Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2021 Feb 1.
Published in final edited form as: J Neurochem. 2020 Jan 7;152(4):425–448. doi: 10.1111/jnc.14919

Neurochemical Organization of the Ventral Striatum’s Olfactory Tubercle

Hillary L Cansler 1,2, Katherine N Wright 1,2,3, Lucas A Stetzik 1,2, Daniel W Wesson 1,2,3,4
PMCID: PMC7042089  NIHMSID: NIHMS1060843  PMID: 31755104

Abstract

The ventral striatum is a collection of brain structures, including the nucleus accumbens, ventral pallidum, and the olfactory tubercle (OT). While much attention has been devoted to the nucleus accumbens, a comprehensive understanding of the ventral striatum and its contributions to neurological diseases requires an appreciation for the complex neurochemical makeup of the ventral striatum’s other components. This review summarizes the rich neurochemical composition of the OT, including the neurotransmitters, neuromodulators, and hormones present. We also address the receptors and transporters involved in each system, as well as their putative functional roles. Finally, we end with briefly reviewing select literature regarding neurochemical changes in the OT in the context of neurological disorders, specifically neurodegenerative disorders. By overviewing the vast literature on the neurochemical composition of the OT, this review will serve to aid future research into the neurobiology of the ventral striatum.

Keywords: Striatum, sensory, motivation, behavior, dopamine, acetylcholine, GABA, Parkinson’s, dementia, neurodegeneration

Graphical Abstract

The olfactory tubercle is a structure within the ventral striatum that has received relatively little attention. In this review, we provide a detailed summary of the olfactory tubercle’s rich neurochemistry, including the neurotransmitters present, as well as their associated receptors, transporters, and putative functions. In addition, we briefly review the literature pertaining to neurochemical changes in the olfactory tubercle in neurodegenerative disorders. This review will serve to aid future research into the olfactory tubercle and the ventral striatum as a whole.

graphic file with name nihms-1060843-f0002.jpg

Introduction

The ventral striatum is a collection of brain structures, including the nucleus accumbens, ventral pallidum, and the olfactory tubercle (OT) (De Olmos and Heimer 1999). While much attention has been devoted to the nucleus accumbens, a comprehensive understanding of the ventral striatum and its contributions to neurological diseases requires an appreciation for the complex neurochemical makeup of the ventral striatum’s other components, including the OT.

Before addressing the neurochemical composition of the OT, first some general orientation to this structure is warranted. The OT is located in the basal forebrain of all mammals, situated within the ventral striatum immediately beneath the ventral pallidum and nucleus accumbens (Fig. 1). Interestingly, it also serves as an olfactory cortex. Indeed, while the OT shares cellular features with other regions of the ventral striatum (Martin-Lopez et al. 2019; Ikemoto 2007; De Olmos and Heimer 1999), it also receives primary sensory input directly from the olfactory bulb, and has a distinctively non-uniform trilaminar structure (Pigache 1970; Millhouse and Heimer 1984). The OT’s layer 1 is the most superficial layer, also referred to as the molecular or plexiform layer, and is the site of olfactory bulb input (Scott et al. 1980). Layer 2, also referred to as the dense cell layer, is (unsurprisingly) densely populated with cell bodies. Layer 3 is the deepest and thickest layer, often called the multiform or polymorph layer, and is the site of association fiber input from the piriform cortex (Luskin and Price 1983; Schwob and Price 1984a). Throughout this review, we will simply refer to these as layers 1-3. The OT also possesses dense, distinctive clusters of granule cells, most often found in layer 3, referred to as the Islands of Calleja (ICj) (Millhouse and Heimer 1984; Millhouse 1987; Hsieh and Puche 2013; Adjei et al. 2013; Adjei and Wesson 2015).

Figure 1. The localization of the olfactory tubercle in a rodent brain.

Figure 1.

(Top) Diagram of a rat brain, including a coronal section indicating the nucleus accumbens, ventral pallidum, and olfactory tubercle (beige shaded region). (Bottom) Enlarged view of the olfactory tubercle, with layer 1 depicted in pink, layer 2 depicted in orange, and layer 3 depicted in yellow. An island of Calleja is depicted in red. Some examples of different cell types found in the olfactory tubercle are illustrated. Based in part upon (Xiong and Wesson 2016). M, medial; D, dorsal; MSN, medium spiny neuron; GC, granule cell; CIN, cholinergic interneuron.

Most of our current understanding of OT neurochemistry comes from classic autoradiography, immunohistochemistry, and in situ hybridization studies characterizing expression of neurotransmitters and receptors throughout the whole rodent brain. While these studies provided important data on the neurochemical organization of the brain, they often did not focus on the OT, and therefore do not always provide information regarding layer-, cell-type-, and cell-compartment-specific expression in the OT specifically. We thus avoid being overly declarative in stating a receptor does or does not exist in specific OT cells (for instance) but refer to whether or not a method uncovered it. Similarly, minimal information is available regarding the physiological role of each of these systems for OT function. That stated, this review will provide a primary resource for our current knowledge of OT neurochemistry by summarizing its neurochemical composition. This includes its neurotransmitters, neuromodulators, and hormones present, as well as the receptors and transporters involved in each system. We also provide brief discussion on the functional roles of these systems. Finally, we discuss the ways in which the neurochemical composition of the OT may change in cases of neurological disease. Throughout this review, we identify several knowledge gaps, which will be important to fill in order to advance our understanding of this brain system. Notably, this review is not meant to cover all known aspects of the OT, and readers are encouraged to seek other content from these sources which provide additional information on the OT’s anatomy and physiology (Wesson and Wilson 2011; Ikemoto 2007; Xiong and Wesson 2016; Millhouse and Heimer 1984; Scott et al. 1980; Murata et al. 2015; Wieland et al. 2015; Schwob and Price 1984b; Payton et al. 2012; Carlson et al. 2014; Carlson et al. 2018; Gadziola et al. 2015; Gadziola and Wesson 2016; Agustín-Pavón et al. 2014).

GABA

Sources of GABA in the OT

Like other striatal regions, the OT is composed primarily of GABAergic neurons, including local interneurons and medium spiny neurons (MSNs, also called ‘spiny projection neurons’) which may provide long-range projections outside of the OT itself, including into other striatal regions and into midbrain structures (Millhouse and Heimer 1984; Zhang et al. 2017b). While MSNs are located in all layers of the OT, many are packed densely into layer 2. Granule cells within the ICj are also GABAergic, exhibiting strong immunoreactivity for GAD65 and GAD67 (Riedel et al. 2002; Hsieh and Puche 2013; Krieger et al. 1983a). These cells possess small, thin axons and dendrites, which appear to primarily project locally within the ICj (Millhouse 1987; Millhouse and Heimer 1984), suggesting they mostly provide intra-ICj GABAergic signaling.

GABAergic interneurons in the OT are immunoreactive for multiple interneuron markers including parvalbumin (Zahm et al. 2003; Koos and JM 1999), calretinin (Riedel et al. 2002; Seifert et al. 1998; Helboe et al. 2015), calbindin (Riedel et al. 2002; Seifert et al. 1998; Helboe et al. 2015), vasoactive intestinal peptide (VIP) (Brunjes et al. 2011), neuropeptide Y (NPY) (Woodhams et al. 1985; Ubeda-Bañon et al. 2008), somatostatin (Finley et al. 1978), reelin (Ramos-Moreno et al. 2006; Martin-Lopez et al. 2019), and cholecystokinin (CCK) (Brunjes et al. 2011). Calbindin-expressing interneurons are scarce within the OT, and are found primarily in layer 3, with some cells also in layer 2 (Martin-Lopez et al. 2019). Notably, calbindin-immunoreactive fibers in the OT tend to be spatially segregated from areas which are void of tyrosine hydroxylase (TH) (Riedel et al. 2002). Parvalbumin-expressing interneurons are also found throughout the OT (Zahm et al. 2003), with the majority of these neurons in layer 3, and some in layer 2 (Martin-Lopez et al. 2019). The majority of reelin-immunoreactive neurons are found in layer 1 in the most medial portion of the OT (Martin-Lopez et al. 2019). While some CCK-positive neurons are found within the OT (Brunjes et al. 2011), the majority of CCK immunoreactivity in the OT appears in terminal fibers arising from other regions, including the ventral tegmental area (VTA) (Záborszky et al. 1985).

GABA receptor expression in the OT

There are two major classes of GABA receptors in the brain (GABAA and GABAB), both of which are abundantly expressed within the OT (Table 1). GABAA receptors are ligand-gated Cl channels composed of 5 homologous or heterologous subunits, of which there are 19 varieties (Olsen and Sieghart 2008). The rat OT shows strong diffuse immunoreactivity for the α2, α4, and β3 subunits, moderate immunoreactivity for the δ subunit, and weak immunoreactivity for other GABAA subunits (Pirker et al. 2000; Schwarzer et al. 2001). Strong immunoreactivity for the α1 and β2 subunits is restricted to the ventral pallidum and possibly the ICj, but is weak to moderate throughout the rest of the OT (Schwarzer et al. 2001). At the level of individual soma, interneurons show immunoreactivity for α1, β2, α5, and δ, while MSNs are immunoreactive for γ3 (Schwarzer et al. 2001). Neuronal processes exhibit immunoreactivity for the α1 and β2 subunits (Schwarzer et al. 2001).

Table 1.

Relative expression of GABA receptors in the OT

Class Receptor type Subunit mRNA Protein
GABAA Ligand-gated Cl channel alpha1 +++ (Duncan et al. 1995) +++ ICj*, + L1-3 (Schwarzer et al. 2001; Pirker et al. 2000)
alpha2 +++ L1-3 (Schwarzer et al. 2001; Pirker et al. 2000)
alpha3 + (Schwarzer et al. 2001)
alpha4 +++ L1-3 (Schwarzer et al. 2001; Pirker et al. 2000)
alpha5 + (Schwarzer et al. 2001)
alpha6
beta1 - (Zhang et al. 1990) + (Schwarzer et al. 2001)
beta2 +++ (Duncan et al. 1995), + (Zhang et al. 1990) ++ ICj*, + L1-3 (Schwarzer et al. 2001; Pirker et al. 2000)
beta3 +++ L2-3 (Zhang et al. 1990) +++ L1-3 (Schwarzer et al. 2001; Pirker et al. 2000)
gamma1 - (Schwarzer et al. 2001)
gamma2 ++ (Duncan et al. 1995) + (Schwarzer et al. 2001)
gamma3 - (Schwarzer et al. 2001)
delta ++ L1-3 (Schwarzer et al. 2001)
GABAB GPCR 1a +++ (Charles et al. 2001)
1b +++ (Charles et al. 2001)
2 + (Durkin et al. 1999) +++ (Charles et al. 2001)

- none, + modest, ++ moderate, +++ high

Cells absent of data reflect no literature available/found, or region not specified. All data cited is from non-human animal models.

*

limited data to confirm the original source, but likely

L1-3, Layers 1-3.

GABAB receptors are G-protein-coupled receptors (GPCRs) that regulate Ca2+ and K+ channels. There are 3 GABAB receptor genes (GABAB1a, GABAB1b, and GABAB2), which heterodimerize to form a functional GABAB receptor (Kaupmann et al. 1998; Marshall et al. 1999). GABAB2 receptor mRNA is present at weak levels throughout the OT, but is completely absent from the ICj (Durkin et al. 1999). The OT also shows strong immunoreactivity for all subunits throughout all layers, which appears to be strongest in layer 2 for the GABAB1b and GABAB2 subunits (Charles et al. 2001).

GABA transporters in the OT

Of the 4 GABA transporters, GAT1 and GAT3 are the two expressed in the central nervous system (for review, see (Zhou and Danbolt 2013)). GAT1 is primarily found presynaptically in GABAergic neurons, as well as some astrocytes (Minelli et al. 1995). In the OT, GAT1 mRNA is found at high levels in layer 1, and weakly in the ICj (Durkin et al. 1995). GAT3 expression is mostly restricted to astrocytic processes throughout the brain, although the OT has not been specifically investigated (Minelli et al. 1996).

Functional role of GABA signaling in the OT

In addition to its role as the major inhibitory neurotransmitter in the brain, GABA also plays an important role as a neural migration guidance cue during early postnatal development of the ICj (Hsieh and Puche 2015). ICj neurons are derived from subventricular zone neurons which migrate along a ventral pathway to the ventral striatum during the first postnatal week (De Marchis et al. 2004). Migrating neurons exclusively express the α3 subunit of the GABAA receptor during migration, and switch to expressing the α1 and α5 subunits once they form the ICj (Hsieh and Puche 2015). Further, disruption of GABA signaling with muscimol results in slower migration, indicating the functional importance of GABA signaling through GABAA receptors during the development of the ICj (Hsieh and Puche 2015). Similarly, mRNA for the the β1 and β2 subunits of the GABAA receptor are transiently expressed in the OT during development, suggesting they may also play a role in the development of the OT (Zhang et al. 1991).

Glutamate

Sources of glutamate in the OT

One major source of glutamatergic input to the OT comes from olfactory bulb projections to OT layer 1 (White 1965; Fuller and Price 1988), primarily the anterolateral portion (Scott et al. 1980). These terminals are easily visualized by their calretinin immunoreactivity (Martin-Lopez et al. 2019). While mitral and tufted cells may both target the OT (Igarashi et al. 2012), the majority of inputs come from tufted cells (Scott et al. 1980; Nagayama et al. 2010; Scott 1981). Another major glutamatergic projection to the OT comes from the piriform cortex, which targets primarily the anterolateral OT (Luskin and Price 1983; Carriero et al. 2009; Schwob and Price 1984b; White et al. 2019). The bulk of this projection originates in the ventral-caudal aspect of the anterior piriform cortex (White et al. 2019) and targets all layers of the OT, with the densest fibers in layer 2 (Schwob and Price 1984b; Schwob and Price 1984a; White et al. 2019). Aside from these major sources of glutamatergic input, midbrain dopamine (DA) neurons co-release glutamate in the OT(Wieland et al. 2014; Mingote et al. 2015). Glutamate co-release also likely occurs from cholinergic interneurons (CINs) and brainstem serotonergic neurons (Gras et al. 2002; Trudeau and El Mestikawy 2018), but the functional role for this in the OT is not known.

Glutamate receptor expression in the OT.

There are two major classes of glutamate receptors in the brain: ionotropic and metabotropic. The OT abundantly expresses glutamate receptors from both classes (Table 2). Ionotropic glutamate receptors include AMPA receptors (AMPARs) and Kainate receptors, which are ligand-gated ion channels, and NMDA receptors (NMDARs), which in addition to being ligand-gated are voltage-gated by a Mg2+ plug. Autoradiographic analysis of each family of glutamate receptors revealed high expression of each of these types of glutamate receptor in the OT (Albin et al. 1992). AMPARs are homo- or hetero-tetramers assembled from a combination of four subunits, which are abbreviated as GluA1-4 (for review see (Greger et al. 2017)). AMPARs are present at high levels in the OT, as evidenced by autoradiographic (Albin et al. 1992) and immunohistochemical studies (Petralia and Wenthold 1992). Specifically, GluA1-3 protein is present at high levels, and GluA4 at moderate levels, in the OT (Petralia and Wenthold 1992), but cell-type- and layer-specific expression of AMPAR subunits has not been specifically investigated.

Table 2.

Relative expression of glutamate receptors in the OT

Class Receptor type Subunit/receptor mRNA Protein
NMDAR Ligand/voltage-gated ion channel GluN1 +++ (Standaert et al. 1999)
GluN2A +++ (Standaert et al. 1999) ++ (Wang et al. 1995)
GluN2B +++ (Standaert et al. 1999) +++ (Wang et al. 1995)
GluN2C - (Standaert et al. 1999)
GluN2D + (Standaert et al. 1999)
GluN3A
GluN3B
AMPAR Ligand-gated ion channel GluA1 (prev. GluR1) + (Pellegrini-Giampietro et al. 2006) +++ (Petralia and Wenthold 1992)
GluA2 (prev. GluR2) ++ (Pellegrini-Giampietro et al. 2006) +++ (Petralia and Wenthold 1992)
GluA3 (prev. GluR3) ++ (Pellegrini-Giampietro et al. 2006) +++ (Petralia and Wenthold 1992)
GluA4 (prev. GluR4) ++ (Petralia and Wenthold 1992)
Kainate Ligand-gated ion channel GluK1 (prev. GluR5) +++ ICj; - OT (Bischoff et al. 1997)
GluK2 (prev. GluR6) +++ L2-3 (Bischoff et al. 1997)
GluK3 (prev. GluR7) ++ L2-3 (Bischoff et al. 1997)
GluK4 (prev. KA1) + L2-3 (Bischoff et al. 1997)
GluK5 (prev. KA2) +++ L2 (Bischoff et al. 1997)
mGluR GPCR mGluR1 (group 1) ++ (Lavreysen et al. 2003; Mutel et al. 2000) +++ ICj (Wada et al. 1998)
mGluR2 (group 2) ++ L1-3 (Mutel et al. 2000; Wada et al. 1998)
mGluR3 (group 2) +++ L1-3 (Tamaru et al. 2001)
mGluR4 (group 3) ++ (Ohishi et al. 1995)
mGluR5 (group 1) ++ L1-3 (Mutel et al. 2000; Wada et al. 1998)
mGluR6 (group 3)
mGluR7 (group 3) ++ (Ohishi et al. 1995) mGluR71 +++ L1, ICj (Kinoshita et al. 1998; Wada et al. 1998)
mGluR7b +++ ICj (Kinoshita et al. 1998)
mGluR8 (group 3) ++ L1 (Wada et al. 1998)

- none, + modest, ++ moderate, +++ high

Cells absent of data reflect no literature available/found, or region not specified. All data cited is from non-human animal models.

L1-3, Layers 1-3.

