Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Mar 1.
Published in final edited form as: Basal Ganglia. 2012 Mar 1;2(1):5–13. doi: 10.1016/j.baga.2011.11.004

Regulation of striatal dopamine release by presynaptic auto- and heteroreceptors

Hui Zhang 1, David Sulzer 1
PMCID: PMC3375990  NIHMSID: NIHMS342997  PMID: 22712055

Summary

Striatal dopamine neurotransmission is critical for normal voluntary movement, affect and cognition. Dysfunctions of its regulation are implicated in a broad range of behaviors and disorders including Parkinson’s disease, schizophrenia and drug abuse. Extracellular dopamine levels result from a dynamic equilibrium between release and reuptake by dopaminergic terminals. Both processes are regulated by multiple mechanisms. Here we review data characterizing how dopamine levels are regulated by presynaptic autoreceptors and heteroreceptors, an area intensively investigated due to advances in real time electrochemical detection of extracellular dopamine, i.e., fast-scan cyclic voltammetry and amperometry, and the development of mutant mouse lines with deletions for specific receptors.

Keywords: dopamine, release, autoreceptor, heteroreceptor, cyclic voltammetry, striatum

1. Introduction

The neurotransmitter dopamine is important for movement, motivation, and cognition. Acting as a neuromodulator of ionotropic synapses, dopamine sets a threshold for striatal activity that underlies motor learning, motivation, and decision making [1-3]. As might be presumed from these complex and important functions, there are multiple presynaptic mechanisms that regulate dopamine release, particularly due to receptor-mediated mechanisms driven by additional neural pathways present in these brain regions.

Dopaminergic neurotransmission is generally initiated by synaptic vesicle fusion, which can be modulated at different levels including dopamine synthesis, uptake and vesicular transport as well as Ca2+ -homeostasis and exocytotic proteins [4]. In addition, dopamine autoreceptors and heteroreceptors expressed on dopamine somatodendritic regions [5-9] and presynaptic axon terminals provide feedback and regulate dopamine release.

Advances in electrochemical detection of extracellular dopamine using carbon fiber electrodes in acute brain striatal brain slices [10, 11], along with an explosion of receptor deficient mutant (“knockout”) mouse lines [12, 13], has vastly contributed to the characterization of autoreceptor and heteroreceptor regulation of this system. While newer approaches, including the use of optical methods such as channelrhodopsins and fluorescent false neurotransmitters may lead to further insights in the future, we review advances from the past decade at the terminal level in the striatum since it is the major point of entry into the basal ganglia for information emanating from the cortex, and plays important roles in motor control, cognition, learning and addiction. In addition, the regulation of dopamine release in this region is best understood since it has the richest dopamine innervation in the central nervous system (CNS).

2. Regulation of release by autoreceptors

It has long been known that dopamine autoreceptors expressed on dopamine presynaptic axon terminals provide feedback and regulate dopamine release. There are multiple pathways or mechanisms at different time scales to achieve this. Activation of dopamine autoreceptors can decrease release by inhibiting dopamine synthesis and enhancing dopamine reuptake by the dopamine transporter as well as regulating vesicular monoamine transporter (VMAT2) expression [4, 14, 15]. These are slow processes that may last for minutes to hours. Here we specifically review newer data on quick effects on evoked dopamine release by modulating ion channel conductance.

2.1. Dopamine autoreceptors

Dopamine autoreceptors belong to the D2-family (D2, D3, D4) of dopamine receptors that are coupled to inhibitory G proteins and modulate ion channel activity and/or inhibit adenylyl cyclase. D2 receptors are expressed along the somatodendritic extent of midbrain dopamine neurons, as well as at their axon terminals in the dorsal striatum and nucleus accumbens (also called the ventral striatum) [16]. Of two isoforms of the D2 receptors generated by alternative splicing, the shorter version (D2-S) may be the dominant presynaptic autoreceptor [17-19].

D2 receptor knockout mice exhibited no detectable autoreceptor response to D2-class receptor agonists in firing rate, dopamine release or dopamine synthesis. Recently, fast-scan cyclic voltammetry (FCV) recordings in a mouse line that is deficient for D2 receptors selectively in dopamine neurons confirmed that it is the autoreceptor regulating dopamine neuron firing and release [20]. Thus, results implicated the D2 receptor as the only functional autoreceptor in the D2-family [20-24].

In contrast, a role for axonal D3 autoreceptors remains unproven. D3 immunoreactivity has been found in almost all midbrain dopamine neurons, but is undetectable in the terminal regions [25]. While it has been suggested that D2 and D3 receptors may both function as autoreceptors [26-29], no clear deficits have been identified in D3 receptor knockout mice, although extracellular dopamine was elevated in the ventral striatum [30] and one study of D3 KO mice using FCV in striatal slices has demonstrated a small D3 role for regulation of secretion but not synthesis in the striatum [31].

Studies on transfected cells demonstrated that D3 receptors can modulate the same G protein-activated inwardly rectifying potassium channels (GIRKs) as D2 receptors [32-34], and have the potential to affect dopamine release [35]. However, a study from D2 receptor knockout mice did not find evidence for D3 receptor activation of GIRK currents on genuine ventral midbrain dopamine neurons [36].

There is little evidence supporting a role of D4 receptors as autoreceptors beyond an immunohistochemistry study that demonstrated presynaptic D4 receptor localization in a subset of mesoaccumbal terminals in the nucleus accumbens shell [37]. In summary, dopamine autoreceptor function is predominantly carried out by D2 receptors.

2.2. In vitro studies

The regulation of dopamine release by dopaminergic autoreceptors was initially studied in vitro with neurochemical approaches [38]. D2 autoreceptor activation inhibits axon terminal [39-45, 46] and somatodendritic dopamine release [41].

The molecular mechanism underlying the inhibition of dopamine release through terminal D2 autoreceptor is not fully understood. One possibility is that D2 autoreceptors inhibit voltage-gated Ca2+ channels in axon terminals, thus directly inhibiting Ca2+-dependent exocytosis. Patch clamp studies on dopamine midbrain neurons in vitro have revealed a D2 receptor regulation of voltage-gated calcium currents [47] and axon terminal dopamine release is dependent on N and P/Q type calcium channels [11, 48]. A presynaptic effect of D2 on these channels has not been proven directly, however, and a study on autapses (i.e., synapses that a neuron makes on itself) of midbrain neurons in culture found no evidence for a D2 autoreceptor regulation of calcium influx [49] but that 4-aminopyridine-sensitive K+ channels acting downstream from calcium influx are involved. Autapses of cultured dopamine neurons are glutamatergic [50], however, and it is not known whether this finding can be extrapolated to dopamine release. Nevertheless, the broad-spectrum K+ channel blockers 4-aminopyridine and tetraethylammonium reduce the ability of a D2 agonist, quinpirole, to inhibit evoked dopamine release in slices [51], supporting a role for presynaptic K+ channels. Using Kv subtype-selective blockers, it was recently shown that Kv1.2, Kv1.3 and Kv1.6 are located in dopamine axonal processes in the dorsal striatum. These channels are likely to be key players in regulating axonal dopamine release and may act as mediators or regulators of presynaptic D2 receptor function [52, 53]. Nevertheless, given that the non-linear nature of the threshold for action potential initiation and consequent dopamine release, blocking any class of K+ channels could alter the efficacy of D2 agonists to inhibit release. Thus, such data are not necessarily mechanistically informative. Further studies are required to evaluate whether the D2 receptor can directly regulate the function of Kv1 channel subtypes in DA neurons. Furthermore, as Kv1 potassium channel blockers can only partially block the inhibition effect of quinpirole, it is possible that other classes of potassium channels are also involved. GIRK2 potassium channels are expressed in dopamine neurons [36, 54] and KATP potassium channels are expressed in dopamine terminals [55], however, they appear to only play minor roles in regulating presynaptic dopamine release [53]. Therefore, other additional mechanisms need to be explored.

2.3. In vivo studies

Consistent with the in vitro studies, the regulation of dopamine release by dopamine autoreceptors was subsequently shown in vivo using microdialysis. Indeed, systemic administration of D2 antagonists enhances the extracellular dopamine level in the dorsal striatum [56]. Moreover, intrastriatal infusion of D2 agonists or antagonists decreases or enhances extracellular dopamine, respectively [57]. This autoregulation acts on the impulse flow-dependent dopamine release [56] and is not mediated by an indirect action on striatal neurons [58]. However, the extracellular dopamine level results from a dynamic equilibrium between dopamine release and dopamine reuptake and microdialysis does not clearly distinguish between changes in release and reuptake.

In contrast, in vivo electrochemical studies have shed light on the D2 autoregulation of dopamine release and of dopamine reuptake. These studies confirmed that the tonic level of extracellular dopamine concentration is sufficient to exert a tonic stimulation of D2 autoreceptors, which inhibits the impulse flow-dependent dopamine release [59, 60]. However, the electrochemical techniques used in these initial studies were too slow to describe the kinetics of the D2 autoregulation. The use of a faster technique (i.e., amperometry) and of control experiments with mice lacking D2 receptors made it possible to describe the dynamic characteristics of D2 autoregulation [21]. The onset of the D2 inhibition is between 50 ms and 100 ms, reaching a maximum between 150 and 300 ms after the end of the conditioning stimulation and disappearing by 800 ms [21]. Similar kinetics was described with in vitro experiments [24, 46] with a maximum effect around 500 ms and duration of less than 5 sec. The slightly longer time course determined in the in vitro studies is likely due to the larger amount of dopamine released per stimulation in slice preparations.

3. Regulation of dopamine release by heteroreceptors

In addition to D2 autoreceptors, there are additional neurotransmitter receptors on dopamine cell bodies and terminals, but their modulatory roles in dopamine release are less well characterized. Many addictive drugs, including nicotine, ethanol, opioids, morphine, and cannabinoids appear to enhance dopamine neuronal activity by acting directly on dopamine neurons or indirectly on GABA interneurons through the mechanism of disinhibition [9, 61, 62].

