You are on page 1of 13

Neuron

Review

Role of Dopamine Neurons in Reward and Aversion:


A Synaptic Plasticity Perspective
Marco Pignatelli1 and Antonello Bonci1,2,3,*
1Intramural Research Program, Synaptic Plasticity Section, National Institute on Drug Abuse, Baltimore, MD 21224, USA
2Solomon H. Snyder Neuroscience Institute
3Department of Psychiatry

Johns Hopkins University School of Medicine, Baltimore, MD 21205, USA


*Correspondence: antonello.bonci@nih.gov
http://dx.doi.org/10.1016/j.neuron.2015.04.015

The brain is wired to predict future outcomes. Experience-dependent plasticity at excitatory synapses within
dopamine neurons of the ventral tegmental area, a key region for a broad range of motivated behaviors, is
thought to be a fundamental cellular mechanism that enables adaptation to a dynamic environment. Thus,
depending on the circumstances, dopamine neurons are capable of processing both positive and negative
reinforcement learning strategies. In this review, we will discuss how changes in synaptic plasticity of dopa-
mine neurons may affect dopamine release, as well as behavioral adaptations to different environmental con-
ditions falling at opposite ends of a saliency spectrum ranging from reward to aversion.

The ventral tegmental area (VTA) is a heterogeneous midbrain In this review, we will focus on the significance of LTP and LTD
structure that plays a major role in regulating different adaptive within DA neurons of the VTA, with respect to neural processing
brain functions related to reward and motivation processing. and coding of both rewarding and aversive stimuli. First, we will
It is predominantly composed of dopaminergic neurons (55%– provide an overview of the role of DA in learning and memory
65%), while the rest are mainly GABAergic (30%) and a small processes that occur when organisms adapt their behavior on
portion are glutamatergic cells (5%) (Hnasko et al., 2012; Stef- the basis of associations with reward and aversion. Next, we
fensen et al., 1998). Here, we will focus on the main neuronal will present recent findings that shed light on the circuit connec-
population of the VTA: dopamine neurons. tivity and functional heterogeneity of the VTA, highlighting the
Dopamine (DA) release from the VTA efferent projections takes complexity and the diversity of a myriad of cellular regulators
place in a broad range of structures such as prefrontal cortex of reward and aversion. Finally, we will review the general rules
(PFC), nucleus accumbens (NAc), amygdala, and hippocampus governing synaptic plasticity of VTA DA neurons and the poten-
(Barrot, 2014). tial impact of a defined synaptic state in terms of DA release, or
One key feature of this dopaminergic pathway, known as the co-release capability and behavioral flexibility.
meso-cortico-limbic circuit, is its ability to reinforce natural
rewarding behavior and attribute motivational salience to other- Encoding and Modulation of Appetitive and Aversive
wise neutral environmental stimuli (Berridge and Robinson, Learning in VTA Dopamine Neurons
1998; Smith et al., 2011). Here, we propose that two forms of Dopamine and Goal-Directed Behaviors
synaptic plasticity, known as long-term potentiation (LTP) and DA is implicated in the execution of goal-directed behaviors
long-term depression (LTD), occurring onto DA neurons, repre- and in the regulation of reward-related vigor (Beierholm et al.,
sent a key fundamental mechanism capable of affecting reward 2013; Ranaldi, 2014). These concepts have emerged mostly
and aversion processes, as well as changes in motivational from behavioral studies showing that DA antagonists are
salience accomplished via effects on VTA DA release exerted capable of attenuating motivation to perform an action, even
in the target areas. before motor responses occur (Wise, 2004). Interestingly, condi-
LTP and LTD are widely recognized as cellular substrates tioned learning precedes and guides the performance of an
for learning and memory in the brain (Anggono and Huganir, instrumental act, as well as being DA-dependent (Darvas et al.,
2012; Bliss and Collingridge, 1993; Collingridge et al., 2010). 2014). In particular, fundamental goal-directed behaviors such
Since the discovery of synaptic plasticity, these forms of as seeking food or water are not innate, but learned (Changizi
cellular memory have been the subject of intensive investi- et al., 2002). It is widely believed that stimuli provided by the sur-
gation in several brain regions that support cognitive pro- rounding environment can be selectively reinforced and thus
cesses, such as hippocampus and cortex. The functional guide the proper execution of directed and motivated behavior,
significance of synaptic plasticity in the aforementioned struc- even for a naive organism (Hall and Mayer, 1975; Johanson and
tures has been extensively characterized (Bliss et al., 2014). Hall, 1979). Once stimulus-reward associations are formed, they
In contrast, the cellular events underlying synaptic plasticity can last long after the appropriate motivational drive states are
in DA neurons of the VTA, and the role of LTP and present (Balleine and Dickinson, 1992; Mendelson, 1966; Mor-
LTD in shaping DA-dependent behaviors, are still poorly gan, 1974). Nevertheless, behavioral effects to conditioned re-
understood. sponses are flexible and subject to extinction if paired with

Neuron 86, June 3, 2015 ª2015 Elsevier Inc. 1145


Neuron

Review

unrewarded trials (Wise et al., 1978) or if the experimental setting (Chen et al., 2008; Stuber et al., 2008; Ungless et al., 2001). Using
occurs without a supportive drive state (Mendelson, 1966; behavioral sensitization, a phenomenon that is based on esca-
Morgan, 1974). In general terms, consumption of reward (food, lating behavioral responses to repeated exposure to a drug, it
mating, and drugs of abuse) produces a hedonic state that is was discovered that N-methyl-D-aspartate receptor (NMDAR)
thought to initiate learning processes devoted to consolidate antagonists delivered into the VTA prevented the development
specific goals or cues that predict availability of a reward or of this behavioral phenomenon. In addition, alpha-amino-3-hy-
actions that permit its consumption (Wise and Kiyatkin, 2011). droxy-5-methyl-4-isoxazolepropionate receptor (AMPAR)
The overall concept is to promote a motivational state to allow blockade is capable of affecting the maintenance of behavioral
an organism to purposely select the best way to achieve the sensitization (Jackson et al., 2000; Wolf, 1998). Therefore, phar-
successful procurement of its essential needs. Based on this macological compounds administered within the VTA that
assumption, the brain seems to store and estimate the value of antagonize the activity of AMPARs and NMDARs severely
certain actions based on prior experience. In other words, an disrupt classical Pavlovian associative learning (Harris and As-
animal uses stored values to predict and plan actions, based ton-Jones, 2003; Kelley et al., 2003; Stuber et al., 2008).
on past events that involved a wide range of variables, including The functional interpretation of these findings opens up several
reward and aversion (Montague et al., 2004). possible cellular scenarios, including disruption of cellular pro-
Dopamine Neuron Activity and Prediction Error cesses related to synaptic plasticity, intrinsic properties, or burst
The brain is wired to predict future outcomes. This fundamental firing modalities of DA neurons. In fact, both LTP and the ability to
extraordinary feature occurs in response to a myriad of discrete generate burst firing rely on the integrity of NMDAR activity in
stimuli or situations, and it is mainly based on prior experience DA neurons of the VTA. It is tempting to speculate that in these
and specific patterns of response. However, when the outcome experimental settings a glutamatergic antagonist could affect
differs from what is expected, a ‘‘prediction error’’ occurs. Such both processes. Moreover, burst firing of DA neurons in the sub-
prediction error, mainly supported by dopaminergic activity, is stantia nigra can play a permissive role in the proper occurrence
generally used by the brain to refine and optimize its future re- of NMDA synaptic plasticity (Harnett et al., 2009). Presumably,
sponses, as well as learn new behavioral strategies for all needs this phenomenon might happen also for DA neurons of the VTA,
that are essential to ensure survival (Cohen et al., 2012; Holler- thus engaging the first fundamental step toward the expression
man and Schultz, 1998; Schultz, 2006; Schultz et al., 1993, of the AMPAR mediated LTP during cocaine exposure (Argilli
1997, 1998). et al., 2008), or potentially during the occurrence of natural
The neural language of DA neurons at resting condition ex- reward-related association.
hibits a consistent tonic pattern of firing. On top of this tonic ac- Aversion: The Other Face of Dopamine
tivity, a brief phasic burst of spike activity can appear and be DA neurons are capable of encoding both reward and aversion.
superimposed based on prior history of reward experience. De- There is solid evidence that several types of rewarding stimuli,
pending on the circumstances, these neurons are capable of like food (McCutcheon et al., 2012; Roitman et al., 2004), intrao-
coding three different functional states: a reward that is ‘‘as ral infusion of a sucrose based solution (Roitman et al., 2008),
expected,’’ ‘‘better than expected,’’ or ‘‘worse than expected.’’ drugs of abuse (Aragona et al., 2009; Stuber et al., 2005a,
Reward falling in the category of ‘‘as expected’’ produce tonic 2005b) and cues predicting their delivery (Daberkow et al.,
activity, reward that are ‘‘better than expected’’ result in phasic 2013; Phillips et al., 2003), all strongly, albeit transiently, increase
bursts signal, and a pause in firing parallels reward that are the probability of DA release events. Given that animals are
‘‘worse than expected.’’ Interestingly, a burst of action potentials willing to perform tasks that lead to the consumption of
is capable of releasing more DA in specific projection areas reward, but are generally reluctant to deal with aversive stimuli
than the same number of spikes organized in spaced action (Miller and Hunt, 1944), it will be crucial to unequivocally differen-
potentials do (Schultz, 1986). These cellular phenotypes recapit- tiate between opposite hedonic evaluations, like reward and
ulate the classic concept that, in mammals, single spikes aversion, as a direct effect of selective coding occurring in
firing in DA neurons are thought to play a permissive role in DA neurons.
initiating movements (Romo and Schultz, 1990), while burst There are several approaches that can be used to mimic aver-
firing is correlated with arousal states and motivation. sive situations in experimental animal models. The methodolo-
More importantly, mimicking a ‘‘better than expected’’ reward gies involve a variety of conditions ranging from acute or chronic
prediction error signal by optogenetic activation of DA neurons exposure to aversive stimuli (foot shocks, forced swimming test,
during the occurrence of reward delivery was sufficient to estab- hind paw injection of formalin, social defeat stress, fear condi-
lish a long-lasting increase in reward-seeking behavior caused tioning) to stereotypical orofacial responses evoked by intraor-
by cue-reward learned associations (Steinberg et al., 2013). ally delivered solutions (Berridge, 2000; Chaudhury et al., 2013;
Many studies have shown that burst firing activity is highly regu- Fadok et al., 2009; Friedman et al., 2014; Lammel et al., 2012;
lated by glutamatergic afferent inputs (Grace and Bunney, Tye et al., 2013). From these studies, a role for DA and DA neu-
1984a, 1984b; Grace et al., 2007). Not surprisingly, another rons, in the modulation and coding of aversive life experiences
substantial body of evidence supports a role for glutamate clearly emerges. Importantly, DA neurons are active players dur-
in learning mechanisms related to reward adaptive behaviors ing fear-related experiences (Fadok et al., 2010), facilitating the
(Kauer, 2004; Lüscher and Malenka, 2011; Schultz, 2011). early stabilization of the memory trace of fear-related learning.
In fact, glutamatergic afferents reaching the VTA are suscep- Additionally, VTA DA neurons undergo activity-dependent plas-
tible to plasticity during reward and drug-associated learning ticity mechanisms (Gore et al., 2014) and actively participate in

