You are on page 1of 12

REVIEWS

STRESS

Stress-related synaptic plasticity


in the hypothalamus
Jaideep S. Bains1, Jaclyn I. Wamsteeker Cusulin1,2 and Wataru Inoue1,3
Abstract | Stress necessitates an immediate engagement of multiple neural and endocrine
systems. However, exposure to a single stressor causes adaptive changes that modify
responses to subsequent stressors. Recent studies examining synapses onto
neuroendocrine cells in the paraventricular nucleus of the hypothalamus demonstrate that
stressful experiences leave indelible marks that alter the ability of these synapses to
undergo plasticity. These adaptations include a unique form of metaplasticity at
glutamatergic synapses, bidirectional changes in endocannabinoid signalling and
bidirectional changes in strength at GABAergic synapses that rely on distinct temporal
windows following stress. This rich repertoire of plasticity is likely to represent an
important building block for dynamic, experience-dependent modulation of
neuroendocrine stress adaptation.

A stressor is a real or imagined threat that is inter­ integrate stress-relevant signals from multiple brain
preted by an organism as requiring immediate, adaptive regions and launch the neuroendocrine response to
responses. In animals, these include behavioural (fear stress 17–20. CRH-producing neurons are embedded
and aggression) and visceral (autonomic and neuro­ within a population of parvocellular neuroendocrine
endocrine) actions (FIG. 1). Stress has also been implicated cells (PNCs) in the PVN that release their contents into
in the emergence of numerous pathologies, including the hypophyseal portal circulation. CRH, along with
depression and anxiety disorders, and as a behavioural other co‑released secretagogues, stimulates the release
modifier that affects fundamental processes, such as of adrenocorticotropic hormone (ACTH) from the
1
Hotchkiss Brain Institute and decision making 1–4. In addition to these direct effects of pituitary corticotropes, which in turn drives the release
the Department of Physiology
and Pharmacology, Cumming
stress on other systems, it is clear that exposure to one of corticosteroids (such as corticosterone (CORT) in
School of Medicine, University stressor often results in altered responses to subsequent rats and mice) from the adrenal cortex (FIG. 1). The activ­
of Calgary, 3330 Hospital stressors1–3,5–12. This implies that the neural circuits that ity of PNCs is controlled by synaptic inputs that convey
Drive NW, Calgary, Alberta launch the response to stress are capable of learning and both interoceptive and exteroceptive stressor-related
T2N 4N1, Canada.
remembering. More precisely, these circuits can extract information (FIG. 1). Consequently, these synapses are
2
Present address: Roche
Pharmaceutical Research and information about an individual stress event, store ideally positioned to not only integrate different stress
Early Development, Roche this information and then use it to modify subsequent signals but also extract information from one stressful
Innovation Center Basel, responses. What are the mechanisms involved in stor­ event and use it to modify the responsiveness of the
F. Hoffmann‑La Roche, ing this information? Where is this information stored? hypothalamic–pituitary–adrenal (HPA) axis.
Grenzacherstrasse 124, 4070
Basel, Switzerland.
How does the brain make causal links between stressful This Review focuses largely on recent studies that
3
Present address: Robarts experiences and embed these links within the neurons have examined changes at glutamatergic and GABAergic
Research Institute, and synapses that drive stress responses? synapses onto PNCs. These studies have revealed that
Department of Physiology In an attempt to address these questions, investiga­ several unique forms of short- and long-term plasticity
and Pharmacology, Schulich
tors have examined synapses and plasticity in several occur at these synapses. To place these findings within
School of Medicine and
Dentistry, University of key brain structures that gate stress responsiveness, the appropriate circuit framework, we first briefly
Western Ontario, 1151 including the hippocampus, prefrontal cortex and describe the source of glutamatergic and GABAergic
Richmond Street North, amygdala13–16. More recently, focus has turned towards synaptic inputs to the PVN (FIG. 2) (for a more extensive
London, Ontario N6A 5B7, understanding synaptic function and plasticity in the review of the anatomical inputs to the PVN, see REF. 21)
Canada.
Correspondence to J.S.B. 
paraventricular nucleus of the hypothalamus (PVN). and then we describe studies that have provided the
e‑mail: jsbains@ucalgary.ca Neurons in the PVN, and more precisely those that first set of rules that define the underlying principles
doi:10.1038/nrn3881 synthesize corticotropin-releasing hormone (CRH), of transmission at synapses on PNCs. Building on this

NATURE REVIEWS | NEUROSCIENCE VOLUME 16 | JULY 2015 | 377

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

a Figure 1 | Stress and functional connectivity. 


Psychological stress Physiological stress or a | Psychological and physiological threats to survival
homeostatic challenge
are relayed via specific brain structures to the major
stress-integrative brain centre, located in the
PFC, amygdala paraventricular nucleus of the hypothalamus (PVN).
Brainstem
and hippocampus
Corticotropin-releasing hormone (CRH)-producing
neurons in the PVN serve as primary coordinators of the
BNST endocrine stress response, which culminates with the
release of corticosteroids (‘glucocorticoids’ in the
figure) from the adrenal cortex. Corticosteroids have
Hypothalamus
multiple effects on peripheral systems but also feed
back to help quench the stress response through
CRH actions in the hippocampus (and other brain regions)
and modify subsequent responses to stress through
Anterior actions in the PVN. The square denotes that
pituitary glucocorticoids are responsible for negative feedback
and can promote more-complex adaptive changes. The
ACTH
dashed arrow denotes that brainstem neurons signal via
the amygdala in response to some homeostatic
Adrenal cortex challenges. b | Cascade of events illustrating
recruitment of the neuroendocrine response to stress.
Glucocorticoids Parvocellular neuroendocrine cells (PNCs; putative
CRH-synthesizing neurons) in the PVN are activated
b CRH Glucocorticoids (plasma) and release CRH from axon endings located at the
ACTH Glucocorticoids (brain) median eminence, part of pituitary portal circulation.
Released Released from Released from CRH stimulates release of adrenocorticotropic
hormone (ACTH) from pituitary corticotrophs in the
Hormone level

from PNCs anterior pituitary adrenal cortex


anterior pituitary, which in turn act on the adrenal
cortex to increase glucocorticoid release into the
general circulation. Glucocorticoids act peripherally
and slowly cross the blood–brain barrier to elicit diverse
CNS effects, including negative feedback on the
Time hypothalamic–pituitary–adrenal (HPA) axis. BNST, bed
Stress nucleus of the stria terminalis; PFC, prefrontal cortex.

foundation, we examine recentNaturefindings


Reviewsthat have dem­
| Neuroscience on fast glutamate or GABA signals36,37 to communicate
onstrated embedded forms of plasticity, or metaplasticity, with PVN neurons. These findings also support the idea
that are only evident after an animal has been subjected that peptidergic synapses can use either glutamate or
to stress. Finally, we discuss findings relating to a type of GABA as a co-transmitter 36. The potential interactions
metaplasticity following stress that incorporates a time between slow-acting neuromodulatory systems and fast
domain, which we call kairoplasticity. ionotropic signals are discussed in detail below.
As in other brain regions, both asymmetrical (glu­
Synapses of the PVN tamatergic) and symmetrical (GABAergic) synapses
CRH-synthesizing PNCs are innervated by glutamater­ are present in the PVN25,29,38,39. Quantitative immuno­
gic and GABAergic fibres that, along with monoamin­ electron microscopy studies indicate that approximately
ergic and peptidergic inputs, are crucial for regulating half of all synapses onto CRH-synthesizing neurons are
HPA-axis output 21. Although reports supported the GABAergic and that the others are non-GABAergic,
idea of glutamate and GABA regulation of the output asymmetrical and presumably glutamatergic29. The
of the PVN22,23, it was the seminal findings by van den proportion of GABAergic synapses in the PVN is sub­
Pol and colleagues that provided the first demonstra­ stantially higher than that in other brain regions, such
tion of fast synaptic transmission mediated by glutamate as the CA1 in the hippocampus (3% GABA) and the
Metaplasticity
A change in synapses that
and GABA in the PVN24–27. At the ultrastructural level, visual cortex (6% GABA)40. Consequently, the PVN
alters their ability to express nearly all presynaptic terminals onto PVN neurons con­ may be ideally suited for studying mechanisms of plas­
plasticity. tain numerous small, clear, round vesicles that closely ticity at GABAergic synapses, which have proven dif­
appose synaptic specializations (which are indicative ficult to investigate in other systems. Approximately
Kairoplasticity
of glutamate or GABA packaging)24,26–28. These small 30% of all synapses onto PVN neurons are also adren­
Derived from the Greek
‘kairos’, meaning the vesicles are often accompanied by dense-core vesicles ergic or noradrenergic. These synapses have an asym­
opportune or correct moment, (which usually contain peptides and amines) 25,26,29. metric morphology 41 but it is unclear whether they,
it refers to observations that Based on numerous electrophysiological studies, it is like noradrenergic synapses in other brain areas, also
different forms of now accepted that virtually all PVN neurons receive release glutamate42,43.
metaplasticity following stress
that can be induced only
fast glutamatergic and GABAergic synaptic inputs30–35. Both glutamatergic and GABAergic synapses
during specific and distinct Moreover, recent optogenetic studies have shown that make axosomatic as well as axodendritic contacts onto
temporal windows. genetically identified peptidergic cell populations rely PNCs25,27,29,44. They are intermingled synaptic contacts