Electron microscopic demonstration of heteroreceptor immunolabel in dopamine terminals is invaluable for establishing their presence, but there are relatively few studies in this area due to technical challenges. To date, only the GDNF receptor [63], nicotinic receptors (nAChRs) [64], delta and kappa opioid receptors [65, 66], the metabotropic receptor mGluR1 [67], and possibly the GABAB receptor [68] have been observed by ultrastructural immunolabel in dopamine terminals.

Although the presence of other presynaptic receptors in dopamine terminals remains to be fully elucidated, studies using synaptosomes preparations, microdialysis, and CV recordings in vivo and in vitro suggest important roles for heteroreceptor modulation on dopamine release, albeit with conflicting results.

3.1. Glutamate receptors

The effects of ionotropic glutamate-receptors (iGluRs) on dopamine release are most likely indirect given that mature dopamine terminals appear to lack these receptors [69, 70], although axonal growth cones possess NMDA receptors [71]. While microdialysis studies indicated a stimulatory effect for iGluRs on dopamine release, FCV recordings demonstrated an inhibitory role on evoked dopamine release in the terminal region [72-74]. Direct perfusion of the acute striatal slices with iGluR agonists (NMAD, AMPA, and KA) causes a profound inhibition in evoked dopamine release in the first 4 minutes, but then a huge release of dopamine, similar to massive release that occurs with 70mM KCl [75, Zhang and Sulzer, unpublished results]. This is consistent with the suggestion that iGluR agonists may induce wide-spread depolarization and then spreading depression [76, 77].

Presynaptic regulation of dopamine release by iGluRs has been more clearly resolved in striatal slices using train stimulation using FCV [72, 78] since the effect of concurrently released glutamate during train stimulation on dopamine release could be revealed by corresponding antagonists. Recent studies by Margaret Rice’s group suggest that glutamatergic regulation of dopamine release is indeed inhibitory. The effect is indirect and mediated by AMPA receptors on striatal cells, which is in turn mediated through retrograde signaling by diffusible H2O2 generated in striatal cells, rather than in dopamine axons [78]. Activity–dependent H2O2 generated in striatal cells inhibits dopamine release via ATP-sensitive K+ channels located on dopamine terminals [55].

In contrast to iGluRs, the metabotropic receptor mGluR1 has been detected by ultrastructural immunolabel on striatal dopamine terminals [79], and evoked dopamine release can be inhibited via group I metabotropic glutamate receptors on dopamine terminals [80]. By monitoring evoked dopamine release directly using FCV, we found a reciprocal modulation by glutamate spillover on evoked striatal DA release, induced by either glutamate uptake blockade or high-frequency stimulation of corticostriatal tracts. Therefore, glutamate released from the cortical terminals during short bursts of cortical activity may escape the confines of its synapse in the striatum and depress dopaminergic transmission by activating presynaptic group I mGluRs. This inhibition is mediated by the mobilization of internal Ca2+ stores which in turn activates apamin-sensitive Ca2+ -dependent potassium currents.

3.2. GABA receptors

By local infusion of GABAA and GABAB receptor agonists and antagonists in the striatum of intact and kainic acid lesioned rats, microdialysis data support a direct influence of GABA on the dopaminergic terminals via presynaptic GABAB receptors, while the effects via the GABAA receptors seem to be postsynaptic and mediated by striatal interneurons [81]. A direct effect of GABAB receptors on dopamine release is further supported by a study using FCV recordings, which shows that GABAB receptor agonists inhibit single-pulse-evoked dopamine release in the striatal slice with kinetic parameters similar to those of the D2 autoreceptor [24]. There is ultrastructural evidence consistent with expression of GABAB receptors on dopamine terminals [68].

Functional studies suggest that GABAA receptors might be colocalized on the dopamine terminals. Muscimol, a GABAA receptor agonist, inhibits the evoked dopamine release in striatal synaptosomes [82] and also inhibits evoked dopamine release by single pulse stimulation measured by FCV in striatal slices (Zhang and Sulzer, unpublished results). However, the results could be complicated by possible actions on cholinergic interneurons. Further experiments with depletion of ACh tone are needed to determine whether the regulation is direct or not (please also refer to the conclusion section).

3.3. Acetylcholine receptors

Classical pharmacological studies of the effects of muscarinic acetylcholine receptors (mAChRs) on dopamine release in the striatum have led to contradictory results [83-89], due in part to the diversity of the mAChR subtypes and the limited receptor subtype selectivity of the muscarinic agonists and antagonists [90]. Activation of mAChRs was shown to inhibit dopamine release in dorsal striatal slices examined by FCV [91]. A human-brain imaging study indicated a tonic muscarinic inhibition of dopamine release in striatum [92]. A study using genetically altered mice that lacked functional M1-M5 mAChRs provides evidence of the different physiological roles of individual mAChRs in a direct manner [13]. The results show that M3 receptors inhibit potassium-stimulated [3H] dopamine release, whereas M4 and M5 receptors facilitate release, and M1 and M2 receptors have no effect. It seems that the modulating effects of M3 and M4 receptors are mediated via striatal GABA release. This view however has been recently challenged by studies by Cragg’s group suggesting that modulating effects of mAChRs are mediated by effects on cholinergic interneurons [93, 94]. Application of a mixture of antagonists for GABA and glutamate receptors does not alter the inhibitory effect of oxotremorine-M (Oxo-M), a broad mAChR agonist on dopamine release, ruling out the involvement of GABA or glutamate as the intermediary. In contrast, blockade of β2-nAChRs completely precludes the effect of Oxo-M, suggesting mAChRs govern dopamine release via manipulating ACh tone at β2-nAChRs on DA terminals [93].

M5 receptor mRNA is the only mAChR subtype mRNA detectable in the dopamine-containing cells of the substantia nigra pars compacta [95, 96], strongly suggesting that the dopamine release-facilitating M5 receptors are located on dopaminergic nerve terminals [97, 98]. Recently FCV studies in striatal slices show that evoked dopamine release in M5 mAChR KO mice is reduced, whereas inhibition of evoked dopamine release by oxotremorine-M (Oxo-M), a broader mAChR agonist is enhanced, consistent with a release-potentiating function of dopaminergic terminal M5 receptors [98]. Oxo-M can inhibit dopamine release indirectly by acting on M2, M4 receptors on cholinergic interneurons, whereas enhance dopamine release directly by acting on M5 receptor on dopamine axons [93]. The overall effect of Oxo-M on dopamine release is inhibitory as shown by experiments [13, 93, 98] probably due to the greater affinity of ACh for nAChRs than mAChRs. In the M5 mAChR KO mice, the inhibition effect of Oxo-M is enhanced because of the loss of the excitatory component. Although this explanation seems reasonable, to some extent, it conflicts with the fact that blockade of β2-nAChRs completely precludes the effect of Oxo-M [93]. Other possibilities such as down-regulation of β2-nAChRs in the M5 KO mice should be explored in the future.

Nicotinic acetylcholine receptors possessing β2-subunits are expressed on dopamine terminals in the striatum all [64, 99, 100]. Nicotine enhanced the extracellular dopamine level by microdialysis [101] and results in vivo indicate that nicotine, like cocaine and alcohol, increase the frequency of non-evoked dopamine transients in the nucleus accumbens [102]. FCV studies demonstrated an inhibition of evoked dopamine release in slices by nicotinic agonists [91, 103]. Because nAChR antagonists also inhibit dopamine release, it appears that activation of nAChRs by endogenous ACh is excitatory for dopamine release, but the receptor is rapidly desensitized by nicotine application, and so nicotine inhibits evoked dopamine release by single pulse stimulation [103-105]. In summary, activation of nAChRs normally keeps dopamine release probability high, but curbs subsequent release during pulse trains due to synaptic depression; nAChR blockade or desensitization decreases dopamine release probability, but facilities subsequent release during high frequencies ( please refer to section 4 for further discussion).

3.4. Opioid receptors

Kappa-opioid receptors are located on dopamine axon terminals [65] while mu-opioid receptors are not expressed on striatal dopamine axon terminals [106]. Using FCV, Schlosser et al. [107] first demonstrated that mu, delta, and kappa opioid receptors inhibit striatal dopamine release. It seems that the effect of kappa opioid receptors on dopamine overflow is likely to be direct, while the influence of mu opioid receptors is indirect and mediated by an inhibition of cholinergic interneuron activity, since nicotine produced negligible effects on dopamine release when applied subsequent to a saturating dose of endomorphin-1, a mu receptor agonist [65, 108]. The similarities between the effects on evoked dopamine overflow elicited by mu-opioid agonists and nicotine suggested a common mechanism of action. Although kappa-agonists profoundly inhibit single-pulse-evoked dopamine overflow throughout the striatum, this inhibition is unusual in that it cannot be overcome by high-frequency stimuli. This is confirmed by the observation that kappa-opioid receptor activation did not alter the paired-pulse ratio of dopamine release, suggesting that the effects are not mediated by changes in either calcium influx or axon terminal membrane potential [108]. Whether the effect of delta opioid receptors on dopamine release is direct remains unclear, but ultrastructural studies show that delta opioid receptors are present on dopamine terminals [66].

3.5. Adenosine receptors

Adenosine, an important neuromodulator in the brain, is a metabolite generated by dephosphoralyation of ATP as a function of neuronal activity. There are four adenosine receptors, A1, A2a, A2b and A3. The A2 subtypes are Gs /Golf /Gq coupled, stimulating adenylyl cyclase, whilst A1 and A3 couple to Gi /Go and inhibit adenylyl cyclase. In the basal ganglia, the receptors of interest are A1 and A2a [109].

