NSC 15510
No. of Pages 16
2 July 2014
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
1
Neuroscience xxx (2014) xxx–xxx
2
NEUROSCIENCE FOREFRONT REVIEW
3
NEUROEPIGENETICS OF STRESS
4
5
6
Q1 B. B. GRIFFITHS AND R. G. HUNTER *
Contents
Introduction
Stress and the hypothalamic–pituitary–adrenal axis
Epigenetic mechanisms
DNA modification
Histone modifications
Non-coding RNA
Other mechanisms and potential
influences on neuroepigenetic marks
Developmental stress
Animal models of developmental stress
Human studies of developmental stress
Prenatal stress in animal models
Prenatal stress in humans
Critical periods
Acute stress
PTSD in humans
Animal models of acute stress
Chronic stress
Animal models of chronic stress
Depression and suicide in humans
Conclusion
Uncited references
References
Q2 University of Massachusetts, Boston, MA, United States
Abstract—Stress, a common if unpredictable life event, can
have pronounced effects on physiology and behavior. Individuals show wide variation in stress susceptibility and
resilience, which are only partially explained by variations
in coding genes. Developmental programing of the
hypothalamic–pituitary–adrenal stress axis provides part
of the explanation for this variance. Epigenetic approaches
have successfully helped fill the explanatory gaps between
the influences of gene and environment on stress responsiveness, and differences in the sequelae of stress across
individuals and generations. Stress and the stress axis
interacts bi-directionally with epigenetic marks within the
brain. It is now clear that exposure to stress, particularly
in early life, has both acute and lasting effects on these
marks. They in turn influence cognitive function and behavior, as well as the risk for suicide and psychiatric disorders
across the lifespan and, in some cases, unto future generations. Ó 2014 Published by Elsevier Ltd. on behalf of
IBRO.
8
00
00
00
00
00
00
9
10
11
12
13
14
15
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
00
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Keywords: histone modifications, DNA methylation, HPA
axis, corticosteroids, non-coding RNA, developmental
programing.
INTRODUCTION
34
Stress is a common, if unpredictable, life event, and can
have both adaptive and maladaptive consequences for
an organism. In the natural environment, many
stressors can have profoundly negative, even lethal,
consequences and for this reason organisms require the
capacity to rapidly and effectively adapt to stressful
circumstances. They also need to keep account of
lessons learned with regard to stress in terms of
memories as well as behavioral and physiologic
adaptations. This requirement for rapid, yet persistent,
change is a challenge both to the largely fixed genome
and to a brain primarily comprised of terminally
differentiated neurons. Mammalian brains meet this
challenge in a variety of interrelated ways, from
structural and functional plasticity to epigenetic
reprograming of neural genomes.
Epigenetics is most broadly defined as transmission of
information above the level of DNA sequence. The term
‘epigenetic’ has evolved substantially since its coinage
decades ago and it now encompasses a range of
effects from behavioral or even cultural transmission of
35
7
*Corresponding author. Address: Department of Psychology, 100
Morrissey Boulevard, Boston, MA 02125, United States. Tel: +1617-287-3198.
E-mail address:
[email protected] (R. G. Hunter).
Abbreviations: ACTH, adrenocorticotropic hormone; AVP, argininevasopressin; BDNF, brain-derived neurotrophic factor; BLA,
basolateral amygdala; COMT, catechol-O-methyltransferase; CpG,
cytosine-guanine
dinucleotides;
CRH,
corticotropin-releasing
hormone; CSD, chronic social defeat; DCS, D-cycloserine; ERK1/2,
extracellular signal-regulated protein kinases 1 and 2; GC,
glucocorticoid; GR, glucocorticoid receptor; H3K4, H3 lysine 4; H3K9,
H3 lysine 9; HATs, histone acetyltransferases; HDACs, histone
deacetylases; HDMs, histone demethylases; HMTs, histone
methyltransferases; HPA, hypothalamic–pituitary–adrenal (HPA); HT,
hypertonic saline solution; lncRNA, long non-coding RNA; LTM, longterm memory; Met, methionine; miRNAs, microRNAs; MR,
mineralocorticoid receptor; MS, maternal separation; NaBu, sodium
butyrate; NMDAR, N-methyl-D-aspartate receptor; OXTR, oxytocin
receptor; piRNA, piwi-interacting RNA; POMC, proopiomelanocortin;
PTSD, post-traumatic stress disorder; PVN, paraventricular nucleus;
SC, skin conductance; snoRNA, small nucleolar RNA; SNP, single
nucleotide polymorphism; TEs, transposable elements; Val, valine;
Q3 VPA, valproic acid.
http://dx.doi.org/10.1016/j.neuroscience.2014.06.041
0306-4522/Ó 2014 Published by Elsevier Ltd. on behalf of IBRO.
1
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
NSC 15510
No. of Pages 16
2 July 2014
2
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
information across and within generations down to the
molecular modifications of nucleic acids and their
packaging proteins. While one might defend making a
hard distinction for molecular epigenetics for the latter
set of phenomena, it is not clear where the line between
the molecular and what it is tempting to call behavioral
epigenetics, properly lies. This is clear even in what is
perhaps the most famous case of epigenetic effects in
the nervous system, maternal transmission of stress
responsiveness described by Meaney and collaborators
(see below) where maternal behavior and molecular
mechanisms are both demonstrably involved. A more
useful distinction might be between the acute changes
in epigenetic marks, which contribute to short-term
defense of homeostasis and longer term responses to
environmental change that might be regarded as
allostatic, to borrow a term from the stress literature
(McEwen and Wingfield, 2010). Allostasis, is defined in
contrast to homeostasis as ‘‘maintaining stability through
change’’ (McEwen and Wingfield, 2003). Homeostatic
epigenetic alterations would return to a baseline level over
a fairly brief period of time (say hours to days), while allostatic changes might persist through the lifespan and into
future generations in the absence of a countermanding
perturbation. It suffices to say that this isnot a definitive
distinction, but it is clear the field is presently large enough
that some more precise terminology is wanting.
The study of neuroepigenetics in general and the
neuroepigenetics of stress has undergone substantial
growth over the last decade (Hunter et al., 2013;
Sweatt, 2013; Reul, 2014). This is due to both the adoption of novel technical approaches such as next generation sequencing, and to the clear need to explain
persistent, environmentally sensitive behavioral variations
that were not due simply to genetic polymorphisms (e.g.
the ‘‘Missing Heredity’’ problem in psychiatric genetics
(Crow, 2011; Danchin et al., 2011)). Stress has long been
known to play a role in brain plasticity (Hunter et al.,
2013). As stress is one of stronger environmental influences on human and animal behavior, it is a logical
means to examine genome-environment interactions with
an epigenetic lens (with the caveat that the distinction
between environmental, genomic and other stochastic
factors is, if anything, less clear in the present age than
previously).
STRESS AND THE HYPOTHALAMIC–
PITUITARY–ADRENAL AXIS
Stress can be conceived of as any threat to bodily
homeostasis—real or imagined—that urges the
organism to act in defense of that homeostasis. These
actions often require behavioral or physiologic changes
on the part of the organism and are therefore referred
to as allostatic (McEwen and Wingfield, 2003), while
the stressor itself represents an allostatic load.
Stressors are defined both by their duration as well as
by the capacity of the organism to respond to them.
Controllable stresses, such as voluntary exercise, can
be both physically and psychologically beneficial
(Adlard and Cotman, 2004; Aschbacher et al., 2013).
Fig. 1. The HPA axis. Stress triggers a cascade of signaling in the
hypothalamic–pituitary–adrenal (HPA) axis. The hypothalamus
releases corticotropin-releasing hormone (CRH) and arginine-vasopressin (AVP) into the pituitary portal. The pituitary converts proopiomelanocortin (POMC) to adrenocorticotropic hormone (ACTH) and
releases it into the bloodstream, where it binds to the adrenal gland.
Corticosteroids released by the adrenal gland bind to glucocorticoid
(GR) and mineralocorticoid (MR) receptors. These receptors inhibit
the HPA axis influencing other brain regions and serotonin receptors
(5-HTR).
Opinion varies on whether such events are properly
called ‘stress’, though they do actuate many of the same
physiologic mechanisms. In contrast, in situations where
the organism does not have meaningful control of the
outcome of the stressor, the effects can be negative
(Maier and Watkins, 2010). This is true of both severe
acute traumas, which can induce Post-Traumatic Stress
Disorder (PTSD), and of more chronically aversive
situations such as social subordination or chronic
unpredictable stress, which can contribute to depressive
behavior. In a successful stress response, once the
individual has escaped the situation, the body will return
to a pre-stress state. In these instances, stress is a
positive response that keeps the individual alive and
well, and can even increase resilience to future
stressors. Both negative and positive stress adaptations
cause the nervous system to undergo epigenetic
changes that influence its future responses.
The mammalian stress response is orchestrated, in
part,
by
the
activity
of
the
hypothalamic–
pituitary–adrenal (HPA) axis (Fig. 1). Perturbation of
glucocorticoid (GC) feedback in the HPA axis is one of
the best-established biomarkers for a number of
complex diseases, including depression and PTSD. The
hypothalamus is the first structure in this common
pathway. Activation of the parvocellular neurons of the
paraventricular nucleus (PVN) of the hypothalamus
induces three distinct responses. Activation of the
pathway begins with the release of corticotropinreleasing hormone (CRH) and arginine-vasopressin
(AVP) into the pituitary portal. This triggers the next step
in the HPA cascade. CRH and AVP release into the
pituitary
portal
increases
the
production
of
proopiomelanocortin (POMC) in the anterior pituitary.
POMC is converted to adrenocorticotropic hormone
(ACTH), which is then released into general circulation
in the blood (Aguilera, 2012).
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
NSC 15510
No. of Pages 16
2 July 2014
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
The adrenal gland produces corticosteroids
(predominantly cortisol in humans, corticosterone in
rodents) in the presence of ACTH. Corticosteroids bind
to mineralocorticoid (MR) and glucocorticoid (GR)
receptors. In the PVN and pituitary GRs have an
inhibitory effect, and this negative feedback closes the
loop and represents the main mechanism of restoring
the HPA axis to baseline levels after a stress response.
MR and GR in other brain regions, such as the
hippocampus, amygdala and prefrontal cortex, also
contribute to the modulation of the HPA axis, spatial
cognition and anxiety behavior (Rozeboom et al., 2007;
Harris et al., 2013). Other portions of the nervous system
also mediate responses to stress, in particular, the locus
coeruleus has noradrenergic projections directly to the
PVN (Sawchenko and Swanson, 1981). Further, the autonomic nervous system also plays a significant role,
releasing catecholamines into the blood and tuning down
housekeeping activities like digestion, while increasing
physiologic responses favorable to fighting or fleeing.
The developmental programing of the HPA axis in
response to stress has been documented for decades
(Liu et al., 1997), although understanding the mechanisms for these stable behavioral modifications remains
incomplete.
177
EPIGENETIC MECHANISMS
178
The term ‘‘epigenetic’’ was initially coined by Waddington
to describe how organisms were able to selectively
modulate a fixed genome to develop a variety of tissues
and cell fates (Waddington, 1942). The meaning of the
term has broadened substantially since the 40’s to
encompass non-genetic transgenerational inheritance,
as well as a number of specific molecular mechanisms
that influence the transcriptional phenotype at a level
above that of the DNA sequence. Epigenetic information
can be passed on to subsequent generations through
fetal programing, behavioral intervention or germline
transmission (Bohacek and Mansuy, 2013). Stress influences each of these types of inheritance, and epigenetic
mechanisms can be used to explain the known developmental and transgenerational programing of the HPA
axis.
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
Fig. 2. Epigenetic marks. Epigenetic modifications include covalent
modifications of histone proteins including acetylation, methylation,
and phosphorylation; methylation of DNA, as well as a suite of
mechanisms such as higherorder chromatin structure and non-coding
RNA (e.g. microRNA (miRNA), piwi-interacting RNA (piRNA), small
nucleolar RNA (snoRNA) and long non-coding RNA (lncRNA)), which
can affect the translation and stability of messenger RNA (mRNA).
3
Cells have several methods of regulating the
expression of genes in response to stress without
actually changing the code of DNA itself. These include
both the epigenetic mechanisms we outline below
(Fig. 2) and the classical transcription factors. At the
molecular level, the line between epigenetic mechanism
and the transcriptional machinery is not easily
established, as the two processes are intimately
intertwined, both structurally and functionally. Most of
following mechanisms adjust the likelihood of a gene
being exposed to transcriptional proteins, but there are
several post-translational mechanisms as well (see
‘Non-coding RNA’).
Gene expression and epigenetic marks are cell and
tissue specific, and changes observed in one area of
the brain may not be the same in another brain
structure (Davies et al., 2012). Likewise, epigenetic
changes in the brain may not be the same as those found
in blood (Axelsen et al., 2007). Because of the difficulties
of examining living human brain tissue, most experiments
in humans use blood to observe epigenetic differences.
The correlation between changes seen in the blood, and
those found in the brain is still unclear. Any test for human
stress disorders will need to utilize epigenetic data from
peripheral tissues to be a viable. However, because
stress induces an endocrine response which reaches
most tissues, it may affect the brain and peripheral tissues
in a similar manner. A recent study found DNA methyla- Q4
tion of Fkbp5 was correlated in the dentate gyrus and
blood, though the pattern was not precisely matched
(Ewald et al., 2014).
194
DNA modification
225
Perhaps the most-studied form of molecular epigenetic
modification is DNA methylation. Methylation occurs on
the cytosine base of DNA, and most often on cytosineguanine dinucleotides (CpG). Generally this decreases
the likelihood that a gene will be transcribed, though
methylation can have a more complex role in
transcription (Jones, 2012). CpG repeats are not equally
distributed throughout the genome, instead being located
mainly within CpG islands. The presence of a CpG site in
a gene represents a greater probability of that gene being
methylated, and methylation can happen in the promoter,
exon, or introns of genes, in that relative order of commonality (Gelfman et al., 2013).
