ArticlePDF Available

Genome-Wide DNA Methylation Scan in Major Depressive Disorder

PLOS
PLOS ONE
Authors:

Abstract and Figures

While genome-wide association studies are ongoing to identify sequence variation influencing susceptibility to major depressive disorder (MDD), epigenetic marks, such as DNA methylation, which can be influenced by environment, might also play a role. Here we present the first genome-wide DNA methylation (DNAm) scan in MDD. We compared 39 postmortem frontal cortex MDD samples to 26 controls. DNA was hybridized to our Comprehensive High-throughput Arrays for Relative Methylation (CHARM) platform, covering 3.5 million CpGs. CHARM identified 224 candidate regions with DNAm differences >10%. These regions are highly enriched for neuronal growth and development genes. Ten of 17 regions for which validation was attempted showed true DNAm differences; the greatest were in PRIMA1, with 12–15% increased DNAm in MDD (p = 0.0002–0.0003), and a concomitant decrease in gene expression. These results must be considered pilot data, however, as we could only test replication in a small number of additional brain samples (n = 16), which showed no significant difference in PRIMA1. Because PRIMA1 anchors acetylcholinesterase in neuronal membranes, decreased expression could result in decreased enzyme function and increased cholinergic transmission, consistent with a role in MDD. We observed decreased immunoreactivity for acetylcholinesterase in MDD brain with increased PRIMA1 DNAm, non-significant at p = 0.08. While we cannot draw firm conclusions about PRIMA1 DNAm in MDD, the involvement of neuronal development genes across the set showing differential methylation suggests a role for epigenetics in the illness. Further studies using limbic system brain regions might shed additional light on this role.
Content may be subject to copyright.
Genome-Wide DNA Methylation Scan in Major
Depressive Disorder
Sarven Sabunciyan
1.
, Martin J. Aryee
2,5.
, Rafael A. Irizarry
1,3,5
, Michael Rongione
9
, Maree J. Webster
7
,
Walter E. Kaufman
3,6
, Peter Murakami
3
, Andree Lessard
8
, Robert H. Yolken
1
, Andrew P. Feinberg
3,4
,
James B. Potash
9
*, GenRED Consortium
"
1Department of Pediatrics, Stanley Division of Developmental Neurovirology, Bloomberg School of Public Health, Johns Hopkins University, Baltimore, Maryland, United
States of America, 2Department of Oncology, Bloomberg School of Public Health, Johns Hopkins University, Baltimore, Maryland, United States of America, 3Epigenetics
Center, Bloomberg School of Public Health, Johns Hopkins University, Baltimore, Maryland, United States of America, 4Division of Molecular Medicine, Department of
Medicine, Bloomberg School of Public Health, Johns Hopkins University, Baltimore, Maryland, United States of America, 5Department of Biostatistics, Bloomberg School
of Public Health, Johns Hopkins University, Baltimore, Maryland, United States of America, 6Center for Genetic Disorders of Cognition and Behavior, Kennedy Krieger
Institute, Baltimore, Maryland, United States of America, 7Uniformed Services University of the Health Sciences, Bethesda, Maryland, United States of America, 8Maryland
Psychiatric Research Center, University of Maryland School of Medicine, Baltimore, Maryland, United States of America, 9Department of Psychiatry, University of Iowa,
Iowa City, Iowa, United States of America
Abstract
While genome-wide association studies are ongoing to identify sequence variation influencing susceptibility to major
depressive disorder (MDD), epigenetic marks, such as DNA methylation, which can be influenced by environment, might
also play a role. Here we present the first genome-wide DNA methylation (DNAm) scan in MDD. We compared 39
postmortem frontal cortex MDD samples to 26 controls. DNA was hybridized to our Comprehensive High-throughput
Arrays for Relative Methylation (CHARM) platform, covering 3.5 million CpGs. CHARM identified 224 candidate regions with
DNAm differences .10%. These regions are highly enriched for neuronal growth and development genes. Ten of 17 regions
for which validation was attempted showed true DNAm differences; the greatest were in PRIMA1, with 12–15% increased
DNAm in MDD (p = 0.0002–0.0003), and a concomitant decrease in gene expression. These results must be considered pilot
data, however, as we could only test replication in a small number of additional brain samples (n = 16), which showed no
significant difference in PRIMA1. Because PRIMA1 anchors acetylcholinesterase in neuronal membranes, decreased
expression could result in decreased enzyme function and increased cholinergic transmission, consistent with a role in MDD.
We observed decreased immunoreactivity for acetylcholinesterase in MDD brain with increased PRIMA1 DNAm, non-
significant at p = 0.08. While we cannot draw firm conclusions about PRIMA1 DNAm in MDD, the involvement of neuronal
development genes across the set showing differential methylation suggests a role for epigenetics in the illness. Further
studies using limbic system brain regions might shed additional light on this role.
Citation: Sabunciyan S, Aryee MJ, Irizarry RA, Rongione M, Webster MJ, et al. (2012) Genome-Wide DNA Methylation Scan in Major Depressive Disorder. PLoS
ONE 7(4): e34451. doi:10.1371/journal.pone.0034451
Editor: Tadafumi Kato, RIKEN Brain Science Institution, Japan
Received December 7, 2011; Accepted February 28, 2012; Published April 12, 2012
Copyright: ß2012 Sabunciyan et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits
unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
Funding: This work was supported by a grant from the NIMH (National Institute of Mental Health) to Dr. Potash (R01MH074131), by the NHGRI (National Human
Genome Research Institute) to Dr. Feinberg (2P50HG003233), and by the Margaret Price Investigatorship, and the Stanley Medical Research Institute (SMRI). The
funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
Competing Interests: The authors have declared that no competing interests exist.
* E-mail: james-potash@uiowa.edu
.These authors contributed equally to this work.
"Membership of the GenRED Consortium is provided in the Acknowledgments.
Introduction
Family studies show that siblings of probands with major
depressive disorder (MDD) have about a three-fold elevated risk of
illness, while the estimated heritability of MDD from twin studies
is about 37% [1]. The modest level of heritability suggests that the
DNA sequence does not fully explain the variability in suscepti-
bility to this illness. Indeed, genome-wide association studies have
not yet definitively identified variants implicated in MDD, though
some intriguing results have been reported [2].
There are at least two other major kinds of explanations for this
variation in susceptibility. One is that environmental factors such
as stressful life events play a significant role in triggering MDD [3],
and another is that epigenetic factors are involved. These may be
interdependent as the environment may cause epigenetic changes.
In an animal model of early-life stress characterized by reduced
maternal care, epigenetic changes, including increased DNA
methylation (DNAm), were seen in the promoter region of the
glucocorticoid-receptor gene, and these persisted into adulthood,
where they correlated with disruption of the hypothalamic-
pituitary-adrenal axis [4]. Analogously, DNA from postmortem
hippocampus obtained from suicide victims with a history of
childhood abuse, also showed increased DNAm in the human
version of the same gene [5].
Epigenetics, which has been frequently implicated in cancers
[6], has also been implicated in brain diseases, such as Rett
PLoS ONE | www.plosone.org 1 April 2012 | Volume 7 | Issue 4 | e34451
syndrome [7] and fragile X syndrome [8]. There is now ample
evidence that DNAm plays a critical role in brain development
and function. One study found that abnormally hypomethylated
CNS neurons were impaired functionally and were selected
against in postnatal development [9]. We have shown that DNAm
signatures distinguished three brain regions—cortex, cerebellum,
and pons [10]. A role for epigenetics in MDD and other
psychiatric disorders has been suggested based on factors such as
the lack of complete concordance in monozygotic twins, the onset
of illness in adolescence or adulthood rather than childhood, the
often episodic nature of the illnesses, and the apparent relationship
to environmental factors, including stress [11].
There are several examples of epigenetic variation in candidate
MDD genes and in DNA treated with medications used for MDD.
For example, early life adversity increased DNAm in Bdnf in rats
[12]. Valproate [13], used to treat bipolar depression, and
haloperidol [14], used for psychotic depression, as well as the
antidepressants imipramine [15], tranylcypromine [16], and
fluoxetine [17] have been shown to induce epigenetic changes in
rodent brain. Further, administration of a histone deacetylase
inhibitor, sodium butyrate, produces an antidepressant effect in an
animal model [18].