378 | JULY 2015 | VOLUME 16 www.nature.com/reviews/neuro

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

receptor (NMDAR) and metabotropic glutamate recep­


Hc– tor (mGluR)) and GABA receptor (GABAA receptor
PBN PVT
vSub
(GABAAR) and GABABR) subunits have been detected in
the PVN48–52. Moreover, microinjection of glutamate into
the PVN stimulates HPA-axis output, whereas admin­
istration of an ionotropic glutamate receptor antagonist
before exposure to stress blunts the neuroendocrine stress
NTS CeA LC
response53–55. By contrast, GABAergic synaptic transmis­
sion restrains baseline HPA-axis activity. For example,
BLA
microinjections of GABAAR antagonists in the absence
of a stressor robustly activate PVN neurons and the HPA
axis54, suggesting that the release from the tonic inhibitory
VLM BNST PFC
GABA tone (a phenomenon known as disinhibition) is an
important mechanism for PVN activation. Such in vivo
peri-
fx
MeA experiments in rodents are largely consistent with ex vivo
electrophysiological recordings showing that GABAAR
blockade increases PNC firing 56. However, the activation
MnPO
of PNCs is probably regulated by the balance between
opposing drives from glutamate and GABA inputs, as
disinhibition (that is, GABAAR blockade)-induced activa­
ZI AHA tion of these cells is blunted in the presence of a glutamate
Raphe receptor antagonist57.
PVN
ARC DMH Anatomy
Information from the internal and external environment
is processed in distinct brain regions and eventually
Primarily glutamate Both glutamate and GABA reaches CRH-producing neurons in the PVN to initi­
Primarily GABA Both monomine and glutamate
ate HPA-axis activation. There are several excellent and
detailed reviews on the neuroanatomical architecture of
Monoamine Dense GR expression PVN afferents21,58 and so we will only briefly describe
this architecture here (see also FIG. 2).
Figure 2 | Major direct and indirect CNS connections to the PVN. Compiled from The direct, monosynaptic inputs to the PVN primarily
multiple anatomical studies, this topological map showsNature Reviews
key regions | Neuroscience
implicated in the arise from deep brain structures, including hypothalamic,
brain response to stress, specifically the major glutamate, GABA and monoamine extrahypothalamic and brainstem nuclei that are classi­
projections to the paraventricular nucleus of the hypothalamus (PVN). The halos cally implicated in instinctive behaviour and physiological
represent areas of the brain in which high levels of glucocorticoid receptors (GRs) have
functions. By contrast, limbic forebrain structures, such
been described. Particular attention is paid here to limbic regions of the brain implicated
in regulation of hypothalamic–pituitary–adrenal (HPA)-axis responses to psychogenic
as the hippocampus, amygdala and prefrontal cortex,
stressors and glucocorticoid feedback. AHA, anterior hypothalamus; ARC, arcuate which are critical for relaying information about psycho­
nucleus; BLA, basolateral amygdala; BNST, bed nucleus of the stria terminalis; CeA, logical stressors, send few direct projections to the PVN.
central amygdala; DMH, dorsomedial hypothalamus; Hc, hippocampus; LC, locus Instead, they direct monosynaptic projections to many of
coeruleus; MeA, medial amygdala; MnPO, medial preoptic nucleus; NTS, nucleus tractus the aforementioned deep brain structures, which in turn
solitarus; PBN, parabrachial nucleus; PVT, paraventricular thalamus; peri‑fx, peri-fornical relay information to the PVN.
area; PFC, prefrontal cortex; Raphe, Raphe nuclei; VLM, ventrolateral medulla; vSub, The hierarchical organization of the stress neuro­
ventral subiculum; ZI, zona incerta. circuitry led Herman and colleagues to propose that
stressor-related information is integrated in deep
in close proximity to one another 44. Furthermore, PNCs brain structures projecting to the PVN 21,58. These
(as well as other PVN neurons) have a relatively simple include non-overlapping populations of glutamatergic
dendritic architecture, with a smooth shaft and low spine and GABAergic neurons53 in the arcuate nucleus, the
density 45. As a point of comparison, the spine density dorsomedial hypothalamus and the suprachiasmatic
in neurons in other stress-responsive regions, such nucleus21. The inputs to the PVN from the median pre­
Spillover as the basolateral amygdala or the prefrontal cortex, optic nucleus and the posterior bed nucleus of the stria
When a neurotransmitter from is one order of magnitude higher than that observed terminalis (BNST) are primarily GABAergic. In addi­
one synapse acts at a in PNCs46,47. This physical architecture, featuring a tion, there is substantial GABAergic innervation from
neighbouring synapse; for
dearth of spines combined with the interspersed nature a ‘cloud’ or ‘halo’ of glutamate decarboxylase-expressing
example, when glutamate
escapes the synaptic cleft and of glutamate and GABA contacts, may contribute to neurons that surround the PVN30,59; this potentially
acts on nearby synapses. neurotransmitter spillover, heterosynaptic modulation and originates from neurons in the anterior hypothalamus
intracellular signalling crosstalk between glutamate (AHA) and the perifornical region of the lateral hypo­
Heterosynaptic modulation and GABAergic synapses. thalamus. Given the extensive innervation of these
When one transmitter system
(for example, glutamate) affects
Both glutamate and GABA have key roles in regu­ hypothalamic territories by all major limbic structures,
a neighbouring but different lating the activity of PNCs. Various glutamate receptor understanding the functional circuit formed by these
system (for example, GABA). (AMPA receptor (AMPAR), kainite receptor, NMDA inputs will be an important task for future investigation.

NATURE REVIEWS | NEUROSCIENCE VOLUME 16 | JULY 2015 | 379

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

A recent study using optogenetic methods revealed a of these receptors during stress initiates intracellular
functional septum–AHA–PVN circuit that regulates signalling necessary for the induction of plasticity at
neuroendocrine output following a single stressor 60. GABAergic synapses70,71 (see below). Glutamatergic
Similar approaches will be critical for revealing the synapses are a key target for several modulators that
necessity and sufficiency of different cell populations regulate HPA-axis output. For example, noradrenaline,
for mediating responses to distinct stressors. which potently activates the HPA axis during stress72,73,
The BNST is a major extrahypothalamic source of acts on α1‑adrenergic receptors to robustly increase the
GABAergic inputs. Activity mapping (that is, examin­ frequency of spontaneous glutamatergic synaptic cur­
ing FOS expression) and inactivation and lesion stud­ rents in PNCs74. Moreover, in vivo demonstrations that
ies have demonstrated an important role for BNST increases in circulating ACTH and CORT — induced
substructures in the regulation of the HPA axis61,62, but by electrical stimulation of noradrenergic afferents
a detailed dissection of the BNST–PNC circuit archi­ to the PVN or by intra-PVN administration of an
tecture is still needed to better understand the specific α1‑adrenergic receptor agonist — are blocked by intra-
contributions of defined inputs during various stressful PVN microinjections of glutamate receptor antagonists
situations. The brainstem, especially the A2 and C2 cell (both NMDAR and non-NMDAR antagonists) provide
groups of the nucleus tractus solitarius and the A1 and additional evidence for a pivotal role of glutamate in
C1 cell groups of the ventrolateral medulla, provides noradrenalin-induced HPA-axis activation75.
the main monoaminergic inputs to PVN. The expres­ Endocannabinoids (eCBs) have been implicated
sion of vesicular glutamate transporter 2 (VGLUT2; repeatedly in the negative regulation of HPA-axis
also known as SLC17A6) in subpopulations of noradr­ responses76–78. At the level of the synapse, eCBs are
energic neurons in the brainstem is consistent with a released from postsynaptic cells and act on CB1‑type
glutamate phenotype63, and recent work indicates that receptors that are located in presynaptic terminals.
a subpopulation of glucagon-like peptide 1 neurons in eCBs thus signal retrogradely to inhibit the release of
the caudal nucleus tractus solitarius sends glutamatergic neurotransmitters, including glutamate and GABA.
fibres to the PVN64. PNCs produce eCBs in an activity-dependent fashion79.
The release of these molecules at glutamatergic synapses
Glutamatergic synapses produces depolarization-induced suppression of excita­
Glutamatergic synapses under basal conditions. PNCs tion (DSE)79 that exhibits characteristics that are con­
express a range of ionotropic glutamate receptor subu­ sistent with well-characterized DSE at other synapses
nits that can form AMPARs, kainate receptors and (specifically, the DSE is CB1‑mediated and is dependent
NMDARs49. Electrophysiological recordings from PNCs on Ca2+ influx into the postsynaptic cell through L‑type
in acute brain slices ex vivo show both spontaneous and calcium channels) (FIG. 3). Although the contribution of
evoked synaptic currents that are sensitive to NMDAR this eCB-mediated, activity-dependent modulation of
and non-NMDAR antagonists31,34,35. During repeated glutamate transmission to HPA-axis activity remains to
synaptic recruitment at rates greater than 2 Hz, the release be determined, one possibility is that it curtails excita­
of glutamate from the presynaptic terminals onto PNCs tory synaptic input during periods of high PNC activ­
shows frequency-dependent short-term depression. This ity. In hypothalamic slices, application of CORT or its
means that transmission is more faithful (that is, it shows synthetic analogue, dexamethasone79–81, stimulates the
a higher fidelity) at lower rates of synaptic activity 34. One rapid synthesis of eCBs and inhibits glutamate release
potential consequence of this is that lowering the release from presynaptic terminals. This requires rapid actions
probability at these synapses could allow for more faithful through a non-canonical, putative membrane receptor
transmission at higher frequencies, thereby altering the that signals through a G protein and adenylase cyclase
filtering capacity of these synapses. activation pathway 5–12,80.
In addition to synaptic glutamate currents, extrasyn­
aptic NMDAR-mediated signalling has been described Glutamatergic synapses following acute stress. A sin­
in PNCs65. Histological studies have shown that the par­ gle acute stress alters signalling at glutamatergic syn­
vocellular subdivision of the PVN expresses high levels apses on PNCs. This manifests as an increase in the
of mRNAs for the kainate-preferring receptor subunits ratio of AMPAR- to NMDAR-mediated transmission35.
glutamate receptor ionotropic, kainate 1 (GLUK1; also Surprisingly, this is not due to an increase in signalling
known as GRIK1 and GLUR5) and GLUK5 (also known via AMPARs, but rather to a long-lasting decrease in
as GRIK5 and KA2)49, which are upregulated following NMDAR signalling 35 that results from the downregula­
stress66. Although the synaptic transmission to PNCs tion of postsynaptic NMDARs by local release of CRH
that is mediated by these kainate receptors remains to be during stress. It is unclear whether CRH is released in
characterized, GLUK1 expression is highly enriched in an autocrine fashion or originates from extra-PVN
CRH-releasing axon terminals at the median eminence, sources82,83. This CRH-mediated NMDAR downregula­
and has been implicated in the release of the hormone tion is an important priming step that allows these syn­
during stress67. Histological studies have also reported apses to exhibit a subsequent form of activity-dependent
low to moderate expression of group I mGluRs (mGluR1 plasticity, or metaplasticity 35. This metaplasticity is
and mGluR5) in PNCs68,69. Although the contributions induced by bursts of high-frequency presynaptic stimu­
of mGluR1 and mGluR5 to PNC excitability are largely lation and manifests as a robust short-term potentiation
unknown, recent studies demonstrate that the activation (STP) of excitatory transmission35 (FIG. 3). This dormant