DNA methylation in mammals is regulated by
methyltransferases DNMT1/3a/3b. All three are highly
expressed during postnatal development in rats, and
then decline after 3 weeks, but are still expressed in
mature rat neurons (Miller and Sweatt, 2007; Simmons
et al., 2013). DNMT1 and DNMT3a have been shown to
be active in adult mouse neurons, and loss of function
impairs learning and memory (Feng et al., 2010;
Maddox et al., 2014). DNMT3a is required for de novo
methylation of non-CpG cytosine bases in maturing
mouse neurons (Guo et al., 2014) and emotional regulation (LaPlant et al., 2010). Demethylation is currently less
well understood, with DNA repair mechanisms, GADD45a
and TET family proteins being non-exclusive suspects
(Guo et al., 2011; Niehrs and Schafer, 2012; Delatte
226
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
NSC 15510
No. of Pages 16
2 July 2014
4
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
270
et al., 2014). As yet the former two mechanisms have yet
to be explored in the context of stress neuroepigenetics,
while some suggestive work has been done with the
TET proteins. Knockdown of TET1 results in increased
methylation of brain-derived neurotrophic factor (BDNF)
in the mouse brain (Turner et al., 2010). BDNF is required
for normal brain development and plasticity, and its dysregulation has been noted in several disorders (Castren
and Rantamaki, 2010). There is also evidence that GRs
can influence DNA methylation through their binding to
GC response elements (Thomassin et al., 2001). Hydroxymethylation is a recent addition to the list of known epigenetic mechanisms, and although its exact role is not
quite clear, early evidence suggests it plays an important
role in brain development, and is part of the TET demethylation pathway (Kriaucionis and Heintz, 2009; Wu and
Zhang, 2011; Wang et al., 2012).
271
Histone modifications
272
Histones are the main protein constituents of the
nucleosome, the fundamental packing unit of chromatin.
Nucleosomes are an octamer assembled from four core
histone proteins: H2A, H2B, H3, and H4. Densely packed
and highly methylated chromatin forms transcriptionally
silent heterochromatin, while euchromatin is less tightly
packed and has higher levels of histone acetylation.
Euchromatin is the site of most active gene transcription.
The histone subunits each have an unstructured
N-terminal tail, which may be modified to influence other
proteins’ ability to bind and transcribe DNA. The number
of possible modifications to histones is quite large, and
beyond the scope of this paper (Allis et al., 2007; Tan
et al., 2011), as most neuroepigenetic research focuses
on three main marks: methylation, acetylation, phosphorylation (Jiang et al., 2008; Hunter, 2012). Therefore we will
focus on these marks.
Methylation of histone tails occurs on lysine bases,
which may be mono-, di-, or tri-methylated, or arginine
bases, which may be mono- or di-methylated
(Kouzarides, 2007). Histone methylation is considered a
relatively stable and specific mark, and may either
increase or decrease gene expression, depending on
the specific histone subunit, base residue, and methylation valence. Histone methyltransferases and demethylases (HMTs and HDMs) show a relatively high degree
of specificity with regard to histone subunit and amino
acid residue (Shi et al., 2004). For example, H3 lysine 4
(H3K4) methylation is transcriptionally active (Strahl
et al., 1999), whereas gene silencing occurs from the dior tri-methylation of the H3 lysine 9 (H3K9; Lachner
et al., 2001) or H3K27 (Cao et al., 2002). The histone
demethylase LSD1 acts on H3K4 and is important for
neural differentiation (Sun et al., 2010) and present in
adult rat brains (Zhang et al., 2010). Histone methylation
can vary widely in response to stress, with several histone
subunits becoming more methylated, and others less
methylated (e.g. Hunter et al., 2009).
Histone acetylation generally increases gene
expression by increasing access to genes (Tsankova
et al., 2004). Histone acetyltransferases (HATs) increase
histone acetylation, and histone deacetylases (HDACs)
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
decrease acetylation of histones and other proteins
(Yang and Seto, 2007). It naturally follows that HAT inhibitors decrease histone acetylation and HDAC inhibitors
increase histone acetylation, and both are drugs of
interest for treating stress disorders in different brain
structures. HDAC inhibitors can also contribute to the
demethylation of DNA (Cervoni and Szyf, 2001). Unlike
the specificity of histone methylation, HATs and HDACs
have a broad influence on the entire genome. However,
acetylation control can be linked to other histone marks.
Stress-induced H3K14 acetylation and H3S10 phosphorylation appear to involve actions of GR on the extracellular signal-regulated protein kinases mitogen activated
protein kinase (ERK MAPK) signaling pathway, and this Q5
pathway in turn activates the nuclear kinases MSK1 and
Elk-1 leading to histone phosphor-acetylation and altered
gene expression in the rat dentate (Gutierrez-Mecinas
et al., 2011).
Phosphorylation is a common method of regulating
the activity of proteins in the cell, and can be used to
modify histones as well. Most of what is known about
histone phosphorylation relates to the cell cycle (Nowak
and Corces, 2004), but it has been shown that histones
are differentially phosphorylated in response to stress
(Trollope et al., 2012; Rotllant et al., 2013). Neural activity
can result in phosphorylation of histones in postsynaptic
neurons (Crosio et al., 2003), and forming memories during stressful events is dependent on histone phosphorylation (Reul and Chandramohan, 2007). An important role
of histone phosphorylation may be the effect of cross-talk
with other histone modifications (Ruthenburg et al., 2007;
Banerjee and Chakravarti, 2011). Much more work is
needed to determine the role histone phosphorylation
plays in stress and stress disorders.
314
Non-coding RNA
348
Classically, RNA has been thought of as the product of a
transcribed gene that is eventually translated into a
functional protein. However, there are many RNAs that
do not encode gene products, but still play a regulatory
role in gene expression. MicroRNAs (miRNAs), perhaps
the best known, are present at all stages of neural
development, and contribute to mature neuronal
function (Sun et al., 2013). Their expression can be controlled by behavior, and has been implicated in many
mental disorders (Eacker et al., 2013). Other non-coding
RNAs known to play a role in behavior include piwi-interacting (piRNA; Landry et al., 2013), long non-coding
(nclRNA; Fenoglio et al., 2013), and brain-specific small
nucleolar RNA (snoRNA; Cavaille et al., 2000). Noncoding RNA represents a potential wealth of mechanisms
for stress disorders, and has only just begun to be
studied.
349
Other mechanisms and potential influences on
neuroepigenetic marks
366
Several additional mechanisms, which may either be
regarded as potentially epigenetic or which interact with
epigenetic mechanisms, may play a role in the brain’s
response to stress, but need further elucidation. Prions
368
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
367
369
370
371
NSC 15510
No. of Pages 16
2 July 2014
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
have been implicated in the persistence of memory
(Kandel et al., 2013), and synapse stability (Blaze and
Roth, 2013; Heinrich and Lindquist, 2011). Chromatin
loops are supranucleosomal organizational structures
which can have an effect on gene regulation (Kadauke
and Blobel, 2009), and which may be associated with
the pathogenesis of Rett syndrome (Horike et al., 2005).
Long thought of as ‘‘junk’’ DNA, transposable elements
(TEs) are potentially mobile genomic constituents, which
have long been thought to become active in response to
stressors (McClintock, 1951, 1984; Faulkner et al.,
2009). Transposition appears to play a role in both nervous system development and in adult neurogenesis
(Gage and Muotri, 2012). Recent work has shown that
TEs are actively transcribed in brain and that their RNAs
can have both gene regulatory and pathogenic effects
(Reilly et al., 2013) and that they can be dynamically regulated by environmental stress via histone modification
(Hunter et al., 2012), in addition to more static regulation
by both DNA and histone methylation. Indeed it has been
argued that at least some of these elements may provide
for increased phenotypic plasticity in response to environmental insults (Hunter et al., 2013; Sharif et al., 2013).
The role of nutrition and the effects the microbiome may
exert on the central nervous system have seen an
increase in interest with regard to their potential effects
on epigenetic marks in the nervous system (Burdge and
Lillycrop, 2010; Stilling et al., 2014). DNA methylation,
histone modification, and non-coding RNA are the best
described mechanisms of molecular epigenetic action,
but the above examples show that new sorts of molecular
machinery have the potential to be added to the list in the
near future.
405
DEVELOPMENTAL STRESS
406
418
Stress during development can change how an individual
will respond to stress in the future, through all periods of
life. Recent years have seen the conclusive
demonstration that childhood adversity has profound,
life-long effects on both physical and mental health
(Larkin et al., 2014). While manipulations such as maternal separation can result in a hyper-responsive HPA axis,
more mild stress can actually increase resiliency and temper some of the effects of later life stress (Francis et al.,
1999; Liu et al., 2000; Champagne et al., 2003). Learning
to manage stress in early development through mild exposure is likely needed for learning to cope with stress as an
adult (DiCorcia and Tronick, 2011).
419
Animal models of developmental stress
420
This effect is most classically associated with mother rats’
pup-grooming behavior. Rat pups with mothers less
receptive to nursing and less active in grooming have
lower levels of GRs in the hippocampus and elevated
stress responses in adulthood (Liu et al., 1997; Francis
et al., 1999). The main mechanism of these epigenetic
changes was the mother’s behavior, which was shown
to increase DNA methylation and decrease histone acetylation of the GR promoter 1–7 in the rat hippocampus,
which led to decreased binding of the transcription factor,
407
408
409
410
411
412
413
414
415
416
417
421
422
423
424
425
426
427
428
429
5
NGFI-A (Weaver et al., 2004, 2007). Switching the pups
of lessattentive mothers to highlyattentive mothers produced the same phenotype as the highlyattentive
mother’s biological offspring, providing strong evidence
that inflated responses to stress are not genetically determined, but programed sometime during development.
Maternal separation (MS) is a model of
developmental stress in rodents in which the mother is
removed from her newborn pups for several hours.
Male MS mice exhibited strain-specific differences on
behavioral tests and in epigenetic marks. Though both
strains had hypermethylation of AVP, only C57BL/6J
mice had hypomethylation of the transcription factor
Nr4a1, and only DBA/2J mice showed hypermethylation
of GR (Kember et al., 2012). Human adults who had
experienced childhood trauma also had hypermethylation
of the GR promoter (Tyrka et al., 2012).
MS resulted in sex-specific differences in the
methylation of the CRH promoter (Chen et al., 2012).
When measured at 60 days, both female and male rats
showed increased ACTH and corticosterone levels in
response to restraint stress. Female MS rats had higher
basal levels of corticosterone. Both females and males
had higher levels of CRH messenger RNA (mRNA) in
the hypothalamus after 30 min of restraint. In parallel with
the corticosterone findings, female MS rats had higher
basal levels of CRH mRNA. These alterations were correlated with hypomethylation of the CRH promoter region in
the PVN of the hypothalamus.
BDNF has been consistently associated with stressinduced neuroplasticity throughout development (Gray
et al., 2013a,b). Thus, it is of interest that BDNF is subject
to epigenetic regulation. For example, MS rats had lower
levels of BDNF mRNA and protein expression in the hippocampus than controls, but higher levels of miR-16,
which had a negative association with the expression of
BDNF (Bai et al., 2012). Although there were several
behavioral differences between MS rats and those subjected to chronic stress at 10 weeks, these did not affect
BDNF expression in the chronically stressed group. MS
rats showed more anhedonia and loss of interest, while
chronically stressed rats showed more anxiety or depression-like behaviors.
Rats with ‘abusive’ mothers have been shown to have
decreased levels of BDNF in the frontal cortex, and
increased BDNF methylation at exon IV (Roth et al.,
2009). This lasted into adulthood, and was reversed with
administration of the methylation inhibitor zebularine. Maltreatment by a maternal figure results in widespread
changes in gene expression, and is dependent on the
sex of the offspring (Blaze and Roth, 2013). These results
parallel observations in human adult PTSD patients,
which demonstrate that those patients with a history of
child abuse show markedly different DNA methylation
levels and almost mutually exclusive patterns of gene
expression (98% dissimilar) from those with no such
history (Mehta et al., 2013).
430
Human studies of developmental stress
487
In humans, developmental stress has a clear impact on
mental health in later life, e.g. (Edwards et al., 2003),
488
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
489
NSC 15510
No. of Pages 16
2 July 2014
6
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
however the biological mechanisms are still not entirely
clarified. Nonetheless, epigenetic factors clearly play a
role in susceptibility to stress-related disorders as the
work of Yehuda and others has made clear. More
recently, compelling evidence for a lasting molecular epigenetic impact of early life trauma and adversity has rapidly accumulated (Radley et al., 2011; Klengel et al., 2013;
McGowan, 2013; Mehta et al., 2013; Yehuda et al., 2014).
Recent work in this vein was published by Mehta and
collaborators suggesting that a history of abuse may
constitute a biologically distinct subset of PTSD patients
Q6 from those without such a history. Individuals diagnosed
with PTSD who have a history of childhood trauma had
different levels of expression for 303 RNA transcripts
compared to those who had some degree of trauma but
had not been diagnosed with PTSD (Mehta et al.,
2013). Individuals diagnosed with PTSD without a history
of childhood trauma had 244 differentially expressed transcription profiles. Remarkably, only 14 transcripts overlapped between the two groups, signifying two distinct
biological signatures for individuals diagnosed with PTSD,
depending on their history of trauma. In individuals with
PTSD and a history of child abuse, 69.3% of the genes
for the differentially expressed transcripts had at least
one CpG methylated site, and individuals with PTSD but
no history of child abuse showed CpG methylation in
33.6% of genes that had been tied to differentially
expressed transcription profiles. For genes with five or
more methylated CpG sites, the gap widens to 11.7 and
0.8% for each group, respectively. It appears as if the differences in transcriptional profiles are mainly due to
hypermethylation of DNA as a consequence of childhood
trauma (Mehta et al., 2013).