Despite the availability of an essentially complete genome
sequence for several years, understanding of the methylome has
progressed more slowly, largely due to limitations in technology
affecting sensitivity, specificity, throughput, quantitation, and cost
among the previously used detection methods. Microarray-based
methods can interrogate much larger numbers of CpGs than other
approaches. One study to date has reported on a genome-wide
DNAm study in psychiatric disorders, demonstrating differences in
the 4–9% range between DNA from bipolar disorder or
schizophrenia brain samples vs. controls [19]. This study used
the methylation-sensitive restriction enzymes HpaII and McrBC to
prepare DNA, which they hybridized to a 12,192 CpG-island
microarray.
We have similarly used a methylation-sensitive restriction
enzyme-based method focused on McrBC, though we have
implemented it on a microarray platform (CHARM), which is
not biased towards CpG islands, but rather has features chosen
agnostically based on high CpG density. We have shown that
CHARM robustly distinguishes tissue types based on differential
DNAm profiles, and can also discriminate between colon cancer
and normal colon tissues [20].
We have now used CHARM analysis to study genome-wide
DNAm variation in 39 MDD and 26 control brains. Here we
report results of this experiment and of follow-up pyrosequencing
experiments to attempt to validate the initial findings and to
correlate DNAm differences with gene expression. While these
data should be followed up on a much larger replication set, the
absence of large DNAm differences in the brains of MDD patients
is itself important in considering the epigenetic hypothesis. These
results suggest that if DNAm plays a role in MDD, the most
critical target may not be the frontal cortex, but other regions,
such as hippocampus and amygdala, key components of the limbic
system, in which epigenetic changes have been shown to influence
cognitive and behavioral phenotypes [21].
Materials and Methods
Ethics Statement
The Johns Hopkins University IRB approved all research
involving human participants. Subjects gave written informed
consent under the IRB-approved protocol.
Brain DNA
Postmortem frontal cortex brain tissue, Brodmann area 10,
from 39 individuals with MDD and 27 matched controls were
donated by the Stanley Medical Research Institute, in two batches.
The first sample set consisted of 12 psychotic depression cases, 12
non-psychotic depression cases and 12 age and sex matched
controls. A second set consisted of 15 non-psychotic depression
cases and 15 age and sex matched controls. To increase power, the
two samples were analyzed together. A structured interview-based
DSM-IV diagnosis was assigned to each sample independently by
two senior psychiatrists, based on available medical records and a
series of interviews conducted with the family [22]. For each brain,
the cerebrum was hemisected, and one half was fixed in formalin
while the other was cut into 1.5 cm thick coronal slices and frozen
in a mixture of isopentane and dry ice. Right and left brain
hemispheres were randomly alternated for formalin fixing or
freezing. Frozen tissues were used for the DNAm studies.
Formalin-fixed, paraffin-embedded sections were employed for
the immunohistochemical analysis of acetylcholinesterase (AChE).
All frozen tissue was stored at 270uC. DNA was extracted using
the MasterPure DNA Purification kit (Epicentre Biotechnologies).
A replication sample set was provided by the Maryland Psychiatric
Research Center. This consisted of post-mortem BA10 samples
from 16 subjects with MDD and 13 controls. These samples were
age, sex, and race matched.
Lymphoblastoid cell line DNA
Cases of MDD (N = 30) were selected from the Genetics of
Recurrent Early Onset Depression (GenRED) study. Clinical
methods have been described elsewhere [23]. MDD cases had two
or more episodes of DSM-IV MDD with onset before age 31.
Subjects gave written informed consent under IRB-approved
protocols. European-American controls (N = 30) selected from the
NIMH Genetics Initiative repository had no MDD. DNA for
GenRED cases and MGS controls was provided from EBV-
transformed lymphoblastoid cell lines by the NIMH Center for
Collaborative Genetics Studies. A replication set of 90 MDD cases
and 90 controls were also run.
CHARM platform
The CHARM assay was performed as described previously
[20]. Briefly, 10 mg of DNA were sheared in 100 ml using a
Hydroshear device (Genomic Solution) to 1.6 kb–3 kb. Sheared
DNA was then divided into two fractions. One fraction was
digested overnight at 37uC with the methyl-sensitive enzyme
McrBC (NEB). Following digestion cut and uncut fractions from
the same sample were electrophoresed in adjacent wells of a 1%
agarose gel. Areas corresponding to the 1.6 kb–3 kb regions were
excised and purified using Qiagen Spin Gel Purification columns.
The gel-purified DNA was quantified on a spectrophotometer and
30 ng of DNA from each fraction was amplified using a
GenomePlex Whole Genome Amplification Kit (SIGMA). The
amplified DNA was then isolated with a Qiagen PCR Purification
column, then quantified using a spectrophotometer. The untreat-
ed, total DNA fraction was labeled with Cy3 and the methyl-
depleted DNA fraction was labeled with Cy5 and hybridized onto
the custom NimbleGen 2.1 M feature CHARM microarray
(design previously described [24]).
Pyrosequencing
1mg of genomic DNA was bisulfite treated using the Epitect kit
(Qiagen). CpG unbiased primers were designed to PCR amplify
92 CpG sites in 17 genes. Nested PCR was performed. Amplicons
DNA Methylation in Major Depressive Disorder
PLoS ONE | www.plosone.org 2 April 2012 | Volume 7 | Issue 4 | e34451
DNA Methylation in Major Depressive Disorder
PLoS ONE | www.plosone.org 3 April 2012 | Volume 7 | Issue 4 | e34451
were analyzed on a PSQ HS 96 pyrosequencer (Biotage), and
CpG sites were quantified, from 0% to 100% methylation, using
Pyro Q-CpG software [25].
Real-time gene expression
RNA was extracted from frontal cortex using the RNAeasy Kit
(Qiagen). MonsterScript 1st – Strand cDNA Synthesis Kit
(Epicentre) was used to generate cDNA for subsequent quantita-
tive real-time PCR. Negative RT samples were used to ensure the
absence of contamination. All reactions were carried out in
triplicate using 16TaqMan master mix (Applied Biosystems), 16
TaqMan probe for each gene, and 10 ng of template in a volume
of 20 mL. Real-time reactions were performed on an Applied
Biosystems 7900HT Real-Time PCR System. Each set of
triplicates was checked to ensure that the threshold cycle (Ct)
values were all within 1 Ct of each other. The delta-delta-Ct
method was used to determine sample quantity.
Microarray data preprocessing
Hybridization quality was assessed by comparing the untreated
fraction signal intensity for each genomic probe to that of
background (anti-genomic) probes, with the expectation that the
genomic probes should register significantly higher signals. Poor
hybridization was indicated by genomic probe signal levels not
being significantly higher than background probe levels. Using this
metric eight arrays were identified as having failed hybridization
and discarded.
Detection of differentially methylated regions (DMRs)
Normalized methylation log-ratios were smoothed using a
weighted sliding window as previously described [24]. For each
probe, the average log-ratio and standard deviation were
computed for cases and controls allowing a Z-score to be
calculated for each probe. Under the assumption that most
regions are not differentially methylated, the median absolute
deviation of t-scores across all probes was used to determine the
standard deviation of the null distribution. Contiguous regions of
$6 smoothed Z-scores with p,0.005 were identified as candidate
DMRs. For these regions, a Bayesian model was used to convert
log ratios of intensities to estimated percent methylation [26]. P-
values were assigned by comparing the DMR areas to a null
distribution generated by permuting sample labels.
Gene Ontology analysis
We sought to determine whether our nominally significant
differentially methylated regions were in or near genes that
clustered together functionally. We determined the nearest gene
for each differential DNAm region and thus created a list of genes
with differential DNAm. We then asked whether this gene list was
enriched for GO Biological Process categories [27] using the NIH
DAVID tool [28]. We calculated an expected number of genes we
would see from our data set in each category under the null
hypothesis and compared that with the observed number to obtain
a p-value using the Fisher exact test. To better determine the
statistical significance of these results we further calculated a False
Discovery Rate using the Benjamini-Hochberg method [29].
Analysis of pyrosequencing data
For each of the 17 most differentially methylated regions, we
assessed pyrosequencing data based on primers designed across
the most CpG dense part of the region implicated by CHARM. A
linear regression model was used to assess the statistical
significance of the effect of case-control status on DNAm. These
were then corrected at two levels of stringency: 1) taking the best p-
value for each gene and correcting for 17 tests (the number of
regions tested); and 2) taking all p-values and correcting for 92 tests
(the number of CpGs tested). We then tested DNAm levels at all
CpGs against a number of additional sample variables including:
pH, postmorterm interval, age, sex, side of brain assayed, smoking
at time of death, and lifetime alcohol use, using a univariate
regression model. Resulting p-values were corrected for the
number of tests performed (92). For PRIMA1 each of these was
added as a covariate into a regression equation with case status as
the primary independent variable and DNAm as the dependent
variable.