Kainate receptors are also homo- or hetero-tetramers assembled from a combination of five subunits, which are abbreviated as GluK1-5 (previously known as GluR5-7 and KA1-2; for review, see (Lerma and Marques 2013)). GluK1 mRNA is high in the granule cells of the ICj and is almost absent from the rest of the OT, while GluK2 is high in the cells just surrounding the ICj (Bischoff et al. 1997). Layer 2 contains high levels of GluK2 and GluK5 mRNA, and moderate levels of GluK3 mRNA (Bischoff et al. 1997). Similarly, layer 3 contains high levels of GluK2 and moderate levels of GluK3 (Bischoff et al. 1997).

NMDA receptors are tetramers composed of four subunits, which are abbreviated as GluN1-3. One NMDA receptor is composed of two GluN1 along with two GluN2 (GluN2A-D) or GluN3 (GluN3A-B) subunits (for review, see (Paoletti et al. 2013)). An in situ hybridization study of the GluN1 and GluN2 subunits revealed high levels of expression of GluN1, GluN2A, and GluN2B, low expression of GluN2D, and no expression of GluN2C in the OT (Standaert et al. 1999). At the level of protein, the OT contains GluN2A at intermediate levels and GluN2B at high levels (Wang et al. 1995).

There are 8 known variants of metabotropic glutamate receptors (mGluRs), which are GPCRs. mGluRs are divided into groups based on the specificity of agonists/antagonists and their downstream effectors: Group 1 mGluRs include mGluR1 and mGluR5, both of which are found in the OT according to autoradiographic (Lavreysen et al. 2003; Mutel et al. 2000) and immunohistochemical evidence (Cowen et al. 2005). Group 2 mGluRs include mGluR2 and mGluR3, which are also both found in the OT. Specifically, mGluR2 is expressed weakly in layer 1 and moderately in layers 2 and 3, while mGluR3 is expressed strongly in all layers of the OT (Tamaru et al. 2001; Wada et al. 1998). Group 3 mGluRs include mGluRs 4, 6, 7, and 8. mGluR7a and 8 are located in the axon terminals of projecting olfactory bulb neurons in layer 1 of the OT (Wada et al. 1998). mGluR4 and mGluR7 mRNA is expressed in the OT (Ohishi et al. 1995), and mGluR7a and 7b protein is expressed within the ICj (Kinoshita et al. 1998).

Glutamate transporters in the OT

There are five known mammalian excitatory amino acid transporters (EAAT1-5; for review see (Zhou and Danbolt 2013)), with EAAT1-4 expressed highly in the brain. EAAT 1 and 2 are strongly expressed selectively in glia (Danbolt et al. 1992; Levy et al. 1993) and EAAT2 mRNA (also known as GLT-1 in humans) is detected in the OT (Berger et al. 2005). EAAT3 is expressed in neurons throughout the whole brain (though this has not been specifically examined in the OT), and is localized to the soma and dendrites, but not axon terminals (Shashidharan et al. 1997). EAAT4 is most highly expressed in the cerebellum, but is also found throughout the rodent forebrain, including the ventral striatum (though the OT was not specifically investigated) (Massie et al. 2008).

Vesicular glutamate transporters 1-3 (VGLUT1-3) are present in the OT (Herzog et al. 2004; Fremeau et al. 2001; Gras et al. 2002). VGLUT1 and 2 are expressed in axon terminals throughout the brain, and exhibit differential expression patterns in the OT, with VGLUT1 highly expressed in layers 1 and 2, while VGLUT2 is expressed at lower levels primarily in layer 3 (Fremeau et al. 2001). VGLUT3 is not expressed as widely throughout the brain, but is nevertheless seen at high levels in the OT, with the highest levels of expression in layers 1 and 3 (Herzog et al. 2004). Interestingly, VGLUT3 is co-expressed with acetylcholine (ACh) and serotonin (5-HT) transporters in the striatum, suggesting that some CINs in the striatum and serotonergic brainstem neurons may co-release glutamate (Gras et al. 2002; Trudeau and El Mestikawy 2018).

Functional role of glutamate in the OT and interactions with other systems

Few studies have investigated the precise mechanisms by which glutamate neurotransmission modulates activity in the OT. White et al. (2019) explored the influence of glutamatergic input from the piriform cortex to the OT, and found that optogenetic stimulation of association fiber terminals in the mouse OT modulates odor-evoked single-unit responses of OT neurons (White et al. 2019). Specifically, odor-excited units are less excited with association fiber stimulation, while odor-suppressed units are less suppressed with stimulation, indicating that these glutamatergic projections may influence the encoding of odors in the OT (White et al. 2019). To better understand the role of glutamate corelease from midbrain dopamine neurons, Wieland et al. (2014) recorded from CINs in acute slices while optogenetically stimulating VTA projections to the OT. This revealed a glutamatergic aspect of the light-evoked response, mediated by both NMDAR and AMPAR components. This glutamate co-release along with DA from the midbrain exerts control over the firing patterns of OT CINs, which may play an important role in olfactory stimulus selection, and allows midbrain dopamine neurons to modulate activity in the OT on fast timescales as well as slow neuromodulatory timescales (Wieland et al. 2014). Mingote et al (2015) recorded from putative MSNs in the OT while stimulating VTA projections and found that the glutamatergic component was completely blocked by AMPAR/Kainate antagonists (Mingote et al. 2015), in contrast to CINs which included an NMDAR component (Wieland et al. 2014).

Dopamine

Sources of Dopamine in the OT

DAergic neurons from the ventral tegmental area (VTA), which form the mesolimbic projection pathway, project to the OT as well as the nucleus accumbens (Watabe-Uchida et al. 2012; Mingote et al. 2015; Zhang et al. 2017a; Del-Fava et al. 2007). These fibers begin to innervate the OT as early as mouse embryonic day 13, and progressively arborize from layer 3 to layer 1 in an ‘inside-out’ gradient (Martin-Lopez et al. 2019). Specifically, DAergic VTA projection fibers innervate the medial OT from the posteromedial VTA, the interfasicular nucleus, and central linear nucleus (Ikemoto 2007). Synaptic connectivity between DAergic projections and OT neurons is increasingly evident (Watabe-Uchida et al. 2012; Ikemoto 2010; Mingote et al. 2015). While DA is certainly a major transmitter in the OT, VTA DAergic neurons also co-release glutamate, thereby yielding more complex effects (Mingote et al. 2015; Wieland et al. 2014).

Dopamine receptor and transporter expression in the OT

DA receptors mediate the diverse functional effects of DA neurotransmission in the brain. There are five types of DA receptors (i.e., D1-D5) which are categorized into two families. The D1-like receptor family (D1 and D5) are GPCRs acting on Gs that ultimately increase intracellular cAMP while the D2 like receptor family (D2, D3, and D4) act on Gi and ultimately reduce cAMP (for review, please see (Missale et al. 1998; Surmeier et al. 2007). The OT contains D1, D2, and D3 receptors and is comprised primarily of D1- or D2-type DA receptor expressing MSNs (Murata et al. 2015) which make up 90-95% of all OT neurons (Le Moine and Bloch 1995; Soares-Cunha et al. 2016). In the OT, D1 and D3 are co-expressed among interneuron populations in the ICj (Le Moine and Bloch 1996; Huang et al. 1992; Diaz et al. 1997) whereas D2 and D3 seem to co-express to a much lesser extent (Bouthenet et al. 1991; Diaz et al. 1995; Murray et al. 1994).

The DA transporter (DAT) is an integral membrane protein allowing for homeostasis of synaptic DA (Torres et al. 2003; Sonders et al. 1997). The OT has robust DAT immunoreactivity, which is consistent with its midbrain DA fiber innervation (Freed et al. 1995; Ciliax et al. 1995; Vaughan et al. 1996; Revay et al. 1996; Vastenhouw et al. 2007).

Functional role of Dopamine signaling in the OT

DA is an important neurotransmitter for motivation (Wise 2004), learning (Redgrave et al. 1999; Pan 2005), reinforcement (Ikemoto 2007), and driving many behavioral responses (Ikemoto 2007). In both rodents and humans, the OT is implicated in odor hedonics and valence, as well as motivated behaviors (DiBenedictis et al. 2015; FitzGerald et al. 2014; Gadziola and Wesson 2016; Gadziola et al. 2015; Murata et al. 2015; Wesson and Wilson 2010; Wesson and Wilson 2011; Zelano et al. 2005).

Rodents readily self-administer intracranial micro-injections of psychostimulants into the OT (Ikemoto 2003; Ikemoto 2005). Specifically to cocaine, intra-OT infusions (through indwelling cannulae) are more robustly self-administered than into either the nucleus accumbens or ventral pallidum (Ikemoto 2003). These findings suggest a role for the OT in mediating the primary reinforcing effects of psychostimulants, which are driven by decreased DA uptake. Furthermore, rodents more greatly self-administer cocaine into the medial portion of the OT versus the lateral, implying unique roles for these OT subregions (Ikemoto 2003). In vivo voltammetry within the OT (e.g., (Park et al. 2017)), or a related method to measure DA levels, may help uncover the basis for this interregional effect. DA release into the OT holds important influences on local OT neurons. For example, in acute slices of the OT, phasic DA release evokes a transient pause in CIN firing via D2 receptor activation (Wieland et al. 2014), which may be important for decision-making and behavior selection.

Consistent with DA’s role in learning and reinforcement in other brain regions, OT DA is associated with the formation of learned odor preferences (Zhang et al. 2017a). Additionally, DAergic neuronal ablation in the medial OT via 6-OHDA microinjections reduce preference of sexually naïve female mice to investigate male urinary odors, while having no effect on the ability to discriminate between the odors (DiBenedictis et al. 2014). Together, these findings suggest that DA in the OT is critical for both learned and innate odor preferences, as well as the primary reinforcing effects of psychostimulants.

Acetylcholine

Sources of acetylcholine in the OT

CINs are large tonically active neurons (Aosaki et al. 1994; Morris et al. 2004) that make up a small fraction of striatal cells, each with a vast synaptic arbor that sends cholinergic projections throughout the striatum (Hebb and Silver 1961; Lim et al. 2014). CINs have been implicated in regulating the activation and modulation of striatal pathways in mice (Centonze et al. 2003; Kaneko et al. 2000; Maurice et al. 2004; Tozzi et al. 2011; Cachope et al. 2012; Mamaligas and Ford 2016) and rats (Pisani et al. 2000). The matrix surrounding ventrostriatal MSNs is rich in cholinergic markers (Graybiel and Ragsdale 1978; Gerfen 1985; Alheid and Heimer 1988) and within the OT, the ICj are exceptionally rich in choline acetyltransferase (ChAT) and acetylcholinesterase (AChE) (Talbot et al. 1988). Cholinergic input to the OT arrives from the horizontal limb of the nucleus of the diagonal band of Broca, nucleus basalis of Meynert (nbM), as well as the laterodorsal tegmentum and the pedunclopontine nuclei in the brainstem (Mesulam et al. 1983; Lauterborn et al. 1993; Koulousakis et al. 2019; Dautan et al. 2014). Specifically, neurons from the horizontal limb of the nucleus of diagonal band of Broca project throughout the ventral striatum, but vertical limb neurons appear to selectively target the OT (Dautan et al. 2016).

Acetylcholine receptor expression in the OT

CINs influence a wide range of postsynaptic targets though the activation of muscarinic and nicotinic acetylcholine receptors (mAChRs and nAChRs, respectively). mAChR activation can either decrease or increase cell excitability, and by virtue of its almost exclusive terminal expression, provides a long-term modulatory role in neurotransmitter release probability (Lim et al. 2014). There are 5 mAChR subtypes, which are broadly grouped as either excitatory (M1, M3, and M5) or inhibitory (M2 and M4) based upon their GPCR effectors (Wess 1996; Lin et al. 2004). Autoradiographic analyses indicate that while the OT has the greatest muscarinic receptor representation within the striatum (Corte 1986), and all subtypes are expressed in the striatum (Yan et al. 2001), only M1, M2, and M4 are reported in the OT, with the largest representation of M4 (Vilaró et al. 1991; Weiner et al. 1990; Vilaró et al. 1992). Specifically, while M4 immunoreactivity is present throughout the OT, its expression is remarkably high within the ICj (Wirtshafter and Osborn 2004). Regional differences in mAChR subunit expression throughout the striatum would convey different functional properties, which should be a subject of future research. Compared to mAChRs, nAChR activation is rapid and always has an excitatory effect. nAchRs are ligand-gated ion channels that can be expressed both pre or post synaptically, and induce depolarization and excitability (Lim et al. 2014). The pentameric structure of nAChRs includes α and β subunits (α2-α10 and β2-β4) (Patrick et al. 1993; McGehee and Role 1995; Dani 2001). The most highly expressed nAChR subunits in the striatum, as indicated by in situ hybridization, are α4, α6, α7, β2,and β3, however all other subunits are present at lower levels (Quik et al. 2007). Specifically, the nAChR subunits α2, α4, α5, α6 and β2 have all been identified in the OT (Brown et al. 2007; Marubio et al. 2003; Drenan et al. 2008; Whiteaker et al. 2006; Metaxas et al. 2013; Ishii et al. 2005; Grady et al. 2002).

Acetylcholine transporter in the OT

ACh is synthesized by ChAT in the cytoplasm of cholinergic neurons and transported into synaptic vesicles by vesicular acetylcholine transporter (VAChT) (Whittaker 1988; Parsons et al. 1993; Usdin et al. 1995). While much of the striatum is rich in VAChT, the OT receives an especially dense innervation of VAChT positive fibers, with few VAChT positive neurons (Schäfer et al. 1994; Arvidsson et al. 1997; Gilmor et al. 1996; Ichikawa et al. 1997; Roghani et al. 1997; Donat et al. 2008; Sorger et al. 2009; Henny and Jones 2006). AChE hydrolyzes ACh in the synaptic cleft yielding choline and acetate, and choline uptake is carried out by the high-affinity choline transporter, which is highly expressed in the OT (Misawa et al. 2001).

Functional roles of acetylcholine in the OT and interactions with other systems

ACh differentially modulates synaptic inputs to the OT: In acute slices, stimulation of either LOT fibers in layer 1 or association fibers in layer 3 evokes extracellular field potentials in layer 2, but layer 3-evoked responses exhibit much more significant depression in the presence of cholinergic agonists carbachol and muscarine (Owen and Halliwell 2001). Further, the authors speculate that this could be due to M4 receptors, which are present at high levels in the OT, but additional work is needed to support this hypothesis (Owen and Halliwell 2001). There are also notable interactions with DA and ACh in the OT, as mentioned above. The α2 and β2 subunits co-localize with DA receptors in the OT, influencing nicotine-induced DA release (Metaxas et al. 2013; Grady et al. 2002). The α4 and α6 subunits are also involved in nicotine-induced DA release in the OT (Drenan et al. 2008; Drenan et al. 2010; Cohen et al. 2012). Additionally, decreased ACh turnover in the OT was found following cocaine self-administration (Smith et al. 2003), which may be an effect of early drug withdrawal. Given DA’s rapid control of OT cholinergic tone discussed above (Wieland et al. 2014), the behavioral significance of the interplay of DA along with ACh in the OT warrants further investigation.

Norepinephrine

Sources of norepinephrine in the OT

Norepinephrine (NE), also known as noradrenaline, is a catecholamine synthesized by noradrenergic neurons. The primary source of NE throughout the brain, and in the OT, comes from the locus coeruleus (LC) (Jones and Moore 1977; Guevara-Aguilar et al. 1982; Solano-Flores et al. 1980).

Norepinephrine receptor expression in the OT

NE receptors are broadly classified as either α1, α2, or β receptors, each of which has three subtypes (α1A-C, α2A-C, and β1-3) (for review, see (Strosberg 1993; Bylund 1992). α1 receptors are excitatory Gi/q-coupled GPCRs, while α2 receptors are inhibitory Gi/o-coupled GPCRs. There are somewhat conflicting reports on the representation of α NE receptors in the OT. Some report weak mRNA and protein expression for α1 receptors in rats and mice (Day et al. 1997; Strazielle et al. 1999), with even weaker α2 protein expression (Strazielle et al. 1999). In contrast, others report quite strong α2C mRNA expression (Scheinin et al. 1994) and weak α2C protein expression in rats (Rosin et al. 1996) in layers 2 and 3. β receptors are Gs-coupled GPCRs that bind NE and signal through adenylate cyclase stimulation, increasing cAMP and downstream Ca2+ (Strosberg 1993). There is evidence for moderate mRNA (Nicholas et al. 1993) and high protein expression of β1 and β2 receptors in the OT (Palacios and Kuhar 1982; Strazielle et al. 1999; Rainbow et al. 1984). The ICj exhibit particularly high β1 expression, with modest β2 expression as well (Rainbow et al. 1984).

Norepinephrine transporter in the OT

NE release is controlled by a balance of monoamine oxidase mediated transmitter degradation and vesicular monoamine transporter (Eiden et al. 2004; Torres et al. 2003). The NE transporter (NET) is confined to adrenergic neurons and it has little-to-no expression in the OT, similar to the dorsal striatum (Hipólide et al. 2005; Strazielle et al. 1999; Sanders et al. 2005). While this is somewhat surprising, it is possible that NE homeostasis in the OT is controlled by a both DAT and monoamine oxidase activity (Demarest and Moore 1979).

Functional roles of norepinephrine signaling in the OT

NE in many brain systems has known roles in learning and memory, as well as arousal and stress (Carter et al. 2010; Uematsu et al. 2017; Morilak et al. 2005). The type and frequency of stressors produces markedly different effects on brain NE levels (Hellriegel and D’Mello 1997). For instance, chronic social isolation results in increased NE levels in the OT of mice that display increased intraspecies aggression (Garris 2003), whereas acute restraint stress decreases NE levels in the OT of pigs (Piekarzewska et al. 2000). While species differences may contribute to these effects, it is apparent that NE in the OT is subject to modulation via different types of stressors. Beyond these studies, however, the functional roles of NE in the OT have not yet been identified.