Hypermethylation of the GR promoter was found in
suicide victims with prior childhood abuse (McGowan
et al., 2009; Labonte et al., 2012a,b). Further, individuals
with depression but no history of child abuse did not show
the same GR promoter methylation (Alt et al., 2010;
Labonte et al., 2012a,b). The severity of traumatic events
experienced in childhood had a direct relationship with the
methylation of GR in adulthood (Perroud et al., 2014).
Child abuse may also result in the up-regulation of
stress-regulatory genes. The AVP promoter and exons
are hypomethylated, leading to decreased binding of the
inhibitory CpG-binding protein MeCP2 and an increase
in HPA activity (Murgatroyd et al., 2009). MeCP2 has also
been linked to Rett syndrome (Kriaucionis and Bird, 2003)
and BDNF expression (Chen et al., 2003; Martinowich
et al., 2003).
539
Prenatal stress in animal models
540
Epigenetic programing is not just limited to post-natal life,
but can begin while still in the womb (Howerton and Bale,
2012). Indeed, the epigenetic influence of stress can exert
effects through multiple generations, as has been shown
by the Meaney group in rodents (Francis et al., 1999).
Prenatal stress follows a similar pattern to newborn
stress: too much leads to structural and behavioral detriments. In response to stress, rat pups of stressed mothers showed a decrease in PVN volume, increased CRF
mRNA expression (Fujioka et al., 1999), decreased
541
542
543
544
545
546
547
548
549
hippocampal neurogenesis and worse outcomes on
memory tasks through all stages of life (Lemaire et al.,
2000). Conversely, rat mothers exposed to mild restraint
stress had pups that were less fearful, learned faster,
and had lower c-Fos expression in the amygdala
(Fujioka et al., 2001).
Mouse fetuses of stressed mothers were shown to
have altered methylation of 11b-hydroxysteroid
dehydrogenase type 2, a corticosterone metabolite
(Jensen Pena et al., 2012), and heightened levels of
DNMT1/3a in GABAergic neurons. Prenatal stress in male
rats produced behavioral deficits and decreased CRF promoter methylation, increased GR exon 1–7 methylation in
the hypothalamus and reduced amygdala CRF promoter
methylation (Mueller and Bale, 2008). Interestingly, these
changes were not observed in females, and the sex-differentiated methylation corresponds to different placental
expression of several genes, such as DNMT1. Prenatal
stress also alters methylation and expression of the structural protein GPM6A (Monteleone et al., 2013). MiR-133b
showed increased levels of expression in the hippocampus, and cultures overexpressing miR-133b showed a
decrease in GPM6A mRNA and function.
550
Prenatal stress in humans
573
The transgenerational effects of stress have also been
observed in humans as a consequence of the
dislocations and atrocities of the Second World War
(Painter et al., 2008). Heightened levels of DNMT1/3a in
GABAergic neurons in human psychiatric patients with
psychotic symptoms were comparable to results found
in prenatally stressed mice (Matrisciano et al., 2013).
Increased HPA activation was also associated with
increased methylation of the GR exon 1–7 homolog (1F)
in human babies exposed to prenatal stress (Oberlander
et al., 2008).
574
Critical periods
585
For both prenatal and neonatal stress, the most important
factor may be the timing of critical periods (Callaghan
et al., 2013). These are set time periods in which stress
causes life-long epigenetic changes and behavioral deficits that may be difficult to overcome, despite how enriching their environment may be later. Critical periods have
been shown to be important in miRNA expression
(Morgan and Bale, 2011), and the methylation of DNA
(Simmons et al., 2012, 2013). If prenatal stress occurs
in the right critical period, epigenetic changes in the male
germline can propagate and spread to the F2 generation
via miRNA expression.
Three non-coding RNAs, miR-322, miR-574–3p, and
miR-873, had higher levels of expression in the males of
the F2 generation (Morgan and Bale, 2011). These three
miRNAs all target the gene for b-glycan, whose function in
the context of stress and plasticity is not completely
understood. Another study using a model of combined
unpredictable maternal stress and maternal separation
found that the F2 generation showed more anxious
behaviors than controls (Gapp et al., 2014). This coincided with the up-regulation of several additional miRNAs
586
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
575
576
577
578
579
580
581
582
583
584
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
NSC 15510
No. of Pages 16
2 July 2014
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
608
609
610
611
612
613
614
in the F1 generation’s sperm: miR-375–3p, miR-375–5p,
miR-200b-3p, miR-672–5p and miR-466–5p. Interestingly, the F3 generation shows similar behavior as the
F2 generation, despite the lack of miRNA upregulation
Q7 in F2 generation’s sperm. Thus the precise mechanisms
for this and other transgenerational effects require further
study.
615
ACUTE STRESS
616
634
Our brains are relatively well-equipped to handle shortterm stress, and normally show few negative
aftereffects, yet even transient stresses can leave a
lasting imprint. Fear conditioning is a robust example of
this, often being acquired in one trial. It can result in
drastic and long-lasting epigenetic changes in several
brain areas including the amygdala (Monsey et al.,
2011), hippocampus (Mizuno et al., 2012), and the mesolimbic dopaminergic system (Pezze and Feldon, 2004).
Several epigenetic regulators have previously been demonstrated to change in response to fear conditioning, such
as methyltransferase expression (Miller and Sweatt,
2007). Though most exposure to fearful situations is
quickly resolved, some individuals continue to relive their
traumatic experiences and suffer from PTSD (Yehuda
and McFarlane, 1995). Because of this discordance,
PTSD is of particular interest to the field of epigenetics
because it has a clear cause-and-effect relationship with
stress (McEwen et al., 2012; Zovkic and Sweatt, 2013).
635
PTSD in humans
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
Epigenetic changes in PTSD have been found at a
number of loci. For example, higher methylation of the
dopamine transporter allele 9r has been associated with
higher incidences of PTSD in human participants
(Chang et al., 2012). In a genome-wide methylation study,
participants with PTSD had different levels of methylation
at 119 CpG sites depending on socioeconomic status,
with 55 sites differing with symptom severity (Uddin
et al., 2013). Further analysis showed that the top interaction differences were involved in synaptic transmission
and neuron projection pathways.
Another study examined veterans with PTSD and
investigated the methylation levels of the GR gene in
response to psychotherapy and found differences in GR
methylation between those who responded to treatment
and those who didnot (Yehuda et al., 2013). The GR promoter region and GR exon 1F were hypermethylated
before treatment in people who later responded to treatment. Furthermore, pre-treatment GR exon 1F methylation was correlated with symptom severity and the
predictive of the amelioration of symptoms post-treatment. Those who responded positively to the treatment
tended to be younger, had the disorder for less time,
and had fewer traumatic events over the course of their
lives. After psychotherapy, the levels of GR methylation
did not change. FKBP5, a regulator of GR binding and
translocation into the nucleus, had lower expression after
treatment in those who responded to therapy.
FKBP5 has several functional mutations (Binder et al.,
Q8 2004). Risk alleles are thought to decrease the functional
7
abilities of GRs in an ultra-short feedback loop, and
carriers of one or more risk allele and exposure to
childhood trauma have an increased risk of PTSD.
Carriers of the risk alleles with no history of child abuse
had decreased diagnoses of PTSD. Risk alleles of
FKBP5 were hypomethylated compared to protective
alleles, with heterozygous individuals resembling riskallele homozygous individuals. This points to
demethylation as the putative source of GR dysregulation
in the risk-allele carriers. The demethylation occurs in GC
response elements of intron 7. Demethylation was also
observed in hippocampal neuronal cultures exposed to
dexamethasone. Pathways affected by the dysregulation
of GRs and FKBP5 include T-cell receptor signaling,
TGF-b signaling, Wnt signaling and pluripotency, and
inflammatory response pathways (Klengel et al., 2013).
Veterans who responded better to PTSD psychotherapy
showed hypomethylation of the FKBP5 promoter
(Yehuda et al., 2013), though the researchers did not look
at different alleles. The same study found that FKBP5
expression was positively correlated with cortisol levels.
Rusiecki et al. (2012) examined the global methylation
of repeating elements in veterans. Those who were diagnosed with PTSD after deployment had stable levels of
LINE-1 methylation, whereas those that did not develop
PTSD had higher levels. Before deployment, veterans
who would later develop PTSD had higher levels of Alu
methylation. These findings suggest a difference in the
global methylation of DNA, and a possible biomarker of
people who might develop, or are currently suffering from,
combat-related PTSD.
Stress has been shown to increase antisocial, and
decrease prosocial, behaviors in part through the
regulation of the oxytocin system (Olff et al., 2013).
Understanding the epigenetic regulation of oxytocin represents a significant step toward elucidating the negative
social effects of stress. The oxytocin receptor (OXTR) is
encoded by a single gene (Zingg and Laporte, 2003).
Expression of OXTR is influenced by two transcription
factors, ERa and SP1 (Safe and Kim, 2008). The OXTR
promoter region has several CpG sites, two of which correspond with ERa and SP1 binding sites (1 and 7, respectively), and methylation of CpG 1 or 7 is tissue specific
and influences the expression of OXTR in different brain
regions (Mamrut et al., 2013). Despite sex-specific roles
of oxytocin, OXTR methylation is sex-independent. OXTR
is differentially methylated in several brain areas, including the ventromedial hypothalamus (Harony-Nicolas
et al., 2014), which has been shown to have inhibitory
effects on the HPA pathway (Suemaru et al., 1995).
Healthy human adults subjected to the Trier social
stress test showed differentially methylated OXTR in
blood samples (Unternaehrer et al., 2012) collected pretest, 10 min post-test, and at 90 min. Post-test samples
showed higher levels of methylated CpG sites associated
with one region of the gene, while the 90-min point
showed a large decrease in methylation at both sites
examined. The extent to which peripheral measures such
as these correlate to central phenomena is a common
question in studies of this sort. However, these results
suggest that the genes responsible for the regulation of
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
691
692
693
694
695
696
697
698
699
700
701
702
703
704
705
706
707
708
709
710
711
712
713
714
715
716
717
718
719
720
721
722
723
724
725
726
NSC 15510
No. of Pages 16
2 July 2014
8
727
728
729
730
731
732
733
734
735
736
737
738
739
740
741
742
743
744
745
746
747
748
749
750
751
752
753
754
755
756
757
758
759
760
761
762
763
764
765
766
767
768
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
affiliative behavior are under dynamic epigenetic control
in response to social stress.
Early epigenetic investigations revealed that a number
of psychiatric drugs in common use, such as valproate,
clozapine and imipramine had epigenetic activities (Boks
et al., 2012). As a consequence, many researchers are
exploring the use of HDAC inhibitors and other epigenetically active drugs to increase acetylation of histones with
the goal of understanding its effect in stress-related disorders. Human participants given the HDAC inhibitor valproic acid (VPA) had lower skin conductance (SC)
responses in response to conditioned fear after sleep,
and when given an hour and a half before testing, VPA
had an anxiolytic effect (Kuriyama et al., 2012). VPA
administration before trials also decreased learning acquisition and increased extinction on a SC response compared VPA administration directly after trials.
Participants who were given VPA before sleep had lower
SC responses than those who were given VPA and
stayed awake and the control sleep group. Conversely,
participants given D-cycloserine (DCS), a partial
N-methyl-D-aspartate receptor receptor (NMDAR) agonist, had higher SC responses in the awake group compared to the sleep group, and higher than the waking
controls, but DCS had no effect on learning and extinction
trials. NMDAR blockade has previously been shown to
prevent memory formation and BDNF expression (Lubin
et al., 2008).
A valine (Val) to methionine (Met) single nucleotide
polymorphism (SNP) in the gene that encodes catecholO-methyltransferase (COMT), which removes dopamine
from synapses that lack dopamine transporters (Chen
et al., 2004), resulted in differentially methylated COMT
gene (Norrholm et al., 2013). This SNP resulted in behavior deficits in the inhibition of fear stimuli. Individuals with
PTSD and homozygous Met/Met responded more fearfully to ‘‘safe’’ stimuli than individuals with Val/Met or
Val/Val genotypes. They also responded more fearfully
to ‘‘safe’’ stimuli than individuals who were not diagnosed
with PTSD and shared the Met/Met genotype. Participants who were more fearful also have higher levels of
COMT DNA methylation.
769
Animal models of acute stress
770
Animal models have contributed substantially to our
understanding of both molecular and behavioral
epigenetics and acutely stressful manipulations have
shown a number of such effects on the brain and
behavior. Many of these studies have used psychiatric
drugs with known interactions with epigenetic writer or
eraser enzymes, however, given that many of these
drugs
have
well
described
interactions
with
neurotransmitter systems, for example, interpreting their
effects in an exclusively epigenetic light not always
tenable. BDNF is a target of GR transactivation and
differentially expressed in response to stress (Bennett
and Lagopoulos, 2014). Histone H4 acetylation and its
influence on BDNF expression may contribute to the
extinction of fear memory associated with VPA administration before sleep (Bredy et al., 2007). Female mice
771
772
773
774
775
776
777
778
779
780
781
782
783
784
785
exhibited increased methylation of BDNF in response to
fear, and activating the expression of BDNF blocked the
return of fear memory (Baker-Andresen et al., 2013).
Similar results were seen in rats treated with
vorinostat, another HDAC inhibitor. In combination with
extinction training, the rats had reduced freezing times,
increased H3 and H4 acetylation, and changes in
NMDAR subunit prevalence (Matsumoto et al., 2013).
Rats in the stress and vorinostat group had increased levels of the NMDAR sub-unit, NR2B, mRNA and associated
protein (but not NR1 or NR2A) in the hippocampus compared to just the vorinostat group. Stress and vorinostat
also increased the levels of CaMKIIa and CaMKIIb
proteins compared to the stress and vehicle group.