It remains unclear whether there is a direct heteroreceptor modulation by adenosine on dopamine release. In the striatum, activation of A1 receptors inhibits [110-112], whereas activation of A2a receptors enhances dopamine release [113]. FCV studies also demonstrate that A1 receptor agonist decrease single-pulse-evoked dopamine release in vitro, but the inhibition is dependent at least in part on the simultaneous activation of D1 dopamine receptors. While the mechanism underlying this interaction remains to be determined, it does not appear to involve an intramembrane interaction between A1 and D1 receptors [112, 114].

To date, the morphological evidence of presynaptic localization of A1 receptors on dopamine terminals is indirect [115-117]. Furthermore, the lesion of glutamatergic, but not dopaminergic, striatal afferents significantly decreases striatal A1 receptor function and agonist binding [118]. In addition, there are no A2a receptors on dopamine terminals [119, 120]. Therefore, the main mechanism underlying adenosine-mediated modulation of striatal dopamine release may be indirect.

3.6. Cannabinoid receptors

These receptors consist of two GPCRs, CB1 and CB2, with the CB1 highly expressed in CNS [121-123]. Marijuana and its main psychoactive component, delta9-tetrahydrocannoabinol (THC) act on CB1 receptors in the CNS. CB1 receptor is coupled through G0 or Gi to inhibit adenylyl cyclase, and activation of presynaptic CB1 receptors is well known to inhibit glutamate and GABA release [124]. FCV studies have not identified a direct presynaptic modulation of dopamine release by CB1 receptors in either dorsal striatum or ventral striatum [80, 125, 126], in agreement with the lack of anatomical evidence of CB1 receptors on dopamine terminals [127-131]. Nevertheless, there are indirect effects by cannabinoids on dopamine release. Ventral tegmental area (VTA) dopamine neurons are thought to produce the cannabinoids [132-134], which would be expected to activate local receptors. CB1 agonists inhibit evoked dopamine release while increasing the frequency of non-evoked dopamine concentration transients in rat striatum in vivo, responses that may be related to effects on dopamine neuronal firing via a reduction of afferent GABAergic transmission [61]. This indirect influence on dopamine cell activity and then release may underlie the psychoactive actions of cannabis.

CB1 agonists alter motor function in a complex way and the suppression of movement might be due to inhibition of dopamine release in the dorsal striatum [135]. Although a CB1 agonist had no effect on single-pulse-evoked-dopamine in striatal slices, it decreased dopamine release over trains of stimuli, suggesting cannabinoids exert indirect changes via a local striatal circuit. Rice and colleagues have suggested this could occur via a nonsynaptic mechanism involving inhibition of GABA release, generation of hydrogen peroxide, and activation of KATP channels to inhibit dopamine release [126].

Recently an intriguing study demonstrates that the inhibition of dopamine release by D2 autoreceptor is attenuated by activation of CB1 receptors [136]. Although the underlying mechanisms are not clear, the results provide a functional correlate to previous observations that CB1 receptors are antagonistic to D2 receptors in the region [137].

In addition, there are additional ionotropic and G-protein-linked receptor candidates that may act as heteroreceptors on dopamine terminals, and elucidation of their effects is fundamental for understanding their roles in modulating dopaminergic transmission.

4. Frequency dependent modulation of dopamine release

Dopaminergic neurons exhibit two patterns of discharge activity: a continuous mode with regularly spaced spikes at a frequency between 2 and 5 Hz, and a bursting activity characterized by brief bursts of 2 to 6 action potentials [138]. Single dopaminergic neurons switch between the patterns. In resting condition and during sleep most dopaminergic neurons discharge with the tonic mode, but rewarding or sensorial stimuli predicting a reward trigger in most dopaminergic neurons a single burst both in rats [139] and monkeys [140, 141]. In rats the intra-burst frequency is 15-30 Hz [138, 139]. Grace and Bunney hypothesized that the bursting mode would be more potent than the tonic pattern to induce dopamine release [138]. Accordingly, electrical stimulations mimicking the bursting mode were twice as potent as regularly spaced stimulation having the same whole duration and number of pul se to enhance the extracellular dopamine level in vivo [142].

As detailed above, numerous studies provide evidence for heteroreceptor regulation of dopamine release, although some results are controversial due in large part to different preparations, stimulation and recording paradigms. In vivo and in vitro data in slices with local stimulation cannot exclude circuit effects. There are striking differences between such results in the slice preparation where there is significant paired pulse depression, and in vivo, where there is typically no detectable depression. While the basis for such differences remains unclear an important factor is presumably due to the loss of most ongoing synaptic activity in the slice. Furthermore, most previous voltammetry studies in striatal slices only examine the effects evoked by single pulse stimuli mimicking the tonic firing mode, while recent studies have demonstrated the frequency dependent modulation of dopamine release by nAChRs [104, 105, 143].

Nicotine shifts VTA neurons from tonic to burst firing modes [5, 144] and enhances extracellular dopamine as measured by microdialysis [101]. On the other hand, nicotine at levels thought to be experienced by smokers (~250-300 nM) desensitizes nAChRs so rapidly that tonic ACh activation is blocked and evoked dopamine release is potently inhibited [91, 103].

In order to resolve the question of how nicotine both elevates extracellular dopamine and depresses evoked dopamine release, the effects of nicotine on the modulation of evoked dopamine release were compared under different firing patterns and found to be dependent on the firing pattern of dopamine neurons. While desensitization of nAChRs indeed curbs dopamine released by stimuli emulating tonic firing, it allows a rapid rise in dopamine from stimuli emulating the phasic firing patterns associated with incentive/salience paradigms. Nicotine may thus enhance the contrast of dopamine signals associated with behavioral cues [104, 105] (Fig. 1). Interestingly, mu opioid agonists inhibit dopamine overflow elicited with single-pulse stimuli while leaving that produced by burst stimuli unaffected, but these differential effects are mediated by nAChRs and caused by inhibition of cholinergic interneurons [108].

Figure 1. Modulation of evoked dopamine (DA) release by nAChRs depends on firing pattern.

Figure 1

(a,b) Voltammetric responses before and after 10-min bath application of nicotine (Nic, 300 nM) or mecamylamine (Meca, 2 μM) at different stimuli. The inhibition of evoked release was not blocked by D2 dopamine, GABAA, GABAB or ionotropic glutamate receptor antagonists. (c,d) Frequency modulation of nicotine and mecamylamine effects on dopamine release (mean ±s.e.m, n = 8 for control; n = 5 for nicotine, n = 4 for mecamylamine; * P < 0.05 compared with respective control values by Student’s t-test). Top panels: evoked dopamine release normalized to that elicited by 1p stimulation under control condition. Bottom panels: relative evoked dopamine release after nicotine and mecamylamine at different stimulation frequencies. (eh) Effects of number of pulses, nicotine and mecamylamine on dopamine release at 20 Hz (e,f) and 100 Hz (g,h) (mean ± s.e.m., n = 4–8, * P < 0.05 compared with respective control values by Student’s t-test). Data from Zhang and Sulzer (2004), copyright by Nature Publishing Group.

Activation of mAChRs also inhibits single-pulse-evoked dopamine release while the depression is alleviated at high frequency stimuli. This frequency dependent modulation is again mediated by nAChRs and cholinergic interneurons: activation of mAChRs on cholinergic interneurons inhibits their firing [145] and thus release, which decreases the activation of nAChRs on dopamine terminals [93, 94].

This frequency dependent modulation may apply to other heteroreceptors, in addition to nAChRs, mAChRs and opioids mu receptors. In order to have much more dopamine to be released during phasic firing compared to tonic firing, the release probability must be low. A broad range of studies by multiple groups using the acute striatal slice find no detectable tonic activation of iGluRs, mGluRs, GABAA receptors (GABAAR) , GABAB receptors (GABABR), mAChRs, adenosine subtype receptor A2aR and A1R, opioid receptors, and D2 autoreceptor (D2R), and tonic activation appears to be limited to nAChRs (Table 1). This is however directly only applicable to the slice, in which with the exception of cholinergic interneurons, there are very few active neurons. This clearly strongly contrasts with the living organism, in which there are certainly highly variable and complex inputs from multiple pathways, and where antagonists might be expected to thus show very complex effects.

Table 1.

Heteroreceptor Regulation of Evoked Dopamine Release by 1 pulse stimulation

Receptors Tonic Tone
(revealed by
antagonists)
Agonist/Effect Notes Subregion
iGluRs
--NMDA
--AMPA
--KA
No NMDA
AMPA
KA
Inhibition within 4
min but mass
release due to
spreading
depression
Dorsal
striatum
mGluRs No Inhibition
(DHPG)
through mGluR1 Dorsal
striatum
GABAAR No Inhibition Dorsal and
ventral
striatum
GABABR No Inhibition Dorsal and
ventral
striatum
Adenosine
(A1R, A2aR)
No A1: Inhibition Dorsal
striatum
Opioid receptors
(mu, delta,
kappa)
No Inhibition ventral
striatum
Cannabinoid
CB1
No NO Dorsal and
ventral
striatum
mAChRs No Inhibition Dorsal and
ventral
striatum
nAChRs Yes
(excitatory)
Dorsal and
ventral
striatum

There is no detectable tonic activation of iGluRs, mGluRs, GABAAR, GABABR, mAChRs, adenosine subtype receptor A2aR and A1R, opioid receptors, and D2 autoreceptors in the striatal slice, except nAChRs. Activation of these receptors inhibits evoked dopamine release elicited by single pulse at various levels of the controls (ref [75, 80, 91, 93, 103, 104, 105, 108, 112, 126, 136, 146]).

Remarkably, with the exception of ACh receptors, the effects of receptor agonists indicate that most heteroreceptors studied to date have been found to depress dopamine release (Table 1). Therefore, in addition to D2Rs, most heteroreceptors in vivo may maintain the dopamine release probability at a low level, which may then provide a marked and rapid increase in dopamine concentration during phasic firing. We speculate the outcome of change of dopamine release by the various heteroreceptors will depend critically on the accompanying changes in the cholinergic interneurons. While dopamine neurons signal salient contextual stimuli by brief burst firing, simultaneously striatal cholinergic interneurons typically pause, which will lead to decreased ACh tone and less activation of nAChRs on dopamine terminals. Thus, the pauses in cholinergic neurons enhance the sensitivity of dopamine synapse to temporally correlated changes in dopamine neurons activity. A range of drugs that affect the dopamine system may exert their actions via altering the signal/noise ratio of dopamine by affecting heteroreceptors on dopamine terminals as well as on cell bodies [146].