380 | JULY 2015 | VOLUME 16 www.nature.com/reviews/neuro

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

b Naive and HFS c Acute stress and HFS (NMDAR)

a Naive
Glutamate
synapse
CB1

↑ Ca2+
Glutamate
CRH
AMPAR NMDAR
eCB d Acute stress and e Repeated stress
depolarization (eCB)

L-type Retrograde
VGCC signal

PNC
mGR
Depolarization CORT

Figure 3 | Glutamatergic synapses — function and plasticity.  a | Diagram depicts a glutamatergic synapse onto a
parvocellular neuroendocrine cell (PNC) from a naive (unstressed) animal. Fast transmissionNature Reviews
is mediated | Neuroscience
by AMPA
receptors (AMPARs) and NMDA receptors (NMDARs)34,35. Synapses also exhibit endocannabinoid (eCB)-mediated
depression. eCBs, recruited either by the postsynaptic depolarization and subsequent Ca2+ influx through L‑type
voltage-gated Ca2+ channels (VGCCs)79 or by the rapid actions of glucocorticoids (such as corticosterone (CORT))81, act at
presynaptic CB1 to decrease glutamate release. b | During bursts of afferent activity, NMDAR activation liberates a
retrograde signal that curtails release from glutamate terminals. c | Acute stress causes activation of postsynaptic cortico-
tropin-releasing hormone (CRH) receptors, resulting in a decrease in NMDAR signalling. This eliminates the retrograde
signal. As a result, glutamatergic synapses, when recruited with high activity rates, can release multiple vesicles of
transmitter. This results in a short-term potentiation that persists for a few minutes after the synaptic burst. d | Acute stress
also amplifies retrograde eCB signalling at these synapses. e | Repeated homotypic stress causes functional
downregulation of CB1. This manifests as a progressive loss of depolarization-induced suppression of excitation (DSE).
CORT is also ineffective in recruiting eCB signalling under these conditions. HFS, high-frequency stimulation; mGR,
membrane glucocorticoid receptor.

STP is not due to an increase in postsynaptic glutamate release84, but experiments have ruled out eCBs acting at
signalling or an increase in release probability; instead, it CB1, opioids or adenosine35. Future studies are needed
is due to a switch in the mode of release at glutamatergic to determine the identity of the retrograde messenger.
synapses from univesicular (that is, a presynaptic action More information is also needed to better understand
potential releases, at most, a single vesicle of glutamate) how this stress-induced metaplasticity at glutamatergic
to multivesicular (that is, a presynaptic action potential synapses may contribute to the adaptation of HPA-
can release more than one vesicle). Importantly, STP can axis responses to repeated stress. Pre-injection of the
also be unmasked in naive slices (slices from animals not NMDAR antagonists MK‑801 (also known as dizocil­
exposed to stress)35 by postsynaptic blockade of NMDAR, pine) or memantine increase Fos and Crh mRNA expres­
Ca2+ signalling and SNARE (soluble N‑ethylmaleimide- sion in the PVN and increase ACTH and CORT levels in
sensitive factor attachment protein receptor)-dependent the blood following two homotypic stressors5,85–87. This
exocytosis (by including pharmacological inhibitors in is consistent with the hypothesis that NMDAR down­
the recording pipette). These observations suggest that regulation plays a key part in amplifying responses to
CRH-mediated decreases in NMDAR-dependent Ca2+ subsequent stressors. Interestingly, glutamatergic syn­
entry during stress prevent the release of a retrograde apses in brain slices continue to exhibit a capacity for
messenger that normally blocks the multivesicular release multivesicular release for more than 72 hours after a sin­
of glutamate following high levels of presynaptic activ­ gle acute restraint stress35. This time course is consistent
ity (FIG. 3). There are various dendritically released fac­ with that of stress-induced HPA-axis sensitization, as
tors that could be responsible for suppressing glutamate reported elsewhere88.

NATURE REVIEWS | NEUROSCIENCE VOLUME 16 | JULY 2015 | 381

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

In addition to CRH, several other stress mediators by NMDARs65. These findings highlight a potentially
act both independently and in concert to drive complex important, but under-investigated, role for glial cells in
forms of metaplasticity at PNC synapses. For exam­ stress-induced glutamatergic synapse plasticity in the
ple, repeated exposure to the same stressor gradually PVN. The effects of chronic stress on PNC glutamatergic
diminishes CB1 signalling and this results in the loss of synapses raise the question of whether reducing stress
eCB-mediated inhibition of neurotransmitter release, levels may have the opposite effect. This was assessed in
including activity-dependent DSE (and depolarization- a study 93 in which rats were exposed to an enriched early
induced suppression of inhibition)79 (FIG. 3). This func­ life environment (augmented maternal care), which is
tional downregulation of CB1 requires the activation of known to persistently attenuate neuroendocrine stress
cytosolic glucocorticoid receptors. As discussed below, responses94–96. Under these conditions, there is a decrease
the actions of noradrenaline and CORT during stress in mEPSC frequency in PNCs that is accompanied by a
functionally upregulate mGluR1 and mGluR5, respec­ decrease in the number of glutamatergic synapses and
tively, at PNC synapses. This stress-induced modulation reduced VGLUT2 expression93. This reduction in the
of mGluRs primes PNC synapses for subsequent activity- number of glutamatergic synapses is transient; that is, it
dependent plasticity of GABAergic synapses70,71, thus in is evident in neonates immediately after early life expe­
a way resembling CRH-mediated metaplasticity at gluta­ riences, but resolves by adolescence, even though other
matergic synapses. Clearly, a state-dependent change in changes relevant to HPA-axis responses (for example,
synaptic plasticity is an emerging idea that needs further transcriptional repression of Crh) persist until adult­
investigation. hood. The authors proposed that synaptic changes in
early life initiate lifelong epigenetic changes that affect
Long-term changes in glutamate transmission after subsequent transcription events in PNCs.
stress. Exposure to chronic stress, or even to a single
stressor, can cause long-term changes in HPA-axis GABAergic synapses
activity and responsiveness, resulting in, for example, GABAergic synapses under basal conditions. GABA
baseline hyperactivity and hypersensitivity to a novel transmission mediated by anion-permeable, pentameric
stressor 88–90. Like other types of experience-dependent GABAA receptors (GABAARs) consists of distinct phasic
learning, such changes in the sensitivity and/or gain of and tonic currents. PVN neurons express α1‑2, β1‑3, and
HPA-axis responses probably involve activity-depend­ γ1‑2 GABAAR subunits and exhibit phasic GABAAR-
ent, long-term synaptic changes. Neuroanatomical gated synaptic currents97. In addition, tonic (extrasyn­
and electrophysiological evidence collectively suggests aptic) inhibition mediated by δ‑subunits has also been
that experience (stress)-dependent plasticity occurs demonstrated98. The specific localization and function
at PNC glutamatergic synapses. Only a few examples of other GABAAR subunits remains unexplored. GABA
of this type of plasticity have been described to date, release at synaptic contacts can occur spontaneously
and most of these involve changes that are consistent — that is, in the absence of action potentials (observed
with an increase in the number of glutamatergic syn­ as miniature inhibitory PSCs) — and following elec­
apses. For example, repeated variable stress in male trical stimulation to drive action potential-dependent
rats for 1 week increases the number of boutons that release events (for example, paired-pulse depression,
are immunopositive for VGLUT2 on identified CRH which is observed at over 80% of synapses)79,99. In
neurons91. Repeated variable stress over 3 weeks causes cell-attached recordings in which anion gradients are
approximately a twofold increase in the number of syn­ unperturbed, GABAAR blockade increases the activity
aptic contacts terminating on CRH-expressing PNCs44; of PNC56. These data are largely consistent with in vivo
this study reported increases in both GABA and non- studies showing that ongoing GABA transmission in
GABA (that is, putative glutamatergic) terminals. the absence of a stressor constrains PNC activity and,
Functionally, repeated homotypic stress (restraint) for thereby, HPA-axis output 54.
3 days in rats increases the frequency of spontaneous
excitatory postsynaptic currents (sEPSCs)92. Likewise, GABAergic synapses following acute stress. Given the
early life stress (due to poor maternal care) increases the evidence for tonic constraint of HPA-axis activity by
frequency of sEPSCs and miniature EPSCs (mEPSCs), GABA, it is worth considering possible mechanisms
which is paralleled by an increase in VGLUT2 expres­ through which disinhibition may occur at the onset of
sion in the PVN in animals examined after weaning 65. a stressor. One possibility is that GABA neurons that
These findings collectively showing an increase in directly synapse onto PNCs may become less active,
excitatory glutamate inputs to PNCs begin to provide a although currently there is no direct evidence in sup­
foundation for earlier studies indicating that these stress port of this idea. This may be due to a lack of definitive
exposures cause a lasting sensitization and/or baseline in vivo recording data and the limitations of immedi­
hyperactivity of the HPA axis. ate early gene mapping. For example, FOS mapping
In addition to inducing the changes in phasic (synap­ studies are instructive, but fail to reflect short-lived
tic) transmission described above, early-life stress results increases in activity or decreases in activity, for exam­
in impairment of astrocytic glutamate transporter 1 (also ple, in nuclei that provide GABA inputs to PVN59. A
Homotypic stress
known as excitatory amino acid transporter 2) func­ second possible mechanism for disinhibition is that
Repeated administration of the tion, which may contribute to the unmasking of a tonic GABAergic synapses become less effective, either
same stressor to an animal. (extrasynaptic) conductance that is primarily mediated through a decrease in the probability of GABA release