Upregulation of NR2B has been previously shown to be
associated with fear extinction (Tang et al., 1999), and
CaMKII is known to cooperate with NMDARs.
Rats given a high dose (1200 mg/kg) of the HDAC
inhibitor sodium butyrate (NaBu) had increased plasma
ACTH, corticosterone, and glucose levels when
measured 60 min after administration compared to
controls (Gagliano et al., 2013). To test NaBu’s dose effect,
the researchers administered a low dose of NaBu (200 mg/
kg), high dose of NaBu (1200 mg/kg), and a hypertonic saline solution (HT). Only the high-dose group showed
increased plasma levels of ACTH, corticosterone, and glucose 120 min post-injection compared to a low dose and
the HT group. C-Fos expression was elevated for the high
dose of NaBu compared to controls, but did not differ from
the HT group in the central amygdala and supraoptic
nucleus. NaBu did, however, result in elevated c-Fos
expression above the HT group in the PVN and the ventral
lateral septum, pointing to these brain areas as putative
sites for NaBu’s influence on the stress response.
As one might anticipate, HAT inhibition produces
cognitive and behavioral effects generally opposite of
those produced by HDAC inhibition. Garcinol is a
naturallyoccurring HAT inhibitor derived from a species
of mangosteen used in South Asian cuisine. Rats
injected with garcinol in the lateral nucleus of the
amygdala had lower levels of histone H3 acetylation,
impaired long-term memory (LTM), and failed to
consolidate previously learned material (Maddox et al.,
2013). When injected 1 h after fear conditioning, rats
showed no difference in freezing behavior. They did show
impaired LTM a day later, but not when injected with garcinol 6 h after the fear-condition training. Likewise, rats
subjected to fear conditioning the following day and then
given an injection of garcinol showed impaired LTM of
the recalled fear memory. As in the original condition, rats
that received the injection and failed to consolidate memory also had lower levels of H3 acetylation.
Manipulating histone modifications through drugs is
an exciting new field of research for those suffering from
stress and anxiety disorders. While HATs and HDACs
are not fully understood, especially since part of their
mechanism of action may involve the acetylation of nonhistone proteins, they have proven to be effective at
altering learning and memory. HAT inhibitors seem to
interfere with the consolidation or reconsolidation of
memories, while HDAC inhibitors increase the
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
786
787
788
789
790
791
792
793
794
795
796
797
798
799
800
801
802
803
804
805
806
807
808
809
810
811
812
813
814
815
816
817
818
819
820
821
822
823
824
825
826
827
828
829
830
831
832
833
834
835
836
837
838
839
840
841
842
843
844
845
846
NSC 15510
No. of Pages 16
2 July 2014
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
847
848
849
850
851
852
853
854
855
856
857
858
859
860
861
862
863
864
865
866
867
868
869
870
871
872
873
874
875
876
877
878
879
880
881
882
883
884
885
886
887
888
889
890
891
892
893
894
895
896
897
898
899
900
901
902
903
904
905
906
907
production of neuroprotective proteins, and have even
been shown to reverse some effects of neuronal loss
and recover some brain plasticity (Fischer et al., 2007).
The HDAC2 gene has been shown to regulate memory
and plasticity through control of glutamate receptors and
BDNF (Guan et al., 2009), and HDAC activity is required
for BDNF functioning (Calfa et al., 2012).
NF-jB is a master transcriptional regulator of
inflammatory genes and implicated in depression (Miller
et al., 2009). The NF-jB pathway is highly regulated by
stress, and different NF-jB family members are differentially expressed in the brain as a consequence of acute
and chronic stress (Gray et al., 2013a,b). Nfkbia showed
increased transcription in male mice in response to a
force swim test. Auditory fear memory was blocked from
reconsolidation by IjB kinase inhibitor (sulfasalazine) or
NF-jB inhibitor (SN50) administration in the basolateral
amygdala (BLA), and was rescued by pretreatment with
NaBu (Si et al., 2012). Rats spent less time freezing
24 h after intra-BLA sulfasalazine, but not after treatment
of the central nucleus of the amygdala, and only when
administered directly after re-exposure to the fear stimulus. When rats were treated with SN50 2 h before reexposure, it had a similar effect. Both drugs reduced
freezing behaviors through the presumed increased activation of NF-jB. Treatment with the HDAC inhibitor NaBu
restored memory retrieval and brought freezing behavior
back up to the level seen in controls.
Two hours of restraint produced a substantial increase
in corticosterone levels in rats, and resulted in the
decreased expression of miR-135a and miR-124 in the
amygdala (Mannironi et al., 2013), both of which had previously been shown to block the translation of the MR
mRNA, though not its transcription (Sõber et al., 2010).
However, Mannironi et al. found that overexpression of
miR-135a led to a marked decrease in the expression of
the Nr3c2 mRNA, which encodes the MR protein
(Mannironi et al., 2013). Transfecting cells that did not
endogenously express either miR-124 or miR-135a
resulted in the decrease in MR levels. Transfected cells
expressing either miR-124 or miR-135a had lower levels
of MR protein, and perhaps more importantly, cells with
both showed a further decrease beyond either one alone.
Inhibiting the miRNAs resulted in the increased expression of MR proteins. In addition to MR inhibitory action,
miR-124 has been shown to directly inhibit the translation
of GR mRNA (Vreugdenhil et al., 2009).
Other miRNAs have been shown to be affected by
the acute stress response. The GR mRNA is blocked
from translation by miR-18a in rats and humans
(Turner and Muller, 2005; Uchida et al., 2008). The
amygdala is the target of several differentiallyexpressed
miRNAs in response to stress, including miR-34, miR134, and miR-183 (Meerson et al., 2010; Haramati
et al., 2011). Increased miR-132 expression in response
to stress reduced levels of acetylcholinesterase (AChE)
in rats (Shaltiel et al., 2013). Stress in mice and rats
has also been shown to have an effect on several miRNAs in the hippocampus, frontal cortex, and cerebellum
(Mongrain et al., 2010; Rinaldi et al., 2010; Babenko
et al., 2012).
9
Transposable elements represent a new frontier for
both epigenetics and genomics in the brain, and recent
work has shown that they are clearly regulated by
stress. Acute stress also has impacts on TEs. Stress
increases
expression
of
the
methyltransferase
Suv39h2, increasing H3K9 trimethylation at tens of
thousands of TEs in the hippocampus, in less than 2 h,
consequently decreasing TE RNA expression (Hunter
et al., 2012). In a model of PTSD, stress also results
in the differential regulation of TEs in the amygdala
(Ponomarev et al., 2010).
908
CHRONIC STRESS
919
Many stressors are temporary in nature; however, some
are not. Chronic and acute stress affects both common
and divergent pathways and mechanisms, including
epigenetic ones such as differential HDAC activity
(Renthal et al., 2007). When the organism is overtaxed
through chronic stress, it can have deleterious effects
on mood and behavior. Sometimes the stressful event is
external, such as poverty, and sometimes it is a persistent
internal state, such as major depression. Stable changes
to the epigenome represent a likely mechanism for
chronic stress disorders, and an interesting avenue for
possible therapeutic intervention.
920
Animal models of chronic stress
932
Chronic social defeat (CSD) is a model of chronic stress
and depression with high face and ethological validity
(Nestler and Hyman, 2010), which results in widespread
changes in gene expression, histone methylation, and
phospho-CREB binding in the nucleus accumbens, all of
which showed reversal through treatment with imipramine
(Wilkinson et al., 2009). CSD increased repressive methylation of the BDNF promoter in the hippocampus, which
was reversed by antidepressants (Tsankova et al., 2006).
CSD has also been shown to regulate DNMT3a in the
nucleus accumbens, and exogenous DNMT3a administration causes depressive-like behavior in rats (LaPlant
et al., 2010). CSD decreased methylation of CRF is n
mice. CRF is not hypomethylated in resilient mice
(Elliott et al., 2010). The epigenetic and behavioral
changes that followed treatment with imipramine were
similar to that in resilient mice (Wilkinson et al., 2009).
Submaximal social defeat was not enough to produce
the changes seen in CSD, unless accompanied by
cocaine administration (Covington et al., 2011). Cocaine’s
action on H3K9me2 and G9a mirrored a local knockout of
G9a, which showed behavioral similarity to CSD and
depression in humans. G9a is also responsible for the
antidepressant-caused
inhibition
of
DMNT1
(Zimmermann et al., 2012).
Rats exposed to chronic variable stress spent less
time exploring novel objects, had a corresponding
reduction in extracellular signal-regulated protein
kinases 1 and 2 (ERK1/2) in the hippocampus (which
leads to lower levels of downstream products such as
the anti-apoptotic protein Bcl-2 and an increase in
SIRT1, an HDAC) and a decrease in histone H4
933
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
909
910
911
912
913
914
915
916
917
918
921
922
923
924
925
926
927
928
929
930
931
934
935
936
937
938
939
940
941
942
943
944
945
946
947
948
949
950
951
952
953
954
955
956
957
958
959
960
961
962
963
964
NSC 15510
No. of Pages 16
2 July 2014
10
965
966
967
968
969
970
971
972
973
974
975
976
977
978
979
980
981
982
983
984
985
986
987
988
989
990
991
992
993
994
995
996
997
998
999
1000
1001
1002
1003
1004
1005
1006
1007
1008
1009
1010
1011
1012
1013
1014
1015
1016
1017
1018
1019
1020
1021
1022
1023
1024
1025
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
acetylation. These changes are reversed by treatment
with sirtinol, a SIRT1 inhibitor (Ferland et al., 2013). Rats
exposed to chronic stress spent less time exploring a
novel item in their cage compared to non-stressed animals, and showed no preference for it over a previously
seen object. This behavioral difference was matched with
a reduction in ERK1/2 in the CA1, CA3, and dentate gyrus
regions of the hippocampus. The reduction in Bcl-2 is
associated with an increase in the acetylation of the
Bcl-2 promoter in the dentate gyrus, but not in other
genes important for long-term memory. Previous
research had shown an increase in SIRT1 activity in the
dentate gyrus in response to chronic stress (Ferland
and Schrader, 2011). The current study found an injection
of sirtinol into the dentate gyrus did not affect the HPA
stress response, but did result in decreased SIRT1 activity. Sirtinol caused a further decrease in SIRT1 activity of
chronically stressed animals relative to sirtinol-treated
controls. The injection of sirtinol reversed the decrease
in H4 acetylation and the decrease in ERK1/2 levels seen
in chronically stressed animals, and increased Bcl-2
expression. Sirtinol also rescued the behavioral deficits
seen in novel object exploration and restored preference
for sucrose. Overall, this points to SIRT1 activity as a
mechanism of chronic stress-induced deficits in behavior,
ERK1/2 activity, Bcl-2 expression, and the acetylation of
histone H4.
Male rats subjected to 6 weeks of stress before being
bred sired offspring with blunted stress responses (lower
corticosterone levels) which could not be accounted for
by changes in single gene expression, but did
correspond in increased levels of nine miRNAs
(Rodgers et al., 2013). Despite the lower levels of corticosterone in response to brief restraint, the offspring of
stressed rats did not behave differently in a battery of
behavioral tests compared to controls, nor did they have
changes in gene expression of CRFr1, POMC, Mc2r,
11bHSD-1. Treatment with SSRIs did not affect the offspring of paternally stressed rats differently than controls.
Paternal stress enriched GSEA gene sets c3 and c5. Nine
miRNAs had higher levels of expression in offspring of
paternally stressed rats. These results were similar both
in paternal rats stressed throughout puberty, or only in
adulthood. Sperm do not have many of the epigenetic regulatory elements that other cells have, but do carry miRNAs that are transmitted during fertilization. Some
miRNAs are necessary for successful development (Liu
et al., 2012). Male mice who experienced MS also had
changes in sperm DNA methylation (Franklin et al.,
2010), although it is still unclear on how much effect this
method of epigenetic transmission has on offspring. MiRNA also presents itself as a target of anti-depressants,
and there is some evidence that electroconvulsive shock
therapy influences expression of miRNAs (O’Connor
et al., 2013).
A mouse model of psychotic depression with nonfunctioning DISC1 alleles was exposed to isolation
stress in adolescence for 3 weeks, and as an adult,
showed a GR-controlled hypermethylation of the gene
that encodes tyrosine hydroxylase, an important enzyme
in the production of dopamine (Niwa et al., 2013).
Depression and suicide in humans
1026
In depressed patients, BDNF promoter methylation was
positively associated with suicide ideation and attempts.
Lower methylation of the BDNF promoter reflects poorer
treatment outcomes (Kang et al., 2013). Previous studies
have shown higher methylation of the BDNF promoter
region in suicide completers (Keller et al., 2010). Chronic
stress can result in both hyper- and hypomethylation of
BDNF in different areas of the same brain structure
(Roth et al., 2011), but generally lower BDNF methylation
in blood samples is associated with depressive symptoms
(Fuchikami et al., 2011).
1027
CONCLUSION
1038
The study of epigenetics has substantially enriched our
understanding of the mechanisms by which the brain
adapts to stress. Particularly striking are the numerous
observations that some epigenetic changes caused by
stressors and environmental insults may be passed on
to subsequent generations (Skinner et al., 2008; Morgan
and Bale, 2011; Crews et al., 2012), either through maternal care (Francis et al., 1999) or other epigenetic mechanisms (Guerrero-Bosagna et al., 2010; Rodgers et al.,
2013). Paternal contributions to the development of a
fetus are often overlooked, but recent findings of spermbased epigenetic inheritance may provide insight into previously unexplained issues, e.g. (Dietz et al., 2011).
