ArticlePDF Available

Dopaminergic Terminals in the Nucleus Accumbens But Not the Dorsal Striatum Corelease Glutamate

Authors:

Abstract and Figures

Coincident signaling by dopamine and glutamate is thought to be crucial for a variety of motivated behaviors. Previous work has suggested that some midbrain dopamine neurons are themselves capable of glutamate corelease, but this phenomenon remains poorly understood. Here, we expressed the light-activated cation channel Channelrhodopsin-2 (ChR2) in genetically defined midbrain dopamine neurons to stimulate exocytosis specifically from dopaminergic terminals in both the nucleus accumbens (NAc) shell and dorsal striatum of brain slices from adult mice. Optical activation resulted in robust glutamate-mediated EPSCs in all medium spiny neurons examined in the NAc shell. In contrast, optically evoked glutamatergic currents were nearly undetectable in the dorsal striatum. Further, we used a conditional knock-out mouse lacking vesicular glutamate transporter 2 (VGLUT2) specifically in dopamine neurons to determine whether VGLUT2 is required for the exocytotic release of glutamate from dopamine neurons. Our data show that conditional knock-out completely abolished all optically evoked glutamate release. These results provide definitive physiological evidence for VGLUT2-mediated glutamate release by mature dopamine neurons projecting to the NAc shell, but not to the dorsal striatum. Thus, the unique ability of NAc-projecting dopamine neurons to synchronously activate both dopamine and glutamate receptors may have crucial implications for the ability to respond to motivationally significant stimuli.
Content may be subject to copyright.
Brief Communications
Dopaminergic Terminals in the Nucleus Accumbens But Not
the Dorsal Striatum Corelease Glutamate
Garret D. Stuber,
1,2
* Thomas S. Hnasko,
3
* Jonathan P. Britt,
1,2
Robert H. Edwards,
2,3
and Antonello Bonci
1,2
1
Ernest Gallo Clinic and Research Center, University of California, San Francisco, Emeryville, California 94608, and Departments of
2
Neurology,
and
3
Physiology, University of California, San Francisco, San Francisco, California 94143
Coincident signaling by dopamine and glutamate is thought to be crucial for a variety of motivated behaviors. Previous work has
suggested that some midbrain dopamine neurons are themselves capable of glutamate corelease, but this phenomenon remains
poorly understood. Here, we expressed the light-activated cation channel Channelrhodopsin-2 (ChR2) in genetically defined
midbrain dopamine neurons to stimulate exocytosis specifically from dopaminergic terminals in both the nucleus accumbens
(NAc) shell and dorsal striatum of brain slices from adult mice. Optical activation resulted in robust glutamate-mediated EPSCs in
all medium spiny neurons examined in the NAc shell. In contrast, optically evoked glutamatergic currents were nearly undetect-
able in the dorsal striatum. Further, we used a conditional knock-out mouse lacking vesicular glutamate transporter 2 (VGLUT2)
specifically in dopamine neurons to determine whether VGLUT2 is required for the exocytotic release of glutamate from dopamine
neurons. Our data show that conditional knock-out completely abolished all optically evoked glutamate release. These results
provide definitive physiological evidence for VGLUT2-mediated glutamate release by mature dopamine neurons projecting to the
NAc shell, but not to the dorsal striatum. Thus, the unique ability of NAc-projecting dopamine neurons to synchronously activate
both dopamine and glutamate receptors may have crucial implications for the ability to respond to motivationally significant
stimuli.
Introduction
The role that a neuron plays within a circuit depends to a large extent
on the neurotransmitter(s) released. Midbrain dopamine neurons
project to a variety of forebrain targets, including the dorsal striatum
and nucleus accumbens, where the release of dopamine is thought to
contribute to the generation of motor function and motivated be-
haviors (Wise, 2004). As a variety of adaptive behaviors are ascribed
to specific striatal subregions, it is possible that the striatal microcir-
cuits that underlie these specific behavioral processes have distinct
properties, such as differences in afferent or efferent connectivity.
Accumulating evidence suggests that a subset of midbrain do-
pamine neurons may also corelease the excitatory neurotrans-
mitter glutamate both in vitro (Sulzer et al., 1998; Joyce and
Rayport, 2000; Dal Bo et al., 2004) and in vivo (Chuhma et al.,
2004; Lavin et al., 2005; Hnasko et al., 2010). However, this pos-
sibility has generated controversy due to difficulty stimulating
selectively dopamine neurons or their terminals.
Vesicular glutamate transporters package glutamate into syn-
aptic vesicles and are both necessary and sufficient for the exocy-
totic release of glutamate from neurons (Reimer and Edwards,
2004; Takamori, 2006). In culture, dopamine neurons immuno-
stain for the vesicular glutamate transporter VGLUT2 (Dal Bo et
al., 2004). However the histochemical evidence for VGLUT2 ex-
pression by dopamine neurons in vivo remains less clear. One
study found substantial colocalization of mRNAs encoding the
dopaminergic marker tyrosine hydroxylase (TH) and VGLUT2
in the ventral midbrain, with a higher incidence of colocalization
in medial and caudal regions of the A10 dopamine cell group
(Kawano et al., 2006). A second report detected colocalization only
very rarely, but medial aspects were not examined (Yamaguchi et al.,
2007). Using single-cell RT-PCR and immuno-electron microscopy,
Trudeau and colleagues detected substantial colocalization of
VGLUT2 transcript in dopamine neurons (Mendez et al., 2008) and
VGLUT2 protein in THterminals of the nucleus accumbens (Des-
carries et al., 2008; Bérubé-Carriére et al., 2009). However, the level
of colocalization declined significantly across development (Mendez
et al., 2008; Be´rube´ -Carrie`re et al., 2009), suggesting that glutamate
corelease may similarly downregulate in adulthood.
To determine whether dopamine neurons release synapti-
cally relevant amounts of glutamate in the adult brain, a
method to selectively stimulate exocytotic release from dopa-
mine terminals is required. We therefore used a multifaceted
strategy, combining optogenetics and conditional gene dis-
ruption to selectively stimulate dopaminergic terminals while
recording postsynaptic currents from medium spiny neurons
(MSNs) in either the ventral or dorsal striatum of control or
conditional knock-out mice lacking VGLUT2 specifically in
dopamine neurons.
Received April 6, 2010; revised May 5, 2010; accepted May 11, 2010.
Thisworkwas supportedby ABMRFand National Alliancefor Researchon SchizophreniaandDepression (G.D.S.),
the A. P. Giannini Foundation (T.S.H.), and the State of California for Medical Research on Alcohol and Substance
Abuse through the University of California, San Francisco (UCSF) (A.B.). We thank Kurt Thorn and the Nikon Imaging
Center at UCSF for assistance and Karl Deisseroth for the AAV-DIO-ChR2-mcherry construct.
*G.D.S. and T.S.H. contributed equally to this work.
Correspondenceshould be addressedto Dr.Antonello Bonci,Department ofNeurology andErnest GalloClinic and
Research Center, University of California, San Francisco, Emeryville, CA 94608. E-mail: antonello.bonci@ucsf.edu.
DOI:10.1523/JNEUROSCI.1754-10.2010
Copyright © 2010 the authors 0270-6474/10/308229-05$15.00/0
The Journal of Neuroscience, June 16, 2010 30(24):8229 – 8233 8229
Materials and Methods
Experimental subjects. Age-matched, sex-matched
adult male and female mice (25 g) were used
as subjects. Control mice carried one copy of
cre recombinase driven by dopamine trans-
porter regulatory elements (Zhuang et al.,
2005) and one conditional VGLUT2 allele
(Hnasko et al., 2010) (Slc6a3
/cre
; Slc17a6
/lox
).
Conditional VGLUT2 knock-out mice were
identical but carried two conditional VGLUT2
alleles (Slc6a3
/cre
; Slc17a6
lox/lox
). Mice were
group housed in a colony maintained with a
standard 12 h light/dark cycle and given food
and water ad libitum. Experiments were con-
ducted in accordance with the Guide for the
Care and Use of Laboratory Animals, as adopted
by the National Institutes of Health, and with
approval of the University of California, San
Francisco Institutional Animal Care and Use
Committee.
Stereotaxic recombinant adeno-associated vi-
rus injection. Methods were adapted from (Tsai
et al., 2009). Briefly, mice were anesthetized
with ketamine/xylazine, placed in a stereotaxic
frame (Kopf), the skull leveled, small holes
drilled, and 1
l (0.25
l/min) of AAV5-EF1
-
DIO-ChR2-mcherry (310
12
genomes/ml)
was injected using a Hamilton syringe bilater-
ally into the VTA (coordinates in mm relative
to bregma: 3.25 anteroposterior, 4.50 dor-
soventral, and 0.5 mediolateral). Mice were
allowed to recover for at least 3 weeks before
further procedures.
Immunohistochemistry. Mice were perfused
with cold PBS followed by 4% paraformalde-
hyde; their brains were removed, postfixed,
cryoprotected in 30% sucrose, and frozen in
superchilled isopentane; and 35
m sections
were cut on a cryostat and floated in PBS. Sec-
tions were rinsed with PBS containing 0.2%
Triton X-100, blocked 1 h with 4% normal
donkey serum, and incubated overnight at 4°C
in sheep anti-TH (Pel-Freze) and rabbit anti-
DsRed (Clontech), both at 1:2000 dilution in
blocking solution. Sections were rinsed, incubated 2 h with Cy2-
conjugated anti-sheep and DyLite549-conjugated anti-rabbit secondar-
ies (1:500, Jackson ImmunoResearch), rinsed again, mounted on slides,
dehydrated through alcohol/xylenes, and coverslipped with DPX.
Images were collected using a Nikon Eclipse Ti-E motorized inverted
microscope equipped with epifluorescence and a Photometrics Coolsnap
HQ2 camera or a Nikon FN1 upright C1si spectral confocal microscope.
For quantification of TH/ChR2-mcherry
neurons in the VTA, confocal
images were taken using a 60objective. ChR2-mcherry
neurons were
identified and then scored for TH content.
Brain slice preparation. Mice were anesthetized and rapidly decapi-
tated. Dissected brains were transferred to ice-cold artificial CSF (ACSF)
containing the following (in mM): 75 sucrose, 87 NaCl, 2.5 KCl, 7 MgCl
2
,
0.5 CaCl
2
, 25 NaHCO
3
, 1.25 NaH
2
PO
4
, and 1 ascorbic acid (saturated
with 95% O
2
and 5% CO
2
). Coronal sections of the striatum (200
m)
were cut with a vibratome (VT1200, Leica). Slices were incubated for at
least 30 min in a holding chamber containing the following (in mM): 126
NaCl, 2.5 KCl, 1.2 MgCl
2
, 2.5 CaCl
2
, 26 NaHCO
3
, 1.2 NaH
2
PO
4
,11
glucose, and 1 ascorbic acid (saturated with 95% O
2
and 5% CO
2
). Dur-
ing recording, slices were superfused (2 ml/min) with this same ACSF at
32°C but with picrotoxin (100
M, to block GABA
A
receptor-mediated
synaptic currents) and without ascorbic acid.
Patch-clamp electrophysiology. Whole-cell voltage-clamp recordings
from MSNs located in the nucleus accumbens (NAc) shell and dorsal
striatum (DS) were obtained under visual control on a differential inter-
ference contrast, upright microscope with infrared illumination. Record-
ings were obtained using 3– 6 Mresistance pipettes backfilled with
internal solution containing the following (in mM): 120 CsCH
3
SO
3
,20
HEPES, 0.4 EGTA, 2.8 NaCl, 5 N(CH
2
CH
3
)
4
Cl, 2.5 Mg-ATP, and 0.25
Mg-GTP, pH 7.3. Currents were measured using either an Axopatch 1D
or 200A amplifier (2 kHz low-pass Bessel filter) with a DigiData 1440
interface (5 kHz digitization) and pClamp software (Molecular Devices).
MSNs, identified by their morphology and hyperpolarized resting mem-
brane potential, were voltage clamped at 70 mV. An optical fiber (200
m core diameter, 0.2 numerical aperture) coupled to a diode-pumped
solid-state 473 nm laser was placed 200
m from the site of recording,
and was used to deliver optical stimulation at a frequency of 0.1 Hz.
Stimulus intensity ranged from 1 to 30 mW with a pulse duration of 5 ms.
Series resistance (5–20 M) was monitored online witha5mVhyper-
polarizing step (50 ms) given after each stimulus.
To quantify EPSC amplitudes, six sweeps were collected from each cell
at each light intensity (0, 1, 2, 5, and 10 mW for input/output experiment,
30 mW for maximal stimulation and pharmacology experiments).
Sweeps were then averaged offline to determine EPSC amplitude at a
particular intensity. Light stimulus intensities were presented in a ran-
dom order and counterbalanced across recordings. For AMPA receptor
(AMPAR) antagonism experiments, 10
M6,7-dinitroquinoxaline-2,3-
dione (DNQX) was used, and for dopamine (DA) receptor antagonism
experiments, 2
MR-()-7-chloro-8-hydroxy-3-methyl-1-phenyl-2,3,4,5-
Figure 1. Expression of ChR2-mcherry in midbrain dopamine neurons and their projections to the striatum. A, Immunostaining
for ChR2-mcherry (red) and TH (green) in the VTA/SN 30 d after injection of AAV-DIO-ChR2-mcherry in a mouse expressing cre
recombinase under the control of the DAT promoter. B, Confocal microscopy shows that nearly all ChR2-mcherry-expressing
neurons in the VTA colabel for TH. C, Staining as described above in a coronal plane including both DS and NAc shell. D, Confocal
images of dopaminergic fibers in the NAc shell indicate that essentially all ChR2-mcherry-labeled puncta also contain TH.
8230 J. Neurosci., June 16, 2010 30(24):8229 – 8233 Stuber et al. Optogenetic Glutamate Corelease from Dopamine Terminals
tetrahydro-1H-3-benzazepine (SCH23390) and
2
Mraclopride were used to block D
1
and D
2
receptors, respectively. Experimenters were blind
to the animal’s genotype during data acquisition.
Fast-scan cyclic voltammetry. Methods were
adapted from previous work (Stuber et al.,
2008). T-650 carbon fiber microelectrodes
(100 –200
m in length) were used for detec-
tion of dopamine. Electrodes were placed in
the NAc shell or dorsal striatum of Slc6a3
/cre
;
Slc17a6
/lox
mice. The potential applied to the
electrode was ramped from 0.4Vto1.3 V
to 0.4 V versus a Ag/AgCl reference every 100
ms at a rate of 400 V/s, and resulting electro-
chemical data were acquired using custom-
written software in LabVIEW. Electrochemical
data were low-pass filtered at 1 kHz offline.
Background-subtracted cyclic voltammograms
were generated immediately after optical stimu-
lation of the slice, and were characteristic of do-
pamine (peak oxidation potential 600 –700 mV).