Immunohistochemistry
Ten micron-thick paraffin sections from four subjects with
MDD and five controls were processed for AChE immunostain-
ing. Sections were incubated with a rabbit polyclonal antibody
targeting a signature epitope of an AChE precursor recombinant
protein, particularly suitable for tissue immunohistochemistry
(HPA019704; Sigma, St. Louis, MO), at a 1:25 dilution and
subsequently processed by a modification of the avidin-biotin-
peroxidase method as we have previously described [30]. Several
AChE immunostaining parameters were measured semi-quanti-
tatively in a blind fashion, using Likert scale scores (0–4) as
reported [30]: overall intensity of staining, degree of reticular
neuropil staining, and density of perikaryal neurite clusters. Scores
were compared by the Mann-Whitney-U test. In addition, we
performed qualitative evaluations of neuronal perikaryal and
nuclear staining.
Figure 1. Examples of CHARM results for two of the regions showing greatest DNAm differences between MDD cases and controls.
The plots show percent methylation versus genomic location with each point representing the methylation level of an individual sample for a given
probe. The curve represents averaged smoothed percent methylation values. The locations of CpG dinucleotides are indicated with black tickmarks
on the X-axis. CpG density was calculated across the region using a standard density estimator and is represented by the smoothed black line. The
location of the CpG island is denoted on the X-axis as an orange line. Gene annotation is indicated, showing LASS2 in (a) and PRIMA1 in (b). The thin
outer grey line represents the transcript, while the thin inner lines represent a coding region. Filled in grey boxes represent exons.
doi:10.1371/journal.pone.0034451.g001
Table 1. Stanley Medical Research Institute MDD and control brain samples.
N Age M F PMI (hr) pH % suicide L R
Control 27 48.2610.5 23 4 26.5615.5 6.660.5 0 12 15
MDD 39 44.6610.6 28 11 44.5632.8 6.660.5 53.8 21 18
doi:10.1371/journal.pone.0034451.t001
DNA Methylation in Major Depressive Disorder
PLoS ONE | www.plosone.org 4 April 2012 | Volume 7 | Issue 4 | e34451
Results
Characteristics of postmortem brain samples are provided in
Table 1. Of these 66 samples, 58 were used in our analyses. Data
for eight were removed because of inadequate quality of array
hybridization. CHARM analysis identified 438 nominally signif-
icant candidate DMRs between MDD and controls (Table S1). Of
these, 224 DMRs showed differences .10%; the largest difference
was 22%. Figure 1 shows examples of two regions with the greatest
DNAm differences. We note that their magnitude was modest
compared to another disease vs. control CHARM experiment in
which we observed colon cancer vs. normal colon DNAm
differences of up to 52%. Nonetheless, their magnitude was not
unexpected given the results of a comparable study of psychiatric
brain samples with DNAm differences in the single digits [19]. We
calculated a false discovery rate (FDR) for each DMR to account
for multiple testing. None of the DNAm differences reached the
threshold for statistical significance (q-value,0.1) after correcting
for multiple testing. However, we sought to further characterize
the results with additional exploratory analyses.
We assessed the DNAm differences between MDD and controls
using the Biological Processes categories of the Gene Ontology
database [27]. The set of overrepresented categories includes
many processes related to neurogenesis and central nervous system
development (Table S2). These categories are intriguing given the
neurotrophic model of MDD that posits a critical role for deficits
in neuronal growth in the etiopathogenesis of the illness [31].
We attempted validation for 17 DMRs chosen because they
were among those showing the greatest DNAm differences
between MDD and controls, and were in or near genes. Within
these regions bisulfite pyrosequencing was conducted across 92
CpG dinucleotides. We observed nominally significant DNAm
differences in 10 of the regions. The four regions with the strongest
results, those in or near the genes LASS2,CPSF3,ZNF263, and
PRIMA1 (Figure 2), remained statistically significant after
correcting for 17 tests. The greatest DNAm difference for each
gene was 4, 8, 8, and 15 percent, respectively, with the MDD
samples being the more highly methylated for each of the four.
When we corrected for 92 CpGs tested, only four consecutive
CpGs in PRIMA1, with 12–15% increased DNAm in MDD,
remained significant (p = 0.00019–0.00028).
For all of the 17 regions tested, we tested the impact of
additional demographic, clinical, and biologic variables on DNAm
(Table S3). After correction for 17 regions tested, DNAm was not
predicted by pH, post-mortem interval, age, sex, side of brain,
smoking, psychotic status, or alcohol use. For PRIMA1, two
variables predicted DNAm for CpG-2 at a nominal level of
significance: increased age was associated with decreased DNAm
(p = 0.04), as was lower pH (p = 0.03) (Table 2). When these two
variables were included as covariates in a regression the
relationship between MDD and DNAm remained significant
(p = 0.008–0.02). We further examined whether medication usage
might account for the increased DNAm at PRIMA1 in MDD
samples by focusing on the subset of seven samples that were
medication free. DNAm for these were 3–6% greater than for the
remaining 32 MDD samples (p = 0.15), suggesting that medication
was not responsible for the difference between MDD and controls.
Figure 2. Results of bisulfite pyrosequencing for validation of
CHARM in brain samples. Regions in or near four genes showed
differences that remained statistically significant after correction for
having tested 17 genes: (a) LASS2, (b) CPSF3, (c) ZNF263, (d) PRIMA1. The
grey bars represent values from control brain sample DNA, while the
black bars represent those from MDD brain samples. The Y-axis is
percent DNA methylation, while the X-axis shows distance along the
chromosome for each CpG dinucleotide assayed. One asterisk indicates
a difference between MDD and control of p,0.05. Two asterisks
indicates p,0.0029 (a correction for 17 regions tested). Three asterisks
indicates p,0.00054 (a correction for 92 CpGs tested).
doi:10.1371/journal.pone.0034451.g002
DNA Methylation in Major Depressive Disorder
PLoS ONE | www.plosone.org 5 April 2012 | Volume 7 | Issue 4 | e34451
To assess the potential functional impact of increased PRIMA1
DNAm in MDD, we tested mRNA levels of the gene in the same
brain samples that were used for the DNAm experiments. Levels
were altered in the MDD brain samples in the expected direction,
being decreased 53% (p = 0.047).
Because of the potential clinical value of blood-derived
biomarkers, we sought to determine whether PRIMA1 DNAm
Table 2. PRIMA1 DNAm by diagnosis, and by covariate status
a.
MDD vs. control Covariates (p-values)
CpG
Control %
DNAm MDD % DNAm Dx (p-value) PMI Brain pH Side of Brain Age Sex Smoking Alcohol
1 40.3 50.9 0.0062 0.66 0.062 0.99 0.053 0.57 0.30 0.74
2 43.4 58.7 0.00027 0.50 0.031 0.90 0.044 0.88 0.51 0.71
3 51.9 67.2 0.00028 0.76 0.052 0.72 0.063 0.64 0.58 0.55
4 57.3 71.4 0.00026 0.76 0.052 0.80 0.083 0.85 0.52 0.60
5 64.4 76.7 0.00019 0.91 0.083 0.96 0.072 0.95 0.70 0.46
6 93.5 95.6 0.050 0.06 0.034 0.77 0.010 0.50 0.36 0.89
a
Dx = diagnosis; PMI = postmortem interval; DNAm = DNA methylation; p-values,0.05 are italicized for clarity.
doi:10.1371/journal.pone.0034451.t002
Figure 3. Immunohistochemical pattern of AChE in frontal cortex. (A) In controls there is diffuse and intense pattern of immunoreactivity
involving mainly the neuropil. (B) In MDD subjects, though variable, immunostaining was reduced. Both 2006. (C) In controls, there is virtually no
perikaryal staining. (D) The latter contrasts with the pattern observed in some areas in MDD subjects, in which groups of pyramidal neurons display
intense perikaryal staining, suggesting redistribution of the enzyme to the cell body. The red circles highlight examples. Both 6406.
doi:10.1371/journal.pone.0034451.g003
DNA Methylation in Major Depressive Disorder
PLoS ONE | www.plosone.org 6 April 2012 | Volume 7 | Issue 4 | e34451
differences could be detected between subjects from our GenRED
study as compared to normal controls collected for genetic studies.
We used DNA from these subjects’ lymphoblastoid cell lines and
saw results similar to those in brain. DNAm was increased in
MDD subjects as compared to controls (by 7–10%, p = 0.0006–
0.01) for three of the four PRIMA1 CpGs (Figure S1).
We attempted to replicate both the brain and the blood results
using independent sample sets. In 16 MDD postmortem brain
samples and 13 controls, we failed to detect a significant difference
in DNAm at any of the four previously implicated PRIMA1 CpGs.
DNAm levels were virtually identical between groups (Supporting
Information S1). The biggest difference was a 4.4% decrease in
methylation for the cases at the third CpG (p = 0.16). Similarly
DNAm in CpGs in LASS2,CPSF3,ZNF263 did not differ
significantly between groups (Supporting Information S1). When
we examined lymphoblastoid cell line DNA from an additional 90
MDD cases and 90 controls, we could not replicate the DNAm
differences in PRIMA1 observed in the prior sample set
(Supporting Information S1).
Using immunohistochemistry, we investigated whether MDD
subjects with high DNAm and low expression for PRIMA1 would
show reduced immunoreactivity for AChE as compared to
controls with the opposite pattern. Such a result would be
consistent with the changes we observed in PRIMA1 DNAm
influencing cholinergic transmission. In a semi-quantitative
comparison between frontal cortex tissues from four MDD
subjects and five controls, we found that overall AChE staining
intensity was reduced in the MDD subjects on average 42%,
however, this difference did not reach statistical significance
(p = 0.08). We also observed that subjects with MDD had a larger
number of superficial pyramidal neuron perikaryal staining,
despite overall reduction in neuropil immunoreactivity, suggesting
redistribution of AChE towards the cell bodies (Figure 3).
Discussion
We report here on the first genome-wide DNA methylation
comparison between MDD and control brain. Although the
magnitude of DNAm differences we observed was relatively small
and did not survive correction for multiple testing, the DMRs
identified were in or near genes enriched for roles in neuronal
growth and development, suggesting that the differences picked up
by our CHARM experiment, despite being relatively small, might
be biologically meaningful. Our validation experiment showed the
greatest differences in PRIMA1, with 12–15% increased DNAm in
MDD. Consistent with this result, PRIMA1 expression was
decreased in MDD brain samples. The DNAm changes in the
brain were also reflected in DNA from an initial set of
lymphoblastoid cell lines, with MDD cases again showing greater
DNAm than controls. However, we were unable to replicate
PRIMA1 DNAm differences in additional sample sets of brain and
lymphoblastoid cell lines. Further, although we observed de-
creased immunoreactivity for AChE in MDD tissues that had
increased PRIMA1 DNAm, this change did not reach statistical
significance. Therefore, we cannot draw firm conclusions about a
potential role for PRIMA1 DNAm in MDD.
PRIMA1 is of substantial biological interest in MDD because of
its relationship to cholinergic neurotransmission. The gene
encodes a protein that both guides the transport of acetylcholin-
esterase to neuronal membranes [32] and anchors it there [33].
When PRIMA1 is knocked down by antisense cDNA [33] or
knocked out [32], there is a decrease in localization of AChE at the
neuronal membrane, or of AChE activity, respectively. AChE
hydrolyzes acetylcholine, thus less of its activity means more
cholinergic transmission. Janowsky and colleagues proposed that
increased cholinergic transmission is a central mechanism in
depression, noting that reserpine, which can cause depression, is
cholinomimetic, and the tricyclic antidepressants are anticholin-
ergic [34]. Additional evidence in support of this hypothesis
includes the induction of depressive symptoms by the administra-
tion of physostigmine, a more specific cholinometic agent [35],
and the alleviation of such symptoms by the use of more specific
anticholinergic medications such as scopolamine [36], a musca-
rinic acetylcholine receptor antagonist, and mecamylamine, a
nicotinic acetycholine receptor antagonist [37]. Intriguingly, stress,
which plays a key role in MDD etiology, has been shown to
influence cholinergic gene expression in mouse brain [38].
Compared to DNAm differences seen in prior studies using the
CHARM platform to compare tissue or cell types, or colon cancer
vs. normal colon, the magnitude of those seen in our study was
modest. This is, perhaps, not surprising given the findings of the
only other genome-wide DNAm studies in psychiatric illness, that
of Mill et al [19] and Dempster et al [39], which similarly found
small, though statistically significant, differences between cases and
controls. It is likely because the magnitude of our DNAm
differences hovered around the limit of resolution of CHARM
that a number of our candidate DMRs did not validate. Since
completing this experiment, we have developed improvements to
CHARM that increase its signal-to-noise ratio. In addition, the
next generation of CHARM includes coverage of a greater
number of CpGs, augmenting beyond the ,20% of all CpGs that
were initially on the array. In the current experiment we employed
a conservative statistical threshold to guard against false positives.
It is possible that a relaxed threshold might have captured more
signals reflecting true biological differences between depression
and controls.
Our failure to detect a robustly replicating signal makes it hard
to draw firm conclusions about a role for DNAm in the frontal
cortex of subjects with MDD. It is possible that larger
etiopathologically relevant DNAm changes might exist in other
brain regions known to be involved in MDD, such as the limbic
regions anterior cingulate cortex [40], amygdala, and hippocam-
pus [41]. We have previously shown brain region-specific variation
in DNAm [10]. Further, disease-related DNAm variation might be
restricted to particular cell types, such as neurons only or even,
more narrowly, subtypes of neurons, such as pyramidal cells.
However, there may be a substantial portion of DMRs that are not
cell type- or tissue-specific. We note that these generalized MDD
DMRs might be the most valuable as they both shed light on
etiopathogenesis, and also potentially provide biomarkers that can
be studied in living patients. Blood-based DMRs would also allow
for much larger numbers of samples to be assayed and for
correlation on a large scale with genotype.
Supporting Information
Figure S1 Results of bisulfite pyrosequencing of six
PRIMA1 CpGs in lymphablastoid cell line samples. The
grey bars represent values from control sample DNA, while the
black bars represent those from MMD samples. The Y-axis is
percent DNA methylation, while the X-axis shows each CpG
arrayed along the chromosome. Asterisks indicate a difference
between MDD and control of p,0.01.
(TIFF)
Table S1 The results of the primary experiment using
CHARM to compare postmortem brain samples be-
tween MDD cases and controls.
(DOC)
DNA Methylation in Major Depressive Disorder
PLoS ONE | www.plosone.org 7 April 2012 | Volume 7 | Issue 4 | e34451
Table S2 The result of taking all genes in or near
nominally significant differentially methylated regions
and examining their representation in Gene Ontology
Categories.
(DOC)
Table S3 Bisulfite pyrosequencing was used to experi-
mentally validate some of the regions that showed
differential methylation between MDD and controls by
CHARM analysis. This table shows those that were nominally
validated. P-values for regression of pyrosequencing methylation at
individual CpGs (rows) on 6 covariates (columns). The last column
shows the F-statistic p-value for the multiple regression of
methylation on all 6 covariates.
(DOC)
Supporting Information S1 Supplementary Tables S4,
S5, S6 show results of replication attempts in postmor-
tem brain and in lymphoblastoid cell lines.
(DOCX)
Acknowledgments
The authors would like to thank SMRI and Dr. Fuller Torrey for providing
postmortem MDD brain samples. GenRED consortium co-authors are:
James A. Knowles, Myrna M. Weissman, William Coryell, William A.
Scheftner, and Douglas F. Levinson. Drs. Potash and Feinberg made equal
contributions and should be considered co-senior authors. Some of this
material was presented at the World Congress of Psychiatric Genetics, San
Diego, November 4–8, 2009. Data and biomaterials were collected in six
projects that participated in the National Institute of Mental Health
(NIMH) Genetics of Recurrent Early-Onset Depression (GenRED) project.
From 1999–2003, the Principal Investigators and Co-Investigators were:
New York State Psychiatric Institute, New York, NY, Myrna M.