While transgenerational effects are remarkable for
their persistence, epigenetic mechanisms also underlie
some of our ability to flexibly adapt to and overcome
stressful circumstances. Resilience to stress can be
instilled through controlled and controllable exposure to
stress (Russo et al., 2012). For example, rats exposed
to levels of prenatal stress that would normally result in
behavioral deficits later in life can have those deficits rescued by postnatal handling (Lemaire et al., 2006).
We have chosen to organize this review on a temporal
basis between acute and chronic stressors as the study
has long made clear that the two types of stress have
distinct and often opposed effects on the organism. It is
apparent that the epigenetic impacts of stress differ
between acute and chronic manipulations, often
producing opposing effects, and this seems to be the
case for the effects of some drugs of abuse as well
(Nestler, 2014). Indeed, recent work from our group has
clearly shown that there is very little overlap in the
changes in gene expression produced by acute and
chronic stress in the rodent hippocampus, as the two conditions overlap by less than 4% (Gray et al., 2013a,b).
This suggests, given the impact of epigenetic marks on
transcriptional landscapes, that the distinction between
acute and chronic exposures is born out in epigenetics
as it has been in other aspects of stress biology. Of
course, these effects, though opposed in direction may
involve differing levels of the same mark and global levels
of a particular mark are likely not as informative as the
overall architecture of interaction between epigenetic
modifications, transcription factors etc. which are highly
computationally complex and as yet poorly understood.
1039
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
1028
1029
1030
1031
1032
1033
1034
1035
1036
1037
1040
1041
1042
1043
1044
1045
1046
1047
1048
1049
1050
1051
1052
1053
1054
1055
1056
1057
1058
1059
1060
1061
1062
1063
1064
1065
1066
1067
1068
1069
1070
1071
1072
1073
1074
1075
1076
1077
1078
1079
1080
1081
1082
1083
NSC 15510
No. of Pages 16
2 July 2014
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
1109
The extent to which this adaptive flexibility can be
enhanced
with
epigenetically
targeted
pharmacotherapies in concert with structured behavioral
interventions such as cognitive behavioral therapy
remains to be seen, but the fact that so many of the
drugs in common psychiatric use have epigenetic
activity (Boks et al., 2012) suggests that an answer could
be extracted via meta-analysis of existing clinical data
sets, though these will not be adequate to determine
mechanism. It is worth noting that the discovery of the epigenetic activity of a number of psychiatric drugs only adds
to the ambiguities in our understanding of their actions. An
example of this might be any of a number of antidepressants, which, despite their immediate action on what are
thought to be their clinically important molecular targets
(e.g. the serotonin transporter), still take weeks to act.
Much mechanistic pharmacology remains to be done
before epigenetic pharmacology can be put on a firm footing, yet it is clear that the field is worthy of the effort.
Much systematic work remains to be done to
understand the complex relations between the genome,
environment and epigenetic factors can be fully
appreciated in the context of the neurobiology of stress,
but the past decade of work has clearly established the
importance of these mechanisms to our understanding
of the impact of stress in disease and heredity.
1110
UNCITED REFERENCES
1084
1085
1086
1087
1088
1089
1090
1091
1092
1093
1094
1095
1096
1097
1098
1099
1100
1101
1102
1103
1104
1105
1106
1107
1108
1111
Q9 Francis and Meaney (1999), Hunter and McEwen (2013).
1112
REFERENCES
1113
1114
1115
1116
1117
1118
1119
1120
1121
1122
1123
1124
1125
1126
1127
1128
1129
1130
1131
1132
1133
1134
1135
1136
1137
1138
1139
1140
1141
1142
1143
1144
1145
Adlard PA, Cotman CW (2004) Voluntary exercise protects against
stress-induced decreases in brain-derived neurotrophic factor
protein expression. Neuroscience 124(4):985–992.
Aguilera G. (2012) The hypothalamic-pituitary-adrenal axis and the
neuroendocrine response to stress. In: Handbook of
neuroendocrinology, (Fink G, Pfaff DW, Levine JE, eds), pp.
175–196. New York: Elsevier.
Allis CD, Berger SL, Cote J, Dent S, Jenuwien T, Kouzarides T, Pillus
L, Reinberg D, Shi Y, Shiekhattar R, Shilatifard A, Workman J,
Zhang Y (2007) New nomenclature for chromatin-modifying
enzymes. Cell 131(4):633–636.
Alt SR, Turner JD, Klok MD, Meijer OC, Lakke EA, Derijk RH, Muller
CP (2010) Differential expression of glucocorticoid receptor
transcripts in major depressive disorder is not epigenetically
programmed. Psychoneuroendocrinology 35(4):544–556.
Aschbacher K, O’Donovan A, Wolkowitz OM, Dhabhar FS, Su Y,
Epel E (2013) Good stress, bad stress and oxidative stress:
insights
from
anticipatory
cortisol
reactivity.
Psychoneuroendocrinology 38(9):1698–1708.
Axelsen JB, Lotem J, Sachs L, Domany E (2007) Genes
overexpressed in different human solid cancers exhibit different
tissue-specific expression profiles. Proc Natl Acad Sci USA
104(32):13122–13127.
Babenko O, Golubov A, Ilnytskyy Y, Kovalchuk I, Metz GA (2012)
Genomic and epigenomic responses to chronic stress involve
miRNA-mediated programming. PLoS One 7(1):e29441.
Bai M, Zhu X, Zhang Y, Zhang S, Zhang L, Xue L, Yi J, Yao S, Zhang
X (2012) Abnormal hippocampal BDNF and miR-16 expression is
associated with depression-like behaviors induced by stress
during early life. PLoS One 7(10):e46921.
Baker-Andresen D, Flavell CR, Li X, Bredy TW (2013) Activation of
BDNF signaling prevents the return of fear in female mice. Learn
Mem 20(5):237–240.
11
Banerjee T, Chakravarti D (2011) A peek into the complex realm of
histone phosphorylation. Mol Cell Biol 31(24):4858–4873.
Bennett MR, Lagopoulos J (2014) Stress and trauma: BDNF control
of dendritic-spine formation and regression. Prog Neurobiol
112:80–99.
Blaze J, Roth TL (2013) Exposure to caregiver maltreatment alters
expression levels of epigenetic regulators in the medial prefrontal
cortex. Int J Dev Neurosci 31(8):804–810.
Bohacek J, Mansuy IM (2013) Epigenetic inheritance of disease and
disease risk. Neuropsychopharmacology 38(1):220–236.
Boks MP, de Jong NM, Kas MJ, Vinkers CH, Fernandes C, Kahn RS,
Mill J, Ophoff RA (2012) Current status and future prospects for
epigenetic psychopharmacology. Epigenetics 7(1):20–28.
Bredy TW, Wu H, Crego C, Zellhoefer J, Sun YE, Barad M (2007)
Histone modifications around individual BDNF gene promoters in
prefrontal cortex are associated with extinction of conditioned
fear. Learn Mem 14(4):268–276.
Burdge GC, Lillycrop KA (2010) Nutrition, epigenetics, and
developmental plasticity: implications for understanding human
disease. Annu Rev Nutr 30:315–339.
Calfa G, Chapleau CA, Campbell S, Inoue T, Morse SJ, Lubin FD,
Pozzo-Miller L (2012) HDAC activity is required for BDNF to
increase quantal neurotransmitter release and dendritic spine
density
in
CA1
pyramidal
neurons.
Hippocampus
22(7):1493–1500.
Callaghan BL, Graham BM, Li S, Richardson R (2013) From
resilience to vulnerability: mechanistic insights into the effects of
stress on transitions in critical period plasticity. Front Psychiatry
4:90.
Cao R, Wang L, Wang H, Xia L, Erdjument-Bromage H, Tempst P,
Jones RS, Zhang Y (2002) Role of histone H3 lysine 27
methylation
in
Polycomb-group
silencing.
Science
298(5595):1039–1043.
Castren E, Rantamaki T (2010) The role of BDNF and its receptors in
depression and antidepressant drug action: reactivation of
developmental plasticity. Dev Neurobiol 70(5):289–297.
Cavaille J, Buiting K, Kiefmann M, Lalande M, Brannan CI,
Horsthemke B, Bachellerie JP, Brosius J, Huttenhofer A (2000)
Identification of brain-specific and imprinted small nucleolar RNA
genes exhibiting an unusual genomic organization. Proc Natl
Acad Sci USA 97(26):14311–14316.
Cervoni N, Szyf M (2001) Demethylase activity is directed by histone
acetylation. J Biol Chem 276(44):40778–40787.
Champagne FA, Francis D, Mar A, Meaney MJ (2003) Variations in
maternal care in the rat as a mediating influence for the effects of
environment on development. Physiol Behav 79(3):359–371.
Chang SC, Koenen KC, Galea S, Aiello AE, Soliven R, Wildman DE,
Uddin M (2012) Molecular variation at the SLC6A3 locus predicts
lifetime risk of PTSD in the Detroit Neighborhood Health Study.
PLoS One 7(6):e39184.
Chen WG, Chang Q, Lin Y, Meissner A, West AE, Griffith EC,
Jaenisch R, Greenberg ME (2003) Derepression of BDNF
transcription involves calcium-dependent phosphorylation of
MeCP2. Science 302(5646):885–889.
Chen J, Lipska BK, Halim N, Ma QD, Matsumoto M, Melhem S,
Kolachana BS, Hyde TM, Herman MM, Apud J, Egan MF,
Kleinman JE, Weinberger DR (2004) Functional analysis of
genetic variation in catechol-O-methyltransferase (COMT):
effects on mRNA, protein, and enzyme activity in postmortem
human brain. Am J Hum Genet 75(5):807–821.
Chen J, Evans AN, Liu Y, Honda M, Saavedra JM, Aguilera G (2012)
Maternal deprivation in rats is associated with corticotrophinreleasing hormone (CRH) promoter hypomethylation and
enhances CRH transcriptional responses to stress in adulthood.
J Neuroendocrinol 24(7):1055–1064.
Covington 3rd HE, Maze I, Sun H, Bomze HM, DeMaio KD, Wu EY,
Dietz DM, Lobo MK, Ghose S, Mouzon E, Neve RL, Tamminga
CA, Nestler EJ (2011) A role for repressive histone methylation in
cocaine-induced vulnerability to stress. Neuron 71(4):656–670.
Crews D, Gillette R, Scarpino SV, Manikkam M, Savenkova MI,
Skinner MK (2012) Epigenetic transgenerational inheritance of
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
1146
1147
1148
1149
1150
1151
1152
1153
1154
1155
1156
1157
1158
1159
1160
1161
1162
1163
1164
1165
1166
1167
1168
1169
1170
1171
1172
1173
1174
1175
1176
1177
1178
1179
1180
1181
1182
1183
1184
1185
1186
1187
1188
1189
1190
1191
1192
1193
1194
1195
1196
1197
1198
1199
1200
1201
1202
1203
1204
1205
1206
1207
1208
1209
1210
1211
1212
1213
1214
1215
1216
NSC 15510
No. of Pages 16
2 July 2014
12
1217
1218
1219
1220
1221
1222
1223
1224
1225
1226
1227
1228
1229
1230
1231
1232
1233
1234
1235
1236Q10
1237
1238
1239
1240
1241
1242
1243
1244
1245
1246
1247
1248
1249
1250
1251
1252
1253
1254
1255
1256
1257
1258
1259
1260
1261
1262
1263
1264
1265
1266
1267
1268
1269
1270
1271
1272
1273
1274
1275
1276
1277
1278
1279
1280
1281
1282
1283
1284
1285
1286
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
altered stress responses. Proc Natl Acad Sci USA
109(23):9143–9148.
Crosio C, Heitz E, Allis CD, Borrelli E, Sassone-Corsi P (2003)
Chromatin remodeling and neuronal response: multiple signaling
pathways induce specific histone H3 modifications and early gene
expression in hippocampal neurons. J Cell Sci 116(Pt
24):4905–4914.
Crow TJ (2011) ’The missing genes: what happened to the heritability
of psychiatric disorders?’. Mol Psychiatry 16(4):362–364.
Danchin E, Charmantier A, Champagne FA, Mesoudi A, Pujol B,
Blanchet S (2011) Beyond DNA: integrating inclusive inheritance
into an extended theory of evolution. Nat Rev Genet
12(7):475–486.
Davies MN, Volta M, Pidsley R, Lunnon K, Dixit A, Lovestone S,
Coarfa C, Harris RA, Milosavljevic A, Troakes C, Al-Sarraj S,
Dobson R, Schalkwyk LC, Mill J (2012) Functional annotation of
the human brain methylome identifies tissue-specific epigenetic
variation across brain and blood. Genome Biol 13(6):R43.
Delatte B, Deplus R, Fuks F (2014) Playing TETris with DNA
modifications. EMBO J.
DiCorcia JA, Tronick E (2011) Quotidian resilience: exploring
mechanisms that drive resilience from a perspective of everyday
stress and coping. Neurosci Biobehav Rev 35(7):1593–1602.
Dietz DM, Laplant Q, Watts EL, Hodes GE, Russo SJ, Feng J,
Oosting RS, Vialou V, Nestler EJ (2011) Paternal transmission of
stress-induced pathologies. Biol Psychiatry 70(5):408–414.
Eacker SM, Dawson TM, Dawson VL (2013) The interplay of
microRNA and neuronal activity in health and disease. Front
Cell Neurosci 7:136.
Edwards VJ, Holden GW, Felitti VJ, Anda RF (2003) Relationship
between multiple forms of childhood maltreatment and adult
mental health in community respondents: results from the adverse
childhood experiences study. Am J Psychiatry 160(8):1453–1460.
Elliott E, Ezra-Nevo G, Regev L, Neufeld-Cohen A, Chen A (2010)
Resilience to social stress coincides with functional DNA
methylation of the Crf gene in adult mice. Nat Neurosci
13(11):1351–1353.