1146 Neuron 86, June 3, 2015 ª2015 Elsevier Inc.


Neuron

Review

the modulation of amygdala action potentials during Pavlovian (Lammel et al., 2011; Margolis et al., 2008). Using this approach,
fear conditioning and stress (Inglis and Moghaddam, 1999; it has been shown that VTA neurons projecting to the lateral shell
Rosenkranz and Grace, 2002). of NAc reside within the lateral portion of the VTA and exhibit
Along these lines, a mouse model carrying a genetic inactiva- large Ih currents, a hyperpolarization-activated cationic current,
tion of functional NMDA receptors on DA neurons displays a mal- which represents an electrophysiological feature traditionally
adaptive conditioning to a cue predicting an aversive outcome, used to identify DA neurons. In addition, a recent study found
and concurrently this impairment in contingency alertness is that glutamatergic synapses display a low AMPAR/NMDAR ratio
associated with the development of a generalized anxiety-like at resting state. Conversely, DA neurons projecting to the baso-
phenotype (Zweifel et al., 2011). Consistent with a direct involve- lateral amygdala, PFC, and core of the NAc originate from the
ment of DA neurons in the coding and modulation of emotional medial portion of the VTA. Moreover, these neurons do not
memory formation and adaptive behavior, bidirectional control show Ih currents, and the glutamatergic synaptic configuration
(via optogenetic inhibition or excitation) of genetically targeted displayed a high AMPAR/NMDAR ratio at resting conditions.
midbrain DA neurons rapidly modulates prototypical symptoms Interestingly, salient appetitive events, like a passive exposure
caused by chronic stress (Chaudhury et al., 2013; Friedman to cocaine, induced synaptic potentiation mostly within DA neu-
et al., 2014; Tye et al., 2013). The latter approaches can directly rons projecting to the lateral shell, while an aversive event (hind
test the involvement of the DA system in the behavioral control paw injection of formalin) was capable of inducing synaptic
of aversion. However, optogenetic manipulations, like the ones potentiation, mostly within PFC projecting neurons (Lammel
performed in the previous studies, can indeed rescue pheno- et al., 2011) (Figure 1).
typic alterations in animals that underwent behavioral manipula- These distinct subsets of neurons receive different afferent
tion, such as depressive-like phenotypes, but such rescue first synapses. NAc shell projection neurons receive excitatory affer-
requires the induction of a non-physiological behavior (Friedman ents, both glutamatergic and cholinergic inputs, from the latero-
et al., 2014; Tye et al., 2013). dorsal tegmentum (LDTg) and inhibitory afferents from the rostro-
In summary, to gain full knowledge about the exquisitely intri- medial tegmental nucleus (RMTg). In the case of dopaminergic
cate physiology of DA neurons, it will be important in the future neurons projecting to the PFC, the excitatory afferents are pro-
to compare reward and aversive responses within an ‘‘intact’’ vided by glutamatergic input emanating from the lateral habe-
mesolimbic system. Furthermore, the many groundbreaking nula. The lateral habenula also sends glutamatergic synapses
contributions, briefly discussed in this review, shed light toward onto GABAergic neurons of the RMTg that project back to NAc
a better understanding of how, and where, reward and aversion shell-projecting neurons. The behavioral implication of this intri-
processing takes place within the CNS. Scientists are not far cate circuit has been determined in vivo by using optogenetic
from giving a precise identity and role to the various neuronal stimulation of specific presynaptic terminals, allowing an exami-
sub-populations located within the VTA, that are engaged in nation of the functional significance of input specificity, relative to
the coding of reward, aversion, or in facilitating contingency the VTA. Optical stimulation of channelrhodopsin-2 expressing
awareness, with the ultimate goal to identify potential targets terminals in the VTA produced a robust conditioned place prefer-
for therapeutic intervention. ence (CPP), a form of Pavlovian conditioning used to measure the
motivational effects of experiences. On the other hand, light stim-
Cracking the Dopamine Code with New Approaches to ulation of the lateral habenula and RMTg terminals in the VTA pro-
Understand Neuronal Heterogeneity duced conditioned place aversion (Lammel et al., 2012).
DA neurons located within the VTA are clustered into specific Providing proper control of DA neuron firing rates in VTA is
anatomical niches and project to a single target region such as essential for the physiological expression of reward processing
nucleus accumbens, prefrontal cortex, hypothalamus, basolat- and motivational salience. The balance between excitatory and
eral amygdala, lateral habenula, pallidum, and bed nucleus inhibitory drive has emerged as a fundamental mechanism for
of stria terminalis (Haber, 2014). The VTA is a very heteroge- the control of firing rates and activity patterns of DA neurons.
neous region in which dopaminergic, GABAergic and gluta- Glutamatergic input via AMPA and NMDA receptors expressed
matergic neurons intermingle. As previously mentioned in the on DA neurons can shape DA release events. GABAergic synap-
introduction, dopaminergic neurons are the most abundant ses are also capable of regulating firing activity via activation of
population (65%), followed by GABA neurons (30%), while gluta- GABAA receptors located on DA neurons, thereby profoundly
matergic neurons (5%) are the most underrepresented popula- reducing their firing rate. Interestingly, there are several glutama-
tion (Morales and Root, 2014; Yamaguchi et al., 2007, 2011). tergic and GABAegic inputs, in addition to those already
Such anatomical complexity is paralleled by a functional and, mentioned, orchestrating the activity of dopaminergic neurons
perhaps, vesicular promiscuity, since dopaminergic neurons of the VTA.
can co-release glutamate or GABA. Activation of pyramidal neurons of the PFC in vivo can induce
The most powerful approaches used to disentangle the bursting of VTA DA neurons (Tong et al., 1996). Lateral hy-
anatomical and physiological intricacies of the VTA are derived pothalamus (Kempadoo et al., 2013) and BNST (Jennings et al.,
from the proper combination of transgenic technologies, opto- 2013) can also provide glutamatergic inputs. GABA neurons are
genetics, electrophysiology, and immunocytochemistry. For located within the VTA, as well. Coordinated activity of these
example, the use of retrograde tracing allowed for a detailed GABA cells, via selective optogenetic stimulation, can support
functional identification of DA neurons residing within the VTA conditioned place aversion and disrupt reward consumption
and projecting to NAc, PFC and basolateral amygdala (BLA) (Tan et al., 2012; van Zessen et al., 2012). Surprisingly, this

Neuron 86, June 3, 2015 ª2015 Elsevier Inc. 1147


Neuron

Review

Figure 1. Synaptic Plasticity and Dopaminergic Activity in Response to Rewarding or Aversive Processes
Rewarding experiences like a passive exposure to cocaine are capable of producing synaptic potentiation mostly within VTA DA neurons projecting to NAc lateral
shell (left) while aversive events, hind paw injection of formalin, cause synaptic potentiation mostly within PFC projecting neurons (right). These distinct subsets of
DA neurons receive different afferents. NAc shell projection neurons receive afferents from LDTg and RMTg and modulate reward processing, while PFC
projecting neurons receive excitatory afferents provided by LHb and they are implicated in aversion. The observed enhanced expression of AMPA receptors
occurring onto specific subsets of DA neurons, respectively supporting reward and aversion, might be a cellular mechanism for burst generation and concurrent
increased DA release.