Weissman, Ph.D. and James K. Knowles, M.D., Ph.D.; University of
Pittsburgh, Pittsburgh, PA, George S. Zubenko, M.D., Ph.D. and Wendy
N. Zubenko, Ed.D., R.N., C.S.; Johns Hopkins University, Baltimore, J.
Raymond DePaulo, M.D., Melvin G. McInnis, M.D. and Dean
MacKinnon, M.D.; University of Pennsylvania, Philadelphia, PA, Douglas
F. Levinson, M.D. (GenRED coordinator), Madeleine M. Gladis, Ph.D.,
Kathleen Murphy-Eberenz, Ph.D. and Peter Holmans, Ph.D. (University
of Wales College of Medicine); University of Iowa, Iowa City, IW,
Raymond R. Crowe, M.D. and William H. Coryell, M.D.; Rush
University Medical Center, Chicago, IL, William A. Scheftner, M.D.
Rush-Presbyterian.
Author Contributions
Conceived and designed the experiments: APF JBP RHY. Performed the
experiments: SS MR MJW WEK AL. Analyzed the data: MJA RAI PM.
Contributed reagents/materials/analysis tools: GC. Wrote the paper: SS
MJA WEK APF JBP.
References
1. Sullivan PF, Neale MC, Kendler KS (2000) Genetic epidemiology of major
depression: Review and meta-analysis. Am J Psychiatry 157: 1552–1562.
2. Shyn SI, Shi J, Kraft JB, Potash JB, Knowles JA, et al. (2011) Novel loci for
major depression identified by genome-wide association study of sequenced
treatment alternatives to relieve depression and meta-analysis of three studies.
Mol Psychiatry 16: 202–15.
3. Paykel ES, Myers JK, Dienelt MN, Klerman GL, Lindenthal JJ, et al. (1969) Life
events and depression. A controlled study. Arch Gen Psychiatry 21: 753–760.
4. Weaver IC, Cervoni N, Champagne FA, D’Alessio AC, Sharma S, et al. (2004)
Epigenetic programming by maternal behavior. Nat Neurosci 7: 847–54.
5. McGowan PO, Sasaki A, D’Alessio AC, Dymov S, Labonte B, et al. (2009)
Epigenetic regulation of the glucocorticoid receptor in human brain associates
with childhood abuse. Nat Neurosci 12: 342–348.
6. Feinberg AP, Vogelstein B (1983) Hypomethylation distinguishes genes of some
human cancers from their normal counterparts. Nature 301: 89–92.
7. Amir RE, Van den Veyver IB, Wan M, Tran CQ, Francke U, et al. (1999) Rett
syndrome is caused by mutations in X-linked MECP2, encoding methyl-CpG-
binding protein 2. Nat Genet 23: 185–188.
8. Oberle I, Rousseau F, Heitz D, Kretz C, Devys D, et al. (1991) Instability of a
550-base pair DNA segment and abnormal methylation in fragile X syndrome.
Science 252: 1097–1102.
9. Fan G, Beard C, Chen RZ, Csankovszki G, Sun Y, et al. (2001) DNA
hypomethylation perturbs the function and survival of CNS neurons in postnatal
animals. J Neurosci 21: 788–797.
10. Ladd-Acosta C, Pevsner J, Sabunciyan S, Yolken RH, Webster MJ, et al. (2007)
DNA methylation signatures within the human brain. Am J Hum Genet 81:
1304–1315.
11. Mill J, Petronis A (2007) Molecular studies of major depressive disorder: The
epigenetic perspective. Mol Psychiatry 12: 799–814.
12. Roth TL, Lubin FD, Funk AJ, Sweatt JD (2009) Lasting epigeneti c influence of
early-life adversity on the BDNF gene. Biol Psychiatry 65: 760–769.
13. Milutinovic S, D’Alessio AC, Detich N, Szyf M: Valproate induces widespread
epigenetic reprogramming which involves demethylation of specific genes.
Carcinogenesis (2007) ; 28: 560–571.
14. Shimabukuro M, Jinno Y, Fuke C, Okazaki Y (2006) Haloperidol treatment
induces tissue- and sex-specific changes in DNA methylation: A control study
using rats. Behav Brain Funct 2: 37.
15. Tsankova NM, Berton O, Renthal W, Kumar A, Neve RL, et al. (2006)
Sustained hippocampal chromatin regulation in a mouse model of depression
and antidepressant action. Nat Neurosci 9: 519–525.
16. Lee MG, Wynder C, Schmidt DM, McCafferty DG, Shiekhattar R (2006)
Histone H3 lysine 4 demethylation is a target of nonselective antidepressive
medications. Chem Biol 13: 563–567.
17. Cassel S, Carouge D, Gensburger C, Anglard P, Burgun C, et al. (2006)
Fluoxetine and cocaine induce the epigenetic factors MeCP2 and MBD1 in
adult rat brain. Mol Pharmacol 70: 487–492.
18. Schroeder FA, Lin CL, Crusio WE, Akbarian S (2007) Antidepressant-like
effects of the histone deacetylase inhibitor, sodium butyrate, in the mouse. Biol
Psychiatry 62: 55–64.
19. Mill J, Tang T, Kaminsky Z, Khare T, Yazdanpanah S, et al. (2008)
Epigenomic profiling reveals DNA-methylation changes associated with major
psychosis. Am J Hum Genet 82: 696–711.
20. Irizarry RA, Ladd-Acosta C, Wen B, Wu Z, Montano C, et al. (2009) The
human colon cancer methylome shows similar hypo- and hypermethylation at
conserved tissue-specific CpG island shores. Nat Genet 41: 178–186.
21. Sweatt JD (2009) Experience-dependent epigenetic modifications in the central
nervous system. Biol Psychiatry 65: 191–197.
22. Torrey EF, Webster M, Knable M, Johnston N, Yolken RH (2000) The Stanley
Foundation brain collection and neuropathology consortium. Schizophr Res 44:
151–155.
23. Levinson DF, Zubenko GS, Crowe RR, DePaulo RJ, Scheftner WS, et al. (2003)
Genetics of recurrent early-onset depression (GenRED): Design and preliminary
clinical characteristics of a repository sample for genetic linkage studies.
Am J Med Genet 119B: 118–130.
24. Irizarry RA, Ladd-Acosta C, Carvalho B, Wu H, Brandenburg SA, et al. (2008)
Comprehensive high-throughput arrays for relative methylation (CHARM).
Genome Res 18: 780–790.
25. Tost J, Gut IG (2007) DNA methylation analysis by pyrosequencing. Nat Protoc
2: 2265–2275.
26. Aryee MJ, Wu Z, Ladd-Acosta C, Herb B, Feinberg AP, et al. (2011) Accurate
genome-scale percentage DNA methylation estimates from microarray data.
Biostatistics 12: 197–210.
27. Harris MA, Clark J, Ireland A, Lomax J, Ashburner M, et al. (2004) The gene
ontology (GO) database and informatics resource. Nucleic Acids Res 32:
D258–61.
28. Huang da W, Sherman BT, Lempicki RA (2009) Systematic and integrative
analysis of large gene lists using DAVID bioinformatics resources. Nat Protoc 4:
44–57.
29. Benjamini Y, Hochberg Y (1995) Controlling the false discovery rate: A practical
and powerful approach to multiple testing. J Royal Stat Soc,
Series B (Methodological) 57: 289–300.
30. Kaufmann WE, MacDonald SM, Altamura CR (2000) Dendritic cytoskeletal
protein expression in mental retardation: An immunohistochemical study of the
neocortex in Rett syndrome. Cereb Cortex 10: 992–1004.
31. Duman RS, Heninger GR, Nestler EJ (1997) A molecular and cellular theory of
depression. Arch Gen Psychiatry 54: 597–606.
32. Dobbertin A, Hrabovska A, Dembele K, Camp S, Taylor P, et al. (2009)
Targeting of acetylcholinesterase in neurons in vivo: A dual processing function
for the proline-rich membrane anchor subunit and the attachment domain on
the catalytic subunit. J Neurosci 29: 4519–4530.
33. Perrier AL, Massoulie J, Krejci E (2002) PRiMA: The membrane anchor of
acetylcholinesterase in the brain. Neuron 33: 275–285.
34. Janowsky DS, el-Yousef MK, Davis JM, Sekerke HJ (1972) A cholinergic-
adrenergic hypothesis of mania and depression. Lancet 2: 632–635.
35. Risch SC, Kalin NH, Janowsky DS (1981) Cholinergic challenges in affective
illness: Behavioral and neuroendocrine correlates. J Clin Psychopharmacol 1:
186–192.
DNA Methylation in Major Depressive Disorder
PLoS ONE | www.plosone.org 8 April 2012 | Volume 7 | Issue 4 | e34451
36. Furey ML, Drevets WC (2006) Antidepressant efficacy of the antimuscarinic
drug scopolamine: A randomized, placebo-controlled clinical trial. Arch Gen
Psychiatry 63: 1121–1129.
37. Bacher I, Wu B, Shytle DR, George TP (2009) Mecamylamine – a nicotinic
acetylcholine receptor antagonist with potential for the treatment of neuropsy-
chiatric disorders. Expert Opin Pharmacother 10: 2709–2721.
38. Kaufer D, Friedman A, Seidman S, Soreq H (199 8) Acute stress facilitates long-
lasting changes in cholinergic gene expression. Nature 393: 373–377.
39. Dempster EL, Pidsley R, Schalkwyk LC, Owens S, Georgiades A, et al. (2011)
Disease-associated epigenetic changes in monozygotic twins discordant for
schizophrenia and bipolar disorder. Hum Mol Genet. 20: 4786–4796.
40. Mayberg HS, Lozano AM, Voon V, McNeely HE, Seminowicz D, et al. (2005)
Deep brain stimulation for treatment-resistant depression. Neuron 45: 651–660.
41. Campbell S, Marriott M, Nahmias C, MacQueen GM (2004) Lower
hippocampal volume in patients suffering from depression: A meta-analysis.
Am J Psychiatry 161: 598–607.
DNA Methylation in Major Depressive Disorder
PLoS ONE | www.plosone.org 9 April 2012 | Volume 7 | Issue 4 | e34451
... It is also important to note that none of these existing studies have examined the association between neighborhood deprivation and DNAm in brain tissue. DNAm changes in the brain specifically are important to study because they can provide indications of neuropathology outcomes such as Alzheimer's disease (AD) [16][17][18][19][20][21][22] and depression [23,24]. Many of these brain health outcomes have themselves been associated with neighborhood deprivation [2,25,26]. ...
... One major strength is that our study is the first known study on the association between neighborhood deprivation and DNAm in brain tissue, which is difficult to obtain and the most relevant tissue to study brain-related health outcomes. Studying DNAm changes in brain tissue is especially important because it can provide insight into neuropathology outcomes such as Alzheimer's disease [16][17][18][19][20][21][22] and depression [23,24]. These brain health outcomes have themselves been associated with neighborhood deprivation [2,25,26], which is the reason more research on neighborhood deprivation using brain tissue is needed. ...
Article
Previous research has found that living in a disadvantaged neighborhood is associated with poor health outcomes. Living in disadvantaged neighborhoods may alter inflammation and immune response in the body, which could be reflected in epigenetic mechanisms such as DNA methylation (DNAm). We used robust linear regression models to conduct an epigenome-wide association study examining the association between neighborhood deprivation (Area Deprivation Index; ADI), and DNAm in brain tissue from 159 donors enrolled in the Emory Goizueta Alzheimer's Disease Research Center (Georgia, USA). We found one CpG site (cg26514961, gene PLXNC1) significantly associated with ADI after controlling for covariates and multiple testing (p-value=5.0e-8). Effect modification by APOE ε4 was statistically significant for the top ten CpG sites from the EWAS of ADI, indicating that the observed associations between ADI and DNAm were mainly driven by donors who carried at least one APOE ε4 allele. Four of the top ten CpG sites showed a significant concordance between brain tissue and tissues that are easily accessible in living individuals (blood, buccal cells, saliva), including DNAm in cg26514961 (PLXNC1). Our study identified one CpG site (cg26514961, PLXNC1 gene) that was significantly associated with neighborhood deprivation in brain tissue. PLXNC1 is related to immune response, which may be one biological pathway how neighborhood conditions affect health. The concordance between brain and other tissues for our top CpG sites could make them potential candidates for biomarkers in living individuals.
... Increased oxidative stress can increase DNA oxidation (e.g., as 8-OHdG and 5-hmc) to induce DNA hypomethylation, or produce elevated prooxidants that can alter DNA methyltransferase (DNMT) protein expression and form DNMT-containing complexes that facilitate DNA hypermethylation (30)(31)(32). It is known that DNA methylation of cytosine residues by DNMTs and histone deacetylation by hypoacetylhistone deacetylases (HDACs) regulate expression of key catecholaminergic genes TH, D1R, D2R, DAT, and NET (32)(33)(34)(35)(36)(37)(38)(39)(40)(41)(42)(43)(44)(45)(46)(47)(48)(49). However, there is limited data on whether developmental Mn exposure may alter the epigenetic profile of catecholaminergic system genes (49)(50)(51)(52)(53). ...
... Protein levels of the key catecholaminergic system genes Th, Dat, and Drd2, which are known to play a role in attentional function, are also known to be altered by developmental Mn exposure (22,(27)(28)(29). Thus, we sought to assess whether developmental Mn exposure leads to lasting changes in the expression of these genes, as well as their known epigenetic gene expression regulators Dnmt1, Dnmt3a, Dnmt3b, Hdac3, and Hdac4 in the prefrontal cortex (PFC) of young adult animals (32)(33)(34)(35)(36)(37)(38)(39)(40)(41)(42)(43)(44)(45)(46)(47)(48)(49). Results show that developmental Mn exposure causes lasting ~50% and ~83% reductions in Th and Dat gene expression, respectively, relative to controls (p's=0.05 ...
Preprint
Elevated manganese (Mn) exposure is associated with attentional deficits in children, and is an environmental risk factor for attention deficit hyperactivity disorder (ADHD). We have shown that developmental Mn exposure causes lasting attention and sensorimotor deficits in a rat model of early childhood Mn exposure, and that these deficits are associated with a hypofunctioning catecholaminergic system in the prefrontal cortex (PFC), though the mechanistic basis for these deficits is not well understood. To address this, male Long-Evans rats were exposed orally to Mn (50 mg/kg/d) over PND 1-21 and attentional function was assessed in adulthood using the 5-Choice Serial Reaction Time Task. Targeted catecholaminergic system and epigenetic gene expression, followed by unbiased differential DNA methylation and gene regulation expression transcriptomics in the PFC, were performed in young adult littermates. Results show that developmental Mn exposure causes lasting focused attention deficits that are associated with reduced gene expression of tyrosine hydroxylase, dopamine transporter, and DNA methyltransferase 3a. Further, developmental Mn exposure causes broader lasting methylation and gene expression dysregulation associated with epigenetic regulation, inflammation, cell development, and hypofunctioning catecholaminergic neuronal systems. Pathway enrichment analyses uncovered mTOR and Wnt signaling pathway genes as significant transcriptomic regulators of the Mn altered transcriptome, and Western blot of total, C1 and C2 phospho-mTOR confirmed mTOR pathway dysregulation. Our findings deepen our understanding of the mechanistic basis of how developmental Mn exposure leads to lasting catecholaminergic dysfunction and attention deficits, which may aid future therapeutic interventions of environmental exposure associated disorders. Significance Statement Attention deficit hyperactivity disorder (ADHD) is associated with environmental risk factors, including exposure to neurotoxic agents. Here we used a rodent model of developmental manganese (Mn) exposure producing lasting attention deficits to show broad epigenetic and gene expression changes in the prefrontal cortex, and to identify disrupted mTOR and Wnt signaling pathways as a novel mechanism for how developmental Mn exposure may induce lasting attention and catecholaminergic system impairments. Importantly, our findings establish early development as a critical period of susceptibility to lasting deficits in attentional function caused by elevated environmental toxicant exposure. Given that environmental health threats disproportionately impact communities of color and low socioeconomic status, our findings can aid future studies to assess therapeutic interventions for vulnerable populations.
... Regardless, these genes may prove useful as screening tools. Sabunciyan et al. [65] performed an MDD study analyzing methylation in 3.5 million CpG sites. The author's classified 224 sites, with the bulk of the genes expressed in neuronal growth and development regions being categorized as having a methylation status of greater than 10% in MDD samples compared to controls. ...
... As PRIMA1 methylation increased in the MDD brain, acetylcholinesterase immunoreactivity decreased. However, the evidence presented is inconclusive but highlights the role epigenetics plays in neuronal development and suggests a potential way to detect MDD [65]. ...