Ewald ER, Wand GS, Seifuddin F, Yang X, Tamashiro KL, Potash JB,
Zandi P, Lee RS (2014) Alterations in DNA methylation of Fkbp5
as a determinant of blood-brain correlation of glucocorticoid
exposure. Psychoneuroendocrinology 44:112–122.
Faulkner GJ, Kimura Y, Daub CO, Wani S, Plessy C, Irvine KM,
Schroder K, Cloonan N, Steptoe AL, Lassmann T, Waki K, Hornig
N, Arakawa T, Takahashi H, Kawai J, Forrest AR, Suzuki H,
Hayashizaki Y, Hume DA, Orlando V, Grimmond SM, Carninci P
(2009) The regulated retrotransposon transcriptome of
mammalian cells. Nat Genet 41(5):563–571.
Feng J, Zhou Y, Campbell SL, Le T, Li E, Sweatt JD, Silva AJ, Fan G
(2010) Dnmt1 and Dnmt3a maintain DNA methylation and
regulate synaptic function in adult forebrain neurons. Nat
Neurosci 13(4):423–430.
Fenoglio C, Ridolfi E, Galimberti D, Scarpini E (2013) An emerging
role for long non-coding RNA dysregulation in neurological
disorders. Int J Mol Sci 14(10):20427–20442.
Ferland CL, Schrader LA (2011) Regulation of histone acetylation in
the hippocampus of chronically stressed rats: a potential role of
sirtuins. Neuroscience 174:104–114.
Ferland CL, Hawley WR, Puckett RE, Wineberg K, Lubin FD,
Dohanich GP, Schrader LA (2013) Sirtuin activity in dentate
gyrus contributes to chronic stress-induced behavior and
extracellular signal-regulated protein kinases 1 and 2 cascade
changes in the hippocampus. Biol Psychiatry 74(12):927–935.
Fischer A, Sananbenesi F, Wang X, Dobbin M, Tsai LH (2007)
Recovery of learning and memory is associated with chromatin
remodelling. Nature 447(7141):178–182.
Francis D, Meaney M (1999) Maternal care and the development of
stress responses. Curr Opin Neurobiol 9(1):128–134.
Francis D, Diorio J, Liu D, Meaney M (1999) Nongenomic
transmission across generations of maternal behavior and
stress responses in the rat. Science 286(5442):1155–1158.
Franklin TB, Russig H, Weiss IC, Graff J, Linder N, Michalon A, Vizi
S, Mansuy IM (2010) Epigenetic transmission of the impact of
early stress across generations. Biol Psychiatry 68(5):408–415.
Fuchikami M, Morinobu S, Segawa M, Okamoto Y, Yamawaki S,
Ozaki N, Inoue T, Kusumi I, Koyama T, Tsuchiyama K, Terao T
(2011) DNA methylation profiles of the brain-derived neurotrophic
factor (BDNF) gene as a potent diagnostic biomarker in major
depression. PLoS One 6(8):e23881.
Fujioka T, Sakata Y, Yamaguchi K, Shibasaki T, Kato H, Nakamura S
(1999) The effects of prenatal stress on the development of
hypothalamic paraventricular neurons in fetal rats. Neuroscience
92(3):1079–1088.
Fujioka T, Fujioka A, Tan N, Chowdhury GM, Mouri H, Sakata Y,
Nakamura S (2001) Mild prenatal stress enhances learning
performance in the non-adopted rat offspring. Neuroscience
103(2):301–307.
Gage FH, Muotri AR (2012) What makes each brain unique. Sci Am
306(3):26–31.
Gagliano H, Delgado-Morales R, Sanz-Garcia A, Armario A (2013)
High doses of the histone deacetylase inhibitor sodium butyrate
trigger a stress-like response. Neuropharmacology 79C:75–82.
Gapp K, Jawaid A, Sarkies P, Bohacek J, Pelczar P, Prados J,
Farinelli L, Miska E, Mansuy IM (2014) Implication of sperm RNAs
in transgenerational inheritance of the effects of early trauma in
mice. Nat Neurosci.
Gelfman S, Cohen N, Yearim A, Ast G (2013) DNA-methylation effect
on cotranscriptional splicing is dependent on GC architecture of
the exon-intron structure. Genome Res 23(5):789–799.
Gray JD, Rubin TG, Hunter RG, McEwen BS (2013a) Hippocampal
gene expression changes underlying stress sensitization and
recovery. Mol Psychiatry.
Gray JD, Milner TA, McEwen BS (2013b) Dynamic plasticity: the role
of glucocorticoids, brain-derived neurotrophic factor and other
trophic factors. Neuroscience 239:214–227.
Guan JS, Haggarty SJ, Giacometti E, Dannenberg JH, Joseph N,
Gao J, Nieland TJ, Zhou Y, Wang X, Mazitschek R, Bradner JE,
DePinho RA, Jaenisch R, Tsai LH (2009) HDAC2 negatively
regulates memory formation and synaptic plasticity. Nature
459(7243):55–60.
Guerrero-Bosagna C, Settles M, Lucker B, Skinner MK (2010)
Epigenetic transgenerational actions of vinclozolin on promoter
regions of the sperm epigenome. PLoS One 5(9).
Guo JU, Su Y, Zhong C, Ming GL, Song H (2011) Hydroxylation of 5methylcytosine by TET1 promotes active DNA demethylation in
the adult brain. Cell 145(3):423–434.
Guo JU, Su Y, Shin JH, Shin J, Li H, Xie B, Zhong C, Hu S, Le T, Fan
G, Zhu H, Chang Q, Gao Y, Ming GL, Song H (2014) Distribution,
recognition and regulation of non-CpG methylation in the adult
mammalian brain. Nat Neurosci 17(2):215–222.
Gutierrez-Mecinas M, Trollope AF, Collins A, Morfett H, Hesketh SA,
Kersante F, Reul JM (2011) Long-lasting behavioral responses to
stress involve a direct interaction of glucocorticoid receptors with
ERK1/2-MSK1-Elk-1 signaling. Proc Natl Acad Sci USA
108(33):13806–13811.
Haramati S, Navon I, Issler O, Ezra-Nevo G, Gil S, Zwang R,
Hornstein E, Chen A (2011) MicroRNA as repressors of stressinduced anxiety: the case of amygdalar miR-34. J Neurosci
31(40):14191–14203.
Harony-Nicolas H, Mamrut S, Brodsky L, Shahar-Gold H, BarkiHarrington L, Wagner S (2014) Brain region-specific methylation in
the promoter of the murine oxytocin receptor gene is involved in its
expression regulation. Psychoneuroendocrinology 39:121–131.
Harris AP, Holmes MC, de Kloet ER, Chapman KE, Seckl JR (2013)
Mineralocorticoid and glucocorticoid receptor balance in control of
HPA
axis
and
behaviour.
Psychoneuroendocrinology
38(5):648–658.
Heinrich SU, Lindquist S (2011) Protein-only mechanism induces
self-perpetuating changes in the activity of neuronal Aplysia
cytoplasmic polyadenylation element binding protein (CPEB).
Proc Natl Acad Sci USA 108(7):2999–3004.
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
1287
1288
1289
1290
1291
1292
1293
1294
1295
1296
1297
1298
1299
1300
1301
1302
1303
1304
1305
1306
1307
1308
1309
1310
1311
1312
1313
1314
1315
1316
1317
1318
1319
1320
1321
1322
1323
1324
1325
1326
1327
1328
1329
1330
1331
1332
1333
1334
1335
1336
1337
1338
1339
1340
1341
1342
1343
1344
1345
1346
1347
1348
1349
1350
1351
1352
1353
1354
1355
1356
NSC 15510
No. of Pages 16
2 July 2014
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
1357
1358
1359
1360
1361
1362
1363
1364
1365
1366
1367
1368
1369
1370
1371
1372
1373
1374
1375
1376
1377
1378
1379
1380
1381
1382
1383
1384
1385
1386
1387
1388
1389
1390
1391
1392
1393
1394
1395
1396
1397
1398
1399
1400
1401
1402
1403
1404
1405
1406
1407
1408
1409
1410
1411
1412
1413
1414
1415
1416
1417
1418
1419
1420
1421
1422
1423
1424
1425
1426
1427
Horike S, Cai S, Miyano M, Cheng JF, Kohwi-Shigematsu T (2005)
Loss of silent-chromatin looping and impaired imprinting of DLX5
in Rett syndrome. Nat Genet 37(1):31–40.
Howerton CL, Bale TL (2012) Prenatal programing: at the intersection
of maternal stress and immune activation. Horm Behav
62(3):237–242.
Hunter RG (2012) Epigenetic effects of stress and corticosteroids in
the brain. Front Cell Neurosci 6:18.
Hunter RG, McEwen BS (2013) Stress and anxiety across the
lifespan: structural plasticity and epigenetic regulation.
Epigenomics 5(2):177–194.
Hunter RG, McCarthy KJ, Milne TA, Pfaff DW, McEwen BS (2009)
Regulation of hippocampal H3 histone methylation by acute and
chronic stress. Proc Natl Acad Sci USA 106(49):20912–20917.
Hunter RG, Murakami G, Dewell S, Seligsohn M, Baker ME, Datson
NA, McEwen BS, Pfaff DW (2012) Acute stress and hippocampal
histone H3 lysine 9 trimethylation, a retrotransposon silencing
response. Proc Natl Acad Sci USA 109(43):17657–17662.
Hunter RG, McEwen BS, Pfaff DW (2013) Environmental stress and
transposon transcription in the mammalian brain. Mob Genet
Elements 3(2):e24555.
Jensen Pena C, Monk C, Champagne FA (2012) Epigenetic effects of
prenatal stress on 11beta-hydroxysteroid dehydrogenase-2 in the
placenta and fetal brain. PLoS One 7(6):e39791.
Jiang Y, Langley B, Lubin FD, Renthal W, Wood MA, Yasui DH,
Kumar A, Nestler EJ, Akbarian S, Beckel-Mitchener AC (2008)
Epigenetics
in
the
nervous
system.
J
Neurosci
28(46):11753–11759.
Jones PA (2012) Functions of DNA methylation: islands, start sites,
gene bodies and beyond. Nat Rev Genet 13(7):484–492.
Kadauke S, Blobel GA (2009) Chromatin loops in gene regulation.
Biochim Biophys Acta 1789(1):17–25.
Kandel ER, Derkatch I, Pavlopoulos E (2013) The role of functional
prions in the persistence of memory storage. Proteopathic seeds
and neurodegenerative diseases. Springer. p. 131–152.
Kang HJ, Kim JM, Lee JY, Kim SY, Bae KY, Kim SW, Shin IS, Kim
HR, Shin MG, Yoon JS (2013) BDNF promoter methylation and
suicidal behavior in depressive patients. J Affect Disord
151(2):679–685.
Keller S, Sarchiapone M, Zarrilli F, Videtic A, Ferraro A, Carli V,
Sacchetti S, Lembo F, Angiolillo A, Jovanovic N, Pisanti F,
Tomaiuolo R, Monticelli A, Balazic J, Roy A, Marusic A, Cocozza
S, Fusco A, Bruni CB, Castaldo G, Chiariotti L (2010) Increased
BDNF promoter methylation in the Wernicke area of suicide
subjects. Arch Gen Psychiatry 67(3):258–267.
Kember RL, Dempster EL, Lee TH, Schalkwyk LC, Mill J, Fernandes
C (2012) Maternal separation is associated with strain-specific
responses to stress and epigenetic alterations to Nr3c1, Avp, and
Nr4a1 in mouse. Brain Behav 2(4):455–467.
Klengel T, Mehta D, Anacker C, Rex-Haffner M, Pruessner JC,
Pariante CM, Pace TW, Mercer KB, Mayberg HS, Bradley B,
Nemeroff CB, Holsboer F, Heim CM, Ressler KJ, Rein T, Binder
EB (2013) Allele-specific FKBP5 DNA demethylation mediates
gene-childhood trauma interactions. Nat Neurosci 16(1):33–41.
Kouzarides T (2007) Chromatin modifications and their function. Cell
128(4):693–705.
Kriaucionis S, Bird A (2003) DNA methylation and Rett syndrome.
Hum Mol Genet 12(Spec No 2):R221–R227.
Kriaucionis S, Heintz N (2009) The nuclear DNA base 5hydroxymethylcytosine is present in Purkinje neurons and the
brain. Science 324(5929):929–930.
Kuriyama K, Honma M, Yoshiike T, Kim Y (2012) Valproic acid but
not d-cycloserine facilitates sleep-dependent offline learning of
extinction and habituation of conditioned fear in humans.
Neuropharmacology.
Labonte B, Suderman M, Maussion G, Navaro L, Yerko V, Mahar I,
Bureau A, Mechawar N, Szyf M, Meaney MJ, Turecki G (2012a)
Genome-wide epigenetic regulation by early-life trauma. Arch
Gen Psychiatry 69(7):722–731.
Labonte B, Yerko V, Gross J, Mechawar N, Meaney MJ, Szyf M,
Turecki G (2012b) Differential glucocorticoid receptor exon 1(b),
13
1(c), and 1(h) expression and methylation in suicide completers
with a history of childhood abuse. Biol Psychiatry 72(1):41–48.
Lachner M, O’Carroll D, Rea S, Mechtler K, Jenuwein T (2001)
Methylation of histone H3 lysine 9 creates a binding site for HP1
proteins. Nature 410(6824):116–120.
Landry CD, Kandel ER, Rajasethupathy P (2013) New mechanisms
in memory storage: piRNAs and epigenetics. Trends Neurosci
36(9):535–542.