population of cells can also innervate NAc. In addition, it has been induced a reduction in the activity of GABAergic neurons, thus
recently demonstrated that these cells synapse onto accumbal causing a subsequent disinhibition of DA neurons. The circuits
cholinergic interneurons (CINs). More importantly, optogenetic and neuronal population described above belong to anatomi-
manipulation of these terminals can induce a pause in the firing cally and molecularly defined neurons.
pattern of CINs in behaving mice and enhanced discrimination It is important to emphasize that beside the biological com-
of stimulus previously paired with an aversive outcome (Brown plexity conferred by the promiscuous release of neurotrans-
et al., 2012). Thus, these data support the intriguing idea that mitter, the existence of a ‘‘hybrid’’ population of neurons within
GABA neurons within the VTA can ‘‘teach’’ NAc how to respond the VTA emerges as a very prominent reality that strongly deter-
and promote actions, in regard to salient events. mines the intrinsic circuitry of this system. These populations of
Given the functional importance of GABAergic neurons cells reside within the VTA and project to the lateral habenula, ex-
located within the VTA, it is important to note that they receive pressing the prototypical markers of DA neurons, but are unable
inhibitory input from medium spiny neurons of the NAc to release detectable levels of DA in the target region. Surpris-
(Bocklisch et al., 2013), BNST (Jennings et al., 2013), and ventral ingly, they can release GABA in the LH and, accordingly, pro-
pallidum (Hjelmstad et al., 2013). Interestingly, it has been mote reward-seeking behaviors (Stamatakis et al., 2013).
discovered that synapses between medium spiny neurons ex- It is remarkable to point out that, during the last few years, sci-
pressing DA receptor type 1 of the NAc projecting to GABA entists have been able to add several fundamental anatomical
neurons of the VTA can undergo synaptic plasticity, namely and functional details to an already complicated and heteroge-
potentiation, after repeated in vivo experiences with cocaine. neous region. We are now facing a fascinating era in which we
This synaptic potentiation is also capable of occluding the are combining together all these pieces of information, with the
occurrence of homosynaptic, inhibitory long-term potentiation aim of clarifying and understanding the rules governing the
(Bocklisch et al., 2013). At the circuit level, this plastic adaptation reward circuits in physiological and pathological states.

1148 Neuron 86, June 3, 2015 ª2015 Elsevier Inc.


Neuron

Review

General Rules Governing Synaptic Plasticity in DA mGlu1-mediated LTD is also supported by elevation of intracel-
Neurons lular Ca2+ levels, since having BAPTA in the pipette completely
In vitro brain slice electrophysiology has been an invaluable tool prevents LTD (Bellone and Lüscher, 2005). The expression
to address synaptic level changes in the function of glutamater- mechanism of mGlu1-LTD in the VTA is quite peculiar. In fact,
gic synapses occurring in the brain, following different circum- instead of relying on a reduced number of postsynaptic AMPA
stances. Several electrophysiological studies have revealed receptors, it relies on an exchange of AMPA receptors with a
that learning induces long-lasting changes in the synaptic lower conductance, those that contain the GluA2 subunit (Bel-
strength of central glutamatergic synapses. As discussed earlier, lone and Lüscher, 2005). We will discuss the importance of
this activity-dependent modification of the efficacy of synaptic AMPAR composition in further detail below. Ultimately, the mo-
transmission is generally termed synaptic plasticity. It is widely lecular players regulating the increase or decrease in synaptic
recognized that this plasticity mechanism is a fundamental strengths observed in LTP and LTD, respectively, are the
cellular candidate for the acquisition and maintenance of AMPARs. This class of receptors can undergo differential traf-
learning and memory traces in the brain. In the mammalian ficking or a change in subunit composition at the cell surface,
brain, given the diversity of the properties and functions of depending on the type of synaptic plasticity occurring. In
each region, several different forms of synaptic plasticity have fact, depending on the synaptic mechanism engaged for the
been described. In this section, we attempt to provide a broad synaptic potentiation or depression of AMPAR-mediated excit-
overview of the mechanisms of synaptic plasticity that have atory synaptic transmission, we could imagine the following
been described at excitatory synapses in the VTA. Glutamatergic molecular scenarios: an increase or decrease in sensitivity,
synapses reaching the VTA are capable of expressing plasticity number of receptors, or changes in the expression of AMPAR
properties. Indeed, several studies have found that both LTP and subunits.
LTD caused by evoked AMPAR currents can be induced in the AMPARs are fundamental ionotropic transmembrane recep-
VTA (Bonci and Malenka, 1999; Thomas et al., 2000). Older in- tors mediating fast excitatory synaptic transmission in the
duction protocols, adapted from the hippocampus literature, CNS. Assembly of these receptors requires four subunits
held DA neurons at a depolarized potential for a long period of (GluA1–GluA4), which can form hetero- or homomeric com-
time, while stimulating afferents at high or low frequencies. plexes (Gan et al., 2015). AMPARs are thought to exist as hetero-
Despite the fact that the older induction protocols were capable meric complexes containing both GluA2/3 and GluA1 subunits.
of producing plasticity, they were unable to recapitulate biolog- At resting state, the vast majority of AMPARs contain the
ically relevant patterns of synaptic activity. GluA2 subunit, forming heteromeric receptors with GluA1 or
Theoretically speaking, perhaps a more physiologically rele- GluA3. However, GluA2-lacking AMPARs can form, in which
vant induction protocol is represented by spike-timing–depen- the tetrameters are composed of GluA1/1 or GluA1/3 AMPARs,
dent plasticity (STDP). In STDP, the induction of plasticity is in other words ‘‘GluA1-type.’’ GluA1-type AMPARs are perme-
governed by specific rules that organize the order and precise able to calcium, which can be a permissive signaling event for
temporal interval between pre and post-synaptic spikes, resem- the establishment of plasticity, while GluA2 containing receptors
bling the physiological activity of the neuron (Feldman, 2012). are Ca2+ impermeable (Isaac et al., 2007). Our understanding of
Based on these assumptions, the use of STDP has emerged in how AMPAR subunit composition is regulated has been aided by
the VTA plasticity field over the last decade. Interestingly, both both biochemical (crosslinking methods) and electrophysiolog-
classical and STDP induction protocols are capable of producing ical techniques. The use of crosslinking assays allowed the res-
LTP and, similarly to what has been reported in other brain re- olution of AMPAR subunit type present on the cell surface and
gions (i.e., hippocampus), rely on NMDAR activation, as well as also revealed the phosphorylation state of specific sets of AM-
a subsequent increase in postsynaptic calcium. Conversely, PARs subunits.
LTD in the VTA is NMDAR independent. In fact, bath application Importantly, depending on their phosphorylation state, AMPARs
of APV, the NMDA receptor antagonist, is not able to prevent low- can increase or decrease channel conductance. Electrophysiolog-
frequency stimulation-induced LTD (LFS-LTD). Surprisingly, ical techniques have also taken advantage of a specific feature
however, Ca2+ signaling in the postsynaptic compartment is of GluA2-lacking AMPARs: that functional block occurs when
essential for the proper expression of LFS-LTD; in fact having intracellular polyamines are applied at a positive potential. This
the Ca2+ chelator BAPTA in the recording pipette prevents LFS- biophysical feature parallels a very unique inward rectifying
LTD. Furthermore, bath application of DA or the D2 receptor current-voltage relationship that can be used to determine the
agonist quinpirole blocks LFS-LTD (Thomas et al., 2000). The mo- presence or absence of GluA2-lacking AMPARs. Biochemical
lecular mechanisms underlying this synaptic process are still and electrophysiological approaches can then be used in combi-
poorly understood. Certainly, the functional recruitment of den- nation, via the use of specific subunit-selective peptide antago-
dritic potassium channels by DA, or by quinpirole, is unlikely since nists. The elegance of the use of the latter approach resides
these experiments were performed with TEA and cesium in the in the possibility to dissect out the relative contribution of GluA1
whole cell pipette, both potassium channel blockers. and GluA2 AMPAR subunits in both in vitro and in vivo experi-
It is very intriguing that activation of mGlu1 (Group I metabo- mental conditions.
tropic glutamate receptors) is also capable of producing LTD in A powerful and convenient behavioral framework for studying
the VTA and, more importantly, mGlu and LFS-LTD coexist synaptic plasticity and the electrophysiological impact, in terms
and do not occlude each other, thus supporting the idea that of behavioral adaptation, has been offered by the drug addiction
they are mechanistically distinct (Bellone and Lüscher, 2005). field. Evidence that drugs of abuse are capable of producing

Neuron 86, June 3, 2015 ª2015 Elsevier Inc. 1149


Neuron

Review
Figure 2. Cocaine Exposure ‘‘Hijacks’’
Synaptic Plasticity Rules of VTA Dopamine
Neurons
Synaptic plasticity involves several orchestrated
changes of synaptic efficacy supporting experi-
ence-dependent modifications of brain function.
The increased efficacy of glutamatergic synapses
onto DA neurons observed after cocaine exposure
is supported by a redistribution of AMPA (GluR2
Lacking) and NMDA receptors (GluN2B and
GluN3A containing). The observed changes in
AMPA and NMDA subunit composition lead to
changes in Ca2+ permeability that are capable of
affecting synaptic rules governing activity-
dependent plasticity. As a consequence, anti-
Hebbian pairing protocols, not sufficient to induce
LTP in naive conditions, can produce LTP after
cocaine exposure. Interestingly, both in vitro and
in vivo activation of mGluR1 can reverse cocaine-
induced synaptic adaptation via synaptic incor-
poration of GluR2 containing AMPA receptors.
Whether activation of mGluR1 is also capable of
reversing NMDA receptors redistribution is still
unknown.