Serotonin

Source of serotonin in the OT

The raphe nuclei are the site of production of 5-HT, and both dorsal and ventral subregions of the raphe project widely throughout the brain. The OT and ICj receive dense serotonergic innervation from the ascending serotonergic pathway, originating in the raphe nuclei (Hillegaart et al. 1990; Fallon 1983).

Serotonin receptors in the OT

The 5-HT neurotransmitter system contains 14 membrane-bound receptors, which are divided into 7 families. These receptors are differentially expressed throughout the brain, peripheral nervous system, and non-neuronal tissue. All 5-HT receptors are GPCRs except for the 5-HT3 receptor, which is a ligand-gated ion channel. Diverse receptor mechanisms and patterns of expression of 5-HT receptors contribute to the wide range of functions in the nervous system, but a lack of pharmacological specificity has posed a significant challenge to attributing specific functions to each receptor. A comprehensive discussion on the structural and functional profiles of these receptors, which is outside of the scope of this review, can be found here (Barnes and Sharp 1999; Hoyer et al. 2002).

Table 3 summarizes the profiles of mRNA (by in situ hybridization) and protein expression (by immunohistochemistry or autoradiography) of 5-HT receptors in the OT. Notably, 5-HT1B receptor mRNA (which is homologous to human 5-HT1A) is highly expressed in the OT, while protein levels are low (Bonaventure et al. 1998; Boschert et al. 1994; Voigt et al. 1991). This pattern of expression is also observed in the basal ganglia, while its major output regions express the converse (high protein and low mRNA) (Boschert et al. 1994; Voigt et al. 1991). This mismatch is in accordance with the established role of 5-HT1B as an autoreceptor, where its regulation of serotonergic reuptake is mediated by its interaction with the serotonin transporter (SERT) at presynaptic terminals (Hagan et al. 2012).

Table 3.

Relative expression of 5-HT receptors in the OT

Family GPCR effector Receptor mRNA Protein
1 Gi/o 5-HT1A +, ICj (Wright et al. 1995)
5-HT1B ++++ (Bonaventure et al. 1998; Boschert et al. 1994; Voigt et al. 1991) + (Boschert et al. 1994)
5-HT1D +++, L2 (Boschert et al. 1994; Bach et al. 1993) +++, L2 (Hadley and Halliwell 2010)
5-HT1E*
5-HT1F ++ (Bruinvels et al. 1994)
2 Gq 5-HT2A +++ (Mijnster et al. 1997) +++, L2 (López-Giménez et al. 1997)
5-HT2B
5-HT2C +++ (Wright et al. 1995)
3 n/a (ligand-gated ion channel) 5-HT3 +, ICj (Kilpatrick et al. 1988; Koyama et al. 2017)
4 Gs 5-HT4 ++++ (Vilaró et al. 1996) +++, OT & ICj (Grossman et al. 1993; Waeber et al. 1996)
5 Gi/o 5-HT5A + (Kinsey et al. 2001)
5-HT5B - (Kinsey et al. 2001)
6 Gs 5-HT6 ++++, OT & ICj (Ruat et al. 1993; Kinsey et al. 2001; Helboe et al. 2015) ++++, L1 (Gérard et al. 1997)
7 Gs 5-HT7 - (Kinsey et al. 2001)

- none, + modest, ++ moderate, +++ high, ++++ extremely high

*

5-HT1E receptor is not present in rats or mice.

Cells absent of data reflect no literature available/found, or region not specified. All data cited is from non-human animal models.

L1-3, Layers 1-3.

Layer 2 of the OT contains concordantly high protein and mRNA for both 5-HT1D (Bach et al. 1993; Hadley and Halliwell 2010) and 5-HT2A receptors (López-Giménez et al. 1997; Mijnster et al. 1997). Additionally, the OT has extremely dense 5-HT6 receptor mRNA expression (Helboe et al. 2015; Ruat et al. 1993; Kinsey et al. 2001), with protein that appears to be localized to layer 1 (Gérard et al. 1997), where the OT receives olfactory bulb input (Imamura et al. 2011).

The 5-HT system plays a critical role in neurodevelopment (Bonnin and Levitt 2011; Gaspar et al. 2003). In the OT and ICj, Waeber and colleagues (1996) reported that 5-HT4 receptor expression is low in the prenatal brain, increases significantly during postnatal development to surpass other brain regions studied (including the hippocampus, caudate putamen, and septum), and plateaus in adulthood (Waeber et al. 1996; Grossman et al. 1993; Vilaró et al. 1996). Developmental and functional profiles of other 5-HT receptors in the OT are unknown.

Serotonin transporters in the OT

SERT is highly expressed in the OT relative to other brain regions, second only to the dorsal raphe (Chen et al. 1992; Choi et al. 2000; Ase et al. 2000). Given the high levels of 5-HT receptors in the OT, high expression of SERT is unsurprising and suggests a robust mechanism for the regulation of 5-HT neurotransmission in the OT. Additionally, the OT exhibits immunoreactivity for VMAT2 (Cliburn et al. 2017) and the low-affinity plasma membrane monoamine transporter (PMAT) (Dahlin et al. 2007).

Functional role of serotonin in the OT

Several different functions of 5-HT in the OT have been identified. One area of research focuses on the role of 5-HT in the regulation of mood and emotional reactivity, given the use of selective-serotonin reuptake inhibitors (SSRIs) for the treatment of depression and anxiety. Olfactory bulbectomy, which is used to model depressive-like behavior in rodents, produces overt intraspecies aggression and increased 5-HT levels in the OT compared to socially-isolated mice (Garris 2003), suggesting a role for OT 5-HT in aggressive behavior.

Widespread serotonergic innervation of the olfactory system suggests an important role for olfaction and odor-guided behavior (McLean and Shipley 1987; Petzold et al. 2009). However, conditional depletion of 5-HT in the adult forebrain (including the OT) has no effect on odor discrimination or conditioned odor learning (Carlson et al. 2016). While a role for 5-HT in olfactory system development cannot be ruled out by this work, more research is needed to define the contributions of 5-HT release into the OT on olfaction.

There is also evidence that 5-HT in the OT plays a role in the reinforcing effects of drugs of abuse (as also discussed in DA and opioids sections), neurotoxicity, and obesity. A rodent model of novelty-seeking behavior, which predicts sensation-seeking and substance abuse based on a heightened locomotor response (Kabbaj 2006), indicates that 5-HT6 receptor mRNA expression in the OT is inversely correlated to locomotor response to novelty (Ballaz et al. 2007), suggesting a possible role of OT 5-HT6 in this behavioral phenotype. 3,4-Methylenedioxymethamphetamine (MDMA, which reduces 5-HT reuptake through its interaction with SERT) induces activation of the immediate early gene Fos in the OT, an effect that is dependent on both D1 receptor and NMDAR signaling (Hashimoto et al. 1997). Further, intracranial self-administration of MDMA into the medial portion of the OT is behaviorally reinforcing (Shin et al. 2008). Additionally, neurotoxic doses of methamphetamine results in decreased 5-HT concentrations in combined OT + nucleus accumbens tissue (Sabol et al. 2001). The OT also exhibits decreased levels of 5-HT when treated with a combination of the anorectic drugs phentermine and fenfluramine (Phen/Fen, which enhances weight loss through reductions in SERT and depletion of axonal 5-HT), indicative of neurotoxicity (Lew et al. 1997). Finally, mice that are resistant to high-fat diet-induced obesity display decreased SERT binding in the OT compared to susceptible mice and mice on a low-fat diet (Huang et al. 2004). Together, these studies indicate that the 5-HT system is altered in the OT by exposure to drugs of abuse, diet-related alterations, and neurotoxicity.

Interactions between systems

The 5-HT system is highly interconnected with other neurotransmitter systems. In the OT, DA and 5-HT activity are closely coupled in a behaviorally-relevant manner, such that 5-HT2 receptor antagonists interact with DA-targeted typical antipsychotics in the OT to reduce negative symptoms of schizophrenia (Cools et al. 1992). There is high co-expression of 5-HT1B with the cannabinoid receptor CB1 mRNA (Hermann et al. 2002). There is also co-expression of the 5-HT2A receptor with both dynorphin (68%) and enkephalin (45%) in the OT, which differs from levels of co-expression reported in other areas of the ventral striatum (Mijnster et al. 1997).

5-HT also interacts with glutamatergic systems in the OT: local excitatory post-synaptic potentials are inhibited through agonism of the 5-HT1B/1D receptors (Hadley and Halliwell 2010; Owen and Halliwell 2001). Finally, GABA systems are modulated by SSRIs in the OT, where chronic sertraline treatment decreased GAD mRNA expression in the OT as well as other brain areas that contribute to emotional processing (e.g., nucleus accumbens, prefrontal cortex, and reticular formation) (Giardino et al. 1996). Taken together, 5-HT in the OT is poised to exert profound effects on physiology and behavior, but there are many gaps in our knowledge of the specific role of 5-HT in the OT.

Opioids

Source of endogenous opioids in the OT

Endogenous opioids (endorphins, enkephalins, and dynorphins) are locally synthesized from the transcription of protein-coding genes for their precursors, called prepropeptides. These prepropeptides undergo N-terminus cleaving upon translation, resulting in propeptides, followed by posttranslational modifications that produce the active neuropeptides. The rat OT contains dense expression of preprodynorphin and preproenkephalin in layer 2, weak expression in layer 3, and no expression in the ICj (Furuta et al. 2002; Harlan et al. 1987). There is very little (2-3%) co-expression of preprodynorphin and preproenkephalin in the OT (Furuta et al. 2002), in accordance with the expression patterns of their respective end products characterized in D1 and D2 receptor-expressing MSNs (see above section). Additionally, the most ventral part of layer 2 contains preprodynorphin, but not preproenkephalin (Furuta et al. 2002). Endogenous morphine-like immunoreactivity has also been detected in the cell bodies of the OT with moderate intensity (Laux et al. 2011).

Opioid receptors in the OT

There are three major Gi/o-coupled opioid receptors: endorphins share the strongest affinity for the mu opioid receptor (MOR), dynorphins for the kappa opioid receptor (KOR), and enkephalins for the delta opioid receptor (DOR), respectively (Waldhoer et al. 2004). Similar to other areas of the dorsal and ventral striatum, DOR and KOR mRNA and protein are densely distributed throughout the OT, whereas MOR mRNA and protein are sparsely distributed (Mansour et al. 1987; Mansour et al. 1994; Furuta et al. 2002).

Interactions with other systems

Throughout the dorsal and ventral striatum, D1-expressing MSNs primarily express dynorphin and D2-expressing MSNs primarily express enkephalin (Lobo et al. 2006; Surmeier et al. 1996; Le Moine and Bloch 1995). OT neurons projecting to the ventrolateral portion of the ventral pallidum (also shown by (Heimer et al. 1987)) contain preprodynorphin (60%) and preproenkephalin (40%), which is the reverse for other areas of the ventral striatum like the nucleus accumbens (Zhou et al. 2003). These are primarily non-overlapping cell populations expressing one or the other, with shared high co-expression of dopamine- and cAMP-regulated neuronal phosphoprotein (DARPP-32; 74-85%) and no co-expression with markers for interneurons such as parvalbumin, calretinin, ChAT, and somatostatin (Furuta et al. 2002), consistent with D1-expressing MSNs and D2-expressing MSNs.

Functional role of opioids in the OT

The opioid system in the OT may influence the reinforcing effects of drugs of abuse. Opioid receptor subtypes have opposing behavioral effects such that MOR agonists are rewarding and KOR agonists are not only aversive, but they also block the rewarding effects of MOR agonism (Mucha and Herz 1985; Funada et al. 1993). In tissue including the OT and nucleus accumbens, systemic MOR agonists increases levels of DA metabolites and KOR agonists blocks this effect (Funada et al. 1993). Additionally, voluntary ethanol consumption in rats decreases preproenkephalin and proenkephalin mRNA expression in the OT; this effect is prevented by treatment with naltrexone, which reduces ethanol consumption through both MOR and KOR antagonism (Oliva and Manzanares 2007; Cowen and Lawrence 2001).

Hormones and neurosteroids

Source and receptors of hormones and neurosteroids in the OT

Sex steroid hormones, particularly brain-derived estrogen, plays a prominent role supporting synaptic plasticity, transcriptional activity, and learning and memory (Tuscher et al. 2016; Saldanha et al. 2011). The mouse OT is equipped to synthesize estrogen locally due to high expression of aromatase (Stanić et al. 2014), the rate-limiting enzyme that catalyzes androgens into estrogens. There is also moderately high expression of estrogen receptor α (ERα), the G protein-coupled estrogen receptor, and androgen receptors in the OT and ICj (Stanić et al. 2014; Rainbow et al. 1982; Roselli and Stormshak 2012; Brailoiu et al. 2007). Other neurosteroids in the OT have been less thoroughly studied. There are relatively low levels of 5α-reductase activity, which converts testosterone into dihydrotestosterone, in the OT (Krieger et al. 1983b). Additionally, while progesterone in the OT has not yet been characterized, its metabolite allopregnanolone, which is a modulator of the GABAA receptor (Schumacher et al. 2014), is densely expressed in OT layer 2 (Saalmann et al. 2007).

Functional role of hormones and neurosteroids in the OT

Although sex steroid hormones have potent and far-reaching effects on behavior, development, and neural function, little is known about their specific effects in the OT. 17β-estradiol augments evoked firing of neurons in the OT upon stimulation of the posterior hypothalamus in cats (Cartas-Heredia et al. 1978). This is in line with findings from other brain regions displaying an excitatory response to estrogens, including striatal MSNs (Huang and Woolley 2012; Cao et al. 2018; Proaño et al. 2018).

Interactions between systems

Sex steroid hormones interact with virtually all other neurotransmitter systems. In the OT specifically, interactions with 5-HT, DA, and oxytocin (which is discussed in more detail below) have been noted. Briefly, 5-HT and estrogen interact in the OT via the 5-HT2A receptor, which is upregulated following the surge of estradiol that follows proestrus in rats (Sumner and Fink 1997). Additionally, raloxifene (which is a selective-estrogen receptor modulator) blocks the estradiol-induced increase in 5-HT2A receptor expression in the OT and D3 receptor binding in the ICj, perhaps contributing to estrogen’s effects on mood and mental state (Sumner et al. 2007; Landry et al. 2002).

Oxytocin

Source of oxytocin in the OT

The supraoptic nucleus, accessory magnocellular nuclei, and paraventricular nucleus of the hypothalamus (PVN) are the main sources of oxytocin in the mammalian brain (Sofroniew 1983; Swanson and Sawchenko 1983). Oxytocin-releasing neurons in the PVN send dense projections to the posterior pituitary and the ICj, as well as less-dense projections to limbic regions, the basal ganglia, cortex, and other olfactory regions (Knobloch et al. 2012; Choe et al. 2015; Prasada Rao and Kanwal 2004).

Oxytocin receptor expression in the OT

The oxytocin receptor is a GPCR that binds oxytocin in the presence of Mg2+ and cholesterol, and it exerts its action through its Gq effector protein (Gimpl and Fahrenholz 2001; Devost et al. 2008). Oxytocin signaling dynamically changes throughout the lifespan and across species (for review, see (Vaidyanathan and Hammock 2017). There are two primary patterns of oxytocin receptor expression in the brain—the “adult” pattern which emerges post-weaning and remains stable throughout adulthood, and the more transient “infant” pattern (Tribollet et al. 1989). The ICj displays a uniquely late pattern of oxytocin receptor expression that does not emerge until postnatal day 40 in the rat (Tribollet et al. 1989); however, see (Hammock and Levitt 2013). The OT contains both oxytocin receptor mRNA and protein (Vaccari et al. 1998; Choe et al. 2015), with the strongest expression in the ICj (Tribollet et al. 1988).

Functional role of oxytocin in the OT and interactions with other systems

Given the high density of estrogen receptors in these areas (see section on hormones & neurosteroids) and the appearance of oxytocin receptors in the ICj around puberty, oxytocin in the OT may contribute to sociosexual behavior through its interaction with sex steroid hormones. In fact, the ICj as well as the ventromedial hypothalamus (VMH) are unique “hot spots” in the brain where systemic estrogen increases oxytocin receptor binding (de Kloet et al. 1986; Krémarik et al. 1995). While estrogen-oxytocin interactions in the VMH have been established for their regulation of female sexual behavior (McCarthy et al. 1994), the functional significance of this interaction in the ICj is poorly understood. It is likely that odor-guided social behaviors are impacted, since the basolateral amygdala (which projects to the OT) exhibits ERα-dependent increases in oxytocin receptor binding following estrogen treatment and plays a role in a variety of social behaviors (Young et al. 1998). There are also testosterone-oxytocin interactions, since reduced oxytocin receptor binding in the OT of male rats is linked to both aging- and castration-induced decreases in testosterone levels (Arsenijevic et al. 1995; Arsenijevic and Tribollet 1998; Johnson et al. 1991; Tribollet et al. 1990).

There may also be interactions with oxytocin and DA in the OT, since an acute infusion of oxytocin into the OT (through indwelling cannulae) blocks cocaine-induced stereotyped sniffing behavior (Sarnyai et al. 1991).

Other neurochemicals

An array of neuropeptides and neurochemicals are expressed in the OT, especially interneurons as discussed in the GABA section. Neurotensin is highly expressed in the OT, specifically among DA D2-type receptor expressing MSNs and the ICj (Schroeder et al. 2019). Projections to the OT and ICj arising from the posteromedial cortical amygdala (which receives input from the accessory olfactory bulb) are immunoreactive for neuropeptide Y and Substance-P (as well as 5-HT and TH) (Ubeda-Bañon et al. 2008), suggesting a role for these peptides in rodent sociosexual behaviors. Autoradiography of corticotropin releasing factor receptors indicates that they are localized in the OT (Weathington et al. 2014). Further, orexin receptor mRNA is highly expressed in the OT as well as faint immunoreactivity found in fibers, suggesting local synthesis of orexin in the OT (Caillol et al. 2003).

Nitric oxide (NO) is a free radical signaling molecule that can act as a neurotransmitter in the central nervous system, where it plays a role in neural plasticity, neurogenesis, and excitotoxicity (for review, see (Zhou and Zhu 2009)). NO is synthesized from L-arginine by neuronal nitric oxide synthase (nNOS), which is abundant within the OT of rodents (Rodrigo et al. 1994) and many other species (Menéndez et al. 2006; Brüning et al. 1994a; Brüning et al. 1994b). Specifically, nNOS immunoreactivity is present within fibers in layer 1 (which likely arise from olfactory bulb mitral and tufted cells (Kosaka and Kosaka 2007)), neurons in layer 3, and fibers and granule cells within the ICj (Rodrigo et al. 1994; Bredt et al. 1990; Vincent and Kimura 1992). NO largely exerts its effects by activating soluble guanylyl cyclase, which is expressed at very high levels in the OT, especially in layer 2 cell bodies (Ding et al. 2004). Guanylyl cyclase activation by NO results in upregulation of cyclic GMP synthesis (Murad et al. 1978). This is observed in acute rat brain slices within the ICj but not the surrounding regions (De Vente et al. 1998), suggesting that NO exerts its effects locally, despite its ability to freely diffuse across cell membranes. The functional role of NO signaling in the OT is unknown, but in vitro studies of OT tissue suggest that NO may inhibit dopamine uptake (Pogun et al. 1994). In vivo studies are needed to improve our understanding of NO’s function in the OT.

Neurochemical alterations in the OT in the context of neurological disorders

Striatal dysfunction is a hallmark of several neurological disorders (Murray et al. 1995; Kreitzer and Malenka 2008; Foltz et al. 2004; McGregor and Nelson 2019; Gerfen et al. 1990; Ouchi 2001). Likewise, olfactory dysfunction is observed in both aging and in numerous neurological disorders (Doty 2017; Murphy 2019; Devanand et al. 2000; Doty et al. 1984; Wesson et al. 2010b). It follows that the OT, as a component of both neural systems, would also be vulnerable to neurological disorders in ways that would influence its neurochemical composition. Here we review a few selected reports uncovering an influence of neurological disease on the OT, which may arise due to insults upon the OT itself, or due to insults upon neural systems that innervate the OT (e.g., the mesolimbic DAergic system, forebrain cholinergic innervation) and thereby affect OT neurochemistry.

Parkinson’s disease and some other movement disorders are characterized by dysfunction in basal ganglia systems, which include the ventral striatum. Lewy bodies, a key pathological hallmark of Parkinson’s disease, are made of the aggregated protein α-synuclein and are found post-mortem within the OT of humans with Parkinson’s disease (Ubeda-Bañon et al. 2010). The number of Lewy bodies reported in the OT however is considerably less than that found in the olfactory bulb itself, but is comparable to Lewy body presence in other olfactory cortices (Ubeda-Bañon et al. 2010). This accumulation of Lewy bodies and the accumulation of phosphorylated α-synuclein can be generated pre-clinically in rodent models by a single injection of α-synuclein preformed fibrils into the olfactory bulb (Rey et al. 2016). In this seeding model, phosphorylated α-synuclein is detected in the OT just weeks after a single injection into the olfactory bulb, suggesting transsynaptic propagation of Parkinson’s pathology in the brain (Rey et al. 2016; Rey et al. 2018). The direct influence of α-synuclein aggregation in the OT on olfactory perception is unclear.

The nucleus basalis of Meynert (nbM) contains a large collection of cholinergic neurons which project into several structures, including the OT (Koulousakis et al. 2019). Lewy bodies are observed in the nbM in humans diagnosed as having dementia with Lewy bodies, and this corresponds to low levels of ChAT (Lippa et al. 1999). It is possible that in the context of Lewy body disease, the OT is subject to aberrant influx of ACh from the nbM, which may alter OT function and subsequently influence disease states.

Alzheimer’s disease is characterized by several pathological features, including the accumulation of amyloid-β aggregates known as plaques and the presence of neurofibrillary tangles. Early in disease progression, these pathological hallmarks are observed in the olfactory system (Braak and Braak 1991). Amyloid-β plaques are observed in the OT in mouse models of Alzheimer’s disease, including in models overexpressing human mutations in the amyloid precursor protein (Wesson et al. 2010a) and in mice co-expressing human mutations in both the amyloid precursor protein and presenilin-1 (Saiz-Sanchez et al. 2013). In the latter mouse model, somatostatin and calretinin expression decline alongside increases in amyloid-β burden (Saiz-Sanchez et al. 2013), suggesting that somatostatin and calretinin interneurons may be particularly vulnerable to amyloidosis and/or other aspects of Alzheimer’s pathogenesis. Whether these changes within the OT may contribute to olfactory dysfunction is yet to be determined.

As a final example, loss of neuromodulatory inputs to the OT are observed following head trauma. Reduced DAergic fiber innervation is detected in the OT in a rat model of mild traumatic brain injury (Haar et al. 2019). This is accompanied by OT neuroinflammation (i.e., gliosis) and heightened levels of phosphorylated tau, the microtubule associated protein (Haar et al. 2019). Interestingly, the reduction in DAT fibers following mild traumatic head injury were not observed in the nigrostriatal pathway, suggesting unique roles of mesolimbic fiber degeneration, including in the OT, in the cognitive and psychiatric effects of brain injury (Haar et al. 2019).

The above results indicate how the several neurological disorders, and/or their pathogens in experimental models, impact the OT. It is possible that these changes in the OT may influence features and possibly the progression of these disorders. Determining, in a causal manner, whether the OT itself is a direct contributor to clinical manifestations of any of these disorders is an important major goal for future investigations.

Summary

Here we summarized the rich neurochemical composition of the OT, including the neurotransmitters, neuromodulators, and hormones present. We also addressed what receptors and transporters are involved in each system as well as their putative functional roles. Finally, we briefly reviewed select literature regarding neurochemical changes in the OT in the context of neurological disorders, specifically neurodegenerative disorders. We predict this will serve to aid future research into the neurobiology of the ventral striatum.

Acknowledgements:

This work was supported by NIH grants R01DC014443, R01DC016519, R01DA049545 to DW, and NIH F32DC018232 to HC. The figure was re-drawn by Marco Bazelmans in BioRender (https://biorender.com/) on the basis of a draft provided by the author.

List of abbreviations

5-HT

5-hydroxytryptamine, serotonin

Ach

acetylcholine

AChE

acetylcholinesterase

AMPAR

α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor

AON

anterior olfactory nucleus

CCK

cholecystokinin

ChAT

choline acetyl transferase

CIN

cholinergic interneuron

DA

dopamine

DARPP-32

dopamine- and cAMP-regulated neuronal phosphoprotein

DOR

delta opioid receptor

EAAT

excitatory amino acid transporter

ERα

estrogen receptor α

GABA

γ-aminobutyric acid

GAD

glutamic acid decarboxylase

GAT

GABA transporter

GluA

subunits of the glutamate AMPA receptor

GluK

subunits of the glutamate Kainate receptor

GluN

subunits of the glutamate NMDA receptor

GPCR

G-protein-coupled receptor

ICj

Islands of Calleja

KOR

kappa opioid receptor

LC

locus coeruleus

mAChR

muscarinic acetylcholine receptor

MDMA

3,4-Methylenedioxymethamphetamine

mGluR

metabotropic glutamate receptor

MOR

mu opioid receptor

MSN

medium spiny neuron

mAChR

muscarinic acetylcholine receptor

nAChR

nicotinic acetylcholine receptor

nbM

nucleus basalis of Meynert

NE

norepinephrine

NET

norepinephrine transporter

NMDAR

N-methyl-D-aspartate receptor

NO

nitric oxide

nNOS

neuronal nitric oxide synthase

NPY

neuropeptide Y

OT

olfactory tubercle

PMAT

plasma membrane monoamine transporter

PVN

paraventricular nucleus of the hypothalamus

SERT

serotonin transporter

SSRI

selective serotonin reuptake inhibitor

TH

tyrosine hydroxylase

VAChT

vesicular acetylcholine transporter

VGLUT

vesicular glutamate transporter

VIP

vasoactive intestinal peptide

VMAT2

vesicular monoamine transporter 2

VMH

ventromedial hypothalamus

VTA

ventral tegmental area.

Footnotes

Conflicts of Interest: The authors have no perceived or real conflicts of interest to declare.

References

  1. Adjei S, Houck AL, Ma K, Wesson DW (2013) Age-dependent alterations in the number, volume, and localization of islands of Calleja within the olfactory tubercle. Neurobiol. Aging 34, 2676–2682. [DOI] [PubMed] [Google Scholar]
  2. Adjei S, Wesson DW (2015) Laminar and spatial localization of the islands of Calleja in mice. Neuroscience 287, 137–143. [DOI] [PubMed] [Google Scholar]
  3. Agustín-Pavón C, Martínez-García F, Lanuza E (2014) Focal lesions within the ventral striato-pallidum abolish attraction for male chemosignals in female mice. Behav. Brain Res 259, 292–296. [DOI] [PubMed] [Google Scholar]
  4. Albin RL, Makowiec RL, Hollingsworth ZR, Dure LS, Penney JB, Young AB (1992) Excitatory amino acid binding sites in the basal ganglia of the rat: A quantitative autoradiographic study. Neuroscience 46, 35–48. [DOI] [PubMed] [Google Scholar]
  5. Alheid GF, Heimer L (1988) New perspectives in basal forebrain organization of special relevance for neuropsychiatric disorders: the striatopallidal, amygdaloid, and corticopetal components of substantia innominata. Neuroscience 27, 1–39. [DOI] [PubMed] [Google Scholar]
  6. Aosaki T, Tsubokawa H, Ishida A, Watanabe K, Graybiel AM, Kimura M (1994) Responses of tonically active neurons in the primate’s striatum undergo systematic changes during behavioral sensorimotor conditioning. J. Neurosci 14, 3969–3984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Arsenijevic Y, Dreifuss J, Vallet P, Marguerat A, Tribollet E (1995) Reduced binding of oxytocin in the rat brain during aging. Brain Res. 698, 275–279. [DOI] [PubMed] [Google Scholar]
  8. Arsenijevic Y, Tribollet E (1998) Region-specific effect of testosterone on oxytocin receptor binding in the brain of the aged rat. Brain Res. 785, 167–70. [DOI] [PubMed] [Google Scholar]
  9. Arvidsson U, Riedl M, Elde R, Meister B (1997) Vesicular acetylcholine transporter (VAChT) protein: A novel and unique marker for cholinergic neurons in the central and peripheral nervous systems. J. Comp. Neurol 378, 454–467. [PubMed] [Google Scholar]
  10. Ase AR, Strazielle C, Hébert C, Botez MI, Lalonde R, Descarries L, Reader TA (2000) Central Serotonin System in Dystonia Musculorum Mutant Mice: Biochemical, Autoradiographic and Immunocytochemical Data. Synapse 37, 179–193. [DOI] [PubMed] [Google Scholar]
  11. Bach AW, Unger L, Sprengel R, Mengod G, Palacios J, Seeburg PH, Voigt MM (1993) Structure functional expression and spatial distribution of a cloned cDNA encoding a rat 5-HT1D-like receptor. J. Recept. Res 13, 479–502. [DOI] [PubMed] [Google Scholar]
  12. Ballaz SJ, Akil H, Watson SJ (2007) Analysis of 5-HT6 and 5-HT7 receptor gene expression in rats showing differences in novelty-seeking behavior. Neuroscience 147, 428–438. [DOI] [PubMed] [Google Scholar]
  13. Barnes NM, Sharp T (1999) A review of central 5-HT receptors and their function. Neuropharmacology 38, 1083–1152. [DOI] [PubMed] [Google Scholar]
  14. Berger UV, DeSilva TM, Chen W, Rosenberg PA (2005) Cellular and subcellular mRNA localization of glutamate transporter isoforms GLT1a and GLT1b in rat brain by in situ hybridization. J. Comp. Neurol 492, 78–89. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Bischoff S, Barhanin J, Bettler B, Mulle C, Heinemann S (1997) Spatial distribution of kainate receptor subunit mRNA in the mouse basal ganglia and ventral mesencephalon. J. Comp. Neurol 379, 541–62. [DOI] [PubMed] [Google Scholar]
  16. Bonaventure P, Voorn P, Luyten WH, Jurzak M, Schotte A, Leysen JE (1998) Detailed mapping of serotonin 5-HT1B and 5-HT1D receptor messenger RNA and ligand binding sites in guinea-pig brain and trigeminal ganglion: clues for function. Neuroscience 82, 469–84. [DOI] [PubMed] [Google Scholar]
  17. Bonnin A, Levitt P (2011) Fetal, maternal, and placental sources of serotonin and new implications for developmental programming of the brain. Neuroscience 197, 1–7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Boschert U, Aït. Amara D, Segu L, Hen R. (1994) The mouse 5-hydroxytryptamine 1B receptor is localized predominantly on axon terminals. Neuroscience 58, 167–182. [DOI] [PubMed] [Google Scholar]
  19. Bouthenet ML, Souil E, Martres MP, Sokoloff P, Giros B, Schwartz JC (1991) Localization of dopamine D3 receptor mRNA in the rat brain using in situ hybridization histochemistry: comparison with dopamine D2 receptor mRNA. Brain Res. 564, 203–219. [DOI] [PubMed] [Google Scholar]
  20. Braak H, Braak E (1991) Neuropathological stageing of Alzheimer-related changes. Acta Neuropathol. 82, 239–259. [DOI] [PubMed] [Google Scholar]
  21. Brailoiu E, Dun SL, Brailoiu GC, Mizuo K, Sklar LA, Oprea TI, Prossnitz ER, Dun NJ (2007) Distribution and characterization of estrogen receptor G protein-coupled receptor 30 in the rat central nervous system. J. Endocrinol 193, 311–321. [DOI] [PubMed] [Google Scholar]
  22. Bredt DS, Hwang PM, Snyder SH (1990) Localization of nitric oxide synthase indicating a neural role for nitric oxide. Nature 347, 768–770. [DOI] [PubMed] [Google Scholar]
  23. Brown RWB, Collins AC, Lindstrom JM, Whiteaker P (2007) Nicotinic α5 subunit deletion locally reduces high-affinity agonist activation without altering nicotinic receptor numbers. J. Neurochem 103, 204–215. [DOI] [PubMed] [Google Scholar]
  24. Bruinvels AT, Landwehrmeyer B, Gustafson EL, Durkin MM, Mengod G, Branchek TA, Hoyer D, Palacios JM (1994) Localization of 5-HT1B, 5-HT1Dα, 5-HT1E and 5-HT1F receptor messenger RNA in rodent and primate brain. Neuropharmacology 33, 367–386. [DOI] [PubMed] [Google Scholar]
  25. Brüning G, Funk U, Mayer B (1994a) Immunocytochemical localization of nitric oxide synthase in the brain of the chicken. Neuroreport 5, 2425–2428. [DOI] [PubMed] [Google Scholar]
  26. Brüning G, Wiese S, Mayer B (1994b) Nitric oxide synthase in the brain of the turtle Pseudemys scripta elegans. J. Comp. Neurol 348, 183–206. [DOI] [PubMed] [Google Scholar]
  27. Brunjes PC, Kay RB, Arrivillaga JP (2011) The mouse olfactory peduncle. J. Comp. Neurol 519, 2870–2886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Bylund DB (1992) Subtypes of alpha-1 and alpha-2 adrenergic receptors. Eur. Neuropsychopharmacol 6, 832–839. [DOI] [PubMed] [Google Scholar]
  29. Cachope R, Mateo Y, Mathur BN, Irving J, Wang H-L, Morales M, Lovinger DM, Cheer JF (2012) Selective activation of cholinergic interneurons enhances accumbal phasic dopamine release: setting the tone for reward processing. Cell Rep. 2, 33–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Caillol M, Aioun J, Baly C, Persuy MA, Salesse R (2003) Localization of orexins and their receptors in the rat olfactory system: possible modulation of olfactory perception by a neuropeptide synthetized centrally or locally. Brain Res 960, 48–61. [DOI] [PubMed] [Google Scholar]
  31. Cao J, Willett JA, Dorris DM, Meitzen J (2018) Sex Differences in Medium Spiny Neuron Excitability and Glutamatergic Synaptic Input: Heterogeneity Across Striatal Regions and Evidence for Estradiol-Dependent Sexual Differentiation. Front. Endocrinol. (Lausanne). 9, 173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Carlson KS, Dillione M, Wesson DW (2014) Odor- and state-dependent olfactory tubercle local field potential dynamics in awake rats. J. Neurophysiol 111, 2109–2123. [DOI] [PubMed] [Google Scholar]
  33. Carlson KS, Gadziola MA, Dauster ES, Wesson DW (2018) Selective Attention Controls Olfactory Decisions and the Neural Encoding of Odors. Curr. Biol 28, 2195–2205.e4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Carlson KS, Whitney MS, Gadziola MA, Deneris ES, Wesson DW (2016) Preservation of essential odor-guided behaviors and odor-based reversal learning after targeting adult brain serotonin synthesis. eNeuro 3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Carriero G, Uva L, Gnatkovsky V, de Curtis M (2009) Distribution of the Olfactory Fiber Input Into the Olfactory Tubercle of the In Vitro Isolated Guinea Pig Brain. J Neurophysiol 101, 1613–1619. [DOI] [PubMed] [Google Scholar]
  36. Cartas-Heredia L, Guevara-Aguilar R, Aguilar-Baturoni HU (1978) Oestrogenic influences on the electrical activity of the olfactory pathway. Brain Res. Bull 3, 623–30. [DOI] [PubMed] [Google Scholar]
  37. Carter ME, Yizhar O, Chikahisa S, Nguyen H, Adamantidis A, Nishino S, Deisseroth K, de Lecea L (2010) Tuning arousal with optogenetic modulation of locus coeruleus neurons. Nat. Neurosci 13, 1526–1533. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Centonze D, Grande C, Usiello A, Gubellini P, Erbs E (2003) Receptor subtypes involved in the presynaptic and postsynaptic actions of dopamine on striatal interneurons. J. Neurosci 23, 6245–6254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Charles K ., Evans M ., Robbins M ., Calver A ., Leslie R ., Pangalos M . (2001) Comparative immunohistochemical localisation of GABAB1a, GABAB1b and GABAB2 subunits in rat brain, spinal cord and dorsal root ganglion. Neuroscience 106, 447–467. [DOI] [PubMed] [Google Scholar]
  40. Chen HT, Clark M, Goldman D (1992) Quantitative autoradiography of 3H-paroxetine binding sites in rat brain. J. Pharmacol. Toxicol. Methods 27, 209–16. [DOI] [PubMed] [Google Scholar]
  41. Choe HK, Reed MD, Benavidez N, Montgomery D, Soares N, Yim YS, Choi GB (2015) Oxytocin Mediates Entrainment of Sensory Stimuli to Social Cues of Opposing Valence. Neuron 87, 152–163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Choi S-R, Hou C, Oya S, Mu M, Kung M-P, Siciliano M, Acton PD, Kung HF (2000) Selective in vitro and in vivo binding of [125I]ADAM to serotonin transporters in rat brain. Synapse 38, 403–412. [DOI] [PubMed] [Google Scholar]
  43. Ciliax BJ, Heilman C, Demchyshyn LL, Pristupa ZB, Ince E, Hersch SM, Niznik HB, Levey AI (1995) The dopamine transporter: immunochemical characterization and localization in brain. J. Neurosci 15, 1714–1723. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Cliburn RA, Dunn AR, Stout KA, Hoffman CA, Lohr KM, Bernstein AI, Winokur EJ, et al. (2017) Immunochemical localization of vesicular monoamine transporter 2 (VMAT2) in mouse brain. J. Chem. Neuroanat 83–84, 82–90. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Cohen BN, Mackey EDW, Grady SR, Mckinney S, Patzlaff NE, Wageman CR, Mcintosh JM, Marks MJ, Lester HA, Drenan RM (2012) Nicotinic cholinergic mechanisms causing elevated dopamine release and abnormal locomotor behavior. Neuroscience 200, 31–41. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Cools AR, Prinssen E, Ellenbroek BA (1992) Interaction between aminergic systems in mesolimbic structures: the importance of 5-HT2, D2 and D1 receptors in the olfactory tubercle for the atypical profile of neuroleptics. Clin. Neuropharmacol 15 Suppl 1 Pt A, 618A–619A. [DOI] [PubMed] [Google Scholar]
  47. Corte R (1986) Muscarinic cholinergic receptor subtypes in the rat brain. I. Quantitative autoradiographic studies. Brain Res. 362, 227–238. [DOI] [PubMed] [Google Scholar]
  48. Cowen MS, Djouma E, Lawrence AJ (2005) The Metabotropic Glutamate 5 Receptor Antagonist 3-[(2-Methyl-1,3-thiazol-4-yl)ethynyl]-pyridine Reduces Ethanol Self-Administration in Multiple Strains of Alcohol-Preferring Rats and Regulates Olfactory Glutamatergic Systems. J. Pharmacol. Exp. Ther 315, 590–600. [DOI] [PubMed] [Google Scholar]
  49. Cowen MS, Lawrence AJ (2001) Alterations in central preproenkephalin mRNA expression after chronic free-choice ethanol consumption by fawn-hooded rats. Alcohol. Clin. Exp. Res 25, 1126–33. [PubMed] [Google Scholar]
  50. Dahlin A, Xia L, Kong W, Hevner R, Wang J (2007) Expression and immunolocalization of the plasma membrane monoamine transporter in the brain. Neuroscience 146, 1193–211. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Danbolt NC, Storm-Mathisen J, Kanner BI (1992) An [Na+ + K+]coupled I-glutamate transporter purified from rat brain is located in glial cell processes. Neuroscience 94, 225–41. [DOI] [PubMed] [Google Scholar]
  52. Dani JA (2001) Overview of nicotinic receptors and their roles in the central nervous system. Biol. Psychiatry 49, 166–174. [DOI] [PubMed] [Google Scholar]
  53. Dautan D, Hacioğlu Bay H, Bolam JP, Gerdjikov TV, Mena-Segovia J (2016) Extrinsic sources of cholinergic innervation of the striatal complex: a whole-brain mapping analysis. Front. Neuroanat 10, 1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Dautan D, Huerta-Ocampo I, Witten IB, Deisseroth K, Bolam JP, Gerdjikov T, Mena-Segovia J (2014) A major external source of cholinergic innervation of the striatum and nucleus accumbens originates in the brainstem. J. Neurosci 34, 4509–4518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Day HEW, Campeau S, Watson SJ, Akil H (1997) Distribution of α1a-, α1b- and α1d-adrenergic receptor mRNA in the rat brain and spinal cord. J. Chem. Neuroanat 13, 115–139. [DOI] [PubMed] [Google Scholar]
  56. Del-Fava F, Hasue RH, Ferreira JGP, Shammah-Lagnado SJ (2007) Efferent connections of the rostral linear nucleus of the ventral tegmental area in the rat. Neuroscience 145, 1059–1076. [DOI] [PubMed] [Google Scholar]
  57. Demarest KT, Moore KE (1979) Comparison of dopamine synthesis regulation in the terminals of nigrostriatal, mesolimbic, tuberoinfundibular and tuberohypophyseal neurons. J. Neural Transm 46, 263–277. [DOI] [PubMed] [Google Scholar]
  58. Devanand DP, Michaels-Marston KS, Liu X, Pelton GH, Padilla M, Marder K, Bell K, Stern Y, Mayeux R (2000) Olfactory deficits in patients with mild cognitive impairment predict Alzheimer’s disease at follow-up. Am J Psychiatry 157, 1399–1405. [DOI] [PubMed] [Google Scholar]
  59. Devost D, Wrzal P, Zingg HH (2008) Oxytocin receptor signalling, in Adv. Vasopressin Oxytocin — From Genes to Behav. to Dis, (Neumann ID, Landgraf R. B. T.-P. in B. R., eds), Vol. 170, pp. 167–176. Elsevier. [DOI] [PubMed] [Google Scholar]
  60. Diaz J, Levesque D, Lammers CH, Griffon N, Martres M-P, Schwartz J-C, Sokoloff P (1995) Phenotypical characterization of neurons expressing the dopamine D3 receptor in the rat brain. Neuroscience 65, 731–745. [DOI] [PubMed] [Google Scholar]
  61. Diaz J, Ridray S, Mignon V, Griffon N, Schwartz J-C, Sokoloff P (1997) Selective Expression of Dopamine D3 Receptor mRNA in Proliferative Zones during Embryonic Development of the Rat Brain. J. Neurosci 17, 4282–4292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. DiBenedictis BT, Olugbemi AO, Baum MJ, Cherry JA (2014) 6-Hydroxydopamine lesions of the anteromedial ventral striatum impair opposite-sex urinary odor preference in female mice. Behav. Brain Res 274, 243–247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. DiBenedictis BT, Olugbemi AO, Baum MJ, Cherry JA (2015) DREADD-Induced Silencing of the Medial Olfactory Tubercle Disrupts the Preference of Female Mice for Opposite-Sex Chemosignals. eNeuro 2, 1–16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Ding JD, Burette A, Nedvetsky PI, Schmidt HHHW, Weinberg RJ (2004) Distribution of Soluble Guanylyl Cyclase in the Rat Brain. J. Comp. Neurol 472, 437–448. [DOI] [PubMed] [Google Scholar]
  65. Donat CK, Schuhmann MU, Voigt C, Nieber K, Deuther-Conrad W, Brust P (2008) Time-dependent alterations of cholinergic markers after experimental traumatic brain injury. Brain Res. 1246, 167–177. [DOI] [PubMed] [Google Scholar]
  66. Doty RL (2017) Olfactory dysfunction in neurodegenerative diseases: is there a common pathological substrate? Lancet Neurol. 16, 478–488. [DOI] [PubMed] [Google Scholar]
  67. Doty RL, Shaman P, Applebaum SL, Giberson R, Sikorski L, Rosenberg L (1984) Smell identification ability: Changes with age. Science (80-. ). 226, 1441–1443. [DOI] [PubMed] [Google Scholar]
  68. Drenan RM, Grady SR, Steele AD, McKinney S, Patzlaff NE, McIntosh JM, Marks MJ, Miwa JM, Lester HA (2010) Cholinergic Modulation of Locomotion and Striatal Dopamine Release Is Mediated by α6α4* Nicotinic Acetylcholine Receptors. J. Neurosci 30, 9877 LP–9889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Drenan RM, Grady SR, Whiteaker P, McClure-Begley T, McKinney S, Miwa JM, Bupp S, Heintz N, McIntosh JM, Bencherif M (2008) In vivo activation of midbrain dopamine neurons via sensitized, high-affinity α6* nicotinic acetylcholine receptors. Neuron 60, 123–136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Duncan GE, Breese GR, Criswell HE, McCown TJ, Herbert JS, Devaud LL, Morrow AL (1995) Distribution of [3H]zolpidem binding sites in relation to messenger RNA encoding the α1, β2 and γ2 subunits of GABAA receptors in rat brain. Neuroscience 64, 1113–28. [DOI] [PubMed] [Google Scholar]
  71. Durkin MM, Gunwaldsen CA, Borowsky B, Jones KA, Branchek TA (1999) An in situ hybridization study of the distribution of the GABA(B2) protein mRNA in the rat CNS. Brain Res. Mol. Brain Res 71, 185–200. [DOI] [PubMed] [Google Scholar]
  72. Durkin MM, Smith KE, Borden LA, Weinshank RL, Branchek TA, Gustafson EL (1995) Localization of messenger RNAs encoding three GABA transporters in rat brain: an in situ hybridization study. Mol. Brain Res 33, 7–21. [DOI] [PubMed] [Google Scholar]
  73. Eiden LE, Schäfer MK-H, Weihe E, Schütz B (2004) The vesicular amine transporter family (SLC18): amine/proton antiporters required for vesicular accumulation and regulated exocytotic secretion of monoamines and acetylcholine. Pflügers Arch. 447, 636–640. [DOI] [PubMed] [Google Scholar]
  74. Fallon JH (1983) The islands of Calleja complex of rat basal forebrain II: Connections of medium and large sized cells. Brain Res. Bull 10, 775–793. [DOI] [PubMed] [Google Scholar]
  75. Finley JCW, Grossman GH, Dimeo P, Petrusz P (1978) Somatostatin‐containing neurons in the rat brain: Widespread distribution revealed by immunocytochemistry after pretreatment with pronase. Am. J. Anat 153, 483–8. [DOI] [PubMed] [Google Scholar]
  76. FitzGerald BJ, Richardson K, Wesson DW (2014) Olfactory tubercle stimulation alters odor preference behavior and recruits forebrain reward and motivational centers. Front. Behav. Neurosci 8, 81. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Foltz TL, Snow DM, Strupp BJ, Booze RM, Mactutus CF (2004) Prenatal intravenous cocaine and the heart rate-orienting response: a dose-response study. Int. J. Dev. Neurosci 22, 285–296. [DOI] [PubMed] [Google Scholar]
  78. Freed C, Revay R, Vaughan RA, Kriek E, Grant S, Uhl GR, Kuhar MJ (1995) Dopamine transporter immunoreactivity in rat brain. J. Comp. Neurol 359, 340–349. [DOI] [PubMed] [Google Scholar]
  79. Fremeau RT, Troyer MD, Pahner I, Nygaard GO, Tran CH, Reimer RJ, Bellocchio EE, Fortin D, Storm-Mathisen J, Edwards RH (2001) The expression of vesicular glutamate transporters defines two classes of excitatory synapse. Neuron 31, 247–60. [DOI] [PubMed] [Google Scholar]
  80. Fuller TA, Price JL (1988) Putative glutamatergic and/or aspartatergic cells in the main and accessory olfactory bulbs of the rat. J. Comp. Neurol 276, 209–18. [DOI] [PubMed] [Google Scholar]
  81. Funada M, Suzuki T, Narita M, Misawa M, Nagase H (1993) Blockade of morphine reward through the activation of kappa-opioid receptors in mice. Neuropharmacology 32, 1315–23. [DOI] [PubMed] [Google Scholar]
  82. Furuta T, Zhou L, Kaneko T (2002) Preprodynorphin-, preproenkephalin-, preprotachykinin A- and preprotachykinin B-immunoreactive neurons in the accumbens nucleus and olfactory tubercle: double-immunofluorescence analysis. Neuroscience 114, 611–27. [DOI] [PubMed] [Google Scholar]
  83. Gadziola MA, Tylicki KA, Christian DL, Wesson DW (2015) The Olfactory Tubercle Encodes Odor Valence in Behaving Mice. J. Neurosci 35, 4515–4527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Gadziola MA, Wesson DW (2016) The Neural Representation of Goal-Directed Actions and Outcomes in the Ventral Striatum’s Olfactory Tubercle. J. Neurosci 36, 548–560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Garris DR (2003) Aggression-associated changes in murine olfactory tubercle bioamines. Brain Res. 963, 150–5. [DOI] [PubMed] [Google Scholar]
  86. Gaspar P, Cases O, Maroteaux L (2003) The developmental role of serotonin: news from mouse molecular genetics. Nat Rev Neurosci 4, 1002–1012. [DOI] [PubMed] [Google Scholar]
  87. Gérard C, Martres M-P, Lefèvre K, Miquel M-C, Vergé D, Lanfumey L, Doucet E, Hamon M, El Mestikawy S (1997) Immuno-localization of serotonin 5-HT6 receptor-like material in the rat central nervous system. Brain Res. 746, 207–219. [DOI] [PubMed] [Google Scholar]
  88. Gerfen CR (1985) The neostriatal mosaic. I. Compartmental organization of projections from the striatum to the substantia nigra in the rat. J. Comp. Neurol 236, 454–476. [DOI] [PubMed] [Google Scholar]
  89. Gerfen CR, Engber TM, Mahan LC, Susel Z, Chase TN, Monsma FJ, Sibley DR (1990) D1 and D2 dopamine receptor-regulated gene expression of striatonigral and striatopallidal neurons. Science (80-. ). 250, 1429–1432. [DOI] [PubMed] [Google Scholar]
  90. Giardino L, Zanni M, Bettelli C, Savina MA, Calzà L (1996) Regulation of glutamic acid decarboxylase mRNA expression in rat brain after sertraline treatment. Eur. J. Pharmacol 312, 183–187. [DOI] [PubMed] [Google Scholar]
  91. Gilmor ML, Nash NR, Roghani A, Edwards RH, Yi H, Hersch SM, Levey AI (1996) Expression of the putative vesicular acetylcholine transporter in rat brain and localization in cholinergic synaptic vesicles. J. Neurosci 16, 2179–2190. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Gimpl G, Fahrenholz F (2001) The oxytocin receptor system: structure, function, and regulation. Physiol. Rev 81, 629–683. [DOI] [PubMed] [Google Scholar]
  93. Grady SR, Murphy KL, Cao J, Marks MJ, McIntosh JM, Collins AC (2002) Characterization of nicotinic agonist-induced [3H] dopamine release from synaptosomes prepared from four mouse brain regions. J. Pharmacol. Exp. Ther 301, 651–660. [DOI] [PubMed] [Google Scholar]
  94. Gras C, Herzog E, Bellenchi GC, Bernard V, Ravassard P, Pohl M, Gasnier B, Giros B, El Mestikawy S (2002) A third vesicular glutamate transporter expressed by cholinergic and serotoninergic neurons. J. Neurosci 22, 5442–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Graybiel AM, Ragsdale CW (1978) Histochemically distinct compartments in the striatum of human, monkeys, and cat demonstrated by acetylthiocholinesterase staining. Proc. Natl. Acad. Sci 75, 5723–5726. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Greger IH, Watson JF, Cull-Candy SG (2017) Structural and Functional Architecture of AMPA-Type Glutamate Receptors and Their Auxiliary Proteins. Neuron 94, 713–730. [DOI] [PubMed] [Google Scholar]
  97. Grossman CJ, Kilpatrick GJ, Bunce KT (1993) Development of a radioligand binding assay for 5-HT4 receptors in guinea-pig and rat brain. Br. J. Pharmacol 109, 618–624. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Guevara-Aguilar R, Solano-Flores LP, Donatti-Albarran OA, Aguilar-Baturoni HU (1982) Differential projections from locus coeruleus to olfactory bulb and olfactory tubercle: an HRP study. Brain Res Bull 8, 711–719. [DOI] [PubMed] [Google Scholar]
  99. Haar CV, Martens KM, Bashir A, McInnes KA, Cheng WH, Cheung H, Stukas S, et al. (2019) Repetitive closed-head impact model of engineered rotational acceleration (CHIMERA) injury in rats increases impulsivity, decreases dopaminergic innervation in the olfactory tubercle and generates white matter inflammation, tau phosphorylation and degener. Exp. Neurol 317, 87–99. [DOI] [PubMed] [Google Scholar]
  100. Hadley JK, Halliwell JV (2010) Serotonin modulates glutamatergic transmission in the rat olfactory tubercle. Eur. J. Neurosci 31, 659–672. [DOI] [PubMed] [Google Scholar]
  101. Hagan CE, McDevitt RA, Liu Y, Furay AR, Neumaier JF (2012) 5-HT(1B) autoreceptor regulation of serotonin transporter activity in synaptosomes. Synapse 66, 1024–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  102. Hammock EAD, Levitt P (2013) Oxytocin receptor ligand binding in embryonic tissue and postnatal brain development of the C57BL/6J mouse. Front. Behav. Neurosci 7, 195. [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Harlan RE, Shivers BD, Romano GJ, Pfaff DW, Howells RD (1987) Localization of preproenkephalin mRNA in the rat brain and spinal cord byin situ hybridization. J. Comp. Neurol 258, 159–184. [DOI] [PubMed] [Google Scholar]
  104. Hashimoto K, Tomitaka S-I, Narita N, Minabe Y, Iyo M (1997) Induction of Fos protein by 3,4-methylenedioxymethamphetamine (Ecstasy) in rat brain: regional differences in pharmacological manipulation. Addict. Biol 2, 317–326. [DOI] [PubMed] [Google Scholar]
  105. Hebb CO, Silver A (1961) Gradient of cholinesterase activity and of choline acetylase activity in nerve fibres: gradient of choline acetylase activity. Nature 189, 123. [DOI] [PubMed] [Google Scholar]
  106. Heimer L, Zaborszky L, Zahm DS, Alheid GF (1987) The ventral striatopallidothalamic projection: I. The striatopallidal link originating in the striatal parts of the olfactory tubercle. J. Comp. Neurol 255, 571–591. [DOI] [PubMed] [Google Scholar]
  107. Helboe L, Egebjerg J, de Jong IEM (2015) Distribution of serotonin receptor 5-HT6 mRNA in rat neuronal subpopulations: A double in situ hybridization study. Neuroscience 310, 442–454. [DOI] [PubMed] [Google Scholar]
  108. Hellriegel ET, D’Mello AP (1997) The effect of acute, chronic and chronic intermittent stress on the central noradrenergic system. Pharmacol. Biochem. Behav 57, 207–14. [DOI] [PubMed] [Google Scholar]
  109. Henny P, Jones BE (2006) Vesicular glutamate (VGlut), GABA (VGAT), and acetylcholine (VACht) transporters in basal forebrain axon terminals innervating the lateral hypothalamus. J. Comp. Neurol 496, 453–467. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Hermann H, Marsicano G, Lutz B (2002) Coexpression of the cannabinoid receptor type 1 with dopamine and serotonin receptors in distinct neuronal subpopulations of the adult mouse forebrain. Neuroscience 109, 451–460. [DOI] [PubMed] [Google Scholar]
  111. Herzog E, Gilchrist J, Gras C, Muzerelle A, Ravassard P, Giros B, Gaspar P, El Mestikawy S (2004) Localization of VGLUT3, the vesicular glutamate transporter type 3, in the rat brain. Neuroscience 123, 983–1002. [DOI] [PubMed] [Google Scholar]
  112. Hillegaart V, Hjorth S, Ahlenius S (1990) Effects of 5-HT and 8-OH-DPAT on forebrain monoamine synthesis after local application into the median and dorsal raphe nuclei of the rat. J. Neural Transm 81, 131–145. [DOI] [PubMed] [Google Scholar]
  113. Hipólide DC, Moreira KM, Barlow KBL, Wilson AA, Nobrega JN, Tufik S (2005) Distinct effects of sleep deprivation on binding to norepinephrine and serotonin transporters in rat brain. Prog. Neuro-Psychopharmacology Biol. Psychiatry 29, 297–303. [DOI] [PubMed] [Google Scholar]
  114. Hoyer D, Hannon JP, Martin GR (2002) Molecular, pharmacological and functional diversity of 5-HT receptors. Pharmacol. Biochem. Behav 71, 533–554. [DOI] [PubMed] [Google Scholar]
  115. Hsieh Y-C, Puche AC (2013) Development of the Islands of Calleja. Brain Res. 1490, 52–60. [DOI] [PubMed] [Google Scholar]
  116. Hsieh Y-C, Puche AC (2015) GABA modulation of SVZ-derived progenitor ventral cell migration. Dev. Neurobiol 75, 791–804. [DOI] [PubMed] [Google Scholar]
  117. Huang GZ, Woolley CS (2012) Estradiol Acutely Suppresses Inhibition in the Hippocampus through a Sex-Specific Endocannabinoid and mGluR-Dependent Mechanism. Neuron 74, 801–808. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Huang Q, Zhou D, Chase K, Gusella JF, Aronin N, DiFiglia M (1992) Immunohistochemical localization of the D1 dopamine receptor in rat brain reveals its axonal transport, pre-and postsynaptic localization, and prevalence in the basal ganglia, limbic system, and thalamic reticular nucleus. Proc. Natl. Acad. Sci 89, 11988–11992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Huang X-F, Huang X, Han M, Chen F, Storlien L, Lawrence AJ (2004) 5-HT2A/2C receptor and 5-HT transporter densities in mice prone or resistant to chronic high-fat diet-induced obesity: a quantitative autoradiography study. Brain Res. 1018, 227–235. [DOI] [PubMed] [Google Scholar]
  120. Ichikawa T, Ajiki K, Matsuura J, Misawa H (1997) Localization of two cholinergic markers, choline acetyltransferase and vesicular acetylcholine transporter in the central nervous system of the rat: in situ hybridization histochemistry and immunohistochemistry. J. Chem. Neuroanat 13, 23–39. [DOI] [PubMed] [Google Scholar]
  121. Igarashi KM, Ieki N, An M, Yamaguchi Y, Nagayama S, Kobayakawa K, Kobayakawa R, et al. (2012) Parallel Mitral and Tufted Cell Pathways Route Distinct Odor Information to Different Targets in the Olfactory Cortex. J. Neurosci 32, 7970–7985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Ikemoto S (2003) Involvement of the Olfactory Tubercle in Cocaine Reward: Intracranial Self-Administration Studies. J. Neurosci 23, 9305–9311. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Ikemoto S (2005) The Functional Divide for Primary Reinforcement of D-Amphetamine Lies between the Medial and Lateral Ventral Striatum: Is the Division of the Accumbens Core, Shell, and Olfactory Tubercle Valid? J. Neurosci 25, 5061–5065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Ikemoto S (2007) Dopamine reward circuitry: Two projection systems from the ventral midbrain to the nucleus accumbens-olfactory tubercle complex. Brain Res Rev 56, 27–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Ikemoto S (2010) Brain reward circuitry beyond the mesolimbic dopamine system: A neurobiological theory. Neurosci. Biobehav. Rev 35, 129–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Imamura F, Ayoub AE, Rakic P, Greer CA (2011) Timing of neurogenesis is a determinant of olfactory circuitry. Nat Neurosci 14, 331–337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Ishii K, Wong JK, Sumikawa K (2005) Comparison of α2 nicotinic acetylcholine receptor subunit mRNA expression in the central nervous system of rats and mice. J. Comp. Neurol 493, 241–260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Johnson AE, Coirini H, Insel TR, McEwen BS (1991) The Regulation of Oxytocin Receptor Binding in the Ventromedial Hypothalaimic Nucleus by Testosterone and Its Metabolites*. Endocrinology 128, 891–896. [DOI] [PubMed] [Google Scholar]
  129. Jones BE, Moore RY (1977) Ascending projections of the locus coeruleus in the rat. II. Autoradiographic study. Brain Res. 127, 23–53. [PubMed] [Google Scholar]
  130. Kabbaj M (2006) Individual Differences in Vulnerability to Drug Abuse: The High Responders/Low Responders Model. CNS Neurol. Disord. - Drug Targets 5, 513–520. [DOI] [PubMed] [Google Scholar]
  131. Kaneko S, Hikida T, Watanabe D, Ichinose H, Nagatsu T, Kreitman RJ, Pastan I, Nakanishi S (2000) Synaptic integration mediated by striatal cholinergic interneurons in basal ganglia function. Science (80-. ). 289, 633–637. [DOI] [PubMed] [Google Scholar]
  132. Kaupmann K, Malitschek B, Schuler V, Heid J, Froestl W, Beck P, Mosbacher J, et al. (1998) GABA(B)-receptor subtypes assemble into functional heteromeric complexes. Nature 396, 683–7. [DOI] [PubMed] [Google Scholar]
  133. Kilpatrick G, Jones B, Tyers M (1988) The distribution of specific binding of the 5-HT3 receptor ligand [3H]GR65630 in rat brain using quantitative autoradiography. Neuroscie Lett. 94, 156–60. [DOI] [PubMed] [Google Scholar]
  134. Kinoshita A, Shigemoto R, Ohishi H, Van DerPutten H , Mizuno N. (1998) Immunohistochemical localization of metabotropic glutamate receptors, mGluR7a and mGluR7b, in the central nervous system of the adult rat and mouse: A light and electron microscopic study. J. Comp. Neurol 58, 1021–6. [PubMed] [Google Scholar]
  135. Kinsey AM, Wainwright A, Heavens R, Sirinathsinghji DJ, Oliver KR (2001) Distribution of 5-ht(5A), 5-ht(5B), 5-ht(6) and 5-HT(7) receptor mRNAs in the rat brain. Brain Res. Mol. Brain Res 88, 194–8. [DOI] [PubMed] [Google Scholar]
  136. Kloet R. de, Voorhuis DAM, Boschma Y, Elands J (1986) Estradiol Modulates Density of Putative Oxytocin Receptors’ in Discrete Rat Brain Regions. Neuroendocrinology 44, 415–421. [DOI] [PubMed] [Google Scholar]
  137. Knobloch HS, Charlet A, Hoffmann LC, Eliava M, Khrulev S, Cetin AH, Osten P, et al. (2012) Evoked Axonal Oxytocin Release in the Central Amygdala Attenuates Fear Response. Neuron 73, 553–566. [DOI] [PubMed] [Google Scholar]
  138. Koos T, JM T (1999) Inhibitory control of neostriatal projection neurons by GABAergic interneurons. Nat. Neurosci 2, 467–472. [DOI] [PubMed] [Google Scholar]
  139. Kosaka T, Kosaka K (2007) Heterogeneity of nitric oxide synthase-containing neurons in the mouse main olfactory bulb. Neurosci. Res 57, 165–178. [DOI] [PubMed] [Google Scholar]
  140. Koulousakis P, Andrade P, Visser-Vandewalle V, Sesia T (2019) The Nucleus Basalis of Meynert and Its Role in Deep Brain Stimulation for Cognitive Disorders: A Historical Perspective. J. Alzheimer’s Dis 69, 905–919. [DOI] [PubMed] [Google Scholar]
  141. Koyama Y, Kondo M, Shimada S (2017) Building a 5-HT3A Receptor Expression Map in the Mouse Brain. Sci. Rep 7, 42884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Kreitzer AC, Malenka R (2008) Striatal plasticity and basal ganglia circuit function. Neuron 60, 543–554. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Krémarik P, Freund-Mercier MJ, Stoeckel ME (1995) Estrogen-sensitive oxytocin binding sites are differently regulated by progesterone in the telencephalon and the hypothalamus of the rat. J. Neuroendocrinol 7, 281–9. [DOI] [PubMed] [Google Scholar]
  144. Krieger NR, Megill JR, Sterling P (1983a) Granule cells in the rat olfactory tubercle accumulate 3H-gamma-aminobutyric acid. J. Comp. Neurol 215, 465–471. [DOI] [PubMed] [Google Scholar]
  145. Krieger NR, Scott RG, Jurman ME (1983b) Testosterone 5 alpha-reductase in rat brain. J. Neurochem 40, 1460–4. [DOI] [PubMed] [Google Scholar]
  146. Landry M, Lévesque D, Di Paolo T (2002) Estrogenic Properties of Raloxifene, but Not Tamoxifen, on D2 and D3 Dopamine Receptors in the Rat Forebrain. Neuroendocrinology 76, 214–222. [DOI] [PubMed] [Google Scholar]
  147. Lauterborn JC, Isackson PJ, Montalvo R, Gall CM (1993) In situ hybridization localization of choline acetyltransferase mRNA in adult rat brain and spinal cord. Mol. brain Res 17, 59–69. [DOI] [PubMed] [Google Scholar]
  148. Laux A, Muller AH, Miehe M, Dirrig-Grosch S, Deloulme JC, Delalande F, Stuber D, et al. (2011) Mapping of endogenous morphine-like compounds in the adult mouse brain: Evidence of their localization in astrocytes and GABAergic cells. J. Comp. Neurol 519, 2390–2416. [DOI] [PubMed] [Google Scholar]
  149. Lavreysen H, Janssen C, Bischoff F, Langlois X, Leysen JE, Lesage ASJ (2003) [3H]R214127: a novel high-affinity radioligand for the mGlu1 receptor reveals a common binding site shared by multiple allosteric antagonists. Mol. Pharmacol 63, 1082–93. [DOI] [PubMed] [Google Scholar]
  150. Lerma J, Marques JM (2013) Kainate receptors in health and disease. Neuron 80, 292–311. [DOI] [PubMed] [Google Scholar]
  151. Levy LM, Lehre KP, Rolstad B, Danbolt NC (1993) A monoclonal antibody raised against an [Na+K+]coupled I-glutamate transporter purified from rat brain confirms glial cell localization. FEBS Lett. 317, 79–84. [DOI] [PubMed] [Google Scholar]