5. Conclusions

The striatum is a highly heterogeneous structure organized into different but overlapping territories [147]. Modulation of dopamine release by heteroreceptors moreover appears to be region specific and dependent. For example, although a similar frequency dependent modulation by nicotine is found in the dorsal striatum as in the shell, the increased ratio of phasic burst relative to tonic firing caused by nicotine is smaller compared to that in the shell [143]. In the dorsal striatum, a region associated with sensorimotor function, both M2 and M4 mAChRs are necessary for mAChR regulation of dopamine release, whereas in the ventral striatum, a region associated with limbic function, only the M4 mAChR is necessary for mAChRs regulation of dopamine release [93]. The modulation of dopamine release by mu opioid receptors only exists in specific subregions of the nucleus accumbens [108].

We have reviewed here most of the data using FCV in striatal slices. Whereas the effect on dopamine release by single pulse stimulation usually suggests a direct action, however, the indirect circuitry effect cannot be completely excluded given the tonic cholinergic tone in the striatum. It is challenging to distinguish a direct effect on the heteroreceptors expressed on dopamine terminals from an indirect effect through the control of ACh release by influencing heteroreceptors expressed on the cholinergic interneurons, unless completely depleting the cholinergic tone first.

Optical techniques appear poised to further define effects of presynaptic receptors on release probability [148-150]. Using a fluorescent false neurotransmitter to directly visualize dopamine release from individual terminals, we have found a frequency-dependent heterogeneity of presynaptic terminals, which is dependent in part on D2 receptors, indicating a mechanism for frequency-dependent coding of presynaptic selection [151]. These data suggest that modulation of dopamine release is not fixed; it could be frequency dependent, region specific, sub-territory specific, even individual terminal specific. We should therefore consider all of these when considering potential therapeutic targets for treating dopamine related disorders.

Acknowledgements

This work was supported by DA007418, the Picower Foundation, and the Parkinson’s Disease Foundation.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • [1].Berke JD, Hyman SE. Addiction, dopamine, and the molecular mechanisms of memory. Neuron. 2000;25:515–32. doi: 10.1016/s0896-6273(00)81056-9. [DOI] [PubMed] [Google Scholar]
  • [2].Wickens JR, Horvitz JC, Costa RM, Killcross S. Dopaminergic mechanisms in actions and habits. J Neurosci. 2007;27:8181–3. doi: 10.1523/JNEUROSCI.1671-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [3].Wise RA. Dopamine, learning and motivation. Nat Rev Neurosci. 2004;5:483–94. doi: 10.1038/nrn1406. [DOI] [PubMed] [Google Scholar]
  • [4].Schmitz Y, Benoit-Marand M, Gonon F, Sulzer D. Presynaptic regulation of dopaminergic neurotransmission. J Neurochem. 2003;87:273–89. doi: 10.1046/j.1471-4159.2003.02050.x. [DOI] [PubMed] [Google Scholar]
  • [5].Mansvelder HD, De Rover M, McGehee DS, Brussaard AB. Cholinergic modulation of dopaminergic reward areas: Upstream and downstream targets of nicotine addiction. Eur J Pharmacol. 2003;480:117–23. doi: 10.1016/j.ejphar.2003.08.099. [DOI] [PubMed] [Google Scholar]
  • [6].Cruz HG, Ivanova T, Lunn ML, Stoffel M, Slesinger PA, Luscher C. Bi-directional effects of GABAB receptor agonists on the mesolimbic dopamine system. Nat Neurosci. 2004;7:153–9. doi: 10.1038/nn1181. [DOI] [PubMed] [Google Scholar]
  • [7].Bonci A, Bernardi G, Grillner P, Mercuri NB. The dopamine-containing neuron: Maestro or simple musician in the orchestra of addiction? Trends Pharmacol Sci. 2003;24:172–7. doi: 10.1016/S0165-6147(03)00068-3. [DOI] [PubMed] [Google Scholar]
  • [8].Margolis EB, Hjelmstad GO, Bonci A, Fields HL. Kappa-opioid agonists directly inhibit midbrain dopaminergic neurons. J Neurosci. 2003;23:9981–6. doi: 10.1523/JNEUROSCI.23-31-09981.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [9].Sulzer D. How addictive drugs disrupt presynaptic dopamine neurotransmission. Neuron. 2011;69:628–49. doi: 10.1016/j.neuron.2011.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [10].Kelly RS, Wightman RM. Detection of dopamine overflow and diffusion with voltammetry in slices of rat brain. Brain Res. 1987;423:79–87. doi: 10.1016/0006-8993(87)90827-4. [DOI] [PubMed] [Google Scholar]
  • [11].Phillips PE, Stamford JA. Differential recruitment of N-, P- and Q-type voltage-operated calcium channels in striatal dopamine release evoked by ‘regular’ and ‘burst’ firing. Brain Res. 2000;884:139–46. doi: 10.1016/s0006-8993(00)02958-9. [DOI] [PubMed] [Google Scholar]
  • [12].Baik JH, Picetti R, Saiardi A, Thiriet G, Dierich A, Depaulis A, Le Meur M, Borrelli E. Parkinsonian-like locomotor impairment in mice lacking dopamine D2 receptors. Nature. 1995;377:424–8. doi: 10.1038/377424a0. [DOI] [PubMed] [Google Scholar]
  • [13].Zhang W, Yamada M, Gomeza J, Basile AS, Wess J. Multiple muscarinic acetylcholine receptor subtypes modulate striatal dopamine release, as studied with M1-M5 muscarinic receptor knock-out mice. J Neurosci. 2002;22:6347–52. doi: 10.1523/JNEUROSCI.22-15-06347.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [14].Cass WA, Gerhardt GA. Direct in vivo evidence that D2 dopamine receptors can modulate dopamine uptake. Neurosci Lett. 1994;176:259–63. doi: 10.1016/0304-3940(94)90096-5. [DOI] [PubMed] [Google Scholar]
  • [15].Rothblat DS, Schneider JS. Regionally specific effects of haloperidol and clozapine on dopamine reuptake in the striatum. Neurosci Lett. 1997;228:119–22. doi: 10.1016/s0304-3940(97)00377-7. [DOI] [PubMed] [Google Scholar]
  • [16].Sesack SR, Aoki C, Pickel VM. Ultrastructural localization of D2 receptor-like immunoreactivity in midbrain dopamine neurons and their striatal targets. J Neurosci. 1994;14:88–106. doi: 10.1523/JNEUROSCI.14-01-00088.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Centonze D, Usiello A, Gubellini P, Pisani A, Borrelli E, Bernardi G, Calabresi P. Dopamine D2 receptor-mediated inhibition of dopaminergic neurons in mice lacking D2l receptors. Neuropsychopharmacology. 2002;27:723–6. doi: 10.1016/S0893-133X(02)00367-6. [DOI] [PubMed] [Google Scholar]
  • [18].Usiello A, Baik JH, Rouge-Pont F, Picetti R, Dierich A, LeMeur M, Piazza PV, Borrelli E. Distinct functions of the two isoforms of dopamine D2 receptors. Nature. 2000;408:199–203. doi: 10.1038/35041572. [DOI] [PubMed] [Google Scholar]
  • [19].Wang Y, Xu R, Sasaoka T, Tonegawa S, Kung MP, Sankoorikal EB. Dopamine D2 long receptor-deficient mice display alterations in striatum-dependent functions. J Neurosci. 2000;20:8305–14. doi: 10.1523/JNEUROSCI.20-22-08305.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [20].Bello EP, Mateo Y, Gelman DM, Noain D, Shin JH, Low MJ, Alvarez VA, Lovinger DM, Rubinstein M. Cocaine supersensitivity and enhanced motivation for reward in mice lacking dopamine D2 autoreceptors. Nat Neurosci. 2011;14:1033–8. doi: 10.1038/nn.2862. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [21].Benoit-Marand M, Borrelli E, Gonon F. Inhibition of dopamine release via presynaptic D2 receptors: Time course and functional characteristics in vivo. J Neurosci. 2001;21:9134–41. doi: 10.1523/JNEUROSCI.21-23-09134.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [22].L’hirondel M, Cheramy A, Godeheu G, Artaud F, Saiardi A, Borrelli E, Glowinski J. Lack of autoreceptor-mediated inhibitory control of dopamine release in striatal synaptosomes of D2 receptor-deficient mice. Brain Res. 1998;792:253–62. doi: 10.1016/s0006-8993(98)00146-2. [DOI] [PubMed] [Google Scholar]
  • [23].Mercuri NB, Saiardi A, Bonci A, Picetti R, Calabresi P, Bernardi G, Borrelli E. Loss of autoreceptor function in dopaminergic neurons from dopamine D2 receptor deficient mice. Neuroscience. 1997;79:323–7. doi: 10.1016/s0306-4522(97)00135-8. [DOI] [PubMed] [Google Scholar]
  • [24].Schmitz Y, Schmauss C, Sulzer D. Altered dopamine release and uptake kinetics in mice lacking D2 receptors. J Neurosci. 2002 doi: 10.1523/JNEUROSCI.22-18-08002.2002. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [25].Diaz J, Pilon C, Le Foll B, Gros C, Triller A, Schwartz JC, Sokoloff P. Dopamine D3 receptors expressed by all mesencephalic dopamine neurons. J Neurosci. 2000;20:8677–84. doi: 10.1523/JNEUROSCI.20-23-08677.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [26].Gainetdinov RR, Grekhova TV, Sotnikova TD, Rayevsky KS. Dopamine D2 and D3 receptor preferring antagonists differentially affect striatal dopamine release and metabolism in conscious rats. Eur J Pharmacol. 1994;261:327–31. doi: 10.1016/0014-2999(94)90125-2. [DOI] [PubMed] [Google Scholar]
  • [27].Sokoloff P, Giros B, Martres M, Bouthene M, Schwartz J. Molecular cloning and characterization of a novel dopamine receptor (D3) as a target for neuroleptics. Nature. 1990;347:146–51. doi: 10.1038/347146a0. [DOI] [PubMed] [Google Scholar]
  • [28].Tepper JM, Sun BC, Martin LP, Creese I. Functional roles of dopamine D2 and D3 autoreceptors on nigrostriatal neurons analyzed by antisense knockdown in vivo. J Neurosci. 1997;17:2519–30. doi: 10.1523/JNEUROSCI.17-07-02519.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [29].Zapata A, Witkin JM, Shippenberg TS. Selective D3 receptor agonist effects of (+)-PD 128907 on dialysate dopamine at low doses. Neuropharmacology. 2001;41:351–9. doi: 10.1016/s0028-3908(01)00069-7. [DOI] [PubMed] [Google Scholar]
  • [30].Koeltzow TE, Xu M, Cooper DC, Hu XT, Tonegawa S, Wolf ME, White FJ. Alterations in dopamine release but not dopamine autoreceptor function in dopamine D3 receptor mutant mice. J Neurosci. 1998;18:2231–8. doi: 10.1523/JNEUROSCI.18-06-02231.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [31].Joseph JD, Wang YM, Miles PR, Budygin EA, Picetti R, Gainetdinov RR, Caron MG, Wightman RM. Dopamine autoreceptor regulation of release and uptake in mouse brain slices in the absence of D3 receptors. Neuroscience. 2002;112:39–49. doi: 10.1016/s0306-4522(02)00067-2. [DOI] [PubMed] [Google Scholar]
  • [32].Kuzhikandathil EV, Oxford GS. Activation of human D3 dopamine receptor inhibits P/Q-type calcium channels and secretory activity in AtT-20 cells. J Neurosci. 1999;19:1698–707. doi: 10.1523/JNEUROSCI.19-05-01698.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [33].Kuzhikandathil EV, Oxford GS. Dominant-negative mutants identify a role for GIRK channels in D3 dopamine receptor-mediated regulation of spontaneous secretory activity. J Gen Physiol. 2000;115:697–706. doi: 10.1085/jgp.115.6.697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [34].Kuzhikandathil EV, Yu W, Oxford GS. Human dopamine D3 and D2L receptors couple to inward rectifier potassium channels in mammalian cell lines. Mol Cell Neurosci. 1998;12:390–402. doi: 10.1006/mcne.1998.0722. [DOI] [PubMed] [Google Scholar]
  • [35].Tang L, Todd RD, O’Malley KL. Dopamine D2 and D3 receptors inhibit dopamine release. J Pharmacol Exp Ther. 1994;270:475–9. [PubMed] [Google Scholar]
  • [36].Davila V, Yan Z, Craciun LC, Logothetis D, Sulzer D. D3 dopamine autoreceptors do not activate G-protein-gated inwardly rectifying potassium channel currents in substantia nigra dopamine neurons. J Neurosci. 2003;23:5693–7. doi: 10.1523/JNEUROSCI.23-13-05693.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [37].Svingos AL, Periasamy S, Pickel VM. Presynaptic dopamine D(4) receptor localization in the rat nucleus accumbens shell. Synapse. 2000;36:222–32. doi: 10.1002/(SICI)1098-2396(20000601)36:3<222::AID-SYN6>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
  • [38].Starke K, Gother M, Kilbinger H. Modulation of neurotransmitter release by presynaptic autoreceptors. Physiol Rev. 1989;69:864–989. doi: 10.1152/physrev.1989.69.3.864. [DOI] [PubMed] [Google Scholar]
  • [39].Dwoskin LP, Zahniser NR. Robust modulation of [3H]dopamine release from rat striatal slices by D-2 dopamine receptors. J Pharmacol Exp Ther. 1986;239:442–53. [PubMed] [Google Scholar]