Current resulting from the oxidation of dopa-
mine (monitored at the peak oxidation potential
determined from cyclic voltammograms) were
then converted into concentration changes using
a calibration factor of 10 nA/
MDA.
Results
An adeno-associated virus encoding a
conditional allele of the light-activated
cation channel channelrhodopsin-2 (ChR2)
activated by cre-mediated recombination
(AAV5-DIO-ChR2-mcherry) (Tsai et al.,
2009) was bilaterally injected into the ven-
tral tegmental area (VTA) of adult mice
(2– 4 months old) expressing cre recom-
binase under the control of dopamine
transporter (DAT) regulatory sequences
(Zhuang et al., 2005). Four weeks after in-
jection, immunohistochemistry revealed
robust, selective expression of ChR2-mch-
erry in VTA dopamine neurons (Fig. 1 A,B).
Indeed, 99.5% of ChR2-mcherry
neurons
identified in the VTA (n409 cells) were
also labeled for the catecholamine biosyn-
thetic enzyme TH. Furthermore, essen-
tially all ChR2-mcherry-labeled fibers and
puncta in striatal terminals costained for
TH (Fig. 1C,D).
We then prepared coronal sections
through striatal regions from injected
mice and light-evoked stimulation was
used to depolarize and stimulate neuro-
transmitter release from the dopaminer-
gic terminals selectively expressing ChR2.
Whole-cell voltage-clamp recordings from
visually identified MSNs revealed that op-
tical stimulation of dopaminergic termi-
nals in the NAc shell resulted in EPSCs
10 pA in 24/24 neurons, which in-
creased in amplitude as a function of light
stimulus intensity (Fig. 2A,B). Light-
evoked EPSCs in the NAc shell were
blocked by AMPAR antagonism (Fig.
3A,C). The rapid response to light sug-
Figure 2. Dopaminergic terminals in the nucleus accumbens shell but not dorsal striatum corelease glutamate. A, Example
EPSCsrecordedin theNAcshell at70 mV followinga5msbluelight pulseatmaximal intensity.B,Light-evoked EPSCamplitudes
recorded in the NAc shell were significantly greater in DAT:VGLUT2 heterozygous mice than in conditional knock-outs (n9 –24
cells per group; 2-way ANOVA for the interaction of genotype and light intensity: F
(4,128)
2.56, p0.05). C, Example EPSCs
recorded in the DS at 70 mV followinga5msblue light pulse at maximal intensity. D, Light-evoked EPSC amplitudes recorded
in the DS were not significantly different in DAT:VGLUT2 heterozygous mice versus conditional knock-outs as a function of light
intensity (2-way ANOVA for the interaction of genotype and light intensity: F
(4,87)
0.31, p0.91). EPSC amplitudes (indepen-
dent of light intensity) were significantly different across genotypes (F
(1,87)
12.80, p0.006).
Figure 3. EPSCs evoked by dopamine terminal stimulation in the nucleus accumbens shell are blocked by AMPAR antagonism
but unaffected by D
1
R/D
2
R antagonism. A, Example EPSCs recorded at 70 mV in the NAc shell from DAT:VGLUT2 heterozygous
mice before and after bath application of DNQX. B, Example EPSCs recorded at 70 mV in the NAc shell from DAT:VGLUT2
heterozygous mice before and after bath application of SCH23390/raclopride. C, Light-evoked EPSCs from dopamine terminals
were significantly reduced by DNQX bath application (n5 cells; t
(4)
2.61; p0.05) but not by SCH23390/raclopride (n9
cells; t
(8)
0.40; p0.70). All data indicate mean SEM.
Stuber et al. Optogenetic Glutamate Corelease from Dopamine Terminals J. Neurosci., June 16, 2010 30(24):8229 – 8233 • 8231
gests a monosynaptic event, but to ensure
that the glutamate-mediated currents did
not reflect the postsynaptic action of do-
pamine, we applied D
1
- and D
2
-type do-
pamine receptor antagonists to the bath.
The addition of SCH23390 and raclopride
did not significantly affect light-evoked
EPSCs (Fig. 3 B,C), indicating that they do
not require the light-evoked release of do-
pamine. Rather, dopamine neurons them-
selves appear to release glutamate.
The ability of dopamine neurons to re-
lease glutamate implies that they express a
vesicular glutamate transporter. To deter-
mine whether VGLUT2 expression by do-
pamine neurons is required for their
release of glutamate, we performed the
same experiments in conditional knock-
out (cKO) mice (Slc6a3
/cre
; Slc17a6
lox/lox
)
where cre recombinase drives both selec-
tive excision in dopamine neurons of the
gene encoding VGLUT2 (Hnasko et al.,
2010) and, following virus injection, the
selective expression of ChR2 in dopamine
neurons. In contrast to control mice, light
never evoked EPSCs in cKO mice (Fig.
2A,B), demonstrating that VGLUT2 is re-
quired for glutamate packaging and re-
lease from dopaminergic terminals.
Although previous results have sug-
gested that mesolimbic dopamine neu-
rons release glutamate (Chuhma et al.,
2004; Hnasko et al., 2010), it is unknown
whether dopamine neurons that project
to the DS also have the capacity to release
glutamate. Therefore, we also examined
brain slices through the DS. Optical stim-
ulation of dopaminergic terminals in the
DS resulted in 10-fold smaller mean glutamate-mediated cur-
rents than in the NAc shell of control mice, with no detectable
release in the DS from mice lacking VGLUT2 in dopamine neu-
rons (Fig. 2C,D). Importantly, light-evoked dopamine release as
measured by fast-scan cyclic voltammetry was detected in both
the DS (Fig. 4A,C) and NAc shell (Fig. 4B,D) of virally injected
mice. Within the striatum, glutamate corelease is thus relatively
restricted to dopamine terminals in the ventral striatum.
Discussion
Our results provide direct physiological evidence that mature
dopamine neurons projecting to the NAc shell corelease gluta-
mate. Previous studies have suggested that mesolimbic dopamine
neurons release glutamate (Chuhma et al., 2004, 2009; Lavin et
al., 2005; Hnasko et al., 2010), but electrical stimulation nonspe-
cifically activates all VTA neurons, making it difficult to exclude
glutamate release from nondopaminergic neurons. Indeed, there
exist VGLUT2-positive neurons in the VTA that do not colabel
for TH (Kawano et al., 2006; Yamaguchi et al., 2007; Nair-
Roberts et al., 2008). Although some of these neurons appear to
project locally (Dobi et al., 2010), it is unknown whether they
may also project to striatal and/or limbic brain regions. Our abil-
ity to drive ChR2 expression selectively in dopamine neurons of
the VTA and thus to selectively activate dopamine terminals in
striatal regions with light eliminates the possibility of inadver-
tently activating these nondopaminergic VTA glutamate neu-
rons. Essentially all ChR2-mcherry-expressing neurons in the
midbrain and puncta in the striatum were also labeled for TH
by confocal microscopy (Fig. 1B,D). Additionally, no ChR2-
mcherry-expressing neurons were seen in brain regions known to
send glutamatergic projections to the NAc shell, such as cortex,
thalamus, hippocampus, amygdala, or lateral hypothalamus.
Surprisingly, we did observe small clusters of ChR2-expressing
neurons in both the lateral habenula and in the ventral premam-
millary nucleus (data not shown), but these neurons are not
known to project to the NAc shell.