LaPlant Q, Vialou V, Covington 3rd HE, Dumitriu D, Feng J, Warren
BL, Maze I, Dietz DM, Watts EL, Iniguez SD, Koo JW, Mouzon E,
Renthal W, Hollis F, Wang H, Noonan MA, Ren Y, Eisch AJ,
Bolanos CA, Kabbaj M, Xiao G, Neve RL, Hurd YL, Oosting RS,
Fan G, Morrison JH, Nestler EJ (2010) Dnmt3a regulates
emotional behavior and spine plasticity in the nucleus
accumbens. Nat Neurosci 13(9):1137–1143.
Larkin H, Felitti VJ, Anda RF (2014) Social work and adverse
childhood experiences research: implications for practice and
health policy. Soc Work Public Health 29(1):1–16.
Lemaire V, Koehl M, Le Moal M, Abrous DN (2000) Prenatal stress
produces learning deficits associated with an inhibition of
neurogenesis in the hippocampus. Proc Natl Acad Sci USA
97(20):11032–11037.
Lemaire V, Lamarque S, Le Moal M, Piazza PV, Abrous DN (2006)
Postnatal stimulation of the pups counteracts prenatal stressinduced deficits in hippocampal neurogenesis. Biol Psychiatry
59(9):786–792.
Liu D, Diorio J, Tannenbaum B, Caldji C, Francis D, Freedman A,
Sharma S, Pearson D, Plotsky PM, Meaney MJ (1997) Maternal
care, hippocampal glucocorticoid receptors, and hypothalamicpituitary-adrenal
responses
to
stress.
Science
277(5332):1659–1662.
Liu D, Caldji C, Sharma S, Plotsky PM, Meaney MJ (2000) Influence
of
neonatal
rearing
conditions
on
stress-induced
adrenocorticotropin responses and norepinepherine release in
the hypothalamic paraventricular nucleus. J Neuroendocrinol
12(1):5–12.
Liu WM, Pang RT, Chiu PC, Wong BP, Lao K, Lee KF, Yeung WS
(2012) Sperm-borne microRNA-34c is required for the first
cleavage division in mouse. Proc Natl Acad Sci USA
109(2):490–494.
Lubin FD, Roth TL, Sweatt JD (2008) Epigenetic regulation of BDNF
gene transcription in the consolidation of fear memory. J Neurosci
28(42):10576–10586.
Maddox SA, Watts CS, Schafe GE (2013) DNA methyltransferase
activity is required for memory-related neural plasticity in the
lateral amygdala. Neurobiol Learn Mem 107C:93–100.
Maddox SA, Watts CS, Schafe GE (2014) DNA methyltransferase
activity is required for memory-related neural plasticity in the
lateral amygdala. Neurobiol Learn Mem 107:93–100.
Maier SF, Watkins LR (2010) Role of the medial prefrontal cortex in
coping and resilience. Brain Res 1355:52–60.
Mamrut S, Harony H, Sood R, Shahar-Gold H, Gainer H, Shi YJ,
Barki-Harrington L, Wagner S (2013) DNA methylation of specific
CpG sites in the promoter region regulates the transcription of the
mouse oxytocin receptor. PLoS One 8(2):e56869.
Mannironi C, Camon J, De Vito F, Biundo A, De Stefano ME,
Persiconi I, Bozzoni I, Fragapane P, Mele A, Presutti C (2013)
Acute stress alters amygdala microRNA miR-135a and miR-124
expression: inferences for corticosteroid dependent stress
response. PLoS One 8(9):e73385.
Martinowich K, Hattori D, Wu H, Fouse S, He F, Hu Y, Fan G, Sun YE
(2003) DNA methylation-related chromatin remodeling in activitydependent BDNF gene regulation. Science 302(5646):
890–893.
Matrisciano F, Tueting P, Dalal I, Kadriu B, Grayson DR, Davis JM,
Nicoletti F, Guidotti A (2013) Epigenetic modifications of
GABAergic interneurons are associated with the schizophrenialike phenotype induced by prenatal stress in mice.
Neuropharmacology 68:184–194.
Matsumoto Y, Morinobu S, Yamamoto S, Matsumoto T, Takei S,
Fujita Y, Yamawaki S (2013) Vorinostat ameliorates impaired fear
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
1428
1429
1430
1431
1432
1433
1434
1435
1436
1437
1438
1439
1440
1441
1442
1443
1444
1445
1446
1447
1448
1449
1450
1451
1452
1453
1454
1455
1456
1457
1458
1459
1460
1461
1462
1463
1464
1465
1466
1467
1468
1469
1470
1471
1472
1473
1474
1475
1476
1477
1478
1479
1480
1481
1482
1483
1484
1485
1486
1487
1488
1489
1490
1491
1492
1493
1494
1495
1496
1497
1498
NSC 15510
No. of Pages 16
2 July 2014
14
1499
1500
1501
1502
1503
1504
1505
1506
1507
1508
1509
1510
1511
1512
1513
1514
1515
1516
1517
1518
1519
1520
1521
1522
1523
1524
1525
1526
1527
1528
1529
1530
1531
1532
1533
1534
1535
1536
1537
1538
1539
1540
1541
1542
1543
1544
1545
1546
1547
1548
1549
1550
1551
1552
1553
1554
1555
1556
1557
1558
1559
1560
1561
1562
1563
1564
1565
1566
1567
1568
1569
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
extinction possibly via the hippocampal NMDA-CaMKII pathway
in an animal model of posttraumatic stress disorder.
Psychopharmacology (Berl) 229(1):51–62.
McClintock B (1951) Chromosome organization and genic
expression. Cold Spring Harb Symp Quant Biol 16:13–47.
McClintock B (1984) The significance of responses of the genome to
challenge. Science 226(4676):792–801.
McEwen BS, Wingfield JC (2003) The concept of allostasis in biology
and biomedicine. Horm Behav 43(1):2–15.
McEwen BS, Wingfield JC (2010) What is in a name? Integrating
homeostasis, allostasis and stress. Horm Behav 57(2):105–111.
McEwen BS, Eiland L, Hunter RG, Miller MM (2012) Stress and
anxiety: structural plasticity and epigenetic regulation as a
consequence of stress. Neuropharmacology 62(1):3–12.
McGowan PO (2013) Epigenomic mechanisms of early adversity and
HPA dysfunction: considerations for PTSD research. Front
Psychiatry 4:110.
McGowan PO, Sasaki A, D’Alessio AC, Dymov S, Labonte B, Szyf M,
Turecki G, Meaney MJ (2009) Epigenetic regulation of the
glucocorticoid receptor in human brain associates with
childhood abuse. Nat Neurosci 12(3):342–348.
Meerson A, Cacheaux L, Goosens KA, Sapolsky RM, Soreq H,
Kaufer D (2010) Changes in brain MicroRNAs contribute to
cholinergic stress reactions. J Mol Neurosci 40(1–2):47–55.
Mehta D, Klengel T, Conneely KN, Smith AK, Altmann A, Pace TW,
Rex-Haffner M, Loeschner A, Gonik M, Mercer KB, Bradley B,
Muller-Myhsok B, Ressler KJ, Binder EB (2013) Childhood
maltreatment is associated with distinct genomic and epigenetic
profiles in posttraumatic stress disorder. Proc Natl Acad Sci USA
110(20):8302–8307.
Miller CA, Sweatt JD (2007) Covalent modification of DNA regulates
memory formation. Neuron 53(6):857–869.
Miller AH, Maletic V, Raison CL (2009) Inflammation and its
discontents: the role of cytokines in the pathophysiology of
major depression. Biol Psychiatry 65(9):732–741.
Mizuno K, Dempster E, Mill J, Giese KP (2012) Long-lasting
regulation of hippocampal Bdnf gene transcription after
contextual fear conditioning. Genes Brain Behav 11(6):651–659.
Mongrain V, Hernandez SA, Pradervand S, Dorsaz S, Curie T,
Hagiwara G, Gip P, Heller HC, Franken P (2010) Separating the
contribution of glucocorticoids and wakefulness to the molecular
and electrophysiological correlates of sleep homeostasis. Sleep
33(9):1147–1157.
Monsey MS, Ota KT, Akingbade IF, Hong ES, Schafe GE (2011)
Epigenetic alterations are critical for fear memory consolidation
and synaptic plasticity in the lateral amygdala. PLoS One
6(5):e19958.
Monteleone MC, Adrover E, Pallares ME, Antonelli MC, Frasch AC,
Brocco MA (2013) Prenatal stress changes the glycoprotein
GPM6A gene expression and induces epigenetic changes in rat
offspring brain. Epigenetics 9(1).
Morgan CP, Bale TL (2011) Early prenatal stress epigenetically
programs dysmasculinization in second-generation offspring via
the paternal lineage. J Neurosci 31(33):11748–11755.
Mueller BR, Bale TL (2008) Sex-specific programming of offspring
emotionality after stress early in pregnancy. J Neurosci
28(36):9055–9065.
Murgatroyd C, Patchev AV, Wu Y, Micale V, Bockmuhl Y, Fischer D,
Holsboer F, Wotjak CT, Almeida OF, Spengler D (2009) Dynamic
DNA methylation programs persistent adverse effects of early-life
stress. Nat Neurosci 12(12):1559–1566.
Nestler EJ (2014) Epigenetic mechanisms of drug addiction.
Neuropharmacology 76(Pt B):259–268.
Nestler EJ, Hyman SE (2010) Animal models of neuropsychiatric
disorders. Nat Neurosci 13(10):1161–1169.
Niehrs C, Schafer A (2012) Active DNA demethylation by Gadd45
and DNA repair. Trends Cell Biol 22(4):220–227.
Niwa M, Jaaro-Peled H, Tankou S, Seshadri S, Hikida T, Matsumoto
Y, Cascella NG, Kano S, Ozaki N, Nabeshima T, Sawa A (2013)
Adolescent stress-induced epigenetic control of dopaminergic
neurons via glucocorticoids. Science 339(6117):335–339.
Norrholm SD, Jovanovic T, Smith AK, Binder E, Klengel T, Conneely
K, Mercer KB, Davis JS, Kerley K, Winkler J, Gillespie CF,
Bradley B, Ressler KJ (2013) Differential Genetic and Epigenetic
regulation of catechol-O-methyltransferase is associated with
impaired fear inhibition in posttraumatic stress disorder. Front
Behav Neurosci 7:30.
Nowak SJ, Corces VG (2004) Phosphorylation of histone H3: a
balancing act between chromosome condensation and
transcriptional activation. Trends Genet 20(4):214–220.
Oberlander TF, Weinberg J, Papsdorf M, Grunau R, Misri S, Devlin
AM (2008) Prenatal exposure to maternal depression, neonatal
methylation of human glucocorticoid receptor gene (NR3C1) and
infant cortisol stress responses. Epigenetics 3(2):97–106.
O’Connor RM, Grenham S, Dinan TG, Cryan JF (2013). microRNAs
as novel antidepressant targets: converging effects of ketamine
and electroconvulsive shock therapy in the rat hippocampus. Int J
Neuropsychopharmacol: 1–8.
Olff M, Frijling JL, Kubzansky LD, Bradley B, Ellenbogen MA,
Cardoso C, Bartz JA, Yee JR, van Zuiden M (2013) The role of
oxytocin in social bonding, stress regulation and mental health:
an update on the moderating effects of context and
interindividual
differences.
Psychoneuroendocrinology
38(9):1883–1894.
Painter RC, Osmond C, Gluckman P, Hanson M, Phillips DI,
Roseboom TJ (2008) Transgenerational effects of prenatal
exposure to the Dutch famine on neonatal adiposity and health
in later life. BJOG 115(10):1243–1249.
Perroud N, Dayer A, Piguet C, Nallet A, Favre S, Malafosse A, Aubry
JM (2014) Childhood maltreatment and methylation of the
glucocorticoid receptor gene NR3C1 in bipolar disorder. Br J
Psychiatry 204:30–35.
Pezze MA, Feldon J (2004) Mesolimbic dopaminergic pathways in
fear conditioning. Prog Neurobiol 74(5):301–320.
Ponomarev I, Rau V, Eger EI, Harris RA, Fanselow MS (2010)
Amygdala transcriptome and cellular mechanisms underlying
stress-enhanced fear learning in a rat model of posttraumatic
stress disorder. Neuropsychopharmacology 35(6):1402–1411.
Radley JJ, Kabbaj M, Jacobson L, Heydendael W, Yehuda R,
Herman JP (2011) Stress risk factors and stress-related
pathology: neuroplasticity, epigenetics and endophenotypes.
Stress 14(5):481–497.
Reilly MT, Faulkner GJ, Dubnau J, Ponomarev I, Gage FH (2013)
The role of transposable elements in health and diseases of the
central nervous system. J Neurosci 33(45):17577–17586.
Renthal W, Maze I, Krishnan V, Covington 3rd HE, Xiao G, Kumar A,
Russo SJ, Graham A, Tsankova N, Kippin TE, Kerstetter KA,
Neve RL, Haggarty SJ, McKinsey TA, Bassel-Duby R, Olson EN,
Nestler EJ (2007) Histone deacetylase 5 epigenetically controls
behavioral adaptations to chronic emotional stimuli. Neuron
56(3):517–529.
Reul JM (2014) Making memories of stressful events: a journey along
epigenetic, gene transcription, and signaling pathways. Front
Psychiatry 5:5.
Reul JM, Chandramohan Y (2007) Epigenetic mechanisms in stressrelated memory formation. Psychoneuroendocrinology 32(Suppl
1):S21–S25.
Rinaldi A, Vincenti S, De Vito F, Bozzoni I, Oliverio A, Presutti C,
Fragapane P, Mele A (2010) Stress induces region specific
alterations in microRNAs expression in mice. Behav Brain Res
208(1):265–269.
Rodgers AB, Morgan CP, Bronson SL, Revello S, Bale TL (2013)
Paternal stress exposure alters sperm microRNA content and
reprograms offspring HPA stress axis regulation. J Neurosci
33(21):9003–9012.