  • [40].Cubeddu LX, Hoffmann IS. Operational characteristics of the inhibitory feedback mechanism for regulation of dopamine release via presynaptic receptors. J Pharmacol Exp Ther. 1982;223:497–501. [PubMed] [Google Scholar]
  • [41].Cragg SJ, Greenfield SA. Differential autoreceptor control of somatodendritic and axon terminal dopamine release in substantia nigra, ventral tegmental area, and striatum. J Neurosci. 1997;17:5738–46. doi: 10.1523/JNEUROSCI.17-15-05738.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [42].Starke K, Reimann W, Zumstein A, Hertting G. Effect of dopamine receptor agonists and antagonists on release of dopamine in the rabbit caudate nucleus in vitro. Naunyn Schmiedebergs Arch Pharmacol. 1978;305:27–36. doi: 10.1007/BF00497003. [DOI] [PubMed] [Google Scholar]
  • [43].Kennedy RT, Jones SR, Wightman RM. Dynamic observation of dopamine autoreceptor effects in rat striatal slices. J Neurochem. 1992;59:449–55. doi: 10.1111/j.1471-4159.1992.tb09391.x. [DOI] [PubMed] [Google Scholar]
  • [44].Palij P, Bull DR, Sheehan MJ, Millar J, Stamford J, Kruk ZL, Humphrey PP. Presynaptic regulation of dopamine release in corpus striatum monitored in vitro in real time by fast cyclic voltammetry. Brain Res. 1990;509:172–4. doi: 10.1016/0006-8993(90)90329-a. [DOI] [PubMed] [Google Scholar]
  • [45].Mayer A, Limberger N, Starke K. Transmitter release patterns of noradrenergic, dopaminergic and cholinergic axons in rabbit brain slices during short pulse trains, and the operation of presynaptic autoreceptors. Naunyn Schmiedebergs Arch Pharmacol. 1988;338:632–43. doi: 10.1007/BF00165627. [DOI] [PubMed] [Google Scholar]
  • [46].Phillips PE, Hancock PJ, Stamford JA. Time window of autoreceptor-mediated inhibition of limbic and striatal dopamine release. Synapse. 2002;44:15–22. doi: 10.1002/syn.10049. [DOI] [PubMed] [Google Scholar]
  • [47].Cardozo DL, Bean BP. Voltage-dependent calcium channels in rat midbrain dopamine neurons modulation by dopamine and GABAB receptors. J Neurophysiol. 1995;74:1137–48. doi: 10.1152/jn.1995.74.3.1137. [DOI] [PubMed] [Google Scholar]
  • [48].Chen BT, Moran KA, Avshalumov MV, Rice ME. Limited regulation of somatodendritic dopamine release by voltage-sensitive Ca2+ channels contrasted with strong regulation of axonal dopamine release. J Neurochem. 2006;96:645–55. doi: 10.1111/j.1471-4159.2005.03519.x. [DOI] [PubMed] [Google Scholar]
  • [49].Congar P, Bergevin A, Trudeau LE. D2 receptors inhibit the secretory process downstream from calcium influx in dopaminergic neurons: Implication of K+ channels. J Neurophysiol. 2002;87:1046–56. doi: 10.1152/jn.00459.2001. [DOI] [PubMed] [Google Scholar]
  • [50].Sulzer D, Joyce MP, Lin L, Geldwert D, Haber SN, Hattori T, Rayport S. Dopamine neurons make glutamatergic synapses in vitro. J Neurosci. 1998;18:4588–602. doi: 10.1523/JNEUROSCI.18-12-04588.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [51].Cass WA, Zahniser NR. Potassium channel blockers inhibit D2 dopamine, but not A1 adenosine, receptor-mediated inhibition of striatal dopamine release. J Neurochem. 1991;57:147–52. doi: 10.1111/j.1471-4159.1991.tb02109.x. [DOI] [PubMed] [Google Scholar]
  • [52].Fulton S, Thibault D, Mendez JA, Lahaie N, Tirotta E, Borrelli E, Bouvier M, Tempel BL, Trudeau LE. Contribution of Kv1.2 voltage-gated potassium channel to D2 autoreceptor regulation of axonal dopamine overflow. J Biol Chem. 2011;286:9360–72. doi: 10.1074/jbc.M110.153262. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [53].Martel P, Leo D, Fulton S, Berard M, Trudeau LE. Role of Kv1 potassium channels in regulating dopamine release and presynaptic D2 receptor function. PLoS One. 2011;6:e20402. doi: 10.1371/journal.pone.0020402. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [54].Webb CK, McCudden CR, Willard FS, Kimple RJ, Siderovski DP, Oxford GS. D2 dopamine receptor activation of potassium channels is selectively decoupled by Gαi-specific GoLoco motif peptides. J Neurochem. 2005;92:1408–18. doi: 10.1111/j.1471-4159.2004.02997.x. [DOI] [PubMed] [Google Scholar]
  • [55].Patel JC, Witkovsky P, Coetzee WA, Rice ME. Subsecond regulation of striatal dopamine release by presynaptic KATP channels. J Neurochem. 2011 doi: 10.1111/j.1471-4159.2011.07358.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [56].Imperato A, Di Chiara G. Dopamine release and metabolism in awake rats after systemic neuroleptics as studied by trans-striatal dialysis. J Neurosci. 1985;5:297–306. doi: 10.1523/JNEUROSCI.05-02-00297.1985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [57].Imperato A, Di Chiara G. Effects of locally applied D-1 and D-2 receptor agonists and antagonists studied with brain dialysis. Eur J Pharmacol. 1988;156:385–93. doi: 10.1016/0014-2999(88)90284-1. [DOI] [PubMed] [Google Scholar]
  • [58].Westerink BH, de Vries JB. On the mechanism of neuroleptic induced increase in striatal dopamine release: Brain dialysis provides direct evidence for mediation by autoreceptors localized on nerve terminals. Neurosci Lett. 1989;99:197–202. doi: 10.1016/0304-3940(89)90289-9. [DOI] [PubMed] [Google Scholar]
  • [59].May LJ, Wightman RM. Effects of D-2 antagonists on frequency-dependent stimulated dopamine overflow in nucleus accumbens and caudate-putamen. J Neurochem. 1989;53:898–906. doi: 10.1111/j.1471-4159.1989.tb11789.x. [DOI] [PubMed] [Google Scholar]
  • [60].Suaud-Chagny MF, Ponec J, Gonon F. Presynaptic autoinhibition of the electrically evoked dopamine release studied in the rat olfactory tubercle by in vivo electrochemistry. Neuroscience. 1991;45:641–52. doi: 10.1016/0306-4522(91)90277-u. [DOI] [PubMed] [Google Scholar]
  • [61].Cheer JF, Wassum KM, Heien ML, Phillips PE, Wightman RM. Cannabinoids enhance subsecond dopamine release in the nucleus accumbens of awake rats. J Neurosci. 2004;24:4393–400. doi: 10.1523/JNEUROSCI.0529-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [62].Luscher C, Ungless MA. The mechanistic classification of addictive drugs. PLoS Med. 2006;3:e437. doi: 10.1371/journal.pmed.0030437. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [63].Araujo DM, Hilt DC, Miller PJ, Wen D, Jiao S, Lapchak PA. Ret receptor tyrosine kinase immunoreactivity is altered in glial cell line-derived neurotrophic factor-responsive neurons following lesions of the nigrostriatal and septohippocampal pathways. Neuroscience. 1997;80:9–16. doi: 10.1016/s0306-4522(97)00122-x. [DOI] [PubMed] [Google Scholar]
  • [64].Wonnacott S, Kaiser S, Mogg A, Soliakov L, Jones IW. Presynaptic nicotinic receptors modulating dopamine release in the rat striatum. Eur J Pharmacol. 2000;393:51–8. doi: 10.1016/s0014-2999(00)00005-4. [DOI] [PubMed] [Google Scholar]
  • [65].Svingos AL, Chavkin C, Colago EE, Pickel VM. Major coexpression of kappa-opioid receptors and the dopamine transporter in nucleus accumbens axonal profiles. Synapse. 2001;42:185–92. doi: 10.1002/syn.10005. [DOI] [PubMed] [Google Scholar]
  • [66].Svingos AL, Clarke CL, Pickel VM. Localization of the delta-opioid receptor and dopamine transporter in the nucleus accumbens shell: Implications for opiate and psychostimulant cross-sensitization. Synapse. 1999;34:1–10. doi: 10.1002/(SICI)1098-2396(199910)34:1<1::AID-SYN1>3.0.CO;2-H. [DOI] [PubMed] [Google Scholar]
  • [67].Paquet M, Tremblay M, Soghomonian JJ, Smith Y. AMPA and NMDA glutamate receptor subunits in midbrain dopaminergic neurons in the squirrel monkey: An immunohistochemical and in situ hybridization study. J Neurosci. 1997;17:1377–96. doi: 10.1523/JNEUROSCI.17-04-01377.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [68].Charara A, Heilman C, Levey AI, Smith Y. Pre- and postsynaptic localization of GABAB receptors in the basal ganglia in monkeys. Neuroscience. 2000;95:127–40. doi: 10.1016/s0306-4522(99)00409-1. [DOI] [PubMed] [Google Scholar]
  • [69].Bernard V, Bolam JP. Subcellular and subsynaptic distribution of the NR1 subunit of the NMDA receptor in the neostriatum and globus pallidus of the rat: Co-localization at synapses with the GluR2/3 subunit of the AMPA receptor. Eur J Neurosci. 1998;10:3721–36. doi: 10.1046/j.1460-9568.1998.00380.x. [DOI] [PubMed] [Google Scholar]
  • [70].Anderson BB, Chen G, Gutman DA, Ewing AG. Dopamine levels of two classes of vesicles are differentially depleted by amphetamine. Brain Res. 1998;788:294–301. doi: 10.1016/s0006-8993(98)00040-7. [DOI] [PubMed] [Google Scholar]
  • [71].Schmitz Y, Luccarelli J, Kim M, Wang M, Sulzer D. Glutamate controls growth rate and branching of dopaminergic axons. J Neurosci. 2009;29:11973–81. doi: 10.1523/JNEUROSCI.2927-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [72].Avshalumov MV, Chen BT, Marshall SP, Pena DM, Rice ME. Glutamate-dependent inhibition of dopamine release in striatum is mediated by a new diffusible messenger, H2O2. J Neurosci. 2003;23:2744–50. doi: 10.1523/JNEUROSCI.23-07-02744.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [73].Kulagina NV, Zigmond MJ, Michael AC. Glutamate regulates the spontaneous and evoked release of dopamine in the rat striatum. Neuroscience. 2001;102:121–8. doi: 10.1016/s0306-4522(00)00480-2. [DOI] [PubMed] [Google Scholar]
  • [74].Wu Y, Pearl SM, Zigmond MJ, Michael AC. Inhibitory glutamatergic regulation of evoked dopamine release in striatum. Neuroscience. 2000;96:65–72. doi: 10.1016/s0306-4522(99)00539-4. [DOI] [PubMed] [Google Scholar]
  • [75].Iravani MM, Kruk ZL. Real-time effects of N-methyl-D-aspartic acid on dopamine release in slices of rat caudate putamen: A study using fast cyclic voltammetry. J Neurochem. 1996;66:1076–85. doi: 10.1046/j.1471-4159.1996.66031076.x. [DOI] [PubMed] [Google Scholar]
  • [76].Moghaddam B, Gruen RJ, Roth RH, Bunney BS, Adams RN. Effect of L-glutamate on the release of striatal dopamine: In vivo dialysis and electrochemical studies. Brain Res. 1990;518:55–60. doi: 10.1016/0006-8993(90)90953-9. [DOI] [PubMed] [Google Scholar]
  • [77].Svensson L, Zhang J, Johannessen K, Engel JA. Effect of local infusion of glutamate analogues into the nucleus accumbens of rats: An electrochemical and behavioural study. Brain Res. 1994;643:155–61. doi: 10.1016/0006-8993(94)90021-3. [DOI] [PubMed] [Google Scholar]
  • [78].Avshalumov MV, Patel JC, Rice ME. AMPA receptor-dependent H2O2 generation in striatal medium spiny neurons but not dopamine axons: One source of a retrograde signal that can inhibit dopamine release. J Neurophysiol. 2008;100:1590–601. doi: 10.1152/jn.90548.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [79].Paquet M, Smith Y. Group I metabotropic glutamate receptors in the monkey striatum: Subsynaptic association with glutamatergic and dopaminergic afferents. J Neurosci. 2003;23:7659–69. doi: 10.1523/JNEUROSCI.23-20-07659.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [80].Zhang H, Sulzer D. Glutamate spillover in the striatum depresses dopaminergic transmission by activating group I metabotropic glutamate receptors. J Neurosci. 2003;23:10585–92. doi: 10.1523/JNEUROSCI.23-33-10585.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [81].Smolders I, De Klippel N, Sarre S, Ebinger G, Michotte Y. Tonic GABA-ergic modulation of striatal dopamine release studied by in vivo microdialysis in the freely moving rat. Eur J Pharmacol. 1995;284:83–91. doi: 10.1016/0014-2999(95)00369-v. [DOI] [PubMed] [Google Scholar]
  • [82].Ronken E, Mulder AH, Schoffelmeer AN. Interacting presynaptic kappa-opioid and GABAA receptors modulate dopamine release from rat striatal synaptosomes. J Neurochem. 1993;61:1634–9. doi: 10.1111/j.1471-4159.1993.tb09797.x. [DOI] [PubMed] [Google Scholar]
  • [83].De Klippel N, Sarre S, Ebinger G, Michotte Y. Effect of M1- and M2-muscarinic drugs on striatal dopamine release and metabolism: An in vivo microdialysis study comparing normal and 6-hydroxydopamine-lesioned rats. Brain Res. 1993;630:57–64. doi: 10.1016/0006-8993(93)90642-z. [DOI] [PubMed] [Google Scholar]
  • [84].Kemel ML, Desban M, Glowinski J, Gauchy C. Distinct presynaptic control of dopamine release in striosomal and matrix areas of the cat caudate nucleus. Proc Natl Acad Sci U S A. 1989;86:9006–10. doi: 10.1073/pnas.86.22.9006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [85].Lehmann J, Langer SZ. Muscarinic receptors on dopamine terminals in the cat caudate nucleus: Neuromodulation of [3H]dopamine release in vitro by endogenous acetylcholine. Brain Res. 1982;248:61–9. doi: 10.1016/0006-8993(82)91147-7. [DOI] [PubMed] [Google Scholar]
  • [86].Raiteri M, Leardi R, Marchi M. Heterogeneity of presynaptic muscarinic receptors regulating neurotransmitter release in the rat brain. J Pharmacol Exp Ther. 1984;228:209–14. [PubMed] [Google Scholar]
  • [87].Schoffelmeer AN, Van Vliet BJ, Wardeh G, Mulder AH. Muscarine receptor-mediated modulation of [3H]dopamine and [14C]acetylcholine release from rat neostriatal slices: Selective antagonism by gallamine but not pirenzepine. Eur J Pharmacol. 1986;128:291–4. doi: 10.1016/0014-2999(86)90781-8. [DOI] [PubMed] [Google Scholar]
  • [88].Smolders I, Bogaert L, Ebinger G, Michotte Y. Muscarinic modulation of striatal dopamine, glutamate, and GABA release, as measured with in vivo microdialysis. J Neurochem. 1997;68:1942–8. doi: 10.1046/j.1471-4159.1997.68051942.x. [DOI] [PubMed] [Google Scholar]
  • [89].Xu M, Mizobe F, Yamamoto T, Kato T. Differential effects of M1- and M2-muscarinic drugs on striatal dopamine release and metabolism in freely moving rats. Brain Res. 1989;495:232–42. doi: 10.1016/0006-8993(89)90217-5. [DOI] [PubMed] [Google Scholar]
  • [90].Wess J. Molecular biology of muscarinic acetylcholine receptors. Crit Rev Neurobiol. 1996;10:69–99. doi: 10.1615/critrevneurobiol.v10.i1.40. [DOI] [PubMed] [Google Scholar]
  • [91].Kudernatsch M, Sutor B. Cholinergic modulation of dopamine overflow in the rat neostriatum: A fast cyclic voltammetric study in vitro. Neurosci Lett. 1994;181:107–12. doi: 10.1016/0304-3940(94)90571-1. [DOI] [PubMed] [Google Scholar]
  • [92].Dewey SL, Smith GS, Logan J, Brodie JD, Simkowitz P, MacGregor RR, Fowler JS, Volkow ND, Wolf AP. Effects of central cholinergic blockade on striatal dopamine release measured with positron emission tomography in normal human subjects. Proc Natl Acad Sci U S A. 1993;90:11816–20. doi: 10.1073/pnas.90.24.11816. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [93].Threlfell S, Clements MA, Khodai T, Pienaar IS, Exley R, Wess J, Cragg SJ. Striatal muscarinic receptors promote activity dependence of dopamine transmission via distinct receptor subtypes on cholinergic interneurons in ventral versus dorsal striatum. J Neurosci. 2010;30:3398–408. doi: 10.1523/JNEUROSCI.5620-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [94].Threlfell S, Cragg SJ. Dopamine signaling in dorsal versus ventral striatum: The dynamic role of cholinergic interneurons. Front Syst Neurosci. 2010;5:11. doi: 10.3389/fnsys.2011.00011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [95].Vilaro MT, Palacios JM, Mengod G. Localization of M5 muscarinic receptor mRNA in rat brain examined by in situ hybridization histochemistry. Neurosci Lett. 1990;114:154–9. doi: 10.1016/0304-3940(90)90064-g. [DOI] [PubMed] [Google Scholar]
  • [96].Weiner DM, Levey AI, Brann MR. Expression of muscarinic acetylcholine and dopamine receptor mRNAs in rat basal ganglia. Proc Natl Acad Sci U S A. 1990;87:7050–4. doi: 10.1073/pnas.87.18.7050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [97].Martire M, D’Amico M, Panza E, Miceli F, Viggiano D, Lavergata F, Iannotti FA, Barrese V, Preziosi P, Annunziato L, Taglialatela M. Involvement of KCNQ2 subunits in [3H]dopamine release triggered by depolarization and pre-synaptic muscarinic receptor activation from rat striatal synaptosomes. J Neurochem. 2007;102:179–93. doi: 10.1111/j.1471-4159.2007.04562.x. [DOI] [PubMed] [Google Scholar]
  • [98].Bendor J, Lizardi-Ortiz JE, Westphalen RI, Brandstetter M, Hemmings HC, Jr., Sulzer D, Flajolet M, Greengard P. AGAP1/AP-3-dependent endocytic recycling of M5 muscarinic receptors promotes dopamine release. EMBO J. 2010;29:2813–26. doi: 10.1038/emboj.2010.154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [99].Zoli M, Moretti M, Zanardi A, McIntosh JM, Clementi F, Gotti C. Identification of the nicotinic receptor subtypes expressed on dopaminergic terminals in the rat striatum. J Neurosci. 2002;22:8785–9. doi: 10.1523/JNEUROSCI.22-20-08785.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [100].Salminen O, Murphy KL, McIntosh JM, Drago J, Marks MJ, Collins AC, Grady SR. Subunit composition and pharmacology of two classes of striatal presynaptic nicotinic acetylcholine receptors mediating dopamine release in mice. Mol Pharmacol. 2004;65:1526–35. doi: 10.1124/mol.65.6.1526. [DOI] [PubMed] [Google Scholar]
  • [101].Pontieri FE, Tanda G, Orzi F, Di Chiara G. Effects of nicotine on the nucleus accumbens and similarity to those of addictive drugs. Nature. 1996;382:255–7. doi: 10.1038/382255a0. [DOI] [PubMed] [Google Scholar]
  • [102].Cheer JF, Wassum KM, Sombers LA, Heien ML, Ariansen JL, Aragona BJ, Phillips PE, Wightman RM. Phasic dopamine release evoked by abused substances requires cannabinoid receptor activation. J Neurosci. 2007;27:791–5. doi: 10.1523/JNEUROSCI.4152-06.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [103].Zhou FM, Liang Y, Dani JA. Endogenous nicotinic cholinergic activity regulates dopamine release in the striatum. Nat Neurosci. 2001;4:1224–9. doi: 10.1038/nn769. [DOI] [PubMed] [Google Scholar]
  • [104].Rice ME, Cragg SJ. Nicotine amplifies reward-related dopamine signals in striatum. Nat Neurosci. 2004;7:583–4. doi: 10.1038/nn1244. [DOI] [PubMed] [Google Scholar]
  • [105].Zhang H, Sulzer D. Frequency-dependent modulation of dopamine release by nicotine. Nat Neurosci. 2004;7:581–2. doi: 10.1038/nn1243. [DOI] [PubMed] [Google Scholar]
  • [106].Trovero F, Herve D, Desban M, Glowinski J, Tassin JP. Striatal opiate mu-receptors are not located on dopamine nerve endings in the rat. Neuroscience. 1990;39:313–21. doi: 10.1016/0306-4522(90)90270-e. [DOI] [PubMed] [Google Scholar]
  • [107].Schlosser B, Kudernatsch MB, Sutor B, ten Bruggencate G. Delta, mu and kappa opioid receptor agonists inhibit dopamine overflow in rat neostriatal slices. Neurosci Lett. 1995;191:126–30. doi: 10.1016/0304-3940(94)11552-3. [DOI] [PubMed] [Google Scholar]
  • [108].Britt JP, McGehee DS. Presynaptic opioid and nicotinic receptor modulation of dopamine overflow in the nucleus accumbens. J Neurosci. 2008;28:1672–81. doi: 10.1523/JNEUROSCI.4275-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [109].Ferre S, Fredholm BB, Morelli M, Popoli P, Fuxe K. Adenosine-dopamine receptor-receptor interactions as an integrative mechanism in the basal ganglia. Trends Neurosci. 1997;20:482–7. doi: 10.1016/s0166-2236(97)01096-5. [DOI] [PubMed] [Google Scholar]
  • [110].Jin S, Johansson B, Fredholm BB. Effects of adenosine A1 and A2 receptor activation on electrically evoked dopamine and acetylcholine release from rat striatal slices. J Pharmacol Exp Ther. 1993;267:801–8. [PubMed] [Google Scholar]
  • [111].Quarta D, Borycz J, Solinas M, Patkar K, Hockemeyer J, Ciruela F, Lluis C, Franco R, Woods AS, Goldberg SR, Ferre S. Adenosine receptor-mediated modulation of dopamine release in the nucleus accumbens depends on glutamate neurotransmission and N-methyl-D-aspartate receptor stimulation. J Neurochem. 2004;91:873–80. doi: 10.1111/j.1471-4159.2004.02761.x. [DOI] [PubMed] [Google Scholar]
  • [112].Cechova S, Elsobky AM, Venton BJ. A1 receptors self-regulate adenosine release in the striatum: Evidence of autoreceptor characteristics. Neuroscience. 2010;171:1006–15. doi: 10.1016/j.neuroscience.2010.09.063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [113].Golembiowska K, Zylewska A. Agonists of A1 and A2A adenosine receptors attenuate methamphetamine-induced overflow of dopamine in rat striatum. Brain Res. 1998;806:202–9. doi: 10.1016/s0006-8993(98)00743-4. [DOI] [PubMed] [Google Scholar]