Another level of complexity has been


revealed by photo-uncaging glutamate
at single synapses of DA neurons in ani-
mals that underwent in vivo cocaine
plasticity within DA neurons of the VTA was shown with a single exposure. The experiments showed reduced NMDAR-mediated
non-contingent injection of cocaine. Our group discovered an currents (Mameli et al., 2011). Reduction in NMDAR-mediated
increased AMPAR/NMDAR ratio, a ‘‘surrogate’’ measure of currents can concurrently support and greatly amplify the
LTP, 24 hr after the treatment (Ungless et al., 2001). Additionally, observed increased in AMPA/NMDA ratio. It is important to point
it was demonstrated that excitatory synapses were not further out that NMDARs are fundamental players in the induction phase
strengthened, supporting the idea that the potentiation, and of LTP and in several other integrative cellular processes. Thus,
the subsequent occlusion phenomena, was due to the fact that their expression or subunit composition can be a crucial
cocaine-induced LTP shared the same mechanism as physio- component for the proper tuning of synaptic plasticity. DA
logical LTP. This plastic neuroadaptation was observed 24h after neurons of animals exposed to cocaine display a larger sensi-
the injection and persists for up to 5 days. tivity to the selective GluN2B inhibitor ifenprodil, compared to
Interestingly, several other salient events like food and sucrose, the selective GluN2A inhibitor zinc (Yuan et al., 2013). The
delivered after self-administration procedures, are capable of latter discovery leads to the consideration of the ratio of
producing the same neuroadaptation (Chen et al., 2008). The GluN2A/GluN2B receptors at synapses. It is important to
fundamental difference is revealed by the duration of the synaptic appreciate that switches in NMDAR subunit composition can
potentiation. Plasticity triggered by sucrose, food, or pure appe- potently regulate Ca2+ entry into the post-synaptic compart-
titive Pavlovian learning and memory mechanisms is a very tran- ment and, more importantly, can affect the thresholds required
sient phenomenon that decays rapidly over time. In the case of for inducing synaptic plasticity (Kopp et al., 2007). Yuan et al.
cocaine, and particularly in animals trained to self-administer (2013) discovered that the observed change in subunit
the drug, this LTP-like state persists for several weeks. Interest- composition was strictly dependent on the insertion of GluN3A
ingly, cocaine-induced synaptic potentiation of AMPARs is receptors, shown via a combination of GluN3A knockout ani-
mediated by prior recruitment of NMDA receptors. In fact, in vivo mals and short hairpin RNA approaches. Synaptic insertion
treatment with MK-801, the noncompetitive antagonist of the of GluN3A receptors lead to reduced Ca2+ permeability and
NMDARs, prevented the occurrence of synaptic plasticity. magnesium sensitivity and is required for the development of
Notably, the increase in AMPAR/NMDAR ratio parallels an a cocaine-induced increase in AMPAR/NMDAR ratio. The
increased sensitivity to polyamines, which is indicative of observed changes in AMPA and NMDA subunit composition,
the presence of GluA2-lacking AMPARs (Argilli et al., 2008; in terms of Ca2+ permeability, can reconcile the observation
Bellone and Luscher, 2006). In addition, both in vitro and in vivo that after cocaine exposure, anti-Hebbian pairing protocols,
activation of mGlu1 are capable of reversing the early synaptic not sufficient to induce LTP in naive conditions, are capable
effects triggered by cocaine exposure, namely GluA2-lacking of inducing LTP (Mameli et al., 2011).
AMPARs, via de novo synthesis and synaptic incorporation Thanks to the advent of optogenetics, circuit level analysis has
of GluA2 containing receptors (Bellone and Luscher, 2006; Ma- become a very fruitful and important way to investigate how
meli et al., 2007) (Figure 2). neuroadaptations caused by appetitive events, like cocaine

1150 Neuron 86, June 3, 2015 ª2015 Elsevier Inc.


Neuron

Review

experience, occur within DA neurons. DA neurons can be classi- transients, the aversive quinine-based oral solution induced a
fied depending on their different projection targets. While VTA DA reduction in DA release (Roitman et al., 2008). Interestingly, DA
neurons projecting to NAc medial shell display a robust increased release does not only depend on the hedonic or the motivational
AMPAR/NMDAR ratio 24 hr after cocaine injection, VTA DA neu- states engaged by the behavioral framework used, but it is highly
rons projecting to a lateral shell show only a subtle change in the regulated by learned associations (Roitman et al., 2004). When
AMPA/NMDA ratio. Surprisingly, VTA DA neurons projecting to animals received extensive Pavlovian conditioning training,
the prefrontal cortex are insensitive to cocaine exposure and where a conditioned stimulus (CS+) predicts the delivery of a
have unchanged AMPAR/NMDAR ratio 24h after cocaine treat- reinforcer-unconditioned stimulus (US), phasic DA release within
ment. Instead, they respond vigorously to an aversive experience the NAc, that early in training appears to be primarily associated
(hind paw injection of formalin) with a remarkable increase in to the delivery of the reward, moves to the onset of the CS.
AMPAR/NMDAR ratio (Lammel et al., 2011). The functional The idea that learned associations can affect the occurrence
segregation of DA neurons based on the projection targets of DA release within the NAc has also been shown for aversive
is also supported by behavioral optogenetic manipulations. experiences. In fact, phasic DA release events are suppressed
In fact, optogenetic stimulation of accumbens-projecting DA by stimuli that have been associated with an aversive outcome.
neurons support conditioned place preference, while optoge- Conversely, phasic DA release events are increased by stimuli
netic stimulation of prefrontal cortex-projecting DA neurons sup- that precede the successful avoidance of an aversive event
ports a vigorous conditioned place aversion (Lammel et al., (Oleson et al., 2012). The lessons that we have learned so far
2012). Establishment of a maladaptive process of reward-related account for an active interplay between learning and memory
learning has been suggested as a core feature of drug addiction. mechanisms and the regulation of specific patterns of DA
Therefore, the ability of drugs of abuse to forge synaptic plasticity release. The question left is how learning and memory mecha-
and ‘‘hijack’’ the natural reward process is an appealing reason nisms are connected with the occurrence of DA release. Interest-
for the development and, more importantly, for the persistence ingly, during reward-learned associations, DA release to reward
of behavioral responses associated with drugs of abuse. predictive cues parallel an enhanced synaptic strength onto DA
However, to date, there is no clear evidence that the observed neurons (Stuber et al., 2008). Moreover, the observed synaptic
initial synaptic plasticity changes triggered by substances of potentiation develops over the course of cue-reward learning.
abuse represent a hallmark of addiction, but rather an initial It is worth mentioning that in conditioned appetitive learning,
step that will eventually lead to the induction of the disorder later where environmental cues are associated with the receipt of a
on. Moreover, it has been shown that the persistence of synaptic reward, food restriction can strongly increase the acquisition
plasticity occurring in the VTA is a requirement for the expression rate with which the animals learn how to perform the task.
of a later synaptic adaptation that takes place in the NAc (Mameli Notably, restriction of food intake is also a well-known strategy
et al., 2009). Potentially, synaptic adaptation in the ventral stria- to enhance the reinforcing efficacy of drugs of abuse in experi-
tum might represent a gate for a later establishment of stimulus- mental animal models (Carroll et al., 1979). The overall idea is
response habit formation occurring in the dorsal striatum (Belin that food restriction can enhance motivational or hedonic states
and Everitt, 2008). that are important for learning and acquisition of behavioral
Together, these observations raise the question of whether tasks. Surprisingly, animals under food restriction protocols
this cocaine-induced long-lasting potentiation may also reflect display not just an increase in behavioral responses to drugs of
a disruption in the ability of DA neurons to code for reward and abuse, but also enhanced burst firing of DA neurons, increased
aversion during the occurrence of naturalistic behavior. AMPA/NMDA ratio and D2 autoreceptor-mediated desensitiza-
tion in DA neurons (Branch et al., 2013).
Synaptic Plasticity and Dopaminergic Activity: Shaping The timing in which synaptic plasticity occurs, and the fact that
DA Release Events glutamatergic plasticity might affect the cellular mechanisms by
The majority of information related to DA function in the brain which DA neurons produce DA release, is of great interest to the
has been provided by a combination of behavioral, pharmaco- scientific community. In fact, the enhanced synaptic strength
logical, genetic, and electrophysiological approaches. Surpris- observed in the aforementioned conditions might be the cellular
ingly, none of the cited approaches can selectively provide mechanism lying at the core of the transformation of neutral
subsecond, temporal resolution information related to DA environmental stimuli to salient reward-predictive cues and
release happening within ‘‘intact’’ forebrain terminal regions. potentially supports the establishment of specific hedonic and
The relatively recent introduction of voltammetric approaches, motivational states that are involved in learning processes.
like fast scan cyclic voltammetry (FSCV), has radically changed
the way we approach DA release in the brain, especially in The Dual Role of Glutamatergic Input onto Dopamine
regards to appetitive and aversive experiences. The power Neurons
of this methodology is mostly represented by a capacity to Glutamatergic inputs are not just required for the induction of
quantitate phasic changes in DA concentration, within given synaptic plasticity but they are also a fundamental player in
target regions, occurring on a physiological timescale. the induction of burst firing of DA neurons. In particular, activa-
Detecting DA concentration by using FSCV within the NAc of tion of NMDARs induces burst firing by regulating calcium
animals that are experiencing opposite hedonic valence re- transients that initiate and terminate bursts via Ca2+-activated
vealed an opposite pattern of DA release. While the appetitive potassium channels (Paladini and Roeper, 2014). Functional
sucrose-based oral solution evoked a robust increase in DA alterations, in the previously cited cellular mechanisms, can

