylation27 BeadChip Patients receiving services in the
primary care clinics at Grady Memorial Hospital
Whole blood
[116]OXTR exon 3CpG islands increase in methylation in female PTSDPTSDBisulfite converted analysis67 human subjects (31 PTSD, 36 controls)Whole blood
[117]SLC6A4 promoterNo association, increase in methylation demonstrated protective effectsPTSDHM27 BeadChip 100 human subjectsWhole blood
[108]POMC and CRHBPCpG sites demonstrated increased methylation as predictor for chronic post-traumatic musculoskeletal painPTSDBisulfite conversion and Illumina Infinium Human MethylationEPIC BeadChip arrayHumansBlood
Table 5. Summary of epigenetic changes to genes associated with suicide.
Table 5. Summary of epigenetic changes to genes associated with suicide.
GeneEpigenetic ChangesStress TypeMethodModel OrganismTissue Location
[120]GRIK2 intron 13/BEGAINHypomethylated/increase in methylationSuicide completion w/MDDGenome-wide methylation sequencing via MBD-SeqHumansCortical brain region
[121]BDNF exon 1Decrease in methylationSuicide completion via hangingBisulfite next-generation sequencingHumansBlood
[122]3′UTR TrkB-T1HypermethylationSuicide completion with low TrkB-T1Methylated DNA immunoprecipitation, labeling, and hybridization via microarrayHumansBA 8 and 9
[123]BDNF promoter region IVHypermethylationSuicide completionMassArray methylation analysisHumansWernicke area
[125]BDNF promoterIncreased methylationSuicidal patients with emotionally unstable personality disorderIllumine EPIC BeadChipHumans (women)Blood
[126]Elovl5 upstream/downstream to transcription start siteMethylation/lower CpG methylationSuicide non-completers with MDDBisulfite pyrosequencingHumansBlood
[127]ARHGEF38Hypermethylation across 4 CpG sitesBipolar disorder suicide completionSureSelect(XT) system, methyl-Seq, confirmation via pyrosequencingHumansBA 46
[128]PSORS1C3HypomethylationSuicide completion w/MDDBisulfite conversion followed by Infinium HumanMethylation450 BeadChip Array, then Bisulfite pyrosequencing confirmationHumansBA 11 and 25
[129]OXTR MT2 region (-901, -924, -934)No methylation at sitesSuicidalityMethylation assay via Bisulfite conversion followed by pyrosequencingHumans (male veterans)Saliva
[130]CYP2D6 CpG sitesHypomethylation (males) and hypermethylation (females)Severe suicide behaviorIllumina Infinium Methylation EPIC BeadChipHumans Peripheral blood
[131]ZNF714/NRIP3Hypomethylated/mixed methylation Suicide completion via hangingNext-generation sequencing of genome-wide methylation analysisHumansBA 9
[132]Nr3c1 1B promoterHigher methylation in the insula and blood, lower levels of methylation in hippocampusSuicide completionBisulfite next-generation sequencing Male HumansInsula, blood, and hippocampus
[133]Nr3c1 5′UTRDNA methylationSuicide completionDNA methylation enrichment assayTeenage humansHippocampus and prefrontal cortex
[132]SLC6A4_2 ampliconHypermethylatedSuicide completionBisulfite next-generation sequencingMale HumansBA 46
[132]5-HT1AHypomethylated in blood/hypermethylated in insulaSuicide completionBisulfite next-generation sequencingMale HumansBlood and insula
[132]SKA2Mixed methylation data in all tissuesSuicide completionBisulfite next-generation sequencingMale HumansHippocampus, insula, amygdala, BA 46, and blood
[132]MAOA_2 ampliconHypomethylated in insula and BA 46Suicide completionBisulfite next-generation sequencingMale HumansInsula and BA 46
[132]GABRA1Decreased methylation in hippocampus and blood/increased methylation in insulaSuicide completionBisulfite next-generation sequencingMale HumansHippocampus, blood, insula
[132]NRIP3 ampliconDecrease methylationSuicide completionBisulfite next-generation sequencingMale HumansHippocampus and insula
[135]CERC2 intronic regionHypermethylation across 4 CpG sitesSuicide completersIllumina HumanMethylation450K BeadChip or Infinium MethylationEPIC BeadChipHumanCerebellum
[136]PPAR-αIncreased methylation in CpG regionsSocial isolation (4 weeks) model for suicide-like behavior/PTSDMethyl-DNA immunoprecipitationMale Swiss-webster miceHippocampus
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Dee, G.; Ryznar, R.; Dee, C. Epigenetic Changes Associated with Different Types of Stressors and Suicide. Cells 2023, 12, 1258. https://doi.org/10.3390/cells12091258

AMA Style

Dee G, Ryznar R, Dee C. Epigenetic Changes Associated with Different Types of Stressors and Suicide. Cells. 2023; 12(9):1258. https://doi.org/10.3390/cells12091258

Chicago/Turabian Style

Dee, Garrett, Rebecca Ryznar, and Colton Dee. 2023. "Epigenetic Changes Associated with Different Types of Stressors and Suicide" Cells 12, no. 9: 1258. https://doi.org/10.3390/cells12091258

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop