Nothing Special   »   [go: up one dir, main page]

Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Oct 11.
Published in final edited form as: Neuron. 2017 Oct 11;96(2):267–284. doi: 10.1016/j.neuron.2017.09.019

Striatal local circuitry: a new framework for lateral inhibition

Dennis A Burke 1,2, Horacio G Rotstein 3,4, Veronica A Alvarez 1
PMCID: PMC5649445  NIHMSID: NIHMS906571  PMID: 29024654

Abstract

This Perspective will examine the organization of intrastriatal circuitry, review recent findings in this area, and discuss how the pattern of connectivity between striatal neurons might give rise to the behaviorally observed synergism between the direct/indirect pathway neurons. The emphasis of this Perspective is on the underappreciated role of lateral inhibition between striatal projection cells in controlling neuronal firing and shaping the output of this circuit. We review some classic studies in combination with more recent anatomical and functional findings to lay out a framework for an updated model of the local intra-striatal circuitry and its contribution to the formation of functional units of processing with the striatum and the integration and filtering of inputs to generate motor patterns and learned behaviors.

1. Introduction

The striatum, as the input nucleus to the basal ganglia, is functionally implicated in motor planning, action selection, reward guided learning, and other motivated behaviors and cognitive processes. Not surprisingly given these broad functions, striatal dysfunction is involved in a variety of diverse neurological and psychiatric diseases including, but not limited to, Parkinson’s disease, Huntington’s disease, schizophrenia, depression, and addiction. By receiving glutamatergic inputs from cortical, thalamic, and limbic regions combined with dopaminergic input from the midbrain, the striatum acts as an integrative hub that assists in the selection of appropriate behaviors through its outputs to downstream basal ganglia structures (Redgrave et al., 1999). This integrative function of the striatum that is critical for the processing of input information is thought to happen at the level of the local intrastriatal circuitry. However, it is still unclear exactly how local striatal circuitry transforms information flow and ultimately affects behavioral output. This Perspective summarizes the most recent finding on the properties and organization of the local circuitry in the striatum with special emphasis on the connectivity between the striatal projection neurons. This component of the circuitry has been largely ignored in the past because the connectivity is considered relatively sparse. However, recent studies suggest that this synaptic connection between projections neurons, which is mediated by their local axon collaterals within the striatum, can exert potent lateral inhibition to the striatal circuitry, which can profoundly regulate the excitability of other projection neurons. Focusing on this underappreciated component of the striatal circuitry, this Perspective further speculates that these connections are functionally organized to aid in the information processing that shapes striatal output. In doing so, we provide a conceptual model in which to test how particular patterns of connectivity between projection neurons can affect the firing. This model provides a framework for understanding the role of striatal lateral inhibition with the hope of inspiring future connectomics research into the organization of the intrastriatal connectivity and how it ultimately assists in coordinating behavior and learning both in health and in disease.

The layout of the perspective is as follows: Section 2 provides a general overview of the different cell-types within the striatum as well as its functional role and inputs. Section 3 focuses on the local connectivity within the striatum and describes the wide variety of synaptic connections among striatal neurons. Section 4 concentrates on the synapses formed between projection neurons and offers a historical view from their initial discovery to our current findings. It discusses possible functional roles for this lateral inhibition and presents evidence for the existence of striatal ensembles. Finally, Section 5 builds on this empirical data and lays out a theoretical framework for how the synapses between striatal projection neurons might be organized to aid in formation of ensembles or functional units. Under this hypothetical model, the pattern of connectivity of the lateral inhibition determines the degree of co-activation of striatal projection neurons and, as such, shapes the striatal output. It also demonstrates that different patterns of intrastriatal connectivity will affect the degree of neuronal co-activation, determining which units of cells fire together, ultimately contributing to each ensemble. Further, it opens the door to propositions involving modulators of lateral inhibition in dynamically reshaping the ensembles, which might underlie forms of striatal learning. This last section aims to stimulate discussion and generate testable hypotheses that drive research on the topic of striatal organization.

2. General Overview

Heterogeneity of striatal subregions and cell-types

Striatal subregions

The rodent striatum, is traditionally subdivided into two regions: dorsal and ventral, which contain subdivisions, based on their inputs and/or immunohistochemical markers. The dorsal striatum, or neostriatum, can be divided into the dorsal lateral striatum (DLS) and the dorsal medial striatum (DMS), based on glutamatergic inputs. The ventral striatum includes the olfactory tubercule and the nucleus accumbens (NAc), which can be further subdivided into “core” and “shell” subregions based on inputs/ and immunohistochemical markers.

Another dimension along which the striatum can be divided is the “patch” (striosome) and “matrix” compartments. First identified by the expression patterns of proteins such as mu-opioid receptors and acetylcholinesterase, there are now over 60 genes known to be differentially expressed in patch vs. matrix compartments (Crittenden and Graybiel, 2011). Though patches only represent a small proportion of the striatum by volume, many of the differentially expressed genes are neuromodulators and their receptors suggesting that the patch and matrix have distinct pharmacological properties (Banghart et al., 2015; Brimblecombe and Cragg, 2016). Afferents to patch and matrix are also thought to arise from different sources with patches receiving limbic inputs and matrix neurons receiving sensorimotor inputs (Gerfen, 1984). However recent viral tracing methods have cast doubt on this strict segregation (Smith et al., 2016).

The striatal subregions are also functionally segregated. In general terms, the dorsal striatum is involved in motor planning, action selection, and stimulus-response habit learning, while the NAc plays a role in processing motivated behavior and reward related learning. However, there is much overlap in the known roles of the dorsal and ventral striatum in controlling motor and motivated behaviors and learning (Isomura et al., 2013). Technical limitations make it difficult to distinguish motivation from motor ability and to discriminate learning and memory from performance. Also the overlapping functions between these striatal subregions is hypothesized to be embedded within the circuit connectivity in the form of “ascending spiral loops” that allow for information flow from the ventral to the dorsal striatum via the circuitry within the basal ganglia (Haber et al., 2000).

Spiny projection neurons

In addition to potential functional overlaps, the dorsal and ventral striatum share multiple features including their main cell-type composition. The principal projection neurons in the striatum are the GABAergic spiny projection neurons (SPNs; also referred to as “medium spiny neurons”), which comprise over 95% of the neurons in this region (Graveland and Difiglia, 1985) and are the only projection neurons of the striatum. SPNs are classically divided in two subclasses based on their projection targets and expression of dopamine receptors (Gerfen et al., 1990). Direct pathway SPNs (dSPNs) represent approximately half of the SPN population, mainly express Gs/olf coupled dopamine D1 receptors, and send dense projections to the substantia nigra pars reticulata and internal globus pallidus/entopeduncular nucleus through the “direct” pathway of the basal ganglia. Indirect pathway SPNs (iSPNs) mainly express Gi/o coupled dopamine D2 receptors and project through the “indirect” pathway to the external globus pallidus (GP) and the ventral pallidum (VP). It is clear now that this dichotomy is an over-simplification of the existing subtypes of SPNs and that the segregation is imperfect. For example, a subset of SPNs co-express both D1 and D2 receptors in the adult striatum (~2–5% dorsal striatum,~6–7% in the NAc core, ~12–15% in the NAc shell) that are morphologically distinct from other SPNs and have a poorly understood function (Bertran-Gonzalez et al., 2008; Gagnon et al., 2017). Furthermore, a recent study shows large cell-to-cell variability in the pattern of overall gene expression within each subclass of SPN (Gokce et al., 2015). Another consideration is that the projections are not fully segregated in the sense that D1R expressing SPNs in the dorsal striatum also send collateral axon projections to the GP (Cazorla et al., 2014; Kawaguchi et al., 1990) and those in the NAc send numerous projections the VP (Kupchik et al., 2015). While the classic segregation of SPNs into direct- and indirect-pathway neurons is incomplete and oversimplified, it has proven extremely useful for advancing our understanding of this brain region and the function of the basal ganglia as a whole. With these caveats in mind, keeping with field nomenclature, throughout this manuscript we will use dSPN to refer to SPNs expressing D1 receptors and iSPNs to refer to those expressing D2 receptors.

Interneurons

The remaining 5% of neurons in the striatum are either cholinergic or GABAergic interneurons. Cholinergic interneurons (CIN) are large, tonically active neurons that make up about 1% of the cells in the striatum and, as their name implies, release acetylcholine (Zhou et al., 2002). These cells have been classically referred as TANs for “tonically active neurons” (but see below that they are not the only type of spontaneously active interneurons). Despite their relatively low abundance, CINs extend wide axonal arborizations throughout the region profoundly influencing striatal function through a variety of muscarinic and nicotinic acetylcholine receptors (reviewed in Gonzales and Smith, 2015). Recently, CINs were shown to also co-release glutamate (Guzman et al., 2011; Higley et al., 2011; Nelson et al., 2014a). Further experiments using paired recordings will be useful in confirming this finding.

The GABAergic interneurons of the striatum can be subdivided in three classical groups and at least three new additional reported subpopulations based on a combination of molecular, physiological, and pharmacological properties (Figure 1).

Figure 1. Diversity of cell-types of striatal interneurons.

Figure 1

The plot represents the estimated proportions of each subtype of striatal interneuron identified based on their expression of neurotransmitter (inner donut), other molecular markers (middle donut), and electrophysiological properties (outer donut). Abbreviations: ACh, acetylcholine; ChAT, choline-acetyl transferase; TAN, tonically active neurons; TH, tyrosine hydroxlase; CR, calretinin; SOM, somatostatin; NOS, nitric oxide synthase; NPY, neuropeptide Y; LTS, low threshold spiking; NGF, neurogliaform; 5HT3R, serotonin type-3 receptor; FA, fast-adapting; PV, parvalbumin; FS, fast-spiking. *, SOM/NOS LTS neurons are also tonically active, although not classically referred as TANs.

The three classically described GABAergic interneurons are the fast-spiking (FSI), the low threshold spiking (LTSI), and the calretinin expression (CR) interneuron. FSIs, the best studied subtype of the GABAergic interneurons, express parvalbumin (PV), can fire action potentials at high frequencies, and mediates strong feedforward inhibition onto SPNs. LTSIs are tonically active, and although striatal “TANs” is used in the literature to refer to presumed cholinergic interneurons, this is an important consideration for in vivo electrophysiology experiments. LTSIs exhibit low threshold Ca2+ spikes and can be identified based on the expression of somatostatin (SOM), nitric oxide synthase (NOS), and neuropeptide Y (NPY). Not much is known with regards to the physiology or synaptic function of CR interneurons likely due to their low abundance in mouse compared to primate striatum (Petryszyn et al., 2014).

Recently, BAC transgenic mice have facilitated further classification of novel GABAergic striatal interneuron subtypes based on molecular, physiological, and pharmacological properties (Figure 1). For example, a recently characterized population of striatal interneurons expresses tyrosine hydroxlase (TH). TH interneurons all release GABA, and this group does not appear to overlap with other known subtypes of striatal interneurons. Characterization of these neurons has revealed four distinct subtypes based on electrophysiological properties, termed Type I–IV (Ibáñez-Sandoval et al., 2010). Interestingly, though tyrosine hydroxylase is the rate limiting enzyme in dopamine synthesis and is used as a marker for dopamine neurons, TH interneurons do not appear to release dopamine (Xenias et al., 2015).

Two recently identified novel interneuron subtypes are the NPY-neurogliaform (NPY-NGF) interneuron and the fast-adapting interneuron (FAI). NPY-NGF interneurons express NPY, but not SOM and NOS, as LTSIs do, and share similarities with cortical NPY expressing neurogliaform cells (Ibáñez-Sandoval et al., 2011). NPY-NGF neurons are excited by acetylcholine through nicotinic receptors and differ substantially from SOM/NOS/NPY expressing LTS neurons in both electrophysiological and anatomical characteristics, as well as in the synaptic response they elicit in SPNs.

FAIs have been identified by expression of the serotonin 5-HT3 receptor (5HT3R) and given their name due to a distinct spike-frequency adaptation during repetitive firing (Faust et al., 2015). FAIs express nicotinic acetylcholine receptors. The 5HT3R is not a perfect marker for the FAI, however, as in many interneurons in the striatum the 5HT3R is coexpressed with other interneuron markers, in particular PV. Neurons that express both PV and 5HT3R appear to have FSI electrophysiological characteristics similar to those neurons that express PV alone (Muñoz-Manchado et al., 2016). For reviews of striatal interneurons, see (Tepper et al., 2010; Tepper and Koós, 2017).

Functional overview

Classical models of basal ganglia function posit that activation of dSPNs disinhibits principal thalamic neurons to facilitate behavior, while activation of iSPNs ultimately inhibits thalamic neurons to suppress motor and reward-based behaviors. Dysregulation in the balance of dSPN and iSPN activity is thought to be the underlying cause of several basal ganglia movement disorders (Albin et al., 1989; DeLong, 1990). This model is supported by experiments using cell specific genetic perturbation, cell ablation, or optogenetic activation (Bateup et al., 2010; Bock et al., 2013; Durieux et al., 2009; Kravitz et al., 2012, 2010; Lobo et al., 2010; Tai et al., 2012).

Glutamate inputs drive SPN firing and, without these excitatory inputs, SPNs are quiescent and their membrane potential is very hyperpolarized. During a desirable action, dopamine is thought to modulate the neuronal excitability to acutely raise the signal-to-noise ratio through activation of Gs/olf coupled D1 receptors (D1Rs) on dSPNs and the Gi/o coupled D2 receptors (D2Rs) on iSPNs. Also, by triggering synaptic plasticity and inducing changes in gene expression, dopamine biases future behavior toward repeating desirable actions, mediating processes that underlie learning (Alexander and Crutcher, 1990; Cohen and Frank, 2009; Gerfen and Surmeier, 2011; Surmeier et al., 2014; Tritsch and Sabatini, 2012).

There is accumulating evidence for a more nuanced view of the interaction between the direct and indirect pathways in driving behavior. Recent work in the dorsal striatum using calcium imaging with fiber photometry (Cui et al., 2014; Tecuapetla et al., 2014), deep brain calcium imaging (Barbera et al., 2016), and in vivo electrophysiology in which SPN subtypes were identified with optogenetic tagging (Jin et al., 2014) or post hoc in situ hybridization (Isomura et al., 2013), have all demonstrated that both dSPNs and iSPNs are more active when an animal is performing a task or freely moving and less active when the animal is not moving or engaged in a task (though dSPNs and iSPNs have differing activity patterns). These data imply that iSPNs and dSPNs are coactive during action initiation and that the direct and indirect pathways are both necessary to facilitate actions. Further support for this idea comes from studies showing that inhibition of either iSPNs or dSPNs can disrupt action initiation or ongoing behavior (Tecuapetla et al., 2016, 2014).

Extending this idea to the ventral striatum, optogenetic inhibition of either iSPN or dSPNs in the NAc after trial start cue in an operant food task reduced motivated responding on the task (Natsubori et al., 2017), and inhibiting synaptic transmission from either iSPNs or dSPNs prevents the acute locomotor response to cocaine and methamphetamine, a behavior known to involve the NAc (Hikida et al., 2010). These studies provide evidence that both dSPNs and iSPNs are more active during actions than at rest, and activity in both SPN subtypes is needed for driving normal motivated behavior.

In addition, fiber photometry recordings to examine the activity of dSPNs and iSPNs in the NAc during an operant food responding task show that both SPN subtypes are more active after a trial start cue and reward delivery than at rest (Natsubori et al., 2017). Furthermore photometry recordings in the NAc during a risky decision task have demonstrated that iSPN activity is highest following a risky choice that led to no reward and that this activity predicts future “safe” choices (Zalocusky et al., 2016). When recording calcium transients in iSPNs during a cocaine conditioned place preference paradigm, iSPNs show marked decreases in their activity after the animal moves into the drug paired side of the chamber, similar to their response to cocaine (Calipari et al., 2016). Three main considerations with fiber photometry experiments are: a) this technique measures calcium signals, an indirect assessment of spikes, and while it is capable to detect single spikes under the right conditions, this is almost rarely verified, b) the time course of the photometry signal is largely determined by properties of the calcium indicator, not the spiking and c) the spatial information is lost when collecting photons through a fiber and as such photometry measures bulk calcium signals that are amplified when there is simultaneous activation. Thus, photometry signals in vivo are most likely a reflection of 1–3 action potentials and the signal detection is biased toward synchronized events that can summate in time. Photometry signals might reflect the activity of ensembles rather than single neurons (Alvarez, 2016). While important to consider these factors, neither of these considerations undermines the main findings that co-activation of iSPNs and dSPNs is seen during a wide range of behaviors. Altogether these studies suggest that both the direct and indirect pathways are necessary to drive behavioral output in both the dorsal and the ventral striatum.

An important question then is how to reconcile the findings from optogenetic stimulation experiments suggesting that iSPN activation inhibits movement (Kravitz et al., 2010) with those from calcium dynamics showing iSPN activation during action initiation and behavior (Barbera et al., 2016; Cui et al., 2014; Natsubori et al., 2017). The most parsimonious explanation is that during behavior, dSPNs facilitate the desired behavior, while iSPNs act simultaneously to inhibit all other competing undesirable behaviors (see more in section 3). Indeed, this concept of the basal ganglia serving as a selection mechanism between competing behavioral programs was proposed over twenty years ago by Jonathon Mink (Mink, 2003, 1996; Mink and Thach, 1993) and Okihide Hikosaka (Hikosaka, 1998, 1991), and it has been incorporated into computational models of basal ganglia function (Collins and Frank, 2014; Frank, 2005). Consistent with this, global optogenetic stimulation of iSPNs would therefore suppress all behavioral programs and inhibit movement.

If this model were correct, precise regulation of the activity of dSPNs and iSPNs that control competing behavioral programs would be required (Tecuapetla et al., 2016), making it important to determine how striatal circuitry can give rise to this coordinated activity. In order for the model to work properly, coordinated activation of dSPNs that facilitate a desired behavior and activation of iSPNs that suppress competing behaviors should be achieved. This coordination could be dictated by the inputs to the striatum, the organization of the local intrastriatal connectivity, or a combination of both. In the next sections, we will first review inputs into the striatum and then discuss the local intrastriatal circuity.

Main inputs to the striatum

The striatum receives two main types of inputs: glutamatergic and dopaminergic. Glutamatergic projections arise from most cortical regions, as well as thalamic and limbic regions. Cortical inputs descend in a topographic fashion with neighboring regions of cortex projecting to adjacent regions of the striatum (Bolam et al., 2000; Dudman and Gerfen, 2015). The DLS receives inputs from sensorimotor cortex, while the DMS is innervated by associative regions of cortex. Both DMS and DLS receive inputs from the thalamus. The NAc is targeted by associative and limbic regions of cortex, as well as amygdala, hippocampal, and thalamic nuclei (Britt et al., 2012; Bubser and Deutch, 1998; MacAskill et al., 2012). Cortical inputs target both subtypes of SPNs, with single axons making synapses onto both dSPNs and iSPNs (Doig et al., 2010). While anatomical studies have suggested that dSPNs and iSPNs receive inputs from different types of cortical neurons, with intertelencephalic (IT) neurons more likely to synapse onto dSPNs and pyramidal tract neurons (PT) more likely to synapse onto iSPNs (Lei et al., 2004), physiological studies have shown that both types of cortical neurons synapse onto both iSPNs and dSPNs (Kress et al., 2013). Retrograde viral tracing methods have suggested that there may be differences in dSPN and iSPN innervation by cortical region (Wall et al., 2013). However, a subsequent study, in which the number of presynaptic neurons labeled outside the striatum was an order of magnitude greater, found no difference in dSPN or iSPN innervation by cortical region (Guo et al., 2015). This suggests that the initial findings of segregated inputs could have suffered from sampling bias and cautions against drawing quantitative conclusions from viral methods with low infection rates.

In the dorsal striatum, both dSPN and iSPNs receive glutamatergic inputs from not only cortex, but also thalamus, and single axons of both types contact both dSPNs and iSPNs (Doig et al., 2010). However, cortical and thalamic synapses differ in their properties: cortical synapses onto SPNs have a much higher failure rate and are facilitating, while thalamic synapses have a lower rate of failure and are depressing (Ding et al., 2008). Synapses from different thalamic nuclei display a great deal of heterogeneity in their strength, short term dynamics, and SPN subtype targets (Ellender et al., 2013; Smith et al., 2009). Furthermore, through their interactions with cholinergic interneurons, thalamic neurons appear to “gate” cortical information flow, suggesting an interesting interaction between these two information streams for striatal processing (Ding et al., 2010).

The striatum is densely innervated by dopamine neurons. Dopamine neurons of the substantia nigra pars compacta target the dorsal striatum, while those of the ventral tegmental area (VTA) target the ventral striatum. Dopamine’s influence on striatal function is vast and complex due to the variety of D1-like (D1, D5) and D2-like (D2, D3, D4) receptors located on both axon terminals and postsynaptic sites in the striatum. In addition to differential effects in directly modulating D1 containing dSPN and D2 containing iSPNs, dopamine powerfully influences striatal processing through its actions on cholinergic and GABAergic interneurons as well as presynaptic terminals of inputs to the striatum. These actions of dopamine on the inputs can be acute and reversible or expressed in the form of long-term changes in synaptic strength (For reviews see (Gerfen and Surmeier, 2011; Kreitzer and Malenka, 2008). For review covering afferents to GABAergic interneurons see (Tepper et al., 2010; Tepper and Koós, 2017)).

It is now known that the midbrain provides another source of glutamate to these regions. The VTA contains a population of glutamate neurons that projects to the striatum (Hnasko et al., 2012; Yamaguchi et al., 2015, 2011, 2007) as well as a population of neurons that co-release dopamine and glutamate (Hnasko et al., 2010; Stuber et al., 2010; Tecuapetla et al., 2010). VTA glutamate neurons projecting to the NAc shell synapse onto FSIs and mediate behavioral aversion (Qi et al., 2016). In addition, it has recently been reported that nondopaminergic glutamate neurons in the VTA can also co-release GABA and that stimulation of these neurons can be transiently reinforcing (Yoo et al., 2016).

Further, dopamine neurons have been shown to co-release GABA along with dopamine in the dorsal striatum (Tritsch et al., 2012) and the NAc, though dopamine neuron IPSCs recorded in the NAc are smaller than in the dorsal striatum (Kim et al., 2015). Interestingly, dopamine neurons do not synthesize GABA through canonical GAD65 and GAD67 enzymes, but rather uptake it through the plasma membrane (Tritsch et al., 2014) and synthesize it through a non-canonical aldehyde dehydrogenase 1a1 pathway (Kim et al., 2015).

Functional GABA projections from cortex have recently been described in both NAc and dorsal striatum. The mPFC sends GABA projections to the NAc shell, and activation of these afferents produces avoidance behavior in a real time place preference paradigm (Lee et al., 2014). In the dorsal striatum cortical GABA projections from layer 5 and 6 SOM+ cells to both dSPNs and iSPNs have recently been described in both auditory and motor cortex, implying that this may be a general feature of cortex (Rock et al., 2016). These findings are in agreement with anatomical studies that identified cortical GABA projections to the striatum (Tomioka et al., 2015) and challenge the common assumption that cortical projections to the striatum are exclusively glutamatergic.

Another large source of GABA to the striatum comes from back projections from the GP. Recently identified arkypallidal, or GP-TA, neurons in the GP exclusively target the striatum with ~10,000 boutons per cell and synapse onto SPNs as well as both cholinergic and GABAergic (PV+ and NOS+) interneurons, suggesting that these cells may provide the largest extrinsic source of GABA to the striatum (Mallet et al., 2012). Functionally, in vivo recordings of these neurons during a behavior task, in which animals must abruptly abort an action, imply that arkypallidal cells may be responsible for sending a “stop” signal to the striatum to cancel evolving action plans (Mallet et al., 2016). In addition to arkypallidal cells, GP neurons that project to downstream structures send GABA projections to the striatum that target PV interneurons (Bevan et al., 1998). VTA contains nondopaminergic GABA neurons that project to the NAc (Van Bockstaele and Pickel, 1995). In addition to locally modulating the excitability of VTA dopamine neurons (Tan et al., 2012; Van Zessen et al., 2012), these neurons project to cholinergic interneurons in the NAc and can pause their firing, which plays a role in associative learning (Brown et al., 2012). Thus, in addition to GABA co-released from midbrain dopamine neurons, GABA projection neurons in midbrain also influence striatal circuits via their projections to CINs.

Another novel inhibitory projection from the bed nucleus of the stria terminalis to the striatum has also been described, terminating on patch and exo-patch SPNs in the dorsal striatum (Smith et al., 2016). Using tract tracing, another study identified an extrinsic source of acetylcholine arising from brainstem nuclei, the pedunculopontine nucleus and the lateral dorsal tegmental nucleus (Dautan et al., 2014). The function of these newly identified projections still needs to be determined. Other minor afferents to the striatum release neuromodulators, such as serotonin from the dorsal raphe (Miguelez et al., 2014) as well as histamine and oxytocin from hypothalamus (Bolam and Ellender, 2016; Dölen et al., 2013).

3. Local striatal circuitry

There is increasing appreciation for the role that local striatal circuitry plays in allowing the striatum to compute sensory, motor, and limbic information into behavioral or cognitive output. It can filter, modulate, amplify, and ultimately transform information flow through this region and produce an output that could not be predicted by the summation/integration of the inputs to SPNs alone.

The local connectivity is mediated by the striatal interneurons and by the synapses formed by collateral axons of SPNs that linger within the striatum and contact other SPNs. While some of these local connections are reciprocal, most of them are unidirectional in nature. Local connectivity can shape the output by directly affecting the activity of projection neurons via classic axo-somatic and axo-dendritic synapses onto SPNs that alter neuronal excitability, cell signaling, dendritic integration, and gene expression. Alternatively, local connectivity can act indirectly to affect striatal output via axo-axonal synapses, which can modulate the synaptic inputs to SPNs or the synaptic transmission from SPNs (Figure 2). Below, we will briefly outline the different synaptic connections between local striatal interneurons and projection neurons (Figure 3).

Figure 2. Axo-axonal modulation in the striatum.

Figure 2

Simple diagram of known axo-axonal presynaptic modulation by dopamine (DA) and acetylcholine (ACh) in the striatum. D2, dopamine D2 receptor; M5, M5 muscarinic ACh receptor; M2/4, M2 and M4 muscarinic ACh receptors; nAR, nicotinic ACh receptor.

Figure 3. Main identified synaptic connections between local striatal neurons.

Figure 3

Schematic showing intrastriatal connectivity between interneurons and projection neurons. Interneurons displayed with molecular markers used to identify cells during experiments. Abbreviations: dSPN, direct pathway projecting medium spiny neuron; iSPN, indirect pathway projecting medium spiny neuron; ChAT, choline-acetyl transferase; TH, tyrosine hydroxlase; SOM, somatostatin; NOS, nitric oxide synthase; NPY, neuropeptide Y; NGF, neurogliaform; 5HT3R, serotonin 3 receptor; PV, parvalbumin;

Axo-axonal connections

In axo-axonal synapses, neurotransmitter release from one synaptic bouton activates receptors located presynaptically on another bouton or axon fiber and act directly to suppress, facilitate, or even trigger neurotransmitter release from the “postsynaptic” axon terminal, sometimes without affecting the firing of action potentials from the postsynaptic neuron (Figure 2). While anatomical evidence for axo-axonal synapses in the striatum is still lagging behind, functional data indicates that presynaptic heteroreceptors can be activated quickly and reliable upon neurotransmitter release from other terminals (Cachope et al., 2012; Kosillo et al., 2016; Threlfell et al., 2012; Shin et al., in revisions). A recent example of an axo-axonal connection that was functionally identified in the striatum is between CINs and dopamine (DA) fibers. Synchronized release of ACh from cholinergic interneurons can activate presynaptic nicotinic acetylcholine receptors (nAChRs) located on DA axons and evoke DA release in both ventral and dorsal striatum (Cachope et al., 2012; Threlfell et al., 2012). Via presynaptic muscarinic M5 receptors on DA axon fibers, ACh can further enhance DA transmission in the striatum (Shin et al., 2015). Through activation of presynaptic muscarinic M2/M4 receptors on glutamatergic terminals from cortex and thalamus, ACh can also suppress glutamate transmission onto SPNs (Ding et al., 2010). GABA transmission from FSIs onto SPNs is also inhibited by ACh through presynaptic muscarinic receptors (Koós and Tepper, 2002). In addition, the existence of nAChRs on glutamatergic terminals has been suggested (Kaiser and Wonnacott, 2000). Further confirmation that CIN stimulation and/or ACh release from these interneurons can produce these modulatory effects on GABA and glutamate transmission will be important for the understanding the spatial and temporal limitation of ACh transmission

Other neurotransmitters act presynaptically in the striatum as well: through activation of presynaptic D2 receptors, DA inhibits its own release from DA terminals through autoreceptors, GABA transmission from iSPNs (Dobbs et al., 2016a; Tecuapetla et al., 2009), and acetylcholine release from CINs. It is proposed to inhibit glutamate transmission from cortical afferents (Bamford et al., 2004), though because cortical D2 receptor expression is sparse, postsynaptic effects of dopamine on glutamate transmission likely play a large role in observed modulation (Higley and Sabatini, 2010). Further, single NPY-NGF interneuron firing can inhibit cortical glutamate release through presynaptic GABAB receptors, suggesting an axo-axonal synapse (Logie et al., 2013).

Axo-dendritic and axo-somatic connections

CIN-GABA interneuron

Cholinergic interneurons form synapses onto FSIs and release both glutamate and acetylcholine, though these synapses are weak and sometimes undetected (English et al., 2012; Nelson et al., 2014a). ACh has been shown to both depolarize FSIs through nicotinic receptors and to modulate their synaptic transmission by inhibiting GABA transmission onto SPNs through presynaptic muscarinic receptors (Koós and Tepper, 2002; Luo et al., 2013). Interestingly, despite this evidence of nicotinic stimulation and depolarizing responses to cholinergic stimulation, FSIs do not appear to play a major role in the disynaptic inhibition of SPNs through cholinergic interneurons (Nelson et al., 2014b). CIN to FSI synaptic connections are not reciprocal as it has been shown that FSI cells do not form synapses onto cholinergic interneurons (Straub et al., 2016; Szydlowski et al., 2013).

Cholinergic interneurons excite both 5HT3R expressing and NPY-NGF neurons through nicotinic receptors (English et al., 2012; Faust et al., 2015). It is unclear whether these connections are reciprocal. However, stimulation of a single CIN is capable of eliciting polysynaptic GABAA-mediated synaptic responses both in itself and in nearby cholinergic interneurons (Sullivan et al., 2008). This recurrent feedback inhibition requires activation of nAChRs and while the cell-type identity of this source of GABA is unknown, it is possible that GABA interneurons are responsible for the recurrent inhibition. Though generally not detected in paired recordings between cells in close proximity (Gittis et al., 2010; Ibáñez-Sandoval et al., 2011), optogenetic experiments have shown that LTSIs form GABAergic synapses onto distant cholinergic interneurons inhibiting them through GABAA receptors (Straub et al., 2016). In addition, there is evidence that LTSIs can induce slow depolarizations in cholinergic interneurons mediated by nitric oxide (Elghaba et al., 2016).

TH interneurons are depolarized by nicotinic agonists, indicating the presence of nicotinic receptors and suggesting that TH interneurons receive synaptic contacts from cholinergic interneurons or afferents (Ibáñez-Sandoval et al., 2015; Luo et al., 2013). However, direct synaptic responses from cholinergic neurons have not yet been recorded.

CIN-SPN

Cholinergic neurons acutely modulate SPN excitability and synaptic function directly though muscarinic receptors (for review see (Oldenburg and Ding, 2011)) and have a role in controlling striatal plasticity (Wang et al., 2006). Indirectly cholinergic interneurons can modulate SPNs through disynaptic release of DA from midbrain terminals (Cachope et al., 2012; Threlfell et al., 2012) and release of GABA from FAIs or NPY-NGF interneurons (English et al., 2012; Faust et al., 2016, 2015) as well as from DA terminals (Nelson et al., 2014).

SPNs are known to synapse onto cholinergic interneurons and release GABA (Chuhma et al., 2011). There is some anatomical evidence that dSPNs in the dorsal striatum make more synaptic contacts onto cholinergic interneurons than iSPNs do (Martone et al., 1992); however, rigorous quantification is lacking, and whether this is generalizable to the ventral striatum is unknown.

GABA interneurons-GABA interneurons

FSIs weakly and sparsely inhibit LTSIs (Gittis et al., 2010; Lee et al., 2017; Szydlowski et al., 2013). Paired recordings have failed to detect LTSI synapses onto FSIs; however, as LTSI synapses onto distant CINs and SPNs have also been missed by paired recordings, it remains possible that LTSIs synapse onto distant FSIs. FSIs strongly synapse on other FSIs, communicating chemically through GABA synapses as well as electrically through gap junctions (Gittis et al., 2010; Koós and Tepper, 1999). NPY-NGF interneurons receive GABAergic inhibition from FSIs (Lee et al., 2017).

GABA interneurons-SPNs

FSIs strongly synapse on SPNs (Gittis et al., 2010; Koós and Tepper, 1999; Planert et al., 2010). In vivo, FSIs respond more quickly and to lower thresholds of cortical stimulation than SPNs providing a powerful mechanism for feedforward inhibition from cortex (Mallet and Moine, 2005). FSIs appear to have a slight bias in targeting dSPNs over iSPNs (Gittis et al., 2010; Planert et al., 2010). There is no evidence for SPNs synapsing onto FSIs (Chuhma et al., 2011). LTSIs have been found in paired recordings with SPNs to evoke sparse GABA responses on SPNs that were much weaker than those evoked from FSIs to SPNs (Gittis et al., 2010; Luo et al., 2013). Recent work using viral tracing and optogenetics has shown that this difference in connectivity detected with paired recordings from cells in close proximity is likely due to differences in spatial organization of these circuits: LTSIs target distant SPNs, while FSIs target SPNs in close proximity (Straub et al., 2016). Further, LTSIs synapse on distal dendrites of SPNs, while FSIs target the soma and proximal dendrites of SPNs.

FAIs provide fast GABAA mediated inhibition to SPNs, while NPY-NGF neurons mediate slow GABAA synaptic responses (English et al., 2012; Faust et al., 2015). NPY-NGF activation inhibits SPN firing in vivo (Lee et al., 2017). Because NPY-NGF interneurons and FAIs are excited by cholinergic stimulation, this provides a both quick and prolonged mechanism for disynaptic cholinergic inhibition of SPNS. TH interneurons synapse onto SPNs and release GABA. Interestingly, TH interneurons are the only identified GABAergic interneurons to receive inputs form SPNs (Ibáñez-Sandoval et al., 2010), suggesting a unique role for these interneurons in striatal processing.

Collateral transmission between SPNs

In addition to their long range projection targets, SPNs extend dense axon collaterals that stay within the striatum or nucleus accumbens and target other SPNs (Chang and Kitai, 1985; Pennartz et al., 1991; Somogyi et al., 1981; Wilson and Groves, 1980). These GABAergic synapses are mostly formed on distal dendrites of SPNs (Wilson and Groves, 1980) and are relatively weak compared to FSI inputs that are made at the soma (Czubayko and Plenz, 2002; Koós et al., 2004; Tunstall et al., 2002). The connectivity is sparse in the sense that only ~30% of paired recordings between SPNs produce a synaptic response, and these connections are all unidirectional.

The network of collaterals axons can provide the circuit with a sufficient degree of lateral inhibition to shape the output of the striatum. Indeed, despite the early work showing weak synaptic connections between discrete pairs of SPNs, synchronized stimulation of axon collateral from multiple SPN can provide a strong source of inhibition to other striatal SPNs (Dobbs et al., 2016). The properties and regulation of this network of lateral inhibition are discussed in more detail below. In addition, we present some models for the organization of the lateral inhibition and its implications for circuit level function.

4. SPN collateral transmission drives lateral inhibition

Anatomical and functional evidence

Early anatomical studies identifying SPNs in the dorsal striatum as the projection neurons of this region made note of their dense axons collaterals within the striatum (Grofová, 1975; Kemp and Powell, 1971). Indeed, some of the earliest studies identifying SPNs, including those by Ramon y Cajal (1911), identified them as interneurons due to their intrastriatal axon branches and the limitations of the Golgi technique (Tepper et al., 2010). By combining antidromic stimulation recordings with cell filling and anatomical reconstruction, it became more clear that SPNs have both projections out of the striatum and an extensive collateral plexus within the striatum (Preston et al., 1980). The use of light and electron microscopy in the early 1980s allowed both Wilson and Groves and Somogyi and colleagues to show that SPN axons, in addition to projecting out of the striatum, target other SPNs within the striatum (Somogyi et al., 1981; Wilson and Groves, 1980). As the idea of separate direct/indirect pathway SPN subpopulations began to emerge, Kawaguchi and colleagues demonstrated that SPNs form intrastriatal axon collaterals regardless of the downstream projection target (i.e. whether that cell was an iSPN or dSPN) (Kawaguchi et al., 1990). Electron microscopy was combined with immunocytochemistry to show that iSPNs and dSPNs were synaptically interconnected and synapses were formed between SPNs of different subclasses (Yung et al., 1996). Anatomical studies in the NAc confirmed that similar to the dorsal striatum, NAc SPNs have dense axon collaterals that remain within the accumbens (Chang and Kitai, 1985; Pennartz et al., 1991) in addition to the long range axon projections to the midbrain and pallidum.

Despite the wealth of anatomical evidence, the resulting functional effect of this high density of local synapses between SPNs has been harder to grasp. Electrophysiological studies using antidromic stimulation hinted at early evidence of functional GABAergic synapses from the axon collaterals in both dorsal striatum (Park et al., 1980) and nucleus accumbens (Chang and Kitai, 1985). Despite this encouraging functional evidence, one of the first studies directly addressing the existence of functional synapses using sharp intracellular electrode recordings in pairs of SPNs in the dorsal striatum of rats found no evidence for functional synaptic connections (Jaeger et al., 1994). Another study using paired whole cell recordings between fast-spiking PV interneurons and SPNs found that PV cells form strong GABAergic connections to SPNs (Koós and Tepper, 1999). Together, these findings suggested that PV interneurons provide a larger source of GABAergic inhibition to SPNs than SPNs collaterals do as previous hypotheses based on the anatomical data suggested (Tepper et al., 2008).

In 2002, the first evidence of monosynaptic GABAergic connections between SPNs was demonstrated by Tunstall and colleagues, followed shortly by Czubayko and Plenz (Czubayko and Plenz, 2002; Tunstall et al., 2002). Tunstall and colleagues used sharp intracellular electrode recordings in adult rat striatal slices and found that 9/45 pairs of SPN formed unidirectional synapses that were blocked by GABAA antagonists. In addition to the sparse connectivity (20% of pairs), the synaptic responses were of small size (IPSP amplitudes of a few hundred microvolts) and showed high rate of failure. As a consequence, averaging hundreds of events was required in order to resolve IPSPs, possibly explaining the discrepancy with the previous study of SPN pairs. Similarly, using young rats and organotypic cultures, Czubayko and Plenz found further evidence for the existence of functional synapses between SPNs. They found synaptic connections between 5/38 (13%) of pairs of SPNs with high failure rates (~0.5) in rat slices, in agreement with the connectivity rates reported by Tunstall et al. IPSPs reported in Czubayko and Plenz were an order of magnitude larger than those reported in Tunstall et al. (~2 mV vs. ~250 µV) likely due to difference the recording configurations (sharp electrode vs whole-cell). Two years later, Taverna and colleagues showed that SPNs in the NAc also form functional collateral synapses with each other (Taverna et al., 2004). Like in the dorsal striatum, these connections are relatively sparse (13/38 pairs, 34%), mediated by GABAA receptors, and mainly unidirectional in nature.

Properties and organization of the inhibitory collateral transmission

The creation of BAC transgenic mice expressing fluorescent proteins under the control of the Drd1 or Drd2 promotor (Gong et al., 2003) allowed for the examination of collateral synapses in identified dSPNs and iSPNs. In both the dorsal striatum and the NAc, Taverna and colleagues found asymmetries in both the connection probability and strength of the synapses between the SPN subclasses (Taverna et al., 2008). They found that while iSPNs form synapses on postsynaptic dSPN (13/47, 28%) and iSPNs (14/39, 36%) with a similar probability, dSPNs were much more likely to form synapse onto other dSPNs (5/19, 26%) than onto iSPNs (3/47, 6%). In addition, presynaptic iSPNs generate larger inhibitory postsynaptic currents (IPSCs) than dSPNs do. In similar experiments, Planert and colleagues found slightly different connection probabilities between the subclasses but largely supported the finding that iSPNs are more likely to form collateral synapses than dSPNs are (Planert et al., 2010). Future studies will determine whether the asymmetry in the connection strength and probability is observed in vivo. These findings are important because they imply that the pattern of collateral synapses between SPNs is organized in a specific nonrandom fashion, and they suggest that there is an asymmetry in the strength and probability of contacts among the subclasses.

Though new hypotheses emerged on how SPN collaterals could shape information flow through the striatum (Plenz, 2003), the concept that PV interneurons are the main source of GABAergic inhibition in the striatum persisted in the field (Tepper et al., 2008, 2004). This was due in large part to the fact that PV→SPN synapses are more likely formed on the SPN soma and generate larger IPSPs than SPN →SPN synapses, which form mainly on distal dendrites (Koós et al., 2004). As a result, many broad theories and overviews of basal ganglia function do not incorporate a possible role for SPN collateral synapses in influencing basal ganglia output (Calabresi et al., 2014; Cazorla et al., 2015; Sesack and Grace, 2010; Silberberg and Bolam, 2015). This is in large part because their function remains unclear but also because collaterals are thought to be weak and infrequent. However, estimates suggest there are >2800 SPNs within the dendritic field of a given SPN (Oorschot, 1996) and thus, 28–36% of them represent more than 790–1,000 likely synaptic partners for each SPN. A clever analysis of synapses in the striatum based on electron microscopy data and synaptic density estimates that each SPN receives around 1,200 to 1,800 synaptic contacts from other SPNs, implying that lateral inhibition could strongly affect SPNs and thus shape output (Wilson, 2007).

Indeed, recent studies from our laboratory have demonstrated an important functional role for SPN collaterals in controlling the excitability of SPNs and locomotion. Simultaneous activation of iSPN axon collaterals using channelrhodopsin caused noticeable inhibition of dSPN firing (Dobbs et al., 2016), proving further confirmation of the anatomical predictions. Important to note with this study is that synchronous activation with channelrhodopsin in a slice does not necessarily recapitulate the in vivo behavior of the network. However, these experiments do shed light on what is possible given the existing synapses within the striatal network. Altogether, the anatomical and functional findings show that while connection probability between isolated pairs of SPNs is relatively low, and it involves only a handful of synapses, concerted activation of axon collaterals from multiple connected SPNs can have a significant inhibitory effect on the firing of their synaptic partners. Also, activation of D2Rs expressed in iSPNs suppresses the collateral GABA transmission, and as such, dopamine and D2-like receptor agonists reduce the inhibition from iSPNs to dSPNs (Dobbs et al., 2016; Lemos et al. 2016). By controlling the degree of lateral inhibition from iSPNs to dSPNs, D2Rs can gate dSPN action potential firing and output. In animals with targeted deletion of D2Rs in iSPNs, neither elevation of endogenous dopamine by cocaine nor a D2-like agonist is able to suppress GABAergic transmission from iSPNs → dSPNs, and mice show impaired cocaine-induced locomotion (Dobbs et al., 2016a). In addition, animals lacking D2Rs in iSPNs display hypolocomotion, bradykinesia, and reduced in vivo firing of SPNs throughout the striatum. In slices prepared from these animals, SPNs were shown to receive more potent inhibition from GABA transmission, likely caused in part by an enhanced collateral transmission that followed the chronic loss of D2R mediated suppression of iSPN collaterals (Lemos et al., 2016). Taken together, these studies suggest an important functional role for the network of collateral axons and synapses in regulating the excitability of SPNs within the striatum and in gating behavioral output of the basal ganglia. The behavioral implications of the lateral inhibition are multiple and seem to involve locomotion and stimulant-induced activity.

Other groups have visualized SPN network activity using calcium indicators both in vitro and in vivo to provide evidence that SPNs are organized into neuronal clusters based on synchronized activation that might represent functional ensembles (Barbera et al., 2016; Carrillo-reid et al., 2008; Carrillo-Reid et al., 2011). In light of this, many computational models have attempted to explain how feedforward inhibition, mediated by PV cells, and lateral inhibition, mediated by SPN collaterals, could generate these ensembles and thereby transform information flow through the striatum (Humphries et al., 2010, 2009; Moyer et al., 2014; Ponzi and Wickens, 2012, 2010, 2013; Yim et al., 2011). These models suggest that lateral inhibition mediated by SPN collaterals might play a role in driving ensemble coherence on slower timescales or in synchronizing active ensembles and suppressing unsynchronized cells. However, these models all assume a random connection probability between SPNs and deserve to be reconsidered in light of the recent evidence of asymmetries in both strength and connection probability of the collateral connections between SPN subclasses and the possible existence of neuronal clusters or ensembles.

Evidence of striatal functional units

An important question to ask is whether the fundamental unit of information processing in the striatum. Though not easily visualized, are there functional unit throughout the striatum analogous to the canonical cortical microcircuit columns? A convergence of evidence suggests that ensemble activity in the striatum is important for information transfer through the region and the resulting behavior.

On one hand, little evidence of synchronized firing is found between SPNs during in vivo electrophysiological recordings in anesthetized animals (Stern et al., 1998). This could speak against the existence of ensemble but given the low firing rate of SPNs in vivo, little spontaneous synchrony is expected. On the other hand, in vivo recordings from freely moving animals show greater evidence of synchronized firing among SPNs (Miller et al., 2008). In behaving animals, SPN firing is seen time locked to specific aspects of the behavioral task, suggesting synchronized spiking of functionally related populations (Adler et al., 2012; Barnes et al., 2005; Cacciapaglia et al., 2011; Carelli, 2002; Jin et al., 2014; West and Carelli, 2016) that are not necessarily in close proximity (Bakhurin et al., 2016). Ensembles are likely formed by sparsely distributed rather than tightly clustered neurons, and as such, the chances of simultaneously recording their firing from the same electrode array might be low. In support of this concept, a recent study that used a large-scale recording approach found higher correlation in the resting state firing of neurons that were responsive to reward predictive cue compared to non-responsive neurons when investigating the overall firing of a large number of striatal neurons during a Pavlovian reward conditioning task (Bakhurin et al., 2016).

Fiber photometry recordings of SPNs in the NAc and dorsal striatum also showed large transients of calcium concentrations time locked to cues and specific behaviors during behavioral tasks (Calipari et al., 2016; Cui et al., 2014; Natsubori et al., 2017; Zalocusky et al., 2016), suggestive of population synchrony during actions, though again not necessarily due to cells in close proximity. Moreover, deep brain calcium imaging, which provides a wide field of view, shows clusters of SPNs that display synchronized calcium transients during free locomotion. The clusters are observed in both the dorsomedial and dorsolateral, and they involve both dSPNs and iSPNs, which supports the concept of striatal functional units where iSPNs and dSPNs work together to produce a behavioral output (Barbera et al., 2016).

Further, in the ventral striatum, which has long been hypothesized to be organized into functionally distinct ensembles of cells (Pennartz et al., 1994), further support for the idea that sparse ensembles of cells work together to control striatal behaviors comes from the analysis of immediate early gene expression. For example, examination of immediate early gene c-fos expression after NAc dependent behaviors, such as context based cocaine induced locomotor sensitization has suggested that sparse ensembles of SPNs are necessary for behavioral expression (Hope et al., 2006). Lending support to this, causal manipulations of c-fos expression or cell excitability in the sparsely labeled cells that express c-fos during locomotor sensitization or conditioned place preference has been shown to disrupt that behavior (Koya et al., 2009; Tolliver et al., 2000).

How are these ensembles formed? Are the cells in the striatum simply following cortical commands or does intrastriatal circuitry play a role in their activity? The topographical organization of the cortex is conveyed through its projections to the striatum and is proposed to persist through the basal ganglia in segregated information streams (Alexander et al., 1986; Dudman and Gerfen, 2015). Thus, cortical inputs could contribute to the creation of discrete functional units in part defined by their inputs. However, while cortical inputs are critical for generating spikes in SPN, it is unlikely that they are defining the units, based on the current experimental evidence. Systematic tracing and mapping of cortical projections into the dorsal striatum has identified 29 distinct subregions, suggesting the existence of numerous discrete zones of striatal processing (Hintiryan et al., 2016). In congruence with the anatomy, the in vivo firing of neurons in the dorsolateral striatum, which receives sensorimotor input, correlates with specific body sensations, movements, or even direction of movement of a specific limb (Carelli and West, 1991; West et al., 1990). However, since cortical axons in the striatum display a great deal of divergence and nearby SPNs in general receive few shared cortical inputs (Kincaid et al., 1998), it is unlikely that these inputs completely define striatal ensembles.

Furthermore, in vitro calcium imaging experiments have identified synchronized cell assemblies in the striatum active in the presence of tonic excitatory input (Carrillo-reid et al., 2008; Carrillo-Reid et al., 2011), demonstrating that glutamate is necessary to generate activity in these cell assemblies. However, blocking GABA abolishes much of the sequential activation of the assemblies in the network, implying that local circuitry within the striatum is important for the observed network behavior of the cell assemblies. Note that similar sequentially switching cell assemblies have been identified in the primate striatum in vivo during behavior, suggesting there is a behavioral relevance for the sequential activation pattern of the cell assemblies (Adler et al., 2012). Taken together, all these studies support the idea that SPNs in the striatum are organized into discrete functional units responsible for driving specific behaviors and that local striatal circuitry plays a role in their dynamics.

5. A conceptual model of synaptic organization of the lateral inhibition

If indeed the striatum operates as a collection of functional units that are responsible for driving specific behaviors, as the most recent evidence seems to indicate, then how are these functional units created and organized? What is the synaptic organization of the functional units both within and between units that allow for the selection of appropriate behaviors and other striatal functions? We hypothesize that the lateral inhibition between SPNs is critical in defining the functional units and that the organization of the SPN collateral transmission within the striatum shapes this lateral inhibition to define behaviorally relevant ensembles of SPNs. This Perspective will mainly focus on the role of the most abundant striatal neurons, SPNs; however, each functional unit is likely to also contain interneurons that regulate activity both within and between functional units.

In this section, we will use the current anatomical and functional data on SPN collateral transmission to draw some conceptual models and speculate how the connectivity pattern of collateral transmission can bring functional relevance to the lateral inhibition within the striatum in order to effectively aid in the integration and filtering of inputs. A variety of theories on striatal function emerged early on that incorporated the idea that lateral inhibition shaped information flow through the striatum (Parent et al., 2000; Smith and Bolam, 1990). Groves, drawing on “winner-take-all” principles of lateral inhibition first established in the retina, hypothesized that SPN collaterals project to all their neighbors allowing the network to temporally and spatially sharpen the broad input from the cortex and that dysregulation in this system leads to movement disorders (Groves, 1983). Similarly, Penney and Young posited that lateral inhibition between projection neurons (SPNs) served to suppress or maintain desired cortical input and hypothesized that the extent and patterning of these axon collaterals may develop during acquisition of new motor and cognitive capacities (Penney and Young, 1983). Extending the idea that SPN collaterals sharpen and filter information into the ventral striatum, Swerdlow and Koob proposed that in the NAc, a disruption of this circuitry, and the resulting inability to filter emotional or cognitive inputs, could be an underlying cause of psychiatric diseases like depression and schizophrenia (Swerdlow and Koob, 1987).

Despite the fact that there is no easily visualized pattern of SPN connectivity, several pieces of evidence point to an overall structure: 1) the connection probability is asymmetrical between SPN subclasses where iSPNs are more likely than dSPNs to be connected to both dSPNs and other iSPNs, while dSPNs are more likely to synapse on other dSPNs (Planert et al., 2010; Taverna et al., 2008); 2) the synaptic contacts made by iSPNs produce larger responses than those from dSPNs (Planert et al., 2010; Taverna et al., 2008, 2004; Tunstall et al., 2002); 3) connections are unidirectional in nature and sparse so they do not just contact every nearby SPN (Planert et al., 2010; Taverna et al., 2008, 2004; Tunstall et al., 2002); and 4) SPNs synapse mainly onto distal dendrites of other SPNs, although there is some diversity on location suggesting different types of collateral synapses (Oorschot et al., 2013). These properties suggest a more complex, nonrandom functional organization of lateral inhibition in the striatum. They also argue strongly against the idea of an all-to-all/winner-take-all network of lateral inhibition in which the first, or most strongly activated, neuron is able to inhibit all of its neighboring SPNs like early models of striatal function hypothesized (Groves, 1983).

The wiring diagram in Figure 4A illustrates a possible pattern of connectivity between SPNs in an attempt to merge the current findings at the behavioral level with the circuit level/anatomical data. We know that (i) dSPN activation facilitates behavior, (ii) iSPN activation suppresses behavior, (iii) also that both dSPNs and iSPNs are active during behavior, and (iv) there is collateral transmission between SPNs. This supports a conceptual model in which each behavior is controlled by a functional unit (FU, rectangles) consisting of both dSPN and iSPN clusters (green and red circles), where the dSPN cluster that drives the desired behavior (say, behavior A) and the iSPN cluster that inhibits the competing behavior (say, behavior B) are simultaneously active during a task (dark colored circles), while the complementary clusters are inactive (light colored circles) (Mink, 1996).

Figure 4. Proposed model of organization of striatal functional units with lateral inhibition within and between units.

Figure 4

A, Colored circles represent multiple SPNs from each subclass that are activated together during behavior forming a cluster or ensemble. Each functional unit contains one dSPN cluster and one iSPN cluster. Clusters that are active during execution of behavior A are highlighted and have strong output that enhances behavior A (green arrow) and suppress competing behavior B (red brake). Lateral inhibition from the active clusters further limits activity of the other silent dSPN and iSPN clusters. B, Basic connectivity of the lateral inhibition within and between the two functional units described. Note that all clusters receive excitatory inputs as well. For simplicity, all interneurons are not included in this early conceptual model.

The model then incorporates recent new evidence showing that iSPN collateral transmission within the striatum is important for setting the magnitude of the lateral inhibition onto dSPNs. It proposes that the active cluster of iSPNs within FU2, which suppresses behavior B through long range indirect-pathway projections (into the globus pallidus and ventral pallidum), also extends intrastriatal collateral axons to inhibit the dSPN cluster (in the same functional unit) and diminish its output through long-range direct pathway projections to the midbrain, which would otherwise promote the competing behavior B.

Paired recording data suggests that iSPNs are connected to a third of nearby iSPNs, and so we hypothesize that the active iSPN cluster in each functional unit also extends axon collateral that synapse onto the iSPN cluster in the competing functional unit and therefore exerts lateral inhibition that limits any possible inhibition of, in this case, the desired behavior A. Thus, under this proposed organization of the collateral transmission, the active iSPN cluster not only suppresses the activation of competing behaviors but also suppresses the inhibition of the desired behavior A.

With regards to the active dSPNs cluster in FU1, the model also incorporates the reported dSPN→dSPN collateral transmission and also the existence of weaker dSPN→iSPN transmission. The model proposes that the active dSPNs cluster in FU1 exerts lateral inhibition onto the dSPN cluster in FU2 to further limit the competing behavior B, and via the weaker connections to the iSPN cluster in FU1, the lateral inhibition dampens the suppression of the desired behavior A. In summary, the active dSPN cluster in this FU1 promotes the desired behavior A via the canonical long-range direct pathway projections, and, at the same time via the axon collaterals that remain within the striatum, it provides lateral inhibition within and across functional units to limit competing behaviors and dampen the inhibition of the desired behavior. This conceptual model will provide a framework for future studies

Recent work has shown the strength of local GABA transmission provided by iSPN axon collaterals can be modulated by dopamine via activation of D2 receptors (Dobbs et al., 2016a, 2016b; Lemos et al., 2016). Acute activation of D2 receptors in iSPNs suppresses the lateral inhibition onto dSPNs and chronic loss of the D2 receptors in these neurons causes an increase in the strength of the GABA-mediated tonic current on dMSMs, which reduces the in vivo firing of SPNs and impairs locomotion (Lemos et al., 2016). If we hypothesize that iSPN collaterals then “gate” dSPNs, and when collateral transmission is inhibited by dopamine, dSPNs are better able to fire action potentials and drive behavior (Dobbs et al., 2016a).

How can this hypothesized connectivity pattern aid in the processing of inputs?

The impact that this proposed pattern of axon collateral transmission has on the activity of the SPN clusters will have to be determined and tested in future studies. To illustrate the possible ways in which the connectivity patterns in this conceptual model can aid on the processing and filtering of the excitatory inputs to the striatum, we applied a standard conductance-based formalism for the individual neurons and synapses that reflects the main network properties rather than reproducing the underlying biology. In this simple model all cells are identical and receive identical inputs. We consider two patterns of synaptic connectivity and determine how the different network scenarios affect the firing of action potentials of the SPN clusters in the two representative functional units.

In a model without lateral inhibition, each SPN cluster fires action potentials at relatively high frequency in response to excitatory inputs, and the clusters fire in synchrony because the inputs are identical and the cells are homogeneous (Figure 5A). Addition of lateral inhibition with an all-to-all connectivity pattern and equal strength among connections (symmetric connectivity patterns) results in synchronous firing at lower frequency (Figure 5B). Weakening transmission from dSPNs to introduce the experimentally reported asymmetry in the connectivity strength among SPN subclasses, results in non-trivial firing patterns where iSPN clusters become more dominant (Figure 5C). iSPN clusters fire in anti-phase bursts and dSPN clusters fire a single spike at every cycle. While it is interesting to see that functional units are taking turns, the low activity of SPNs clusters will not be conducive to the generation of motor behaviors. Going back to the pattern of Figure 5B, removal of the crossed connections between the functional leaves the originally proposed organization of the lateral inhibition (structured, Figure 5D) which produces synchrony in all clusters, similar to the firing pattern observed in Figure 5B (with slightly lower frequency). This illustrates the different balances that can be operating in the network to produce the same type of patterns. However, when these structured connections are no longer symmetrical between SPN subclasses, the firing patterns can change drastically. As shown in the example of Figure 5E, the asymmetric structured pattern can result in synchronized activation of dSPN and iSPN clusters belonging to different competing functional units where iSPN and dSPN clusters fire in phase, while the other clusters are silent. In other words, some clusters are shut down while others are activated in phase. Note that the excitatory inputs to the silent SPN clusters are still present and have not changed and as such this shows that, at least in this conceptual model, the lateral inhibition can indeed filter excitatory inputs to SPN to shut down the firing of a selected subset of iSPN and dSPN clusters. This example illustrates how the lateral inhibition organized in the proposed “structured” fashion and with the asymmetry in the connections can generate the predicted pattern of co-activation of iSPNs and dSPNs that is observed in vivo and that is conducive to activation of behavior A and inhibition of behavior B. In this way, lateral inhibition can help to striatum in its important task of selecting motor actions.

Figure 5. Diverse connectivity patterns for the lateral inhibition generate diverse firing patterns.

Figure 5

Left, Diagram of the connectivity pattern for the lateral inhibition within and between the two functional units. Right, raster plots showing the firing of action potentials in response to identical excitatory inputs for each SPN cluster (green for dSPN and red for iSPN clusters) for each functional unit (top raster for FU1, bottom raster for FU2). A, no lateral connections between cells. B, “all-to-all” connections between cells with equal synaptic strengths. C, “all-to-all” connections between cells with asymmetrical connection strengths (iSPN>dSPN). D, proposed organization of lateral inhibition between and within units (“structured”) with equal connection strengths. E, “Structured” connections between cells with asymmetrical connection strengths (iSPN>dSPN).

This model is meant to serve as a proof of concept that the patterning of the lateral inhibition between SPNs would be able to transform synchronous inputs to all cells into selective activation by a subset of SPNs, while silencing others. It is a simplified model with a small number of cells created for illustrative purposes. Thus, it does not capture all known aspects of SPN biology, most importantly, as detailed earlier, the fact that synapses from other SPNs are formed mainly onto the distal dendrites. Dendritic modeling work suggests that inhibition in distal dendrites can exert powerful shunting of excitatory inputs synapsing on dendrites, as glutamate inputs to SPNs tend to do (Gidon and Segev, 2012). However, it is unlikely that a single SPN in vivo is capable of suppressing spikes in other SPNs. In order for the computation illustrated with the model to occur in vivo, there would likely need to be some degree of synchronized activation of axon collaterals, which we have recently demonstrated can shunt action potential firing (Dobbs et al., 2016a). Thus, synchronized firing of ensembles would likely be needed to affect firing. How reasonable is that assumption based on the data? As discussed previously (see “Evidence of striatal functional units” Section 4), in vivo electrophysiology and calcium imaging data indicates evidence for synchronized firing of SPNs during specific aspects of behavior. Though cells in close proximity might not always fire in synchrony (Stern et al., 1998), given the broad reach of SPN axons and dendrites, these units need not consist of cell bodies in close proximity. Indeed, the model predicts that SPNs that are co-active are not connected by collateral synapses and thus co-active SPNs could be spaced further beyond the region of overlap between their axonal arborization and the dendritic field. Considering the sparse, disparate pattern of cells that express immediately early genes during behaviors like context induced cocaine locomotor sensitization and the wide spatial distribution of correlated, task responsive neurons (Bakhurin et al., 2016), it seems likely that the cells in these units would, in fact, not be in close proximity (but note (Barbera et al., 2016).

Another possibility that has been suggested (Koós et al., 2004; Plenz, 2003), is that local collaterals of SPNs are important for controlling transition to depolarized, plateau “upstates”, which prime SPNs for firing, rather than affecting the spiking itself. As SPN collaterals synapse mainly on distal dendrites (Oorschot et al., 2013) , and plateau upstate potentials are generated by regenerative events in the distal dendrites (Plotkin et al., 2011), GABA released from SPN could provide sufficient shunting to prevent this transition. Indeed, in vitro studies show that GABA antagonists prolong SPN upstate duration, suggesting that GABA inputs plays a role in upstate timing (Vergara et al., 2003). By controlling the upstate transitions that precede spiking and bursting, as well as directly shunting excitatory inputs, inhibition in the dendrites would still can provide a powerful means of effecting cell output (Gidon and Segev, 2012; Lovett-Barron et al., 2012). Thus, the local SPN synapses between and within units could be controlling the excitability of neighboring cells in the way proposed by the model, without necessarily directly suppressing spiking.

Open Questions

This proposed model is intended as a preliminary proposition to stimulate discussion and research regarding how local circuitry within the striatum might be organized given known functional and behavioral data. As a result, there are many open questions regarding this framework. How many cells do these units consist of? How much overlap is there between units? How might these units arise? Are they fixed developmentally, completely plastic based on learning, or somewhere in between? Given the known role of the striatum in certain forms of learning and the numerous examples of plasticity in the region, it is likely that any organization in the region is malleable. Dopamine release during behavior can modulate the strength of the lateral inhibition to alter the connectivity pattern in ways that it favors the activation of selected FUs and the inhibition of others to ultimately raise the signal-to-noise ratio and reinforce desirable behaviors. To what degree is this the case, and are certain units more flexible and plastic than others? Furthermore, how would the various interneurons and their numerous known synapses detailed earlier fit within this functional unit framework?

Concluding remarks

Given the growing diversity of striatal inputs, cell-types, and connections described in recent years, it is important to develop a framework for how the circuitry within the striatum is wired to drive processing of information and affect output. Here we present a hypothesis on how one aspect of the striatal circuitry, the lateral inhibition between SPNs, could be organized given available data. It is our hope that this will drive the future work needed to both determine the validity of this model and to understand how other elements of striatal circuitry could fit within this framework.

Acknowledgments

This work was supported by the Intramural Programs of National Institute on Alcohol Abuse and Alcoholism and National Institute of Neurological Disorders and Stroke Grant (ZIA-AA000421) to VAA and by from the National Science Foundation (DMS-1313861) to HGR. We are grateful to Dr. Miriam E. Bocarsly and the other members of the Alvarez laboratory for their valuable comments on this Perspective.

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. Adler A, Katabi S, Finkes I, Israel Z, Prut Y, Bergman H. Temporal Convergence of Dynamic Cell Assemblies in the Striato-Pallidal Network. 2012;32:2473–2484. doi: 10.1523/JNEUROSCI.4830-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Albin RL, Young AB, Penney JB. The functional anatomy of basal ganglia disorders. Trends Neurosci. 1989;12:366–375. doi: 10.1016/0166-2236(89)90074-X. [DOI] [PubMed] [Google Scholar]
  3. Alexander GE, Crutcher MD. Functional architecture of basal ganglia circuits: neural substrates of parallel processing. Trends Neurosci. 1990;13:266–271. doi: 10.1016/0166-2236(90)90107-L. [DOI] [PubMed] [Google Scholar]
  4. Alexander GE, DeLong MR, Strick PL. Parallel organization of functionally segregated circuits linking basal ganglia and cortex. Annu. Rev. Neurosci. 1986;9:357–381. doi: 10.1146/annurev.neuro.9.1.357. [DOI] [PubMed] [Google Scholar]
  5. Alvarez VA. Clues on the coding of reward cues by the nucleus accumbens. Proc. Natl. Acad. Sci. 2016;113:2560–2562. doi: 10.1073/pnas.1601162113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bamford NS, Zhang H, Schmitz Y, Wu NP, Cepeda C, Levine MS, Schmauss C, Zakharenko SS, Zablow L, Sulzer D. Heterosynaptic dopamine neurotransmission selects sets of corticostriatal terminals. Neuron. 2004;42:653–663. doi: 10.1016/S0896-6273(04)00265-X. [DOI] [PubMed] [Google Scholar]
  7. Banghart MR, Neufeld SQ, Wong NC, Sabatini BL. Enkephalin Disinhibits Mu Opioid Receptor-Rich Striatal Patches via Delta Opioid Receptors. Neuron. 2015;88:1227–1239. doi: 10.1016/j.neuron.2015.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Barbera G, Liang B, Zhang L, Gerfen CR, Culurciello E, Chen R, Li Y, Lin D-T. Spatially Compact Neural Clusters in the Dorsal Striatum Encode Locomotion Relevant Information. Neuron. 2016;92:202–213. doi: 10.1016/j.neuron.2016.08.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Barnes TD, Kubota Y, Hu D, Jin DZ, Graybiel AM. Activity of striatal neurons reflects dynamic encoding and recoding of procedural memories. Nature. 2005;437:1158–1161. doi: 10.1038/nature04053. [DOI] [PubMed] [Google Scholar]
  10. Bateup HS, Santini E, Shen W, Birnbaum S, Valjent E, Surmeier DJ, Fisone G, Nestler EJ, Greengard P. Distinct subclasses of medium spiny neurons differentially regulate striatal motor behaviors. Proc. Natl. Acad. Sci. U. S. A. 2010;107:14845–50. doi: 10.1073/pnas.1009874107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Bertran-Gonzalez J, Bosch C, Maroteaux M, Matamales M, Hervé D, Valjent E, Girault J-A. Opposing patterns of signaling activation in dopamine D1 and D2 receptor-expressing striatal neurons in response to cocaine and haloperidol. J. Neurosci. 2008;28:5671–5685. doi: 10.1523/JNEUROSCI.1039-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Bevan MD, Booth Pa, Eaton Sa, Bolam JP. Selective innervation of neostriatal interneurons by a subclass of neuron in the globus pallidus of the rat. J. Neurosci. 1998;18:9438–52. doi: 10.1523/JNEUROSCI.18-22-09438.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Bock R, Shin JH, Kaplan AR, Dobi A, Markey E, Kramer PF, Gremel CM, Christensen CH, Adrover MF, Alvarez VA. Strengthening the accumbal indirect pathway promotes resilience to compulsive cocaine use. Nat. Neurosci. 2013;16:632–638. doi: 10.1038/nn.3369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Bolam JP, Ellender TJ. Histamine and the striatum. Neuropharmacology. 2016;106:74–84. doi: 10.1016/j.neuropharm.2015.08.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Bolam JP, Hanley JJ, Booth Pa, Bevan MD. Synaptic organisation of the basal ganglia. J. Anat. 2000;196(Pt 4):527–542. doi: 10.1046/j.1469-7580.2000.19640527.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Brimblecombe KR, Cragg SJ. The striosome and matrix compartments of the striatum: a path through the labyrinth from neurochemistry towards function. ACS Chem. Neurosci. acschemneuro.6b00333. 2016 doi: 10.1021/acschemneuro.6b00333. [DOI] [PubMed] [Google Scholar]
  17. Britt JP, Benaliouad F, McDevitt RA, Stuber GD, Wise RA, Bonci A. Synaptic and Behavioral Profile of Multiple Glutamatergic Inputs to the Nucleus Accumbens. Neuron. 2012;76:790–803. doi: 10.1016/j.neuron.2012.09.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Brown MTC, Tan KR, O’Connor EC, Nikonenko I, Muller D, Lüscher C. Ventral tegmental area GABA projections pause accumbal cholinergic interneurons to enhance associative learning. Nature. 2012;492:452–6. doi: 10.1038/nature11657. [DOI] [PubMed] [Google Scholar]
  19. Bubser M, Deutch AY. Thalamic paraventricular nucleus neurons collateralize to innervate the prefrontal cortex and nucleus accumbens. Brain Res. 1998;787:304–310. doi: 10.1016/S0006-8993(97)01373-5. [DOI] [PubMed] [Google Scholar]
  20. Cacciapaglia F, Wightman RM, Carelli RM. Microcircuits within the Nucleus Accumbens during Sucrose-Directed behavior. 2011;31:13860–13869. doi: 10.1523/JNEUROSCI.1340-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Cachope R, Mateo Y, Mathur BN, Irving J, Wang HL, Morales M, Lovinger DM, Cheer JF. Selective activation of cholinergic interneurons enhances accumbal phasic dopamine release: Setting the tone for reward processing. Cell Rep. 2012;2:33–41. doi: 10.1016/j.celrep.2012.05.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Calabresi P, Picconi B, Tozzi A, Ghiglieri V, Di Filippo M. Direct and indirect pathways of basal ganglia: a critical reappraisal. Nat. Neurosci. 2014;17:1022–1030. doi: 10.1038/nn.3743. [DOI] [PubMed] [Google Scholar]
  23. Calipari ES, Bagot RC, Purushothaman I, Davidson TJ, Yorgason JT, Peña CJ, Walker DM, Pirpinias ST, Guise KG, Ramakrishnan C, Deisseroth K, Nestler EJ. In vivo imaging identifies temporal signature of D1 and D2 medium spiny neurons in cocaine reward. Proc. Natl. Acad. Sci. U. S. A. 2016;113:2726–2731. doi: 10.1073/pnas.1521238113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Carelli RM. Nucleus accumbens cell firing during goal-directed behaviors for cocaine vs. “natural” reinforcement. Physiol. Behav. 2002;76:379–387. doi: 10.1016/S0031-9384(02)00760-6. [DOI] [PubMed] [Google Scholar]
  25. Carelli RM, West MO. Representation of the body by single neurons in the dorsolateral striatum of the awake, unrestrained rat. J. Comp. Neurol. 1991;309:231–249. doi: 10.1002/cne.903090205. [DOI] [PubMed] [Google Scholar]
  26. Carrillo-Reid L, Hernandez-Lopez S, Tapia D, Galarraga E, Bargas J. Dopaminergic Modulation of the Striatal Microcircuit: Receptor-Specific Configuration of Cell Assemblies. J. Neurosci. 2011;31:14972–14983. doi: 10.1523/JNEUROSCI.3226-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Carrillo-reid L, Tecuapetla F, Tapia D, Herna A, Galarraga E. Encoding Network States by Striatal Cell Assemblies. 2008;1:1435–1450. doi: 10.1152/jn.01131.2007. [DOI] [PubMed] [Google Scholar]
  28. Cazorla M, de Carvalho FD, Chohan MO, Shegda M, Chuhma N, Rayport S, Ahmari SE, Moore H, Kellendonk C. Dopamine D2 receptors regulate the anatomical and functional balance of basal ganglia circuitry. Neuron. 2014;81:153–64. doi: 10.1016/j.neuron.2013.10.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Cazorla M, Kang UJ, Kellendonk C. Balancing the basal ganglia circuitry: A possible new role for dopamine D2 receptors in health and disease. Mov. Disord. 2015;30:895–903. doi: 10.1002/mds.26282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Chang HT, Kitai ST. Projection neurons of the nucleus accumbens: an intracellular labeling study. Brain Res. 1985;347:112–116. doi: 10.1016/0006-8993(85)90894-7. [DOI] [PubMed] [Google Scholar]
  31. Chuhma N, Tanaka KF, Hen R, Rayport S. Functional connectome of the striatal medium spiny neuron. J. Neurosci. 2011;31:1183–1192. doi: 10.1523/JNEUROSCI.3833-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Cohen MX, Frank MJ. Neurocomputational models of basal ganglia function in learning, memory and choice. Behav. Brain Res. 2009;199:141–156. doi: 10.1016/j.bbr.2008.09.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Collins AGE, Frank MJ. Opponent actor learning (OpAL): Modeling interactive effects of striatal dopamine on reinforcement learning and choice incentive. Psychol. Rev. 2014;121:337–66. doi: 10.1037/a0037015. [DOI] [PubMed] [Google Scholar]
  34. Crittenden JR, Graybiel AM. Basal Ganglia Disorders Associated with Imbalances in the Striatal Striosome and Matrix Compartments. Front. Neuroanat. 2011;5:59. doi: 10.3389/fnana.2011.00059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Cui G, Jun SB, Jin X, Pham MD, Vogel SS, Lovinger DM, Costa RM. Concurrent activation of striatal direct and indirect pathways during action initiation. Nature. 2014;494:238–242. doi: 10.1038/nature11846. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Czubayko U, Plenz D. Fast synaptic transmission between striatal spiny projection neurons. Proc. Natl. Acad. Sci. U. S. A. 2002;99:15764–15769. doi: 10.1073/pnas.242428599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Dautan D, Huerta-Ocampo I, Witten IB, Deisseroth K, Bolam JP, Gerdjikov T, Mena-Segovia J. A major external source of cholinergic innervation of the striatum and nucleus accumbens originates in the brainstem. J. Neurosci. 2014;34:4509–18. doi: 10.1523/JNEUROSCI.5071-13.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. DeLong MR. Primate models of movement disorders of basal ganglia origin. Trends Neurosci. 1990;13:281–285. doi: 10.1016/0166-2236(90)90110-V. [DOI] [PubMed] [Google Scholar]
  39. Ding J, Peterson JD, Surmeier DJ. Corticostriatal and Thalamostriatal Synapses Have Distinctive Properties. J. Neurosci. 2008;28:6483–6492. doi: 10.1523/JNEUROSCI.0435-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Ding JB, Guzman JN, Peterson JD, Goldberg JA, Surmeier DJ. Thalamic gating of corticostriatal signaling by cholinergic interneurons. Neuron. 2010;67:294–307. doi: 10.1016/j.neuron.2010.06.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Dobbs LK, Kaplan AR, Lemos JC, Matsui A, Rubinstein M, Alvarez VA. Dopamine Regulation of Lateral Inhibition between Striatal Neurons Gates the Stimulant Actions of Cocaine. Neuron. 2016a:1–14. doi: 10.1016/j.neuron.2016.04.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Dobbs LK, Lemos JC, Alvarez VA. Restructuring of basal ganglia circuitry and associated behaviors triggered by low striatal D2 receptor expression: implications for substance use disorders. Genes, Brain Behav. 2016b:56–70. doi: 10.1111/gbb.12361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Doig NM, Moss J, Bolam JP. Cortical and Thalamic Innervation of Direct and Indirect Pathway Medium-Sized Spiny Neurons in Mouse Striatum. J. Neurosci. 2010;30:14610–14618. doi: 10.1523/JNEUROSCI.1623-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Dölen G, Darvishzadeh A, Huang KW, Malenka RC. Social reward requires coordinated activity of nucleus accumbens oxytocin and serotonin. Nature. 2013;501:179–84. doi: 10.1038/nature12518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Dudman JT, Gerfen CR. The Basal Ganglia, The Rat Nervous System. 2015 doi: 10.1016/B978-0-12-374245-2.00017-6. [DOI] [Google Scholar]
  46. Durieux PF, Bearzatto B, Guiducci S, Buch T, Waisman A, Zoli M, Schiffmann SN, de Kerchove d’Exaerde A. D2R striatopallidal neurons inhibit both locomotor and drug reward processes. Nat. Neurosci. 2009;12:393–5. doi: 10.1038/nn.2286. [DOI] [PubMed] [Google Scholar]
  47. Elghaba R, Vautrelle N, Bracci E. Mutual Control of Cholinergic and Low-Threshold Spike Interneurons in the Striatum. Front. Cell. Neurosci. 2016;10:111. doi: 10.3389/fncel.2016.00111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Ellender TJ, Harwood J, Kosillo P, Capogna M, Bolam JP. Heterogeneous properties of central lateral and parafascicular thalamic synapses in the striatum. J. Physiol. 2013;591:257–72. doi: 10.1113/jphysiol.2012.245233. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. English DF, Ibáñez-Sandoval O, Stark E, Tecuapetla F, Buzsáki G, Deisseroth K, Tepper JM, Koós T. GABAergic circuits mediate the reinforcement-related signals of striatal cholinergic interneurons. Nat. Neurosci. 2012;15:123–130. doi: 10.1038/nn.2984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Faust TW, Assous M, Shah F, Tepper JM, Koós T. Novel fast adapting interneurons mediate cholinergic-induced fast GABA<inf>A</inf> inhibitory postsynaptic currents in striatal spiny neurons. Eur. J. Neurosci. 2015;42:1764–1774. doi: 10.1111/ejn.12915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Faust TW, Assous M, Tepper JM, Koós T. Neostriatal GABAergic Interneurons Mediate Cholinergic Inhibition of Spiny Projection Neurons. J. Neurosci. 2016;36:9505–9511. doi: 10.1523/JNEUROSCI.0466-16.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Frank MJ. Dynamic dopamine modulation in the basal ganglia: a neurocomputational account of cognitive deficits in medicated and nonmedicated Parkinsonism. J. Cogn. Neurosci. 2005;17:51–72. doi: 10.1162/0898929052880093. [DOI] [PubMed] [Google Scholar]
  53. Gagnon D, Petryszyn S, Sanchez MG, Bories C, Beaulieu JM, De Koninck Y, Parent A, Parent M. Striatal Neurons Expressing D1 and D2 Receptors are Morphologically Distinct and Differently Affected by Dopamine Denervation in Mice. Sci. Rep. 2017;7:41432. doi: 10.1038/srep41432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Gerfen CR. The neostriatal mosaic: compartmentalization of corticostriatal input and striatonigral output systems. Nature. 1984;311:461–4. doi: 10.1038/311461a0. [DOI] [PubMed] [Google Scholar]
  55. Gerfen CR, Surmeier DJ. Modulation of striatal projection systems by dopamine. Annu. Rev. Neurosci. 2011;34:441–466. doi: 10.1146/annurev-neuro-061010-113641. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Gerfen CRC, Engber T, Mahan L, Susel Z, Chase T, Monsma F, Sibley D. D1 and D2 dopamine receptor-regulated gene expression of striatonigral and striatopallidal neurons. Science. 1990 doi: 10.1126/science.2147780. [DOI] [PubMed] [Google Scholar]
  57. Gidon A, Segev I. Principles Governing the Operation of Synaptic Inhibition in Dendrites. Neuron. 2012;75:330–341. doi: 10.1016/j.neuron.2012.05.015. [DOI] [PubMed] [Google Scholar]
  58. Gittis AH, Nelson AB, Thwin MT, Palop JJ, Kreitzer AC. Distinct roles of GABAergic interneurons in the regulation of striatal output pathways. J. Neurosci. 2010;30:2223–34. doi: 10.1523/JNEUROSCI.4870-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Gokce O, Stanley GM, Treutlein B, Neff NF, Camp JG, Malenka RC, Rothwell PE, Fuccillo MV, Südhof TC, Quake SR. Cellular Taxonomy of the Mouse Striatum as Revealed by Single-Cell RNA-Seq. Cell Rep. 2015:1126–1137. doi: 10.1016/j.celrep.2016.06.059. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Gong S, Zheng C, Doughty ML, Losos K, Didkovsky N, Schambra UB, Nowak NJ, Joyner A, Leblanc G, Hatten ME, Heintz N. A gene expression atlas of the central nervous system based on bacterial artificial chromosomes. Nature. 2003;425:917–925. doi: 10.1038/nature02033. [DOI] [PubMed] [Google Scholar]
  61. Gonzales KK, Smith Y. Cholinergic interneurons in the dorsal and ventral striatum: Anatomical and functional considerations in normal and diseased conditions. Ann. N. Y. Acad. Sci. 2015;1349:1–45. doi: 10.1111/nyas.12762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Graveland GA, Difiglia M. The frequency and distribution of medium-sized neurons with indented nuclei in the primate and rodent neostriatum. Brain Res. 1985;327:307–311. doi: 10.1016/0006-8993(85)91524-0. [DOI] [PubMed] [Google Scholar]
  63. Grofová I. The identification of striatal and pallidal neurons projecting to substantia nigra. An experimental study by means of retrograde axonal transport of horseradish peroxidase. Brain Res. 1975;91:286–291. doi: 10.1016/0006-8993(75)90550-8. [DOI] [PubMed] [Google Scholar]
  64. Groves PM. A theory of the functional organization of the neostriatum and the neostriatal control of voluntary movement. Brain Res. Rev. 1983;5:109–132. doi: 10.1016/0165-0173(83)90011-5. [DOI] [PubMed] [Google Scholar]
  65. Guo Q, Wang D, He X, Feng Q, Lin R, Xu F, Fu L, Luo M. Whole-brain mapping of inputs to projection neurons and cholinergic interneurons in the dorsal striatum. PLoS One. 2015;10:1–15. doi: 10.1371/journal.pone.0123381. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Guzman MS, de Jaeger X, Raulic S, Souza IA, Li AX, Schmid S, Menon RS, Gainetdinov RR, Caron MG, Bartha R, Prado VF, Prado MAM. Elimination of the vesicular acetylcholine transporter in the striatum reveals regulation of behaviour by cholinergic-glutamatergic co-transmission. PLoS Biol. 2011;9 doi: 10.1371/journal.pbio.1001194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Haber SN, Fudge JL, McFarland NR. Striatonigrostriatal pathways in primates form an ascending spiral from the shell to the dorsolateral striatum. J. Neurosci. 2000;20:2369–2382. doi: 10.1523/JNEUROSCI.20-06-02369.2000. doi: http://www.jneurosci.org/content/20/6/2369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Higley M, Sabatini B. Competitive regulation of synaptic Ca2+ influx by D2 dopamine and A2A adenosine receptors. Nat. Neurosci. 2010;13:958–966. doi: 10.1038/nn.2592.Competitive. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Higley MJ, Gittis AH, Oldenburg IA, Balthasar N, Seal RP, Edwards RH, Lowell BB, Kreitzer AC, Sabatini BL. Cholinergic interneurons mediate fast VGluT3-dependent glutamatergic transmission in the striatum. PLoS One. 2011;6 doi: 10.1371/journal.pone.0019155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Hikida T, Kimura K, Wada N, Funabiki K, Nakanishi Shigetada S. Distinct Roles of Synaptic Transmission in Direct and Indirect Striatal Pathways to Reward and Aversive behavior. Neuron. 2010;66:896–907. doi: 10.1016/j.neuron.2010.05.011. [DOI] [PubMed] [Google Scholar]
  71. Hikosaka O. Neural systems for control of voluntary action--a hypothesis. Adv Biophys. 1998;35:81–102. doi:S0065227X9880004X [pii] [PubMed] [Google Scholar]
  72. Hikosaka O. Basal ganglia--possible role in motor coordination and learning. Curr. Opin. Neurobiol. 1991;1:638–643. doi: 10.1016/S0959-4388(05)80042-X. [DOI] [PubMed] [Google Scholar]
  73. Hintiryan H, Foster NN, Bowman I, Bay M, Song MY, Gou L, Yamashita S, Bienkowski MS, Zingg B, Zhu M, Yang XW, Shih JC, Toga AW, Dong H-W. The mouse cortico-striatal projectome. Nat. Neurosci. 2016;19:1100–1114. doi: 10.1038/nn.4332. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Hnasko TS, Chuhma N, Zhang H, Goh GY, Sulzer D, Palmiter RD, Rayport S, Edwards RH. Vesicular Glutamate Transport Promotes Dopamine Storage and Glutamate Corelease In Vivo. Neuron. 2010;65:643–656. doi: 10.1016/j.neuron.2010.02.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Hnasko TS, Hjelmstad GO, Fields HL, Edwards RH. Ventral Tegmental Area Glutamate Neurons: Electrophysiological Properties and Projections. J. Neurosci. 2012;32:15076–15085. doi: 10.1523/JNEUROSCI.3128-12.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Hope BT, Simmons DE, Mitchell TB, Kreuter JD, Mattson BJ. Cocaine-induced locomotor activity and Fos expression in nucleus accumbens are sensitized for 6 months after repeated cocaine administration outside the home cage. Eur. J. Neurosci. 2006;24:867–875. doi: 10.1111/j.1460-9568.2006.04969.x. [DOI] [PubMed] [Google Scholar]
  77. Humphries MD, Wood R, Gurney K. Reconstructing the three-dimensional gabaergic microcircuit of the striatum. PLoS Comput. Biol. 2010;6 doi: 10.1371/journal.pcbi.1001011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Humphries MD, Wood R, Gurney K. Dopamine-modulated dynamic cell assemblies generated by the GABAergic striatal microcircuit. Neural Networks. 2009;22:1174–1188. doi: 10.1016/j.neunet.2009.07.018. [DOI] [PubMed] [Google Scholar]
  79. Ibáñez-Sandoval O, Tecuapetla F, Unal B, Shah F, Koós T, Tepper JM. A novel functionnally distinct subtype of striatal NPY interneuron. J. Neurosci. 2011;31:16757–16769. doi: 10.1523/JNEUROSCI.2628-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Ibáñez-Sandoval O, Tecuapetla F, Unal B, Shah F, Koós T, Tepper JM. Electrophysiological and morphological characteristics and synaptic connectivity of tyrosine hydroxylase-expressing neurons in adult mouse striatum. J. Neurosci. 2010;30:6999–7016. doi: 10.1523/JNEUROSCI.5996-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Ibáñez-Sandoval O, Xenias HS, Tepper JM, Koós T. Dopaminergic and cholinergic modulation of striatal tyrosine hydroxylase interneurons. Neuropharmacology. 2015;95:468–476. doi: 10.1016/j.neuropharm.2015.03.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Isomura Y, Takekawa T, Harukuni R, Handa T, Aizawa H, Takada M, Fukai T. Reward-Modulated Motor Information in Identified Striatum Neurons. J. Neurosci. 2013;33:10209–10220. doi: 10.1523/JNEUROSCI.0381-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Jaeger D, Kita H, Wilson CJ. Surround inhibition among projection neurons is weak or nonexistent in the rat neostriatum. J. Neurophysiol. 1994;72:2555–2558. doi: 10.1152/jn.1994.72.5.2555. [DOI] [PubMed] [Google Scholar]
  84. Jin X, Tecuapetla F, Costa RM. Basal ganglia subcircuits distinctively encode the parsing and concatenation of action sequences. Nat. Neurosci. 2014;17:423–30. doi: 10.1038/nn.3632. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Kaiser S, Wonnacott S. α-Bungarotoxin-Sensitive Nicotinic Receptors Indirectly Modulate [3H]Dopamine Release in Rat Striatal Slices via Glutamate Release. Mol. Pharmacolgy. 2000;58:312–318. doi: 10.1124/mol.58.2.312. [DOI] [PubMed] [Google Scholar]
  86. Kawaguchi Y, Wilson CJ, Emson PC. Projection subtypes of rat neostriatal matrix cells revealed by intracellular injection of biocytin. J. Neurosci. 1990;10:3421–38. doi: 10.1523/JNEUROSCI.10-10-03421.1990. doi: http://www.jneurosci.org/content/10/10/3421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Kemp JM, Powell TP. The structure of the caudate nucleus of the cat: light and electron microscopy. Philos Trans R Soc L. B Biol Sci. 1971;262:383–401. doi: 10.1098/rstb.1971.0102. [DOI] [PubMed] [Google Scholar]
  88. Kim J-I, Ganesan S, Luo SX, Wu Y-W, Park E, Huang EJ, Chen L, Ding JB. Aldehyde dehydrogenase 1a1 mediates a GABA synthesis pathway in midbrain dopaminergic neurons. Science. 2015;350:102–6. doi: 10.1126/science.aac4690. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Kincaid AE, Zheng T, Wilson CJ. Connectivity and convergence of single corticostriatal axons. J. Neurosci. 1998;18:4722–31. doi: 10.1523/JNEUROSCI.18-12-04722.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Koós T, Tepper JM. Dual cholinergic control of fast-spiking interneurons in the neostriatum. J. Neurosci. 2002;22:529–535. doi: 10.1523/JNEUROSCI.22-02-00529.2002. 22/2/529 [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Koós T, Tepper JM. Inhibitory control of neostriatal projection neurons by GABAergic interneurons. Nat. Neurosci. 1999;2:467–72. doi: 10.1038/8138. [DOI] [PubMed] [Google Scholar]
  92. Koós T, Tepper JM, Wilson CJ. Comparison of IPSCs evoked by spiny and fast-spiking neurons in the neostriatum. J Neurosci. 2004;24:7916–7922. doi: 10.1523/JNEUROSCI.2163-04.2004\r24/36/7916. [pii] [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Kosillo P, Zhang Y-F, Threlfell S, Cragg SJ. Cortical Control of Striatal Dopamine Transmission via Striatal Cholinergic Interneurons. Cereb. Cortex. 2016:1–10. doi: 10.1093/cercor/bhw252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Koya E, Golden Sa, Harvey BK, Guez-Barber DH, Berkow A, Simmons DE, Bossert JM, Nair SG, Uejima JL, Marin MT, Mitchell TB, Farquhar D, Ghosh SC, Mattson BJ, Hope BT. Targeted disruption of cocaine-activated nucleus accumbens neurons prevents context-specific sensitization. Nat. Neurosci. 2009;12:1069–73. doi: 10.1038/nn.2364. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Kravitz AV, Freeze BS, Parker PRL, Kay K, Thwin MT, Deisseroth K, Kreitzer AC. Regulation of parkinsonian motor behaviours by optogenetic control of basal ganglia circuitry. Nature. 2010;466:622–6. doi: 10.1038/nature09159. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Kravitz AV, Tye LD, Kreitzer AC. Distinct roles for direct and indirect pathway striatal neurons in reinforcement. Nat. Neurosci. 2012;15:816–818. doi: 10.1038/nn.3100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Kreitzer AC, Malenka RC. Striatal Plasticity and Basal Ganglia Circuit Function. Neuron. 2008;60:543–554. doi: 10.1016/j.neuron.2008.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Kress GJ, Yamawaki N, Wokosin DL, Wickersham IR, Shepherd GMG, Surmeier DJ. Convergent cortical innervation of striatal projection neurons. Nat. Neurosci. 2013;16:665–7. doi: 10.1038/nn.3397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Kupchik YM, Brown RM, Heinsbroek Ja, Lobo MK, Schwartz DJ, Kalivas PW. Coding the direct/indirect pathways by D1 and D2 receptors is not valid for accumbens projections. Nat. Neurosci. 2015:1–4. doi: 10.1038/nn.4068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Lee AT, Vogt D, Rubenstein JL, Sohal VS. A Class of GABAergic Neurons in the Prefrontal Cortex Sends Long-Range Projections to the Nucleus Accumbens and Elicits Acute Avoidance Behavior. J. Neurosci. 2014;34:11519–11525. doi: 10.1523/JNEUROSCI.1157-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Lee K, Holley SM, Shobe JL, Chong NC, Cepeda C, Levine MS, Masmanidis SC. Parvalbumin Interneurons Modulate Striatal Output and Enhance Performance during Associative Learning. Neuron. 2017;93:1451–1463.e4. doi: 10.1016/j.neuron.2017.02.033. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Lei W, Jiao Y, Del Mar N, Reiner A. Evidence for Differential Cortical Input to Direct Pathway versus Indirect Pathway Striatal Projection Neurons in Rats. J. Neurosci. 2004;24:8289–8299. doi: 10.1523/JNEUROSCI.1990-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Lemos JC, Friend DM, Kaplan AR, Shin JH, Rubinstein M, Kravitz AV, Alvarez VA. Enhanced GABA Transmission Drives Bradykinesia Following Loss of Dopamine D2 Receptor Signaling. Neuron. 2016;90:824–838. doi: 10.1016/j.neuron.2016.04.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Lobo MK, Covington HE, Chaudhury D, Friedman AK, Sun H, Damez-Werno D, Dietz DM, Zaman S, Koo JW, Kennedy PJ, Mouzon E, Mogri M, Neve RL, Deisseroth K, Han M-H, Nestler EJ. Cell type-specific loss of BDNF signaling mimics optogenetic control of cocaine reward. Science. 2010;330:385–90. doi: 10.1126/science.1188472. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Logie C, Bagetta V, Bracci E. Presynaptic control of corticostriatal synapses by endogenous GABA. J. Neurosci. 2013;33:15425–31. doi: 10.1523/JNEUROSCI.2304-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Lovett-Barron M, Turi GF, Kaifosh P, Lee PH, Bolze F, Sun X-H, Nicoud J-F, Zemelman BV, Sternson SM, Losonczy A. Regulation of neuronal input transformations by tunable dendritic inhibition. Nat. Neurosci. 2012;15:423–30. doi: 10.1038/nn.3024. S1-3. [DOI] [PubMed] [Google Scholar]
  107. Luo R, Janssen MJ, Partridge JG, Vicini S. Direct and GABA-mediated indirect effects of nicotinic ACh receptor agonists on striatal neurones. J. Physiol. 2013;591:203–17. doi: 10.1113/jphysiol.2012.241786. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. MacAskill AF, Little JP, Cassel JM, Carter AG. Subcellular connectivity underlies pathway-specific signaling in the nucleus accumbens. Nat Neurosci. 2012;15:1624–1626. doi: 10.1038/nn.3254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Mallet N, Micklem BR, Henny P, Brown MT, Williams C, Bolam JP, Nakamura KC, Magill PJ. Dichotomous Organization of the External Globus Pallidus. Neuron. 2012;74:1075–1086. doi: 10.1016/j.neuron.2012.04.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Mallet N, Le Moine C. Feedforward Inhibition of Projection Neurons by Fast- Spiking GABA Interneurons in the Rat Striatum In Vivo. 2005;25:3857–3869. doi: 10.1523/JNEUROSCI.5027-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Mallet N, Schmidt R, Leventhal D, Chen F, Amer N, Boraud T, Berke JD. Arkypallidal Cells Send a Stop Signal to Striatum. Neuron. 2016;89:308–316. doi: 10.1016/j.neuron.2015.12.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Martone ME, Armstrong DM, Young SJ, Groves PM. Ultrastructural examination of enkephalin and substance P input to cholinergic neurons within the rat neostriatum. Brain Res. 1992;594:253–262. doi: 10.1016/0006-8993(92)91132-X. [DOI] [PubMed] [Google Scholar]
  113. Miguelez C, Morera-Herreras T, Torrecilla M, Ruiz-Ortega Ja, Ugedo L. Interaction between the 5-HT system and the basal ganglia: functional implication and therapeutic perspective in Parkinson’s disease. Front. Neural Circuits. 2014;8:21. doi: 10.3389/fncir.2014.00021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Miller BR, Walker AG, Shah AS, Barton SJ, Rebec GV, Br M, Ag W, As S, Sj B, Gv R. Dysregulated Information Processing by Medium Spiny Neurons in Striatum of Freely Behaving Mouse Models of Huntington ’ s Disease. Manager. 2008:2205–2216. doi: 10.1152/jn.90606.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Mink JW. The Basal Ganglia and involuntary movements: impaired inhibition of competing motor patterns. Arch. Neurol. 2003;60:1365–8. doi: 10.1001/archneur.60.10.1365. [DOI] [PubMed] [Google Scholar]
  116. Mink JW. the Basal Ganglia : Focused Selection and Inhibition of Competing Motor Programs. Prog. Neurobiol. 1996;50:381–425. doi: 10.1016/S0301-0082(96)00042-1. [DOI] [PubMed] [Google Scholar]
  117. Mink JW, Thach WT. Basal ganglia intrinsic circuits and their role in behavior. Curr. Opin. Neurobiol. 1993;3:950–957. doi: 10.1016/0959-4388(93)90167-W. [DOI] [PubMed] [Google Scholar]
  118. Moyer JT, Halterman BL, Finkel LH, Wolf JA. Lateral and feedforward inhibition suppress asynchronous activity in a large, biophysically-detailed computational model of the striatal network. Front. Comput. Neurosci. 2014;8:152. doi: 10.3389/fncom.2014.00152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Muñoz-Manchado AB, Foldi C, Szydlowski S, Sjulson L, Farries M, Wilson C, Silberberg G, Hjerling-Leffler J. Novel Striatal GABAergic Interneuron Populations Labeled in the 5HT3aEGFP Mouse. Cereb. Cortex. 2016;26:96–105. doi: 10.1093/cercor/bhu179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Natsubori A, Tsustui-Kimura I, Nishida H, Bouchekioua Y, Sekiya H, Uchigashima M, Watanabe M, de Kerchove d’Exaerde A, Mimura M, Takata N, Tanaka KF. Ventrolateral striatal medium spiny neurons positively regulate food-incentive, goal-directed behavior independently of D1 and D2 selectivity. J. Neurosci. 2017:3377–16. doi: 10.1523/JNEUROSCI.3377-16.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Nelson AB, Bussert TG, Kreitzer AC, Seal RP. Striatal Cholinergic Neurotransmission Requires VGLUT3. J. Neurosci. 2014a;34:8772–7. doi: 10.1523/JNEUROSCI.0901-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Nelson AB, Hammack N, Yang CF, Shah NM, Seal RP, Kreitzer AC. Striatal cholinergic interneurons drive GABA release from dopamine terminals. Neuron. 2014b;82:63–70. doi: 10.1016/j.neuron.2014.01.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Oldenburg IA, Ding JB. Cholinergic modulation of synaptic integration and dendritic excitability in the striatum. Curr. Opin. Neurobiol. 2011;21:425–432. doi: 10.1016/j.conb.2011.04.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Oorschot DE. Total number of neurons in the neostriatal, pallidal, subthalamic, and substantia nigral nuclei of the rat basal ganglia: A stereological study using the cavalieri and optical disector methods. J. Comp. Neurol. 1996;366:580–599. doi: 10.1002/(SICI)1096-9861(19960318)366:4&#x0003c;580::AID-CNE3&#x0003e;3.0.CO;2-0. [DOI] [PubMed] [Google Scholar]
  125. Oorschot DE, Lin N, Cooper BH, Reynolds JNJ, Sun H, Wickens JR. Synaptic connectivity between rat striatal spiny projection neurons in vivo: Unexpected multiple somatic innervation in the context of overall sparse proximal connectivity. Basal Ganglia. 2013;3:93–108. doi: 10.1016/j.baga.2013.04.001. [DOI] [Google Scholar]
  126. Parent A, Sato F, Wu Y, Gauthier J, Levesque M, Parent M. Organization of the basal ganglia: the importance of axonal collateralization. Trends Neurosci. 2000;23:S20–7. doi: 10.1016/S1471-1931(00)00022-7. [DOI] [PubMed] [Google Scholar]
  127. Park MR, Lighthall JW, Kitai ST. Recurrent inhibition in the rat neostriatum. Brain Res. 1980;194:359–369. doi: 10.1016/0006-8993(80)91217-2. [DOI] [PubMed] [Google Scholar]
  128. Pennartz CMA, Groenewegen HJ, Lopes da Silva FH. The nucleus accumbens as a complex of functionally distinct neuronal ensembles: An integration of behavioural, electrophysiological and anatomical data. Prog. Neurobiol. 1994;42:719–761. doi: 10.1016/0301-0082(94)90025-6. [DOI] [PubMed] [Google Scholar]
  129. Pennartz CMAHBP, Kitai ST, Lopes da Silva FH. Contribution of NMDA receptors to postsynaptic potentials and paired-pulse facilitation in identified neurons of the rat nucleus accumbens in vitro. Exp. brain Res. 1991;86:190–198. doi: 10.1007/BF00231053. [DOI] [PubMed] [Google Scholar]
  130. Penney JB, Young a B. Speculations on the functional anatomy of basal ganglia disorders. Annu. Rev. Neurosci. 1983;6:73–94. doi: 10.1146/annurev.ne.06.030183.000445. [DOI] [PubMed] [Google Scholar]
  131. Petryszyn S, Beaulieu JM, Parent A, Parent M. Distribution and morphological characteristics of striatal interneurons expressing calretinin in mice: A comparison with human and nonhuman primates. J. Chem. Neuroanat. 2014;59–60:51–61. doi: 10.1016/j.jchemneu.2014.06.002. [DOI] [PubMed] [Google Scholar]
  132. Planert H, Szydlowski SN, Hjorth JJJ, Grillner S, Silberberg G. Dynamics of synaptic transmission between fast-spiking interneurons and striatal projection neurons of the direct and indirect pathways. J. Neurosci. 2010;30:3499–3507. doi: 10.1523/JNEUROSCI.5139-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Plenz D. When inhibition goes incognito: Feedback interaction between spiny projection neurons in striatal function. Trends Neurosci. 2003;26:436–443. doi: 10.1016/S0166-2236(03)00196-6. [DOI] [PubMed] [Google Scholar]
  134. Plotkin JL, Day M, Surmeier DJ. Synaptically driven state transitions in distal dendrites of striatal spiny neurons. Nat. Neurosci. 2011;14:881–888. doi: 10.1038/nn.2848. [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Ponzi A, Wickens J. Input dependent cell assembly dynamics in a model of the striatal medium spiny neuron network. Front. Syst. Neurosci. 2012;6:6. doi: 10.3389/fnsys.2012.00006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Ponzi A, Wickens J. Sequentially switching cell assemblies in random inhibitory networks of spiking neurons in the striatum. J. Neurosci. 2010;30:5894–5911. doi: 10.1523/JNEUROSCI.5540-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Ponzi A, Wickens JR. Optimal Balance of the Striatal Medium Spiny Neuron Network. PLoS Comput. Biol. 2013:9. doi: 10.1371/journal.pcbi.1002954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Preston RJ, Bishop GA, Kitai ST. Medium spiny neuron projection from the rat striatum: An intracellular horseradish peroxidase study. Brain Res. 1980;183:253–263. doi: 10.1016/0006-8993(80)90462-X. [DOI] [PubMed] [Google Scholar]
  139. Qi J, Zhang S, Wang H-L, Barker DJ, Miranda-Barrientos J, Morales M. VTA glutamatergic inputs to nucleus accumbens drive aversion by acting on GABAergic interneurons. Nat. Neurosci. 2016;19:725–733. doi: 10.1038/nn.4281. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Redgrave P, Prescott TJ, Gurney K. The basal ganglia: A vertebrate solution to the selection problem? Neuroscience. 1999;89:1009–1023. doi: 10.1016/S0306-4522(98)00319-4. [DOI] [PubMed] [Google Scholar]
  141. Rock C, Zurita H, Wilson C, Apicella A., junior An inhibitory corticostriatal pathway. Elife. 2016;5:1–17. doi: 10.7554/eLife.15890. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Sesack SR, Grace Aa. Cortico-Basal Ganglia reward network: microcircuitry. Neuropsychopharmacology. 2010;35:27–47. doi: 10.1038/npp.2009.93. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Shin JH, Adrover MF, Alvarez VA. n.d. Distinctive modulation of dopamine release in the nucleus accumbens shell mediated by dopamine and acetylcholine receptors. J. Neurosci. doi: 10.1523/JNEUROSCI.0596-17.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Shin JH, Adrover MF, Wess J, Alvarez Va. Muscarinic regulation of dopamine and glutamate transmission in the nucleus accumbens. Proc. Natl. Acad. Sci. 2015:201508846. doi: 10.1073/pnas.1508846112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Silberberg G, Bolam JP. Local and afferent synaptic pathways in the striatal microcircuitry. Curr Opin Neurobiol. 2015;33:182–187. doi: 10.1016/j.conb.2015.05.002. [DOI] [PubMed] [Google Scholar]
  146. Smith AD, Bolam JP. The neural network of the basal ganglia as revealed by the study of synaptic connections of identical neurones. Trends Neurosci. 1990;13:259–265. doi: 10.1016/0166-2236(90)90106-k. [DOI] [PubMed] [Google Scholar]
  147. Smith JB, Klug JR, Ross DL, Howard CD, Hollon NG, Ko VI, Hoffman H, Callaway EM, Gerfen CR, Jin X. Genetic-Based Dissection Unveils the Inputs and Outputs of Striatal Patch and Matrix Compartments. Neuron. 2016;91:1069–1084. doi: 10.1016/j.neuron.2016.07.046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Smith Y, Raju D, Nanda B, Pare JF, Galvan A, Wichmann T. The thalamostriatal systems: Anatomical and functional organization in normal and parkinsonian states. Brain Res. Bull. 2009;78:60–68. doi: 10.1016/j.brainresbull.2008.08.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Somogyi P, Bolam JP, Smith a D. Monosynaptic cortical input and local axon collaterals of identified striatonigral neurons. A light and electron microscopic study using the Golgi-peroxidase transport-degeneration procedure. J. Comp. Neurol. 1981;195:567–584. doi: 10.1002/cne.901950403. [DOI] [PubMed] [Google Scholar]
  150. Stern Ea, Jaeger D, Wilson CJ. Membrane potential synchrony of simultaneously recorded striatal spiny neurons in vivo. Nature. 1998;394:475–478. doi: 10.1038/28848. [DOI] [PubMed] [Google Scholar]
  151. Straub C, Saulnier JL, Bègue A, Feng DD, Huang KW, Sabatini BL. Principles of Synaptic Organization of GABAergic Interneurons in the Striatum. Neuron. 2016;92:84–92. doi: 10.1016/j.neuron.2016.09.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Stuber GD, Hnasko TS, Britt JP, Edwards RH, Bonci A. Dopaminergic Terminals in the Nucleus Accumbens But Not the Dorsal Striatum Corelease Glutamate. J. Neurosci. 2010;30:8229–8233. doi: 10.1523/JNEUROSCI.1754-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Surmeier DJ, Graves SM, Shen W. Dopaminergic modulation of striatal networks in health and Parkinson’s disease. Curr. Opin. Neurobiol. 2014;29:109–117. doi: 10.1016/j.conb.2014.07.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Swerdlow NR, Koob GF. Dopamine, schizophrenia, mania, and depression: Toward a unified hypothesis of cortico-striato- pallido-thalamic function. Behav. Brain Sci. 1987;10:197–245. doi: 10.1017/S0140525&#x000d7;00047488. [DOI] [Google Scholar]
  155. Szydlowski SN, Pollak Dorocic I, Planert H, Carlen M, Meletis K, Silberberg G. Target selectivity of feedforward inhibition by striatal fast-spiking interneurons. J Neurosci. 2013;33:1678–1683. doi: 10.1523/JNEUROSCI.3572-12.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Tai L-H, Lee AM, Benavidez N, Bonci A, Wilbrecht L. Transient stimulation of distinct subpopulations of striatal neurons mimics changes in action value. Nat. Neurosci. 2012;15:1281–9. doi: 10.1038/nn.3188. [DOI] [PMC free article] [PubMed] [Google Scholar]
  157. Tan KR, Yvon C, Turiault M, Mirzabekov JJ, Doehner J, Labouèbe G, Deisseroth K, Tye KM, Lüscher C. GABA Neurons of the VTA Drive Conditioned Place Aversion. Neuron. 2012;73:1173–1183. doi: 10.1016/j.neuron.2012.02.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Taverna S, Ilijic E, Surmeier DJ. Recurrent Collateral Connections of Striatal Medium Spiny Neurons Are Disrupted in Models of Parkinson’s Disease. J. Neurosci. 2008;28:5504–5512. doi: 10.1523/JNEUROSCI.5493-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Taverna S, van Dongen YC, Groenewegen HJ, Pennartz CMa. Direct physiological evidence for synaptic connectivity between medium-sized spiny neurons in rat nucleus accumbens in situ. J. Neurophysiol. 2004;91:1111–1121. doi: 10.1152/jn.00892.2003. [DOI] [PubMed] [Google Scholar]
  160. Tecuapetla F, Jin X, Lima SQ, Costa RM. Complementary Contributions of Striatal Projection Pathways to Action Initiation and Execution. Cell. 2016;166:703–715. doi: 10.1016/j.cell.2016.06.032. [DOI] [PubMed] [Google Scholar]
  161. Tecuapetla F, Koós T, Tepper JM, Kabbani N, Yeckel MF. Differential Dopaminergic Modulation of Neostriatal Synaptic Connections of Striatopallidal Axon Collaterals. J. Neurosci. 2009;29:8977–8990. doi: 10.1523/JNEUROSCI.6145-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Tecuapetla F, Matias S, Dugue GP, Mainen ZF, Costa RM. Balanced activity in basal ganglia projection pathways is critical for contraversive movements. Nat. Commun. 2014;5:4315. doi: 10.1038/ncomms5315. [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Tecuapetla F, Patel JC, Xenias H, English D, Tadros I, Shah F, Berlin J, Deisseroth K, Rice ME, Tepper JM, Koós T. Glutamatergic signaling by mesolimbic dopamine neurons in the nucleus accumbens. J Neurosci. 2010;30:7105–7110. doi: 10.1523/JNEUROSCI.0265-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Tepper JM, Koós T. GABAergic Interneurons of the Striatum. Handb. Behav. Neurosci. 2017;24:157–178. doi: 10.1016/B978-0-12-802206-1.00008-8. [DOI] [Google Scholar]
  165. Tepper JM, Koós T, Wilson CJ. GABAergic microcircuits in the neostriatum. Trends Neurosci. 2004;27:662–669. doi: 10.1016/j.tins.2004.08.007. [DOI] [PubMed] [Google Scholar]
  166. Tepper JM, Tecuapetla F, Koós T, Ibáñez-Sandoval O. Heterogeneity and diversity of striatal GABAergic interneurons. Front. Neuroanat. 2010;4:150. doi: 10.3389/fnana.2010.00150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. Tepper JM, Wilson CJ, Koós T. Feedforward and feedback inhibition in neostriatal GABAergic spiny neurons. Brain Res. Rev. 2008;58:272–281. doi: 10.1016/j.brainresrev.2007.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Threlfell S, Lalic T, Platt NJ, Jennings KA, Deisseroth K, Cragg SJ. Striatal dopamine release is triggered by synchronized activity in cholinergic interneurons. Neuron. 2012;75:58–64. doi: 10.1016/j.neuron.2012.04.038. [DOI] [PubMed] [Google Scholar]
  169. Tolliver BK, Sganga MW, Sharp FR. Suppression of c-fos induction in the nucleus accumbens prevents acquisition but not expression of morphine-conditioned place preference. Eur. J. Neurosci. 2000;12:3399–3406. doi: 10.1046/j.1460-9568.2000.00214.x. [DOI] [PubMed] [Google Scholar]
  170. Tomioka R, Sakimura K, Yanagawa Y. Corticofugal GABAergic projection neurons in the mouse frontal cortex. Front. Neuroanat. 2015;9:133. doi: 10.3389/fnana.2015.00133. [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Tritsch NX, Ding JB, Sabatini BL. Dopaminergic neurons inhibit striatal output through non-canonical release of GABA. Nature. 2012;490:262–6. doi: 10.1038/nature11466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Tritsch NX, Oh WJ, Gu C, Sabatini BL. Midbrain dopamine neurons sustain inhibitory transmission using plasma membrane uptake of GABA, not synthesis. Elife. 2014;3:e01936. doi: 10.7554/eLife.01936. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Tritsch NX, Sabatini BL. Dopaminergic Modulation of Synaptic Transmission in Cortex and Striatum. Neuron. 2012;76:33–50. doi: 10.1016/j.neuron.2012.09.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Tunstall MJ, Oorschot DE, Kean A, Wickens JR. Inhibitory Interactions Between Spiny Projection Neurons in the Rat Striatum. J. Neurophysiol. 2002;88:1263–9. doi: 10.1152/jn.2002.88.3.1263. 12205147. [DOI] [PubMed] [Google Scholar]
  175. Van Bockstaele EJ, Pickel VM. GABA-containing neurons in the ventral tegmental area project to the nucleus accumbens in rat brain. Brain Res. 1995;682:215–221. doi: 10.1016/0006-8993(95)00334-M. [DOI] [PubMed] [Google Scholar]
  176. Van Zessen R, Phillips JL, Budygin EA, Stuber GD. Activation of VTA GABA Neurons Disrupts Reward Consumption. Neuron. 2012;73:1184–1194. doi: 10.1016/j.neuron.2012.02.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Vergara R, Rick C, Hernández-López S, Laville A, Guzman JN, Galarraga E, Surmeier DJ, Bargas J. Spontaneous voltage oscillations in striatal projection neurons in a rat corticostriatal slice. J. Physiol. 2003;553:169–82. doi: 10.1113/jphysiol.2003.050799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Wall NR, De La Parra M, Callaway EM, Kreitzer AC. Differential innervation of direct- and indirect-pathway striatal projection neurons. Neuron. 2013;79:347–360. doi: 10.1016/j.neuron.2013.05.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Wang Z, Kai L, Day M, Ronesi J, Yin HH, Ding J, Tkatch T, Lovinger DM, Surmeier DJ. Dopaminergic Control of Corticostriatal Long-Term Synaptic Depression in Medium Spiny Neurons Is Mediated by Cholinergic Interneurons. Neuron. 2006;50:443–452. doi: 10.1016/j.neuron.2006.04.010. [DOI] [PubMed] [Google Scholar]
  180. West EA, Carelli RM. Nucleus Accumbens Core and Shell Differentially Encode Reward-Associated Cues after Reinforcer Devaluation. J. Neurosci. 2016;36:1128–1139. doi: 10.1523/JNEUROSCI.2976-15.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. West MO, Carelli RM, Pomerantz M, Cohen SM, Gardner JP, Chapin JK, Woodward DJ. A region in the dorsolateral striatum of the rat exhibiting single-unit correlations with specific locomotor limb movements. J. Neurophysiol. 1990;64:1233–46. doi: 10.1152/jn.1990.64.4.1233. [DOI] [PubMed] [Google Scholar]
  182. Wilson CJ. GABAergic inhibition in the neostriatum. Prog. Brain Res. 2007;160:91–110. doi: 10.1016/S0079-6123(06)60006-X. [DOI] [PubMed] [Google Scholar]
  183. Wilson CJ, Groves PM. Fine structure and synaptic connections of the common spiny neuron of the rat neostriatum: a study employing intracellular inject of horseradish peroxidase. J. Comp. Neurol. 1980;194:599–615. doi: 10.1002/cne.901940308. [DOI] [PubMed] [Google Scholar]
  184. Xenias HS, Ibáñez-Sandoval O, Koós T, Tepper JM. Are Striatal Tyrosine Hydroxylase Interneurons Dopaminergic? J. Neurosci. 2015;35:6584–6599. doi: 10.1523/JNEUROSCI.0195-15.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Yamaguchi T, Qi J, Wang HL, Zhang S, Morales M. Glutamatergic and dopaminergic neurons in the mouse ventral tegmental area. Eur. J. Neurosci. 2015;41:760–772. doi: 10.1111/ejn.12818. [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Yamaguchi T, Sheen W, Morales M. Glutamatergic neurons are present in the rat ventral tegmental area. Eur. J. Neurosci. 2007;25:106–118. doi: 10.1111/j.1460-9568.2006.05263.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Yamaguchi T, Wang H-L, Li X, Ng TH, Morales M. Mesocorticolimbic Glutamatergic Pathway. J. Neurosci. 2011;31:8476–8490. doi: 10.1523/JNEUROSCI.1598-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Yim MY, Aertsen A, Kumar A. Significance of input correlations in striatal function. PLoS Comput. Biol. 2011:7. doi: 10.1371/journal.pcbi.1002254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  189. Yoo JH, Zell V, Gutierrez-Reed N, Wu J, Ressler R, Shenasa MA, Johnson A, Fife K, Faget L, Hnasko T. Ventral tegmental area Glutamate Neurons Co-release GABA and Promote Positive Reinforcement. Nat. Commun. Accepted. 2016:1–13. doi: 10.1038/ncomms13697. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Yung KK, Smith a D, Levey a I, Bolam JP. Synaptic connections between spiny neurons of the direct and indirect pathways in the neostriatum of the rat: evidence from dopamine receptor and neuropeptide immunostaining. Eur. J. Neurosci. 1996;8:861–869. doi: 10.1111/j.1460-9568.1996.tb01573.x. [DOI] [PubMed] [Google Scholar]
  191. Zalocusky KA, Ramakrishnan C, Lerner TN, Davidson TJ, Knutson B, Deisseroth K. Nucleus accumbens D2R cells signal prior outcomes and control risky decision-making. Nature. 2016;531:642–6. doi: 10.1038/nature17400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Zhou FM, Wilson CJ, Dani JA. Cholinergic interneuron characteristics and nicotinic properties in the striatum. J. Neurobiol. 2002;53:590–605. doi: 10.1002/neu.10150. [DOI] [PubMed] [Google Scholar]

RESOURCES