Neuron 86, June 3, 2015 ª2015 Elsevier Inc. 1151


Neuron

Review

affect the way DA neurons produce burst firing during motiva- able to identify AMPA receptors in striatal DA terminals (Bernard
tionally relevant conditions. Plasticity mechanisms occurring in and Bolam, 1998; Bernard et al., 1997).
response to aversive, appetitive, and associative learning condi- How can it be possible that AMPA receptors are capable of
tions develop via a selective recruitment of synaptic AMPARs modulating DA release in the striatum via action onto dopami-
(Lammel et al., 2011; Saal et al., 2003; Ungless et al., 2001). nergic terminals if there is no evidence for expression? It has
Thus, it is likely that enhanced expression of AMPARs in DA neu- been suggested that H2O2 can act as a key regulator of this local
rons represents an additional cellular mechanism for promoting AMPAR-mediated DA release (Avshalumov et al., 2000, 2008;
burst generation (Canavier and Landry, 2006; Komendantov Avshalumov and Rice, 2003). In this model, glutamatergic activ-
et al., 2004). It is therefore plausible to hypothesize that a poten- ity occurring onto AMPARs expressed in medium spiny neurons
tiated synaptic state onto DA neurons may allow the generation (MSNs) could lead to the production of H2O2, which could then
of a more depolarized membrane potential under specific cir- diffuse to adjacent DA terminals, open KATP channels and regu-
cumstances and subsequently allow fast Mg2+ unbinding ki- late DA release. While there is no clear evidence for the existence
netics of NMDARs, thus increasing channel open probability of AMPARs in dopaminergic terminals in the NAc, there is solid
and burst firing generation. The latter functional state may allow evidence for the existence of mGlu receptors (mGluRs) (Paquet
several behavioral conditions to generate uncontrolled burst and Smith, 2003). mGluRs, mostly located at the perisynaptic
firing, representing an unchecked route to specific forebrain ef- sites, are engaged in the regulation of DA release in the striatum,
fectors capable of supporting different psychiatric conditions. where they are recruited in the case of glutamate spillover. Once
In support of this prediction is the demonstration that burst firing functionally recruited, they could activate Ca2+-activated K+
of DA neurons is elevated in awake rats that are experiencing channels and thus affect DA release (Zhang and Sulzer, 2003).
passive exposure to cocaine (Koulchitsky et al., 2012). All of these studies however only partially account for a local
Unexpectedly, in a social defeat stress model, a behavioral glutamatergic regulation of DA release into the NAc, at the level
condition with presumably an opposite hedonic valence in res- of dopaminergic terminals, under physiological conditions.
pect to cocaine exposure, susceptible mice showed a robust In fact, little is known about the presence and the function
hyperactivity of VTA DA neurons. This cellular phenotype was of AMPARs onto dopaminergic terminals after exposure to
caused by the upregulation of the hyperpolarization-activated reward-related stimuli, drugs of abuse, or aversive stimuli. Addi-
cyclic nucleotide-gated type 2 (HCN2) channels (Ih), a protein tionally, there seems to be no evidence about the presence of
regulating intrinsic excitability onto DA neurons. On the other AMPA receptors on DA terminals located in the PFC, even
hand, mice expressing a resilient phenotype displayed an even though it is known that AMPA receptors blockade in the PFC
larger Ih and also increased potassium (K+) channel-mediated significantly reduced cortical DA release (Takahata and Moghad-
currents. Pharmacological treatment with the Ih potentiator la- dam, 2000).
motrigine, or increasing activity of DA VTA neurons via optoge- It is plausible to speculate that AMPA receptors may be
netic manipulation in susceptible mice, is capable of reversing actively transported to axonal terminals or incorporated there,
depression-related traits (Friedman et al., 2014). Interestingly, after in loco de novo synthesis from preexisting mRNA, when
optogenetic stimulation occurring within DA neurons of the the dopaminergic system is expressing synaptic plasticity pro-
VTA mediates behavioral rescue via a homeostatic plasticity cesses triggered by salient conditions (Figure 3). The expression
mechanism occurring between VTA-NAc projecting neurons and synaptic recruitment of these AMPA receptors would be
(Friedman et al., 2014). Future studies will likely solve the ques- potentially regulated depending on the biological identity of
tion of how synaptic plasticity and intrinsic mechanisms regu- each individual DA neuron, projection target, specific glutama-
lating cellular excitability can affect, in a cooperative fashion, tergic afferent, or, more importantly, by the emotional nature of
phasic DA release in a population of DA-projecting neurons. the experience an animal faces. Cross talk between AMPARs
A glutamatergic regulation of DA release has been also sug- and mGluRs at the axonal terminals can then shape DA release
gested at the level of nerve terminals for PFC and NAc (Floresco events within a given terminal region in a multifaceted way,
et al., 1998; Jones et al., 2010; Takahata and Moghaddam, thereby adding another level of complexity to the excitatory
1998). For example, dopaminergic and excitatory terminals are events happening at the somatic level.
localized in close apposition both in the PFC and NAc. Anatom- Moreover, neuronal inputs arising from multiple sites of the
ical studies identified that glutamatergic afferents from the brain, and converging at the axonal level, might differentially
BLA terminate in close apposition to DA terminals in the NAc impact DA release and ensure appropriate behavioral response
(Johnson et al., 1994). On the other hand, DA terminals and excit- to the information provided by an ever-changing environment
atory afferents from the hippocampal formation are localized in both physiological and pathological states (Floresco et al.,
in close apposition in the PFC (Carr and Sesack, 1996; Smiley 1998; Jones et al., 2010; Takahata and Moghaddam, 1998).
and Goldman-Rakic, 1993). DA release within NAc and PFC is Thus, mesocortical DA neurons may be engaged in cognitive
highly sensitive to AMPARs blockade, suggesting that AMPARs processing underlying working memory or extinction process-
are regulating DA release in a tonic fashion. This experimental ing, as well as acute stress reaction occurring during aversive ex-
data raise the possibility that AMPA receptors do not only reside periences and so modulate alertness (Abraham et al., 2014).
within the somatic or dendritic shaft compartments of VTA DA Meanwhile, mesoaccumbal DA neurons might be engaged in
neurons (Paquet et al., 1997), but could also be potentially ex- motivational functions like proper orientation of goal directed be-
pressed on DA terminals, where they might exert a control over haviors, in the emotional valence and attribution of saliency to
DA release. However, electron microscopy studies were not the experience (Ungless, 2004).

1152 Neuron 86, June 3, 2015 ª2015 Elsevier Inc.


Neuron

Review

Figure 3. Synaptic Plasticity within Specific Cellular Compartments of VTA Dopamine Neurons: Shaping and Tuning Terminal DA Release
within Specific Target Regions in Response to Rewarding or Aversive Events
DA release is regulated at the level of nerve terminals by glutamatergic afferents in the Nucleus accumbens (NAc) and in the medial prefrontal cortex (mPFC).
Glutamatergic afferents arising from the basolateral amygdala (BLA) terminate in close apposition onto DA terminals in the NAc, while afferents emanated by the
hippocampus (Hipp) localized in close apposition to DA terminals in the mPFC. DA release in both terminal regions is sensitive to AMPA receptors blockade,
presumably because AMPA receptors are exerting a tonic control over DA release. We speculate that AMPA receptors may be actively transported to axonal
terminals or incorporated there after de novo synthesis from pre-existing mRNA when DA neurons undergo plasticity processes. The latter condition could
potentially confer a very sophisticated level of control over terminal DA release within specific target regions that are implicated in reward or aversion processing.

Particularly interesting to this matter is the notion that VTA While a lot of work has been done, solving all the biological
dopaminergic neurons do not code, per se, for cognitive aspects questions discussed in this section will provide a much better un-
of learned associations. As a matter of fact, DA neurons do derstanding of how synaptic plasticity, intrinsic excitability, and
not carry detailed aspects about the nature of the stimuli, but concurrent DA release, occurring in specific circuits engaged
rather they have the ability to add ‘‘color or flavor’’ to experi- in appetitive or aversive conditions, can produce maladaptive,
ence-dependent processes implicated in memory formation, pathological behaviors, such as substance use disorders.
by virtue of their control over ‘‘saliency attribution’’ (Berridge
and Robinson, 1998).
The way DA neurons modulate saliency attribution to Conclusions
behavioral experiences is by controlling, via DA release, gluta- In conclusion, we have reviewed recent findings concerning the
matergic synaptic activity within the ventral striatum (O’Donnell effects of both salient appetitive and aversive stimuli in DA neu-
and Grace, 1995; O’Donnell et al., 1999). Glutamatergic inputs rons of the VTA. Over the past decade, the use of innovative
reaching the NAc (from the ventral hippocampus, basolateral techniques such as transgenic approaches and optogenetics
amygdala, PFC, and thalamus) provide information about con- has radically changed the view of the midbrain DA system. Sci-
textual elements, cues, and even subtle descriptive features ence has come a long way in the description of the VTA that
that belong to a specific environmental condition (Britt et al., is now passing through a novel reclassification based on
2012; Sesack and Grace, 2010). What DA does to these neuron subtypes displaying specific features at the molecular,
afferents is to control the signal-to-noise ratio and the resultant anatomical, and electrophysiological level (Barrot, 2014; Margo-
throughput of the ventral striatum, by maintaining a proper lis et al., 2008; Root et al., 2014; Tritsch et al., 2012). These
tuning between the drive and resultant behavioral responses groundbreaking scientific contributions are reshaping our knowl-
promoted by limbic or cortical inputs, respectively (Goto and edge about synaptic connectivity and molecular profiling of the
Grace, 2008). VTA (Ekstrand et al., 2014).