Due to technical limitations, earlier electrophysiological stud-
ies of glutamate corelease in the NAc shell have reported relatively
small AMPAR currents limited to brain slices made from young
(3 weeks old) animals (Chuhma et al., 2004; Hnasko et al.,
2010). Given that VGLUT2 expression by dopamine neurons
appears to upregulate in culture (Dal Bo et al., 2004; Mendez et
al., 2008) and downregulate over development in vivo (Mendez et
al., 2008; Be´rube´-Carrie`re et al., 2009), it has remained unclear
whether mature dopamine neurons also corelease glutamate.
However, we now find that VGLUT2 in dopamine neurons is
required for glutamate release from dopaminergic terminals in
the NAc shell of adult mice, demonstrating that VGLUT2-
dependent glutamate corelease persists into adulthood. Although
some evidence suggests that only a small fraction of dopamine
Figure 4. Optical stimulation of dopamine terminals in dorsal striatum and nucleus accumbens shell results in dopamine
release. A, Example trace of dopamine release resulting from optical stimulation of dopamine terminals in the DS (1 pulse, 5 ms
pulse duration). Inset shows background-subtracted cyclic voltammograms taken immediately after optical stimulation. The color
plot below shows current measured as the electrode across the entire scanned potential range. B, Example trace of dopamine
release resulting from optical stimulation of dopamine terminals in the NAc shell. C, Average dopamine release (solid line; dashed
lines indicate SEM) in the dorsal striatum following one-pulse optical stimulation (n7 recording sites). D, Average dopamine
release in the nucleus accumbens shell following one-pulse optical stimulation (n5 recording sites).
8232 J. Neurosci., June 16, 2010 30(24):8229 – 8233 Stuber et al. Optogenetic Glutamate Corelease from Dopamine Terminals
neurons express VGLUT2 in adult rodents (Yamaguchi et al.,
2007), we see ubiquitous light-evoked glutamatergic currents in
MSNs of the NAc shell. Therefore, it is likely that a significant
number of dopamine neurons in the adult VTA express VGLUT2
at levels too low to detect by conventional means, or alternatively,
a small percentage of VGLUT2-expressing dopamine neurons
may send axon collaterals to essentially all MSNs in the NAc shell.
Although we were able to observe substantial ChR2-mcherry
expression and light-evoked dopamine release in the DS, gluta-
mate currents were much smaller and less frequent in the DS
relative to the NAc shell. These results are consistent with a pre-
vious report (Kawano et al., 2006) showing the absence of
VGLUT2 in the substantia nigra pars compacta compared to the
VTA. More recently, we reported that cKO mice lacking
VGLUT2 selectively in dopamine neurons show a presynaptic
reduction in dopamine storage and release that is restricted to the
ventral striatum (Hnasko et al., 2010). Together, these findings
strongly suggest that within the striatum, VGLUT2 expression
and glutamate corelease is restricted to mesolimbic projections,
which may also extend to other limbic structures (Lavin et al.,
2005).
In conclusion, our findings provide unequivocal evidence that
a subset of dopaminergic terminals release glutamate capable of
postsynaptic signaling. Mice lacking VGLUT2 in dopamine neu-
rons show deficits in psychostimulant-induced locomotion
(Hnasko et al., 2010; Walle´n-Mackenzie et al., 2010), apparently
due to a presynaptic role for glutamate in vesicular monoamine
filling (Amilhon et al., 2010; Hnasko et al., 2010). Although we
still understand little about the role for glutamate released by
monoamine neurons as an independent signal, rapid, synaptic
signaling mediated by glutamate may provide the temporal spec-
ificity required for reward prediction and/or incentive salience
(Lapish et al., 2006, 2007).
References
Amilhon B, Lepicard E, Renoir T, Mongeau R, Popa D, Poirel O, Miot S, Gras
C, Gardier AM, Gallego J, Hamon M, Lanfumey L, Gasnier B, Giros B, El
Mestikawy S (2010) VGLUT3 (vesicular glutamate transporter type 3)
contribution to the regulation of serotonergic transmission and anxiety.
J Neurosci 30:2198–2210.
Be´ rube´-Carrie` re N, Riad M, Dal Bo G, Le´vesque D, Trudeau LE, Descarries L
(2009) The dual dopamine-glutamate phenotype of growing mesence-
phalic neurons regresses in mature rat brain. J Comp Neurol 517:
873–891.
Chuhma N, Zhang H, Masson J, Zhuang X, Sulzer D, Hen R, Rayport S
(2004) Dopamine neurons mediate a fast excitatory signal via their glu-
tamatergic synapses. J Neurosci 24:972–981.
Chuhma N, Choi WY, Mingote S, Rayport S (2009) Dopamine neuron glu-
tamate cotransmission: frequency-dependent modulation in the
mesoventromedial projection. Neuroscience 164:1068–1083.
Dal Bo G, St-Gelais F, Danik M, Williams S, Cotton M, Trudeau LE (2004)
Dopamine neurons in culture express VGLUT2 explaining their capacity
to release glutamate at synapses in addition to dopamine. J Neurochem
88:1398–1405.
Descarries L, Be´rube´-Carrie`re N, Riad M, Bo GD, Mendez JA, Trudeau LE
(2008) Glutamate in dopamine neurons: synaptic versus diffuse trans-
mission. Brain Res Rev 58:290–302.
Dobi A, Margolis EB, Wang HL, Harvey BK, Morales M (2010) Glutama-
tergic and nonglutamatergic neurons of the ventral tegmental area estab-
lish local synaptic contacts with dopaminergic and nondopaminergic
neurons. J Neurosci 30:218–229.
Hnasko TS, Chuhma N, Zhang H, Goh GY, Sulzer D, Palmiter RD, Rayport S,
Edwards RH (2010) Vesicular glutamate transport promotes dopamine
storage and glutamate corelease in vivo. Neuron 65:643–656.
Joyce MP, Rayport S (2000) Mesoaccumbens dopamine neuron synapses
reconstructed in vitro are glutamatergic. Neuroscience 99:445–456.
Kawano M, Kawasaki A, Sakata-Haga H, Fukui Y, Kawano H, Nogami H,
Hisano S (2006) Particular subpopulations of midbrain and hypotha-
lamic dopamine neurons express vesicular glutamate transporter 2 in the
rat brain. J Comp Neurol 498:581–592.
Lapish CC, Seamans JK, Chandler LJ (2006) Glutamate-dopamine cotrans-
mission and reward processing in addiction. Alcohol Clin Exp Res
30:1451–1465.
Lapish CC, Kroener S, Durstewitz D, Lavin A, Seamans JK (2007) The abil-
ity of the mesocortical dopamine system to operate in distinct temporal
modes. Psychopharmacology (Berl) 191:609–625.
Lavin A, Nogueira L, Lapish CC, Wightman RM, Phillips PE, Seamans JK
(2005) Mesocortical dopamine neurons operate in distinct temporal do-
mains using multimodal signaling. J Neurosci 25:5013–5023.
Mendez JA, Bourque MJ, Dal Bo G, Bourdeau ML, Danik M, Williams S,
Lacaille JC, Trudeau LE (2008) Developmental and target-dependent
regulation of vesicular glutamate transporter expression by dopamine
neurons. J Neurosci 28:6309–6318.
Nair-Roberts RG, Chatelain-Badie SD, Benson E, White-Cooper H, Bolam
JP, Ungless MA (2008) Stereological estimates of dopaminergic,
GABAergic and glutamatergic neurons in the ventral tegmental area, sub-
stantia nigra and retrorubral field in the rat. Neuroscience 152:1024
1031.
Reimer RJ, Edwards RH (2004) Organic anion transport is the primary
function of the SLC17/type I phosphate transporter family. Pflugers Arch
447:629–635.
Stuber GD, Klanker M, de Ridder B, Bowers MS, Joosten RN, Feenstra MG,
Bonci A (2008) Reward-predictive cues enhance excitatory synaptic
strength onto midbrain dopamine neurons. Science 321:1690–1692.
Sulzer D, Joyce MP, Lin L, Geldwert D, Haber SN, Hattori T, Rayport S
(1998) Dopamine neurons make glutamatergic synapses in vitro. J Neu-
rosci 18:4588–4602.
Takamori S (2006) VGLUTs: ‘exciting’ times for glutamatergic research?
Neurosci Res 55:343–351.
Tsai HC, Zhang F, Adamantidis A, Stuber GD, Bonci A, de Lecea L, Deisseroth
K (2009) Phasic firing in dopaminergic neurons is sufficient for behav-
ioral conditioning. Science 324:1080–1084.
Walle´n-Mackenzie A, Wootz H, Englund H (2010) Genetic inactivation of
the vesicular glutamate transporter 2 (VGLUT2) in the mouse: what have
we learnt about functional glutamatergic neurotransmission? Ups J Med
Sci 115:11–20.
Wise RA (2004) Dopamine, learning and motivation. Nat Rev Neurosci
5:483–494.
Yamaguchi T, Sheen W, Morales M (2007) Glutamatergic neurons are
present in the rat ventral tegmental area. Eur J Neurosci 25:106–118.
Zhuang X, Masson J, Gingrich JA, Rayport S, Hen R (2005) Targeted gene
expression in dopamine and serotonin neurons of the mouse brain.
J Neurosci Methods 143:27–32.
Stuber et al. Optogenetic Glutamate Corelease from Dopamine Terminals J. Neurosci., June 16, 2010 30(24):8229 – 8233 • 8233
... Because DA axons co-release glutamate, as well as GABA, 72,73,103,104 it was important to test possible involvement of co-released glutamate. We found that greater PTX-induced enhancement of 10 p-versus 1 p-evoked [DA] o persisted in the presence of a group 1 mGluR antagonist in the dStr and in ionotropic AMPAR and NMDAR antagonists in the NAc, supporting the independence of DA-release regulation by indirect consequences from co-released glutamate acting via local microcircuits. ...
Article
Full-text available
Striatal dopamine axons co-release dopamine and gamma-aminobutyric acid (GABA), using GABA provided by uptake via GABA transporter-1 (GAT1). Functions of GABA co-release are poorly understood. We asked whether co-released GABA autoinhibits dopamine release via axonal GABA type A receptors (GABAARs), complementing established inhibition by dopamine acting at axonal D2 autoreceptors. We show that dopamine axons express α3-GABAAR subunits in mouse striatum. Enhanced dopamine release evoked by single-pulse optical stimulation in striatal slices with GABAAR antagonism confirms that an endogenous GABA tone limits dopamine release. Strikingly, an additional inhibitory component is seen when multiple pulses are used to mimic phasic axonal activity, revealing the role of GABAAR-mediated autoinhibition of dopamine release. This autoregulation is lost in conditional GAT1-knockout mice lacking GABA co-release. Given the faster kinetics of ionotropic GABAARs than G-protein-coupled D2 autoreceptors, our data reveal a mechanism whereby co-released GABA acts as a first responder to dampen phasic-to-tonic dopamine signaling.
... While dopamine (DA) neurons have received the most attention, VTA and SNc neurons that release GABA (γaminobutyric acid) or glutamate can mediate related functions and have been implicated in addiction and Parkinson's disease (Barcomb & Ford, 2023;Buck et al., 2022;Morales & Margolis, 2017). Also present are neurons that co-express multiple neurotransmitter markers and that can release multiple neurotransmitters, supported by electrophysiological evidence demonstrating co-release of DA and glutamate, DA and GABA, or GABA and glutamate (Chuhma et al., 2004;Hnasko et al., 2010Root et al., 2014;Stuber et al., 2010;Tecuapetla et al., 2010;Tritsch et al., 2012;Yoo et al., 2016). ...
Preprint
Full-text available
Most studies on the ventral tegmental area (VTA) and substantia nigra pars compacta (SNc) have focused on dopamine neurons and their role in processes such as motivation, learning, movement, and associated disorders. However there has been increasing attention on other VTA and SNc cell types that release GABA, glutamate, or a combination of these neurotransmitters. Yet the relative distributions and proportions of neurotransmitter-defined cell types across VTA and SNc has remained unclear. Here, we used fluorescent in situ hybridization in male and female mice to label VTA and SNc neurons that expressed mRNA encoding the canonical vesicular transporters for dopamine, GABA, or glutamate: vesicular monoamine transporter VMAT2, vesicular GABA transporter (VGAT), and vesicular glutamate transporter (VGLUT2). Within VTA, we found that no one type was particularly more abundant, instead we observed similar numbers of VMAT2+ (44%), VGAT+ (37%) and VGLUT2+ (41%) neurons. In SNc we found that a slight majority of neurons expressed VMAT2 (54%), fewer were VGAT+ (42%), and VGLUT2+ neurons were least abundant (16%). Moreover, 20% of VTA neurons and 10% of SNc neurons expressed more than one vesicular transporter, including 45% of VGLUT2 neurons. We also assessed within VTA and SNc subregions and found remarkable heterogeneity in cell-type composition. And by quantifying density across both anterior-posterior and medial-lateral axes we generated heatmaps to visualize the distribution of each cell type. Our data complement recent single-cell RNAseq studies and support a more diverse landscape of neurotransmitter-defined cell types in VTA and SNc than is typically appreciated.
... ; https://doi.org/10.1101/2024.02.24.581852 doi: bioRxiv preprint contribution by such processes by reporting that glutamate release from glutamate/dopamine co-releasing VTA NAcS neurons promotes reinforcement in the form of optogenetic selfstimulation (Zell et al., 2020). In addition, DA terminals co-release glutamate preferentially in the ventromedial, but not the dorsal striatum (Mingote et al., 2019;Stuber et al., 2010;Tecuapetla et al., 2010). It is possible that the observed effects after manipulating this region were due to additional gluatamate modulation -perhaps explaining the lack of effect when A further consideration is the potential reinforcing effect of light applied via optogenetics, which could potentially exert rewarding effects. ...
Preprint
Full-text available
Cognitive flexibility, the capacity to adapt behaviour to changes in the environment, is impaired in a range of brain disorders, including substance use disorder and Parkinson’s disease. Putative neural substrates of cognitive flexibility include mesencephalic pathways to the ventral striatum (VS) and dorsomedial striatum (DMS), hypothesised to encode learning signals needed to maximize rewarded outcomes during decision-making. However, it is unclear whether mesencephalic projections to the ventral and dorsal striatum are distinct in their contribution to flexible reward-related learning. Here, rats acquired a two-choice spatial probabilistic reversal learning (PRL) task, reinforced on an 80%:20% basis, that assessed the flexibility of behaviour to repeated reversals of response-outcome contingencies. We report that optogenetic stimulation of projections from the ventral tegmental area (VTA) to the nucleus accumbens shell (NAcbS) in the VS significantly impaired reversal learning when optical stimulation was temporally aligned with negative feedback (i.e., reward omission). Moreover, the exploitation-exploration parameter, β , was increased (indicating greater exploitation of information) when this pathway was optogenetically stimulated after a spurious loss (i.e. an incorrect (20%) response at the 80% reinforrced location) compared to after a spurious win (i.e. a correct (20%) response at the 20% reinforced location). VTA → NAcbS stimulation during other phases of the behavioural task was without effect. Optogenetic stimulation of projection neurons from the substantia nigra (SN) to the DMS, aligned either with reward receipt or omission or prior to making a choice, had no effect on reversal learning. These findings are consistent with the notion that enhanced activity in VTA → NAcbS projections leads to maladaptive perseveration as a consequence of an inappropriate bias to exploitation via positive reinforcement.
... However, the specific role of the VTA-PFC glutamatergic pathway in pain modulation remains unclear. Adding to the complexity, it has been reported that some VTA glutamatergic neurons can co-release dopamine 14,15 , although this phenomenon appears to be specific to certain projections 16 . For instance, within the mesocortical pathway, approximately two-thirds of VTA neurons projecting to the PFC express the type-2 vesicular glutamate transporter (vGluT2), and about 40% of these neurons co-express tyrosine hydroxylase (TH), an enzyme involved in dopamine biosynthesis 15 . ...
Article
Full-text available
Dysfunction in the mesocortical pathway, connecting the ventral tegmental area (VTA) to the prefrontal cortex, has been implicated in chronic pain. While extensive research has focused on the role of dopamine, the contribution of glutamatergic signaling in pain modulation remains unknown. Using in vivo calcium imaging, we observe diminished VTA glutamatergic activity targeting the prelimbic cortex (PL) in a mouse model of neuropathic pain. Optogenetic activation of VTA glutamatergic terminals in the PL alleviates neuropathic pain, whereas inhibiting these terminals in naïve mice induces pain-like responses. Importantly, this pain-modulating effect is independent of dopamine co-release, as demonstrated by CRISPR/Cas9-mediated gene deletion. Furthermore, we show that VTA neurons primarily project to excitatory neurons in the PL, and their activation restores PL outputs to the anterior cingulate cortex, a key region involved in pain processing. These findings reveal a distinct mesocortical glutamatergic pathway that critically modulates neuropathic pain independent of dopamine signaling.
... Glutamate neurons in the VTA mainly express VgluT2 but not VgluT1 or VgluT3 [73,74]. VgluT2-expressing glutamate neurons are mostly located in the anterior and middle line of the VTA [44,73] and project to the NAc, ventral pallidum (VP), PFC, dorsal hippocampus (DH), and lateral habenula (LHb) [13,33,71,74,75] (Figure 3). Optical stimulation of VTA glutamate neurons is rewarding, as assessed by the increased firing of VTA DA neurons, supporting oICSS, and producing conditioned place preference and appetitive instrumental conditioning [13,32,33]. ...
Preprint
Full-text available
Brain-stimulation reward, also known as intracranial self-stimulation (ICSS), is a commonly used procedure for studying brain reward function and drug reward. In electrical ICSS, an electrode is surgically implanted into the medial forebrain bundle in the lateral hypothalamus or the ventral tegmental area in the midbrain. Operant lever responding leads to the delivery of electrical pulse stimulation. The alteration in the stimulation frequency-lever response curve is used to evaluate the impact of pharmacological agents on brain reward function. If a test drug induces a leftward or upward shift in the ICSS response curve, it implies a reward-enhancing or abuse-like effect. Conversely, if a drug causes a rightward or downward shift in the functional response curve, it suggests a reward-attenuating or aversive effect. A significant drawback of electrical ICSS is the lack of cellular selectivity in understanding the neural substrates underlying drug reward versus aversion. Excitingly, recent advancements in optical ICSS (oICSS) have facilitated the development of at least three cell type-specific oICSS models – dopamine-, glutamate-, and GABA-dependent oICSS. In these new models, a comparable frequency-rate response curve has been established and employed to study the substrate-specific mechanisms underlying brain reward function and a drug's rewarding versus aversive effects. In this review article, we summarize recent progress in this exciting research area. The findings in these new behavioral models have not only increased our understanding of the neural mechanisms underlying drug reward and addiction but have also introduced a novel experimental approach in preclinical medication development for treating substance use disorders.
Article
Full-text available
Brain-stimulation reward, also known as intracranial self-stimulation (ICSS), is a commonly used procedure for studying brain reward function and drug reward. In electrical ICSS (eICSS), an electrode is surgically implanted into the medial forebrain bundle (MFB) in the lateral hypothalamus or the ventral tegmental area (VTA) in the midbrain. Operant lever responding leads to the delivery of electrical pulse stimulation. The alteration in the stimulation frequency-lever response curve is used to evaluate the impact of pharmacological agents on brain reward function. If a test drug induces a leftward or upward shift in the eICSS response curve, it implies a reward-enhancing or abuse-like effect. Conversely, if a drug causes a rightward or downward shift in the functional response curve, it suggests a reward-attenuating or aversive effect. A significant drawback of eICSS is the lack of cellular selectivity in understanding the neural substrates underlying this behavior. Excitingly, recent advancements in optical ICSS (oICSS) have facilitated the development of at least three cell type-specific oICSS models—dopamine-, glutamate-, and GABA-dependent oICSS. In these new models, a comparable stimulation frequency-lever response curve has been established and employed to study the substrate-specific mechanisms underlying brain reward function and a drug’s rewarding versus aversive effects. In this review article, we summarize recent progress in this exciting research area. The findings in oICSS have not only increased our understanding of the neural mechanisms underlying drug reward and addiction but have also introduced a novel behavioral model in preclinical medication development for treating substance use disorders.
Article
The role of nitrergic system in modulating the action of psychostimulants on reward processing is well established. However, the relevant anatomical underpinnings and scope of the involved interactions with mesolimbic dopaminergic system have not been clarified. Using immunohistochemistry, we track the changes in neuronal nitric oxide synthase (nNOS) containing cell groups in the animals conditioned to intracranial self‐stimulation (ICSS) via an electrode implanted in the lateral hypothalamus‐medial forebrain bundle (LH‐MFB) area. An increase in the nNOS immunoreactivity was noticed in the cells and fibers in the ventral tegmental area (VTA) and nucleus accumbens shell (AcbSh), the primary loci of the reward system. In addition, nNOS was up‐regulated in the nucleus accumbens core (AcbC), vertical limb of diagonal band (VDB), locus coeruleus (LC), lateral hypothalamus (LH), superficial gray layer (SuG) of the superior colliculus, and periaqueductal gray (PAG). The brain tissue fragments drawn from these areas showed a change in nNOS mRNA expression, but in opposite direction. Intracerebroventricular (icv) administration of nNOS inhibitor, 7‐nitroindazole (7‐NI) showed decreased lever press activity in a dose‐dependent manner in ICSS task. While an increase in the dopamine (DA) and 3, 4‐dihydroxyphenylacetic acid (DOPAC) efflux was noted in the microdialysates collected from the AcbSh of ICSS rats, pre‐administration of 7‐NI (icv route) attenuated the response. The study identifies nitrergic centers that probably mediate sensory, cognitive, and motor components of the goal‐directed behavior.
Preprint
Sniffing is a motivated behavior displayed by all terrestrial vertebrates on the planet. While sniffing is associated with acquiring and processing odors, sniffing is also intertwined with affective and motivated states. The neuromodulatory systems which influence the display of sniffing are unclear. Here, we report that dopamine release into the ventral striatum, with exception of the nucleus accumbens core, is coupled with bouts of sniffing and that stimulation of dopaminergic terminals in those regions initiates sniffing. The activity of post-synaptic D1 and D2 receptor-expressing neurons in the ventral striatum is also coupled with sniffing and antagonism of ventral striatum D1 and D2 receptors squelches sniffing behavior. Together, these results support a model whereby sniffing is initiated by dopamine’s actions upon ventral striatum neurons. The nature of sniffing being integral to both olfaction and motivated behaviors implicates this circuit in a wide array of functions. One-Sentence Summary Mesolimbic dopamine input to the ventral striatum supports the highly conserved behavior of sniffing by promoting both the initiation and vigor of sniffing bouts.
Article
For many years, neuroscientists have investigated the behavioural, computational and neurobiological mechanisms that support value-based decisions, revealing how humans and animals make choices to obtain rewards. However, many decisions are influenced by factors other than the value of physical rewards or second-order reinforcers (such as money). For instance, animals (including humans) frequently explore novel objects that have no intrinsic value solely because they are novel and they exhibit the desire to gain information to reduce their uncertainties about the future, even if this information cannot lead to reward or assist them in accomplishing upcoming tasks. In this Review, I discuss how circuits in the primate brain responsible for detecting, predicting and assessing novelty and uncertainty regulate behaviour and give rise to these behavioural components of curiosity. I also briefly discuss how curiosity-related behaviours arise during postnatal development and point out some important reasons for the persistence of curiosity across generations.
Article
Full-text available
During the past decade, three proteins that possess the capability of packaging glutamate into presynaptic vesicles have been identified and characterized. These three vesicular glutamate transporters, VGLUT1-3, are encoded by solute carrier genes Slc17a6-8. VGLUT1 (Slc17a7) and VGLUT2 (Slc17a6) are expressed in glutamatergic neurons, while VGLUT3 (Slc17a8) is expressed in neurons classically defined by their use of another transmitter, such as acetylcholine and serotonin. As glutamate is both a ubiquitous amino acid and the most abundant neurotransmitter in the adult central nervous system, the discovery of the VGLUTs made it possible for the first time to identify and specifically target glutamatergic neurons. By molecular cloning techniques, different VGLUT isoforms have been genetically targeted in mice, creating models with alterations in their glutamatergic signalling. Glutamate signalling is essential for life, and its excitatory function is involved in almost every neuronal circuit. The importance of glutamatergic signalling was very obvious when studying full knockout models of both VGLUT1 and VGLUT2, none of which were compatible with normal life. While VGLUT1 full knockout mice die after weaning, VGLUT2 full knockout mice die immediately after birth. Many neurological diseases have been associated with altered glutamatergic signalling in different brain regions, which is why conditional knockout mice with abolished VGLUT-mediated signalling only in specific circuits may prove helpful in understanding molecular mechanisms behind such pathologies. We review the recent studies in which mouse genetics have been used to characterize the functional role of VGLUT2 in the central nervous system.
Article
Full-text available
Three different subtypes of H(+)-dependent carriers (named VGLUT1-3) concentrate glutamate into synaptic vesicles before its exocytotic release. Neurons using other neurotransmitter than glutamate (such as cholinergic striatal interneurons and 5-HT neurons) express VGLUT3. It was recently reported that VGLUT3 increases acetylcholine vesicular filling, thereby, stimulating cholinergic transmission. This new regulatory mechanism is herein designated as vesicular-filling synergy (or vesicular synergy). In the present report, we found that deletion of VGLUT3 increased several anxiety-related behaviors in adult and in newborn mice as early as 8 d after birth. This precocious involvement of a vesicular glutamate transporter in anxiety led us to examine the underlying functional implications of VGLUT3 in 5-HT neurons. On one hand, VGLUT3 deletion caused a significant decrease of 5-HT(1A)-mediated neurotransmission in raphe nuclei. On the other hand, VGLUT3 positively modulated 5-HT transmission of a specific subset of 5-HT terminals from the hippocampus and the cerebral cortex. VGLUT3- and VMAT2-positive serotonergic fibers show little or no 5-HT reuptake transporter. These results unravel the existence of a novel subset of 5-HT terminals in limbic areas that might play a crucial role in anxiety-like behaviors. In summary, VGLUT3 accelerates 5-HT transmission at the level of specific 5-HT terminals and can exert an inhibitory control at the raphe level. Furthermore, our results suggest that the loss of VGLUT3 expression leads to anxiety-associated behaviors and should be considered as a potential new target for the treatment of this disorder.
Article
Full-text available
The ventral tegmental area (VTA) contributes to reward and motivation signaling. In addition to the well established populations of dopamine (DA) or GABA VTA neurons, glutamatergic neurons were recently discovered in the VTA. These glutamatergic neurons express the vesicular glutamate transporter 2, VGluT2. To investigate whether VTA glutamatergic neurons establish local synapses, we tagged axon terminals from resident VTA neurons by intra-VTA injection of Phaseolus vulgaris leucoagglutinin (PHA-L) or an adeno-associated virus encoding wheat germ agglutinin (WGA) and by immunoelectron microscopy determined the presence of VGluT2 in PHA-L- or WGA-positive terminals. We found that PHA-L- or WGA-positive terminals from tagged VTA cells made asymmetric or symmetric synapses within the VTA. VGluT2 immunoreactivity was detected in the vast majority of PHA-L- or WGA-positive terminals forming asymmetric synapses. These results indicate that both VTA glutamatergic and nonglutamatergic (likely GABAergic) neurons establish local synapses. To examine the possible DAergic nature of postsynaptic targets of VTA glutamatergic neurons, we did triple immunolabeling with antibodies against VGluT2, tyrosine hydroxylase (TH), and PHA-L. From triple-labeled tissue, we found that double-labeled PHA-L (+)/VGluT2 (+) axon terminals formed synaptic contacts on dendrites of both TH-positive and TH-negative cells. Consistent with these anatomical observations, in whole-cell slice recordings of VTA neurons we observed that blocking action potential activity significantly decreased the frequency of synaptic glutamatergic events in DAergic and non-DAergic neurons. These observations indicate that resident VTA glutamatergic neurons are likely to affect both DAergic and non-DAergic neurotransmission arising from the VTA.
Article
Full-text available
Rewarding Bursts of Dopamine Dopaminergic neurons are thought to be involved in the cognitive and hedonic underpinnings of motivated behaviors. However, it is still unclear whether dopaminergic neuron activation is sufficient to elicit reward-related behavior and which type of neuronal activity pattern serves this purpose. Tsai et al. (p. 1080; published online 23 April) directly compared tonic versus phasic firing of dopaminergic cells in the ventral tegmental area, and the effects on both behavior and dopamine release. Using a transgenic system and virus injection in mice, they targeted the dopaminergic cells with rhodopsin. Light stimulation was then used to drive dopaminergic cells either with a tonic low level of pulses or bursts of high-frequency pulses, with the number of pulses being equal across conditions. Only the high-frequency phasic firing induced a conditioned place preference and dopamine release.
Article
Full-text available
Using sensory information for the prediction of future events is essential for survival. Midbrain dopamine neurons are activated by environmental cues that predict rewards, but the cellular mechanisms that underlie this phenomenon remain elusive. We used in vivo voltammetry and in vitro patch-clamp electrophysiology to show that both dopamine release to reward predictive cues and enhanced synaptic strength onto dopamine neurons develop over the course of cue-reward learning. Increased synaptic strength was not observed after stable behavioral responding. Thus, enhanced synaptic strength onto dopamine neurons may act to facilitate the transformation of neutral environmental stimuli to salient reward-predictive cues.
Article
Full-text available
Interactions between dopamine and glutamate play prominent roles in memory, addiction, and schizophrenia. Several lines of evidence have suggested that the ventral midbrain dopamine neurons that give rise to the major CNS dopaminergic projections may also be glutamatergic. To examine this possibility, we double immunostained ventral midbrain sections from rat and monkey for the dopamine-synthetic enzyme tyrosine hydroxylase and for glutamate; we found that most dopamine neurons immunostained for glutamate, both in rat and monkey. We then used postnatal cell culture to examine individual dopamine neurons. Again, most dopamine neurons immunostained for glutamate; they were also immunoreactive for phosphate-activated glutaminase, the major source of neurotransmitter glutamate. Inhibition of glutaminase reduced glutamate staining. In single-cell microculture, dopamine neurons gave rise to varicosities immunoreactive for both tyrosine hydroxylase and glutamate and others immunoreactive mainly for glutamate, which were found near the cell body. At the ultrastructural level, dopamine neurons formed occasional dopaminergic varicosities with symmetric synaptic specializations, but they more commonly formed nondopaminergic varicosities with asymmetric synaptic specializations. Stimulation of individual dopamine neurons evoked a fast glutamatergic autaptic EPSC that showed presynaptic inhibition caused by concomitant dopamine release. Thus, dopamine neurons may exert rapid synaptic actions via their glutamatergic synapses and slower modulatory actions via their dopaminergic synapses. Together with evidence for glutamate cotransmission in serotonergic raphe neurons and noradrenergic locus coeruleus neurons, the present results suggest that glutamatergic cotransmission may be the rule for central monoaminergic neurons.
Article
There is solid electron microscopic data demonstrating the existence of dopamine (DA) axon terminals (varicosities) with or without synaptic membrane specializations (junctional complexes) in many parts of the CNS, and notably in neostriatum and nucleus accumbens. The dual morphological character of these DA innervations has led to the suggestion that the meso-telencephalic DA system operates by diffuse (or volume) as well as by classical synaptic transmission. In the last decade, electrophysiological and neurochemical evidence has also accumulated indicating that monoamine neurons in various parts of the CNS, and particularly the mesencephalic DA neurons, might release glutamate as a co-transmitter. Following the identification of the vesicular transporters for glutamate (VGluT), in situ hybridization and RT-PCR studies carried out on isolated neurons or standard tissue cultures, and more recently in vivo, have shown that VGluT2 mRNA may be expressed in a significant proportion of mesencephalic DA neurons, at least in the ventral tegmental area. A current study also suggests that the co-expression of tyrosine hydroxylase (TH) and VGluT2 by these neurons is regulated during embryonic development, and may be derepressed or reactivated postnatally following their partial destruction by neonatal administration of 6-hydroxydopamine (6-OHDA). In both 15 day-old and adult rats subjected or not to the neonatal 6-OHDA lesion, concurrent electron microscopic examination of the nucleus accumbens after dual immunocytochemical labeling for TH and VGluT2 reveals the co-existence of the two proteins in a significant proportion of these axon terminals. Moreover, all TH varicosities which co-localize VGluT2 are synaptic, as if there was a link between the potential of DA axon terminals to release glutamate and their establishment of synaptic junctions. Together with the RT-PCR and in situ hybridization data demonstrating the co-localization of TH and VGluT2 mRNA in mesencephalic neurons of the VTA, these observations raise a number of fundamental questions regarding the functioning of the meso-telencephalic DA system in healthy or diseased brain.
Article
Dopamine neurons in the ventral tegmental area (VTA) play an important role in the motivational systems underlying drug addiction, and recent work has suggested that they also release the excitatory neurotransmitter glutamate. To assess a physiological role for glutamate corelease, we disrupted the expression of vesicular glutamate transporter 2 selectively in dopamine neurons. The conditional knockout abolishes glutamate release from midbrain dopamine neurons in culture and severely reduces their excitatory synaptic output in mesoaccumbens slices. Baseline motor behavior is not affected, but stimulation of locomotor activity by cocaine is impaired, apparently through a selective reduction of dopamine stores in the projection of VTA neurons to ventral striatum. Glutamate co-entry promotes monoamine storage by increasing the pH gradient that drives vesicular monoamine transport. Remarkably, low concentrations of glutamate acidify synaptic vesicles more slowly but to a greater extent than equimolar Cl(-), indicating a distinct, presynaptic mechanism to regulate quantal size.
Article
Coexpression of tyrosine hydroxylase (TH) and vesicular glutamate transporter 2 (VGLUT2) mRNAs in the ventral tegmental area (VTA) and colocalization of these proteins in axon terminals of the nucleus accumbens (nAcb) have recently been demonstrated in immature (15-day-old) rat. After neonatal 6-hydroxydopamine (6-OHDA) lesion, the proportion of VTA neurons expressing both mRNAs and of nAcb terminals displaying the two proteins was enhanced. To determine the fate of this dual phenotype in adults, double in situ hybridization and dual immunolabeling for TH and VGLUT2 were performed in 90-day-old rats subjected or not to the neonatal 6-OHDA lesion. Very few neurons expressed both mRNAs in the VTA and substantia nigra (SN) of P90 rats, even after neonatal 6-OHDA. Dually immunolabeled terminals were no longer found in the nAcb of normal P90 rats and were exceedingly rare in the nAcb of 6-OHDA-lesioned rats, although they had represented 28% and 37% of all TH terminals at P15. Similarly, 17% of all TH terminals in normal neostriatum and 46% in the dopamine neoinnervation of SN in 6-OHDA-lesioned rats were also immunoreactive for VGLUT2 at P15, but none at P90. In these three regions, all dually labeled terminals made synapse, in contradistinction to those immunolabeled for only TH or VGLUT2 at P15. These results suggest a regression of the VGLUT2 phenotype of dopamine neurons with age, following normal development, lesion, or sprouting after injury, and a role for glutamate in the establishment of synapses by these neurons.
Article
Mesoventromedial dopamine neurons projecting from the medial ventral tegmental area to the ventromedial shell of the nucleus accumbens play a role in attributing incentive salience to environmental stimuli that predict important events, and appear to be particularly sensitive to the effects of psychostimulant drugs. Despite the observation that these dopamine neurons make up almost the entire complement of neurons in the projection, stimulating their cell bodies evokes a fast glutamatergic response in accumbens neurons. This is apparently due to dopamine neuron glutamate cotransmission, suggested by the extensive coexpression of vesicular glutamate transporter 2 (VGLUT2) in the neurons. To examine the interplay between the dopamine and glutamate signals, we used acute quasi-horizontal brain slices made from DAT-YFP mice in which the intact mesoventromedial projection can be visualized. Under current clamp, when dopamine neurons were stimulated repeatedly, dopamine neuron glutamate transmission showed dopamine-mediated facilitation, solely at higher, burst-firing frequencies. Facilitation was diminished under voltage clamp and flipped to inhibition by intracellular Cs(+) or GDPbetaS, indicating that it was mediated postsynaptically. Postsynaptic facilitation was D1 mediated, required activation of NMDA receptors and closure of voltage gated K(+)-channels. When postsynaptic facilitation was blocked, D2-mediated presynaptic inhibition became apparent. These counterbalanced pre- and postsynaptic actions determine the frequency dependence of dopamine modulation; at lower firing frequencies dopamine modulation is not apparent, while at burst firing frequency postsynaptic facilitation dominates and dopamine becomes facilitatory. Dopamine neuron glutamate cotransmission may play an important role in encoding the incentive salience value of conditioned stimuli that activate goal-directed behaviors, and may be an important subtract for enduring drug-seeking behaviors.