382 | JULY 2015 | VOLUME 16 www.nature.com/reviews/neuro

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

or owing to a decrease in the number or function of kinase II drive the insertion of GABAA receptors into
postsynaptic GABAARs. Observations that acute stress the PNC (that is, the postsynaptic) membrane70 (FIG. 4).
decreases the probability of GABA release are consist­ By contrast, LTDGABA requires CORT acting on cytosolic
ent with this idea100 (but also see REF. 71). An alterna­ glucocorticoid receptors in PNCs71. The activation of glu­
tive explanation is that GABAAR-mediated inhibition is cocorticoid receptors inhibits an intracellular regulator of
reduced at the onset of acute stress through alterations G protein signalling (regulator of G protein signalling 4),
in chloride homeostasis in the postsynaptic cell56,98. which amplifies mGluR5 signalling and culminates in the
This is based on electrophysiological recordings in liberation of the opioid peptide enkephalin from somato-
slices prepared immediately after a single 30‑minute dendritic vesicles71 (FIG. 4). Enkephalin acts as a retrograde
restraint stress, which reported a depolarizing shift in signal and inhibits the release of GABA from presynaptic
chloride reversal potential in PNCs56,98. This altera­ terminals.
tion in chloride homeostasis requires a decrease in
the functional capacity of K‑Cl co‑transporter 2 (KCC2; Changes in GABA transmission following chronic stress.
also known as SLC12A5)56, which is probably due to Chronic or repeated stress can also cause several altera­
decreased membrane expression of the transporter 98. tions in PNC GABAergic synapses. These may include
One consequence of this reduction in chloride extru­ changes in the expression of GABAAR subunits103,104
sion capacity is that engagement of GABA afferents and a reduction in retrograde eCB signalling 79. Perhaps
has a nonlinear effect on PNC firing 101: at low levels of the most striking observations are those of putative
GABA input the decrease in KCC2 function results in a GABAergic synaptogenesis and subcellular synapse
weakening of inhibition, whereas at higher rates of syn­ redistribution. Under non-stress conditions, GABAergic
aptic transmission and/or after enhancement of tonic synapses are uniformly distributed between somatic and
GABAAR currents there is a paradoxical GABAAR- dendritic compartments of PNCs44. Chronic stress causes
dependent excitation that increases action potential two distinct changes. First, there is a dramatic increase
firing 56,98. Consistent with this model, genetic deletion in the total number of GABAergic synaptic contacts;
of GABAAR δ‑subunits102 or pharmacological block­ second, there is a change in relative distribution of these
ade of positive allosteric modulation of extrasynaptic contacts, such that dendritic inputs become more highly
GABAAR δ‑subunits blunts HPA-axis responses98. represented44. The functional consequences of such struc­
tural plasticity warrant further investigation but, interest­
Kairoplasticity at GABAergic synapses. Recent experi­ ingly, chronic stress results in a paradoxical decrease in
ments offer new insights into the rules and expres­ the number of functional GABAergic synapses in elec­
sion mechanisms for synaptic plasticity at GABAergic trophysiological recordings104. One possible explanation
synapses on PNCs. These findings (discussed in detail for this is that chronic stress may lead to the formation
below) describe a form of metaplasticity but are unique of (functionally) silent synapses that are subjected to
from classical metaplasticity in that the timing of the dynamic recruitment by subsequent synaptic activities in
induction after the initial exposure to stress is critical. a mechanism that involves an LTPGABA-like mechanism70.
In naive (not exposed to stress) animals, GABAergic Furthermore, the effect of this synaptic rearrangement on
synapses on PNCs fail to exhibit long-term plasticity 70,71. HPA-axis function remains unknown. In many neuronal
However, after exposure to either a physical or a psy­ circuits, the location of GABAergic synaptic input onto
chological stressor, both long-term potentiation (LTP)70 a principal or projection neuron reflects innervation by
and long-term depression (LTD)71 can be induced at functionally distinct GABA subpopulations, which mod­
GABAergic synapses (LTPGABA and LTD GABA, respec­ ify neuronal firing and/or integration of excitatory input
tively). Importantly (and surprisingly), whether synapses in a very specific fashion. A rearrangement of GABA
exhibit LTP or LTD is not a consequence of the induction inputs onto PNCs may alter the relative contributions
protocol used; rather, it is a function of when slices were of specific nuclei. Such compartmentalization of GABA
prepared after stress. Specifically, LTP can be induced input also determines the efficacy of inhibition through
only in slices prepared immediately after a stress, whereas activity-dependent shifts in the chloride driving force. A
LTD can be induced only in slices prepared 90 minutes collapse of inhibition under high rates of synaptic activ­
after stress. These temporal windows for opposing forms ity occurs preferentially at dendritic GABA inputs, where
of plasticity arise because two specific neuromodulators, chloride extrusion mechanisms and diffusional capacity
noradrenaline and CORT, exhibit discrete time domains are volume-limited105. Further investigation will be nec­
for actions in the PVN13. More precisely, noradrenaline, essary to address how changes at specific GABAergic
which is released at the onset of stress, is the essential synapses contribute to chronic stress-induced HPA-
associative signal for LTPGABA. That is, activation of axis adaptations and the behavioural and physiological
β‑adrenergic receptors and the downstream activation pathologies that accompany them.
K‑Cl co‑transporter 2 of protein kinase A cause a functional upregulation of
(KCC2). A transmembrane (otherwise dormant) mGluR1 on PNCs. This noradren­ Functional implications
potassium–chloride aline-induced priming of mGluR1 creates a time window It is increasingly clear that synapses onto PNCs are
co‑transporter that extrudes in which subsequent bursts of glutamate afferent activ­ dynamic, can be modified by experience and need to
chloride and maintains the
driving force for chloride influx
ity effectively activates mGluR1. An mGluR1‑dependent be an important part of any discussion that invokes
into cells upon the opening of increase in intracellular Ca2+ and the ensuing phospho­ plasticity of responses to stress. The next challenge is to
GABAA receptors. rylation of calcium/calmodulin-dependent protein understand how these mechanistically discrete forms of

NATURE REVIEWS | NEUROSCIENCE VOLUME 16 | JULY 2015 | 383

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

a Naive b HFS or presynaptic c Presynaptic stimulation


stimulation paired with paired with postsynaptic
GABA postsynaptic depolarization depolarization
synapse
CB1

GABAAR GABA
eCB μOR
Nucleus
PNC
Opioid β-AR RGS4
L-type retrograde
VGCC signal NA GR
AMPAR CORT
mGluR5
mGluR1
Glutamate

Depolarization Glutamate
synapse

d Acute stress and postsynaptic e Repeated stress f Novel stress


depolarization (eCB)