Neuron 86, June 3, 2015 ª2015 Elsevier Inc. 1153


Neuron

Review

In this review, we speculate that, depending on the emotional Bernard, V., and Bolam, J.P. (1998). Subcellular and subsynaptic distribution
of the NR1 subunit of the NMDA receptor in the neostriatum and globus pal-
content of the experiences, each individual projection-specific lidus of the rat: co-localization at synapses with the GluR2/3 subunit of the
DA cell might make a unique contribution to inform the brain AMPA receptor. Eur. J. Neurosci. 10, 3721–3736.
how to adapt to behavioral responses. Changes in synaptic
Bernard, V., Somogyi, P., and Bolam, J.P. (1997). Cellular, subcellular, and
configuration, and the resultant patterns of DA cell activity, might subsynaptic distribution of AMPA-type glutamate receptor subunits in the
provide a mechanism by which synaptic neural adaptations neostriatum of the rat. J. Neurosci. 17, 819–833.
occurring in the reward circuit are transferred to forebrain effec- Berridge, K.C. (2000). Measuring hedonic impact in animals and infants: micro-
tors. DA release, in its ‘‘conventional’’ or promiscuous co-release structure of affective taste reactivity patterns. Neurosci. Biobehav. Rev. 24,
with other neurotransmitters, can promote or enhance the occur- 173–198.
rence of spike-time-dependent plasticity at active cortico- Berridge, K.C., and Robinson, T.E. (1998). What is the role of dopamine in
limbic-striatal synapses (Shiflett and Balleine, 2011). The occur- reward: hedonic impact, reward learning, or incentive salience? Brain Res.
Brain Res. Rev. 28, 309–369.
rence of synaptic plasticity within the cortico-limbic-striatal
circuits can affect and shape the function of each individual Bliss, T.V., and Collingridge, G.L. (1993). A synaptic model of memory: long-
region that can now take the lead and impose its own specific term potentiation in the hippocampus. Nature 361, 31–39.
function to orchestrate and guide future adaptive behaviors. Bliss, T.V., Collingridge, G.L., and Morris, R.G. (2014). Synaptic plasticity in
Additionally, DA-mediated synaptic plasticity can support the health and disease: introduction and overview. Philos. Trans. R. Soc. Lond.
B Biol. Sci. 369, 20130129.
maintenance and consolidation of learning, presumably by add-
ing specific emotional weight, depending on the functional con- Bocklisch, C., Pascoli, V., Wong, J.C., House, D.R., Yvon, C., de Roo, M., Tan,
nectivity of the neural network in which neurons reside. K.R., and Lüscher, C. (2013). Cocaine disinhibits dopamine neurons by poten-
tiation of GABA transmission in the ventral tegmental area. Science 341, 1521–
1525.
REFERENCES
Bonci, A., and Malenka, R.C. (1999). Properties and plasticity of excitatory syn-
apses on dopaminergic and GABAergic cells in the ventral tegmental area.
Abraham, A.D., Neve, K.A., and Lattal, K.M. (2014). Dopamine and extinction: J. Neurosci. 19, 3723–3730.
a convergence of theory with fear and reward circuitry. Neurobiol. Learn. Mem.
108, 65–77. Branch, S.Y., Goertz, R.B., Sharpe, A.L., Pierce, J., Roy, S., Ko, D., Paladini,
C.A., and Beckstead, M.J. (2013). Food restriction increases glutamate recep-
Anggono, V., and Huganir, R.L. (2012). Regulation of AMPA receptor trafficking tor-mediated burst firing of dopamine neurons. J. Neurosci. 33, 13861–13872.
and synaptic plasticity. Curr. Opin. Neurobiol. 22, 461–469.
Britt, J.P., Benaliouad, F., McDevitt, R.A., Stuber, G.D., Wise, R.A., and Bonci,
Aragona, B.J., Day, J.J., Roitman, M.F., Cleaveland, N.A., Wightman, R.M., A. (2012). Synaptic and behavioral profile of multiple glutamatergic inputs to
and Carelli, R.M. (2009). Regional specificity in the real-time development of the nucleus accumbens. Neuron 76, 790–803.
phasic dopamine transmission patterns during acquisition of a cue-cocaine
association in rats. Eur. J. Neurosci. 30, 1889–1899. Brown, M.T., Tan, K.R., O’Connor, E.C., Nikonenko, I., Muller, D., and Lüscher,
C. (2012). Ventral tegmental area GABA projections pause accumbal cholin-
Argilli, E., Sibley, D.R., Malenka, R.C., England, P.M., and Bonci, A. (2008). ergic interneurons to enhance associative learning. Nature 492, 452–456.
Mechanism and time course of cocaine-induced long-term potentiation in
the ventral tegmental area. J. Neurosci. 28, 9092–9100. Canavier, C.C., and Landry, R.S. (2006). An increase in AMPA and a decrease
in SK conductance increase burst firing by different mechanisms in a model of
Avshalumov, M.V., and Rice, M.E. (2003). Activation of ATP-sensitive K+ a dopamine neuron in vivo. J. Neurophysiol. 96, 2549–2563.
(K(ATP)) channels by H2O2 underlies glutamate-dependent inhibition of stria-
tal dopamine release. Proc. Natl. Acad. Sci. USA 100, 11729–11734. Carr, D.B., and Sesack, S.R. (1996). Hippocampal afferents to the rat prefron-
tal cortex: synaptic targets and relation to dopamine terminals. J. Comp.
Avshalumov, M.V., Chen, B.T., and Rice, M.E. (2000). Mechanisms underlying Neurol. 369, 1–15.
H(2)O(2)-mediated inhibition of synaptic transmission in rat hippocampal
slices. Brain Res. 882, 86–94. Carroll, M.E., France, C.P., and Meisch, R.A. (1979). Food deprivation in-
creases oral and intravenous drug intake in rats. Science 205, 319–321.
Avshalumov, M.V., Patel, J.C., and Rice, M.E. (2008). AMPA receptor-depen-
dent H2O2 generation in striatal medium spiny neurons but not dopamine Changizi, M.A., McGehee, R.M., and Hall, W.G. (2002). Evidence that appeti-
axons: one source of a retrograde signal that can inhibit dopamine release. tive responses for dehydration and food-deprivation are learned. Physiol.
J. Neurophysiol. 100, 1590–1601. Behav. 75, 295–304.

Balleine, B., and Dickinson, A. (1992). Signalling and incentive processes in Chaudhury, D., Walsh, J.J., Friedman, A.K., Juarez, B., Ku, S.M., Koo, J.W.,
instrumental reinforcer devaluation. Q. J. Exp. Psychol. B 45, 285–301. Ferguson, D., Tsai, H.C., Pomeranz, L., Christoffel, D.J., et al. (2013). Rapid
regulation of depression-related behaviours by control of midbrain dopamine
Barrot, M. (2014). The ventral tegmentum and dopamine: a new wave of diver- neurons. Nature 493, 532–536.
sity. Neuroscience 282C, 243–247.
Chen, B.T., Bowers, M.S., Martin, M., Hopf, F.W., Guillory, A.M., Carelli, R.M.,
Beierholm, U., Guitart-Masip, M., Economides, M., Chowdhury, R., Düzel, E., Chou, J.K., and Bonci, A. (2008). Cocaine but not natural reward self-adminis-
Dolan, R., and Dayan, P. (2013). Dopamine modulates reward-related vigor. tration nor passive cocaine infusion produces persistent LTP in the VTA.
Neuropsychopharmacology 38, 1495–1503. Neuron 59, 288–297.

Belin, D., and Everitt, B.J. (2008). Cocaine seeking habits depend upon dopa- Cohen, J.Y., Haesler, S., Vong, L., Lowell, B.B., and Uchida, N. (2012). Neuron-
mine-dependent serial connectivity linking the ventral with the dorsal striatum. type-specific signals for reward and punishment in the ventral tegmental area.
Neuron 57, 432–441. Nature 482, 85–88.

Bellone, C., and Lüscher, C. (2005). mGluRs induce a long-term depression in Collingridge, G.L., Peineau, S., Howland, J.G., and Wang, Y.T. (2010). Long-
the ventral tegmental area that involves a switch of the subunit composition of term depression in the CNS. Nat. Rev. Neurosci. 11, 459–473.
AMPA receptors. Eur. J. Neurosci. 21, 1280–1288.
Daberkow, D.P., Brown, H.D., Bunner, K.D., Kraniotis, S.A., Doellman, M.A.,
Bellone, C., and Luscher, C. (2006). Cocaine triggered AMPA receptor redistri- Ragozzino, M.E., Garris, P.A., and Roitman, M.F. (2013). Amphetamine para-
bution is reversed in vivo by mG luR-dependent long-term depression. Nat. doxically augments exocytotic dopamine release and phasic dopamine sig-
Neurosci. 9, 636–641. nals. J. Neurosci. 33, 452–463.

1154 Neuron 86, June 3, 2015 ª2015 Elsevier Inc.


Neuron

Review
Darvas, M., Wunsch, A.M., Gibbs, J.T., and Palmiter, R.D. (2014). Dopamine Jackson, A., Mead, A.N., and Stephens, D.N. (2000). Behavioural effects
dependency for acquisition and performance of Pavlovian conditioned of alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionate-receptor antago-
response. Proc. Natl. Acad. Sci. USA 111, 2764–2769. nists and their relevance to substance abuse. Pharmacol. Ther. 88, 59–76.

Ekstrand, M.I., Nectow, A.R., Knight, Z.A., Latcha, K.N., Pomeranz, L.E., and Jennings, J.H., Sparta, D.R., Stamatakis, A.M., Ung, R.L., Pleil, K.E., Kash,
Friedman, J.M. (2014). Molecular profiling of neurons based on connectivity. T.L., and Stuber, G.D. (2013). Distinct extended amygdala circuits for diver-
Cell 157, 1230–1242. gent motivational states. Nature 496, 224–228.

Fadok, J.P., Dickerson, T.M., and Palmiter, R.D. (2009). Dopamine is neces- Johanson, I.B., and Hall, W.G. (1979). Appetitive learning in 1-day-old rat pups.
sary for cue-dependent fear conditioning. J. Neurosci. 29, 11089–11097. Science 205, 419–421.

Fadok, J.P., Darvas, M., Dickerson, T.M., and Palmiter, R.D. (2010). Long-term Johnson, L.R., Aylward, R.L., Hussain, Z., and Totterdell, S. (1994). Input from
memory for pavlovian fear conditioning requires dopamine in the nucleus ac- the amygdala to the rat nucleus accumbens: its relationship with tyrosine
cumbens and basolateral amygdala. PLoS ONE 5, e12751. hydroxylase immunoreactivity and identified neurons. Neuroscience 61,
851–865.
Feldman, D.E. (2012). The spike-timing dependence of plasticity. Neuron 75,
556–571. Jones, J.L., Day, J.J., Aragona, B.J., Wheeler, R.A., Wightman, R.M., and
Carelli, R.M. (2010). Basolateral amygdala modulates terminal dopamine
Floresco, S.B., Yang, C.R., Phillips, A.G., and Blaha, C.D. (1998). Basolateral release in the nucleus accumbens and conditioned responding. Biol. Psychi-
amygdala stimulation evokes glutamate receptor-dependent dopamine efflux atry 67, 737–744.
in the nucleus accumbens of the anaesthetized rat. Eur. J. Neurosci. 10, 1241–
1251. Kauer, J.A. (2004). Learning mechanisms in addiction: synaptic plasticity in the
ventral tegmental area as a result of exposure to drugs of abuse. Annu. Rev.
Friedman, A.K., Walsh, J.J., Juarez, B., Ku, S.M., Chaudhury, D., Wang, J., Li, Physiol. 66, 447–475.
X., Dietz, D.M., Pan, N., Vialou, V.F., et al. (2014). Enhancing depression mech-
anisms in midbrain dopamine neurons achieves homeostatic resilience. Sci- Kelley, A.E., Andrzejewski, M.E., Baldwin, A.E., Hernandez, P.J., and Pratt,
ence 344, 313–319. W.E. (2003). Glutamate-mediated plasticity in corticostriatal networks: role in
adaptive motor learning. Ann. N Y Acad. Sci. 1003, 159–168.
Gan, Q., Salussolia, C.L., and Wollmuth, L.P. (2015). Assembly of AMPA
receptors: mechanisms and regulation. J. Physiol. 593, 39–48. Kempadoo, K.A., Tourino, C., Cho, S.L., Magnani, F., Leinninger, G.M., Stuber,
G.D., Zhang, F., Myers, M.G., Deisseroth, K., de Lecea, L., and Bonci, A.
Gore, B.B., Soden, M.E., and Zweifel, L.S. (2014). Visualization of plasticity in (2013). Hypothalamic neurotensin projections promote reward by enhancing
fear-evoked calcium signals in midbrain dopamine neurons. Learn. Mem. 21, glutamate transmission in the VTA. J. Neurosci. 33, 7618–7626.
575–579.
Komendantov, A.O., Komendantova, O.G., Johnson, S.W., and Canavier, C.C.
Goto, Y., and Grace, A.A. (2008). Limbic and cortical information processing in (2004). A modeling study suggests complementary roles for GABAA and
the nucleus accumbens. Trends Neurosci. 31, 552–558. NMDA receptors and the SK channel in regulating the firing pattern in midbrain
dopamine neurons. J. Neurophysiol. 91, 346–357.
Grace, A.A., and Bunney, B.S. (1984a). The control of firing pattern in nigral
Kopp, C., Longordo, F., and Lüthi, A. (2007). Experience-dependent changes
dopamine neurons: burst firing. J. Neurosci. 4, 2877–2890.
in NMDA receptor composition at mature central synapses. Neuropharma-
cology 53, 1–9.
Grace, A.A., and Bunney, B.S. (1984b). The control of firing pattern in nigral
dopamine neurons: single spike firing. J. Neurosci. 4, 2866–2876.
Koulchitsky, S., De Backer, B., Quertemont, E., Charlier, C., and Seutin, V.
(2012). Differential effects of cocaine on dopamine neuron firing in awake
Grace, A.A., Floresco, S.B., Goto, Y., and Lodge, D.J. (2007). Regulation of
and anesthetized rats. Neuropsychopharmacology 37, 1559–1571.
firing of dopaminergic neurons and control of goal-directed behaviors. Trends
Neurosci. 30, 220–227. Lammel, S., Ion, D.I., Roeper, J., and Malenka, R.C. (2011). Projection-specific
modulation of dopamine neuron synapses by aversive and rewarding stimuli.
Haber, S.N. (2014). The place of dopamine in the cortico-basal ganglia circuit.
Neuron 70, 855–862.
Neuroscience 282C, 248–257.
Lammel, S., Lim, B.K., Ran, C., Huang, K.W., Betley, M.J., Tye, K.M., Deisser-
Hall, M.E., and Mayer, M.A. (1975). Effects of alpha methyl-para-tyrosine on oth, K., and Malenka, R.C. (2012). Input-specific control of reward and aver-
the recall of a passive avoidance response. Pharmacol. Biochem. Behav. 3, sion in the ventral tegmental area. Nature 491, 212–217.
579–582.
Lüscher, C., and Malenka, R.C. (2011). Drug-evoked synaptic plasticity
Harnett, M.T., Bernier, B.E., Ahn, K.C., and Morikawa, H. (2009). Burst-timing- in addiction: from molecular changes to circuit remodeling. Neuron 69, 650–663.
dependent plasticity of NMDA receptor-mediated transmission in midbrain
dopamine neurons. Neuron 62, 826–838. Mameli, M., Balland, B., Luján, R., and Lüscher, C. (2007). Rapid synthesis and
synaptic insertion of GluR2 for mGluR-LTD in the ventral tegmental area. Sci-
Harris, G.C., and Aston-Jones, G. (2003). Altered motivation and learning ence 317, 530–533.
following opiate withdrawal: evidence for prolonged dysregulation of reward
processing. Neuropsychopharmacology 28, 865–871. Mameli, M., Halbout, B., Creton, C., Engblom, D., Parkitna, J.R., Spanagel, R.,
and Lüscher, C. (2009). Cocaine-evoked synaptic plasticity: persistence in the
Hjelmstad, G.O., Xia, Y., Margolis, E.B., and Fields, H.L. (2013). Opioid modu- VTA triggers adaptations in the NAc. Nat. Neurosci. 12, 1036–1041.
lation of ventral pallidal afferents to ventral tegmental area neurons.
J. Neurosci. 33, 6454–6459. Mameli, M., Bellone, C., Brown, M.T., and Lüscher, C. (2011). Cocaine inverts
rules for synaptic plasticity of glutamate transmission in the ventral tegmental
Hnasko, T.S., Hjelmstad, G.O., Fields, H.L., and Edwards, R.H. (2012). Ventral area. Nat. Neurosci. 14, 414–416.
tegmental area glutamate neurons: electrophysiological properties and pro-
jections. J. Neurosci. 32, 15076–15085. Margolis, E.B., Mitchell, J.M., Ishikawa, J., Hjelmstad, G.O., and Fields, H.L.
(2008). Midbrain dopamine neurons: projection target determines action po-
Hollerman, J.R., and Schultz, W. (1998). Dopamine neurons report an error in tential duration and dopamine D(2) receptor inhibition. J. Neurosci. 28, 8908–
the temporal prediction of reward during learning. Nat. Neurosci. 1, 304–309. 8913.

Inglis, F.M., and Moghaddam, B. (1999). Dopaminergic innervation of the McCutcheon, J.E., Beeler, J.A., and Roitman, M.F. (2012). Sucrose-predictive
amygdala is highly responsive to stress. J. Neurochem. 72, 1088–1094. cues evoke greater phasic dopamine release than saccharin-predictive cues.
Synapse 66, 346–351.
Isaac, J.T., Ashby, M.C., and McBain, C.J. (2007). The role of the GluR2
subunit in AMPA receptor function and synaptic plasticity. Neuron 54, Mendelson, J. (1966). The role of hunger in the T-maze learning for food by
859–871. rats. J. Comp. Psychol. 62, 341–349.

Neuron 86, June 3, 2015 ª2015 Elsevier Inc. 1155


Neuron

Review
Miller, N.E., and Hunt, J.M. (1944). Experimental Studies of Conflict. Personal- Schultz, W., Tremblay, L., and Hollerman, J.R. (1998). Reward prediction in
ity and the Behavior Disorders (Oxford, England: Ronald Press), pp. 431–465. primate basal ganglia and frontal cortex. Neuropharmacology 37, 421–429.

Montague, P.R., Hyman, S.E., and Cohen, J.D. (2004). Computational roles for Sesack, S.R., and Grace, A.A. (2010). Cortico-Basal Ganglia reward network:
dopamine in behavioural control. Nature 431, 760–767. microcircuitry. Neuropsychopharmacology 35, 27–47.

Morales, M., and Root, D.H. (2014). Glutamate neurons within the midbrain Shiflett, M.W., and Balleine, B.W. (2011). Molecular substrates of action con-
dopamine regions. Neuroscience 282C, 60–68. trol in cortico-striatal circuits. Prog. Neurobiol. 95, 1–13.