Roth TL, Lubin FD, Funk AJ, Sweatt JD (2009) Lasting epigenetic
influence of early-life adversity on the BDNF gene. Biol Psychiatry
65(9):760–769.
Roth TL, Zoladz PR, Sweatt JD, Diamond DM (2011) Epigenetic
modification of hippocampal Bdnf DNA in adult rats in an animal
model of post-traumatic stress disorder. J Psychiatr Res
45(7):919–926.
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
1570
1571
1572
1573
1574
1575
1576
1577
1578
1579
1580
1581
1582
1583
1584
1585
1586
1587
1588
1589
1590
1591
1592
1593
1594
1595
1596
1597
1598
1599
1600
1601
1602
1603
1604
1605
1606
1607
1608
1609
1610
1611
1612
1613
1614
1615
1616
1617
1618
1619
1620
1621
1622
1623
1624
1625
1626
1627
1628
1629
1630
1631
1632
1633
1634
1635
1636
1637
1638
1639
1640
NSC 15510
No. of Pages 16
2 July 2014
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
1641
1642
1643
1644
1645
1646
1647
1648
1649
1650
1651
1652
1653
1654
1655
1656
1657
1658
1659
1660
1661
1662
1663
1664
1665
1666
1667
1668
1669
1670
1671
1672
1673
1674
1675
1676
1677
1678
1679
1680
1681
1682
1683
1684
1685
1686
1687
1688
1689
1690
1691
1692
1693
1694
1695
1696
1697
1698
1699
1700
1701
1702
1703
1704
1705
1706
1707
1708
1709
1710
1711
Rotllant D, Pastor-Ciurana J, Armario A (2013) Stress-induced brain
histone H3 phosphorylation: contribution of the intensity of
stressors and length of exposure. J Neurochem 125(4):599–609.
Rozeboom AM, Akil H, Seasholtz AF (2007) Mineralocorticoid
receptor overexpression in forebrain decreases anxiety-like
behavior and alters the stress response in mice. Proc Natl Acad
Sci U S A 104(11):4688–4693.
Rusiecki JA, Chen L, Srikantan V, Zhang L, Yan L, Polin ML,
Baccarelli A (2012) DNA methylation in repetitive elements and
post-traumatic stress disorder: a case–control study of US military
service members. Epigenomics 4(1):29–40.
Russo SJ, Murrough JW, Han MH, Charney DS, Nestler EJ (2012)
Neurobiology of resilience. Nat Neurosci 15(11):1475–1484.
Ruthenburg AJ, Li H, Patel DJ, Allis CD (2007) Multivalent
engagement of chromatin modifications by linked binding
modules. Nat Rev Mol Cell Biol 8(12):983–994.
Safe S, Kim K (2008) Non-classical genomic estrogen receptor (ER)/
specificity protein and ER/activating protein-1 signaling pathways.
J Mol Endocrinol 41(5):263–275.
Sawchenko PE, Swanson LW (1981) Central noradrenergic
pathways for the integration of hypothalamic neuroendocrine
and autonomic responses. Science 214(4521):685–687.
Shaltiel G, Hanan M, Wolf Y, Barbash S, Kovalev E, Shoham S,
Soreq H (2013) Hippocampal microRNA-132 mediates stressinducible cognitive deficits through its acetylcholinesterase target.
Brain Struct Funct 218(1):59–72.
Sharif J, Shinkai Y, Koseki H (2013) Is there a role for endogenous
retroviruses to mediate long-term adaptive phenotypic response
upon environmental inputs? Philos Trans R Soc Lond B Biol Sci
368(1609):20110340.
Shi Y, Lan F, Matson C, Mulligan P, Whetstine JR, Cole PA, Casero
RA (2004) Histone demethylation mediated by the nuclear amine
oxidase homolog LSD1. Cell 119(7):941–953.
Si J, Yang J, Xue L, Yang C, Luo Y, Shi H, Lu L (2012) Activation of
NF-jB in basolateral amygdala is required for memory
reconsolidation in auditory fear conditioning. PLoS One
7(9):e43973.
Simmons RK, Howard JL, Simpson DN, Akil H, Clinton SM (2012)
DNA methylation in the developing hippocampus and amygdala
of anxiety-prone versus risk-taking rats. Dev Neurosci
34(1):58–67.
Simmons RK, Stringfellow SA, Glover ME, Wagle AA, Clinton SM
(2013) DNA methylation markers in the postnatal developing rat
brain. Brain Res 1533:26–36.
Skinner MK, Anway MD, Savenkova MI, Gore AC, Crews D (2008)
Transgenerational epigenetic programming of the brain
transcriptome and anxiety behavior. PLoS One 3(11):e3745.
Sõber S, Laan M, Annilo T (2010) MicroRNAs miR-124 and miR-135a
are potential regulators of the mineralocorticoid receptor gene
(NR3C2) expression. Biochem Biophys Res Commun
391(1):727–732.
Stilling RM, Dinan TG, Cryan JF (2014) Microbial genes, brain &
behaviour–epigenetic regulation of the gut–brain axis. Genes
Brain Behav 13(1):69–86.
Strahl BD, Ohba R, Cook RG, Allis CD (1999) Methylation of histone
H3 at lysine 4 is highly conserved and correlates with
transcriptionally active nuclei in tetrahymena. Proc Natl Acad
Sci 96(26):14967–14972.
Suemaru S, Darlington DN, Akana SF, Cascio CS, Dallman MF
(1995) Ventromedial hypothalamic lesions inhibit corticosteroid
feedback regulation of basal ACTH during the trough of the
circadian rhythm. Neuroendocrinology 61(4):453–463.
Sun G, Alzayady K, Stewart R, Ye P, Yang S, Li W, Shi Y (2010)
Histone demethylase LSD1 regulates neural stem cell
proliferation. Mol Cell Biol 30(8):1997–2005.
Sun A, Crabtree G, Yoo A (2013) MicroRNAs: regulators of neuronal
fate. Curr Opin Cell Biol 25(2):215–221.
Sweatt JD (2013) The emerging field of neuroepigenetics. Neuron
80(3):624–632.
Tan M, Luo H, Lee S, Jin F, Yang JS, Montellier E, Buchou T, Cheng
Z, Rousseaux S, Rajagopal N (2011) Identification of 67 histone
15
marks and histone lysine crotonylation as a new type of histone
modification. Cell 146(6):1016–1028.
Tang YP, Shimizu E, Dube GR, Rampon C, Kerchner GA, Zhuo M,
Liu G, Tsien JZ (1999) Genetic enhancement of learning and
memory in mice. Nature 401(6748):63–69.
Thomassin H, Flavin M, Espinas ML, Grange T (2001)
Glucocorticoid-induced DNA demethylation and gene memory
during development. EMBO J 20(8):1974–1983.
Trollope AF, Gutierrez-Mecinas M, Mifsud KR, Collins A, Saunderson
EA, Reul JM (2012) Stress, epigenetic control of gene expression
and memory formation. Exp Neurol 233(1):3–11.
Tsankova NM, Kumar A, Nestler EJ (2004) Histone modifications at
gene promoter regions in rat hippocampus after acute and chronic
electroconvulsive seizures. J Neurosci 24(24):5603–5610.
Tsankova NM, Berton O, Renthal W, Kumar A, Neve RL, Nestler EJ
(2006) Sustained hippocampal chromatin regulation in a mouse
model of depression and antidepressant action. Nat Neurosci
9(4):519–525.
Turner JD, Muller CP (2005) Structure of the glucocorticoid receptor
(NR3C1) gene 5’ untranslated region: identification, and tissue
distribution of multiple new human exon 1. J Mol Endocrinol
35(2):283–292.
Turner JD, Alt SR, Cao L, Vernocchi S, Trifonova S, Battello N, Muller
CP (2010) Transcriptional control of the glucocorticoid receptor:
CpG islands, epigenetics and more. Biochem Pharmacol
80(12):1860–1868.
Tyrka AR, Price LH, Marsit C, Walters OC, Carpenter LL (2012)
Childhood adversity and epigenetic modulation of the leukocyte
glucocorticoid receptor: preliminary findings in healthy adults.
PLoS One 7(1):e30148.
Uchida S, Nishida A, Hara K, Kamemoto T, Suetsugi M, Fujimoto M,
Watanuki T, Wakabayashi Y, Otsuki K, McEwen BS, Watanabe Y
(2008) Characterization of the vulnerability to repeated stress in
Fischer 344 rats: possible involvement of microRNA-mediated
down-regulation of the glucocorticoid receptor. Eur J Neurosci
27(9):2250–2261.
Uddin M, Galea S, Chang SC, Koenen KC, Goldmann E, Wildman
DE, Aiello AE (2013) Epigenetic signatures may explain the
relationship between socioeconomic position and risk of mental
illness: preliminary findings from an urban community-based
sample. Biodemogr Soc Biol 59(1):68–84.
Unternaehrer E, Luers P, Mill J, Dempster E, Meyer AH, Staehli S,
Lieb R, Hellhammer DH, Meinlschmidt G (2012) Dynamic
changes in DNA methylation of stress-associated genes (OXTR,
BDNF) after acute psychosocial stress. Transl Psychiatry 2:e150.
Vreugdenhil E, Verissimo CS, Mariman R, Kamphorst JT, Barbosa
JS, Zweers T, Champagne DL, Schouten T, Meijer OC, de Kloet
ER, Fitzsimons CP (2009) MicroRNA 18 and 124a down-regulate
the glucocorticoid receptor: implications for glucocorticoid
responsiveness in the brain. Endocrinology 150(5):2220–2228.
Waddington CH (1942) The epigenotype. Endeavour 1:18–20.
Wang T, Pan Q, Lin L, Szulwach KE, Song CX, He C, Wu H, Warren ST,
Jin P, Duan R, Li X (2012) Genome-wide DNA hydroxymethylation
changes are associated with neurodevelopmental genes in the
developing human cerebellum. Hum Mol Genet 21(26):5500–5510.
Weaver IC, Cervoni N, Champagne FA, D’Alessio AC, Sharma S,
Seckl JR, Dymov S, Szyf M, Meaney MJ (2004) Epigenetic
programming by maternal behavior. Nat Neurosci 7(8):847–854.
Weaver IC, D’Alessio AC, Brown SE, Hellstrom IC, Dymov S, Sharma
S, Szyf M, Meaney MJ (2007) The transcription factor nerve
growth factor-inducible protein a mediates epigenetic
programming: altering epigenetic marks by immediate-early
genes. J Neurosci 27(7):1756–1768.
Wilkinson MB, Xiao G, Kumar A, LaPlant Q, Renthal W, Sikder D,
Kodadek TJ, Nestler EJ (2009) Imipramine treatment and
resiliency exhibit similar chromatin regulation in the mouse
nucleus accumbens in depression models. J Neurosci
29(24):7820–7832.
Wu H, Zhang Y (2011) Mechanisms and functions of Tet proteinmediated
5-methylcytosine
oxidation.
Genes
Dev
25(23):2436–2452.
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
1712
1713
1714
1715
1716
1717
1718
1719
1720
1721
1722
1723
1724
1725
1726
1727
1728
1729
1730
1731
1732
1733
1734
1735
1736
1737
1738
1739
1740
1741
1742
1743
1744
1745
1746
1747
1748
1749
1750
1751
1752
1753
1754
1755
1756
1757
1758
1759
1760
1761
1762
1763
1764
1765
1766
1767
1768
1769
1770
1771
1772
1773
1774
1775
1776
1777
1778
1779
1780
1781
1782
NSC 15510
No. of Pages 16
2 July 2014
16
1783
1784
1785
1786
1787
1788
1789
1790
1791
1792
1793
1794
1795
1796
1811
1812
1813
B. B. Griffiths, R. G. Hunter / Neuroscience xxx (2014) xxx–xxx
Yang XJ, Seto E (2007) HATs and HDACs: from structure, function
and regulation to novel strategies for therapy and prevention.
Oncogene 26(37):5310–5318.
Yehuda R, McFarlane AC (1995) Conflict between current knowledge
about posttraumatic stress disorder and its original conceptual
basis. Am J Psychiatry 152(12):1705–1713.
Yehuda R, Daskalakis NP, Desarnaud F, Makotkine I, Lehrner AL,
Koch E, Flory JD, Buxbaum JD, Meaney MJ, Bierer LM (2013)
Epigenetic biomarkers as predictors and correlates of symptom
improvement following psychotherapy in combat veterans with
PTSD. Front Psychiatry 4.
Yehuda R, Daskalakis NP, Lehrner A, Desarnaud F, Bader HN,
Makotkine I, Flory JD, Bierer LM, Meaney MJ (2014) Influences of
maternal and paternal PTSD on epigenetic regulation of the
glucocorticoid receptor gene in holocaust survivor offspring. Am J
Psychiatry.
Zhang YZ, Zhang QH, Ye H, Zhang Y, Luo YM, Ji XM, Su YY (2010)
Distribution of lysine-specific demethylase 1 in the brain of rat and
its response in transient global cerebral ischemia. Neurosci Res
68(1):66–72.
Zimmermann N, Zschocke J, Perisic T, Yu S, Holsboer F, Rein T
(2012) Antidepressants inhibit DNA methyltransferase 1 through
reducing G9a levels. Biochem J 448(1):93–102.
Zingg HH, Laporte SA (2003) The oxytocin receptor. Trends
Endocrinol Metab 14(5):222–227.
Zovkic IB, Sweatt JD (2013) Epigenetic mechanisms in learned
fear: implications for PTSD. Neuropsychopharmacology 38(1):
77–93.
(Accepted 16 June 2014)
(Available online xxxx)
Please cite this article in press as: Griffiths BB, Hunter RG. Neuroepigenetics of stress. Neuroscience (2014), http://dx.doi.org/10.1016/
j.neuroscience.2014.06.041
1797
1798
1799
1800
1801
1802
1803
1804
1805
1806
1807
1808
1809
1810