Figure 4 | GABAergic synapses — function and plasticity.  a | Diagram depicts a GABAergic Nature Reviews
synapse on a| Neuroscience
parvocellular
neuroendocrine cell (PNC). Fast transmission is mediated by GABAA receptors (GABAARs). Synapses exhibit
endocannabinoid (eCB)-mediated depression. eCBs, recruited by postsynaptic depolarization and subsequent Ca2+ influx
through L‑type voltage-gated Ca2+ channels (VGCCs)79, act at presynaptic CB1 to decrease GABA release. b | An acute
stress causes local release of noradrenaline (NA) and activation of β‑adrenoceptors (β‑ARs). This primes group I
metabotropic glutamate receptors (mGluRs) in PNCs through a mechanism that is unresolved70. Once primed, mGluR
activation increases intracellular Ca2+, which drives phosphorylation of calcium/calmodulin-dependent protein kinase II
(CAMKII) and the insertion of GABAARs at previously silent GABAergic synapses (not shown). c | If a longer time
(90 minutes) is allowed for the animal to recover after stress, a different form of plasticity is observed. The longer recovery
period allows circulating corticosterone (CORT) to affect biochemical processes in PNCs. Specifically, CORT amplifies
signalling through mGluR5, by inhibiting the regulator of G protein signalling 4 (RGS4) in PNCs71. This removes the ‘brake’
from mGluR5 downstream signalling and allows synaptically released glutamate to liberate an opioid retrograde signal.
The opioid, which is probably enkephalin, binds to presynaptic μ‑opioid receptors (μOR) and decreases GABA release
probability. d | Acute stress also amplifies retrograde eCB signalling at GABAergic synapses. e | Repeated homotypic stress
causes a downregulation of CB1 and consequent loss of eCB signalling. f | Exposure to a novel stressor after repeated
homotypic stress causes an immediate reinstatement of eCB signalling. This is due to the recovery of CB1Rs at presynaptic
terminals. The effects of novel stressors can be mimicked either by globally increasing neural activity using
electroconvulsive seizures or by increasing synaptic activity in situ. AMPAR, AMPA receptor; GR, glucocorticoid receptor;
HFS, high-frequency stimulation.

plasticity contribute to dynamic, experience-dependent prolonged stress65. This has recently been reviewed in
modulation of CRH neuron excitability and the result­ detail elsewhere101. In brief, LTPGABA, which increases
ant HPA-axis activity. At GABAergic synapses, syn­ the efficacy of chloride flow across the membrane,
aptic changes must be considered within a theoretical enhances the robustness of the GABA response. That is,
framework that takes into account the shifts in chlo­ in a simplistic scenario, LTPGABA causes more-effective
ride homeo­stasis that accompany both acute56,98 and inhibition when the chloride reversal potential is below

384 | JULY 2015 | VOLUME 16 www.nature.com/reviews/neuro

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

Box 1 | Plasticity, metaplasticity and kairoplasticity


In the classic interpretation of synaptic plasticity (see the figure, part a), presentation of a stimulus (S1) that engages two synaptically coupled
neurons (shown in green and blue) modifies the efficacy of the synapse between them. Depending on the stimulus, the original synaptic response
(shown by the medium size of the synapse) may
be potentiated (shown as a larger synapse;
synaptic strength >1) or depressed (shown as a a S1 b S1 S2
smaller synapse; synaptic strength <1).
In synaptic metaplasticity between synaptically
coupled neurons (see the figure, part b),
presentation of S1 causes embedded
biochemical changes (shown as a change in the + +
colour of the postsynaptic cell) but does not

Synaptic strength

Synaptic strength
change synaptic strength. Presentation of a
subsequent stimulus (S2) may either potentiate
1 1
(synaptic strength >1) or depress (synaptic
strength <1) the synapse. The direction of the
plasticity depends on the nature of S2.
– –
In kairoplasticity (see the figure, part c), the Time Time
presentation of S1 causes biochemical changes
c S1 S2 S1 S2
but, as in metaplasticity, does not change
synaptic strength. Again, presentation of S2 may
t
either potentiate (synaptic strength >1) or
depress (synaptic strength <1) the synapse.
However, the direction of the plasticity depends
on the temporal window after S1 rather than on + +
Synaptic strength

Synaptic strength
the nature of S2. For example, if S2 is presented
within a narrow temporal window of S1, the
synapse is potentiated (see the figure, part c, left 1 1
panel). If the temporal window is longer (as
represented by the variable t), S2 will cause a
synaptic depression despite being the same – –
stimulus (see the figure, part c, right panel). Time Time
Specific neuromodulators may be important for Biochemically changed
Presynaptic neuron Postsynaptic neuron
determining the temporal windows in postsynaptic neuron
kairoplasticity.
Nature Reviews | Neuroscience

the resting membrane potential (that is, in the naive, Conclusions and future perspectives
unstressed state) and more excitation when the chlo­ The descriptions of multiple forms of plasticity at synapses
ride reversal potential is above the resting membrane onto PNCs indicate that there are multiple mechanisms
potential (that is, after stress). However, it is impor­ by which the brain adapts to stressors. Importantly, they
tant to consider that chloride homeostasis is dynamic provide a framework in which to ask more precise ques­
and is directly influenced by the chloride flux through tions about how organisms use information from one
GABAARs. Thus, during a period of postsynaptic depo­ exposure to stress to make adjustments to ensuing stress­
larization (that is, during persistent spiking activity) ful experiences. Here, we have described multiple embed­
that generates a large driving force for chloride influx, ded biochemical changes following an acute stress that
LTPGABA augments the rate of chloride influx (chlo­ can bias synapses towards plasticity when these synapses
ride loading)56,106. When combined with the impaired are subsequently recruited in an activity-dependent fash­
chloride extrusion mechanism (for example, through ion. This is consistent with the definition of metaplastic­
downregulation of KCC2) during stress, LTP GABA ity 110. We have also noted that the nature of the plasticity
favours the development of subsequent depolarizing is dependent on how soon after stress the synapses are
GABA response and can, paradoxically, contribute to interrogated. In other words, there seem to be opportune
the stress-induced sensitization of PNCs. By contrast, moments for the induction of distinct forms of plastic­
LTD GABA can have the opposite effect. In this model, ity following stress. Temporal windows for bidirectional
the role of LTP GABA and LTD GABA in shaping HPA- plasticity following stress have also been described in the
axis responses fits with existing evidence from in vivo amygdala111, suggesting that it may be important to incor­
studies suggesting that noradrenaline contributes to porate time as a key variable when discussing plasticity or
sensitization of HPA-axis output 107,108 and that CORT metaplasticity. Consequently, we propose a new descriptor
provides negative feedback to prevent overactivation of for the lexicon of plasticity, kairoplasticity, which incorpo­
the HPA axis109. The discovery of LTPGABA and LTDGABA rates distinct temporal windows for bidirectional changes
in PNCs70,71 provides an important foundation to test in synaptic strength (BOX 1).
this model and to tease apart the mechanistic underpin­ Derived from the Greek word ‘kairos’, meaning the
nings of neuroendocrine adaptations during repeated, opportune or correct moment, kairoplasticity refers
and subsequently chronic, stress experiences. to observations that specific and distinct temporal

NATURE REVIEWS | NEUROSCIENCE VOLUME 16 | JULY 2015 | 385

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

windows after stress are instructive for different forms many important and unresolved issues. The adoption
of metaplasticity. One plausible consequence of such a of new tools that enable specific activation of identified
mechanism is that it contributes to the versatility of the cell populations will be extremely beneficial for gain­
system, enabling context-specific adjustment of the stress ing a better understanding of the interactions between
response. These observations also raise new questions: co‑transmitters in the PVN. In this vein, recent studies
for example, how are these distinct modes of plasticity have described new mouse models that allow for both
encoded by one stressor affected by subsequent experi­ the identification and the genetic targeting of specific
ences? Glutamatergic synapses onto PNCs, for exam­ subsets of PNCs99,113,114. One important limitation that
ple, are primed for multivesicular release for more than must be addressed is the dearth of information about the
3 days after a single stress challenge35; how subsequent activity of PNCs and CRH-producing neurons in vivo,
stressors are integrated during this time window remains particularly in response to stress exposure; developing
unclear. Moreover, it is possible that mechanisms exist and implementing approaches that allow for the use of
to reverse these synaptic changes. For example, living in electrophysiology and/or imaging techniques in awake,
an ‘enriched’ environment could be one way in which behaving animals will be a considerable advance in help­
stress-embedded synaptic changes are erased. Recent ing to make causal links between changes in synaptic
work demonstrating that stress-induced loss of eCB sig­ function or plasticity and the output of stress circuitry. By
nalling at PNC synapses can be restored following expo­ filling these knowledge gaps, we will begin to have a bet­
sure to novel stimuli112 is consistent with the idea that ter understanding of how specific inputs shape outputs
the salience of an experience may be a critical element from the PVN and we will also gain insights into how we
in returning HPA-axis activity to baseline112. Although might exploit the capacity of these inputs to be modified
recent studies have increased our understanding of syn­ to help reset dysfunctional neural circuits in individuals
aptic function and plasticity in the PVN, there are still with stress-related conditions, such as depression.

1. McEwen, B. S. Mood disorders and allostatic load. 17. Bruhn, T. O., Plotsky, P. M. & Vale, W. W. Effect of demonstrated by quantitative immunoelectron
Biol. Psychiatry 54, 200–207 (2003). paraventricular lesions on corticotropin-releasing microscopy. Neuroscience 113, 581–592 (2002).
2. Dias, C. et al. β‑catenin mediates stress resilience factor (CRF)-like immunoreactivity in the stalk- This paper provides electron microscopy evidence
through Dicer1/microRNA regulation. Nature 516, median eminence: studies on the that GABA innervation of CRH-producing neurons
51–55 (2014). adrenocorticotropin response to ether stress and exhibits morphological plasticity.
3. Chaudhury, D. et al. Rapid regulation of depression- exogenous CRF. Endocrinology 114, 30. Boudaba, C., Szabo, K. & Tasker, J. G. Physiological
related behaviours by control of midbrain dopamine 57–62 (1984). mapping of local inhibitory inputs to the hypothalamic
neurons. Nature 493, 532–536 (2013). 18. Swanson, L. W. & Sawchenko, P. E. Paraventricular paraventricular nucleus. J. Neurosci. 16, 7151–7160
4. Dias-Ferreira, E. et al. Chronic stress causes nucleus: a site for the integration of neuroendocrine (1996).
frontostriatal reorganization and affects decision- and autonomic mechanisms. Neuroendocrinology 31, Using glutamate microdrops, this study maps the
making. Science 325, 621–625 (2009). 410–417 (1980). potential locations of GABA cells that provide input
5. Armario, A., Escorihuela, R. M. & Nadal, R. Long-term 19. Denver, R. J. Structural and functional evolution of to the PVN.
neuroendocrine and behavioural effects of a single vertebrate neuroendocrine stress systems. Ann. NY 31. Boudaba, C., Schrader, L. A. & Tasker, J. G.
exposure to stress in adult animals. Neurosci. Acad. Sci. 1163, 1–16 (2009). Physiological evidence for local excitatory synaptic
Biobehav. Rev. 32, 1121–1135 (2008). 20. Makara, G. B., Stark, E., Kapocs, G. & Antoni, F. A. circuits in the rat hypothalamus. J. Neurophysiol. 77,
6. Andrés, R., Martí, O. & Armario, A. Direct evidence of Long-term effects of hypothalamic paraventricular 3396–3400 (1997).
acute stress-induced facilitation of ACTH response to lesion on CRF content and stimulated ACTH secretion. 32. Gordon, G. R. & Bains, J. S. Noradrenaline triggers
subsequent stress in rats. Am. J. Physiol. 277, Am. J. Physiol. 250, E319–E324 (1986). multivesicular release at glutamatergic synapses in the
R863–R868 (1999). 21. Ulrich-Lai, Y. M. & Herman, J. P. Neural regulation of hypothalamus. J. Neurosci. 25, 11385–11395
7. Ons, S., Rotllant, D., Marín-Blasco, I. J. & endocrine and autonomic stress responses. Nat. Rev. (2005).
Armario, A. Immediate-early gene response to Neurosci. 10, 397–409 (2009). 33. Iremonger, K. J. & Bains, J. S. Integration of
repeated immobilization: Fos protein and arc mRNA 22. Makara, G. B. & Stark, E. Effects of gamma- asynchronously released quanta prolongs the
levels appear to be less sensitive than c‑fos mRNA to aminobutyric acid (GABA) and GABA antagonist drugs postsynaptic spike window. J. Neurosci. 27,
adaptation. Eur. J. Neurosci. 31, 2043–2052 on ACTH release. Neuroendocrinology 16, 178–190 6684–6691 (2007).
(2010). (1974). 34. Marty, V., Kuzmiski, J. B., Baimoukhametova, D. V. &
8. Dallman, M. F. et al. Stress, feedback and facilitation 23. Makara, G. B. & Stark, E. Effect of intraventricular Bains, J. S. Short-term plasticity impacts information
in the hypothalamo-pituitary-adrenal axis. glutamate on ACTH release. Neuroendocrinology 18, transfer at glutamate synapses onto parvocellular
J. Neuroendocrinol. 4, 517–526 (1992). 213–216 (1975). neuroendocrine cells in the paraventricular nucleus of
9. Bhatnagar, S. & Dallman, M. Neuroanatomical basis 24. Decavel, C. & van den Pol, A. N. Converging GABA- the hypothalamus. J. Physiol. 589, 4259–4270
for facilitation of hypothalamic-pituitary-adrenal and glutamate-immunoreactive axons make (2011).
responses to a novel stressor after chronic stress. synaptic contact with identified hypothalamic 35. Kuzmiski, J. B., Marty, V., Baimoukhametova, D. V. &
Neuroscience 84, 1025–1039 (1998). neurosecretory neurons. J. Comp. Neurol. 316, Bains, J. S. Stress-induced priming of glutamate
10. Keller-Wood, M. E. & Dallman, M. F. Corticosteroid 104–116 (1992). synapses unmasks associative short-term plasticity.
inhibition of ACTH secretion. Endocr. Rev. 5, 1–24 25. Decavel, C. & van den Pol, A. N. GABA: a dominant Nat. Neurosci. 13, 1257–1264 (2011).
(1984). neurotransmitter in the hypothalamus. J. Comp. 36. Atasoy, D., Betley, J. N., Su, H. H. & Sternson, S. M.
11. Girotti, M. et al. Habituation to repeated restraint Neurol. 302, 1019–1037 (1990). Deconstruction of a neural circuit for hunger. Nature
stress is associated with lack of stress-induced c‑fos 26. van den Pol, A. N., Wuarin, J. P. & Dudek, F. E. 488, 172–177 (2012).
expression in primary sensory processing areas of the Glutamate, the dominant excitatory transmitter in 37. Betley, J. N., Cao, Z. F. H., Ritola, K. D. &
rat brain. Neuroscience 138, 1067–1081 (2006). neuroendocrine regulation. Science 250, 1276–1278 Sternson, S. M. Parallel, redundant circuit
12. Grissom, N. & Bhatnagar, S. Habituation to repeated (1990). organization for homeostatic control of feeding
stress: get used to it. Neurobiol. Learn. Mem. 92, This was the first demonstration of glutamate as a behavior. Cell 155, 1337–1350 (2013).
215–224 (2009). fast, excitatory synaptic signal in the PVN. 38. Kiss, J. Z. et al. Quantitative histological studies on the
13. Joels, M. & Baram, T. Z. The neuro-symphony of 27. van den Pol, A. N. Glutamate and aspartate hypothalamic paraventricular nucleus in rats: I.
stress. Nat. Rev. Neurosci. 10, 459–466 (2009). immunoreactivity in hypothalamic presynaptic axons. Number of cells and synaptic boutons. Brain Res.
14. Sapolsky, R. M., Krey, L. C. & McEwen, B. S. Prolonged J. Neurosci. 11, 2087–2101 (1991). 262, 217–224 (1983).
glucocorticoid exposure reduces hippocampal neuron 28. Meister, B., Hokfelt, T., Geffard, M. & Oertel, W. 39. Kiss, J. Z. et al. Quantitative histological studies on
number: implications for aging. J. Neurosci. 5, Glutamic acid decarboxylase- and γ-aminobutyric the hypothalamic paraventricular nucleus in rats. II.
1222–1227 (1985). acid-like immunoreactivities in corticotropin- Number of local and certain afferent nerve
15. Sapolsky, R. M. & Pulsinelli, W. A. Glucocorticoids releasing factor-containing parvocellular neurons of terminals. Brain Res. 265,
potentiate ischemic injury to neurons: therapeutic the hypothalamic paraventricular nucleus. 11–20 (1983).
implications. Science 229, 1397–1400 (1985). Neuroendocrinology 48, 516–526 (1988). 40. Hiscock, J. J., Murphy, S. & Willoughby, J. O. Confocal
16. Pavlides, C., Watanabe, Y. & McEwen, B. S. Effects of 29. Miklos, I. H. & Kovacs, K. J. GABAergic innervation of microscopic estimation of GABAergic nerve terminals
glucocorticoids on hippocampal long-term corticotropin-releasing hormone (CRH)-secreting in the central nervous system. J. Neurosci. Methods
potentiation. Hippocampus 3, 183–192 (1993). parvocellular neurons and its plasticity as 95, 1–11 (2000).