Morgan, M.J. (1974). Resistance to satiation. Anim. Behav. 22, 449–466. Smiley, J.F., and Goldman-Rakic, P.S. (1993). Heterogeneous targets of dopa-
mine synapses in monkey prefrontal cortex demonstrated by serial section elec-
O’Donnell, P., and Grace, A.A. (1995). Synaptic interactions among excitatory tron microscopy: a laminar analysis using the silver-enhanced diaminobenzidine
afferents to nucleus accumbens neurons: hippocampal gating of prefrontal sulfide (SEDS) immunolabeling technique. Cereb. Cortex 3, 223–238.
cortical input. J. Neurosci. 15, 3622–3639.
Smith, K.S., Berridge, K.C., and Aldridge, J.W. (2011). Disentangling pleasure
O’Donnell, P., Greene, J., Pabello, N., Lewis, B.L., and Grace, A.A. (1999). from incentive salience and learning signals in brain reward circuitry. Proc.
Modulation of cell firing in the nucleus accumbens. Ann. N Y Acad. Sci. 877, Natl. Acad. Sci. USA 108, E255–E264.
157–175.
Stamatakis, A.M., Jennings, J.H., Ung, R.L., Blair, G.A., Weinberg, R.J., Neve,
Oleson, E.B., Gentry, R.N., Chioma, V.C., and Cheer, J.F. (2012). Subsecond R.L., Boyce, F., Mattis, J., Ramakrishnan, C., Deisseroth, K., and Stuber, G.D.
dopamine release in the nucleus accumbens predicts conditioned punishment (2013). A unique population of ventral tegmental area neurons inhibits the
and its successful avoidance. J. Neurosci. 32, 14804–14808. lateral habenula to promote reward. Neuron 80, 1039–1053.

Paladini, C.A., and Roeper, J. (2014). Generating bursts (and pauses) in the Steffensen, S.C., Svingos, A.L., Pickel, V.M., and Henriksen, S.J. (1998). Elec-
dopamine midbrain neurons. Neuroscience 282C, 109–121. trophysiological characterization of GABAergic neurons in the ventral
tegmental area. J. Neurosci. 18, 8003–8015.
Paquet, M., and Smith, Y. (2003). Group I metabotropic glutamate receptors in
the monkey striatum: subsynaptic association with glutamatergic and dopami- Steinberg, E.E., Keiflin, R., Boivin, J.R., Witten, I.B., Deisseroth, K., and Janak,
nergic afferents. J. Neurosci. 23, 7659–7669. P.H. (2013). A causal link between prediction errors, dopamine neurons and
learning. Nat. Neurosci. 16, 966–973.
Paquet, M., Tremblay, M., Soghomonian, J.J., and Smith, Y. (1997). AMPA and
NMDA glutamate receptor subunits in midbrain dopaminergic neurons in the Stuber, G.D., Roitman, M.F., Phillips, P.E., Carelli, R.M., and Wightman, R.M.
squirrel monkey: an immunohistochemical and in situ hybridization study. (2005a). Rapid dopamine signaling in the nucleus accumbens during contin-
J. Neurosci. 17, 1377–1396. gent and noncontingent cocaine administration. Neuropsychopharmacology
30, 853–863.
Phillips, P.E., Stuber, G.D., Heien, M.L., Wightman, R.M., and Carelli, R.M.
(2003). Subsecond dopamine release promotes cocaine seeking. Nature Stuber, G.D., Wightman, R.M., and Carelli, R.M. (2005b). Extinction of cocaine
422, 614–618. self-administration reveals functionally and temporally distinct dopaminergic
signals in the nucleus accumbens. Neuron 46, 661–669.
Ranaldi, R. (2014). Dopamine and reward seeking: the role of ventral tegmental
area. Rev. Neurosci. 25, 621–630. Stuber, G.D., Klanker, M., de Ridder, B., Bowers, M.S., Joosten, R.N., Feen-
stra, M.G., and Bonci, A. (2008). Reward-predictive cues enhance excitatory
Roitman, M.F., Stuber, G.D., Phillips, P.E., Wightman, R.M., and Carelli, R.M. synaptic strength onto midbrain dopamine neurons. Science 321, 1690–
(2004). Dopamine operates as a subsecond modulator of food seeking. 1692.
J. Neurosci. 24, 1265–1271.
Takahata, R., and Moghaddam, B. (1998). Glutamatergic regulation of basal and
Roitman, M.F., Wheeler, R.A., Wightman, R.M., and Carelli, R.M. (2008). Real- stimulus-activated dopamine release in the prefrontal cortex. J. Neurochem. 71,
time chemical responses in the nucleus accumbens differentiate rewarding 1443–1449.
and aversive stimuli. Nat. Neurosci. 11, 1376–1377.
Takahata, R., and Moghaddam, B. (2000). Target-specific glutamatergic regu-
Romo, R., and Schultz, W. (1990). Dopamine neurons of the monkey midbrain: lation of dopamine neurons in the ventral tegmental area. J. Neurochem. 75,
contingencies of responses to active touch during self-initiated arm move- 1775–1778.
ments. J. Neurophysiol. 63, 592–606.
Tan, K.R., Yvon, C., Turiault, M., Mirzabekov, J.J., Doehner, J., Labouèbe, G.,
Root, D.H., Mejias-Aponte, C.A., Zhang, S., Wang, H.L., Hoffman, A.F., Lup- Deisseroth, K., Tye, K.M., and Lüscher, C. (2012). GABA neurons of the VTA
ica, C.R., and Morales, M. (2014). Single rodent mesohabenular axons release drive conditioned place aversion. Neuron 73, 1173–1183.
glutamate and GABA. Nat. Neurosci. 17, 1543–1551.
Thomas, M.J., Malenka, R.C., and Bonci, A. (2000). Modulation of long-term
Rosenkranz, J.A., and Grace, A.A. (2002). Dopamine-mediated modulation of depression by dopamine in the mesolimbic system. J. Neurosci. 20, 5581–
odour-evoked amygdala potentials during pavlovian conditioning. Nature 417, 5586.
282–287.
Tong, Z.Y., Overton, P.G., and Clark, D. (1996). Antagonism of NMDA recep-
Saal, D., Dong, Y., Bonci, A., and Malenka, R.C. (2003). Drugs of abuse and tors but not AMPA/kainate receptors blocks bursting in dopaminergic neurons
stress trigger a common synaptic adaptation in dopamine neurons. Neuron induced by electrical stimulation of the prefrontal cortex. J. Neural Transm.
37, 577–582. 103, 889–904.

Schultz, W. (1986). Responses of midbrain dopamine neurons to behavioral Tritsch, N.X., Ding, J.B., and Sabatini, B.L. (2012). Dopaminergic neurons
trigger stimuli in the monkey. J. Neurophysiol. 56, 1439–1461. inhibit striatal output through non-canonical release of GABA. Nature 490,
262–266.
Schultz, W. (2006). Behavioral theories and the neurophysiology of reward.
Annu. Rev. Psychol. 57, 87–115. Tye, K.M., Mirzabekov, J.J., Warden, M.R., Ferenczi, E.A., Tsai, H.C., Finkel-
stein, J., Kim, S.Y., Adhikari, A., Thompson, K.R., Andalman, A.S., et al.
Schultz, W. (2011). Potential vulnerabilities of neuronal reward, risk, and deci- (2013). Dopamine neurons modulate neural encoding and expression of
sion mechanisms to addictive drugs. Neuron 69, 603–617. depression-related behaviour. Nature 493, 537–541.

Schultz, W., Apicella, P., and Ljungberg, T. (1993). Responses of monkey Ungless, M.A. (2004). Dopamine: the salient issue. Trends Neurosci. 27,
dopamine neurons to reward and conditioned stimuli during successive steps 702–706.
of learning a delayed response task. J. Neurosci. 13, 900–913.
Ungless, M.A., Whistler, J.L., Malenka, R.C., and Bonci, A. (2001). Single
Schultz, W., Dayan, P., and Montague, P.R. (1997). A neural substrate of pre- cocaine exposure in vivo induces long-term potentiation in dopamine neurons.
diction and reward. Science 275, 1593–1599. Nature 411, 583–587.

1156 Neuron 86, June 3, 2015 ª2015 Elsevier Inc.


Neuron

Review
van Zessen, R., Phillips, J.L., Budygin, E.A., and Stuber, G.D. (2012). Activation Yamaguchi, T., Wang, H.L., Li, X., Ng, T.H., and Morales, M. (2011). Mesocor-
of VTA GABA neurons disrupts reward consumption. Neuron 73, 1184–1194. ticolimbic glutamatergic pathway. J. Neurosci. 31, 8476–8490.

Wise, R.A. (2004). Dopamine, learning and motivation. Nat. Rev. Neurosci. 5,
483–494. Yuan, T., Mameli, M., O’ Connor, E.C., Dey, P.N., Verpelli, C., Sala, C., Perez-
Otano, I., Lüscher, C., and Bellone, C. (2013). Expression of cocaine-evoked
Wise, R.A., and Kiyatkin, E.A. (2011). Differentiating the rapid actions of synaptic plasticity by GluN3A-containing NMDA receptors. Neuron 80,
cocaine. Nat. Rev. Neurosci. 12, 479–484. 1025–1038.

Wise, R.A., Spindler, J., deWit, H., and Gerberg, G.J. (1978). Neuroleptic- Zhang, H., and Sulzer, D. (2003). Glutamate spillover in the striatum depresses
induced ‘‘anhedonia’’ in rats: pimozide blocks reward quality of food. Science dopaminergic transmission by activating group I metabotropic glutamate re-
201, 262–264. ceptors. J. Neurosci. 23, 10585–10592.
Wolf, M.E. (1998). The role of excitatory amino acids in behavioral sensitization
to psychomotor stimulants. Prog. Neurobiol. 54, 679–720. Zweifel, L.S., Fadok, J.P., Argilli, E., Garelick, M.G., Jones, G.L., Dickerson,
T.M., Allen, J.M., Mizumori, S.J., Bonci, A., and Palmiter, R.D. (2011). Activa-
Yamaguchi, T., Sheen, W., and Morales, M. (2007). Glutamatergic neurons are tion of dopamine neurons is critical for aversive conditioning and prevention of
present in the rat ventral tegmental area. Eur. J. Neurosci. 25, 106–118. generalized anxiety. Nat. Neurosci. 14, 620–626.

Neuron 86, June 3, 2015 ª2015 Elsevier Inc. 1157

You might also like