  152. Lew R, Weisenberg B, Vosmer G, Seiden LS (1997) Combined phentermine/fenfluramine administration enhances depletion of serotonin from central terminal fields. Synapse 26, 36–45. [DOI] [PubMed] [Google Scholar]
  153. Lim SAO, Kang UJ, McGehee DS (2014) Striatal cholinergic interneuron regulation and circuit effects. Front. Synaptic Neurosci 6, 22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Lin JY, Chung KKH, Castro D. De, Funk GD, Lipski J. (2004) Effects of muscarinic acetylcholine receptor activation on membrane currents and intracellular messengers in medium spiny neurones of the rat striatum. Eur. J. Neurosci 20, 1219–1230. [DOI] [PubMed] [Google Scholar]
  155. Lippa CF, Smith TW, Perry E (1999) Dementia with Lewy bodies: choline acetyltransferase parallels nucleus basalis pathology. J. Neural Transm 106, 525–535. [DOI] [PubMed] [Google Scholar]
  156. Lobo MK, Karsten SL, Gray M, Geschwind DH, Yang XW (2006) FACS-array profiling of striatal projection neuron subtypes in juvenile and adult mouse brains. Nat. Neurosci 9, 443–52. [DOI] [PubMed] [Google Scholar]
  157. López-Giménez JF, Mengod G, Palacios JM, Vilaró MT (1997) Selective visualization of rat brain 5-HT2A receptors by autoradiography with [3H]MDL 100,907. Naunyn. Schmiedebergs. Arch. Pharmacol 356, 446–454. [DOI] [PubMed] [Google Scholar]
  158. Luskin MB, Price JL (1983) The topographic organization of associational fibers of the olfactory system in the rat, including centrifugal fibers to the olfactory bulb. J. Comp. Neurol 216, 264–291. [DOI] [PubMed] [Google Scholar]
  159. Mamaligas AA, Ford CP (2016) Spontaneous Synaptic Activation of Muscarinic Receptors by Striatal Cholinergic Neuron Firing. Neuron 91, 574–586. [DOI] [PMC free article] [PubMed] [Google Scholar]
  160. Mansour A, Fox CA, Burke S, Meng F, Thompson RC, Akil H, Watson SJ (1994) Mu, delta, and kappa opioid receptor mRNA expression in the rat CNS: An in situ hybridization study. J. Comp. Neurol 350, 412–438. [DOI] [PubMed] [Google Scholar]
  161. Mansour A, Khachaturian H, Lewis ME, Akil H, Watson SJ (1987) Autoradiographic differentiation of mu, delta, and kappa opioid receptors in the rat forebrain and midbrain. J. Neurosci 7, 2445–64. [PMC free article] [PubMed] [Google Scholar]
  162. De Marchis S, Fasolo A, Puche AC (2004) Subventricular zone-derived neuronal progenitors migrate into the subcortical forebrain of postnatal mice. J. Comp. Neurol 476, 290–300. [DOI] [PubMed] [Google Scholar]
  163. Marshall FH, Jones KA, Kaupmann K, Bettler B (1999) GABA(B) receptors - The first 7TM heterodimers. Trends Pharmacol. Sci 20, 396–9. [DOI] [PubMed] [Google Scholar]
  164. Martin-Lopez E, Xu C, Liberia T, Meller SJ, Greer CA (2019) Embryonic and postnatal development of mouse olfactory tubercle. Mol. Cell. Neurosci 11, 82–96. [DOI] [PMC free article] [PubMed] [Google Scholar]
  165. Marubio LM, Gardier AM, Durier S, David D, Klink R, Arroyo-Jimenez MM, McIntosh JM, et al. (2003) Effects of nicotine in the dopaminergic system of mice lacking the alpha4 subunit of neuronal nicotinic acetylcholine receptors. Eur. J. Neurosci 17, 1329–1337. [DOI] [PubMed] [Google Scholar]
  166. Massie A, Cnops L, Smolders I, McCullumsmith R, Kooijman R, Kwak S, Arckens L, Michotte Y (2008) High-affinity Na+/K+-dependent glutamate transporter EAAT4 is expressed throughout the rat fore- and midbrain. J. Comp. Neurol 511, 155–72. [DOI] [PubMed] [Google Scholar]
  167. Maurice N, Mercer J, CS C, Hernandez-Lopez S, Held J (2004) D2 dopamine receptor-mediated modulation of voltage-dependent Na+ channels reduces autonomous activity in striatal cholinergic interneurons. J. Neurosci 24, 10289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. McCarthy MM, Kleopoulos SP, Mobbs CV, Pfaff DW (1994) Infusion of Antisense Oligo-deoxynucleotides to the Oxytocin Receptor in the Ventromedial Hypothalamus Reduces Estrogen-Induced Sexual Receptivity and Oxytocin Receptor Binding in the Female Rat. Neuroendocrinology 59, 432–440. [DOI] [PubMed] [Google Scholar]
  169. McGehee DS, Role LW (1995) Physiological Diversity of Nicotinic Acetylcholine Receptors Expressed by Vertebrate Neurons. Annu. Rev. Physiol 57, 521–546. [DOI] [PubMed] [Google Scholar]
  170. McGregor MM, Nelson AB (2019) Circuit Mechanisms of Parkinson’s Disease. Neuron 101, 1042–1056. [DOI] [PubMed] [Google Scholar]
  171. McLean JH, Shipley MT (1987) Serotonergic afferents to the rat olfactory bulb: I. Origins and laminar specificity of serotonergic inputs in the adult rat. J Neurosci 7, 3016–3028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Menéndez L, Insua D, Rois JL, Santamarina G, Suárez ML, Pesini P (2006) The immunohistochemical localization of neuronal nitric oxide synthase in the basal forebrain of the dog. J. Chem. Neuroanat 31, 200–209. [DOI] [PubMed] [Google Scholar]
  173. Mesulam M-M, Mufson EJ, Wainer BH, Levey AI (1983) Central cholinergic pathways in the rat: An overview based on an alternative nomenclature (Ch1–Ch6). Neuroscience 10, 1185–1201. [DOI] [PubMed] [Google Scholar]
  174. Metaxas A, Al-Hasani R, Farshim P, Tubby K, Berwick A, Ledent C, Hourani S, Kitchen I, Bailey A (2013) Genetic deletion of the adenosine A2Areceptor prevents nicotine-induced upregulation of α7, but not α4β2* nicotinic acetylcholine receptor binding in the brain. Neuropharmacology 71, 228–236. [DOI] [PubMed] [Google Scholar]
  175. Mijnster MJ, Raimundo AGV, Koskuba K, Klop H, Docter GJ, Groenewegen HJ, Voorn P (1997) Regional and cellular distribution of serotonin 5-hydroxytryptamine2a receptor mRNA in the nucleus accumbens, olfactory tubercle, and caudate putamen of the rat. J. Comp. Neurol 389, 1–11. [PubMed] [Google Scholar]
  176. Millhouse OE (1987) Granule cells of the olfactory tubercle and the question of the islands of calleja. J Comp Neurol 265, 1–24. [DOI] [PubMed] [Google Scholar]
  177. Millhouse OE, Heimer L (1984) Cell configurations in the olfactory tubercle of the rat. J Comp Neurol 228, 571–597. [DOI] [PubMed] [Google Scholar]
  178. Minelli A, Brecha NC, Karschin C, DeBiasi S, Conti F (1995) GAT-1, a high-affinity GABA plasma membrane transporter, is localized to neurons and astroglia in the cerebral cortex. J. Neurosci 15, 7734–46. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Minelli A, DeBiasi S, Brecha NC, Vitellaro Zuccarello L, Conti F (1996) GAT-3, a High-Affinity GABA Plasma Membrane Transporter, Is Localized to Astrocytic Processes, and It Is Not Confined to the Vicinity of GABAergic Synapses in the Cerebral Cortex. J. Neurosci 16, 6255–64. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Mingote S, Chuhma N, Kusnoor SV, Field B, Deutch AY, Rayport S (2015) Functional Connectome Analysis of Dopamine Neuron Glutamatergic Connections in Forebrain Regions. J. Neurosci 35, 16259–16271. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Misawa H, Nakata K, Matsuura J, Nagao M, Okuda T, Haga T (2001) Distribution of the high-affinity choline transporter in the central nervous system of the rat. Neuroscience 105, 87–98. [DOI] [PubMed] [Google Scholar]
  182. Missale C, Nash SR, Robinson SW, Jaber M, Caron MG (1998) Dopamine receptors: from structure to function. Physiol. Rev 78, 189–225. [DOI] [PubMed] [Google Scholar]
  183. Le Moine C, Bloch B (1995) D1 and D2 dopamine receptor gene expression in the rat striatum: Sensitive cRNA probes demonstrate prominent segregation of D1 and D2 mRNAS in distinct neuronal populations of the dorsal and ventral striatum. J. Comp. Neurol 355, 418–426. [DOI] [PubMed] [Google Scholar]
  184. Le Moine C, Bloch B (1996) Expression of the D3 dopamine receptor in peptidergic neurons of the nucleus accumbens: Comparison with the D1 and D2 dopamine receptors. Neuroscience 73, 131–143. [DOI] [PubMed] [Google Scholar]
  185. Morilak DA, Barrera G, Echevarria DJ, Garcia AS, Hernandez A, Ma S, Petre CO (2005) Role of brain norepinephrine in the behavioral response to stress. Prog. Neuro-Psychopharmacology Biol. Psychiatry 29, 1214–1224. [DOI] [PubMed] [Google Scholar]
  186. Morris G, Arkadir D, Nevet A, Vaadia E, Bergman H (2004) Coincident but distinct messages of midbrain dopamine and striatal tonically active neurons. Neuron 43, 133–143. [DOI] [PubMed] [Google Scholar]
  187. Mucha RF, Herz A (1985) Motivational properties of kappa and mu opioid receptor agonists studied with place and taste preference conditioning. Psychopharmacology (Berl). 86, 274–80. [DOI] [PubMed] [Google Scholar]
  188. Murad F, Mittal CK, Arnold WP, Katsuki S, Kimura H (1978) Guanylate cyclase: activation by azide, nitro compounds, nitric oxide, and hydroxyl radical and inhibition by hemoglobin and myoglobin. Adv. Cyclic Nucleotide Res 9, 145–158. [PubMed] [Google Scholar]
  189. Murata K, Kanno M, Ieki N, Mori K, Yamaguchi M (2015) Mapping of Learned Odor-Induced Motivated Behaviors in the Mouse Olfactory Tubercle. J. Neurosci 35, 10581–10599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Murphy C (2019) Olfactory and other sensory impairments in Alzheimer disease. Nat. Rev. Neurol 15, 11–24. [DOI] [PubMed] [Google Scholar]
  191. Murray AM, Ryoo HL, Gurevich E, Joyce JN (1994) Localization of dopamine D3 receptors to mesolimbic and D2 receptors to mesostriatal regions of human forebrain. Proc. Natl. Acad. Sci 91, 11271–11275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Murray A, Weihueller F, Marshall J, Hurtig H, Gottleib G, Joyce J (1995) Damage to dopamine systems differs between Parkinson’s disease and Alzheimer’s disease with parkinsonism. Ann Neurol 37, 300–12. [DOI] [PubMed] [Google Scholar]
  193. Mutel V, Ellis GJ, Adam G, Chaboz S, Nilly A, Messer J, Bleuel Z, et al. (2000) Characterization of [(3)H]Quisqualate binding to recombinant rat metabotropic glutamate 1a and 5a receptors and to rat and human brain sections. J. Neurochem 75, 2590–601. [DOI] [PubMed] [Google Scholar]
  194. Nagayama S, Enerva A, Fletcher ML, Masurkar AV, Igarashi KM, Mori K, Chen WR (2010) Differential Axonal Projection of Mitral and Tufted Cells in the Mouse Main Olfactory System. Front. Neural Circuits 4, 1–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Nicholas AP, Pieribone VA, Hökfelt T (1993) Cellular localization of messenger RNA for beta-1 and beta-2 adrenergic receptors in rat brain: An in situ hybridization study. Neuroscience 56, 1023–1039. [DOI] [PubMed] [Google Scholar]
  196. Ohishi H, Akazawa C, Shigemoto R, Nakanishi S, Mizuno N (1995) Distributions of the mRNAs for L-2-amino-4-phosphonobutyrate-sensitive metabotropic glutamate receptors, mGluR4 and mGluR7, in the rat brain. J. Comp. Neurol 360, 555–570. [DOI] [PubMed] [Google Scholar]
  197. Oliva JM, Manzanares J (2007) Gene Transcription Alterations Associated with Decrease of Ethanol Intake Induced by Naltrexone in the Brain of Wistar Rats. Neuropsychopharmacology 32, 1358–1369. [DOI] [PubMed] [Google Scholar]
  198. De Olmos JS, Heimer L (1999) The Concepts of the Ventral Striatopallidal System and Extended Amygdala. Ann. N. Y. Acad. Sci 877, 1–32. [DOI] [PubMed] [Google Scholar]
  199. Olsen RW, Sieghart W (2008) International Union of Pharmacology. LXX. Subtypes of gamma-aminobutyric acid(A) receptors: classification on the basis of subunit composition, pharmacology, and function. Update. Pharmacol. Rev 60, 243–60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Ouchi Y (2001) Alterations in binding site density of dopamine transporter in the striatum, orbitofrontal cortex, and amygdala in early Parkinson’s disease: compartment analysis for beta-CFT binding with positron emission tomography. Ann. Neurol 45, 601–610. [PubMed] [Google Scholar]
  201. Owen GS, Halliwell JV (2001) Electrophysiological characterization of laminar synaptic inputs to the olfactory tubercle of the rat studied in vitro: modulation of glutamatergic transmission by cholinergic agents is pathway-specific. Eur. J. Neurosci 13, 1767–1780. [DOI] [PubMed] [Google Scholar]
  202. Palacios J, Kuhar MJ (1982) Beta adrenergic receptor localization in rat brain by light microscopic autoradiography. Neurochem. Int 4, 473–490. [DOI] [PubMed] [Google Scholar]
  203. Pan W-X (2005) Dopamine Cells Respond to Predicted Events during Classical Conditioning: Evidence for Eligibility Traces in the Reward-Learning Network. J. Neurosci 25, 6235–6242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Paoletti P, Bellone C, Zhou Q (2013) NMDA receptor subunit diversity: Impact on receptor properties, synaptic plasticity and disease. Nat. Rev. Neurosci 14, 383–400. [DOI] [PubMed] [Google Scholar]
  205. Park J, Wakabayashi KT, Szalkowski C, Bhimani RV (2017) Heterogeneous Extracellular Dopamine Regulation in the Subregions of the Olfactory Tubercle. J. Neurochem 142, 365–377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Parsons SM, Prior C, Marshall IG (1993) Acetylcholine transport, storage, and release, in Int. Rev. Neurobiol, Vol. 35, pp. 279–390. Elsevier. [DOI] [PubMed] [Google Scholar]
  207. Patrick J, Séquéla P, Vernino S, Amador M, Luetje C, Dani JA (1993) Chapter 11: Functional diversity of neuronal nicotinic acetylcholine receptors, in Cholinergic Funct. Dysfunct, (Cuello A. C. B. T.-P. in B. R., ed), Vol. 98, pp. 113–120. Elsevier. [DOI] [PubMed] [Google Scholar]
  208. Payton CA, Wilson DA, Wesson DW (2012) Parallel Odor Processing by Two Anatomically Distinct Olfactory Bulb Target Structures. PLoS One 7, e34926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Pellegrini-Giampietro DE, Bennett MV, Zukin RS (2006) Differential expression of three glutamate receptor genes in developing rat brain: an in situ hybridization study. Proc. Natl. Acad. Sci 88, 4157–61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Petralia RS, Wenthold RJ (1992) Light and electron immunocytochemical localization of AMPA-selective glutamate receptors in the rat brain. J. Comp. Neurol 318, 329–354. [DOI] [PubMed] [Google Scholar]
  211. Petzold GC, Hagiwara A, Murthy VN (2009) Serotonergic modulation of odor input to the mammalian olfactory bulb. Nat Neurosci 12, 784–791. [DOI] [PubMed] [Google Scholar]
  212. Piekarzewska AB, Rosochacki SJ, Sender G (2000) The effect of acute restraint stress on regional brain neurotransmitter levels in stress-susceptible pietrain pigs. J. Vet. Med. A. Physiol. Pathol. Clin. Med 47, 257–69. [DOI] [PubMed] [Google Scholar]
  213. Pigache RM (1970) The anatomy of “paleocortex”. A critical review. Ergeb Anat Entwicklungsgesch 43, 3–62. [PubMed] [Google Scholar]
  214. Pirker S, Schwarzer C, Wieselthaler A, Sieghart W, Sperk G (2000) GABA(A) receptors: immunocytochemical distribution of 13 subunits in the adult rat brain. Neuroscience 101, 815–50. [DOI] [PubMed] [Google Scholar]
  215. Pisani A, Bonsi P, Centonze D, Calabresi P, Bernardi G (2000) Activation of D2-like dopamine receptors reduces synaptic inputs to striatal cholinergic interneurons. J. Neurosci 20, RC69–RC69. [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Pogun S, Baumann MH, Kuhar MJ (1994) Nitric oxide inhibits [3H]dopamine uptake. Brain Res. 641, 83–91. [DOI] [PubMed] [Google Scholar]
  217. Prasada Rao PD, Kanwal JS (2004) Oxytocin and Vasopressin Immunoreactivity within the Forebrain and Limbic-Related Areas in the Mustached Bat, Pteronotus parnellii. Brain. Behav. Evol 63, 151–168. [DOI] [PubMed] [Google Scholar]
  218. Proaño SB, Morris HJ, Kunz LM, Dorris DM, Meitzen J (2018) Estrous cycle-induced sex differences in medium spiny neuron excitatory synaptic transmission and intrinsic excitability in adult rat nucleus accumbens core. J. Neurophysiol 120, 1356–1373. [DOI] [PMC free article] [PubMed] [Google Scholar]
  219. Quik M, Bordia T, O’Leary K (2007) Nicotinic receptors as CNS targets for Parkinson’s disease. Biochem. Pharmacol 74, 1224–1234. [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Rainbow TC, Parsons B, MacLusky NJ, McEwen BS (1982) Estradiol receptor levels in rat hypothalamic and limbic nuclei. J. Neurosci 2, 1439–1445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  221. Rainbow TC, Parsons B, Wolfe BB (1984) Quantitative autoradiography of beta 1- and beta 2-adrenergic receptors in rat brain. Proc. Natl. Acad. Sci 81, 1585–1589. [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. Ramos-Moreno T, Galazo MJ, Porrero C, Martínez-Cerdeño V, Clascá F (2006) Extracellular matrix molecules and synaptic plasticity: Immunomapping of intracellular and secreted Reelin in the adult rat brain. Eur. J. Neurosci 23, 401–22. [DOI] [PubMed] [Google Scholar]
  223. Redgrave P, Prescott TJ, Gurney K (1999) Is the short-latency dopamine response too short to signal reward error? Trends Neurosci. 22, 146–151. [DOI] [PubMed] [Google Scholar]
  224. Revay R, Vaughan R, Grant S, Kuhar MJ (1996) Dopamine transporter immunohistochemistry in median eminence, amygdala, and other areas of the rat brain. Synapse 22, 93–99. [DOI] [PubMed] [Google Scholar]
  225. Rey NL, Steiner JA, Maroof N, Luk KC, Madaj Z, Trojanowski JQ, Lee VM-Y, Brundin P (2016) Widespread transneuronal propagation of α-synucleinopathy triggered in olfactory bulb mimics prodromal Parkinson’s disease. J. Exp. Med 213, 1759–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  226. Rey NL, Wesson DW, Brundin P (2018) The olfactory bulb as the entry site for prion-like propagation in neurodegenerative diseases. Neurobiol. Dis 109, 226–248. [DOI] [PMC free article] [PubMed] [Google Scholar]