Chapter
Major Depressive Disorder (MDD) or depression is a pathological mental condition affecting millions of people worldwide. Identification of objective biological markers of depression can provide for a better diagnostic and intervention criteria; ultimately aiding to reduce its socioeconomic health burden. This review provides a comprehensive insight into the major biomarker candidates that have been implicated in depression neurobiology. The key biomarker categories are covered across all the "omics" levels. At the epigenomic level, DNA-methylation, non-coding RNA and histone-modifications have been discussed in relation to depression. The proteomics system shows great promise with inflammatory markers as well as growth factors and neurobiological alterations within the endocannabinoid system. Characteristic lipids implicated in depression together with the endocrine system are reviewed under the metabolomics section. The chapter also examines the novel biomarkers for depression that have been proposed by studies in the microbiome. Depression affects individuals differentially and explicit biomarkers identified by robust research criteria may pave the way for better diagnosis, intervention, treatment, and prediction of treatment response.
... Regardless, these genes may prove useful as screening tools. Sabunciyan et al. [65] performed an MDD study analyzing methylation in 3.5 million CpG sites. The author's classified 224 sites, with the bulk of the genes expressed in neuronal growth and development regions being categorized as having a methylation status of greater than 10% in MDD samples compared to controls. ...
... As PRIMA1 methylation increased in the MDD brain, acetylcholinesterase immunoreactivity decreased. However, the evidence presented is inconclusive but highlights the role epigenetics plays in neuronal development and suggests a potential way to detect MDD [65]. ...
Article
Full-text available
Major Depressive Disorder (MDD) or depression is a pathological mental condition affecting millions of people worldwide. Identification of objective biological markers of depression can provide for a better diagnostic and intervention criteria; ultimately aiding to reduce its socioeconomic health burden. This review provides a comprehensive insight into the major biomarker candidates that have been implicated in depression neurobiology. The key biomarker categories are covered across all the ‘omics’ levels. At the epigenomic level, DNA-methylation, non-coding RNA and histone-modifications have been discussed in relation to depression. The proteomics system shows great promise with inflammatory markers as well as growth factors and neurobiological alterations within the endocannabinoid system. Characteristic lipids implicated in depression together with the endocrine system are reviewed under the metabolomics section. The chapter also examines the novel biomarkers for depression that have been proposed by studies in the microbiome. Depression affects individuals differentially and explicit biomarkers identified by robust research criteria may pave the way for better diagnosis, intervention, treatment, and prediction of treatment response.
... We first measured fndc5 expression in dlPFC samples of MDD patients from a well characterized cohort composed of adult subjects (Martins-De-Souza et al., 2012;Sabunciyan et al., 2012). We observed a significant reduction in fndc5 mRNA content in the dlPFC of individuals with depression, regardless of the presence or absence of psychosis, compared with healthy control subjects. ...
Article
Full-text available
Major depressive disorder (MDD) is a major cause of disability in adults. MDD is both a comorbidity and a risk factor for Alzheimer’s disease (AD), and regular physical exercise has been associated with reduced incidence and severity of MDD and AD. Irisin is an exercise-induced myokine derived from proteolytic processing of fibronectin type III domain-containing protein 5 (FNDC5). FNDC5/irisin is reduced in the brains of AD patients and mouse models. However, whether brain FNDC5/irisin expression is altered in depression remains elusive. Here, we investigate changes in fndc5 expression in postmortem brain tissue from MDD individuals and mouse models of depression. We found decreased fndc5 expression in the MDD prefrontal cortex, both with and without psychotic traits. We further demonstrate that the induction of depressive-like behavior in male mice by lipopolysaccharide decreased fndc5 expression in the frontal cortex, but not in the hippocampus. Conversely, chronic corticosterone administration increased fndc5 expression in the frontal cortex, but not in the hippocampus. Social isolation in mice did not result in altered fndc5 expression in either frontal cortex or hippocampus. Finally, fluoxetine, but not other antidepressants, increased fndc5 gene expression in the mouse frontal cortex. Results indicate a region-specific modulation of fndc5 in depressive-like behavior and by antidepressant in mice. Our finding of decreased prefrontal cortex fndc5 expression in MDD individuals differs from results in mice, highlighting the importance of carefully interpreting observations in mice. The reduction in fndc5 mRNA suggests that decreased central FNDC5/irisin could comprise a shared pathologic mechanism between MDD and AD.
Article
Full-text available
Major depressive disorder (MDD) is a chronic, generally episodic and debilitating disease that affects an estimated 300 million people worldwide, but its pathogenesis is poorly understood. The heritability estimate of MDD is 30-40%, suggesting that genetics alone do not account for most of the risk of major depression. Another factor known to associate with MDD involves environmental stressors such as childhood adversity and recent life stress. Recent studies have emerged to show that the biological impact of environmental factors in MDD and other stress-related disorders is mediated by a variety of epigenetic modifications. These epigenetic modification alterations contribute to abnormal neuroendocrine responses, neuroplasticity impairment, neurotransmission and neuroglia dysfunction, which are involved in the pathophysiology of MDD. Furthermore, epigenetic marks have been associated with the diagnosis and treatment of MDD. The evaluation of epigenetic modifications holds promise for further understanding of the heterogeneous etiology and complex phenotypes of MDD, and may identify new therapeutic targets. Here, we review preclinical and clinical epigenetic findings, including DNA methylation, histone modification, noncoding RNA, RNA modification, and chromatin remodeling factor in MDD. In addition, we elaborate on the contribution of these epigenetic mechanisms to the pathological trait variability in depression and discuss how such mechanisms can be exploited for therapeutic purposes.
Chapter
Full-text available
There is growing demand for novel biomarkers that detect early stage disease as well as monitor clinical management and therapeutic strategies. Exosome analysis could provide the next advance in attaining that goal. Exosomes are membrane encapsulated biologic nanometric-sized particles of endocytic origin which are released by all cell types. Unfortunately, exosomes are exceptionally challenging to characterize with current technologies. Exosomes are between 30 and 200nm in diameter, a size that makes them out of the sensitivity range to most cell-oriented sorting or analysis platforms, i.e., traditional flow cytometers. The most common methods for targeting exosomes to date typically involve purification followed by the characterization and the specific determination of their cargo. The whole procedure is time consuming, requiring thus skilled personnel as well as laboratory facilities and benchtop instrumentation. The most relevant methodology for exosome isolation, characterization and quantification is addressed in this chapter, including the most up-to-date approaches to explore the potential usefulness of exosomes as biomarkers in liquid biopsies and in advanced nanomedicine.
Article
Full-text available
Cellular processes that control transcription of genetic information are critical for cellular function, and are often implicated in psychiatric and neurological disease states. Among the most critical of these processes are epigenetic mechanisms, which serve to link the cellular environment with genomic material. Until recently our understanding of epigenetic mechanisms has been limited by the lack of tools that can selectively manipulate the epigenome with genetic, cellular, and temporal precision, which in turn diminishes the potential impact of epigenetic processes as therapeutic targets. This review highlights an emerging suite of tools that enable robust yet selective interrogation of the epigenome. In addition to allowing site-specific epigenetic editing, these tools can be paired with optogenetic approaches to provide temporal control over epigenetic processes, allowing unparalleled insight into the function of these mechanisms. This improved control promises to revolutionize our understanding of epigenetic modifications in human health and disease states.
Article
Although the etiopathogenesis of mental disorders is not fully understood, accumulating evidence support that clinical symptomatology cannot be assigned to a single gene mutation, but it involves several genetic factors. More specifically, a tight association between genes and environmental risk factors, which could be mediated by epigenetic mechanisms, may play a role in the development of mental disorders. Several data suggest that epigenetic modifications such as DNA methylation, post-translational histone modification and interference of microRNA (miRNA) or long non-coding RNA (lncRNA) may modify the severity of the disease and the outcome of the therapy. Indeed, these changes may help to identify patients particularly vulnerable to mental disorders and may have potential utility as biomarkers to facilitate diagnosis and treatment of psychiatric disorders. This article summarizes the most relevant preclinical and human data showing how epigenetic modifications can be central to the therapeutic efficacy of antidepressant and/or antipsychotic agents, as possible predictor of drugs response.
Article
Full-text available
Background: Major depressive disorder (MDD) is a common psychiatric entity, being characterized by alterations in mood and in other clinical dimensions. Several epigenome-wide association studies (EWAS) for MDD have been published. Here, we aimed to identify common genes in EWAS and their convergence with multiple lines of genomic evidence. Methods: We carried out a computational analysis using data of EWAS, which included a meta-analysis for brain samples of MDD, a convergence analysis for brain and blood samples, and top results from available genome-wide expression and association data. Functional enrichment and protein-protein interaction network analyses were also done. Results: The meta-analysis for brain samples detected a significant gene, FAM53B. A list of forty-four top differentially methylated (DM) candidate genes was found, including GRM8, NOTCH4 and SEMA6A, in addition to known druggable genes. The binding-sites for brain-expressed transcription factors, CREB and FOXO1, were enriched in the top DM genes. The protein-protein interaction networks showed that DM genes for MDD, such as RPRM and TMEM14B, play a central role. Conclusion: In this study, we found integrative evidence for the possible role of novel candidate genes and pathways. These genes are involved in mechanisms of synaptic plasticity, which have been associated with several psychiatric disorders. Analysis of epigenetic factors have a great potential for the identification of the mechanisms involved in the pathogenesis of MDD, taking into account their possible role in the interaction between genetic factors and the environment.
Article
Full-text available
DAVID bioinformatics resources consists of an integrated biological knowledgebase and analytic tools aimed at systematically extracting biological meaning from large gene/protein lists. This protocol explains how to use DAVID, a high-throughput and integrated data-mining environment, to analyze gene lists derived from high-throughput genomic experiments. The procedure first requires uploading a gene list containing any number of common gene identifiers followed by analysis using one or more text and pathway-mining tools such as gene functional classification, functional annotation chart or clustering and functional annotation table. By following this protocol, investigators are able to gain an in-depth understanding of the biological themes in lists of genes that are enriched in genome-scale studies.
Article
Full-text available
The Gene Ontology (GO) project (http://www. geneontology.org/) provides structured, controlled vocabularies and classifications that cover several domains of molecular and cellular biology and are freely available for community use in the annotation of genes, gene products and sequences. Many model organism databases and genome annotation groups use the GO and contribute their annotation sets to the GO resource. The GO database integrates the vocabularies and contributed annotations and provides full access to this information in several formats. Members of the GO Consortium continually work collectively, involving outside experts as needed, to expand and update the GO vocabularies. The GO Web resource also provides access to extensive documentation about the GO project and links to applications that use GO data for functional analyses.
Article
Full-text available
The common approach to the multiplicity problem calls for controlling the familywise error rate (FWER). This approach, though, has faults, and we point out a few. A different approach to problems of multiple significance testing is presented. It calls for controlling the expected proportion of falsely rejected hypotheses – the false discovery rate. This error rate is equivalent to the FWER when all hypotheses are true but is smaller otherwise. Therefore, in problems where the control of the false discovery rate rather than that of the FWER is desired, there is potential for a gain in power. A simple sequential Bonferroni-type procedure is proved to control the false discovery rate for independent test statistics, and a simulation study shows that the gain in power is substantial. The use of the new procedure and the appropriateness of the criterion are illustrated with examples.
Article
Full-text available
Studies of the major psychoses, schizophrenia (SZ) and bipolar disorder (BD), have traditionally focused on genetic and environmental risk factors, although more recent work has highlighted an additional role for epigenetic processes in mediating susceptibility. Since monozygotic (MZ) twins share a common DNA sequence, their study represents an ideal design for investigating the contribution of epigenetic factors to disease etiology. We performed a genome-wide analysis of DNA methylation on peripheral blood DNA samples obtained from a unique sample of MZ twin pairs discordant for major psychosis. Numerous loci demonstrated disease-associated DNA methylation differences between twins discordant for SZ and BD individually, and together as a combined major psychosis group. Pathway analysis of our top loci highlighted a significant enrichment of epigenetic changes in biological networks and pathways directly relevant to psychiatric disorder and neurodevelopment. The top psychosis-associated, differentially methylated region, significantly hypomethylated in affected twins, was located in the promoter of ST6GALNAC1 overlapping a previously reported rare genomic duplication observed in SZ. The mean DNA methylation difference at this locus was 6%, but there was considerable heterogeneity between families, with some twin pairs showing a 20% difference in methylation. We subsequently assessed this region in an independent sample of postmortem brain tissue from affected individuals and controls, finding marked hypomethylation (>25%) in a subset of psychosis patients. Overall, our data provide further evidence to support a role for DNA methylation differences in mediating phenotypic differences between MZ twins and in the etiology of both SZ and BD.
Article
The common approach to the multiplicity problem calls for controlling the familywise error rate (FWER). This approach, though, has faults, and we point out a few. A different approach to problems of multiple significance testing is presented. It calls for controlling the expected proportion of falsely rejected hypotheses — the false discovery rate. This error rate is equivalent to the FWER when all hypotheses are true but is smaller otherwise. Therefore, in problems where the control of the false discovery rate rather than that of the FWER is desired, there is potential for a gain in power. A simple sequential Bonferronitype procedure is proved to control the false discovery rate for independent test statistics, and a simulation study shows that the gain in power is substantial. The use of the new procedure and the appropriateness of the criterion are illustrated with examples.
Article
AN extensive literature exists regarding the relationship of life events and depression. The largest group of studies has concerned the descriptive characterization of those events occurring at the onset of depression. There has been particular emphasis on actual or symbolic losses,1-4 including loss of self-esteem.5 Others have been concerned with the general presence or absence of stress at onset and have attempted to define a group of endogenous depressions, occurring in the absence of stress, and showing characteristic clinical features.6-8 Although it has been generally assumed that most depressions are reactions to events, some dissent has been expressed. Hudgens9 and his colleagues found events uncommon in the six months prior to onset of illness in 40 patients hospitalized with affective disorders. Winokur and Pitts10 reported reactive depressions to be infrequent and threw doubt on the validity of
Article
Recent studies have begun to characterize the actions of stress and antidepressant treatments beyond the neurotransmitter and receptor level. This work has demonstrated that long-term antidepressant treatments result in the sustained activation of the cyclic adenosine 3', 5'-monophosphate system in specific brain regions, including the increased function and expression of the transcription factor cyclic adenosine monophosphate response element-binding protein. The activated cyclic adenosine 3', 5'-monophosphate system leads to the regulation of specific target genes, including the increased expression of brain-derived neurotrophic factor in certain populations of neurons in the hippocampus and cerebral cortex. The importance of these changes is highlighted by the discovery that stress can decrease the expression of brain-derived neurotrophic factor and lead to atrophy of these same populations of stressvulnerable hippocampal neurons. The possibility that the decreased size and impaired function of these neurons may be involved in depression is supported by recent clinical imaging studies, which demonstrate a decreased volume of certain brain structures. These findings constitute the framework for an updated molecular and cellular hypothesis of depression, which posits that stressinduced vulnerability and the therapeutic action of antidepressant treatments occur via intracellular mechanisms that decrease or increase, respectively, neurotrophic factors necessary for the survival and function of particular neurons. This hypothesis also explains how stress and other types of neuronal insult can lead to depression in vulnerable individuals and it outlines novel targets for the rational design of fundamentally new therapeutic agents.
Article
Millennial-scale variability in the flux of ice-rafted detritus to North Atlantic sediments during the last glacial period has been interpreted to reflect a climate-forced increase in the discharge of icebergs from ice-sheet margins surrounding the northern North Atlantic Ocean. But the relationship between ice-sheet variability and climate change is not clear, as both the sources of ice-rafted detritus and the ice-marginal processes are varied and complex. Terrestrial records are helpful in unravelling this complexity because they can demonstrate the scale of ice-sheet oscillations, and whether the ice sheet (or sector) was advancing or retreating with respect to climate change. Here we constrain the age and anatomy of a prominent readvance of the British Ice Sheet in the northern Irish Sea region at ~1414CkyrBP(~16.4 calendar kyr BP). The analysis indicates that the British Ice Sheet participated in an iceberg discharge episode known as Heinrich event 1. Comparison with other terrestrial and marine ice-sheet records suggests that the dynamic collapse of the Laurentide Ice Sheet beginning at 14.6-15.014CkyrBP, (~17.2-17.6 calendar kyr BP) initiated varied responses from other ice-sheet margins around the northern North Atlantic region. These observations support the argument that the release of icebergs and meltwater during Heinrich event 1 disrupted the North Atlantic thermohaline circulation, leading to a delay or reversal of deglaciation of the Northern Hemisphere and at least as far south as 40°S for two to three thousand years,,, suggesting a climate forcing and response similar to that of the ensuing Younger Dryas `cold snap',.