386 | JULY 2015 | VOLUME 16 www.nature.com/reviews/neuro

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

41. Liposits, Z., Phelix, C. & Paull, W. K. Adrenergic 60. Anthony, T. E. et al. Control of stress-induced persistent This paper shows that homotypic stress causes a
innervation of corticotropin releasing factor (CRF)- anxiety by an extra-amygdala septohypothalamic collapse of eCB signalling at GABA and
synthesizing neurons in the hypothalamic circuit. Cell 156, 522–536 (2014). glutamatergic synapses.
paraventricular nucleus of the rat. A combined light 61. Zhu, W., Umegaki, H., Suzuki, Y., Miura, H. & Iguchi, A. 80. Di, S., Maxson, M. M., Franco, A. & Tasker, J. G.
and electron microscopic immunocytochemical study. Involvement of the bed nucleus of the stria terminalis Glucocorticoids regulate glutamate and GABA
Histochemistry 84, 201–205 (1986). in hippocampal cholinergic system-mediated activation synapse-specific retrograde transmission via divergent
42. DePuy, S. D. et al. Glutamatergic neurotransmission of the hypothalamo–pituitary–adrenocortical axis in nongenomic signaling pathways. J. Neurosci. 29,
between the C1 neurons and the parasympathetic rats. Brain Res. 916, 101–106 (2001). 393–401 (2009).
preganglionic neurons of the dorsal motor nucleus of 62. Mongeau, R., Miller, G. A., Chiang, E. & 81. Di, S., Malcher-Lopes, R., Halmos, K. C. & Tasker, J. G.
the vagus. J. Neurosci. 33, 1486–1497 (2013). Anderson, D. J. Neural correlates of competing fear Nongenomic glucocorticoid inhibition via
43. Holloway, B. B. et al. Monosynaptic glutamatergic behaviors evoked by an innately aversive stimulus. endocannabinoid release in the hypothalamus: a fast
activation of locus coeruleus and other lower J. Neurosci. 23, 3855–3868 (2003). feedback mechanism. J. Neurosci. 23, 4850–4857
brainstem noradrenergic neurons by the C1 cells in 63. Stornetta, R. L., Macon, C. J., Nguyen, T. M., Coates, M. B. (2003).
mice. J. Neurosci. 33, 18792–18805 (2013). & Guyenet, P. G. Cholinergic neurons in the mouse rostral This was the frst demonstration of the rapid effects
44. Miklos, I. H. & Kovacs, K. J. Reorganization of synaptic ventrolateral medulla target sensory afferent areas. Brain CORT has on eCB signalling.
inputs to the hypothalamic paraventricular nucleus Struct. Funct. 218, 455–475 (2013). 82. Ono, N., Bedran de Castro, J. C. & McCann, S. M.
during chronic psychogenic stress in rats. Biol. 64. Zheng, H., Stornetta, R. L., Agassandian, K. & Ultrashort-loop positive feedback of corticotropin (ACTH)-
Psychiatry 71, 301–308 (2012). Rinaman, L. Glutamatergic phenotype of glucagon-like releasing factor to enhance ACTH release in stress. Proc.
This paper provides anatomical evidence for peptide 1 neurons in the caudal nucleus of the solitary Natl Acad. Sci. USA 82, 3528–3531 (1985).
changes in synaptic inputs to PVN neurons tract in rats. Brain Struct. Funct. http://dx.doi. 83. Parkes, D. G., Yamamoto, G. Y., Vaughan, J. M. &
following stress. org/10.1007/s00429-014-0841-6 (2014). Vale, W. W. Characterization and regulation of
45. Rho, J. H. & Swanson, L. W. A morphometric analysis 65. Gunn, B. G. et al. Dysfunctional astrocytic and synaptic corticotropin-releasing factor in the human hepatoma
of functionally defined subpopulations of neurons in regulation of hypothalamic glutamatergic transmission NPLC‑KC cell line. Neuroendocrinology 57, 663–669
the paraventricular nucleus of the rat with in a mouse model of early-life adversity: relevance to (1993).
observations on the effects of colchicine. J. Neurosci. neurosteroids and programming of the stress 84. Regehr, W. G., Carey, M. R. & Best, A. R. Activity-
9, 1375–1388 (1989). response. J. Neurosci. 33, 19534–19554 (2013). dependent regulation of synapses by retrograde
46. Mitra, R., Jadhav, S., McEwen, B. S., Vyas, A. & This was the first demonstration of synaptic effects messengers. Neuron 63, 154–170 (2009).
Chattarji, S. Stress duration modulates the in the PVN of early life stress. 85. Givalois, L., Arancibia, S. & Tapia-Arancibia, L.
spatiotemporal patterns of spine formation in the 66. Koenig, J. I. & Cho, J. Y. Provocation of kainic acid Concomitant changes in CRH mRNA levels in rat
basolateral amygdala. Proc. Natl Acad. Sci. USA 102, receptor mRNA changes in the rat paraventricular hippocampus and hypothalamus following
9371–9376 (2005). nucleus by insulin-induced hypoglycaemia. immobilization stress. Brain Res. Mol. Brain Res. 75,
47. Radley, J. J., Anderson, R. M., Hamilton, B. A., J. Neuroendocrinol. 17, 111–118 (2005). 166–171 (2000).
Alcock, J. A. & Romig-Martin, S. A. Chronic stress- 67. Evanson, N. K., Van Hooren, D. C. & Herman, J. P. 86. Pechnick, R. N., Bresee, C. J. & Poland, R. E. The role
induced alterations of dendritic spine subtypes predict GluR5‑mediated glutamate signaling regulates of antagonism of NMDA receptor-mediated
functional decrements in an hypothalamo-pituitary- hypothalamo-pituitary-adrenocortical stress responses neurotransmission and inhibition of the dopamine
adrenal-inhibitory prefrontal circuit. J. Neurosci. 33, at the paraventricular nucleus and median eminence. reuptake in the neuroendocrine effects of
14379–14391 (2013). Psychoneuroendocrinology 34, 1370–1379 (2009). phencyclidine. Life Sci. 78, 2006–2011 (2006).
48. Aubry, J. M., Bartanusz, V., Pagliusi, S., Schulz, P. & 68. van den Pol, A. N. Metabotropic glutamate receptor 87. Lee, S., Rivier, C. & Torres, G. Induction of c‑fos and
Kiss, J. Z. Expression of ionotropic glutamate receptor mGluR1 distribution and ultrastructural localization in CRF mRNA by MK‑801 in the parvocellular
subunit mRNAs by paraventricular corticotropin- hypothalamus. J. Comp. Neurol. 349, 615–632 (1994). paraventricular nucleus of the rat hypothalamus. Brain
releasing factor (CRF) neurons. Neurosci. Lett. 205, 69. Kocsis, K., Kiss, J., Gorcs, T. & Halasz, B. Metabotropic Res. Mol. Brain Res. 24, 192–198 (1994).
95–98 (1996). glutamate receptor in vasopressin, CRF and VIP 88. Armario, A., Martí, O., Vallès, A., Dal-Zotto, S. &
49. Herman, J. P., Eyigor, O., Ziegler, D. R. & Jennes, L. hypothalamic neurones. Neuroreport 9, 4029–4033 Ons, S. Long-term effects of a single exposure to
Expression of ionotropic glutamate receptor subunit (1998). immobilization on the hypothalamic-pituitary-adrenal
mRNAs in the hypothalamic paraventricular nucleus of 70. Inoue, W. et al. Noradrenaline is a stress-associated axis: neurobiologic mechanisms. Ann. NY Acad. Sci.
the rat. J. Comp. Neurol. 422, 352–362 (2000). metaplastic signal at GABA synapses. Nat. Neurosci. 1018, 162–172 (2004).
50. Eyigor, O., Centers, A. & Jennes, L. Distribution of 16, 605–612 (2013). 89. Akana, S. F. et al. Feedback and facilitation in the
ionotropic glutamate receptor subunit mRNAs in the This paper shows that stress causes a postsynaptic adrenocortical system: unmasking facilitation by
rat hypothalamus. J. Comp. Neurol. 434, 101–124 LTP due to GABAA receptor insertion that requires partial inhibition of the glucocorticoid response to
(2001). priming by noradrenaline. prior stress. Endocrinology 131, 57–68 (1992).
51. Ziegler, D. R., Cullinan, W. E. & Herman, J. P. 71. Wamsteeker Cusulin, J. I., Füzesi, T., Inoue, W. & 90. Akana, S. F. et al. Feedback sensitivity of the rat
Organization and regulation of paraventricular nucleus Bains, J. S. Glucocorticoid feedback uncovers hypothalamo-pituitary-adrenal axis and its capacity to
glutamate signaling systems: N‑methyl-d‑aspartate retrograde opioid signaling at hypothalamic synapses. adjust to exogenous corticosterone. Endocrinology
receptors. J. Comp. Neurol. 484, 43–56 (2005). Nat. Neurosci. 16, 596–604 (2013). 131, 585–594 (1992).
52. Margeta-Mitrovic, M., Mitrovic, I., Riley, R. C., This paper demonstrates that stress causes a 91. Flak, J. N., Ostrander, M. M., Tasker, J. G. &
Jan, L. Y. & Basbaum, A. I. Immunohistochemical presynaptic LTD that relies on CORT. Herman, J. P. Chronic stress-induced neurotransmitter
localization of GABAB receptors in the rat central 72. Pacak, K., Palkovits, M., Kopin, I. J. & Goldstein, D. S. plasticity in the PVN. J. Comp. Neurol. 517, 156–165
nervous system. J. Comp. Neurol. 405, 299–321 Stress-induced norepinephrine release in the (2009).
(1999). hypothalamic paraventricular nucleus and pituitary- 92. Kusek, M., Tokarski, K. & Hess, G. Repeated restraint
53. Ziegler, D. R. & Herman, J. P. Local integration of adrenocortical and sympathoadrenal activity: in vivo stress enhances glutamatergic transmission in the
glutamate signaling in the hypothalamic paraventricular microdialysis studies. Front. Neuroendocrinol. 16, paraventricular nucleus of the rat hypothalamus.
region: regulation of glucocorticoid stress responses. 89–150 (1995). J. Physiol. Pharmacol. 64, 565–570 (2013).
Endocrinology 141, 4801–4804 (2000). 73. Pacak, K. & Palkovits, M. Stressor specificity of central 93. Korosi, A. et al. Early-life experience reduces
54. Cole, R. L. & Sawchenko, P. E. Neurotransmitter neuroendocrine responses: implications for stress- excitation to stress-responsive hypothalamic neurons
regulation of cellular activation and neuropeptide related disorders. Endocr. Rev. 22, 502–548 (2001). and reprograms the expression of corticotropin-
gene expression in the paraventricular nucleus of the 74. Boudaba, C., Di, S. & Tasker, J. G. Presynaptic releasing hormone. J. Neurosci. 30, 703–713 (2010).
hypothalamus. J. Neurosci. 22, 959–969 (2002). noradrenergic regulation of glutamate inputs to 94. Levine, S. Maternal and environmental influences on
55. Darlington, D. N., Miyamoto, M., Keil, L. C. & hypothalamic magnocellular neurones. the adrenocortical response to stress in weanling rats.
Dallman, M. F. Paraventricular stimulation with J. Neuroendocrinol. 15, 803–810 (2003). Science 156, 258–260 (1967).
glutamate elicits bradycardia and pituitary responses. 75. Feldman, S. & Weidenfeld, J. Involvement of 95. Francis, D. D., Champagne, F. A., Liu, D. &
Am. J. Physiol. 256, R112–R119 (1989). endogeneous glutamate in the stimulatory effect of Meaney, M. J. Maternal care, gene expression, and
56. Hewitt, S. A., Wamsteeker, J. I., Kurz, E. U. & norepinephrine and serotonin on the hypothalamo- the development of individual differences in stress
Bains, J. S. Altered chloride homeostasis removes pituitary-adrenocortical axis. Neuroendocrinology 79, reactivity. Ann. NY Acad. Sci. 896, 66–84 (1999).
synaptic inhibitory constraint of the stress axis. Nat. 43–53 (2004). 96. Avishai-Eliner, S., Eghbal-Ahmadi, M., Tabachnik, E.,
Neurosci. 12, 438–443 (2009). 76. Steiner, M. A., Marsicano, G., Wotjak, C. T. & Lutz, B. Brunson, K. L. & Baram, T. Z. Down-regulation of
This was the first demonstration of changes in Conditional cannabinoid receptor type 1 mutants hypothalamic corticotropin-releasing hormone messenger
chloride gradients as a possible cause of reveal neuron subpopulation-specific effects on ribonucleic acid (mRNA) precedes early-life experience-
disinhibition during stress. behavioral and neuroendocrine stress responses. induced changes in hippocampal glucocorticoid receptor
57. Bartanusz, V. et al. Local γ-aminobutyric acid and Psychoneuroendocrinology 33, 1165–1170 (2008). mRNA. Endocrinology 142, 89–97 (2001).
glutamate circuit control of hypophyseotrophic 77. Steiner, M. A. & Wotjak, C. T. Role of the 97. Cullinan, W. E., Ziegler, D. R. & Herman, J. P.
corticotropin-releasing factor neuron activity in the endocannabinoid system in regulation of the Functional role of local GABAergic influences on the
paraventricular nucleus of the hypothalamus. Eur. hypothalamic-pituitary-adrenocortical axis. Prog. HPA axis. Brain Struct. Funct. 213, 63–72 (2008).
J. Neurosci. 19, 777–782 (2004). Brain Res. 170, 397–432 (2008). 98. Sarkar, J., Wakefield, S., MacKenzie, G., Moss, S. J. &
58. Herman, J. P. et al. Central mechanisms of stress 78. Lu, A. et al. Conditional mouse mutants highlight Maguire, J. Neurosteroidogenesis is required for the
integration: hierarchical circuitry controlling mechanisms of corticotropin-releasing hormone physiological response to stress: role of neurosteroid-
hypothalamo-pituitary-adrenocortical responsiveness. effects on stress-coping behavior. Mol. Psychiatry 13, sensitive GABAA receptors. J. Neurosci. 31,
Front. Neuroendocrinol. 24, 151–180 (2003). 1028–1042 (2008). 18198–18210 (2011).
59. Bali, B., Erdélyi, F., Szabó, G. & Kovács, K. J. 79. Wamsteeker, J. I., Kuzmiski, J. B. & Bains, J. S. This paper shows that stress causes a
Visualization of stress-responsive inhibitory circuits in Repeated stress impairs endocannabinoid signaling in post-translational modification of KCC2 to modify
the GAD65‑eGFP transgenic mice. Neurosci. Lett. the paraventricular nucleus of the hypothalamus. chloride gradients and also demonstrates a clear
380, 60–65 (2005). J. Neurosci. 30, 11188–11196 (2010). role for extrasynaptic GABAA receptors.