  • [114].O’Neill C, Nolan BJ, Macari A, O’Boyle KM, O’Connor JJ. Adenosine A1 receptor-mediated inhibition of dopamine release from rat striatal slices is modulated by D1 dopamine receptors. Eur J Neurosci. 2007;26:3421–8. doi: 10.1111/j.1460-9568.2007.05953.x. [DOI] [PubMed] [Google Scholar]
  • [115].Glass M, Faull RL, Dragunow M. Localisation of the adenosine uptake site in the human brain: A comparison with the distribution of adenosine A1 receptors. Brain Res. 1996;710:79–91. doi: 10.1016/0006-8993(95)01318-0. [DOI] [PubMed] [Google Scholar]
  • [116].Mahan LC, McVittie LD, Smyk-Randall EM, Nakata H, Monsma FJ, Jr., Gerfen CR, Sibley DR. Cloning and expression of an A1 adenosine receptor from rat brain. Mol Pharmacol. 1991;40:1–7. [PubMed] [Google Scholar]
  • [117].Rivkees SA, Price SL, Zhou FC. Immunohistochemical detection of A1 adenosine receptors in rat brain with emphasis on localization in the hippocampal formation, cerebral cortex, cerebellum, and basal ganglia. Brain Res. 1995;677:193–203. doi: 10.1016/0006-8993(95)00062-u. [DOI] [PubMed] [Google Scholar]
  • [118].Alexander SP, Reddington M. The cellular localization of adenosine receptors in rat neostriatum. Neuroscience. 1989;28:645–51. doi: 10.1016/0306-4522(89)90011-0. [DOI] [PubMed] [Google Scholar]
  • [119].Hettinger BD, Lee A, Linden J, Rosin DL. Ultrastructural localization of adenosine A2A receptors suggests multiple cellular sites for modulation of GABAergic neurons in rat striatum. J Comp Neurol. 2001;431:331–46. doi: 10.1002/1096-9861(20010312)431:3<331::aid-cne1074>3.0.co;2-w. [DOI] [PubMed] [Google Scholar]
  • [120].Rosin DL, Hettinger BD, Lee A, Linden J. Anatomy of adenosine A2A receptors in brain: Morphological substrates for integration of striatal function. Neurology. 2003;61:S12–8. doi: 10.1212/01.wnl.0000095205.33940.99. [DOI] [PubMed] [Google Scholar]
  • [121].Mailleux P, Vanderhaeghen JJ. Localization of cannabinoid receptor in the human developing and adult basal ganglia. Higher levels in the striatonigral neurons. Neurosci Lett. 1992;148:173–6. doi: 10.1016/0304-3940(92)90832-r. [DOI] [PubMed] [Google Scholar]
  • [122].Mailleux P, Vanderhaeghen JJ. Distribution of neuronal cannabinoid receptor in the adult rat brain: A comparative receptor binding radioautography and in situ hybridization histochemistry. Neuroscience. 1992;48:655–68. doi: 10.1016/0306-4522(92)90409-u. [DOI] [PubMed] [Google Scholar]
  • [123].Matsuda LA. Molecular aspects of cannabinoid receptors. Crit Rev Neurobiol. 1997;11:143–66. doi: 10.1615/critrevneurobiol.v11.i2-3.30. [DOI] [PubMed] [Google Scholar]
  • [124].Schlicker E, Kathmann M. Modulation of transmitter release via presynaptic cannabinoid receptors. Trends Pharmacol Sci. 2001;22:565–72. doi: 10.1016/s0165-6147(00)01805-8. [DOI] [PubMed] [Google Scholar]
  • [125].Szabo B, Muller T, Koch H. Effects of cannabinoids on dopamine release in the corpus striatum and the nucleus accumbens in vitro. J Neurochem. 1999;73:1084–9. doi: 10.1046/j.1471-4159.1999.0731084.x. [DOI] [PubMed] [Google Scholar]
  • [126].Sidlo Z, Reggio PH, Rice ME. Inhibition of striatal dopamine release by CB1 receptor activation requires nonsynaptic communication involving GABA, H2O2, and KATP channels. Neurochem Int. 2008;52:80–8. doi: 10.1016/j.neuint.2007.07.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [127].Fusco FR, Martorana A, Giampa C, De March Z, Farini D, D’Angelo V, Sancesario G, Bernardi G. Immunolocalization of CB1 receptor in rat striatal neurons: A confocal microscopy study. Synapse. 2004;53:159–67. doi: 10.1002/syn.20047. [DOI] [PubMed] [Google Scholar]
  • [128].Matsuda LA, Bonner TI, Lolait SJ. Localization of cannabinoid receptor mRNA in rat brain. J Comp Neurol. 1993;327:535–50. doi: 10.1002/cne.903270406. [DOI] [PubMed] [Google Scholar]
  • [129].Romero J, Garcia-Palomero E, Castro JG, Garcia-Gil L, Ramos JA, Fernandez-Ruiz JJ. Effects of chronic exposure to delta9-tetrahydrocannabinol on cannabinoid receptor binding and mRNA levels in several rat brain regions. Brain Res Mol Brain Res. 1997;46:100–8. doi: 10.1016/s0169-328x(96)00277-x. [DOI] [PubMed] [Google Scholar]
  • [130].Julian MD, Martin AB, Cuellar B, Rodriguez De Fonseca F, Navarro M, Moratalla R, Garcia-Segura LM. Neuroanatomical relationship between type 1 cannabinoid receptors and dopaminergic systems in the rat basal ganglia. Neuroscience. 2003;119:309–18. doi: 10.1016/s0306-4522(03)00070-8. [DOI] [PubMed] [Google Scholar]
  • [131].Lau T, Schloss P. The cannabinoid CB1 receptor is expressed on serotonergic and dopaminergic neurons. Eur J Pharmacol. 2008;578:137–41. doi: 10.1016/j.ejphar.2007.09.022. [DOI] [PubMed] [Google Scholar]
  • [132].Lupica CR, Riegel AC. Endocannabinoid release from midbrain dopamine neurons: A potential substrate for cannabinoid receptor antagonist treatment of addiction. Neuropharmacology. 2005;48:1105–16. doi: 10.1016/j.neuropharm.2005.03.016. [DOI] [PubMed] [Google Scholar]
  • [133].Matyas F, Urban GM, Watanabe M, Mackie K, Zimmer A, Freund TF, Katona I. Identification of the sites of 2-arachidonoylglycerol synthesis and action imply retrograde endocannabinoid signaling at both GABAergic and glutamatergic synapses in the ventral tegmental area. Neuropharmacology. 2008;54:95–107. doi: 10.1016/j.neuropharm.2007.05.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [134].Riegel AC, Lupica CR. Independent presynaptic and postsynaptic mechanisms regulate endocannabinoid signaling at multiple synapses in the ventral tegmental area. J Neurosci. 2004;24:11070–8. doi: 10.1523/JNEUROSCI.3695-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [135].Fernandez-Ruiz J, Gonzales S. Cannabinoid control of motor function at the basal ganglia. Handb Exp Pharmacol. 2005:479–507. doi: 10.1007/3-540-26573-2_16. [DOI] [PubMed] [Google Scholar]
  • [136].O’Neill C, Evers-Donnelly A, Nicholson D, O’Boyle KM, O’Connor JJ. D2 receptor-mediated inhibition of dopamine release in the rat striatum in vitro is modulated by CB1 receptors: Studies using fast cyclic voltammetry. J Neurochem. 2009;108:545–51. doi: 10.1111/j.1471-4159.2008.05782.x. [DOI] [PubMed] [Google Scholar]
  • [137].Fuxe K, Marcellino D, Rivera A, Diaz-Cabiale Z, Filip M, Gago B, Roberts DC, Langel U, Genedani S, Ferraro L, de la Calle A, Narvaez J, Tanganelli S, Woods A, Agnati LF. Receptor-receptor interactions within receptor mosaics. Impact on neuropsychopharmacology. Brain Res Rev. 2008;58:415–52. doi: 10.1016/j.brainresrev.2007.11.007. [DOI] [PubMed] [Google Scholar]
  • [138].Grace AA, Bunney BS. The control of firing pattern in nigral dopamine neurons : Burst firing. J Neurosci. 1984;4:2877–90. doi: 10.1523/JNEUROSCI.04-11-02877.1984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [139].Hyland BI, Reynolds JN, Hay J, Perk CG, Miller R. Firing modes of midbrain dopamine cells in the freely moving rat. Neuroscience. 2002;114:475–92. doi: 10.1016/s0306-4522(02)00267-1. [DOI] [PubMed] [Google Scholar]
  • [140].Schultz W, Apicella P, Ljungberg T. Responses of monkey dopamine neurons to reward and conditioned stimuli during successive steps of learning a delayed response task. J Neurosci. 1993;13:900–13. doi: 10.1523/JNEUROSCI.13-03-00900.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [141].Mirenowicz J, Schultz W. Preferential activation of midbrain dopamine neurons by appetitive rather than aversive stimuli. Nature. 1996;379:449–51. doi: 10.1038/379449a0. [DOI] [PubMed] [Google Scholar]
  • [142].Gonon F. Nonlinear relationship between impulse flow and dopamine release by rat midbrain dopaminergic neurons as studied by in vivo electrochemistry. Neuroscience. 1988;24:19–28. doi: 10.1016/0306-4522(88)90307-7. [DOI] [PubMed] [Google Scholar]
  • [143].Zhang T, Zhang L, Liang Y, Siapas AG, Zhou FM, Dani JA. Dopamine signaling differences in the nucleus accumbens and dorsal striatum exploited by nicotine. J Neurosci. 2009;29:4035–43. doi: 10.1523/JNEUROSCI.0261-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [144].Grenhoff J, Aston-Jones G, Svensson TH. Nicotinic effects on the firing pattern of midbrain dopamine neurons. Acta Physiol Scand. 1986;128:351–8. doi: 10.1111/j.1748-1716.1986.tb07988.x. [DOI] [PubMed] [Google Scholar]
  • [145].Ding J, Guzman JN, Tkatch T, Chen S, Goldberg JA, Ebert PJ, Levitt P, Wilson CJ, Hamm HE, Surmeier DJ. RGS4-dependent attenuation of M4 autoreceptor function in striatal cholinergic interneurons following dopamine depletion. Nat Neurosci. 2006;9:832–42. doi: 10.1038/nn1700. [DOI] [PubMed] [Google Scholar]
  • [146].Zhang H, Sulzer D. Possible filtering property of heteroreceptors on nigrostriatal terminals on evoked dopamine release in striatal slice. 2004 Neurosci Abst.
  • [147].Haber SN, Fudge JL, McFarland NR. Striatonigrostriatal pathways in primates form an ascending spiral from the shell to the dorsolateral striatum. J Neurosci. 2000;20:2369–82. doi: 10.1523/JNEUROSCI.20-06-02369.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [148].Mani M, Ryan TA. Live imaging of synaptic vesicle release and retrieval in dopaminergic neurons. Front Neural Circuits. 2009;3:3. doi: 10.3389/neuro.04.003.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [149].Daniel JA, Galbraith S, Iacovitti L, Abdipranoto A, Vissel B. Functional heterogeneity at dopamine release sites. J Neurosci. 2009;29:14670–80. doi: 10.1523/JNEUROSCI.1349-09.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [150].Onoa B, Li H, Gagnon-Bartsch JA, Elias LA, Edwards RH. Vesicular monoamine and glutamate transporters select distinct synaptic vesicle recycling pathways. J Neurosci. 2010;30:7917–27. doi: 10.1523/JNEUROSCI.5298-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [151].Gubernator NG, Zhang H, Staal RG, Mosharov EV, Pereira DB, Yue M, Balsanek V, Vadola PA, Mukherjee B, Edwards RH, Sulzer D, Sames D. Fluorescent false neurotransmitters visualize dopamine release from individual presynaptic terminals. Science. 2009;324:1441–4. doi: 10.1126/science.1172278. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES