Neuropharmacology: Bryony Laura Winters, Christopher Walter Vaughan

You might also like

You are on page 1of 23

Neuropharmacology 197 (2021) 108736

Contents lists available at ScienceDirect

Neuropharmacology
journal homepage: www.elsevier.com/locate/neuropharm

Invited review

Mechanisms of endocannabinoid control of synaptic plasticity


Bryony Laura Winters *, Christopher Walter Vaughan
Pain Management Research Institute, Kolling Institute of Medical Research, Northern Clinical School, University of Sydney at Royal North Shore Hospital, NSW, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: The endogenous cannabinoid transmitter system regulates synaptic transmission throughout the nervous system.
Cannabinoids Unlike conventional transmitters, specific stimuli induce synthesis of endocannabinoids (eCBs) in the post­
Endocannabinoids synaptic neuron, and these travel backwards to modulate presynaptic inputs. In doing so, eCBs can induce short-
Synaptic plasticity
term changes in synaptic strength and longer-term plasticity. While this eCB regulation is near ubiquitous, it
displays major regional and synapse specific variations with different synapse specific forms of short-versus long-
term plasticity throughout the brain. These differences are due to the plethora of pre- and postsynaptic mech­
anisms which have been implicated in eCB signalling, the intricacies of which are only just being realised. In this
review, we shall describe the current understanding and highlight new advances in this area, with a focus on the
retrograde action of eCBs at CB1 receptors (CB1Rs).
This article is part of the special Issue on ‘Cannabinoids’.

Constituents of the plant Cannabis sativa produce many of their well- analogues of phytocannabinoids which mimic the actions of plant
known effects by acting on the endogenous cannabinoid transmitter derived phytocannabinoids. These include THC analogues such as
system. Over the past 30 years it has been shown that this endocanna­ nabilone and CP55940, plus more recently characterised drugs of abuse
binoid system acts widely throughout the nervous system. This has shed such as XLR-11 and AB-CHMINACA (Pertwee, 2015; Wiley et al., 2017).
light on the role of the endogenous cannabinoid system in various pa­ The basic components of the endogenous cannabinoid system
thologies and has provided leads for novel therapeutic targets. This re­ include: endogenous molecules called endocannabinoids (eCBs),
view summarises the synaptic mechanisms by which the endogenous cellular proteins involved in their production, the receptors and chan­
cannabinoid system controls brain activity, information which has nels upon which they act, plus other proteins involved in their uptake
largely been obtained from the ex vivo brain slice preparation. Firstly, we and subsequent breakdown. Of the eCBs, N-arachidonoylethanolamide
briefly describe cannabinoids and aspects of their targets which are (anandamide) and 2-Arachidonoylglycerol (2-AG) are the best charac­
relevant to their synaptic actions. terised in terms of their production, targets, and degradation. Briefly,
these eCBs are largely produced ‘on demand’ from membrane bound
1. Cannabinoids and their targets phospholipids in the postsynaptic cell body via an, as yet to be identified
non-vesicular manner, or potentially from non-synaptic extracellular
Cannabinoids are an incredibly diverse range of plant derived, syn­ vesicles (Haj-Dahmane et al., 2018; Nakamura et al., 2019) (see also
thetic and endogenous chemicals which act on a diverse range of targets. section 4.1.1). Either way, eCB production differs fundamentally from
most neurotransmitters which are stored in presynaptic vesicles and
1.1. Cannabinoids and their targets released via Ca2+-dependent and independent mechanisms.
Many of the actions of these cannabinoids are mediated by a specific
The plant Cannabis sativa is well known for its recreational use, plus class of Gαi/o-coupled receptors, cannabinoid CB1Rs and CB2Rs, which
its potential therapeutic actions and side-effects. The plant derived were identified in the 1990s (Pertwee, 2015). CB1Rs are one of the most
constituents of cannabis include its main psychoactive constituent, Δ9- widely expressed Gαi/o-coupled class of G-protein coupled receptors
tetrahydrocannabinol (THC), the major non-psychoactive constituent, (GPCRs) within the nervous system, while CB2Rs are more restricted to
cannabidiol, plus over five hundred other cannabinoids, terpenes and the immune system. Thus, CB1Rs are ideally placed to modulate synaptic
other natural products. There are also many synthetic drugs and activity within the brain and thereby control a vast array of behavioural,

* Corresponding author.
E-mail address: bryony.winters@sydney.edu.au (B.L. Winters).

https://doi.org/10.1016/j.neuropharm.2021.108736
Received 29 March 2021; Received in revised form 27 July 2021; Accepted 28 July 2021
Available online 31 July 2021
0028-3908/© 2021 Elsevier Ltd. All rights reserved.
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

autonomic, and somatic functions. It should be noted that there is evi­ from membrane phospholipids in a Ca2+ dependent manner via
dence of cannabinoid CB2R expression within the nervous system, NAPE-phospholipase D (PLD), PLA2 and/or other enzymes (Fig. 1). By
particularly in pathological states (Jordan and Xi, 2019; Miller and Devi, contrast, 2-AG is formed by Ca2+-independent
2011). Thus, plant derived and synthetically produced cannabinoids, Gαq/11-GPCR/phospholipase C (PLCβ)-diacylglycerol lipase (DAGLα)
such as THC and CP55940, produce many of their effects by mimicking cascade and by a Ca2+-dependent PLC-DAGL pathway (Fig. 1). The ac­
the actions of eCBs at CB1Rs and/or CB2Rs. Natural, synthetic, and eCBs, tions of eCBs are terminated by their uptake via passive diffusion, or as
to a varying extent, also target other receptors and ion channels. These yet to be defined mechanisms (Fowler, 2013; Fu et al., 2011; Glaser
‘non-cannabinoid receptor’ targets include GPCRs such as GPR55, et al., 2003; Kaczocha et al., 2009), such as reuptake via a putative
ligand gated ion channels such TRPV1, 5HT3R, GlyRs and GABAARs, eCB-transporter (discussed further in section 4.1.2). Once within the
plus sodium and calcium voltage-gated ion channels (De Petrocellis cell, eCB degradation is catalysed by the enzymes, fatty acid amide
et al., 2017; Morales and Jagerovic, 2016; Muller et al., 2018). hydrolase (FAAH), monoacylglycerol lipase (MAGL), and α/β-Hydrolase
domain-containing 6/12 (ABHD6/12) (Blankman and Cravatt, 2013;
Fowler, 2012). While anandamide is largely degraded by FAAH, 2-AG is
1.2. eCB production and degradation degraded by MAGL and to a lesser extent ABHD6/12. It is also important
to note that while FAAH is mainly located postsynaptically, MAGL is
The eCBs anandamide and 2-AG are produced from membrane expressed in presynaptic nerve terminals. There are several recently
phospholipids by distinct stimuli and biosynthetic pathways (Cascio and characterised agents which inhibit the breakdown of eCBs by FAAH and
Marini, 2015; Katona and Freund, 2012; Pertwee, 2015; Piomelli, 2003; MAGL. In doing so, these drugs can enhance the ‘natural’ activation of
Piscitelli and Di Marzo, 2012). Anandamide is thought to be formed

Fig. 1. Endocannabinoids act as retrograde transmitters to transiently reduce the probability of neurotransmitter release. Simplified schematic depicting
the pathways involved in short-term retrograde endocannabinoid signalling. The endocannabinoid 2-arachidonylglycerol (2-AG) is generated by the conversion of
diacylglycerol into 2-AG via diacylglycerol lipase α (DAGLα). This DAGLα induced production of 2-AG can be triggered by physiological stimuli: (1) neuronal
depolarisation induced Ca2+ influx via voltage gated Ca2+ channels (VGCCs) via an unknown mechanism (2) activation of postsynaptic Gαq/11-GPCRs (e.g. group I
mGluRs, M1/3 mAChRs) by transmitters stimulates phospholipase C (PLCβ) in a Ca2+-independent manner; (3) low level Gαq/11-GPCR activation of PLCβ is enhanced
by lesser increases in Ca2+. Although its physiological origin is less clear, (4) neuronal depolarisation induced Ca2+ influx can also lead to the production of
anandamide (AEA) via enzymes such as NAPE-phospholipase D and/or phospholipase A2. (5) Following their de novo synthesis, 2-AG and AEA are then released into
the synapse, via as yet to be confirmed mechanisms. (6) These endocannabinoids ‘travel backwards’ to activate cannabinoid CB1 Gαi/o-GPCRs located on presynaptic
terminals. (7) This initiates Gβγ-subunit inhibition of presynaptic VGCCs which leads to a transient reduction in the release of neurotransmitter onto the postsynaptic
neuron which generates endocannabinoids. (8) Endocannabinoids are taken back up into neuronal and glial cells, possibly by a selective carrier-mediated transporter.
2-AG is taken up into presynaptic terminals and degraded by MAG-lipase (MAGL), while anandamide is taken up into the postsynaptic neuron and glia, then degraded
by fatty acid amide hydrolase (FAAH).

2
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

the eCB system, as opposed to receptor agonists. Together, the differ­ 2.2. Endocannabinoids
ential location and targets of these enzymes has consequences for the
actions of these eCBs and the therapeutic potential of FAAH and MAGL Numerous studies have demonstrated that exogenous application of
blockers (Piscitelli and Di Marzo, 2012). the eCBs, anandamide, and 2-AG, presynaptically inhibits synaptic
transmission within a number of regions throughout the brain and spinal
2. Short-term presynaptic actions of exogenously applied cord. However, there are regional and synapse specific variations. In
cannabinoids some brain regions anandamide inhibits GABAgeric, but not gluta­
matergic synaptic transmission while synthetic agonists inhibit both
Many neurotransmitters exert their physiological effects by acting on GABAergic and glutamatergic synaptic transmission (Adermark and
pre- and postsynaptic neuronal elements, and in some cases non- Lovinger, 2007b; Hentges, 2007; Hentges et al., 2005; Jennings et al.,
neuronal elements within the nervous system. This is also the case for 2001, 2003; Kawahara et al., 2011; Lau et al., 2014). These differences
cannabinoids, although the cannabinoid CB1Rs actions are largely are likely to be due to several factors, in addition to the level of CB1R
presynaptic. expression between excitatory and inhibitory synapses described above
(Shen et al., 1996). Firstly, synthetic cannabinoid agonists generally
have greater efficacy and potency at cannabinoid receptors than eCBs.
2.1. Phytocannabinoids and synthetic cannabinoids Secondly, eCBs undergo uptake and breakdown which greatly limits
their access to the synapse following exogenous administration. Thus,
There is a wealth of anatomical evidence that cannabinoid CB1Rs are the anandamide and 2-AG induced inhibition of synaptic transmission is
largely located on presynaptic nerve terminals throughout the brain enhanced by blocking their degradation with FAAH and MAGL in­
(Ohno-Shosaku and Kano, 2014). Thus, numerous electrophysiological hibitors, respectively (Bajo et al., 2009; Kawahara et al., 2011; Lau et al.,
studies have reported that THC and several synthetic non-selective 2014; Lee et al., 2015; Vaughan et al., 2000; Yasmin et al., 2020).
cannabinoid agonists produce CB1R mediated inhibition of GABAergic Another potential factor is that anandamide acts at other presynaptic
and glutamatergic synaptic transmission via a presynaptic mechanism targets within the brain, such as the non-selective cation channel,
(Alger, 2002; Araque et al., 2017; Hoffman et al., 2017; Kano et al., TRPV1 (Marinelli et al., 2002, 2003), plus a range of postsynaptic ele­
2009; Laaris et al., 2010; Schlicker and Kathmann, 2001; Shen and ments (section 1.1). Thus, TRPV1 mediated enhancement of synaptic
Thayer, 1999) (Fig. 1). In several brain regions, however, cannabinoid transmission functionally opposes CB1R mediated inhibition (Bhaskaran
agonists inhibit GABAergic synaptic transmission with greater poten­ and Smith, 2010; Kawahara et al., 2011; Lee et al., 2015; Morisset et al.,
cy/efficacy than glutamatergic transmission. This has been attributed to 2001). In addition to these factors, anandamide enhances presynaptic
the observation that GABAergic synapses express considerably higher excitability via intracellular cascades involving PKA and IP3 (Hofmann
levels of CB1Rs at their presynaptic terminals than their glutamatergic et al., 2011; Sang et al., 2010).
counterparts (Katona et al., 2001; Marsicano and Lutz, 1999; Penasco
et al., 2019; Uchigashima et al., 2007) although see (Katona et al., 2.3. Tonic presynaptic endocannabinoid control
2006). It should be emphasised that cannabinoid receptors are also
expressed in non-neuronal elements where they have complex actions; In addition to their actions following exogenous administration,
this is addressed in another review in this issue. tonically released eCBs exert a basal presynaptic CB1R mediated sup­
The presynaptic nature of cannabinoid inhibition has been delin­ pression of GABAergic, and to a lesser extent, glutamatergic synaptic
eated by using a variety of electrophysiological techniques such as the transmission in several brain regions. In these studies, it has been
examination of paired evoked synaptic currents, quantal TTX-resistant demonstrated that CB1R antagonists, such as rimonabant (SR141716)
synaptic currents, and presynaptic calcium imaging. It has generally and AM251, produce an increase in evoked and spontaneous TTX-
been assumed that voltage-gated calcium channels (VGCCs) have a resistant synaptic currents in the slice preparation, and even in iso­
crucial role in the modulation of synaptic transmission by cannabinoids lated neurons with intact presynaptic boutons (Ferraro et al., 2020;
because Ca2+ entry into the nerve terminal is required for presynaptic Hentges et al., 2005; Huang and Woolley, 2012; Lau et al., 2014;
vesicle fusion and the subsequent release of neurotransmitter (see Losonczy et al., 2003; Manz et al., 2020; Marcus et al., 2020; Melis et al.,
Gandini and Zamponi, 2021; Zamponi and Currie, 2013 for compre­ 2004a; Neu et al., 2007; Oliet et al., 2007; Song et al., 2015; Yasmin
hensive reviews on GPCR modulation of VGCCs). Accordingly, CB1R et al., 2020; Zhu and Lovinger, 2005). This tonic suppression is likely to
activation inhibits VGCCs and reduces the influx of Ca2+ into presyn­ be due to postsynaptic release of eCBs because it is blocked by chelating
aptic nerve terminals (Daniel and Crepel, 2001; Kreitzer and Regehr, postsynaptic Ca2+ with BAPTA (Hentges et al., 2005; Losonczy et al.,
2001; Twitchell et al., 1997) (Fig. 1). In addition, cannabinoid 2003; Manz et al., 2020; Neu et al., 2007; Oliet et al., 2007; Zhu and
agonist-induced inhibition of synaptic transmission is reduced by Lovinger, 2005). It is also likely that the tonic suppression is related to
blockade of N- and P/Q-type VGCCs (Hoffman and Lupica, 2000; Huang the basal excitability of neurons because it can be enhanced by
et al., 2003; Sullivan, 1999; Varma et al., 2002). Indeed, it has been increasing neuronal firing (Drew et al., 2008; Lee et al., 2010; Oliet
shown that presynaptic CB1R induced inhibition via VGCCs is mediated et al., 2007; Song et al., 2015). Thus, the level of the basal ‘endo­
via the Gβγ, rather than the Gαi/o subunit (Jensen et al., 2021; Roloff and cannabinoid tone’ maybe be reflected by the overall level of neuronal
Thayer, 2009) (Fig. 1). While it has been shown that cannabinoid re­ excitability in different brain regions and experimental conditions.
ceptors couple more directly to presynaptic release mechanisms, there Interpretation of these tonic eCB experiments is, however, compli­
are several other less direct presynaptic mechanisms (Robbe et al., cated because cannabinoid ligands such as rimonabant and AM251 are
2003). It has been demonstrated that presynaptic 4-AP sensitive inverse agonists. Indeed, a role for constitutive CB1R activation has been
K+-channels have a role in cannabinoid CB1R induced inhibition (Daniel demonstrated in the hippocampus because while basal GABAergic syn­
and Crepel, 2001; Robbe et al., 2003; Varma et al., 2002). Mitochondrial aptic transmission is enhanced by AM251, it is unaffected by the neutral
CB1Rs have a role in cannabinoid inhibition of glutamatergic synaptic antagonist NESS0327 (Jensen et al., 2021; Lee et al., 2015). Recently, it
transmission in the hippocampus via soluble adenylyl cyclase-protein has been shown that this constitutive activity is mediated by the Gβγ
kinase A (PKA) control of mitochondrial energetic activity (Hebert-Ch­ subunit of CB1Rs (Jensen et al., 2021). By contrast, in some regions, such
atelain et al., 2016). Finally, while the probability of neurotransmitter as the midbrain PAG and prefrontal cortex both inverse CB1R agonists
release can also be controlled by the size of the ready releasable pool of and neutral CB1R antagonists enhance synaptic transmission, indicating
vesicles, there are conflicting reports as to whether this is modulated by that tonic inhibition is due to the release of eCBs which are acting on
cannabinoids (García-Morales et al., 2015; Sullivan, 1999). CB1Rs (Lau et al., 2014; Marcus et al., 2020).

3
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

The activity of the eCBs is tightly regulated by their uptake and evoked by minimal stimulation, (iii) a reduction in the frequency,
degradation via FAAH and MAGL (Blankman and Cravatt, 2013; Di but not amplitude of spontaneous TTX-resistant, or evoked Sr2+-
Marzo, 2018; Fowler et al., 2017). A number of studies have demon­ mediated quantal synaptic currents, but (iv) does not affect the
strated that tonic suppression of synaptic transmission can be unmasked, current induced by exogenous application of the modulated trans­
or enhanced by blocking the metabolism of eCBs, using selective in­ mitter, GABA or glutamate. Thus, postsynaptic depolarisation pro­
hibitors of FAAH (e.g. URB597, PF3845) and MAGL (e.g. JZL184), or duces a presynaptic reduction in the probability of release of that
genetic deletion of these enzymes (Ferraro et al., 2020; Kawahara et al., neurotransmitter from nerve terminals impinging upon that neuron.
2011; Lau et al., 2014; Lee et al., 2010, 2015; Liu et al., 2016; Yasmin C. Mediated by eCBs: The depolarisation induced reduction in evoked
et al., 2020). However, there are substantial regional and synapse spe­ synaptic currents is (i) abolished by CB1R antagonism and knockout,
cific variations in the relative roles of these enzymes, in addition to (ii) unaffected by blocking vesicular endocytosis with postsynaptic
gender and species related differences. In some cases, dual blockade of botulinum neurotoxin (Wilson and Nicoll, 2001), and (iii) is abol­
both FAAH and MAGL (e.g. with JZL195) leads to a greater increase in ished by disrupting postsynaptic signalling pathways involved in the
tonic inhibition of synaptic transmission than individual blockade of production of eCBs (see below). Thus, retrograde eCB signalling in­
FAAH, or MAGL (Fucich et al., 2020; Lau et al., 2014). This indicates volves ‘on-demand’ postsynaptic production and non-vesicular
that 2-AG and anandamide are tonically released, and that their actions release of eCBs.
can be enhanced by overcoming their uptake and breakdown with
FAAH/MAGL inhibitors. It should also be noted that, in some regions, Since these initial studies, depolarisation induced retrograde sig­
MAGL has been reported to make a greater contribution to tonic sup­ nalling has been detected in numerous brain regions. Some of these
pression than FAAH, and this has been attributed to functionally regions include the amygdala (Kamprath et al., 2011; Kodirov et al.,
opposing pre- and postsynaptic actions of anandamide on TRPV1 2010; Zhu and Lovinger, 2005), dorsal raphe nucleus (Haj-Dahmane and
channels (Kawahara et al., 2011; Lee et al., 2015). Shen, 2009), cortex (Bodor et al., 2005; Fortin et al., 2004; Marcus et al.,
2020; Terral et al., 2020), habenula (Vickstrom et al., 2020), hypo­
3. Short-term synaptic plasticity thalamus (Colmers and Bains, 2018; Hirasawa et al., 2004; Jo et al.,
2005), striatum (Uchigashima et al., 2007), substantia nigra (Wall­
It is now well established that eCBs act as retrograde messengers michrath and Szabo, 2002; Yanovsky et al., 2003), and ventral
throughout the brain. Unlike most neurotransmitters, they are produced tegmental area (Melis et al., 2004b).
‘on-demand’ in the postsynaptic neuron and travel backwards to control
the release of neurotransmitters onto that neuron. Generally, this short- 3.1.2. G-protein coupled receptors
term depression is expressed by activation of presynaptic CB1Rs. Following on from the phenomenon of DSI/DSE, it was shown that
Following is a summary of the mechanisms underlying short-term eCB short-term retrograde eCB plasticity could also be induced by activation
induced plasticity. of postsynaptic Gαq/11-coupled group 1 metabotropic glutamate
(mGluR) and M1/M3 Gαq/11-coupled cholinergic muscarinic receptors
3.1. The triggers for retrograde eCB signalling (M1/M3 mAChRs) in the hippocampus and cerebellum (Kim et al., 2002;
Maejima et al., 2001; Varma et al., 2001) (Fig. 1). A parallel experi­
Retrograde eCB signalling can be induced by the engagement of a mental approach to that described above for depolarisation-induced
range of distinct postsynaptic voltage/ligand gated ion channels and suppression demonstrated Gαq/11-GPCR induced suppression was
receptors. These are discussed below. mediated by retrograde eCB signalling. The postsynaptic locus of in­
duction was demonstrated by the observation that postsynaptic GDPβS
3.1.1. Neuronal depolarisation abolished the inhibition produced by group 1 mGluR and mAChR ago­
The first evidence for a physiological role for eCBs was based on the nists, but not that produced by cannabinoid agonists which act directly
phenomenon of ‘depolarisation-induced suppression of inhibition’ on presynaptic CB1Rs.
(DSI), previously observed in the hippocampus and cerebellum (Fig. 1). Since these initial studies in the hippocampus and cerebellum, group
In these intriguing studies, brief (seconds) postsynaptic depolarisation 1 mGluR-induced, retrograde eCB signalling has been shown to occur at
produced a short-lasting inhibition of transmitter release onto that numerous GABAergic and glutamatergic synapses throughout the brain.
neuron, observed as inhibition of electrically evoked GABAergic inhib­ Some of these regions include the amygdala (Sheinin et al., 2008; Zhu
itory postsynaptic currents (IPSCs) (Llano et al., 1991; Pitler and Alger, and Lovinger, 2005), calyx of Held (Kushmerick et al., 2004), cortex
1992). It was not until a decade later that eCBs were identified as the (Kiritoshi et al., 2016), hypothalamus (Iremonger et al., 2011), midbrain
retrograde signalling agent involved in DSI, and the parallel phenome­ periaqueductal grey (Drew et al., 2008), striatum (Centonze et al., 2007;
non of ‘depolarisation-induced suppression of excitation’ (DSE, for Kreitzer and Malenka, 2005; Narushima et al., 2006), substantia nigra
glutamatergic excitatory postsynaptic currents, EPSCs), both in brain (Freestone et al., 2014), and ventral tegmental area (Melis et al., 2004a).
slice and cultured neuron preparations (Diana et al., 2002; Kreitzer and It should be noted that the relative role of the postsynaptic group 1 re­
Regehr, 2001; Ohno-Shosaku et al., 2001; Wilson and Nicoll, 2001). ceptor subtypes, mGluR1 and mGluR5, in retrograde eCB signalling
Briefly, the key observations which identified eCBs as retrograde varies throughout the brain. In addition to this, agonist induced acti­
signalling agents in this form of short-term plasticity are: vation of postsynaptic M1/M3 Gαq/11-coupled mAChRs has also been
shown to induced retrograde eCB inhibition of synaptic transmission in
A. Postsynaptic induction: The depolarisation induced reduction in several brain regions (Edwards et al., 2006; Freestone et al., 2015;
evoked synaptic currents is abolished by rapid buffering of post­ Fukudome et al., 2004; Lau and Vaughan, 2008; Martin et al., 2015;
synaptic Ca2+ with BAPTA, while that produced by cannabinoid Narushima et al., 2006; Neuhofer et al., 2018).
CB1R agonists is unaffected by postsynaptic BAPTA. Furthermore, Activation of the retrograde eCB signalling system is not exclusive to
the depolarisation induced inhibition is associated with an increase metabotropic glutamate and cholinergic receptors. Several other classes
in postsynaptic Ca2+ and this is abolished by subtype selective VGCC of neurotransmitter-GPCR systems have been shown to inhibit
blockers. Thus, DSI and DSE are generated by a depolarisation GABAergic and glutamatergic synaptic transmission via retrograde eCB
induced postsynaptic Ca2+ influx via VGCCs. signalling. Some of these transmitter-GPCR systems include cholecys­
B. Presynaptic expression: Depolarisation produces (i) an increase in the tokinin-CCK1R (Földy et al., 2007; Mitchell et al., 2011), dopamine-D2R
paired pulse ratio and co-efficient of variation of electrically evoked (Yin and Lovinger, 2006), neurotensin-NT1/2R (Kortleven et al., 2012;
synaptic currents, (ii) an increase in failure rate of synaptic currents Mitchell et al., 2009; Yin et al., 2008), orexin-OX1/2R (Haj-Dahmane and

4
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

Shen, 2005; Ho et al., 2011; Kargar et al., 2018; Tung et al., 2016), presynaptic sensitivity is a major determinant of depolarisation-induced
oxytocin-OTR (Hirasawa et al., 2004; Oliet et al., 2007), seroto­ suppression at some synapses (Ohno-Shosaku et al., 2002b), DSI/DSE
nin-5HT2R (Best and Regehr, 2008), and substance P-NK1/2R (Drew are weak, or absent in some cannabinoid-sensitive synapses (Drew et al.,
et al., 2009). However, several lines of evidence indicate that some of 2008; Engler et al., 2006; Freiman et al., 2006; Hentges et al., 2005;
these neurotransmitters do not directly engage the eCB system by their Kreitzer and Malenka, 2005; Sheinin et al., 2008). This heterogeneity
respective postsynaptic receptors. Firstly, the cholecystokinin, neuro­ may be more widespread than currently thought because the use of gross
tensin, and substance P induced inhibition of evoked synaptic currents is electrical stimulation to examine synaptic transmission can mask syn­
abolished, not only by cannabinoid receptor antagonists, but also by apse specific differences. Thus, the use of minimal stimulation, paired
group 1 mGluR antagonists in the striatum and midbrain periaqueductal recordings, and optogenetic approaches to activate specific synaptic
grey (Drew et al., 2009; Mitchell et al., 2009, 2011; Yin et al., 2018). inputs has revealed an increasing diversity in the actions of endoge­
Secondly, while these transmitters produce an increase in the nously released cannabinoids (Colmers and Bains, 2018; Deroche et al.,
paired-pulse ratio of evoked synaptic currents, they do not inhibit the 2020; Lines et al., 2017; Wilson and Nicoll, 2001). In addition to
rate of spontaneous (Ca2+-dependent and -independent) miniature ‘normal’ variations, it might be noted that the presence and extent of
synaptic currents. This suggests that these peptides act upstream to the retrograde eCB signalling can be altered in pathological states (Di et al.,
cannabinoid-sensitive synapse (Drew et al., 2009; Mitchell et al., 2009, 2013; Kiritoshi et al., 2016; Marcus et al., 2020).
2011). Together, these observations indicate that these neurotransmit­ Most of the above studies have used stimuli which are less physio­
ters increase the release of glutamate, which then engages postsynaptic logically relevant to engage retrograde eCB signalling, such as prolonged
mGluR-induced retrograde eCB signalling. depolarisation and application of synthetic GPCR/LGIC agonists. How­
Other classes of receptors can also drive eCB retrograde inhibition. ever, several studies have shown that eCB signalling can be driven by
For example, the inhibition of synaptic transmission induced by BDNF physiologically relevant stimuli. Depolarisation induced retrograde
(via the tyrosine kinase TrkB receptor), glucocorticoids and estrogen signalling can be induced by postsynaptic action potential firing at
(via glucocorticoid and E-alpha nuclear receptors) is mediated by physiological rates, at least in some brain regions (Dubruc et al., 2013;
retrograde eCB signalling (Di et al., 2003; Huang and Woolley, 2012; Fortin et al., 2004). Likewise, group 1 mGluR signalling can be induced
Tabatadze et al., 2015; Wu et al., 2020; Yeh et al., 2017). As for some of in some brain regions by stimulation of glutamatergic afferent inputs at
the above Gαq/11-coupled receptors, however, the inhibition of synaptic physiologically relevant rates (Brown et al., 2003; Chen et al., 2016;
transmission by estrogen has been shown to be via mobilisation of Galante and Diana, 2004; Maejima et al., 2001; Marcaggi and Attwell,
postsynaptic mGluR receptors (Huang and Woolley, 2012). 2005; Melis et al., 2004a; Yin and Lovinger, 2006).

3.1.3. G-protein coupled receptor and depolarisation induced interactions 3.2. Postsynaptic signalling mechanisms involved in short-term plasticity
In addition to their individual actions, there are interactions between
Gαq/11-receptor and depolarisation induced retrograde eCB inhibition. The postsynaptic mechanisms involved in short-term retrograde eCB
In the hippocampus, sub-threshold concentrations of mGluR1/5 and M1/3 signalling have been driven by our understanding of the pathways
mAChR agonists potentiate DSI and DSE (Hoffman et al., 2003; Kim involved in their production and degradation (section 1.1). Kano and
et al., 2002; Ohno-Shosaku et al., 2002a, 2003; Varma et al., 2001). colleagues have categorised short-term eCB plasticity induced by the
Gαq/11-induced interactions have since been observed for mGluR, above triggers into three basic signalling modes: (i) Ca2+-dependent
mAChR and D1Rs in the amygdala, substantia nigra and ventral signalling which is induced by depolarisation, (ii) Ca2+-independent
tegmental area (Kiritoshi et al., 2016; Tong et al., 2017; Yanovsky et al., signalling which is induced by Gαq/11-coupled GPCRs, and (iii) Ca2+-
2003). There are some interesting drug targets which can be used to assisted signalling which is induced by an interaction between depo­
enhance the natural engagement of the Gαq/11-receptor eCB system and larisation and Gαq/11-coupled GPCRs (Fig. 1) (Araque et al., 2017;
its enhancement of depolarisation-induced signalling. For example, in Castillo et al., 2012; Kano et al., 2009; Katona and Freund, 2012;
some regions mGluR5 negative allosteric modulators reduce DSI/DSE, Ohno-Shosaku and Kano, 2014). These are described below.
indicating that there is tonic mGluR-driven eCB signalling (Kiritoshi
et al., 2013; Straiker et al., 2018). Thus, besides directly acting agonists, 3.2.1. Ca2+-dependent eCB signalling
DSI and DSE can be enhanced by an mGluR5 positive allosteric modu­ The postsynaptic signalling pathways underlying depolarisation-
lator, or by knockout of SAP90/PSD-95-associated proteins, both of induced retrograde inhibition have only been partly resolved.
which enhance mGluR5 expression/signalling (Chen et al., 2011; Kir­ Depolarisation-induced retrograde eCB signalling requires a large in­
itoshi et al., 2016) see also section 4.1.1). crease in postsynaptic Ca2+ to micromolar levels (Fig. 1). Thus, both DSI
and DSE can be mimicked by uncaging postsynaptic Ca2+ and abolished
3.1.4. Ligand gated ion channels (LGIC) by chelating postsynaptic Ca2+ with BAPTA (Kim et al., 2002; Kreitzer
In addition to depolarisation induced Ca2+ increases via VGCCs, it and Regehr, 2001; Ohno-Shosaku et al., 2001; Pitler and Alger, 1992;
has also been shown that eCB signalling can be induced by a range of Trettel and Levine, 2003; Wilson and Nicoll, 2001). The main source of
Ca2+-permeable ligand gated ion channels. These include ionotropic postsynaptic Ca2+ for DSI and DSE is entry via L-type VGCCs triggered
glutamate receptors, such as NMDA receptors (NMDAR), kainite re­ by neuronal depolarisation, although release from intracellular Ca2+
ceptors and Ca2+-permeable AMPA receptors (Lourenço et al., 2011; stores has also been reported to have a role in some regions (Isokawa and
Manz et al., 2020; Marshall et al., 2018; Ohno-Shosaku et al., 2007) Alger, 2005; Straiker and Mackie, 2005) (Fig. 1).
(Fig. 1). Furthermore, activation of TRPV1 has been shown to induce The eCB involved in depolarisation-induced retrograde eCB signal­
retrograde eCB inhibition of synaptic transmission in the periaqueductal ling is likely to be 2-AG because it is prolonged by blockade and
grey (Liao et al., 2011). However, this is indirectly mediated by TRPV1 knockout of MAGL, but not FAAH in numerous brain regions (Hashi­
enhancement of glutamate release, activation of postsynaptic mGluRs motodani et al., 2007b; Kiritoshi et al., 2016; Makara et al., 2005;
and retrograde CB1R inhibition of GABAergic synaptic transmission, as Marcus et al., 2020; Straiker and Mackie, 2005; Szabo et al., 2006;
described above for cholecystokinin. Vickstrom et al., 2020). While native expression of FAAH does appear to
have a role DSI and DSE, it has been shown that overexpression of FAAH
3.1.5. Retrograde eCB signalling is heterogeneous and is engaged by can influence hippocampal DSE in cultured neurons (Straiker et al.,
physiological stimuli 2011; Zimmermann et al., 2019). There is also some evidence of a role
While retrograde eCB signalling is observed in numerous brain re­ for COX-2 and ABHD6 in the control of DSI and DSE (Kim and Alger,
gions, it is not an ubiquitous, homogenous phenomenon. While 2004; Kiritoshi et al., 2016; Straiker et al., 2009; Zhong et al., 2011).

5
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

While 2-AG can be formed by the PLC-DAGL cascade, the specific demonstrated that both PLCβ and DAGLα are involved in Gαq/11-induced
signalling cascade underlying depolarisation-induced production of 2- retrograde eCB signalling in several brain regions (Chen et al., 2016;
AG remains unclear (Fig. 1). DSI and DSE are unaffected by the PLC Hashimotodani et al., 2013; Ho et al., 2011; Kargar et al., 2018; Kortl­
inhibitors U73122 and ET18 (Brenowitz and Regehr, 2003; Chevaleyre even et al., 2012; Lau and Vaughan, 2008; Maejima et al., 2005; Melis
and Castillo, 2003; Edwards et al., 2006; Hashimotodani et al., 2008; et al., 2004a; Selvam et al., 2018; Sheinin et al., 2008; Tanimura et al.,
Szabo et al., 2006), or by knockout of PLCβ1/4 and PLCδ1/3/4 2010; Tung et al., 2016; Wu et al., 2020; Zhong et al., 2011). Together,
(Hashimotodani et al., 2005; Maejima et al., 2005). By contrast, DSI and these studies indicate that Gαq/11-receptor induced retrograde eCB sig­
DSE are abolished by the DAGL inhibitors tetrahydrolipstatin and nalling is mediated by Ca2+-independent production of 2-AG via a
RHC80267 (Kano et al., 2009), and this has been confirmed more PLCβ-DAGLα pathway.
recently with a more selective DAGL inhibitor, DO34, and knockout of
DAGLα (Gao et al., 2010; Hashimotodani et al., 2013; Marcus et al., 3.2.3. Ca2+-assisted eCB signalling
2020; Tong et al., 2017; Vickstrom et al., 2020; Wang et al., 2012; The postsynaptic signalling pathways underlying Gαq/11-receptor
Yoshino et al., 2011). Furthermore, the threshold for induced enhancement of depolarisation induced retrograde inhibition of
depolarisation-induced suppression is correlated to the level of DAGL synaptic transmission have been characterised (Fig. 1). The mechanisms
expression (Yoshida et al., 2011). Together, these observations indicate underlying this synergistic interaction slightly differ to those involved in
that DSI and DSE are mediated by an unknown PLCβ-independent Gαq/11-induced eCB signalling. Like Gαq/11-receptor induced eCB sig­
pathway which involves DAGLα dependent production of 2-AG. There nalling, the Gαq/11-induced enhancement of DSI/DSE involves the PLCβ-
are potential alternative pathways for depolarisation-induced eCB sig­ DAGLα cascade as it is abolished by their knockout (Hashimotodani
nalling. It has recently been shown that activation of PLCε by the ex­ et al., 2005; Maejima et al., 2005). Unlike Gαq/11-receptor induced eCB
change protein directly activated by cAMP (Epac) induces DSI in the signalling, the Gαq/11-induced enhancement of DSI/DSE is a
VTA (Tong et al., 2017). It has also been shown that DSE is abolished by Ca2+-dependent process. This has been demonstrated by the correlation
knockout of cytosolic PLA2 and DAGL in the cerebellum (Wang et al., between the degree of enhancement with the intracellular level of
2012). However, the role of the enzymes, PLCε and PLA2, remains to be free-Ca2+, and the role of PLCβ by knockout. Thus, PLCβ acts to integrate
verified. different forms of eCB signalling, including those triggered by
In addition to direct postsynaptic neuronal depolarisation, several Gαq/11-coupled receptors and neuronal depolarisation. Finally, other
ligand-gated ion channels have been shown to induce retrograde eCB interactions between eCB signalling systems have been observed. For
signalling (section 3.1.3). Indeed, kainite, NMDA and Ca2+-permeable example, Gαi/o-coupled D2R actions enhance DSI and DSE (Melis et al.,
AMPA receptor-induced retrograde eCB signalling, which is Ca2+- 2004b). It remains to be seen whether other neurotransmitters systems
dependent because it is blocked by postsynaptic BAPTA (Lourenço et al., interact with eCBs.
2011; Manz et al., 2020; Marshall et al., 2018; Ohno-Shosaku et al.,
2007) (Fig. 1). There are, however, some differences. In two of these 3.3. Termination of short-term plasticity
studies, the kainite and Ca2+-permeable AMPA receptor induced inhi­
bition appears to be mediated by anandamide, rather than 2-AG because Short-term retrograde endocannabinoid signalling is temporally
it is prolonged by URB597, but not by JZL184, and is unaffected by restricted and spatially confined to the postsynaptic neurons where they
DAGL inhibitors (Lourenço et al., 2011; Manz et al., 2020). In the other are produced, and in some cases, adjacent synapses (Galante and Diana,
study, kainite and NMDA receptor induced inhibition involves DAGL 2004; Kreitzer et al., 2002; Maejima et al., 2001; Wilson and Nicoll,
(Marshall et al., 2018; Ohno-Shosaku et al., 2007). The relative roles of 2001). This temporal and spatial restriction is influenced not only by the
anandamide and 2-AG in these forms of retrograde signalling, and their fate of endocannabinoids, but also by that of the neurotransmitters
signalling pathways remain to be verified. driving their production.

3.2.2. Ca2+-independent eCB signalling 3.3.1. eCB uptake and degradation


The postsynaptic signalling pathways underlying Gαq/11-receptor Basal and stimulus-evoked eCB signalling are controlled by the lipid
induced retrograde inhibition of synaptic transmission have been well nature of eCBs which limits their spread in the extracellular environ­
characterised in numerous brain regions (Fig. 1). Firstly, mGluR1/5 and ment, and by the uptake and breakdown of eCBs which limits their
M1/3 mAChR induced inhibition is G-protein mediated because it is temporal and spatial activity (Fig. 1). As noted above, degradation via
abolished by postsynaptic GDPβS in the hippocampus and cerebellum both enzymes FAAH and MAGL limits basal eCB inhibition in some brain
(Galante and Diana, 2004; Kim et al., 2002; Maejima et al., 2001, 2005; regions (section 2.3). By contrast, degradation via the enzyme MAGL
Neu et al., 2007). These studies have also shown that mGluR/mAChR limits depolarisation and Gαq/11-receptor induced retrograde eCB inhi­
induced inhibition is Ca2+-independent because it is unaffected by bition, both temporally and spatially (sections 3.2.1 and 3.2.2). This
postsynaptic BAPTA. This coupling has also been observed for retro­ breakdown has a crucial physiological role because MAGL inhibition
grade eCB signalling in other brain regions, and for a range of and knockout greatly enhance mGluR-mediated retrograde inhibition
Gαq/11-coupled GPCRs (for reference, see studies described in section elicited by ‘physiological’ stimulation evoked glutamate release (Chen
3.1.2). It might be noted, however, that Ca2+-dependent coupling has et al., 2016; Kiritoshi et al., 2016; Zhong et al., 2011). Using cell-specific
been observed in some preparations which involves the release of Ca2+ knockout, it has been shown that both neuronal and glial MAGL control
from intracellular stores (Melis et al., 2004a; Robbe et al., 2002; Straiker retrograde eCB signalling within both the cerebellum and hippocampus,
and Mackie, 2005). although their relative contributions vary with different synapses (Chen
Like DSI and the DSE, the eCB involved in Gαq/11-coupled GPCR et al., 2016; Liu et al., 2016; Viader et al., 2015). Interestingly, the up­
induced retrograde inhibition is also likely to be 2-AG because Gαq/11 take and degradation of eCBs can be altered in pathological states which
agonist induced inhibition is prolonged by blockade and knockout of affect glia (and consequently their uptake of endocannabinoids) and
MAGL, but not FAAH (Kiritoshi et al., 2016; Selvam et al., 2018; Tung levels of eCB breakdown enzymes (Di et al., 2013; Longaretti et al.,
et al., 2016; Zhong et al., 2011). It is well known that Gαq/11-GPCRs, 2020). It has therefore been suggested that inhibitors of endocannabi­
such as group 1 mGluRs, activate PLCβ which hydrolyses membrane noid degradation are likely to have important implication for numerous
phospholipids to produce inositol 1,4,5-trisphosphate (IP3) and 1,2-diac­ diseases and disorders (Cristino et al., 2020; Fowler, 2015; Senst and
ylglycerol (DAG), the latter of which is then converted by DAGL into Bains, 2014).
2-AG (section 1.1) (Fig. 1). Despite some earlier controversies (Kano
et al., 2009), more recent pharmacological and knockout studies have

6
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

3.3.2. Glutamate uptake and degradation synaptic efficacy, the direction of plasticity depends on the type of
Activation of Gαq/11-coupled retrograde endocannabinoid signalling, presynaptic input being stimulated and whether there is a homo- or
such as that by group 1 mGluRs, is also controlled by the uptake of hetero-synaptic contribution. For instance, eCB-LTP is primarily a form
glutamate. Normally, the spread of glutamate is tightly controlled by of heterosynaptic or metaplastic plasticity in which potentiation is
transporters, but spill-over from the synaptic cleft can lead to an facilitated by the long-lasting disinhibition of excitatory synapses
enhancement and spread of mGluR induced eCB mediated inhibition. attributed to eCB-mediated LTD of inhibitory inputs (iLTD; Basu et al.,
This spill-over can be caused by concurrent activation of nearby (rather 2013; Chevaleyre and Castillo, 2003, 2004; Kim et al., 2019; Monory
than spatially dispersed) synapses which overwhelms glutamate trans­ et al., 2015; Xu et al., 2012; Younts et al., 2013; Zhu and Lovinger,
porters at those synapses. It can also be promoted by drugs which inhibit 2007). It should be noted however, recent evidence indicates eCB-LTP
neuronal and/or glial glutamate transporters (Crepel and Daniel, 2007; can also be homosynaptic with eCBs acting at either TRPV1s or CB1Rs
Drew et al., 2008; Freestone et al., 2014; Linehan et al., 2018; Marcaggi (Cui et al., 2015; Cui et al., 2018; Wang et al., 2016 & see section 4.3.1).
and Attwell, 2005). A similar enhancement of mAChR induced retro­ In contrast, eCB-LTD may be homo- or heterosynaptic and can be
grade endocannabinoid signalling has been observed with cholines­ expressed at both glutamatergic and GABAergic synapses (Adermark
terase inhibitors (Lau and Vaughan, 2008; Narushima et al., 2007). et al., 2009; Arami et al., 2013; Chevaleyre and Castillo, 2003; Chiu
et al., 2010; Penasco et al., 2019; Peterfi et al., 2012). However, given
4. Long-term synaptic plasticity the frequent requirement for Gαq/11-coupled signalling, eCB-plasticity at
GABAergic synapses is often heterosynaptic, although some evidence
In addition to short-term changes in synaptic strength, it is now well indicates strong repetitive postsynaptic depolarisation is sufficient to
established eCBs play a central role in the induction of bidirectional induce homosynaptic eCB-plasticity at GABAergic synapses in the hip­
long-term synaptic plasticity. Long-term potentiation (LTP) and long- pocampus (Younts et al., 2013). In addition to retrograde signalling,
term depression (LTD) are two major forms of long-lasting plasticity eCBs act in an autocrine manner at postsynaptic CB1Rs or TRPV1s to
that describe the activity-dependent strengthening or weakening of induce LTD or LTP (Bacci et al., 2004; Chavez et al., 2010, 2014; Gibson
synaptic efficacy (respectively). They can be induced by diverse patterns et al., 2008; Marinelli et al., 2008, 2009), or at autapses (synaptic con­
of activity and unlike short-term plasticity, which results from a tran­ tacts of a neuron’s axon onto its own dendrite or soma, Yin et al., 2018),
sient signal that does not overtly change synaptic architecture, long- in a form of autaptic LTD (Kellogg et al., 2009). Conversely, eCBs might
term plasticity is maintained by changes in the presynapse (e.g. also be released from presynaptic sites to act at presynaptic CB1Rs and
release machinery) or postsynapse (e.g. receptor surface expression) or induce a presynaptic only form of LTD (Khlaifia et al., 2013) or at pre­
both, which fundamentally alters synaptic connectivity. This persistent synaptic TRPV1s to induce LTP (Bialecki et al., 2020). Further, CB1Rs
change in synaptic efficacy can last several hours or days. It is a brain- located on nearby astrocytes facilitate gliotransmission and this can
wide phenomenon that is critical for experience-dependent adaptions result in subsequent long-term plasticity (Gomez-Gonzalo et al., 2015).
of neural circuits, which can underlie a plethora of processes ranging Therefore, eCBs have the capacity to induce bidirectional long-term
from learning and memory formation to complex behavioural adaptions plasticity via multiple pathways and at a variety of different excitatory
to environmental stimuli (see Chistiakova and Volgushev, 2009; Citri and inhibitory synapses throughout the brain.
and Malenka, 2008; Collingridge et al., 2004; Yang and Calakos, 2013 eCB-plasticity can be induced by a plethora of stimulation protocols
for comprehensive reviews on synaptic plasticity). including: afferent only with or without postsynaptic depolarisation [e.
eCBs play a major role in regulating long-term plasticity and are g. theta burst stimulus (TBS; Arami et al., 2013; Zhao et al., 2015); low
necessary for the induction of several forms LTP and LTD (Araque et al., or high frequency stimuli (LFS, HFS respectively; Baca et al., 2015;
2017; Castillo et al., 2012; Piette et al., 2020). This has been demon­ Chevaleyre and Castillo, 2003; Gerdeman et al., 2002; Park et al.,
strated in numerous regions within the CNS including: the hippocampus 2017)], spike-timing dependent (coordinated pre-/post-synaptic firing;
(Chevaleyre and Castillo, 2003), striatum (Gerdeman et al., 2002), Andrade-Talavera et al., 2016; Sjostrom et al., 2003), and input-timing
amygdala (Azad et al., 2004), nucleus accumbens (Robbe et al., 2002), dependent (coordinated activity of more than one afferent input; Basu
nucleus tractus solitarius (Khlaifia et al., 2013), ventral tegmental area et al., 2013) stimuli, as well as repetitive strong postsynaptic depolar­
(Friend et al., 2017), cerebellum (Soler-Llavina and Sabatini, 2006), isations (Younts et al., 2013). Thus, eCB release and subsequent induc­
several cortical regions (e.g. prefrontal, Martin et al., 2015; somato­ tion of eCB-plasticity can occur under diverse conditions of neuronal and
sensory, Bender et al., 2006; visual, Joo et al., 2019 & insular, Liu et al., network activity. In addition, agonist activation of VGCCs and
2013; cortices) and the spinal cord (Kato et al., 2012). In addition, eCBs Gαq/11-coupled receptors can induce persistent eCB-plasticity (Ader­
have been implicated in metaplasticity (the plasticity of plasticity; mark and Lovinger, 2007a; Azad et al., 2004; Haj-Dahmane and Shen,
Chevaleyre and Castillo, 2004; Jensen et al., 2021; Melis et al., 2014; Xu 2014; Izumi and Zorumski, 2008, 2016, 2016; Martin et al., 2015).
et al., 2014; Yang et al., 2014); and synaptic scaling (Kim and Alger, Pinpointing the underlying mechanisms that govern eCB-plasticity
2010; Zhang et al., 2009), both of which are powerful homeostatic however has been challenging, in part due to the diverse signalling
processes that can set the gain of network activity and govern associative conditions that favour eCB-plasticity induction, which can vary sub­
information processing (Abraham, 2008; Turrigiano, 2008). eCBs can stantially depending on the microcircuit and brain region studied (see
therefore have a long-term impact on neuronal function in numerous Araque et al., 2017 for review). This is further complicated by ‘atypical’
brain regions and accordingly, eCBs have been linked to multiple pro­ mechanisms of eCB-signalling which include their action at pre- or
cesses including: learning and memory, goal-directed/habitual decision postsynaptically expressed TRPV1s or CB2Rs (Li and Kim, 2016) or their
making, sensory processing, motor control and feeding behaviours action at non-neuronal cells such as astrocytes and microglia (reviewed
(reviewed in Augustin and Lovinger, 2018). elsewhere in this issue). Here we shall concentrate on fundamental
Like short-term plasticity, eCB-dependent long-term plasticity conditions required for eCB-plasticity that depend on the retrograde
(henceforth referred as eCB-plasticity) is primarily a postsynaptically action of eCBs at CB1Rs. We shall discuss the current understanding and
induced but presynaptically expressed phenomenon (Azad et al., 2004; highlight recent advances and potential avenues for further research in
Carey et al., 2011; Chevaleyre and Castillo, 2003; Gerdeman et al., 2002; the mechanisms that govern this form of long-term eCB-plasticity, with a
Wilson and Nicoll, 2001). Through their retrograde action, eCBs that are particular focus on the molecular mechanisms that occur across the
released from the postsynapse, act on presynaptic CB1Rs to initiate a synapse within pre- and postsynaptic domains.
Gαi/o-signalling cascade that persistently reduces the probability of
neurotransmitter release via diverse mechanisms (discussed in detail
below). Since this is fundamentally an inhibitory process that reduces

7
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

4.1. Long-term eCB-plasticity requires extended CB1R activation have a role in the tonic control of excitatory inputs since genetic deletion
of DAGLα increases glutamate drive, particularly to direct pathway
A key difference between short- and long-term eCB-plasticity resides projection neurons (Shonesy et al., 2018). Therefore, while both eCBs
with the duration of CB1R activation. While short-term eCB plasticity is can act to induce long-term plasticity, there are distinct regional dif­
transient and requires CB1R activity on the duration of seconds, long- ferences that favour one or other of the main eCBs for specific roles.
term eCB-plasticity typically requires more prolonged CB1R activation
(Chevaleyre and Castillo, 2003; Ronesi et al., 2004). Importantly, this 4.1.1.1. A role for Ca2+- and GPCR-dependent signalling and their syn­
does not refer to the induction protocol, which are often brief high/low ergistic interaction in eCB-plasticity. Since differential mechanisms of
frequency or spike-timing stimuli (see above). Instead, if CB1R antago­ eCB-synthesis cannot account for the enhanced duration of CB1R acti­
nists are applied up to 5 min after induction, this can significantly vation required for eCB-plasticity, it is possible some other signal may
diminish the level of eCB-plasticity (Chevaleyre and Castillo, 2003; act in conjunction with those canonically associated with short-term
Ronesi et al., 2004), suggesting CB1R activity is required to persist plasticity (i.e. Gαq/11-coupled signalling and VGCCs) to promote suffi­
beyond the duration of the induction protocol. It should be noted cient eCB-synthesis for sustained CB1R activation. Alternatively, the
however, while prolonged eCB-activation of CB1Rs is necessary for the high degree of synergy between these two main stimuli can substantially
induction of plasticity, CB1R activity plays no role in plasticity mainte­ increase eCB-synthesis/release (see section 3.1.3), suggesting eCB-
nance, since antagonists applied >10 min after plasticity induction have plasticity might result from the synergistic interaction of Gαq/11-
no effect on long-term plasticity expression, which has been reported at dependent signalling and Ca2+-dependent processes downstream of
numerous synapses (Adermark and Lovinger, 2007a, b; Adermark et al., postsynaptic Gαq/11-coupled receptors and L-type VGCCs. Indeed, this
2009; Andrade-Talavera et al., 2016; Chevaleyre and Castillo, 2003; has been demonstrated at synapses where DSI/DSE and eCB-LTD are
Ronesi et al., 2004; Sjostrom et al., 2003). dependent on different eCB-synthesis pathways, whereby induction of
Several mechanisms have the potential to underlie sustained CB1R DSI/DSE can either enhance the magnitude of plasticity or reduce the
activation, which loosely fall into two categories: (1) increasing overall threshold for plasticity induction (Carlson et al., 2002; Chevaleyre and
levels of the endogenous agonist (e.g. enhanced eCB synthesis/release, Castillo, 2003; Edwards et al., 2008). Further, multiple studies found
reduced eCB metabolism) or (2) enhancing the receptor (CB1R), which activity-induced eCB-plasticity to be dependent on both Gαq/11 and
includes the more controversial ideas of endosomal CB1R signalling or VGCCs. Interestingly, it has been suggested that PLCβ acts as a coinci­
an endogenous allosteric modulator or binding protein that may alter dence detector due to its Ca2+ dependence, which is required for full
CB1R activity. catalytic function and its simultaneous activation by Gαq/11-receptors, of
which PLC is the main effector (Bender et al., 2006; Fino et al., 2010;
4.1.1. Mechanisms that promote enhanced eCB synthesis and/or release for Nevian and Sakmann, 2006). Thus, PLCβ might be key to mediating the
long-term eCB-plasticity synergistic interaction between VGCC and Gαq/11-dependent processes
Like short-term eCB-plasticity, a key determinant of persistent eCB- and promoting eCB-synthesis sufficient for inducing eCB-plasticity.
plasticity is the ability to induce ‘on-demand’ eCB synthesis and Importantly however, parallel activation of Gαq/11 and VGCCs is not
release. Given the prerequisite for prolonged CB1R activation (minutes always a requirement for eCB-plasticity since numerous instances have
not seconds), this indicates a requirement for higher eCB-concentrations been reported in which eCB-plasticity is either VGCC or G-protein in­
acting at the synapse and increased demands for eCB-synthesis and dependent. In these cases, alternative signalling cascades that substitute
release. Indeed, manipulations that would be expected to reduce eCB for one or other of the eCB-synthesis pathways have been implicated.
levels such as increasing FAAH/MAGL expression or knockdown of key Indeed, multiple candidates that either directly alter intracellular cal­
synthesis enzymes such as DAGLα (i.e. for 2-AG synthesis) can prevent cium (e.g. NMDARs, Djurisic et al., 2019; calcium permeable AMPARs,
eCB-plasticity, whilst inhibiting or genetic deletion of FAAH/MAGL can Manz et al., 2020; Soler-Llavina and Sabatini, 2006; or increased release
promote eCB-plasticity induction (Azad et al., 2004; Friend et al., 2019; of Ca2+ from intracellular stores, Bender et al., 2006; Fino et al., 2010),
Pan et al., 2011; Schurman et al., 2019; Wang et al., 2016; Zimmermann or feed into the downstream signalling cascades that upregulate
et al., 2019). PLC/DAGLα, PLC/PLD/Abdh4 to promote 2AG or AEA synthesis
Since multiple pathways govern eCB-synthesis, it is possible a respectively (e.g. BDNF/TrKB, Gangarossa et al., 2020; Lemtiri-Chlieh
mechanism divide exists between short- and long-term eCB-plasticity. and Levine, 2010; Pan et al., 2019; Zhao and Levine, 2014; see also
However, while early studies indicated this might be the case at single Hashimotodani et al., 2007a for review), have been shown to play a role
synapses (e.g. Chevaleyre and Castillo, 2003; Edwards et al., 2006), in eCB-plasticity. In particular, NMDARs have been shown to be a
there is no consistency across synapses. Instead, eCB-synthesis follows requirement for eCB-plasticity at several synapses (Arami et al., 2013;
the canonical pathways described above, with the relative calcium de­ Djurisic et al., 2019; Joo et al., 2019; Liu et al., 2013; Song et al., 2018),
pendency (i.e. calcium dependent, calcium assisted or calcium inde­ although this differs depending on the synapse studied and the mecha­
pendent) varying depending on the stimulus or brain region studied, nism of induction (Andrade-Talavera et al., 2016; Bender et al., 2006;
rather than being a defining feature of eCB-plasticity duration (see Haj-Dahmane and Shen, 2010; Huang et al., 2008; Penasco et al., 2019).
section 3.2; also reviewed in Hashimotodani et al., 2007a). This con­ Interestingly, a recent study has implicated the paired
trasts from postsynaptically expressed forms of synaptic plasticity in immunoglobulin-like receptor B (PirB) as a downstream target of
which the duration and concentration of the Ca2+ signal often defines NMDAR-dependent signalling that is required for postsynaptic
the direction of plasticity (reviewed in Evans and Blackwell, 2015). eCB-synthesis and/or release (Djurisic et al., 2019). Although typically
Interesting however, the relative contribution of the two main eCBs, associated with immune cell function, PirBs are also expressed in neu­
2-AG and AEA, in short-versus long-term plasticity differs at some syn­ rons and appear to play a role in governing dendritic spine dynamics and
apses. For example, in the hippocampus, 2-AG has been implicated in structural plasticity (Bochner et al., 2014; Djurisic et al., 2013; Vidal
the induction of DSE, DSI, LTD and LTP (Chevaleyre and Castillo, 2003; et al., 2016). At CA3-CA1 hippocampal synapses, eCB-LTD was impaired
Hashimotodani et al., 2013; Pan et al., 2011; Straiker et al., 2009), in PirB− /- mice while LTP was facilitated, an imbalance that could not be
whilst AEA appears to tonically regulate hippocampal activity (Zim­ attributed to impairments in mGluR or CB1R activity (Djurisic et al.,
mermann et al., 2019). Conversely at cortico-striatal synapses, AEA has 2019). It was suggested that PirB forms part of a negative feedback
been implicated in LTD and LTP (Ade and Lovinger, 2007; Adermark system that limits the release of neurotransmitter by supporting eCB
and Lovinger, 2009; Mathur et al., 2013; Ronesi et al., 2004), but 2-AG is retrograde signalling via an unknown mechanism. Further, since genetic
thought to govern short-term plasticity at this synapse (Shonesy et al., deletion or acute interference (using a recombinant ‘decoy’ receptor) of
2018; Tanimura et al., 2010; Uchigashima et al., 2007) and it may also

8
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

PirB results in increased spine density and an increase in functional protein directly activated by cAMP) was intracellularly loaded into
glutamatergic synapses (Bochner et al., 2014; Djurisic et al., 2013; Vidal dopaminergic VTA neurons (Tong et al., 2017). Clearly there is a com­
et al., 2016), this NMDAR/PirB/eCB link may be essential for normal plex interplay between the downstream signalling of Gαq/11-, Gαi/o-,
homeostatic control of LTP and aberrant excitation. While this illustrates Gαs-coupled receptors which can significantly alter eCB-synthesis and
the key role eCBs play in homeostatic control of excitation, further release required for eCB-plasticity. The extent to which likely varies
research is required to pinpoint the mechanism in which PirB might be depending on the microcircuit studied and the degree of integration
acting to support eCB-synthesis or release. with other neurotransmitter systems that signal within a dynamic
In addition to directly affecting Gαq/11 or Ca2+ cascades, alterations network.
to the function of key initiators (e.g. mGluRs, VGCCs, NMDARs) that
interrupt or enhance eCB-synthesis would also be expected to alter eCB- 4.1.1.2. A role for an ‘endocannabinoid signalosome’ and other post­
plasticity induction. Indeed, a recent study implicates changes in synaptic molecular complexes in eCB-plasticity?. Beyond enhancing syn­
NMDAR function can affect eCB signalling (Song et al., 2018). More ergy between the two eCB-synthesis pathways, the molecular
specifically, Song et al. found that RGS9-2− /− mice displayed reduced mechanisms that underlie enhanced eCB synthesis at GABAergic syn­
NMDAR currents and impaired retrograde eCB-signalling resulting in an apses remain poorly understood. More progress has been made at glu­
increase in quantal neurotransmitter release from excitatory inputs tamatergic synapses, where there is good evidence to suggest key
exclusively to striatal dopamine D2R expressing medium spiny neurons, elements required for eCB-synthesis including mGluRs, VGCCs, PLCβ,
which are known to form part of the basal ganglia indirect pathway DAGLα exist in a ‘signalosome’ which is held in close proximity to the
(iMSN; (Song et al., 2018). RGS9-2 is a regulator of G-protein signalling perisynapse in dendritic spines by longer coiled-coiled-forms of homer
protein (RGS) that accelerates intrinsic GTPase activity to inactivate (homer-1b & homer-2), key scaffolding proteins known to directly or
G-protein signalling and has been associated with several Gαi/o-coupled indirectly interact with L-type VGCCs, mGluR5 and DAGLα (Brakeman
receptors including: mu-opioid receptors and D2Rs (Cabrera-Vera et al., et al., 1997; Jung et al., 2007, 2012; Olson et al., 2005; Piomelli, 2014;
2004; Psifogeorgou et al., 2011). It was suggested RGS9-2 decreases Zhang et al., 2005). Although this ‘eCB-signalosome’ has so far only
Gαi/o-dependent signalling, which would increase cAMP/PKA activity been implicated in 2-AG signalling, the supramolecular complex may
(since Gαi/o-control of adenylyl cyclase is reduced) and increase provide a focal point for eCB generation dedicated to retrograde trans­
PKA-phosphorylation of NMDARs, which is known to increase NMDAR mission that is separate from functions such as eicosanoid production or
function and presumably promote eCB-synthesis (Skeberdis et al., 2006; phospholipid remodelling (see in Jung et al., 2012; Piomelli, 2014). It
Song et al., 2018). However, the role of postsynaptic cAMP and PKA may also facilitate synergy between VGCC-independent Ca2+ signalling
with regards to eCB-plasticity is not straightforward. For example, ge­ and Gαq/11 signalling since homer interacts with the Shank family of
netic deletion of adenylyl cyclase 5 (AC5), an isoform enriched in striatal scaffolding proteins, which provides a molecular link with NMDARs via
MSNs, was shown to impair eCB-LTD but via a mechanism of occlusion PSD-95 (Bertaso et al., 2010; Tu et al., 1999, although see below). In this
since postsynaptic loading of cAMP impaired eCB-LTD in wildtype mice regard, an interesting regulatory mechanism was reported, which in­
but induced LTP in AC5− /− mice (Kheirbek et al., 2009). Indeed, while volves BDNF/TrKB-dependent increases in the expression of the imme­
low intracellular cAMP levels are permissive for eCB-LTD, higher levels diate early gene homer1a (Roloff et al., 2010). Homer1a is a truncated
appear to promote NMDAR-dependent LTP at this synapse (Augustin isoform of homer1 that lacks the coiled-coiled C-terminus preventing
et al., 2014). The mechanism described in this instance was attributed to homer multimerization but the N-terminal EVH1-domain, which medi­
PKA-mediated phosphorylation and activation of RGS4, another RGS ates interactions with mGluRs and shank, is preserved (Ango et al.,
protein that inhibits the activity of Gαq/11-coupled receptors, including 2001). Therefore, homer1a acts in a dominant negative capacity to
mGluRs (Lerner and Kreitzer, 2012). In fact, this mechanism has been disassemble crosslinked complexes and prevent mGluRs and possibly
implicated in D2R/mGluR/VGCC-dependent eCB-LTD at cortico-striatal other members of the eCB-signalosome from interacting with longer
synapses, in which activation of Gαi/o-coupled D2Rs is a requirement for isoforms of homer (Ango et al., 2001; Clifton et al., 2019; Ronesi et al.,
eCB-LTD (Augustin et al., 2014; Kheirbek et al., 2009; Kreitzer and 2012). Indeed, BDNF was shown to inhibit mGluR-dependent short-term
Malenka, 2005; Lerner and Kreitzer, 2012; Pawlak and Kerr, 2008; Yin eCB-plasticity at excitatory hippocampal synapses due to increases in
and Lovinger, 2006), whilst activation of Gαs-coupled adenosine A2 homer1a expression (Roloff et al., 2010) and it was previously reported
receptors (which increases cAMP/PKA) impairs eCB-LTD (Tozzi et al., that BDNF impairs eCB-LTD at excitatory synapses in the visual cortex of
2011). The reason for this discrepancy in function for postsynaptic juvenile mice (Huang et al., 2008). However, it is not clear whether
cAMP/PKA signalling is unclear, although it may suggest a degree of these results reflect homer1a interference with the assembly of
compartmentalisation that segregates cAMP/PKA control of NMDARs eCB-synthesis machinery or the resulting constitutive activity of mGluRs
with that of mGluRs, perhaps via differential activation of the various (a widely reported effect of homer1a, Ango et al., 2001; Hu et al., 2010;
RGS proteins or dynamic regulation of the postsynaptic density (see Ronesi et al., 2012), which would occlude eCB-signalling and has been
below). It should also be noted, the precise cellular locus of D2Rs implicated in mouse models of fragile X syndrome (Aloisi et al., 2017;
required for eCB-LTD is under debate and recent evidence suggests D2Rs Guo et al., 2016; Jung et al., 2012; Ronesi et al., 2012; Tang and Alger,
that are expressed on cholinergic interneurons (which reduce acetyl­ 2015). Moreover, the role of BDNF in eCB-plasticity may be bidirec­
choline release and prevent M1AChR-inhibition of L-type VGCCs), play a tional since several other reports indicate BDNF signalling acts to
more significant role in governing eCB-LTD at corticostriatal synapses facilitate eCB-synthesis in various brain regions (Gangarossa et al.,
than those expressed postsynaptically on iMSNs (Augustin et al., 2018; 2020; Lemtiri-Chlieh and Levine, 2010; Maison et al., 2009; Pan et al.,
Wang et al., 2006). However, since much of this work has been con­ 2019; Zhao et al., 2015; Zhong et al., 2015), which likely involves
ducted at striatal synapses, it is not clear whether these mechanisms are cross-talk of TrKB-dependent signalling cascades to increase the activity
specific to this region. In this regard, it is interesting to note PKA has of PLCβ or DAGLα (Pan et al., 2019; Zhao et al., 2015; Zhong et al., 2015)
recently been shown to directly phosphorylate DAGLα which enhances or prolonging intracellular calcium transients (Gangarossa et al., 2020).
DAGLα activity and increases the magnitude of DSE in striatal D1R Another molecular scaffold that controls mGluR surface expression
expressing MSNs, which form part of the basal ganglia direct pathway and sets a threshold for mGluR activation to trigger eCB-plasticity in
(dMSN; Shonesy et al., 2020). Similarly, eCB-iLTD at inhibitory inputs to striatal medium spiny neurons has been described involving SAP90/
principal neurons of the basolateral amygdala was found to be depen­ PSD-95 associated proteins (SAPAP; also known as GKAP, Chen et al.,
dent on mGluR1 and PKA activity, but independent of PLC and DAGLα 2011). SAPAPs are a family of scaffolding proteins that bind to PSD-95 in
activity (Azad et al., 2004). Conversely, eCB-iLTD in the ventral a phosphorylation-dependent manner and to Shank via a PDZ domain
tegmental area (VTA) was impaired when an agonist of Epac (exchange

9
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

(Zeng et al., 2016; Zhu et al., 2017). PSD-95/SAPAP/Shank, together 2005; Ronesi et al., 2004), to date no single candidate has been identi­
with Homer1-3, form a core molecular complex of the postsynaptic fied (reviewed in Fowler, 2013; Nicolussi and Gertsch, 2015), although
density that links ionotropic glutamate receptors (i.e. AMPAR and several targets have been implicated, including: FAAH-like AEA trans­
NMDARs) with mGluRs (reviewed in Kim and Sheng, 2004; Ting et al., porter; (FLAT; Fu et al., 2011), fatty acid binding proteins (FABP;
2012). Intriguingly in SAPAP3 knockout mice, the threshold for Haj-Dahmane et al., 2018; Kaczocha et al., 2009); and most recently,
mGluR-induced (CB1R-dependent) depression was considerably Pannexin-1 (an ion/metabolite channel; Bialecki et al., 2020). Further,
reduced, suggesting SAPAP3 acts to limit eCB-plasticity induction by the precise role of these candidates is unclear, with evidence indicating
somehow limiting the surface availability of mGluRs (Chen et al., 2011). they primarily act as reuptake transporters to promote the clearance and
It is possible activity-dependent regulation of the interaction between catabolism of AEA (Bialecki et al., 2020; Fu et al., 2011; Kaczocha et al.,
SAPAP and PSD-95 may play a role in this gating of eCB-plasticity, 2009). In the case of FLAT, it was suggested this protein may also act as a
although at present the evidence supporting this is incomplete. Firstly, means of translocating AEA across the synaptic cleft due to its weak
it is known that CaMKIIα and β directly phosphorylate SAPAP at membrane binding properties and bidirectional transport of AEA into
different N-terminal residues which can either induce poly­ and out of HEK293 cells (Fu et al., 2011). Similarly, a putative endo­
ubiquitination and presumably degradation of SAPAP or increase SAPAP cannabinoid membrane transporter (EMT) was identified in U937
integration within the postsynaptic density (Shin et al., 2012). Secondly, human monocytes/macrophages that could bidirectionally transport
phosphorylation within an N-terminal ‘GK-binding repeat’-motif both AEA and 2-AG, as well as other structurally related eicosanoid eCBs
(14-amino acid repeats) that corresponds to the site targeted by CaMKII down their concentration gradients (Chicca et al., 2012). Further, FABP5
and increased SAPAP integration with the postsynaptic density, facili­ was recently implicated in controlling the retrograde transport of 2-AG,
tated SAPAP binding to PSD-95 (Shin et al., 2012; Zhu et al., 2017). in which pharmacological inhibition or genetic deletion of FAB5 could
Thirdly, aberrant CaMKII activity associated with stress has been shown impair eCB-LTD and eCB-dependent tonic control of glutamatergic in­
to impair eCB-LTD in the lateral habenula (Park et al., 2017) and it puts to dorsal raphe neurons (Haj-Dahmane et al., 2018). In this regard,
negatively modulates DSE in the striatum (Shonesy et al., 2013), it is interesting to note that a recent study implicates a novel mechanism
although the latter was attributed to CaMKII directly phosphorylating in which 2-AG is contained within non-synaptic extracellular vesicles
DAGLα and inhibiting its activity (Shonesy et al., 2013). Together, these and is released via a process involving vesicle fusion (Nakamura et al.,
findings indicate the molecular scaffolding that links mGluRs and 2019). FABP5 was found to colocalise with these extracellular vesicles
NMDARs hinders eCB-signalling, which appears to contrast with the and although not directly shown, it was suggested 2-AG binding to
findings reported above. It is possible that while long-homer might act as FABP5 may facilitate its release to the extracellular space (Nakamura
a molecular ‘dock’ for eCB-synthesis machinery, the additional inter­ et al., 2019). It is possible one or all these candidates may facilitate
action with SAPAP/PSD95/NMDAR might restrict and compartmen­ either eCB release or the retrograde transport of eCBs across the synaptic
talise the diffusion of rate-limiting signals such as Ca2+. It is tempting to cleft. However, direct observations of eCB transporters within a physi­
speculate that the gating of sufficient eCB-synthesis for eCB-plasticity is ological context are sparse, with FABP5 being the first to be linked to
dynamically regulated via activity-dependent changes in the molecular long-term eCB-plasticity. Further complexity is given by the poor spec­
complexes that constitute the postsynaptic density, which may alter the ificity of the transport inhibitors, which are known to have several
molecular links between Ca2+ and Gαq/11-dependent pathways. How­ off-target effects including activating TRPV1 (Zygmunt et al., 2000),
ever, it is important to note that this potential link can only pertain to sodium channels (Kelley and Thayer, 2004), calcium channels (Alptekin
glutamatergic synapses since the complex postsynaptic architecture et al., 2010), CB2Rs (Sagar et al., 2010) and inhibiting FAAH/MAGL
containing PSD-95 are absent at GABAergic synapses (see Sheng and (Vandevoorde and Fowler, 2005) (also reviewed in Nicolussi and
Kim, 2011 for review). Further work is required to fully elucidate the Gertsch, 2015). It therefore remains to be seen whether a transporter is a
dynamics that regulate the ‘eCB-signalosome’ and other molecular un­ central requirement for enhanced eCB-release to sufficiently sustain
derpinnings that promote eCB synthesis and plasticity in this regard. CB1R activation for long-term eCB-plasticity.

4.1.1.3. Does eCB-plasticity require a regulated eCB-release mechanism?. 4.1.1.4. Activity-dependent regulation of eCB-breakdown?. It is possible
It is possible sufficient eCB-release to induce eCB-plasticity requires a downregulation of the respective degrative enzymes MAGL and FAAH
regulated release mechanism such as through a membrane transporter that hydrolyse 2-AG or AEA (respectively) may underlie sustained CB1R
(Fowler, 2013; Nicolussi and Gertsch, 2015). Since eCBs are lipophilic activity required for plasticity induction, since this would be expected to
signalling molecules that can readily diffuse through lipid membranes, extend the lifetime of the eCB signal. Indeed, as previously mentioned,
facilitated release is not strictly required. However, at striatal and so­ inhibitors of either enzyme can either reduce the threshold for LTD/LTP
matosensory cortical synapses, facilitated eCB release through the pu­ or increase the magnitude of plasticity at numerous synapses throughout
tative AEA membrane transporter (AMT) was shown to be a requirement the brain (see section 4.1.1). Similarly, inhibitors of the more recently
for plasticity induction (Adermark and Lovinger, 2007b; Adermark identified ABHD6, which hydrolyses 2-AG at postsynaptic sites, was
et al., 2009; Maglio et al., 2018; Ronesi et al., 2004). Interestingly, when shown to facilitate LTD induction in the prefrontal cortex to otherwise
AEA was delivered postsynaptically to striatal MSNs to bypass the need subthreshold stimuli (Marrs et al., 2010). Endogenous regulation of
for eCB synthesis, release of this eCB was dependent on the strength of MAGL, FAAH and ABHD6 activity is not well defined, but likely involves
afferent activation (Adermark and Lovinger, 2007b). Whilst gluta­ changes in overall expression, rather than post-translational modifica­
matergic inputs required a stronger pre-synaptic stimulus; at GABAergic tions. In this regard, ketamine (a known modifier of synaptic plasticity;
synapses, spontaneous activity was sufficient to induce eCB-LTD but the Luo et al., 2020; Maren et al., 1991; Salami et al., 2000; Stringer and
magnitude of depression could be enhanced by extracellular stimuli Guyenet, 1983), was shown to increase 2-AG levels in the dorsal stria­
(Adermark and Lovinger, 2007b). These findings indicate eCB release tum by increasing the expression of PR domain protein 5 (PRDM5),
depends on sufficient pre-synaptic activity, which might upregulate the which functions as a transcriptional repressor of MAGL (Xu et al., 2020).
activity of a postsynaptically expressed transporter via an unknown Similarly, a link between presynaptic BDNF signalling, reduced MAGL
mechanism to facilitate eCB release (Adermark and Lovinger, 2007b; expression but increased CB1R expression in cultured cerebellar granule
Adermark et al., 2009; Ronesi et al., 2004). neurons has been reported (Maison et al., 2009). Although as previously
It should be noted however, the involvement of AMT is controversial. stated, the interaction between BDNF and eCB signalling is complex,
While activity and pharmacological assays support the existence of an since BDNF also acts at postsynaptic sites to facilitate or inhibit
eCB transporter (Chicca et al., 2012; Maglio et al., 2018; Moore et al., eCB-plasticity (see section 4.1.1). Conversely, the growth factor

10
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

neuregulin-1 was shown to enhance the degradation of 2-AG by acting at membrane and subcellularly expressed CB1Rs for plasticity induction.
ErbB receptor tyrosine kinases to increase the expression of MAGL, Given the lipophilicity of eCBs and the likelihood of a diffusion delay
which reduced the magnitude of iLTD at hippocampal synapses (Du between membrane expressed receptors and those expressed within
et al., 2013). Similarly, FAAH expression is known to be upregulated by intracellular compartments, it is conceivable the extended activation of
oestrogen (Grimaldi et al., 2012) and progesterone (Maccarrone et al., CB1Rs required for plasticity induction reflects activation of these two
2003) signalling although a link between this and eCB-plasticity has yet different pools of receptors. Further research in the area is required to
to be made. Indeed, while these studies demonstrate the expression and fully elucidate the functional role of CB1Rs signalling from intracellular
putative activity of MAGL/FAAH can be regulated, it is not clear compartments.
whether this a requirement for long-term eCB plasticity. Nevertheless,
given the plethora of evidence illustrating plasticity is facilitated when 4.1.2.2. Activity-dependent changes in the interaction with binding partners
these enzymes are pharmacologically inhibited, any process that results or other modifiers may alter CB1R affinity or efficacy?. Other intriguing
in a reduction in their activity or expression would be expected to pro­ possibilities for modifying CB1R activity are: direct or indirect post-
mote eCB-plasticity. translational modification (Madeo et al., 2016), altered interactions
with binding partners (Niehaus et al., 2007) or endogenous allosteric
4.1.2. Possible mechanisms to enhance CB1R signalling modulators (Lu et al., 2019; Pamplona et al., 2012; Straiker et al., 2015),
each of which might have the capacity to alter the affinity or efficacy of
4.1.2.1. CB1R signalling from intracellular compartments?. An under­ eCBs at CB1Rs or alter CB1R interaction with its downstream effectors.
studied mechanism of enhanced CB1R signalling in relation to long-term Indeed, lipoxin A4, an oxygenated derivative of arachidonic acid and a
eCB plasticity is the possibility of CB1R signalling from intracellular known anti-inflammatory mediator, was shown to bind to and enhance
compartments such as endosomes or mitochondria. As with the majority CB1R signalling in assays of cAMP production and activation of inwardly
of GPCRs, agonist binding to CB1Rs recruits β-arrestin and promotes rectifying K+ channels (Pamplona et al., 2012). Of particular note,
clathrin-dependent receptor endocytosis (Gainetdinov et al., 2004; knock-out mice that were deficient for the PTEN-induced putative ki­
Magalhaes et al., 2012). Traditionally, this process is viewed as a nase 1 (PINK1− /− ), were shown to have impaired binding ability of CB1R
mechanism of receptor desensitisation and/or inactivation whereby agonists, which could be restored following D2R activation (Madeo
CB1Rs are excluded from synapses (Mikasova et al., 2008), removed et al., 2016). Although the mechanism is unclear, this could indicate a
from the cell surface and may be targeted to either early or late endo­ role for PINK1 to indirectly promote eCB-CB1R interaction, perhaps
somes for recycling or lysosomal degradation (respectively; Gainetdinov through modification of the interaction of CB1Rs with binding proteins
et al., 2004; Magalhaes et al., 2012). It is also a major caveat to the idea such as CRIP1a/b, which are known to limit CB1R activity (Booth et al.,
that eCB-plasticity requires prolonged CB1R activity since these inacti­ 2019; Madeo et al., 2016; Niehaus et al., 2007). However, to our
vation/desensitisation mechanisms would be expected to be engaged knowledge, no reports have found a link between any of these possible
following eCB binding. Indeed, exogenous CB1R agonists such as THC modifications to CB1R function and long-term eCB-plasticity. In fact, in
have been shown to desensitise CB1Rs and impair eCB-LTD (Neuhofer the case of DSE, lipoxin A4 was found to have no effect on the magnitude
et al., 2020). Similarly, chronic inhibition of MAGL or genetic deletion, of DSE in cultured hippocampal neurons (Straiker et al., 2015). Never­
which produces an excess of 2AG, was shown to downregulate CB1Rs in theless, these may provide compelling avenues for further research.
numerous brain regions and impair eCB-plasticity (Imperatore et al.,
2015; Schlosburg et al., 2010). However, it is not clear whether condi­
tions that promote eCB-plasticity also promote CB1Rs desensitisation or 4.2. eCB-plasticity depends on highly regulated spatiotemporal
whether ‘prolonged’ activation represents a sustained eCB signal that coordination of pre- and postsynaptic activity
engages a greater number of CB1Rs over a larger time-space period (e.g.
extrasynaptic CB1Rs). As an interesting alternative, there is good evi­ 4.2.1. CB1R activation provides a temporal window for eCB-plasticity
dence to indicate CB1Rs are capable of signalling from endosomal lo­ induction
cations (Brailoiu et al., 2011; Rozenfeld and Devi, 2008; see also Jong Although CB1R activation is necessary for long-term eCB-plasticity
et al., 2018 for a review on intracellular GPCRs in synaptic plasticity). induction, a plethora of evidence indicates it is not sufficient and there is
Indeed, CB1Rs were found to be associated with Gαi/o in Rab7 positive a clear presynaptic requirement that has not been well defined. Indeed,
late-endosomal fractions and agonists such as WIN55,212–2 can induce multiple reports indicate activation of CB1Rs either directly (agonist) or
ERK1/2 phosphorylation (and activation) in the presence of membrane indirectly (upregulation of eCB synthesis) in the absence of adequate
impermeable CB1R antagonists (Rozenfeld and Devi, 2008). Further, afferent activity fails to produce long-term plasticity (Adermark and
given CB1Rs are enriched within the endosomal system, while plasma Lovinger, 2007a, b; Edwards et al., 2006; Heifets et al., 2008; Martin
membrane CB1Rs comprises ~20 % of total CB1Rs (Leterrier et al., 2006; et al., 2015; Ronesi et al., 2004; Singla et al., 2007; Sjostrom et al.,
Rozenfeld and Devi, 2008), it is possible endosomal signalling of CB1Rs 2003). It has been suggested that presynaptic activity may act as a
is a common feature of their activity. Aside from endosomes, CB1Rs are cofactor for plasticity induction, which is independent of eCB release
also expressed at mitochondrial membranes (so called mtCB1Rs) and and CB1R activation (Heifets and Castillo, 2009; Heifets et al., 2008).
have been shown to reduce cellular respiration and energy production in This was demonstrated in CA1 pyramidal neurons where iLTD, which
neuronal mitochondria (Benard et al., 2012). Remarkably, these was inhibited when spontaneous activity was reduced using low dose
mtCB1Rs we found to contribute to DSI (Benard et al., 2012) and a TTX, could be rescued by an afferent stimulus delivered immediately
longer form of agonist-induced depression of excitatory synaptic trans­ after the TBS induction protocol (Heifets et al., 2008). Further, paired
mission in hippocampal slices, which was linked to acute recordings between hippocampal interneurons and pyramidal neurons
cannabinoid-induced memory impairments in novel object recognition showed DHPG-induced iLTD, which would be expected to induce the
(Hebert-Chatelain et al., 2016). It is therefore possible mtCB1Rs release of eCBs, was abolished if the presynaptic interneuron was held
contribute towards the induction of long-term eCB-plasticity. This is silent (Heifets et al., 2008). Yet evidence from striatal synapses indicates
particularly compelling given mitochondria are known to support the adequate presynaptic activity is required to facilitate postsynaptic eCB
energy demand of highly active synapses (Sheng, 2014) and recent ev­ release, which is independent of eCB synthesis (Adermark and Lovinger,
idence suggests mitochondria may be dynamically regulated to affect 2007a; Adermark et al., 2009; Ronesi et al., 2004; although see discus­
synaptic strength (Cserep et al., 2018; Sheng, 2014). In this regard, it sion above on eCB transporters). Therefore, it is not clear whether pre­
would be tempting to propose eCBs may be required to activate both synaptic activity is required to enhance postsynaptic eCB release or to
facilitate an unknown presynaptic mechanism that acts in conjunction

11
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

with presynaptically expressed CB1Rs to persistently reduce the proba­ an additional role outside of their canonical receptor function. For
bility of neurotransmitter release. Given the nature of eCB-plasticity (i.e. instance, providing a reserve pool of receptors that might facilitate rapid
postsynaptically induced, but presynaptically expressed), it is likely the eCB-signalling under certain conditions or perhaps to sequester Gαi/o
coordinated activity of both pre- and postsynaptic activity is required. proteins from other GPCRs expressed at the same terminal (Vasquez and
Indeed, all forms of associative plasticity (including eCB-plasticity) Lewis, 1999; and also discussed in Busquets-Garcia et al., 2018).
require some form of coordinated and temporally correlated pre-/­
post-synaptic activity, the order of which determines the direction of 4.2.2. eCB-plasticity is spatially constrained to active synapses
plasticity as defined by Hebbian or anti-Hebbian learning rules (Feld­ Another feature of long-term eCB-plasticity is that it is relatively
man, 2012). Hebbian plasticity describes when synaptic strengthening input specific and expressed at synapses that are highly localised to
(LTP) results from activity in which presynaptic inputs lead or are syn­ those engaged by the site of stimulation (Chevaleyre and Castillo, 2003,
chronous with postsynaptic spikes (pre-post), while synaptic weakening 2004, 2004; Soler-Llavina and Sabatini, 2006). While precise temporal
(LTD) results from postsynaptic spikes preceding presynaptic input pre-/postsynaptic coordination plays a part in this specificity, it also
(post-pre). In contrast, anti-Hebbian plasticity reflects the opposite of concerns the degree of eCB diffusion from its release site, which provides
these conditions (LTP: post-pre; LTD: pre-post). It is interesting to note spatial constraints to eCB signalling. For eCBs to induce long-term
eCBs have been implicated in both Hebbian and anti-Hebbian forms of plasticity at inhibitory synapses, a degree of diffusion is expected be­
plasticity (Andrade-Talavera et al., 2016; Bender et al., 2006; Cui et al., tween the site of eCB-release and inhibitory terminals. Particularly since
2015, 2018; Fino et al., 2010; Tzounopoulos et al., 2007; Zhao and eCB-plasticity of inhibitory synapses is largely heterosynaptic and re­
Tzounopoulos, 2011). Further, the intensity of afferent activity used to quires non-GABAergic inputs (e.g. cholinergic, glutamatergic etc.) to
induce eCB-plasticity can vary widely even within the same synapses, activate Gαq/11-coupled signalling and initiate eCB synthesis to act at
but it is required within a limited time frame (Martin et al., 2015; Singla CB1Rs that are presumably located outside of the site of release (i.e. at
et al., 2007; Sjostrom et al., 2003). For example, when endogenous inhibitory terminals). However, whilst eCBs may travel considerable
acetylcholine release was evoked optogenetically to act a distances (up to 100 μm) under certain conditions (Kreitzer et al., 2002),
Gαq/11-coupled M1AChRs, eCB-LTD of excitatory inputs to prelimbic long-term eCB-plasticity is restricted to inputs located within 10–20 μm
cortical neurons was only induced when the light stimulus was delivered of the site of afferent stimulation (Chevaleyre and Castillo, 2003, 2004,
in conjunction with an appropriately timed extracellular afferent 2004; Soler-Llavina and Sabatini, 2006). This suggests eCB signalling
(Martin et al., 2015). Therefore, it is not the intensity of presynaptic induced by patterns of presynaptic activity, is restricted to a small region
activity per se, rather the duration of eCB-CB1R activation provides a along the somatodendritic axis. Yet, when eCB-iLTD was induced with
window in which temporally appropriate presynaptic activity is multiple postsynaptic depolarisations delivered at theta frequency, iLTD
required to induce long-term eCB-plasticity (Martin et al., 2015; Sjos­ was expressed at both somatic and dendritic inputs to CA1 pyramidal
trom et al., 2003). Although there appears to be no single rule for neurons, indicating at actively firing neurons, eCB-plasticity might be
afferent activity or the duration of CB1R action, variations in the pat­ induced across the whole somatodendritic axis (Younts et al., 2013).
terns, timing and order of pre- and/or post-synaptic stimulation can Then again, while eCB-plasticity was more globally expressed in depo­
drastically alter expression of eCB-plasticity; indeed, this temporal larised neurons, there were limitations to this: (i) iLTD was specific to
requirement has been widely reported at several synapses (Adermark the depolarised neuron, (ii) it was compartmentalised to GABAergic
and Lovinger, 2009; Adermark et al., 2009; Cui et al., 2016; Heifets inputs and (iii) postsynaptic firing was required at specific theta fre­
et al., 2008; Martin et al., 2015; Singla et al., 2007; Sjostrom et al., quencies (a network oscillation displayed during active waking (Younts
2003). et al., 2013). Nevertheless, these findings are particularly intriguing as
The degree of pre-/postsynaptic coordination required will likely they suggest a functional difference between afferent and
depend on several synaptic attributes. Firstly, as previously described, somatic-depolarisation induced eCB-plasticity. Highly localised
patterns of activity that upregulate both eCB-synthesis pathways would eCB-plasticity induced by patterns of afferent activity, likely plays a role
be expected to generate higher eCB concentrations (see sections 3.1.3 & in fine tuning synaptic strength in an active network. This specificity to
4.1.1; also reviewed in Hashimotodani et al., 2007a). As a result, this active synapses has significant implications for the integration of
may either overcome enzymatic degradation, thus increasing the chance different inputs to overall somatic output. While eCB-signalling at
of eCBs interacting with their receptors or eCBs may diffuse greater proximal dendrites might be expected to alter somatic output, given the
distances from their release site to engage CB1Rs expressed extra­ distance required for signal propagation to the soma, eCB-action at distal
synaptically. Secondly, the relative expression level of CB1Rs at axon dendrites might have less of an effect (although see Menon et al., 2013;
terminals and whether there is a receptor reserve (i.e. an excess of re­ Nicholson et al., 2006). Instead, local signalling at distal sites might alter
ceptor number required for full activation of the effector) will determine the threshold of plasticity at proximal dendrites (Dudman et al., 2007).
how strongly these receptors engage downstream signalling following In this regard however, the metaplastic eCB-dependent facilitation of
activation. This is particularly relevant to the differences between LTP in CA1 pyramidal neurons was restricted to pathways located
eCB-plasticity at GABAergic and glutamatergic synapses, arising from within 10 μm of one another (Younts et al., 2013). It is also interesting to
their relative expression of CB1Rs (see sections 2.1 & 2.2). While note that super-resolution imaging of hippocampal GABAergic axons,
eCB-plasticity is readily expressed at glutamatergic synapses, induction indicates perisomatic interneurons (i.e. interneurons located close to the
protocols often require stronger stimuli or more coordinated patterns of CA1 pyramidal cell layer) contain a more complex active zone with a
pre-/post-synaptic activity or activation of multiple inputs (Adermark higher number of CB1Rs and greater receptor/effector ratio compared
and Lovinger, 2007b, 2009, 2009; Adermark et al., 2009; Penasco et al., with dendritic interneurons (Dudok et al., 2015); although the func­
2019; Peterfi et al., 2012). Thirdly, the strength of the signal following tional consequence of this has yet to be investigated. Nevertheless, the
CB1R activation will depend on how tightly these GPCRs are coupled to high specificity of afferent-induced eCB-plasticity to active synapses
their effectors (i.e., VGCCs, Kirs and adenylyl cyclase). In this regard, it is indicates a subtle modulatory role of eCBs over neuronal activity.
interesting to note that while glutamatergic terminals express fewer Conversely, when neurons are actively firing at frequencies such as those
CB1Rs, there is evidence to suggest they maybe more tightly coupled to used in the theta burst firing protocol, which relies on
their downstream effectors (Steindel et al., 2013). At present, these somatic-depolarisation (Younts et al., 2013), eCBs may act to alter
differences between glutamatergic and GABAergic CB1R expression synaptic inputs across the entire somatodendritic axis. In the case of CA1
levels and their relative coupling with effectors are difficult to reconcile. pyramidal neurons, in which only inhibitory inputs are affected by
It has been suggested at synapses expressing high levels of CB1Rs that postsynaptic depolarisation-induced eCB-plasticity (Younts et al., 2013),
exhibit lower efficiency signalling, the excess CB1Rs may be performing this could play a role in sustaining excitatory rhythms of activity that

12
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

have been associate with behaviourally relevant processes such as theta activate inwardly rectifying K+ channels (Kir/GIRKs) via βγ-subunit
and active waking. Indeed, in instances when eCB-signalling has become signalling (see section 2.1). In addition, via a mechanism that likely
overactivated and dysregulated, the resulting hyperexcitability has been involves either Gαi/o-inhibition of c-Raf (which is usually phosphory­
linked to early epileptiform activity in the hippocampus, although it lated by PKA and inhibits Raf kinase/MEK) or βγ-recruitment of the
should be noted that eCB signalling at later stages supresses seizure PI3K-Akt pathway, CB1Rs activate MAPK signalling pathways including
generation (see Sugaya and Kano, 2018 for review). ERK and p38 (Derkinderen et al., 2001, 2003; Galve-Roperh et al., 2002;
see also Howlett, 2005 for review). Thus, CB1Rs have the capacity to
upregulate numerous downstream signalling pathways. Despite these
4.3. Presynaptic mechanisms underlying induction and expression of
advances, the precise mechanisms responsible for persistent reduction in
long-term eCB-plasticity
presynaptic activity are not well characterised. This is because the study
of real time presynaptic dynamics has been problematic due to limita­
Substantial progress has been made in understanding signal trans­
tions in available technology able to access the presynaptic bouton
duction pathways that are upregulated following CB1Rs activation, the
(discussed below). Nevertheless, substantial evidence implicates
consequence of which ranges from alterations in neuronal excitability to
eCB-plasticity depends on CB1R-dependent alterations in the activity of
changes in nuclear transcription/translation (see Howlett, 2005;
adenylyl cyclase and/or VGCCs, the relative contribution of which may
Howlett et al., 2002; Turu and Hunyady, 2010 for reviews). To briefly
vary depending on the synapse studied (discussed below). In contrast, to
recap, CB1Rs are canonically coupled to Gαi/o G-proteins which down­
our knowledge, little evidence has implicated a mechanism involving
regulate the activity of adenylyl cyclase and can also inhibit VGCCs and

Fig. 2. CB1Rs alter active zone signalling at presynaptic terminals to persistently reduce the probability of neurotransmitter release. Schematic depicting
putative pathways in which the active zone matrix might be altered downstream of presynaptic CB1R activation. Activation of CB1Rs initiates Gαi/o to inhibit adenylyl
cyclase (AC) and downregulate cAMP/PKA dependent signalling. (1) This prevents PKA from phosphorylating RIM1α, which reduces its interaction with other active
zone matrix proteins, including Rab3 located on synaptic vesicles, an interaction that is required to prime synaptic vesicle docking. (2) Once activated, CB1Rs
somehow interact with the GTPase Rac1 to inhibit GTP binding and activation of Rac1. (3) This prevents GTP-bound Rac1 from interacting with the WAVE1 complex,
exposing the VCA region (where activated ARP 2/3 binds), and allowing WAVE1/Arp 2/3 mediated assembly of F-actin from G-actin monomers. This inhibition of
actin polymerisation likely impairs synaptic vesicle transport from reserve pools to the active zone matrix and thus reduces the availability of neurotransmitter for
release. (4) Meanwhile, downstream of Gβγ signalling, voltage gated calcium channels (VGCC) are inhibited whilst mTOR is activated likely via Gβγ upregulation of
PI3K/Akt signalling pathways. mTOR initiates the nucleation of mTORC1/mTORC2 complexes, which are involved in numerous cellular processes including local
cap-dependent protein synthesis. (5) Newly synthesised proteins include ubiquitin-proteosome-system (UPS)-related proteins, whilst WAVE1, ARP2/3 (Arpc2),
synapsin, MUNC18, SNAP-25 and other active zone proteins are ubiquitinated in functional domains, which impairs key interactions and may target the proteins for
proteosomal degradation (6). Importantly, whilst protein ubiquitination is required of eCB-iLTD expression, proteosomal degradation was not but it may promote
subsequent reduction in presynaptic bouton volume and long-term weakening of the synapse.

13
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

GIRKs, suggesting persistent CB1R-βγ activation of these channels likely to decreased actin assembly, resulting in a reduced active zone matrix,
does not determine eCB-plasticity (see also in Lovinger, 2008). reduction in axon bouton density and eCB-iLTD in the hippocampus
(Fig. 2; Monday et al., 2020; Younts et al., 2016). It was found that
4.3.1. Alterations in the presynaptic active zone matrix downstream of eCB-iLTD requires cap-dependent protein synthesis, which was upre­
CB1R activation may underlie eCB-plasticity gulated by a mechanism involving mTOR (mammalian target of rapa­
The reduction in neurotransmitter release associated with CB1R- mycin) (Younts et al., 2016), likely via CB1R-activation of PI3K/Akt
induced inhibition of presynaptic VGCCs is typically associated with pathways (Gomez et al., 2011). mTOR is a serine/threonine kinase that
short-term depression of synaptic transmission (section 2.1). By upon activation, nucleates two distinct multiprotein complexes
contrast, only a handful of studies have implicated presynaptic VGCCs as (mTORC1 and mTORC2), which are involved in numerous cell processes
a requirement for long-term eCB-plasticity (Huang et al., 2003; Mato including cap-dependent protein synthesis and ribosome biogenesis
et al., 2008) and it is not clear whether this is due to persistent CB1R-βγ (Laplante and Sabatini, 2009). Intriguingly, downstream of CB1R-mTOR
effects or whether some other mechanism is at play. More progress has activation, the expression level of several proteins was increased, many
been made in the understanding of Gαi/o-downregulation of adenylyl of which play key roles in ubiquitin-proteosomal degradation (Monday
cyclase activity, which reduces cAMP synthesis and PKA-dependent et al., 2020). Conversely, expression of proteins associated with the
signalling (Fig. 2). Indeed, activators of adenylyl cyclase or inhibitors WAVE1 regulatory complex (a complex of proteins that promotes the
of PKA have been shown to inhibit or occlude long-term eCB-plasticity at assembly of actin filaments), such as ARPC2 (an essential component of
several synapses throughout the brain (Ahumada et al., 2013; Cheva­ the Arp2/3 complex required for actin polymerisation and assembly),
leyre et al., 2007; Haj-Dahmane and Shen, 2010; Mato et al., 2008; Park were decreased by CB1R activation (Monday et al., 2020). This finding is
et al., 2017). PKA is known to phosphorylate the active zone protein particularly intriguing as it suggests eCB-CB1R controls presynaptic
RIM1α (Rab3-interacting molecule 1α), which is thought to promote its proteostasis, a homeostatic mechanism that balances protein synthesis
interaction with other members of the active zone matrix and has been and degradation, which is vital for neuronal function and synaptic
linked to promoting presynaptic LTP (Lonart et al., 2003) although see plasticity (Hanus and Schuman, 2013). Further, it was found that actin
(Kaeser et al., 2008a). RIM1α, together with RIM1β (an isoform encoded remodelling via CB1R-inhibition of Rac1, a GTPase that is required to
by the same gene; Kaeser et al., 2008b), are part of a diverse family of expose the VCA region of the WAVE1 complex to promote Arp2/3 as­
multidomain adapter proteins that form a scaffold for various key pre­ sembly and actin polymerisation, is a requirement for iLTD expression
synaptic proteins including: 14-3-3, Munc13-1, Munc18-1, synapto­ (Monday et al., 2020). Of particular note, whilst CB1R activation
tagmin and VGCCs, which together dock synaptic vesicles to exocytotic induced a reduction in synaptic bouton size, this was not a requirement
machinery and prime them for fusion and release (Fig. 2; see also Gar­ of eCB-iLTD. Instead, inhibitors of ubiquitination completely prevented
cia-Junco-Clemente et al., 2005; Ziv and Garner, 2004 for comprehen­ iLTD expression. It was suggested that ubiquitination may prevent
sive reviews on active zones). Interestingly, eCB-iLTD in the Arp2/3 and other integral presynaptic proteins from interacting with
hippocampus was abolished in RIM1α− /- and RIM1αβ− /- mice, suggest­ their binding partners, which leads to actin depolymerisation and
ing CB1R-regulation of RIM1α (via inhibition of PKA) might be eventually a decrease in bouton density (Monday et al., 2020). This
responsible for underlying the sustained decrease in probability of pre­ process involves highly intricate signalling pathways that are not fully
synaptic release (Chevaleyre et al., 2007; Kaeser et al., 2008b). This was defined; however, it demonstrates the potential for eCB-CB1R signalling
the first direct indication that eCB-plasticity might result from altered to significantly alter both the signalling within presynaptic active zones
presynaptic release machinery and it is now becoming increasingly clear as well as presynaptic bouton density (Fig. 2). This would be expected to
that CB1Rs can significantly alter the active zone matrix (Fig. 2; Alonso markedly downregulate presynaptic function and could play a role in
et al., 2017; Glebov et al., 2017; Monday et al., 2020; Njoo et al., 2015; retraction of synaptic inputs. It remains to be seen whether this process
Wang et al., 2018). However, RIM1 does not appear to be a requirement is specific to inhibitory hippocampal synapses or whether it is common
for all eCB-plasticity, for example at corticostriatal synapses, while to brain wide eCB-plasticity.
conditional knock-out of RIM1αβ from excitatory neurons impairs gen­
eral synaptic transmission, eCB-LTD remained comparable to controls 4.3.1.2. Can CB1Rs bidirectionally alter the active zone matrix to promote
(Kupferschmidt et al., 2019). It is therefore likely that the precise mo­ eCB-LTP?. Intriguingly, and in contrast to the findings detailed above,
lecular underpinnings of eCB-plasticity may vary between synapses at recent evidence suggests downstream CB1R signalling may produce
different brain regions. Other mechanisms that involve changes to active bidirectional effects on the active zone matrix. Indeed, eCB-facilitation
zone proteins downstream of CB1R activation and may be implicated in of LTP at lateral perforant path inputs to dentate gyrus neurons of the
eCB-plasticity include: ERK-dependent phosphorylation of Munc18-1 hippocampus was shown to be dependent on a novel signalling pathway
(Schmitz et al., 2016; Wang et al., 2018) and reduced cAMP activation that involves CB1R activation of the integrin-associated tyrosine kinase
of Epac2 (guanine nucleotide exchange proteins activated by FAK (focal adhesion kinase) and subsequent activation of Rho/ROCK
cAMP)/PLC signalling, which was linked to decreased RIM1α content at (RhoA associated coiled coil containing kinase 2) signalling to promote
active zones and CB1R-dependent silencing of cerebellar granule cells presynaptic actin remodelling and presumably enhance neurotrans­
(Alonso et al., 2017; Ramirez-Franco et al., 2014). mitter release (Wang et al., 2016, 2018; see Hanna and El-Sibai, 2013;
Schmandke et al., 2007 for comprehensive reviews on Rho/ROCK sig­
4.3.1.1. A role for protein synthesis-dependent disassembly of the active nalling pathways and the control of actin dynamics). Remarkably,
zone matrix in eCB-LTD?. Of particular interest is the finding that long- facilitated LTP was independent of GABAergic activity and thus could
term eCB-plasticity is dependent on presynaptic protein synthesis, a not be explained by a disinhibition mechanism (Wang et al., 2016).
characteristic that is distinct from short-term eCB-plasticity (Adermark Rather, these findings indicate a novel mechanism in which CB1R ac­
et al., 2009; Monday et al., 2020; Navakkode and Korte, 2014; Yin et al., tivity may directly enhance the probability of presynaptic neurotrans­
2006; Younts et al., 2016). This has been demonstrated in the hippo­ mitter release. These findings are surprising given the wealth of
campus (Monday et al., 2020; Navakkode and Korte, 2014; Younts et al., evidence indicating CB1R activation reduces presynaptic strength (see
2016) and the striatum (Adermark et al., 2009; Yin et al., 2006), discussion above). Whether this is a brain-wide mechanism or restricted
although see (Jung et al., 2012) who find eCB-LTD in the nucleus to this particular synapse, remains to be seen. It is also not clear whether
accumbens is independent of protein synthesis. In two sequential studies this is mediated by Gαi/o-signalling or whether it results from an alter­
by the same group, an intricate signalling pathway has recently been native G-protein coupling to CB1Rs. In this regard, CB1R-dependent LTP
identified downstream of CB1R activation, which links protein synthesis at corticostriatal synapses was found to be dependent on presynaptic

14
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

activation of PKA, suggesting CB1Rs may couple to Gαs in this instance some other presynaptic mechanism that acts synergistically with CB1R
(Cui et al., 2016). Alternatively, CB1Rs expressed in astrocytes are signalling is likely required. However, this ‘additional’ presynaptic
coupled to Gαq/11 and induce LTP of neighbouring hippocampal CA1 signal is poorly defined. Firstly, it is not clear whether this signal is
neurons via a mechanism that involves Ca2+-dependent glutamate homosynaptic (i.e. contained within the synapse that is undergoing eCB-
release from astrocytes, which acts at presynaptically expressed plasticity) or whether other neuromodulator signalling from hetero­
NMDARs or group I mGluRs to increase the probability of neurotrans­ synaptic inputs might play a role. Secondly, it is not clear whether the
mitter release (Gomez-Gonzalo et al., 2015; Navarrete and Araque, signal is required to work in concert with CB1R signalling to facilitate the
2008, 2010). Indeed, substantial evidence supports the idea that CB1Rs induction of eCB-plasticity (e.g. by synergistically feeding into Gαi/o-
are promiscuous with their G-protein coupling, with evidence support­ pathways) or whether it indirectly alters mechanisms that facilitate the
ing coupling to Gαq/11, Gαs, Gαz and Gα12/13 throughout various brain expression of eCB-plasticity (e.g. by altering or priming the active zone/
regions (Caballero-Floran et al., 2016; Diez-Alarcia et al., 2016; Glass release machinery). It is likely a combination of all possible factors plays
and Felder, 1997; Navarrete and Araque, 2008; Prather et al., 2000). a role, and this varies depending on both overall network activity and
Despite this, the functional relevance of this heterogeneity in G-protein the neuronal activity profile within a particular microcircuit. Thus,
coupling is not well defined and is likely cell-type specific (discussed in identifying the exact presynaptic conditions that favour eCB-plasticity
Busquets-Garcia et al., 2018; Busquets-Garcia et al., 2015). Adding has been particularly challenging, which likely reflects the subtle role
further complexity to the downstream signalling of CB1Rs is the possi­ eCBs play in regulating a dynamic network.
bility of heterodimerisation with other GPCRs. In this regard, the func­ Despite these challenges, a prominent candidate for this additional
tional interplay between dopamine D2R determines whether CB1Rs presynaptic signal is the influx of presynaptic calcium either via pre­
signal via Gαi/o or Gαs (Glass and Felder, 1997). Therefore, while most of synaptically expressed NMDARs (Andrade-Talavera et al., 2016; Bender
the evidence supports canonical CB1R Gαi/o-signalling as the main et al., 2006; Heifets et al., 2008; Singla et al., 2007; Sjostrom et al.,
mechanism underlying eCB-plasticity, there is considerable propensity 2003), or perhaps through patterns of action-potential driven activation
for atypical CB1R signalling to distinctly alter synaptic efficacy. Future of presynaptic VGCCs (see in Adermark et al., 2009). Although the
studies that address the nuances of CB1R coupling under different precise function of this calcium signal is not well defined, it likely in­
physiological and pathological conditions would considerably increase volves calcium sensors such as the calcium/calmodulin-dependent
our understanding of the eCB system function and could uncover pos­ phosphatase calcineurin (also known as PP2B), the activity of which
sibilities for specific pharmacological interventions against neurological was required for eCB-LTD at both hippocampal and corticostriatal syn­
disorders. apses (Andrade-Talavera et al., 2016; Cui et al., 2016; Heifets et al.,
2008). Calcineurin is a serine/threonine phosphatase that is often
4.3.1.3. Towards a better understanding of presynaptically expressed eCB- associated with postsynaptic signalling and is involved in multiple
plasticity. To date, studying mechanisms that underlie changes in pre­ pathways implicated in synaptic plasticity, one of which is the regula­
synaptic function has been inherently difficult. This is due to the paucity tion of postsynaptically expressed VGCCs (L-type, CaV1.2/1.3) (Peterson
of techniques that are able to resolve presynaptic nanoarchitecture and et al., 1999; Yakel, 1997; Zuhlke et al., 1999). Both calcineurin and
signalling dynamics in real time. Traditional techniques such as electron CaV1.2 exist in a molecular complex mediated by the scaffolding protein
microscopy provide ultrastructural resolution of machinery contained AKAP79/150 (A-Kinase anchoring proteins), which together with PKA
within pre- and postsynaptic compartments but are limited in their and calmodulin bidirectionally regulate the activity of CaV1.2 (Dittmer
ability to give functional context. Conversely, paired electrophysiolog­ et al., 2014; Oliveria et al., 2012). However, at presynaptic sites, the role
ical recordings of synaptically connected neurons allows selective of calcineurin is less clear. Calcineurin, RIM1 and other components of
(usually pharmacological) manipulations to the pre- and post-synapse the active zone matrix are known to form part of a macromolecular
and provides functional context. However, this technique is extremely complex that constitutes so called the ‘VGCC nanodomain’, suggesting
challenging, particularly in brain slices which – outside of in vivo re­ they may act to regulate VGCC activity as well as synaptic vesicle
cordings – offers the closest to physiologically relevant conditions to exo/endocytosis (Muller et al., 2010; see also in Nanou and Catterall,
study synaptic plasticity. Indeed, the synapse of interest is limited to 2018 & Gandini and Zamponi, 2021). Indeed, calcineurin – together
between two neurons that are contained within the sliced tissue and with cyclin-dependent kinase 5 (cdk5), are known to play a role in
manipulations are limited by the selectivity and availability of phar­ governing synaptic vesicle cycling, which may involve changing the
macological (or other) reagents. Following the advent of super- interaction between P/Q- (CaV2.1) or N-type (CaV2.2) VGCCs with RIM1
resolution fluorescent imaging techniques such as STORM (stochastic or other release machinery (Kim and Ryan, 2013; Su et al., 2012;
optical reconstruction microscopy), PALM (photoactivatable local­ Tomizawa et al., 2002; also reviewed in Dolphin and Lee, 2020; Fassio
isation microscopy), STED (stimulated emission depletion microscopy) et al., 2016). Yet these reports disagree on whether calcineurin/cdk5
and SIM (structured illumination microscopy), live cell imaging of pre­ facilitate or impair VGCC activity (Kim and Ryan, 2013; Su et al., 2012).
synaptic machinery is now possible (see Nosov et al., 2020; Tonnesen In fact, inhibition of calcineurin was shown to dramatically reduce
and Nagerl, 2013 for reviews). Although these technologies are >10–15 action-potential induced calcium influx and exocytosis in cultured hip­
years old (Tonnesen and Nagerl, 2013), their use in studying presynaptic pocampal neurons (Kim and Ryan, 2013), whilst in dorsal root ganglion
dynamics during eCB-plasticity is still in its infancy. Combining these neurons, calcineurin acted to downregulate P/Q- and N-type currents
techniques with sophisticated genetic manipulations, will offer following TRPV1 activation (Wu et al., 2005). The reason for this
extremely powerful methods to fully interrogate the presynaptic discrepancy is unclear and might reflect the different routes of calcium
mechanisms underlying eCB-plasticity. Therefore, while our current entry (voltage-gated vs ligand-gated ion-channels). Nevertheless, it
understanding is relatively limited, future studies in this regard are demonstrates the complex relationship between the regulatory mecha­
eagerly anticipated to resolve the many unanswered questions regarding nisms that govern both VGCC activity and the active zone matrix at
the intricate presynaptic signalling mechanisms governing presynaptic sites. Thus, relating this back to the context of eCB-plasticity
eCB-plasticity. is difficult to reconcile and requires further work to fully elucidate the
function of pre-synaptic calcium outside its role in stimulating neuro­
4.3.2. Integration with other presynaptic signalling mechanisms transmitter release and the role calcium-dependent proteins, such as
Whilst it is important to fully characterise downstream signalling of calcineurin, may play in this.
CB1Rs, as we previously stated, in many cases, activation of CB1Rs alone
is not sufficient to induce eCB-plasticity (see section 4.2.1). Therefore,

15
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

5. Summary & conclusions Ango, F., Prezeau, L., Muller, T., Tu, J.C., Xiao, B., Worley, P.F., Pin, J.P., Bockaert, J.,
Fagni, L., 2001. Agonist-independent activation of metabotropic glutamate receptors
by the intracellular protein Homer. Nature 411, 962–965.
Research into the endogenous cannabinoid system has revealed that Arami, M.K., Sohya, K., Sarihi, A., Jiang, B., Yanagawa, Y., Tsumoto, T., 2013. Reciprocal
it is a near ubiquitous regulator of neuronal communication throughout Homosynaptic and heterosynaptic long-term plasticity of corticogeniculate
the brain. This regulation is observed as the retrograde control of syn­ projection neurons in layer VI of the mouse visual cortex. J. Neurosci. 33,
7787–7798.
aptic transmission over short-time scales, and by more complex mech­ Araque, A., Castillo, P.E., Manzoni, O.J., Tonini, R., 2017. Synaptic functions of
anisms over longer-time scales. Despite the widespread nature of eCB endocannabinoid signaling in health and disease. Neuropharmacology 124, 13–24.
retrograde control, there are substantial regional and synapse-specific Augustin, S.M., Beeler, J.A., McGehee, D.S., Zhuang, X., 2014. Cyclic AMP and afferent
activity govern bidirectional synaptic plasticity in striatopallidal neurons.
differences. The forms of eCB short- and long-term plasticity differ J. Neurosci. 34, 6692–6699.
across synapses and brain regions. Furthermore, eCB short- and long- Augustin, S.M., Chancey, J.H., Lovinger, D.M., 2018. Dual dopaminergic regulation of
term plasticity are not always readily expressed at the same synapse. corticostriatal plasticity by cholinergic interneurons and indirect pathway medium
spiny neurons. Cell Rep. 24, 2883–2893.
Nonetheless, the key difference between short- and long-term eCB- Augustin, S.M., Lovinger, D.M., 2018. Functional relevance of endocannabinoid-
plasticity appears to reside with the duration of CB1R activation and dependent synaptic plasticity in the central nervous system. ACS Chem. Neurosci. 9,
other yet to be defined presynaptic signalling cascades. 2146–2161.
Azad, S.C., Monory, K., Marsicano, G., Cravatt, B.F., Lutz, B., Zieglgansberger, W.,
We are only beginning to fully understand the processes involved in Rammes, G., 2004. Circuitry for associative plasticity in the amygdala involves
promoting long-term changes in synaptic efficacy that are dependent on endocannabinoid signaling. J. Neurosci. 24, 9953–9961.
eCB-signalling. While the mechanisms of eCB synthesis are relatively Baca, M., Schiess, A.R., Jelenik, D., James, C.D., Donald Partridge, L., 2015. Induction
frequency affects cortico-striatal synaptic plasticity with implications for frequency
well defined, beyond the synergistic interaction of these mechanisms, it
filtering. Brain Res. 1615, 80–88.
is not clear what else across the pre-/post-synaptic domains may pro­ Bacci, A., Huguenard, J.R., Prince, D.A., 2004. Long-lasting self-inhibition of neocortical
mote sufficient eCB synthesis or sustained CB1R activity necessary for interneurons mediated by endocannabinoids. Nature 431, 312–316.
induction of eCB-plasticity. Similarly, the mechanisms downstream of Bajo, M., Roberto, M., Schweitzer, P., 2009. Differential alteration of hippocampal
excitatory synaptic transmission by cannabinoid ligands. J. Neurosci. Res. 87,
CB1R activation that result in persistent reduction in neurotransmitter 766–775.
release remain poorly understood. Further research is needed to fully Basu, J., Srinivas, K.V., Cheung, S.K., Taniguchi, H., Huang, Z.J., Siegelbaum, S.A., 2013.
understand the integration of CB1R signalling with other presynaptic A cortico-hippocampal learning rule shapes inhibitory microcircuit activity to
enhance hippocampal information flow. Neuron 79, 1208–1221.
activity which might alter assembly of the active zone matrix. Here we Benard, G., Massa, F., Puente, N., Lourenco, J., Bellocchio, L., Soria-Gomez, E., Matias, I.,
have detailed several key avenues for further research that have sig­ Delamarre, A., Metna-Laurent, M., Cannich, A., Hebert-Chatelain, E., Mulle, C.,
nificant potential in governing eCB-plasticity. Those that concern Ortega-Gutierrez, S., Martin-Fontecha, M., Klugmann, M., Guggenhuber, S., Lutz, B.,
Gertsch, J., Chaouloff, F., Lopez-Rodriguez, M.L., Grandes, P., Rossignol, R.,
‘atypical’ CB1R signalling, such as from intracellular compartments, or Marsicano, G., 2012. Mitochondrial CB(1) receptors regulate neuronal energy
via mechanisms that are outside of Gαi/o-dependent pathways are metabolism. Nat. Neurosci. 15, 558–564.
particularly intriguing and attest to the intricacy of the eCB system. Bender, V.A., Bender, K.J., Brasier, D.J., Feldman, D.E., 2006. Two coincidence detectors
for spike timing-dependent plasticity in somatosensory cortex. J. Neurosci. 26,
4166–4177.
Acknowledgements Bertaso, F., Roussignol, G., Worley, P., Bockaert, J., Fagni, L., Ango, F., 2010. Homer1a-
dependent crosstalk between NMDA and metabotropic glutamate receptors in mouse
neurons. PloS One 5, e9755.
This study was supported by the Ernest Heine Family Foundation and
Best, A.R., Regehr, W.G., 2008. Serotonin evokes endocannabinoid release and
Pain Foundation. retrogradely suppresses excitatory synapses. J. Neurosci. 28, 6508–6515.
Bhaskaran, M.D., Smith, B.N., 2010. Effects of TRPV1 activation on synaptic excitation in
References the dentate gyrus of a mouse model of temporal lobe epilepsy. Exp. Neurol. 223,
529–536.
Bialecki, J., Werner, A., Weilinger, N.L., Tucker, C.M., Vecchiarelli, H.A., Egana, J.,
Abraham, W.C., 2008. Metaplasticity: tuning synapses and networks for plasticity. Nat. Mendizabal-Zubiaga, J., Grandes, P., Hill, M.N., Thompson, R.J., 2020. Suppression
Rev. Neurosci. 9, 387. of presynaptic glutamate release by postsynaptic metabotropic NMDA receptor
Ade, K.K., Lovinger, D.M., 2007. Anandamide regulates postnatal development of long- signalling to pannexin-1. J. Neurosci. 40, 729–742.
term synaptic plasticity in the rat dorsolateral striatum. J. Neurosci. 27, 2403–2409. Blankman, J.L., Cravatt, B.F., 2013. Chemical probes of endocannabinoid metabolism.
Adermark, L., Lovinger, D.M., 2007a. Combined activation of L-type Ca2+ channels and Pharmacol. Rev. 65, 849–871.
synaptic transmission is sufficient to induce striatal long-term depression. Bochner, D.N., Sapp, R.W., Adelson, J.D., Zhang, S., Lee, H., Djurisic, M., Syken, J.,
J. Neurosci. 27, 6781–6787. Dan, Y., Shatz, C.J., 2014. Blocking PirB up-regulates spines and functional synapses
Adermark, L., Lovinger, D.M., 2007b. Retrograde endocannabinoid signaling at striatal to unlock visual cortical plasticity and facilitate recovery from amblyopia. Sci.
synapses requires a regulated postsynaptic release step. Proc. Natl. Acad. Sci. U. S. A. Transl. Med. 6, 258ra140.
104, 20564–20569. Bodor, A.L., Katona, I., Nyíri, G., Mackie, K., Ledent, C., Hájos, N., Freund, T.F., 2005.
Adermark, L., Lovinger, D.M., 2009. Frequency-dependent inversion of net striatal Endocannabinoid signaling in rat somatosensory cortex: laminar differences and
output by endocannabinoid-dependent plasticity at different synaptic inputs. involvement of specific interneuron types. J. Neurosci. 25, 6845–6856.
J. Neurosci. 29, 1375–1380. Booth, W.T., Walker, N.B., Lowther, W.T., Howlett, A.C., 2019. Cannabinoid receptor
Adermark, L., Talani, G., Lovinger, D.M., 2009. Endocannabinoid-dependent plasticity at interacting protein 1a (CRIP1a): function and structure. Molecules 24.
GABAergic and glutamatergic synapses in the striatum is regulated by synaptic Brailoiu, G.C., Oprea, T.I., Zhao, P., Abood, M.E., Brailoiu, E., 2011. Intracellular
activity. Eur. J. Neurosci. 29, 32–41. cannabinoid type 1 (CB1) receptors are activated by anandamide. J. Biol. Chem. 286,
Ahumada, J., Fernandez de Sevilla, D., Couve, A., Buno, W., Fuenzalida, M., 2013. Long- 29166–29174.
term depression of inhibitory synaptic transmission induced by spike-timing Brakeman, P.R., Lanahan, A.A., O’Brien, R., Roche, K., Barnes, C.A., Huganir, R.L.,
dependent plasticity requires coactivation of endocannabinoid and muscarinic Worley, P.F., 1997. Homer: a protein that selectively binds metabotropic glutamate
receptors. Hippocampus 23, 1439–1452. receptors. Nature 386, 284–288.
Alger, B.E., 2002. Retrograde signaling in the regulation of synaptic transmission: focus Brenowitz, S.D., Regehr, W.G., 2003. Calcium dependence of retrograde inhibition by
on endocannabinoids. Prog. Neurobiol. 68, 247–286. endocannabinoids at synapses onto Purkinje cells. J. Neurosci. 23, 6373–6384.
Aloisi, E., Le Corf, K., Dupuis, J., Zhang, P., Ginger, M., Labrousse, V., Spatuzza, M., Brown, S.P., Brenowitz, S.D., Regehr, W.G., 2003. Brief presynaptic bursts evoke
Georg Haberl, M., Costa, L., Shigemoto, R., Tappe-Theodor, A., Drago, F., Vincenzo synapse-specific retrograde inhibition mediated by endogenous cannabinoids. Nat.
Piazza, P., Mulle, C., Groc, L., Ciranna, L., Catania, M.V., Frick, A., 2017. Altered Neurosci. 6, 1048–1057.
surface mGluR5 dynamics provoke synaptic NMDAR dysfunction and cognitive Busquets-Garcia, A., Bains, J., Marsicano, G., 2018. CB1 receptor signaling in the brain:
defects in Fmr1 knockout mice. Nat. Commun. 8, 1103. extracting specificity from ubiquity. Neuropsychopharmacology 43, 4–20.
Alonso, B., Bartolome-Martin, D., Ferrero, J.J., Ramirez-Franco, J., Torres, M., Sanchez- Busquets-Garcia, A., Desprez, T., Metna-Laurent, M., Bellocchio, L., Marsicano, G., Soria-
Prieto, J., 2017. CB1 receptors down-regulate a cAMP/Epac2/PLC pathway to Gomez, E., 2015. Dissecting the cannabinergic control of behavior: the where
silence the nerve terminals of cerebellar granule cells. J. Neurochem. 142, 350–364. matters. Bioessays 37, 1215–1225.
Alptekin, A., Galadari, S., Shuba, Y., Petroianu, G., Oz, M., 2010. The effects of Caballero-Floran, R.N., Conde-Rojas, I., Oviedo Chavez, A., Cortes-Calleja, H., Lopez-
anandamide transport inhibitor AM404 on voltage-dependent calcium channels. Eur. Santiago, L.F., Isom, L.L., Aceves, J., Erlij, D., Floran, B., 2016. Cannabinoid-induced
J. Pharmacol. 634, 10–15. depression of synaptic transmission is switched to stimulation when dopaminergic
Andrade-Talavera, Y., Duque-Feria, P., Paulsen, O., Rodriguez-Moreno, A., 2016. tone is increased in the globus pallidus of the rodent. Neuropharmacology 110,
Presynaptic spike timing-dependent long-term depression in the mouse 407–418.
Hippocampus. Cerebr. Cortex 26, 3637–3654.

16
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

Cabrera-Vera, T.M., Hernandez, S., Earls, L.R., Medkova, M., Sundgren-Andersson, A.K., Di, S., Malcher-Lopes, R., Halmos, K.C., Tasker, J.G., 2003. Nongenomic glucocorticoid
Surmeier, D.J., Hamm, H.E., 2004. RGS9-2 modulates D2 dopamine receptor- inhibition via endocannabinoid release in the hypothalamus: a fast feedback
mediated Ca2+ channel inhibition in rat striatal cholinergic interneurons. Proc. Natl. mechanism. J. Neurosci. 23, 4850–4857.
Acad. Sci. U. S. A. 101, 16339–16344. Di, S., Popescu, I.R., Tasker, J.G., 2013. Glial control of endocannabinoid heterosynaptic
Carey, M.R., Myoga, M.H., McDaniels, K.R., Marsicano, G., Lutz, B., Mackie, K., modulation in hypothalamic magnocellular neuroendocrine cells. J. Neurosci. 33,
Regehr, W.G., 2011. Presynaptic CB1 receptors regulate synaptic plasticity at 18331–18342.
cerebellar parallel fiber synapses. J. Neurophysiol. 105, 958–963. Diana, M.A., Levenes, C., Mackie, K., Marty, A., 2002. Short-term retrograde inhibition of
Carlson, G., Wang, Y., Alger, B.E., 2002. Endocannabinoids facilitate the induction of GABAergic synaptic currents in rat Purkinje cells is mediated by endogenous
LTP in the hippocampus. Nat. Neurosci. 5, 723–724. cannabinoids. J. Neurosci. 22, 200–208.
Cascio, M.G., Marini, P., 2015. Biosynthesis and fate of endocannabinoids. Handb. Exp. Diez-Alarcia, R., Ibarra-Lecue, I., Lopez-Cardona, A.P., Meana, J., Gutierrez-Adan, A.,
Pharmacol. 231, 39–58. Callado, L.F., Agirregoitia, E., Uriguen, L., 2016. Biased agonism of three different
Castillo, P.E., Younts, T.J., Chávez, A.E., Hashimotodani, Y., 2012. Endocannabinoid cannabinoid receptor agonists in mouse brain cortex. Front. Pharmacol. 7, 415.
signaling and synaptic function. Neuron 76, 70–81. Dittmer, P.J., Dell’Acqua, M.L., Sather, W.A., 2014. Ca2+/calcineurin-dependent
Centonze, D., Rossi, S., Prosperetti, C., Gasperi, V., De Chiara, V., Bari, M., Tscherter, A., inactivation of neuronal L-type Ca2+ channels requires priming by AKAP-anchored
Febbraro, F., Bernardi, G., Maccarrone, M., 2007. Endocannabinoids limit protein kinase A. Cell Rep. 7, 1410–1416.
metabotropic glutamate 5 receptor-mediated synaptic inhibition of striatal principal Djurisic, M., Brott, B.K., Saw, N.L., Shamloo, M., Shatz, C.J., 2019. Activity-dependent
neurons. Mol. Cell. Neurosci. 35, 302–310. modulation of hippocampal synaptic plasticity via PirB and endocannabinoids. Mol.
Chavez, A.E., Chiu, C.Q., Castillo, P.E., 2010. TRPV1 activation by endogenous Psychiatr. 24, 1206–1219.
anandamide triggers postsynaptic long-term depression in dentate gyrus. Nat. Djurisic, M., Vidal, G.S., Mann, M., Aharon, A., Kim, T., Ferrao Santos, A., Zuo, Y.,
Neurosci. 13, 1511–1518. Hubener, M., Shatz, C.J., 2013. PirB regulates a structural substrate for cortical
Chavez, A.E., Hernandez, V.M., Rodenas-Ruano, A., Chan, C.S., Castillo, P.E., 2014. plasticity. Proc. Natl. Acad. Sci. U. S. A. 110, 20771–20776.
Compartment-specific modulation of GABAergic synaptic transmission by TRPV1 Dolphin, A.C., Lee, A., 2020. Presynaptic calcium channels: specialized control of
channels in the dentate gyrus. J. Neurosci. 34, 16621–16629. synaptic neurotransmitter release. Nat. Rev. Neurosci. 21, 213–229.
Chen, M., Wan, Y., Ade, K., Ting, J., Feng, G., Calakos, N., 2011. Sapap3 deletion Drew, G.M., Lau, B.K., Vaughan, C.W., 2009. Substance P drives endocannabinoid-
anomalously activates short-term endocannabinoid-mediated synaptic plasticity. mediated disinhibition in a midbrain descending analgesic pathway. J. Neurosci. 29,
J. Neurosci. 31, 9563–9573. 7220–7229.
Chen, Y., Liu, X., Vickstrom, C.R., Liu, M.J., Zhao, L., Viader, A., Cravatt, B.F., Liu, Q.S., Drew, G.M., Mitchell, V.A., Vaughan, C.W., 2008. Glutamate spillover modulates
2016. Neuronal and astrocytic monoacylglycerol lipase limit the spread of GABAergic synaptic transmission in the rat midbrain periaqueductal grey via
endocannabinoid signaling in the cerebellum. eNeuro 3. metabotropic glutamate receptors and endocannabinoid signaling. J. Neurosci. 28,
Chevaleyre, V., Castillo, P.E., 2003. Heterosynaptic LTD of hippocampal GABAergic 808–815.
synapses: a novel role of endocannabinoids in regulating excitability. Neuron 38, Du, H., Kwon, I.K., Kim, J., 2013. Neuregulin-1 impairs the long-term depression of
461–472. hippocampal inhibitory synapses by facilitating the degradation of endocannabinoid
Chevaleyre, V., Castillo, P.E., 2004. Endocannabinoid-mediated metaplasticity in the 2-AG. J. Neurosci. 33, 15022–15031.
hippocampus. Neuron 43, 871–881. Dubruc, F., Dupret, D., Caillard, O., 2013. Self-tuning of inhibition by endocannabinoids
Chevaleyre, V., Heifets, B.D., Kaeser, P.S., Sudhof, T.C., Castillo, P.E., 2007. shapes spike-time precision in CA1 pyramidal neurons. J. Neurophysiol. 110,
Endocannabinoid-mediated long-term plasticity requires cAMP/PKA signaling and 1930–1944.
RIM1alpha. Neuron 54, 801–812. Dudman, J.T., Tsay, D., Siegelbaum, S.A., 2007. A role for synaptic inputs at distal
Chicca, A., Marazzi, J., Nicolussi, S., Gertsch, J., 2012. Evidence for bidirectional dendrites: instructive signals for hippocampal long-term plasticity. Neuron 56,
endocannabinoid transport across cell membranes. J. Biol. Chem. 287, 866–879.
34660–34682. Dudok, B., Barna, L., Ledri, M., Szabo, S.I., Szabadits, E., Pinter, B., Woodhams, S.G.,
Chistiakova, M., Volgushev, M., 2009. Heterosynaptic plasticity in the neocortex. Exp. Henstridge, C.M., Balla, G.Y., Nyilas, R., Varga, C., Lee, S.H., Matolcsi, M.,
Brain Res. 199, 377–390. Cervenak, J., Kacskovics, I., Watanabe, M., Sagheddu, C., Melis, M., Pistis, M.,
Chiu, C.Q., Puente, N., Grandes, P., Castillo, P.E., 2010. Dopaminergic modulation of Soltesz, I., Katona, I., 2015. Cell-specific STORM super-resolution imaging reveals
endocannabinoid-mediated plasticity at GABAergic synapses in the prefrontal cortex. nanoscale organization of cannabinoid signaling. Nat. Neurosci. 18, 75–86.
J. Neurosci. 30, 7236–7248. Edwards, D.A., Kim, J., Alger, B.E., 2006. Multiple mechanisms of endocannabinoid
Citri, A., Malenka, R.C., 2008. Synaptic plasticity: multiple forms, functions, and response initiation in hippocampus. J. Neurophysiol. 95, 67–75.
mechanisms. Neuropsychopharmacology 33, 18–41. Edwards, D.A., Zhang, L., Alger, B.E., 2008. Metaplastic control of the endocannabinoid
Clifton, N.E., Trent, S., Thomas, K.L., Hall, J., 2019. Regulation and function of activity- system at inhibitory synapses in hippocampus. Proc. Natl. Acad. Sci. U. S. A. 105,
dependent homer in synaptic plasticity. Mol Neuropsychiatry 5, 147–161. 8142–8147.
Collingridge, G.L., Isaac, J.T., Wang, Y.T., 2004. Receptor trafficking and synaptic Engler, B., Freiman, I., Urbanski, M., Szabo, B., 2006. Effects of exogenous and
plasticity. Nat. Rev. Neurosci. 5, 952–962. endogenous cannabinoids on GABAergic neurotransmission between the caudate-
Colmers, P.L.W., Bains, J.S., 2018. Presynaptic mGluRs control the duration of putamen and the globus pallidus in the mouse. J. Pharmacol. Exp. Therapeut. 316,
endocannabinoid-mediated DSI. J. Neurosci. 38, 10444–10453. 608–617.
Crepel, F., Daniel, H., 2007. Developmental changes in agonist-induced retrograde Evans, R.C., Blackwell, K.T., 2015. Calcium: amplitude, duration, or location? Biol. Bull.
signaling at parallel fiber-Purkinje cell synapses: role of calcium-induced calcium 228, 75–83.
release. J. Neurophysiol. 98, 2550–2565. Fassio, A., Fadda, M., Benfenati, F., 2016. Molecular machines determining the fate of
Cristino, L., Bisogno, T., Di Marzo, V., 2020. Cannabinoids and the expanded endocytosed synaptic vesicles in nerve terminals. Front. Synaptic Neurosci. 8, 10.
endocannabinoid system in neurological disorders. Nat. Rev. Neurol. 16, 9–29. Feldman, D.E., 2012. The spike-timing dependence of plasticity. Neuron 75, 556–571.
Cserep, C., Posfai, B., Schwarcz, A.D., Denes, A., 2018. Mitochondrial ultrastructure is Ferraro, A., Wig, P., Boscarino, J., Reich, C.G., 2020. Sex differences in endocannabinoid
coupled to synaptic performance at axonal release sites. eNeuro 5. modulation of rat CA1 dendritic neurotransmission. Neurobiol Stress 13, 100283.
Cui, Y., Paille, V., Xu, H., Genet, S., Delord, B., Fino, E., Berry, H., Venance, L., 2015. Fino, E., Paille, V., Cui, Y., Morera-Herreras, T., Deniau, J.M., Venance, L., 2010. Distinct
Endocannabinoids mediate bidirectional striatal spike-timing-dependent plasticity. coincidence detectors govern the corticostriatal spike timing-dependent plasticity.
J. Physiol. 593, 2833–2849. J. Physiol. 588, 3045–3062.
Cui, Y., Perez, S., Venance, L., 2018. Endocannabinoid-LTP mediated by CB1 and TRPV1 Földy, C., Lee, S.Y., Szabadics, J., Neu, A., Soltesz, I., 2007. Cell type-specific gating of
receptors encodes for limited occurrences of coincident activity in neocortex. Front. perisomatic inhibition by cholecystokinin. Nat. Neurosci. 10, 1128–1130.
Cell. Neurosci. 12, 182. Fortin, D.A., Trettel, J., Levine, E.S., 2004. Brief trains of action potentials enhance
Cui, Y., Prokin, I., Xu, H., Delord, B., Genet, S., Venance, L., Berry, H., 2016. pyramidal neuron excitability via endocannabinoid-mediated suppression of
Endocannabinoid dynamics gate spike-timing dependent depression and inhibition. J. Neurophysiol. 92, 2105–2112.
potentiation. Elife 5, e13185. Fowler, C.J., 2012. Anandamide uptake explained? Trends Pharmacol. Sci. 33, 181–185.
Daniel, H., Crepel, F., 2001. Control of Ca(2+) influx by cannabinoid and metabotropic Fowler, C.J., 2013. Transport of endocannabinoids across the plasma membrane and
glutamate receptors in rat cerebellar cortex requires K(+) channels. J. Physiol. 537, within the cell. FEBS J. 280, 1895–1904.
793–800. Fowler, C.J., 2015. The potential of inhibitors of endocannabinoid metabolism for drug
De Petrocellis, L., Nabissi, M., Santoni, G., Ligresti, A., 2017. Actions and regulation of development: a critical review. Handb. Exp. Pharmacol. 231, 95–128.
ionotropic cannabinoid receptors. Adv. Pharmacol. 80, 249–289. Fowler, C.J., Doherty, P., Alexander, S.P.H., 2017. Endocannabinoid turnover. Adv.
Derkinderen, P., Ledent, C., Parmentier, M., Girault, J.A., 2001. Cannabinoids activate Pharmacol. 80, 31–66.
p38 mitogen-activated protein kinases through CB1 receptors in hippocampus. Freestone, P.S., Guatteo, E., Piscitelli, F., di Marzo, V., Lipski, J., Mercuri, N.B., 2014.
J. Neurochem. 77, 957–960. Glutamate spillover drives endocannabinoid production and inhibits GABAergic
Derkinderen, P., Valjent, E., Toutant, M., Corvol, J.C., Enslen, H., Ledent, C., Trzaskos, J., transmission in the Substantia Nigra pars compacta. Neuropharmacology 79,
Caboche, J., Girault, J.A., 2003. Regulation of extracellular signal-regulated kinase 467–475.
by cannabinoids in hippocampus. J. Neurosci. 23, 2371–2382. Freestone, P.S., Wu, X.H., de Guzman, G., Lipski, J., 2015. Excitatory drive from the
Deroche, M.A., Lassalle, O., Castell, L., Valjent, E., Manzoni, O.J., 2020. Cell-type- and Subthalamic nucleus attenuates GABAergic transmission in the Substantia Nigra pars
endocannabinoid-specific synapse connectivity in the adult nucleus accumbens core. compacta via endocannabinoids. Eur. J. Pharmacol. 767, 144–151.
J. Neurosci. 40, 1028–1041. Freiman, I., Anton, A., Monyer, H., Urbanski, M.J., Szabo, B., 2006. Analysis of the
Di Marzo, V., 2018. New approaches and challenges to targeting the endocannabinoid effects of cannabinoids on identified synaptic connections in the caudate-putamen by
system. Nat. Rev. Drug Discov. 17, 623–639. paired recordings in transgenic mice. J. Physiol. 575, 789–806.

17
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

Friend, L., Weed, J., Sandoval, P., Nufer, T., Ostlund, I., Edwards, J.G., 2017. CB1- Haj-Dahmane, S., Shen, R.Y., 2014. Chronic stress impairs alpha1-adrenoceptor-induced
Dependent long-term depression in ventral tegmental area GABA neurons: a novel endocannabinoid-dependent synaptic plasticity in the dorsal raphe nucleus.
target for marijuana. J. Neurosci. 37, 10943–10954. J. Neurosci. 34, 14560–14570.
Friend, L.N., Williamson, R.C., Merrill, C.B., Newton, S.T., Christensen, M.T., Haj-Dahmane, S., Shen, R.Y., Elmes, M.W., Studholme, K., Kanjiya, M.P., Bogdan, D.,
Petersen, J., Wu, B., Ostlund, I., Edwards, J.G., 2019. Hippocampal stratum oriens Thanos, P.K., Miyauchi, J.T., Tsirka, S.E., Deutsch, D.G., Kaczocha, M., 2018. Fatty-
somatostatin-positive cells undergo CB1-dependent long-term potentiation and acid-binding protein 5 controls retrograde endocannabinoid signaling at central
express endocannabinoid biosynthetic enzymes. Molecules 24. glutamate synapses. Proc. Natl. Acad. Sci. U. S. A. 115, 3482–3487.
Fu, J., Bottegoni, G., Sasso, O., Bertorelli, R., Rocchia, W., Masetti, M., Guijarro, A., Hanna, S., El-Sibai, M., 2013. Signaling networks of Rho GTPases in cell motility. Cell.
Lodola, A., Armirotti, A., Garau, G., Bandiera, T., Reggiani, A., Mor, M., Cavalli, A., Signal. 25, 1955–1961.
Piomelli, D., 2011. A catalytically silent FAAH-1 variant drives anandamide Hanus, C., Schuman, E.M., 2013. Proteostasis in complex dendrites. Nat. Rev. Neurosci.
transport in neurons. Nat. Neurosci. 15, 64–69. 14, 638–648.
Fucich, E.A., Stielper, Z.F., Cancienne, H.L., Edwards, S., Gilpin, N.W., Molina, P.E., Hashimotodani, Y., Ohno-Shosaku, T., Kano, M., 2007a. Ca(2+)-assisted receptor-driven
Middleton, J.W., 2020. Endocannabinoid degradation inhibitors ameliorate endocannabinoid release: mechanisms that associate presynaptic and postsynaptic
neuronal and synaptic alterations following traumatic brain injury. J. Neurophysiol. activities. Curr. Opin. Neurobiol. 17, 360–365.
123, 707–717. Hashimotodani, Y., Ohno-Shosaku, T., Kano, M., 2007b. Presynaptic monoacylglycerol
Fukudome, Y., Ohno-Shosaku, T., Matsui, M., Omori, Y., Fukaya, M., Tsubokawa, H., lipase activity determines basal endocannabinoid tone and terminates retrograde
Taketo, M.M., Watanabe, M., Manabe, T., Kano, M., 2004. Two distinct classes of endocannabinoid signaling in the hippocampus. J. Neurosci. 27, 1211–1219.
muscarinic action on hippocampal inhibitory synapses: M2-mediated direct Hashimotodani, Y., Ohno-Shosaku, T., Maejima, T., Fukami, K., Kano, M., 2008.
suppression and M1/M3-mediated indirect suppression through endocannabinoid Pharmacological evidence for the involvement of diacylglycerol lipase in
signalling. Eur. J. Neurosci. 19, 2682–2692. depolarization-induced endocanabinoid release. Neuropharmacology 54, 58–67.
Gainetdinov, R.R., Premont, R.T., Bohn, L.M., Lefkowitz, R.J., Caron, M.G., 2004. Hashimotodani, Y., Ohno-Shosaku, T., Tanimura, A., Kita, Y., Sano, Y., Shimizu, T., Di
Desensitization of G protein-coupled receptors and neuronal functions. Annu. Rev. Marzo, V., Kano, M., 2013. Acute inhibition of diacylglycerol lipase blocks
Neurosci. 27, 107–144. endocannabinoid-mediated retrograde signalling: evidence for on-demand
Galante, M., Diana, M.A., 2004. Group I metabotropic glutamate receptors inhibit GABA biosynthesis of 2-arachidonoylglycerol. J. Physiol. 591, 4765–4776.
release at interneuron-Purkinje cell synapses through endocannabinoid production. Hashimotodani, Y., Ohno-Shosaku, T., Tsubokawa, H., Ogata, H., Emoto, K., Maejima, T.,
J. Neurosci. 24, 4865–4874. Araishi, K., Shin, H.S., Kano, M., 2005. Phospholipase Cbeta serves as a coincidence
Galve-Roperh, I., Rueda, D., Gomez del Pulgar, T., Velasco, G., Guzman, M., 2002. detector through its Ca2+ dependency for triggering retrograde endocannabinoid
Mechanism of extracellular signal-regulated kinase activation by the CB(1) signal. Neuron 45, 257–268.
cannabinoid receptor. Mol. Pharmacol. 62, 1385–1392. Hebert-Chatelain, E., Desprez, T., Serrat, R., Bellocchio, L., Soria-Gomez, E., Busquets-
Gandini, M.A., Zamponi, G.W., 2021. Voltage-gated calcium channel nanodomains: Garcia, A., Pagano Zottola, A.C., Delamarre, A., Cannich, A., Vincent, P., Varilh, M.,
molecular composition and function. FEBS J. https://doi.org/10.1111/febs.15759. Robin, L.M., Terral, G., García-Fernández, M.D., Colavita, M., Mazier, W., Drago, F.,
In press. Puente, N., Reguero, L., Elezgarai, I., Dupuy, J.W., Cota, D., Lopez-Rodriguez, M.L.,
Gangarossa, G., Perez, S., Dembitskaya, Y., Prokin, I., Berry, H., Venance, L., 2020. BDNF Barreda-Gómez, G., Massa, F., Grandes, P., Bénard, G., Marsicano, G., 2016.
controls bidirectional endocannabinoid plasticity at corticostriatal synapses. Cerebr. A cannabinoid link between mitochondria and memory. Nature 539, 555–559.
Cortex 30, 197–214. Heifets, B.D., Castillo, P.E., 2009. Endocannabinoid signaling and long-term synaptic
Gao, Y., Vasilyev, D.V., Goncalves, M.B., Howell, F.V., Hobbs, C., Reisenberg, M., plasticity. Annu. Rev. Physiol. 71, 283–306.
Shen, R., Zhang, M.Y., Strassle, B.W., Lu, P., Mark, L., Piesla, M.J., Deng, K., Heifets, B.D., Chevaleyre, V., Castillo, P.E., 2008. Interneuron activity controls
Kouranova, E.V., Ring, R.H., Whiteside, G.T., Bates, B., Walsh, F.S., Williams, G., endocannabinoid-mediated presynaptic plasticity through calcineurin. Proc. Natl.
Pangalos, M.N., Samad, T.A., Doherty, P., 2010. Loss of retrograde endocannabinoid Acad. Sci. U. S. A. 105, 10250–10255.
signaling and reduced adult neurogenesis in diacylglycerol lipase knock-out mice. Hentges, S.T., 2007. Synaptic regulation of proopiomelanocortin neurons can occur
J. Neurosci. 30, 2017–2024. distal to the arcuate nucleus. J. Neurophysiol. 97, 3298–3304.
Garcia-Junco-Clemente, P., Linares-Clemente, P., Fernandez-Chacon, R., 2005. Active Hentges, S.T., Low, M.J., Williams, J.T., 2005. Differential regulation of synaptic inputs
zones for presynaptic plasticity in the brain. Mol. Psychiatr. 10, 185–200 image 131. by constitutively released endocannabinoids and exogenous cannabinoids.
García-Morales, V., Montero, F., Moreno-López, B., 2015. Cannabinoid agonists J. Neurosci. 25, 9746–9751.
rearrange synaptic vesicles at excitatory synapses and depress motoneuron activity Hirasawa, M., Schwab, Y., Natah, S., Hillard, C.J., Mackie, K., Sharkey, K.A., Pittman, Q.
in vivo. Neuropharmacology 92, 69–79. J., 2004. Dendritically released transmitters cooperate via autocrine and retrograde
Gerdeman, G.L., Ronesi, J., Lovinger, D.M., 2002. Postsynaptic endocannabinoid release actions to inhibit afferent excitation in rat brain. J. Physiol. 559, 611–624.
is critical to long-term depression in the striatum. Nat. Neurosci. 5, 446–451. Ho, Y.C., Lee, H.J., Tung, L.W., Liao, Y.Y., Fu, S.Y., Teng, S.F., Liao, H.T., Mackie, K.,
Gibson, H.E., Edwards, J.G., Page, R.S., Van Hook, M.J., Kauer, J.A., 2008. TRPV1 Chiou, L.C., 2011. Activation of orexin 1 receptors in the periaqueductal gray of
channels mediate long-term depression at synapses on hippocampal interneurons. male rats leads to antinociception via retrograde endocannabinoid (2-
Neuron 57, 746–759. arachidonoylglycerol)-induced disinhibition. J. Neurosci. 31, 14600–14610.
Glaser, S.T., Abumrad, N.A., Fatade, F., Kaczocha, M., Studholme, K.M., Deutsch, D.G., Hoffman, A.F., Lupica, C.R., 2000. Mechanisms of cannabinoid inhibition of GABA(A)
2003. Evidence against the presence of an anandamide transporter. Proc. Natl. Acad. synaptic transmission in the hippocampus. J. Neurosci. 20, 2470–2479.
Sci. U. S. A. 100, 4269–4274. Hoffman, A.F., Lycas, M.D., Kaczmarzyk, J.R., Spivak, C.E., Baumann, M.H., Lupica, C.R.,
Glass, M., Felder, C.C., 1997. Concurrent stimulation of cannabinoid CB1 and dopamine 2017. Disruption of hippocampal synaptic transmission and long-term potentiation
D2 receptors augments cAMP accumulation in striatal neurons: evidence for a Gs by psychoactive synthetic cannabinoid ’Spice’ compounds: comparison with Δ(9)
linkage to the CB1 receptor. J. Neurosci. 17, 5327–5333. -tetrahydrocannabinol. Addiction Biol. 22, 390–399.
Glebov, O.O., Jackson, R.E., Winterflood, C.M., Owen, D.M., Barker, E.A., Doherty, P., Hoffman, A.F., Riegel, A.C., Lupica, C.R., 2003. Functional localization of cannabinoid
Ewers, H., Burrone, J., 2017. Nanoscale structural plasticity of the active zone matrix receptors and endogenous cannabinoid production in distinct neuron populations of
modulates presynaptic function. Cell Rep. 18, 2715–2728. the hippocampus. Eur. J. Neurosci. 18, 524–534.
Gomez-Gonzalo, M., Navarrete, M., Perea, G., Covelo, A., Martin-Fernandez, M., Hofmann, M.E., Bhatia, C., Frazier, C.J., 2011. Cannabinoid receptor agonists potentiate
Shigemoto, R., Lujan, R., Araque, A., 2015. Endocannabinoids induce lateral long- action potential-independent release of GABA in the dentate gyrus through a CB1
term potentiation of transmitter release by stimulation of gliotransmission. Cerebr. receptor-independent mechanism. J. Physiol. 589, 3801–3821.
Cortex 25, 3699–3712. Howlett, A.C., 2005. Cannabinoid receptor signaling. Handb. Exp. Pharmacol. 53–79.
Gomez, O., Sanchez-Rodriguez, A., Le, M., Sanchez-Caro, C., Molina-Holgado, F., Molina- Howlett, A.C., Barth, F., Bonner, T.I., Cabral, G., Casellas, P., Devane, W.A., Felder, C.C.,
Holgado, E., 2011. Cannabinoid receptor agonists modulate oligodendrocyte Herkenham, M., Mackie, K., Martin, B.R., Mechoulam, R., Pertwee, R.G., 2002.
differentiation by activating PI3K/Akt and the mammalian target of rapamycin International union of pharmacology. XXVII. Classification of cannabinoid receptors.
(mTOR) pathways. Br. J. Pharmacol. 163, 1520–1532. Pharmacol. Rev. 54, 161–202.
Grimaldi, P., Pucci, M., Di Siena, S., Di Giacomo, D., Pirazzi, V., Geremia, R., Hu, J.H., Park, J.M., Park, S., Xiao, B., Dehoff, M.H., Kim, S., Hayashi, T., Schwarz, M.K.,
Maccarrone, M., 2012. The faah gene is the first direct target of estrogen in the testis: Huganir, R.L., Seeburg, P.H., Linden, D.J., Worley, P.F., 2010. Homeostatic scaling
role of histone demethylase LSD1. Cell. Mol. Life Sci. 69, 4177–4190. requires group I mGluR activation mediated by Homer1a. Neuron 68, 1128–1142.
Guo, W., Molinaro, G., Collins, K.A., Hays, S.A., Paylor, R., Worley, P.F., Szumlinski, K. Huang, G.Z., Woolley, C.S., 2012. Estradiol acutely suppresses inhibition in the
K., Huber, K.M., 2016. Selective disruption of metabotropic glutamate receptor 5- hippocampus through a sex-specific endocannabinoid and mGluR-dependent
homer interactions mimics phenotypes of fragile X syndrome in mice. J. Neurosci. mechanism. Neuron 74, 801–808.
36, 2131–2147. Huang, Y., Yasuda, H., Sarihi, A., Tsumoto, T., 2008. Roles of endocannabinoids in
Haj-Dahmane, S., Shen, R.Y., 2005. The wake-promoting peptide orexin-B inhibits heterosynaptic long-term depression of excitatory synaptic transmission in visual
glutamatergic transmission to dorsal raphe nucleus serotonin neurons through cortex of young mice. J. Neurosci. 28, 7074–7083.
retrograde endocannabinoid signaling. J. Neurosci. 25, 896–905. Huang, Y.C., Wang, S.J., Chiou, L.C., Gean, P.W., 2003. Mediation of amphetamine-
Haj-Dahmane, S., Shen, R.Y., 2009. Endocannabinoids suppress excitatory synaptic induced long-term depression of synaptic transmission by CB1 cannabinoid receptors
transmission to dorsal raphe serotonin neurons through the activation of presynaptic in the rat amygdala. J. Neurosci. 23, 10311–10320.
CB1 receptors. J. Pharmacol. Exp. Therapeut. 331, 186–196. Imperatore, R., Morello, G., Luongo, L., Taschler, U., Romano, R., De Gregorio, D.,
Haj-Dahmane, S., Shen, R.Y., 2010. Regulation of plasticity of glutamate synapses by Belardo, C., Maione, S., Di Marzo, V., Cristino, L., 2015. Genetic deletion of
endocannabinoids and the cyclic-AMP/protein kinase A pathway in midbrain monoacylglycerol lipase leads to impaired cannabinoid receptor CB(1)R signaling
dopamine neurons. J. Physiol. 588, 2589–2604. and anxiety-like behavior. J. Neurochem. 135, 799–813.

18
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

Iremonger, K.J., Kuzmiski, J.B., Baimoukhametova, D.V., Bains, J.S., 2011. Dual leads to EPSP to spike potentiation in CA1 pyramidal neurons. J. Neurosci. 39,
regulation of anterograde and retrograde transmission by endocannabinoids. 224–237.
J. Neurosci. 31, 12011–12020. Kim, J., Alger, B.E., 2004. Inhibition of cyclooxygenase-2 potentiates retrograde
Isokawa, M., Alger, B.E., 2005. Retrograde endocannabinoid regulation of GABAergic endocannabinoid effects in hippocampus. Nat. Neurosci. 7, 697–698.
inhibition in the rat dentate gyrus granule cell. J. Physiol. 567, 1001–1010. Kim, J., Alger, B.E., 2010. Reduction in endocannabinoid tone is a homeostatic
Izumi, Y., Zorumski, C.F., 2008. Direct cortical inputs erase long-term potentiation at mechanism for specific inhibitory synapses. Nat. Neurosci. 13, 592–600.
Schaffer collateral synapses. J. Neurosci. 28, 9557–9563. Kim, J., Isokawa, M., Ledent, C., Alger, B.E., 2002. Activation of muscarinic acetylcholine
Izumi, Y., Zorumski, C.F., 2016. GABA and endocannabinoids mediate depotentiation of receptors enhances the release of endogenous cannabinoids in the hippocampus.
schaffer collateral synapses induced by stimulation of temperoammonic inputs. PloS J. Neurosci. 22, 10182–10191.
One 11, e0149034. Kim, S.H., Ryan, T.A., 2013. Balance of calcineurin Aalpha and CDK5 activities sets
Jennings, E.A., Vaughan, C.W., Christie, M.J., 2001. Cannabinoid actions on rat release probability at nerve terminals. J. Neurosci. 33, 8937–8950.
superficial medullary dorsal horn neurons in vitro. Journal of Physiology-London Kiritoshi, T., Ji, G., Neugebauer, V., 2016. Rescue of impaired mGluR5-driven
534, 805–812. endocannabinoid signaling restores prefrontal cortical output to inhibit pain in
Jennings, E.A., Vaughan, C.W., Roberts, L.A., Christie, M.J., 2003. The actions of arthritic rats. J. Neurosci. 36, 837–850.
anandamide on rat superficial medullary dorsal horn neurons in vitro. Journal of Kiritoshi, T., Sun, H., Ren, W., Stauffer, S.R., Lindsley, C.W., Conn, P.J., Neugebauer, V.,
Physiology-London 548, 121–129. 2013. Modulation of pyramidal cell output in the medial prefrontal cortex by
Jensen, K.R., Berthoux, C., Nasrallah, K., Castillo, P.E., 2021. Multiple cannabinoid mGluR5 interacting with CB1. Neuropharmacology 66, 170–178.
signaling cascades powerfully suppress recurrent excitation in the hippocampus. Kodirov, S.A., Jasiewicz, J., Amirmahani, P., Psyrakis, D., Bonni, K., Wehrmeister, M.,
Proc. Natl. Acad. Sci. U. S. A. 118, e2017590118. Lutz, B., 2010. Endogenous cannabinoids trigger the depolarization-induced
Jo, Y.H., Chen, Y.J., Chua Jr., S.C., Talmage, D.A., Role, L.W., 2005. Integration of suppression of excitation in the lateral amygdala. Learn. Mem. 17, 43–49.
endocannabinoid and leptin signaling in an appetite-related neural circuit. Neuron Kortleven, C., Bruneau, L.C., Trudeau, L.E., 2012. Neurotensin inhibits glutamate-
48, 1055–1066. mediated synaptic inputs onto ventral tegmental area dopamine neurons through the
Jong, Y.I., Harmon, S.K., O’Malley, K.L., 2018. Intracellular GPCRs play key roles in release of the endocannabinoid 2-AG. Neuropharmacology 63, 983–991.
synaptic plasticity. ACS Chem. Neurosci. 9, 2162–2172. Kreitzer, A.C., Carter, A.G., Regehr, W.G., 2002. Inhibition of interneuron firing extends
Joo, K., Cho, K.H., Youn, S.H., Jang, H.J., Rhie, D.J., 2019. Layer-specific involvement of the spread of endocannabinoid signaling in the cerebellum. Neuron 34, 787–796.
endocannabinoid signaling in muscarinic-induced long-term depression in layer 2/3 Kreitzer, A.C., Malenka, R.C., 2005. Dopamine modulation of state-dependent
pyramidal neurons of rat visual cortex. Brain Res. 1712, 124–131. endocannabinoid release and long-term depression in the striatum. J. Neurosci. 25,
Jordan, C.J., Xi, Z.X., 2019. Progress in brain cannabinoid CB(2) receptor research: from 10537–10545.
genes to behavior. Neurosci. Biobehav. Rev. 98, 208–220. Kreitzer, A.C., Regehr, W.G., 2001. Retrograde inhibition of presynaptic calcium influx
Jung, K.M., Astarita, G., Zhu, C., Wallace, M., Mackie, K., Piomelli, D., 2007. A key role by endogenous cannabinoids at excitatory synapses onto Purkinje cells. Neuron 29,
for diacylglycerol lipase-alpha in metabotropic glutamate receptor-dependent 717–727.
endocannabinoid mobilization. Mol. Pharmacol. 72, 612–621. Kupferschmidt, D.A., Augustin, S.M., Johnson, K.A., Lovinger, D.M., 2019. Active zone
Jung, K.M., Sepers, M., Henstridge, C.M., Lassalle, O., Neuhofer, D., Martin, H., proteins RIM1alphabeta are required for normal corticostriatal transmission and
Ginger, M., Frick, A., DiPatrizio, N.V., Mackie, K., Katona, I., Piomelli, D., action control. J. Neurosci. 39, 1457–1470.
Manzoni, O.J., 2012. Uncoupling of the endocannabinoid signalling complex in a Kushmerick, C., Price, G.D., Taschenberger, H., Puente, N., Renden, R., Wadiche, J.I.,
mouse model of fragile X syndrome. Nat. Commun. 3, 1080. Duvoisin, R.M., Grandes, P., von Gersdorff, H., 2004. Retroinhibition of presynaptic
Kaczocha, M., Glaser, S.T., Deutsch, D.G., 2009. Identification of intracellular carriers for Ca2+ currents by endocannabinoids released via postsynaptic mGluR activation at a
the endocannabinoid anandamide. Proc. Natl. Acad. Sci. U. S. A. 106, 6375–6380. calyx synapse. J. Neurosci. 24, 5955–5965.
Kaeser, P.S., Kwon, H.B., Blundell, J., Chevaleyre, V., Morishita, W., Malenka, R.C., Laaris, N., Good, C.H., Lupica, C.R., 2010. Delta9-tetrahydrocannabinol is a full agonist
Powell, C.M., Castillo, P.E., Sudhof, T.C., 2008a. RIM1alpha phosphorylation at at CB1 receptors on GABA neuron axon terminals in the hippocampus.
serine-413 by protein kinase A is not required for presynaptic long-term plasticity or Neuropharmacology 59, 121–127.
learning. Proc. Natl. Acad. Sci. U. S. A. 105, 14680–14685. Laplante, M., Sabatini, D.M., 2009. mTOR signaling at a glance. J. Cell Sci. 122,
Kaeser, P.S., Kwon, H.B., Chiu, C.Q., Deng, L., Castillo, P.E., Sudhof, T.C., 2008b. 3589–3594.
RIM1alpha and RIM1beta are synthesized from distinct promoters of the RIM1 gene Lau, B.K., Drew, G.M., Mitchell, V.A., Vaughan, C.W., 2014. Endocannabinoid
to mediate differential but overlapping synaptic functions. J. Neurosci. 28, modulation by FAAH and monoacylglycerol lipase within the analgesic circuitry of
13435–13447. the periaqueductal grey. Br. J. Pharmacol. 171, 5225–5236.
Kamprath, K., Romo-Parra, H., Häring, M., Gaburro, S., Doengi, M., Lutz, B., Pape, H.C., Lau, B.K., Vaughan, C.W., 2008. Muscarinic modulation of synaptic transmission via
2011. Short-term adaptation of conditioned fear responses through endocannabinoid endocannabinoid signalling in the rat midbrain periaqueductal gray. Mol.
signaling in the central amygdala. Neuropsychopharmacology 36, 652–663. Pharmacol. 74, 1392–1398.
Kano, M., Ohno-Shosaku, T., Hashimotodani, Y., Uchigashima, M., Watanabe, M., 2009. Lee, S.H., Földy, C., Soltesz, I., 2010. Distinct endocannabinoid control of GABA release
Endocannabinoid-mediated control of synaptic transmission. Physiol. Rev. 89, at perisomatic and dendritic synapses in the hippocampus. J. Neurosci. 30,
309–380. 7993–8000.
Kargar, H.M., Azizi, H., Mirnajafi-Zadeh, J., Mani, A.R., Semnanian, S., 2018. Orexin A Lee, S.H., Ledri, M., Tóth, B., Marchionni, I., Henstridge, C.M., Dudok, B., Kenesei, K.,
presynaptically decreases inhibitory synaptic transmission in rat locus coeruleus Barna, L., Szabó, S.I., Renkecz, T., Oberoi, M., Watanabe, M., Limoli, C.L., Horvai, G.,
neurons. Neurosci. Lett. 683, 89–93. Soltesz, I., Katona, I., 2015. Multiple forms of endocannabinoid and endovanilloid
Kato, A., Punnakkal, P., Pernia-Andrade, A.J., von Schoultz, C., Sharopov, S., Nyilas, R., signaling regulate the tonic control of GABA release. J. Neurosci. 35, 10039–10057.
Katona, I., Zeilhofer, H.U., 2012. Endocannabinoid-dependent plasticity at spinal Lemtiri-Chlieh, F., Levine, E.S., 2010. BDNF evokes release of endogenous cannabinoids
nociceptor synapses. J. Physiol. 590, 4717–4733. at layer 2/3 inhibitory synapses in the neocortex. J. Neurophysiol. 104, 1923–1932.
Katona, I., Freund, T.F., 2012. Multiple functions of endocannabinoid signaling in the Lerner, T.N., Kreitzer, A.C., 2012. RGS4 is required for dopaminergic control of striatal
brain. Annu. Rev. Neurosci. 35, 529–558. LTD and susceptibility to parkinsonian motor deficits. Neuron 73, 347–359.
Katona, I., Urban, G.M., Wallace, M., Ledent, C., Jung, K.M., Piomelli, D., Mackie, K., Leterrier, C., Laine, J., Darmon, M., Boudin, H., Rossier, J., Lenkei, Z., 2006. Constitutive
Freund, T.F., 2006. Molecular composition of the endocannabinoid system at activation drives compartment-selective endocytosis and axonal targeting of type 1
glutamatergic synapses. J. Neurosci. 26, 5628–5637. cannabinoid receptors. J. Neurosci. 26, 3141–3153.
Katona, I.N., Rancz, E.A., Acsady, L., Ledent, C., Mackie, K., Hajos, N., Freund, T.F., Li, Y., Kim, J., 2016. Deletion of CB2 cannabinoid receptors reduces synaptic
2001. Distribution of CB1 cannabinoid receptors in the amygdala and their role in transmission and long-term potentiation in the mouse hippocampus. Hippocampus
the control of GABAergic transmission. J. Neurosci. 21, 9506–9518. 26, 275–281.
Kawahara, H., Drew, G.M., Christie, M.J., Vaughan, C.W., 2011. Inhibition of fatty acid Liao, H.T., Lee, H.J., Ho, Y.C., Chiou, L.C., 2011. Capsaicin in the periaqueductal gray
amide hydrolase unmasks CB1 receptor and TRPV1 channel-mediated modulation of induces analgesia via metabotropic glutamate receptor-mediated endocannabinoid
glutamatergic synaptic transmission in midbrain periaqueductal grey. Br. J. retrograde disinhibition. Br. J. Pharmacol. 163, 330–345.
Pharmacol. 163, 1214–1222. Linehan, V., Fang, L.Z., Hirasawa, M., 2018. Short-term high-fat diet primes excitatory
Kelley, B.G., Thayer, S.A., 2004. Anandamide transport inhibitor AM404 and structurally synapses for long-term depression in orexin neurons. J. Physiol. 596, 305–316.
related compounds inhibit synaptic transmission between rat hippocampal neurons Lines, J., Covelo, A., Gómez, R., Liu, L., Araque, A., 2017. Synapse-specific regulation
in culture independent of cannabinoid CB1 receptors. Eur. J. Pharmacol. 496, 33–39. revealed at single synapses is concealed when recording multiple synapses. Front.
Kellogg, R., Mackie, K., Straiker, A., 2009. Cannabinoid CB1 receptor-dependent long- Cell. Neurosci. 11, 367.
term depression in autaptic excitatory neurons. J. Neurophysiol. 102, 1160–1171. Liu, M.G., Koga, K., Guo, Y.Y., Kang, S.J., Collingridge, G.L., Kaang, B.K., Zhao, M.G.,
Kheirbek, M.A., Britt, J.P., Beeler, J.A., Ishikawa, Y., McGehee, D.S., Zhuang, X., 2009. Zhuo, M., 2013. Long-term depression of synaptic transmission in the adult mouse
Adenylyl cyclase type 5 contributes to corticostriatal plasticity and striatum- insular cortex in vitro. Eur. J. Neurosci. 38, 3128–3145.
dependent learning. J. Neurosci. 29, 12115–12124. Liu, X., Chen, Y., Vickstrom, C.R., Li, Y., Viader, A., Cravatt, B.F., Liu, Q.S., 2016.
Khlaifia, A., Farah, H., Gackiere, F., Tell, F., 2013. Anandamide, cannabinoid type 1 Coordinated regulation of endocannabinoid-mediated retrograde synaptic
receptor, and NMDA receptor activation mediate non-Hebbian presynaptically suppression in the cerebellum by neuronal and astrocytic monoacylglycerol lipase.
expressed long-term depression at the first central synapse for visceral afferent Sci. Rep. 6, 35829.
fibers. J. Neurosci. 33, 12627–12637. Llano, I., Leresche, N., Marty, A., 1991. Calcium entry increases the sensitivity of
Kim, E., Sheng, M., 2004. PDZ domain proteins of synapses. Nat. Rev. Neurosci. 5, cerebellar Purkinje cells to applied GABA and decreases inhibitory synaptic currents.
771–781. Neuron 6, 565–574.
Kim, H.H., Park, J.M., Lee, S.H., Ho, W.K., 2019. Association of mGluR-dependent LTD of
excitatory synapses with endocannabinoid-dependent LTD of inhibitory synapses

19
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

Lonart, G., Schoch, S., Kaeser, P.S., Larkin, C.J., Sudhof, T.C., Linden, D.J., 2003. Mathur, B.N., Tanahira, C., Tamamaki, N., Lovinger, D.M., 2013. Voltage drives diverse
Phosphorylation of RIM1alpha by PKA triggers presynaptic long-term potentiation at endocannabinoid signals to mediate striatal microcircuit-specific plasticity. Nat.
cerebellar parallel fiber synapses. Cell 115, 49–60. Neurosci. 16, 1275–1283.
Longaretti, A., Forastieri, C., Gabaglio, M., Rubino, T., Battaglioli, E., Rusconi, F., 2020. Mato, S., Lafourcade, M., Robbe, D., Bakiri, Y., Manzoni, O.J., 2008. Role of the cyclic-
Termination of acute stress response by the endocannabinoid system is regulated AMP/PKA cascade and of P/Q-type Ca++ channels in endocannabinoid-mediated
through lysine-specific demethylase 1-mediated transcriptional repression of 2-AG long-term depression in the nucleus accumbens. Neuropharmacology 54, 87–94.
hydrolases ABHD6 and MAGL. J. Neurochem. 155, 98–110. Melis, M., Greco, B., Tonini, R., 2014. Interplay between synaptic endocannabinoid
Losonczy, A., Somogyi, P., Nusser, Z., 2003. Reduction of excitatory postsynaptic signaling and metaplasticity in neuronal circuit function and dysfunction. Eur. J.
responses by persistently active metabotropic glutamate receptors in the Neurosci. 39, 1189–1201.
hippocampus. J. Neurophysiol. 89, 1910–1919. Melis, M., Perra, S., Muntoni, A.L., Pillolla, G., Lutz, B., Marsicano, G., Di Marzo, V.,
Lourenço, J., Matias, I., Marsicano, G., Mulle, C., 2011. Pharmacological activation of Gessa, G.L., Pistis, M., 2004a. Prefrontal cortex stimulation induces 2-arachidonoyl-
kainate receptors drives endocannabinoid mobilization. J. Neurosci. 31, 3243–3248. glycerol-mediated suppression of excitation in dopamine neurons. J. Neurosci. 24,
Lovinger, D.M., 2008. Presynaptic modulation by endocannabinoids. Handb. Exp. 10707–10715.
Pharmacol. 435–477. Melis, M., Pistis, M., Perra, S., Muntoni, A.L., Pillolla, G., Gessa, G.L., 2004b.
Lu, D., Immadi, S.S., Wu, Z., Kendall, D.A., 2019. Translational potential of allosteric Endocannabinoids mediate presynaptic inhibition of glutamatergic transmission in
modulators targeting the cannabinoid CB1 receptor. Acta Pharmacol. Sin. 40, rat ventral tegmental area dopamine neurons through activation of CB1 receptors.
324–335. J. Neurosci. 24, 53–62.
Luo, Y., Yu, Y., Zhang, M., He, H., Fan, N., 2020. Chronic administration of ketamine Menon, V., Musial, T.F., Liu, A., Katz, Y., Kath, W.L., Spruston, N., Nicholson, D.A., 2013.
induces cognitive deterioration by restraining synaptic signaling. Mol. Psychiatr. Balanced synaptic impact via distance-dependent synapse distribution and
https://doi.org/10.1038/s41380-020-0793-6. In press. complementary expression of AMPARs and NMDARs in hippocampal dendrites.
Maccarrone, M., Bari, M., Di Rienzo, M., Finazzi-Agro, A., Rossi, A., 2003. Progesterone Neuron 80, 1451–1463.
activates fatty acid amide hydrolase (FAAH) promoter in human T lymphocytes Mikasova, L., Groc, L., Choquet, D., Manzoni, O.J., 2008. Altered surface trafficking of
through the transcription factor Ikaros. Evidence for a synergistic effect of leptin. presynaptic cannabinoid type 1 receptor in and out synaptic terminals parallels
J. Biol. Chem. 278, 32726–32732. receptor desensitization. Proc. Natl. Acad. Sci. U. S. A. 105, 18596–18601.
Madeo, G., Schirinzi, T., Maltese, M., Martella, G., Rapino, C., Fezza, F., Mastrangelo, N., Miller, L.K., Devi, L.A., 2011. The highs and lows of cannabinoid receptor expression in
Bonsi, P., Maccarrone, M., Pisani, A., 2016. Dopamine-dependent CB1 receptor disease: mechanisms and their therapeutic implications. Pharmacol. Rev. 63,
dysfunction at corticostriatal synapses in homozygous PINK1 knockout mice. 461–470.
Neuropharmacology 101, 460–470. Mitchell, V.A., Jeong, H.J., Drew, G.M., Vaughan, C.W., 2011. Cholecystokinin exerts an
Maejima, T., Hashimoto, K., Yoshida, T., Aiba, A., Kano, M., 2001. Presynaptic inhibition effect via the endocannabinoid system to inhibit GABAergic transmission in
caused by retrograde signal from metabotropic glutamate to cannabinoid receptors. midbrain periaqueductal gray. Neuropsychopharmacology 36, 1801–1810.
Neuron 31, 463–475. Mitchell, V.A., Kawahara, H., Vaughan, C.W., 2009. Neurotensin inhibition of GABAergic
Maejima, T., Oka, S., Hashimotodani, Y., Ohno-Shosaku, T., Aiba, A., Wu, D., Waku, K., transmission via mGluR-induced endocannabinoid signalling in rat periaqueductal
Sugiura, T., Kano, M., 2005. Synaptically driven endocannabinoid release requires grey. J. Physiol. 587, 2511–2520.
Ca2+-assisted metabotropic glutamate receptor subtype 1 to phospholipase Cbeta4 Monday, H.R., Bourdenx, M., Jordan, B.A., Castillo, P.E., 2020. CB1-receptor-mediated
signaling cascade in the cerebellum. J. Neurosci. 25, 6826–6835. inhibitory LTD triggers presynaptic remodeling via protein synthesis and
Magalhaes, A.C., Dunn, H., Ferguson, S.S., 2012. Regulation of GPCR activity, trafficking ubiquitination. Elife 9.
and localization by GPCR-interacting proteins. Br. J. Pharmacol. 165, 1717–1736. Monory, K., Polack, M., Remus, A., Lutz, B., Korte, M., 2015. Cannabinoid CB1 receptor
Maglio, L.E., Noriega-Prieto, J.A., Maraver, M.J., Fernandez de Sevilla, D., 2018. calibrates excitatory synaptic balance in the mouse hippocampus. J. Neurosci. 35,
Endocannabinoid-dependent long-term potentiation of synaptic transmission at rat 3842–3850.
barrel cortex. Cerebr. Cortex 28, 1568–1581. Moore, S.A., Nomikos, G.G., Dickason-Chesterfield, A.K., Schober, D.A., Schaus, J.M.,
Maison, P., Walker, D.J., Walsh, F.S., Williams, G., Doherty, P., 2009. BDNF regulates Ying, B.P., Xu, Y.C., Phebus, L., Simmons, R.M., Li, D., Iyengar, S., Felder, C.C., 2005.
neuronal sensitivity to endocannabinoids. Neurosci. Lett. 467, 90–94. Identification of a high-affinity binding site involved in the transport of
Makara, J.K., Mor, M., Fegley, D., Szabó, S.I., Kathuria, S., Astarita, G., Duranti, A., endocannabinoids. Proc. Natl. Acad. Sci. U. S. A. 102, 17852–17857.
Tontini, A., Tarzia, G., Rivara, S., Freund, T.F., Piomelli, D., 2005. Selective Morales, P., Jagerovic, N., 2016. Advances towards the discovery of GPR55 ligands. Curr.
inhibition of 2-AG hydrolysis enhances endocannabinoid signaling in hippocampus. Med. Chem. 23, 2087–2100.
Nat. Neurosci. 8, 1139–1141. Morisset, V., Ahluwalia, J., Nagy, I., Urban, L., 2001. Possible mechanisms of
Manz, K.M., Ghose, D., Turner, B.D., Taylor, A., Becker, J., Grueter, C.A., Grueter, B.A., cannabinoid-induced antinociception in the spinal cord. Eur. J. Pharmacol. 429,
2020. Calcium-permeable AMPA receptors promote endocannabinoid signaling at 93–100.
parvalbumin interneuron synapses in the nucleus accumbens core. Cell Rep. 32, Muller, C., Morales, P., Reggio, P.H., 2018. Cannabinoid ligands targeting TRP channels.
107971. Front. Mol. Neurosci. 11, 487.
Marcaggi, P., Attwell, D., 2005. Endocannabinoid signaling depends on the spatial Muller, C.S., Haupt, A., Bildl, W., Schindler, J., Knaus, H.G., Meissner, M., Rammner, B.,
pattern of synapse activation. Nat. Neurosci. 8, 776–781. Striessnig, J., Flockerzi, V., Fakler, B., Schulte, U., 2010. Quantitative proteomics of
Marcus, D.J., Bedse, G., Gaulden, A.D., Ryan, J.D., Kondev, V., Winters, N.D., Rosas- the Cav2 channel nano-environments in the mammalian brain. Proc. Natl. Acad. Sci.
Vidal, L.E., Altemus, M., Mackie, K., Lee, F.S., Delpire, E., Patel, S., 2020. U. S. A. 107, 14950–14957.
Endocannabinoid signaling collapse mediates stress-induced amygdalo-cortical Nakamura, Y., Dryanovski, D.I., Kimura, Y., Jackson, S.N., Woods, A.S., Yasui, Y., Tsai, S.
strengthening. Neuron 105, 1062–1076 e1066. Y., Patel, S., Covey, D.P., Su, T.P., Lupica, C.R., 2019. Cocaine-induced
Maren, S., Baudry, M., Thompson, R.F., 1991. Differential effects of ketamine and MK- endocannabinoid signaling mediated by sigma-1 receptors and extracellular vesicle
801 on the induction of long-term potentiation. Neuroreport 2, 239–242. secretion. Elife 8.
Marinelli, S., Di Marzo, V., Berretta, N., Matias, I., Maccarrone, M., Bernardi, G., Nanou, E., Catterall, W.A., 2018. Calcium channels, synaptic plasticity, and
Mercuri, N.B., 2003. Presynaptic facilitation of glutamatergic synapses to neuropsychiatric disease. Neuron 98, 466–481.
dopaminergic neurons of the rat substantia nigra by endogenous stimulation of Narushima, M., Hashimoto, K., Kano, M., 2006. Endocannabinoid-mediated short-term
vanilloid receptors. J. Neurosci. 23, 3136–3144. suppression of excitatory synaptic transmission to medium spiny neurons in the
Marinelli, S., Pacioni, S., Bisogno, T., Di Marzo, V., Prince, D.A., Huguenard, J.R., striatum. Neurosci. Res. 54, 159–164.
Bacci, A., 2008. The endocannabinoid 2-arachidonoylglycerol is responsible for the Narushima, M., Uchigashima, M., Fukaya, M., Matsui, M., Manabe, T., Hashimoto, K.,
slow self-inhibition in neocortical interneurons. J. Neurosci. 28, 13532–13541. Watanabe, M., Kano, M., 2007. Tonic enhancement of endocannabinoid-mediated
Marinelli, S., Pacioni, S., Cannich, A., Marsicano, G., Bacci, A., 2009. Self-modulation of retrograde suppression of inhibition by cholinergic interneuron activity in the
neocortical pyramidal neurons by endocannabinoids. Nat. Neurosci. 12, 1488–1490. striatum. J. Neurosci. 27, 496–506.
Marinelli, S., Vaughan, C.W., Christie, M.J., Connor, M., 2002. Capsaicin activation of Navakkode, S., Korte, M., 2014. Pharmacological activation of CB1 receptor modulates
glutamatergic synaptic transmission in the rat locus coeruleus in vitro. Journal of long term potentiation by interfering with protein synthesis. Neuropharmacology 79,
Physiology-London 543, 531–540. 525–533.
Marrs, W.R., Blankman, J.L., Horne, E.A., Thomazeau, A., Lin, Y.H., Coy, J., Bodor, A.L., Navarrete, M., Araque, A., 2008. Endocannabinoids mediate neuron-astrocyte
Muccioli, G.G., Hu, S.S., Woodruff, G., Fung, S., Lafourcade, M., Alexander, J.P., communication. Neuron 57, 883–893.
Long, J.Z., Li, W., Xu, C., Moller, T., Mackie, K., Manzoni, O.J., Cravatt, B.F., Navarrete, M., Araque, A., 2010. Endocannabinoids potentiate synaptic transmission
Stella, N., 2010. The serine hydrolase ABHD6 controls the accumulation and efficacy through stimulation of astrocytes. Neuron 68, 113–126.
of 2-AG at cannabinoid receptors. Nat. Neurosci. 13, 951–957. Neu, A., Földy, C., Soltesz, I., 2007. Postsynaptic origin of CB1-dependent tonic
Marshall, J.J., Xu, J., Contractor, A., 2018. Kainate receptors inhibit glutamate release inhibition of GABA release at cholecystokinin-positive basket cell to pyramidal cell
via mobilization of endocannabinoids in striatal direct pathway spiny projection synapses in the CA1 region of the rat hippocampus. J. Physiol. 578, 233–247.
neurons. J. Neurosci. 38, 3901–3910. Neuhofer, D., Lassalle, O., Manzoni, O.J., 2018. Muscarinic M1 receptor modulation of
Marsicano, G., Lutz, B., 1999. Expression of the cannabinoid receptor CB1 in distinct synaptic plasticity in nucleus accumbens of wild-type and fragile X mice. ACS Chem.
neuronal subpopulations in the adult mouse forebrain. Eur. J. Neurosci. 11, Neurosci. 9, 2233–2240.
4213–4225. Neuhofer, D., Spencer, S.M., Chioma, V.C., Beloate, L.N., Schwartz, D., Kalivas, P.W.,
Martin, H.G., Bernabeu, A., Lassalle, O., Bouille, C., Beurrier, C., Pelissier-Alicot, A.L., 2020. The loss of NMDAR-dependent LTD following cannabinoid self-administration
Manzoni, O.J., 2015. Endocannabinoids mediate muscarinic acetylcholine receptor- is restored by positive allosteric modulation of CB1 receptors. Addiction Biol. 25,
dependent long-term depression in the adult medial prefrontal cortex. Front. Cell. e12843.
Neurosci. 9, 457. Nevian, T., Sakmann, B., 2006. Spine Ca2+ signaling in spike-timing-dependent
plasticity. J. Neurosci. 26, 11001–11013.

20
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

Nicholson, D.A., Trana, R., Katz, Y., Kath, W.L., Spruston, N., Geinisman, Y., 2006. Psifogeorgou, K., Terzi, D., Papachatzaki, M.M., Varidaki, A., Ferguson, D., Gold, S.J.,
Distance-dependent differences in synapse number and AMPA receptor expression in Zachariou, V., 2011. A unique role of RGS9-2 in the striatum as a positive or negative
hippocampal CA1 pyramidal neurons. Neuron 50, 431–442. regulator of opiate analgesia. J. Neurosci. 31, 5617–5624.
Nicolussi, S., Gertsch, J., 2015. Endocannabinoid transport revisited. Vitam. Horm. 98, Ramirez-Franco, J., Bartolome-Martin, D., Alonso, B., Torres, M., Sanchez-Prieto, J.,
441–485. 2014. Cannabinoid type 1 receptors transiently silence glutamatergic nerve
Niehaus, J.L., Liu, Y., Wallis, K.T., Egertova, M., Bhartur, S.G., Mukhopadhyay, S., Shi, S., terminals of cultured cerebellar granule cells. PloS One 9, e88594.
He, H., Selley, D.E., Howlett, A.C., Elphick, M.R., Lewis, D.L., 2007. CB1 cannabinoid Robbe, D., Alonso, G., Manzoni, O.J., 2003. Exogenous and endogenous cannabinoids
receptor activity is modulated by the cannabinoid receptor interacting protein CRIP control synaptic transmission in mice nucleus accumbens. Ann. N. Y. Acad. Sci.
1a. Mol. Pharmacol. 72, 1557–1566. 1003, 212–225.
Njoo, C., Agarwal, N., Lutz, B., Kuner, R., 2015. The cannabinoid receptor CB1 interacts Robbe, D., Kopf, M., Remaury, A., Bockaert, J., Manzoni, O.J., 2002. Endogenous
with the WAVE1 complex and plays a role in actin dynamics and structural plasticity cannabinoids mediate long-term synaptic depression in the nucleus accumbens.
in neurons. PLoS Biol. 13, e1002286. Proc. Natl. Acad. Sci. U. S. A. 99, 8384–8388.
Nosov, G., Kahms, M., Klingauf, J., 2020. The decade of super-resolution microscopy of Roloff, A.M., Anderson, G.R., Martemyanov, K.A., Thayer, S.A., 2010. Homer 1a gates
the presynapse. Front. Synaptic Neurosci. 12, 32. the induction mechanism for endocannabinoid-mediated synaptic plasticity.
Ohno-Shosaku, T., Hashimotodani, Y., Ano, M., Takeda, S., Tsubokawa, H., Kano, M., J. Neurosci. 30, 3072–3081.
2007. Endocannabinoid signalling triggered by NMDA receptor-mediated calcium Roloff, A.M., Thayer, S.A., 2009. Modulation of excitatory synaptic transmission by Delta
entry into rat hippocampal neurons. J. Physiol. 584, 407–418. 9-tetrahydrocannabinol switches from agonist to antagonist depending on firing
Ohno-Shosaku, T., Kano, M., 2014. Endocannabinoid-mediated retrograde modulation of rate. Mol. Pharmacol. 75, 892–900.
synaptic transmission. Curr. Opin. Neurobiol. 29, 1–8. Ronesi, J., Gerdeman, G.L., Lovinger, D.M., 2004. Disruption of endocannabinoid release
Ohno-Shosaku, T., Maejima, T., Kano, M., 2001. Endogenous cannabinoids mediate and striatal long-term depression by postsynaptic blockade of endocannabinoid
retrograde signals from depolarized postsynaptic neurons to presynaptic terminals. membrane transport. J. Neurosci. 24, 1673–1679.
Neuron 29, 729–738. Ronesi, J.A., Collins, K.A., Hays, S.A., Tsai, N.P., Guo, W., Birnbaum, S.G., Hu, J.H.,
Ohno-Shosaku, T., Matsui, M., Fukudome, Y., Shosaku, J., Tsubokawa, H., Taketo, M.M., Worley, P.F., Gibson, J.R., Huber, K.M., 2012. Disrupted Homer scaffolds mediate
Manabe, T., Kano, M., 2003. Postsynaptic M-1 and M-3 receptors are responsible for abnormal mGluR5 function in a mouse model of fragile X syndrome. Nat. Neurosci.
the muscarinic enhancement of retrograde endocannabinoid signalling in the 15, 431–440. S431.
hippocampus. Eur. J. Neurosci. 18, 109–116. Rozenfeld, R., Devi, L.A., 2008. Regulation of CB1 cannabinoid receptor trafficking by
Ohno-Shosaku, T., Shosaku, J., Tsubokawa, H., Kano, M., 2002a. Cooperative the adaptor protein AP-3. Faseb. J. 22, 2311–2322.
endocannabinoid production by neuronal depolarization and group I metabotropic Sagar, D.R., Jhaveri, M.D., Richardson, D., Gray, R.A., de Lago, E., Fernandez-Ruiz, J.,
glutamate receptor activation. Eur. J. Neurosci. 15, 953–961. Barrett, D.A., Kendall, D.A., Chapman, V., 2010. Endocannabinoid regulation of
Ohno-Shosaku, T., Tsubokawa, H., Mizushima, I., Yoneda, N., Zimmer, A., Kano, M., spinal nociceptive processing in a model of neuropathic pain. Eur. J. Neurosci. 31,
2002b. Presynaptic cannabinoid sensitivity is a major determinant of depolarization- 1414–1422.
induced retrograde suppression at hippocampal synapses. J. Neurosci. 22, Salami, M., Fathollahi, Y., Esteky, H., Motamedi, F., Atapour, N., 2000. Effects of
3864–3872. ketamine on synaptic transmission and long-term potentiation in layer II/III of rat
Oliet, S.H., Baimoukhametova, D.V., Piet, R., Bains, J.S., 2007. Retrograde regulation of visual cortex in vitro. Eur J Pharmacol 390, 287–293.
GABA transmission by the tonic release of oxytocin and endocannabinoids governs Sang, N., Zhang, J., Chen, C., 2010. Anandamide potentiation of miniature spontaneous
postsynaptic firing. J. Neurosci. 27, 1325–1333. excitatory synaptic transmission is mediated via IP3 pathway. Neurochem. Int. 56,
Oliveria, S.F., Dittmer, P.J., Youn, D.H., Dell’Acqua, M.L., Sather, W.A., 2012. Localized 590–596.
calcineurin confers Ca2+-dependent inactivation on neuronal L-type Ca2+ channels. Schlicker, E., Kathmann, M., 2001. Modulation of transmitter release via presynaptic
J. Neurosci. 32, 15328–15337. cannabinoid receptors. Trends Pharmacol. Sci. 22, 565–572.
Olson, P.A., Tkatch, T., Hernandez-Lopez, S., Ulrich, S., Ilijic, E., Mugnaini, E., Zhang, H., Schlosburg, J.E., Blankman, J.L., Long, J.Z., Nomura, D.K., Pan, B., Kinsey, S.G.,
Bezprozvanny, I., Surmeier, D.J., 2005. G-protein-coupled receptor modulation of Nguyen, P.T., Ramesh, D., Booker, L., Burston, J.J., Thomas, E.A., Selley, D.E., Sim-
striatal CaV1.3 L-type Ca2+ channels is dependent on a Shank-binding domain. Selley, L.J., Liu, Q.S., Lichtman, A.H., Cravatt, B.F., 2010. Chronic monoacylglycerol
J. Neurosci. 25, 1050–1062. lipase blockade causes functional antagonism of the endocannabinoid system. Nat.
Pamplona, F.A., Ferreira, J., Menezes de Lima Jr., O., Duarte, F.S., Bento, A.F., Forner, S., Neurosci. 13, 1113–1119.
Villarinho, J.G., Bellocchio, L., Wotjak, C.T., Lerner, R., Monory, K., Lutz, B., Schmandke, A., Schmandke, A., Strittmatter, S.M., 2007. ROCK and Rho: biochemistry
Canetti, C., Matias, I., Calixto, J.B., Marsicano, G., Guimaraes, M.Z., Takahashi, R.N., and neuronal functions of Rho-associated protein kinases. Neuroscientist 13,
2012. Anti-inflammatory lipoxin A4 is an endogenous allosteric enhancer of CB1 454–469.
cannabinoid receptor. Proc. Natl. Acad. Sci. U. S. A. 109, 21134–21139. Schmitz, S.K., King, C., Kortleven, C., Huson, V., Kroon, T., Kevenaar, J.T., Schut, D.,
Pan, B., Wang, W., Zhong, P., Blankman, J.L., Cravatt, B.F., Liu, Q.S., 2011. Alterations of Saarloos, I., Hoetjes, J.P., de Wit, H., Stiedl, O., Spijker, S., Li, K.W., Mansvelder, H.
endocannabinoid signaling, synaptic plasticity, learning, and memory in D., Smit, A.B., Cornelisse, L.N., Verhage, M., Toonen, R.F., 2016. Presynaptic
monoacylglycerol lipase knock-out mice. J. Neurosci. 31, 13420–13430. inhibition upon CB1 or mGlu2/3 receptor activation requires ERK/MAPK
Pan, E., Zhao, Z., McNamara, J.O., 2019. LTD at mossy fiber synapses onto stratum phosphorylation of Munc18-1. EMBO J. 35, 1236–1250.
lucidum interneurons requires TrkB and retrograde endocannabinoid signaling. Schurman, L.D., Carper, M.C., Moncayo, L.V., Ogasawara, D., Richardson, K., Yu, L.,
J. Neurophysiol. 121, 609–619. Liu, X., Poklis, J.L., Liu, Q.S., Cravatt, B.F., Lichtman, A.H., 2019. Diacylglycerol
Park, H., Rhee, J., Lee, S., Chung, C., 2017. Selectively impaired endocannabinoid- lipase-alpha regulates hippocampal-dependent learning and memory processes in
dependent long-term depression in the lateral habenula in an animal model of mice. J. Neurosci. 39, 5949–5965.
depression. Cell Rep. 20, 289–296. Selvam, R., Yeh, M.L., Levine, E.S., 2018. Endogenous cannabinoids mediate the effect of
Pawlak, V., Kerr, J.N., 2008. Dopamine receptor activation is required for corticostriatal BDNF at CA1 inhibitory synapses in the hippocampus. Synapse, e22075.
spike-timing-dependent plasticity. J. Neurosci. 28, 2435–2446. Senst, L., Bains, J., 2014. Neuromodulators, stress and plasticity: a role for
Penasco, S., Rico-Barrio, I., Puente, N., Gomez-Urquijo, S.M., Fontaine, C.J., Egana- endocannabinoid signalling. J. Exp. Biol. 217, 102–108.
Huguet, J., Achicallende, S., Ramos, A., Reguero, L., Elezgarai, I., Nahirney, P.C., Sheinin, A., Talani, G., Davis, M.I., Lovinger, D.M., 2008. Endocannabinoid- and
Christie, B.R., Grandes, P., 2019. Endocannabinoid long-term depression revealed at mGluR5-dependent short-term synaptic depression in an isolated neuron/bouton
medial perforant path excitatory synapses in the dentate gyrus. Neuropharmacology preparation from the hippocampal CA1 region. J. Neurophysiol. 100, 1041–1052.
153, 32–40. Shen, M., Piser, T.M., Seybold, V.S., Thayer, S.A., 1996. Cannabinoid receptor agonists
Pertwee, R.G., 2015. Endocannabinoids and their pharmacological actions. Handb. Exp. inhibit glutamatergic synaptic transmission in rat hippocampal cultures. J. Neurosci.
Pharmacol. 231, 1–37. 16, 4322–4334.
Peterfi, Z., Urban, G.M., Papp, O.I., Nemeth, B., Monyer, H., Szabo, G., Erdelyi, F., Shen, M., Thayer, S.A., 1999. Delta9-tetrahydrocannabinol acts as a partial agonist to
Mackie, K., Freund, T.F., Hajos, N., Katona, I., 2012. Endocannabinoid-mediated modulate glutamatergic synaptic transmission between rat hippocampal neurons in
long-term depression of afferent excitatory synapses in hippocampal pyramidal cells culture. Mol. Pharmacol. 55, 8–13.
and GABAergic interneurons. J. Neurosci. 32, 14448–14463. Sheng, M., Kim, E., 2011. The postsynaptic organization of synapses. Cold Spring Harb
Peterson, B.Z., DeMaria, C.D., Adelman, J.P., Yue, D.T., 1999. Calmodulin is the Ca2+ Perspect Biol 3.
sensor for Ca2+ -dependent inactivation of L-type calcium channels. Neuron 22, Sheng, Z.H., 2014. Mitochondrial trafficking and anchoring in neurons: new insight and
549–558. implications. J. Cell Biol. 204, 1087–1098.
Piette, C., Cui, Y., Gervasi, N., Venance, L., 2020. Lights on endocannabinoid-mediated Shin, S.M., Zhang, N., Hansen, J., Gerges, N.Z., Pak, D.T., Sheng, M., Lee, S.H., 2012.
synaptic potentiation. Front. Mol. Neurosci. 13, 132. GKAP orchestrates activity-dependent postsynaptic protein remodeling and
Piomelli, D., 2003. The molecular logic of endocannabinoid signalling. Nat. Rev. homeostatic scaling. Nat. Neurosci. 15, 1655–1666.
Neurosci. 4, 873–884. Shonesy, B.C., Parrish, W.P., Haddad, H.K., Stephenson, J.R., Baldi, R., Bluett, R.J.,
Piomelli, D., 2014. More surprises lying ahead. The endocannabinoids keep us guessing. Marks, C.R., Centanni, S.W., Folkes, O.M., Spiess, K., Augustin, S.M., Mackie, K.,
Neuropharmacology 76 Pt B, 228–234. Lovinger, D.M., Winder, D.G., Patel, S., Colbran, R.J., 2018. Role of striatal direct
Piscitelli, F., Di Marzo, V., 2012. Redundancy" of endocannabinoid inactivation: new pathway 2-arachidonoylglycerol signaling in sociability and repetitive behavior.
challenges and opportunities for pain control. ACS Chem. Neurosci. 3, 356–363. Biol. Psychiatr. 84, 304–315.
Pitler, T.A., Alger, B.E., 1992. Postsynaptic spike firing reduces synaptic GABAA Shonesy, B.C., Stephenson, J.R., Marks, C.R., Colbran, R.J., 2020. Cyclic AMP-dependent
responses in hippocampal pyramidal cells. J. Neurosci. 12, 4122–4132. protein kinase and D1 dopamine receptors regulate diacylglycerol lipase-alpha and
Prather, P.L., Martin, N.A., Breivogel, C.S., Childers, S.R., 2000. Activation of synaptic 2-arachidonoyl glycerol signaling. J. Neurochem. 153, 334–345.
cannabinoid receptors in rat brain by WIN 55212-2 produces coupling to multiple G Shonesy, B.C., Wang, X., Rose, K.L., Ramikie, T.S., Cavener, V.S., Rentz, T.,
protein alpha-subunits with different potencies. Mol. Pharmacol. 57, 1000–1010. Baucum 2nd, A.J., Jalan-Sakrikar, N., Mackie, K., Winder, D.G., Patel, S., Colbran, R.

21
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

J., 2013. CaMKII regulates diacylglycerol lipase-alpha and striatal endocannabinoid Homer and PSD-95 complexes by the Shank family of postsynaptic density proteins.
signaling. Nat. Neurosci. 16, 456–463. Neuron 23, 583–592.
Singla, S., Kreitzer, A.C., Malenka, R.C., 2007. Mechanisms for synapse specificity during Tung, L.W., Lu, G.L., Lee, Y.H., Yu, L., Lee, H.J., Leishman, E., Bradshaw, H., Hwang, L.
striatal long-term depression. J. Neurosci. 27, 5260–5264. L., Hung, M.S., Mackie, K., Zimmer, A., Chiou, L.C., 2016. Orexins contribute to
Sjostrom, P.J., Turrigiano, G.G., Nelson, S.B., 2003. Neocortical LTD via coincident restraint stress-induced cocaine relapse by endocannabinoid-mediated disinhibition
activation of presynaptic NMDA and cannabinoid receptors. Neuron 39, 641–654. of dopaminergic neurons. Nat. Commun. 7, 12199.
Skeberdis, V.A., Chevaleyre, V., Lau, C.G., Goldberg, J.H., Pettit, D.L., Suadicani, S.O., Turrigiano, G.G., 2008. The self-tuning neuron: synaptic scaling of excitatory synapses.
Lin, Y., Bennett, M.V., Yuste, R., Castillo, P.E., Zukin, R.S., 2006. Protein kinase A Cell 135, 422–435.
regulates calcium permeability of NMDA receptors. Nat. Neurosci. 9, 501–510. Turu, G., Hunyady, L., 2010. Signal transduction of the CB1 cannabinoid receptor.
Soler-Llavina, G.J., Sabatini, B.L., 2006. Synapse-specific plasticity and J. Mol. Endocrinol. 44, 75–85.
compartmentalized signaling in cerebellar stellate cells. Nat. Neurosci. 9, 798–806. Twitchell, W., Brown, S., Mackie, K., 1997. Cannabinoids inhibit N- and P/Q-type
Song, C., Anderson, G.R., Sutton, L.P., Dao, M., Martemyanov, K.A., 2018. Selective role calcium channels in cultured rat hippocampal neurons. J. Neurophysiol. 78, 43–50.
of RGS9-2 in regulating retrograde synaptic signaling of indirect pathway medium Tzounopoulos, T., Rubio, M.E., Keen, J.E., Trussell, L.O., 2007. Coactivation of pre- and
spiny neurons in dorsal striatum. J. Neurosci. 38, 7120–7131. postsynaptic signaling mechanisms determines cell-specific spike-timing-dependent
Song, Y., Zhang, J., Chen, C., 2015. Fine-tuning of synaptic upscaling at excitatory plasticity. Neuron 54, 291–301.
synapses by endocannabinoid signaling is mediated via the CB1 receptor. Sci. Rep. 5, Uchigashima, M., Narushima, M., Fukaya, M., Katona, I., Kano, M., Watanabe, M., 2007.
16257. Subcellular arrangement of molecules for 2-arachidonoyl-glycerol-mediated
Steindel, F., Lerner, R., Haring, M., Ruehle, S., Marsicano, G., Lutz, B., Monory, K., 2013. retrograde signaling and its physiological contribution to synaptic modulation in the
Neuron-type specific cannabinoid-mediated G protein signalling in mouse striatum. J. Neurosci. 27, 3663–3676.
hippocampus. J. Neurochem. 124, 795–807. Vandevoorde, S., Fowler, C.J., 2005. Inhibition of fatty acid amide hydrolase and
Straiker, A., Dvorakova, M., Zimmowitch, A., Mackie, K., 2018. Cannabidiol inhibits monoacylglycerol lipase by the anandamide uptake inhibitor VDM11: evidence that
endocannabinoid signaling in autaptic hippocampal neurons. Mol. Pharmacol. 94, VDM11 acts as an FAAH substrate. Br. J. Pharmacol. 145, 885–893.
743–748. Varma, N., Brager, D.H., Morishita, W., Lenz, R.A., London, B., Alger, B., 2002.
Straiker, A., Hu, S.S., Long, J.Z., Arnold, A., Wager-Miller, J., Cravatt, B.F., Mackie, K., Presynaptic factors in the regulation of DST expression in hippocampus.
2009. Monoacylglycerol lipase limits the duration of endocannabinoid-mediated Neuropharmacology 43, 550–562.
depolarization-induced suppression of excitation in autaptic hippocampal neurons. Varma, N., Carlson, G.C., Ledent, C., Alger, B.E., 2001. Metabotropic glutamate receptors
Mol. Pharmacol. 76, 1220–1227. drive the endocannabinoid system in hippocampus. J. Neurosci. 21, RC188.
Straiker, A., Mackie, K., 2005. Depolarization-induced suppression of excitation in Vasquez, C., Lewis, D.L., 1999. The CB1 cannabinoid receptor can sequester G-proteins,
murine autaptic hippocampal neurones. J. Physiol. 569, 501–517. making them unavailable to couple to other receptors. J. Neurosci. 19, 9271–9280.
Straiker, A., Mitjavila, J., Yin, D., Gibson, A., Mackie, K., 2015. Aiming for allosterism: Vaughan, C.W., Connor, M., Bagley, E.E., Christie, M.J., 2000. Actions of cannabinoids
evaluation of allosteric modulators of CB1 in a neuronal model. Pharmacol. Res. 99, on membrane properties and synaptic transmission in rat periaqueductal gray
370–376. neurons in vitro. Mol. Pharmacol. 57, 288–295.
Straiker, A., Wager-Miller, J., Hu, S.S., Blankman, J.L., Cravatt, B.F., Mackie, K., 2011. Viader, A., Blankman, J.L., Zhong, P., Liu, X., Schlosburg, J.E., Joslyn, C.M., Liu, Q.S.,
COX-2 and fatty acid amide hydrolase can regulate the time course of depolarization- Tomarchio, A.J., Lichtman, A.H., Selley, D.E., Sim-Selley, L.J., Cravatt, B.F., 2015.
induced suppression of excitation. Br. J. Pharmacol. 164, 1672–1683. Metabolic interplay between astrocytes and neurons regulates endocannabinoid
Stringer, J.L., Guyenet, P.G., 1983. Elimination of long-term potentiation in the action. Cell Rep. 12, 798–808.
hippocampus by phencyclidine and ketamine. Brain Res. 258, 159–164. Vickstrom, C.R., Liu, X., Liu, S., Hu, M.M., Mu, L., Hu, Y., Yu, H., Love, S.L., Hillard, C.J.,
Su, S.C., Seo, J., Pan, J.Q., Samuels, B.A., Rudenko, A., Ericsson, M., Neve, R.L., Yue, D. Liu, Q.S., 2020. Role of endocannabinoid signaling in a septohabenular pathway in
T., Tsai, L.H., 2012. Regulation of N-type voltage-gated calcium channels and the regulation of anxiety- and depressive-like behavior. Mol. Psychiatr. https://doi.
presynaptic function by cyclin-dependent kinase 5. Neuron 75, 675–687. org/10.1038/s41380-020-00905-1. In press.
Sugaya, Y., Kano, M., 2018. Control of excessive neural circuit excitability and Vidal, G.S., Djurisic, M., Brown, K., Sapp, R.W., Shatz, C.J., 2016. Cell-autonomous
prevention of epileptic seizures by endocannabinoid signaling. Cell. Mol. Life Sci. 75, regulation of dendritic spine density by PirB. eNeuro 3.
2793–2811. Wallmichrath, I., Szabo, B., 2002. Cannabinoids inhibit striatonigral GABAergic
Sullivan, J.M., 1999. Mechanisms of cannabinoid-receptor-mediated inhibition of neurotransmission in the mouse. Neuroscience 113, 671–682.
synaptic transmission in cultured hippocampal pyramidal neurons. J. Neurophysiol. Wang, D.J., Yang, D., Su, L.D., Xie, Y.J., Zhou, L., Sun, C.L., Wang, Y., Wang, X.X.,
82, 1286–1294. Zhou, L., Shen, Y., 2012. Cytosolic phospholipase A2 alpha/arachidonic acid
Szabo, B., Urbanski, M.J., Bisogno, T., Di Marzo, V., Mendiguren, A., Baer, W.U., signaling mediates depolarization-induced suppression of excitation in the
Freiman, I., 2006. Depolarization-induced retrograde synaptic inhibition in the cerebellum. PloS One 7, e41499.
mouse cerebellar cortex is mediated by 2-arachidonoylglycerol. J. Physiol. 577, Wang, W., Jia, Y., Pham, D.T., Palmer, L.C., Jung, K.M., Cox, C.D., Rumbaugh, G.,
263–280. Piomelli, D., Gall, C.M., Lynch, G., 2018. Atypical endocannabinoid signaling
Tabatadze, N., Huang, G., May, R.M., Jain, A., Woolley, C.S., 2015. Sex differences in initiates a new form of memory-related plasticity at a cortical input to Hippocampus.
molecular signaling at inhibitory synapses in the Hippocampus. J. Neurosci. 35, Cerebr. Cortex 28, 2253–2266.
11252–11265. Wang, W., Trieu, B.H., Palmer, L.C., Jia, Y., Pham, D.T., Jung, K.M., Karsten, C.A.,
Tang, A.H., Alger, B.E., 2015. Homer protein-metabotropic glutamate receptor binding Merrill, C.B., Mackie, K., Gall, C.M., Piomelli, D., Lynch, G., 2016. A primary cortical
regulates endocannabinoid signaling and affects hyperexcitability in a mouse model input to Hippocampus expresses a pathway-specific and endocannabinoid-
of fragile X syndrome. J. Neurosci. 35, 3938–3945. dependent form of long-term potentiation. eNeuro 3.
Tanimura, A., Yamazaki, M., Hashimotodani, Y., Uchigashima, M., Kawata, S., Abe, M., Wang, Z., Kai, L., Day, M., Ronesi, J., Yin, H.H., Ding, J., Tkatch, T., Lovinger, D.M.,
Kita, Y., Hashimoto, K., Shimizu, T., Watanabe, M., Sakimura, K., Kano, M., 2010. Surmeier, D.J., 2006. Dopaminergic control of corticostriatal long-term synaptic
The endocannabinoid 2-arachidonoylglycerol produced by diacylglycerol lipase depression in medium spiny neurons is mediated by cholinergic interneurons.
alpha mediates retrograde suppression of synaptic transmission. Neuron 65, Neuron 50, 443–452.
320–327. Wiley, J.L., Marusich, J.A., Thomas, B.F., 2017. Combination chemistry: structure-
Terral, G., Varilh, M., Cannich, A., Massa, F., Ferreira, G., Marsicano, G., 2020. Synaptic activity relationships of novel psychoactive cannabinoids. Curr Top Behav Neurosci
functions of type-1 cannabinoid receptors in inhibitory circuits of the anterior 32, 231–248.
piriform cortex. Neuroscience 433, 121–131. Wilson, R.I., Nicoll, R.A., 2001. Endogenous cannabinoids mediate retrograde signalling
Ting, J.T., Peca, J., Feng, G., 2012. Functional consequences of mutations in postsynaptic at hippocampal synapses. Nature 410, 588–592.
scaffolding proteins and relevance to psychiatric disorders. Annu. Rev. Neurosci. 35, Wu, Y., Liu, Q., Guo, B., Ye, F., Ge, J., Xue, L., 2020. BDNF activates postsynaptic TrkB
49–71. receptors to induce endocannabinoid release and inhibit presynaptic calcium influx
Tomizawa, K., Ohta, J., Matsushita, M., Moriwaki, A., Li, S.T., Takei, K., Matsui, H., at a calyx-type synapse. J. Neurosci. 40, 8070–8087.
2002. Cdk5/p35 regulates neurotransmitter release through phosphorylation and Wu, Z.Z., Chen, S.R., Pan, H.L., 2005. Transient receptor potential vanilloid type 1
downregulation of P/Q-type voltage-dependent calcium channel activity. activation down-regulates voltage-gated calcium channels through calcium-
J. Neurosci. 22, 2590–2597. dependent calcineurin in sensory neurons. J. Biol. Chem. 280, 18142–18151.
Tong, J., Liu, X., Vickstrom, C., Li, Y., Yu, L., Lu, Y., Smrcka, A.V., Liu, Q.S., 2017. The Xu, J., Antion, M.D., Nomura, T., Kraniotis, S., Zhu, Y., Contractor, A., 2014.
epac-phospholipase cε pathway regulates endocannabinoid signaling and cocaine- Hippocampal metaplasticity is required for the formation of temporal associative
induced disinhibition of ventral tegmental area dopamine neurons. J. Neurosci. 37, memories. J. Neurosci. 34, 16762–16773.
3030–3044. Xu, J.Y., Zhang, J., Chen, C., 2012. Long-lasting potentiation of hippocampal synaptic
Tonnesen, J., Nagerl, U.V., 2013. Superresolution imaging for neuroscience. Exp. Neurol. transmission by direct cortical input is mediated via endocannabinoids. J. Physiol.
242, 33–40. 590, 2305–2315.
Tozzi, A., de Iure, A., Di Filippo, M., Tantucci, M., Costa, C., Borsini, F., Ghiglieri, V., Xu, W., Li, H., Wang, L., Zhang, J., Liu, C., Wan, X., Liu, X., Hu, Y., Fang, Q., Xiao, Y.,
Giampa, C., Fusco, F.R., Picconi, B., Calabresi, P., 2011. The distinct role of medium Bu, Q., Wang, H., Tian, J., Zhao, Y., Cen, X., 2020. Endocannabinoid signaling
spiny neurons and cholinergic interneurons in the D(2)/A(2)A receptor interaction regulates the reinforcing and psychostimulant effects of ketamine in mice. Nat.
in the striatum: implications for Parkinson’s disease. J. Neurosci. 31, 1850–1862. Commun. 11, 5962.
Trettel, J., Levine, E.S., 2003. Endocannabinoids mediate rapid retrograde signaling at Yakel, J.L., 1997. Calcineurin regulation of synaptic function: from ion channels to
interneuron -> pyramidal neuron synapses of the neocortex. J. Neurophysiol. 89, transmitter release and gene transcription. Trends Pharmacol. Sci. 18, 124–134.
2334–2338. Yang, K., Lei, G., Xie, Y.F., MacDonald, J.F., Jackson, M.F., 2014. Differential regulation
Tu, J.C., Xiao, B., Naisbitt, S., Yuan, J.P., Petralia, R.S., Brakeman, P., Doan, A., of NMDAR and NMDAR-mediated metaplasticity by anandamide and 2-AG in the
Aakalu, V.K., Lanahan, A.A., Sheng, M., Worley, P.F., 1999. Coupling of mGluR/ hippocampus. Hippocampus 24, 1601–1614.

22
B.L. Winters and C.W. Vaughan Neuropharmacology 197 (2021) 108736

Yang, Y., Calakos, N., 2013. Presynaptic long-term plasticity. Front. Synaptic Neurosci. 5, Zhang, H., Maximov, A., Fu, Y., Xu, F., Tang, T.S., Tkatch, T., Surmeier, D.J.,
8. Bezprozvanny, I., 2005. Association of CaV1.3 L-type calcium channels with Shank.
Yanovsky, Y., Mades, S., Misgeld, U., 2003. Retrograde signaling changes the venue of J. Neurosci. 25, 1037–1049.
postsynaptic inhibition in rat substantia nigra. Neuroscience 122, 317–328. Zhang, S.Y., Xu, M., Miao, Q.L., Poo, M.M., Zhang, X.H., 2009. Endocannabinoid-
Yasmin, F., Colangeli, R., Morena, M., Filipski, S., van der Stelt, M., Pittman, Q.J., dependent homeostatic regulation of inhibitory synapses by miniature excitatory
Hillard, C.J., Teskey, G.C., McEwen, B.S., Hill, M.N., Chattarji, S., 2020. Stress- synaptic activities. J. Neurosci. 29, 13222–13231.
induced modulation of endocannabinoid signaling leads to delayed strengthening of Zhao, L., Levine, E.S., 2014. BDNF-endocannabinoid interactions at neocortical
synaptic connectivity in the amygdala. Proc. Natl. Acad. Sci. U. S. A. 117, 650–655. inhibitory synapses require phospholipase C signaling. J. Neurophysiol. 111,
Yeh, M.L., Selvam, R., Levine, E.S., 2017. BDNF-induced endocannabinoid release 1008–1015.
modulates neocortical glutamatergic neurotransmission. Synapse 71. Zhao, L., Yeh, M.L., Levine, E.S., 2015. Role for endogenous BDNF in endocannabinoid-
Yin, H.H., Adermark, L., Lovinger, D.M., 2008. Neurotensin reduces glutamatergic mediated long-term depression at neocortical inhibitory synapses. eNeuro 2.
transmission in the dorsolateral striatum via retrograde endocannabinoid signaling. Zhao, Y., Tzounopoulos, T., 2011. Physiological activation of cholinergic inputs controls
Neuropharmacology 54, 79–86. associative synaptic plasticity via modulation of endocannabinoid signaling.
Yin, H.H., Davis, M.I., Ronesi, J.A., Lovinger, D.M., 2006. The role of protein synthesis in J. Neurosci. 31, 3158–3168.
striatal long-term depression. J. Neurosci. 26, 11811–11820. Zhong, P., Liu, Y., Hu, Y., Wang, T., Zhao, Y.P., Liu, Q.S., 2015. BDNF interacts with
Yin, H.H., Lovinger, D.M., 2006. Frequency-specific and D2 receptor-mediated inhibition endocannabinoids to regulate cocaine-induced synaptic plasticity in mouse midbrain
of glutamate release by retrograde endocannabinoid signaling. Proc. Natl. Acad. Sci. dopamine neurons. J. Neurosci. 35, 4469–4481.
U. S. A. 103, 8251–8256. Zhong, P., Pan, B., Gao, X.P., Blankman, J.L., Cravatt, B.F., Liu, Q.S., 2011. Genetic
Yin, L., Zheng, R., Ke, W., He, Q., Zhang, Y., Li, J., Wang, B., Mi, Z., Long, Y.S., Rasch, M. deletion of monoacylglycerol lipase alters endocannabinoid-mediated retrograde
J., Li, T., Luan, G., Shu, Y., 2018. Autapses enhance bursting and coincidence synaptic depression in the cerebellum. J. Physiol. 589, 4847–4855.
detection in neocortical pyramidal cells. Nat. Commun. 9, 4890. Zhu, J., Zhou, Q., Shang, Y., Li, H., Peng, M., Ke, X., Weng, Z., Zhang, R., Huang, X., Li, S.
Yoshida, T., Uchigashima, M., Yamasaki, M., Katona, I., Yamazaki, M., Sakimura, K., S.C., Feng, G., Lu, Y., Zhang, M., 2017. Synaptic targeting and function of SAPAPs
Kano, M., Yoshioka, M., Watanabe, M., 2011. Unique inhibitory synapse with mediated by phosphorylation-dependent binding to PSD-95 MAGUKs. Cell Rep. 21,
particularly rich endocannabinoid signaling machinery on pyramidal neurons in 3781–3793.
basal amygdaloid nucleus. Proc. Natl. Acad. Sci. U. S. A. 108, 3059–3064. Zhu, P.J., Lovinger, D.M., 2005. Retrograde endocannabinoid signaling in a postsynaptic
Yoshino, H., Miyamae, T., Hansen, G., Zambrowicz, B., Flynn, M., Pedicord, D., Blat, Y., neuron/synaptic bouton preparation from basolateral amygdala. J. Neurosci. 25,
Westphal, R.S., Zaczek, R., Lewis, D.A., Gonzalez-Burgos, G., 2011. Postsynaptic 6199–6207.
diacylglycerol lipase mediates retrograde endocannabinoid suppression of inhibition Zhu, P.J., Lovinger, D.M., 2007. Persistent synaptic activity produces long-lasting
in mouse prefrontal cortex. J. Physiol. 589, 4857–4884. enhancement of endocannabinoid modulation and alters long-term synaptic
Younts, T.J., Chevaleyre, V., Castillo, P.E., 2013. CA1 pyramidal cell theta-burst firing plasticity. J. Neurophysiol. 97, 4386–4389.
triggers endocannabinoid-mediated long-term depression at both somatic and Zimmermann, T., Bartsch, J.C., Beer, A., Lomazzo, E., Guggenhuber, S., Lange, M.D.,
dendritic inhibitory synapses. J. Neurosci. 33, 13743–13757. Bindila, L., Pape, H.C., Lutz, B., 2019. Impaired anandamide/palmitoylethanolamide
Younts, T.J., Monday, H.R., Dudok, B., Klein, M.E., Jordan, B.A., Katona, I., Castillo, P.E., signaling in hippocampal glutamatergic neurons alters synaptic plasticity, learning,
2016. Presynaptic protein synthesis is required for long-term plasticity of GABA and emotional responses. Neuropsychopharmacology 44, 1377–1388.
release. Neuron 92, 479–492. Ziv, N.E., Garner, C.C., 2004. Cellular and molecular mechanisms of presynaptic
Zamponi, G.W., Currie, K.P., 2013. Regulation of Ca(V)2 calcium channels by G protein assembly. Nat. Rev. Neurosci. 5, 385–399.
coupled receptors. Biochim. Biophys. Acta 1828, 1629–1643. Zuhlke, R.D., Pitt, G.S., Deisseroth, K., Tsien, R.W., Reuter, H., 1999. Calmodulin
Zeng, M., Shang, Y., Guo, T., He, Q., Yung, W.H., Liu, K., Zhang, M., 2016. A binding site supports both inactivation and facilitation of L-type calcium channels. Nature 399,
outside the canonical PDZ domain determines the specific interaction between 159–162.
Shank and SAPAP and their function. Proc. Natl. Acad. Sci. U. S. A. 113, Zygmunt, P.M., Chuang, H., Movahed, P., Julius, D., Hogestatt, E.D., 2000. The
E3081–E3090. anandamide transport inhibitor AM404 activates vanilloid receptors. Eur. J.
Pharmacol. 396, 39–42.

23

You might also like