NATURE REVIEWS | NEUROSCIENCE VOLUME 16 | JULY 2015 | 387

© 2015 Macmillan Publishers Limited. All rights reserved


REVIEWS

99. Wamsteeker Cusulin, J. I., Füzesi, T., Watts, A. G. & 106. Deeb, T. Z., Lee, H. H. C., Walker, J. A., Davies, P. A. & amnesia, flashbulb and traumatic memories, and the
Bains, J. S. Characterization of corticotropin-releasing Moss, S. J. Hyperpolarizing GABAergic transmission Yerkes-Dodson law. Neural Plast. 2007, 60803
hormone neurons in the paraventricular nucleus of the depends on KCC2 function and membrane potential. (2007).
hypothalamus of Crh-IRES-Cre mutant mice. PLoS ONE Channels 5, 475–481 (2011). 112. Wamsteeker Cusulin, J. I., Senst, L., Teskey, G. C. &
8, e64943 (2013). 107. Ma, S. & Morilak, D. A. Chronic intermittent cold Bains, J. S. Experience salience gates
100. Verkuyl, J. M., Karst, H. & Joëls, M. GABAergic stress sensitises the hypothalamic-pituitary-adrenal endocannabinoid signaling at hypothalamic synapses.
transmission in the rat paraventricular nucleus of the response to a novel acute stress by enhancing J. Neurosci. 34, 6177–6181 (2014).
hypothalamus is suppressed by corticosterone and noradrenergic influence in the rat paraventricular This paper demonstrates that a novel stressor,
stress. Eur. J. Neurosci. 21, 113–121 (2005). nucleus. J. Neuroendocrinol. 17, translated through an increase in local synaptic
101. Inoue, W. & Bains, J. S. Beyond inhibition: GABA 761–769 (2005). activity, can reset synapses compromised by
synapses tune the neuroendocrine stress axis. 108. Morilak, D. A. et al. Role of brain norepinephrine in repeated homotypic stress.
Bioessays 36, 561–569 (2014). the behavioral response to stress. Prog. 113. Itoi, K. et al. Visualization of corticotropin-releasing
102. Lee, V., Sarkar, J. & Maguire, J. Loss of Gabrd in CRH Neuropsychopharmacol. Biol. Psychiatry 29, factor neurons by fluorescent proteins in the mouse
neurons blunts the corticosterone response to stress 1214–1224 (2005). brain and characterization of labeled neurons in the
and diminishes stress-related behaviors. 109. Weiser, M. J., Osterlund, C. & Spencer, R. L. Inhibitory paraventricular nucleus of the hypothalamus.
Psychoneuroendocrinology 41, 75–88 (2014). effects of corticosterone in the hypothalamic Endocrinology 155, 4054–4060 (2014).
103. Cullinan, W. E. & Wolfe, T. J. Chronic stress regulates paraventricular nucleus (PVN) on stress-induced 114. Dabrowska, J., Hazra, R., Guo, J.‑D., Dewitt, S. &
levels of mRNA transcripts encoding β subunits of the adrenocorticotrophic hormone secretion and gene Rainnie, D. G. Central CRF neurons are not created
GABAA receptor in the rat stress axis. Brain Res. 887, expression in the PVN and anterior pituitary. equal: phenotypic differences in CRF-containing
118–124 (2000). J. Neuroendocrinol. 23, 1231–1240 (2011). neurons of the rat paraventricular hypothalamus and
104. Verkuyl, J. M., Hemby, S. E. & Joëls, M. Chronic stress This paper provides evidence for CORT as an the bed nucleus of the stria terminalis. Front.
attenuates GABAergic inhibition and alters gene adaptogenic signal in the PVN. Neurosci. 7, 156 (2013).
expression of parvocellular neurons in rat 110. Abraham, W. C. Metaplasticity: tuning synapses and
hypothalamus. Eur. J. Neurosci. 20, 1665–1673 networks for plasticity. Nat. Rev. Neurosci. 9, 387 Acknowledgements
(2004). (2008). Research by the authors is supported by the Canadian
105. Staley, K. J. & Proctor, W. R. Modulation of 111. Diamond, D. M., Campbell, A. M., Park, C. R., Institutes for Health Research.
mammalian dendritic GABAA receptor function by the Halonen, J. & Zoladz, P. R. The temporal dynamics
kinetics of Cl– and HCO3– transport. J. Physiol. 519, model of emotional memory processing: a synthesis Competing interests statement
693–712 (1999). on the neurobiological basis of stress-induced The authors declare no competing interests.

388 | JULY 2015 | VOLUME 16 www.nature.com/reviews/neuro

© 2015 Macmillan Publishers Limited. All rights reserved

You might also like