  227. Riedel A, Härtig W, Seeger G, Gärtner U, Brauer K, Arendt T (2002) Principles of rat subcortical forebrain organization: a study using histological techniques and multiple fluorescence labeling. J. Chem. Neuroanat 23, 75–104. [DOI] [PubMed] [Google Scholar]
  228. Rodrigo J, Springall DR, Uttenthal O, Bentura ML, Abadia-Molina F, Riveros-Moreno V, Martinez-Murillo R, Polak JM, Moncada S (1994) Localization of nitric oxide synthase in the adult rat brain. Philos. Trans. R. Soc. B Biol. Sci 345, 175–221. [DOI] [PubMed] [Google Scholar]
  229. Roghani A, Shirzadi A, Butcher LL, Edwards RH (1997) Distribution of the vesicular transporter for acetylcholine in the rat central nervous system. Neuroscience 82, 1195–1212. [DOI] [PubMed] [Google Scholar]
  230. Roselli CE, Stormshak F (2012) Ontogeny of Cytochrome P450 Aromatase mRNA Expression in the Developing Sheep Brain. J. Neuroendocrinol 24, 443–452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  231. Rosin DL, Talley EM, Lee A, Stornetta RL, Gaylinn BD, Guyenet PG, Lynch KR (1996) Distribution of α2C-adrenergic receptor-like immunoreactivity in the rat central nervous system. J. Comp. Neurol 372, 135–165. [DOI] [PubMed] [Google Scholar]
  232. Ruat M, Traiffort E, Arrang JM, Tardivellacombe J, Diaz J, Leurs R, Schwartz JC (1993) A Novel Rat Serotonin (5-HT6) Receptor: Molecular Cloning, Localization and Stimulation of cAMP Accumulation. Biochem. Biophys. Res. Commun 193, 268–276. [DOI] [PubMed] [Google Scholar]
  233. Saalmann YB, Kirkcaldie MTK, Waldron S, Calford MB (2007) Cellular Distribution of the GABA A Receptor-Modulating 3?-Hydroxy, 5?-Reduced Pregnane Steroids in the Adult Rat Brain. J. Neuroendocrinol 19, 272–284. [DOI] [PubMed] [Google Scholar]
  234. Sabol KE, Roach JT, Broom SL, Ferreira C, Preau MM (2001) Long-term effects of a high-dose methamphetamine regimen on subsequent methamphetamine-induced dopamine release in vivo. Brain Res. 892, 122–9. [DOI] [PubMed] [Google Scholar]
  235. Saiz-Sanchez D, La Rosa-Prieto C. De, Ubeda-Bañon I, Martinez-Marcos A. (2013) Interneurons and Beta-Amyloid in the Olfactory Bulb, Anterior Olfactory Nucleus and Olfactory Tubercle in APPxPS1 Transgenic Mice Model of Alzheimer’s Disease. Anat. Rec 296, 1413–1423. [DOI] [PubMed] [Google Scholar]
  236. Saldanha CJ, Remage-Healey L, Schlinger BA (2011) Synaptocrine Signaling: Steroid Synthesis and Action at the Synapse. Endocr. Rev 32, 532–549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  237. Sanders JD, Happe HK, Bylund DB, Murrin LC (2005) Development of the norepinephrine transporter in the rat CNS. Neuroscience 130, 107–117. [DOI] [PubMed] [Google Scholar]
  238. Sarnyai Z, Babarczy E, Kriván M, Szabó G, Kovács GL, Barth T, Telegdy G (1991) Selective attenuation of cocaine-induced stereotyped behaviour by oxytocin: putative role of basal forebrain target sites. Neuropeptides 19, 51–6. [DOI] [PubMed] [Google Scholar]
  239. Schäfer MKH, Weihe E, Varoqui H, Eiden LE, Erickson JD (1994) Distribution of the vesicular acetylcholine transporter (VAChT) in the central and peripheral nervous systems of the rat. J. Mol. Neurosci 5, 1–26. [DOI] [PubMed] [Google Scholar]
  240. Scheinin M, Lomasney JW, Hayden-Hixson DM, Schambra UB, Caron MG, Lefkowitz RJ, Fremeau RT (1994) Distribution of α2-adrenergic receptor subtype gene expression in rat brain. Mol. Brain Res 21, 133–149. [DOI] [PubMed] [Google Scholar]
  241. Schroeder LE, Furdock R, Quiles CR, Kurt G, Perez-Bonilla P, Garcia A, Colon-Ortiz C, Brown J, Bugescu R, Leinninger GM (2019) Mapping the populations of neurotensin neurons in the male mouse brain. Neuropeptides 76, 101930. [DOI] [PMC free article] [PubMed] [Google Scholar]
  242. Schumacher M, Mattern C, Ghoumari A, Oudinet JP, Liere P, Labombarda F, Sitruk-Ware R, De Nicola AF, Guennoun R (2014) Revisiting the roles of progesterone and allopregnanolone in the nervous system: Resurgence of the progesterone receptors. Prog. Neurobiol 113, 6–39. [DOI] [PubMed] [Google Scholar]
  243. Schwarzer C, Berresheim U, Pirker S, Wieselthaler A, Fuchs K, Sieghart W, Sperk G (2001) Distribution of the major gamma-aminobutyric acid(A) receptor subunits in the basal ganglia and associated limbic brain areas of the adult rat. J. Comp. Neurol 433, 526–49. [DOI] [PubMed] [Google Scholar]
  244. Schwob JE, Price JL (1984a) The development of lamination of afferent fibers to the olfactory cortex in rats, with additional observations in the adult. J Comp Neurol 223, 203–222. [DOI] [PubMed] [Google Scholar]
  245. Schwob JE, Price JL (1984b) The development of axonal connections in the central olfactory system of rats. J Comp Neurol 223, 177–202. [DOI] [PubMed] [Google Scholar]
  246. Scott JW (1981) Electrophysiological identification of mitral and tufted cells and distributions of their axons in olfactory system of the rat. J Neurophysiol 46, 918–931. [DOI] [PubMed] [Google Scholar]
  247. Scott JW, McBride RL, Schneider SP (1980) The organization of projections from the olfactory bulb to the piriform cortex and olfactory tubercle in the rat. J Comp Neurol 194, 519–534. [DOI] [PubMed] [Google Scholar]
  248. Seifert U, Härtig W, Grosche J, Brückner G, Riedel A, Brauer K (1998) Axonal expression sites of tyrosine hydroxylase, calretinin- and calbindin-immunoreactivity in striato-pallidal and septal nuclei of the rat brain: a double-immunolabelling study. Brain Res. 795, 227–246. [DOI] [PubMed] [Google Scholar]
  249. Shashidharan P, Huntley GW, Murray JM, Buku A, Moran T, Walsh MJ, Morrison JH, Plaitakis A (1997) Immunohistochemical localization of the neuron-specific glutamate transporter EAAC1 (EAAT3) in rat brain and spinal cord revealed by a novel monoclonal antibody. Brain Res. 773, 139–48. [DOI] [PubMed] [Google Scholar]
  250. Shin R, Qin M, Liu Z-H, Ikemoto S (2008) Intracranial self-administration of MDMA into the ventral striatum of the rat: differential roles of the nucleus accumbens shell, core, and olfactory tubercle. Psychopharmacology (Berl). 198, 261–270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  251. Smith JE, Vaughn TC, Co C (2003) Acetylcholine turnover rates in rat brain regions during cocaine self-administration. J. Neurochem 88, 502–512. [DOI] [PubMed] [Google Scholar]
  252. Soares-Cunha C, Coimbra B, Sousa N, Rodrigues AJ (2016) Reappraising striatal D1- and D2-neurons in reward and aversion. Neurosci. Biobehav. Rev 68, 370–386. [DOI] [PubMed] [Google Scholar]
  253. Sofroniew MV (1983) Morphology of Vasopressin and Oxytocin Neurones and Their Central and Vascular Projections, in Neurohypophysis Struct. Funct. Control, (Cross BA, Leng G. B. T.-P. in B. R., eds), Vol. 60, pp. 101–114. Elsevier. [DOI] [PubMed] [Google Scholar]
  254. Solano-Flores LP, Aguilar-Baturoni HU, Guevara-Aguilar R (1980) Locus coeruleus influences upon the olfactory tubercle. Brain Res Bull 5, 383–389. [DOI] [PubMed] [Google Scholar]
  255. Sonders MS, Zhu S-J, Zahniser NR, Kavanaugh MP, Amara SG (1997) Multiple Ionic Conductances of the Human Dopamine Transporter: The Actions of Dopamine and Psychostimulants. J. Neurosci 17, 960–974. [DOI] [PMC free article] [PubMed] [Google Scholar]
  256. Sorger D, Scheunemann M, Vercouillie J, Großmann U, Fischer S, Hiller A, Wenzel B, et al. (2009) Neuroimaging of the vesicular acetylcholine transporter by a novel 4-[18F]fluoro-benzoyl derivative of 7-hydroxy-6-(4-phenyl-piperidin-1-yl)-octahydro-benzo[1,4]oxazines. Nucl. Med. Biol 36, 17–27. [DOI] [PubMed] [Google Scholar]
  257. Standaert DG, Friberg IK, Landwehrmeyer GB, Young AB, Penney JB (1999) Expression of NMDA glutamate receptor subunit mRNAs in neurochemically identified projection and interneurons in the striatum of the rat. Mol. Brain Res 64, 11–23. [DOI] [PubMed] [Google Scholar]
  258. Stanić D, Dubois S, Chua HK, Tonge B, Rinehart N, Horne MK, Boon WC (2014) Characterization of Aromatase Expression in the Adult Male and Female Mouse Brain. I. Coexistence with Oestrogen Receptors α and β, and Androgen Receptors. PLoS One 9, e90451. [DOI] [PMC free article] [PubMed] [Google Scholar]
  259. Strazielle C, Lalonde R, Hébert C, Reader TA (1999) Regional brain distribution of noradrenaline uptake sites, and of α1-, α2- and β-adrenergic receptors in PCD mutant mice: a quantitative autoradiographic study. Neuroscience 94, 287–304. [DOI] [PubMed] [Google Scholar]
  260. Strosberg AD (1993) Structure, function, and regulation of adrenergic receptors. Protein Sci. 2, 1198–1209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  261. Sumner BE ., Fink G (1997) The density of 5-hydoxytryptamine2A receptors in forebrain is increased at pro-oestrus in intact female rats. Neurosci. Lett 234, 7–10. [DOI] [PubMed] [Google Scholar]
  262. Sumner BEH, Grant KE, Rosie R, Hegele-Hartung C, Fritzemeier K-H, Fink G (2007) Raloxifene blocks estradiol induction of the serotonin transporter and 5-hydroxytryptamine2A receptor in female rat brain. Neurosci. Lett 417, 95–99. [DOI] [PubMed] [Google Scholar]
  263. Surmeier DJ, Ding J, Day M, Wang Z, Shen W (2007) D1 and D2 dopamine-receptor modulation of striatal glutamatergic signaling in striatal medium spiny neurons. Trends Neurosci. 30, 228–235. [DOI] [PubMed] [Google Scholar]
  264. Surmeier DJ, Song W-J, Yan Z (1996) Coordinated Expression of Dopamine Receptors in Neostriatal Medium Spiny Neurons. J. Neurosci 16, 6579–91. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. Swanson LW, Sawchenko PE (1983) Hypothalamic Integration: Organization of the Paraventricular and Supraoptic Nuclei. Annu. Rev. Neurosci 6, 269–324. [DOI] [PubMed] [Google Scholar]
  266. Talbot K, Woolf NJ, Butcher LL (1988) Feline islands of Calleja complex: II. Cholinergic and cholinesterasic features. J Comp Neurol 275, 580–603. [DOI] [PubMed] [Google Scholar]
  267. Tamaru Y, Nomura S, Mizuno N, Shigemoto R (2001) Distribution of metabotropic glutamate receptor mGluR3 in the mouse CNS: differential location relative to pre- and postsynaptic sites. Neuroscience 106, 481–503. [DOI] [PubMed] [Google Scholar]
  268. Torres GE, Gainetdinov RR, Caron MG (2003) Plasma membrane monoamine transporters: structure, regulation and function. Nat Rev Neurosci 4, 13–25. [DOI] [PubMed] [Google Scholar]
  269. Tozzi A, A de I, M D. F, Tantucci M, Costa C (2011) The distinct role of medium spiny neurons and cholinergic interneurons in the D2/A2A receptor interaction in the striatum: implications for Parkinson’s disease. J. Neurosci 31, 1850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  270. Tribollet E, Audigier S, Dubois-Dauphin M, Dreifuss J (1990) Gonadal steroids regulate oxytocin receptors but not vasopressin receptors in the brain of male and female rats. An autoradiographical study. Brain Res. 511, 129–40. [DOI] [PubMed] [Google Scholar]
  271. Tribollet E, Barberis C, Jard S, Dubois-Dauphin M, Dreifuss J (1988) Localization and pharmacological characterization of high affinity binding sites for vasopressin and oxytocin in the rat brain by light microscopic autoradiography. Brain Res. 442, 105–118. [DOI] [PubMed] [Google Scholar]
  272. Tribollet E, Charpak S, Schmidt A, Dubois-Dauphin M, Dreifuss J (1989) Appearance and transient expression of oxytocin receptors in fetal, infant, and peripubertal rat brain studied by autoradiography and electrophysiology. J. Neurosci 9, 1764–1773. [DOI] [PMC free article] [PubMed] [Google Scholar]
  273. Trudeau L-E, El Mestikawy S (2018) Glutamate Cotransmission in Cholinergic, GABAergic and Monoamine Systems: Contrasts and Commonalities. Front. Neural Circuits 12, 113. [DOI] [PMC free article] [PubMed] [Google Scholar]
  274. Tuscher JJ, Szinte JS, Starrett JR, Krentzel AA, Fortress AM, Remage-Healey L, Frick KM (2016) Inhibition of local estrogen synthesis in the hippocampus impairs hippocampal memory consolidation in ovariectomized female mice. Horm. Behav 83, 60–67. [DOI] [PMC free article] [PubMed] [Google Scholar]
  275. Ubeda-Bañon I, Novejarque A, Mohedano-Moriano A, Pro-Sistiaga P, Insausti R, Martinez-Garcia F, Lanuza E, Martinez-Marcos A (2008) Vomeronasal inputs to the rodent ventral striatum. Brain Res. Bull 75, 467–473. [DOI] [PubMed] [Google Scholar]
  276. Ubeda-Bañon I, Saiz-Sanchez D, de la Rosa-Prieto C, Argandoña-Palacios L, Garcia-Muñozguren S, Martinez-Marcos A (2010) α-Synucleinopathy in the human olfactory system in Parkinson’s disease: involvement of calcium-binding protein- and substance P-positive cells. Acta Neuropathol. 119, 723–735. [DOI] [PubMed] [Google Scholar]
  277. Uematsu A, Tan BZ, Ycu EA, Cuevas JS, Koivumaa J, Junyent F, Kremer EJ, Witten IB, Deisseroth K, Johansen JP (2017) Modular organization of the brainstem noradrenaline system coordinates opposing learning states. Nat. Neurosci 20, 1602. [DOI] [PubMed] [Google Scholar]
  278. Usdin TB, Eiden LE, Bonner TI, Erickson JD (1995) Molecular biology of the vesicular ACh transporter. Trends Neurosci. 18, 218–224. [DOI] [PubMed] [Google Scholar]
  279. Vaccari C, Lolait SJ, Ostrowski NL (1998) Comparative Distribution of Vasopressin V1b and Oxytocin Receptor Messenger Ribonucleic Acids in Brain. Endocrinology 139, 5015–5033. [DOI] [PubMed] [Google Scholar]
  280. Vaidyanathan R, Hammock EAD (2017) Oxytocin receptor dynamics in the brain across development and species. Dev. Neurobiol 77, 143–157. [DOI] [PubMed] [Google Scholar]
  281. Vastenhouw B, Van Der Have F, Van Der Linden AJA, Von Oerthel L, Booij JPH, Burbach J, Smidt MP, Beekman FJ (2007) Movies of dopamine transporter occupancy with ultra-high resolution focusing pinhole SPECT. Mol. Psychiatry 12, 984–987. [DOI] [PubMed] [Google Scholar]
  282. Vaughan RA, Brown VL, McCoy MT, Kuhar MJ (1996) Species and brain region specific dopamine transporters: immunological and glycosylation characteristics. J. Neurochem 66, 2146–2152. [DOI] [PubMed] [Google Scholar]
  283. De Vente J, Hopkins DA, Markerink-Van Ittersum M, Steinbusch HWM (1998) Nitric oxide-mediated cGMP production in the islands of Calleja in the rat. Brain Res. 789, 175–178. [DOI] [PubMed] [Google Scholar]
  284. Vilaró MT, Cortés R, Gerald C, Branchek TA, Palacios JM, Mengod G (1996) Localization of 5-HT4 receptor mRNA in rat brain by in situ hybridization histochemistry. Mol. Brain Res 43, 356–360. [DOI] [PubMed] [Google Scholar]
  285. Vilaró MT, Wiederhold KH, Palacios JM, Mengod G (1991) Muscarinic cholinergic receptors in the rat caudate-putamen and olfactory tubercle belong predominantly to the m4 class: In situ hybridization and receptor autoradiography evidence. Neuroscience 40, 159–167. [DOI] [PubMed] [Google Scholar]
  286. Vilaró MT, Wiederhold K, Palacios JM, Mengod G (1992) Muscarinic M2‐selective ligands also recognize M4 receptors in the rat brain: Evidence from combined in situ hybridization and receptor autoradiography. Synapse 11, 171–183. [DOI] [PubMed] [Google Scholar]
  287. Vincent SR, Kimura H (1992) Histochemical mapping of nitric oxide synthase in the rat brain. Neuroscience 46, 755–784. [DOI] [PubMed] [Google Scholar]
  288. Voigt MM, Laurie DJ, Seeburg PH, Bach A (1991) Molecular cloning and characterization of a rat brain cDNA encoding a 5-hydroxytryptamine1B receptor. EMBO J. 10, 4017–4023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  289. Wada E, Shigemoto R, Kinoshita A, Ohishi H, Mizuno N (1998) Metabotropic glutamate receptor subtypes in axon terminals of projection fibers from the main and accessory olfactory bulbs: a light and electron microscopic immunohistochemical study in the rat. J Comp Neurol 393, 493–504. [PubMed] [Google Scholar]
  290. Waeber C, Sebben M, Bockaert J, Dumuis A (1996) Regional distribution and ontogeny of 5-HT4 binding sites in rat brain. Behav. Brain Res 73, 259–262. [DOI] [PubMed] [Google Scholar]
  291. Waldhoer M, Bartlett SE, Whistler JL (2004) Opioid Receptors. Annu. Rev. Biochem 73, 953–990. [DOI] [PubMed] [Google Scholar]
  292. Wang Y.-h, Bosy TZ, Yasuda RP, Grayson DR, Vicini S, Pizzorusso T, Wolfe BB. (1995) Characterization of NMDA Receptor Subunit-Specific Antibodies: Distribution of NR2A and NR2B Receptor Subunits in Rat Brain and Ontogenic Profile in the Cerebellum. J. Neurochem 65, 176–183. [DOI] [PubMed] [Google Scholar]
  293. Watabe-Uchida M, Zhu L, Ogawa SK, Vamanrao A, Uchida N (2012) Whole-Brain Mapping of Direct Inputs to Midbrain Dopamine Neurons. Neuron 74, 858–873. [DOI] [PubMed] [Google Scholar]
  294. Weathington JM, Hamki A, Cooke BM (2014) Sex- and region-specific pubertal maturation of the forebrain CRF receptor system in the rat. J. Comp. Neurol 522, 1284–98. [DOI] [PubMed] [Google Scholar]
  295. Weiner DM, Levey AI, Brann MR (1990) Expression of muscarinic acetylcholine and dopamine receptor mRNAs in rat basal ganglia. Proc. Natl. Acad. Sci. U. S. A 87, 7050–4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  296. Wess J (1996) Molecular biology of muscarinic acetylcholine receptors. Crit. Rev. Neurobiol 10, 69–99. [DOI] [PubMed] [Google Scholar]
  297. Wesson DW, Levy E, Nixon RA, Wilson DA (2010a) Olfactory Dysfunction Correlates with β-Amyloid Plaque Burden in an Alzheimer’s Disease Mouse Model. J Neurosci 30, 505–514. [DOI] [PMC free article] [PubMed] [Google Scholar]
  298. Wesson DW, Wilson DA (2010) Smelling Sounds: Olfactory-Auditory Sensory Convergence in the Olfactory Tubercle. J Neurosci 30, 3013–3021. [DOI] [PMC free article] [PubMed] [Google Scholar]
  299. Wesson DW, Wilson DA (2011) Sniffing out the contributions of the olfactory tubercle to the sense of smell: hedonics, sensory integration, and more? Neurosci Biobehav Rev 35, 655–668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  300. Wesson DW, Wilson DA, Nixon RA (2010b) Should olfactory dysfunction be used as a biomarker of Alzheimer’s disease? Expert Rev. Neurother 10, 633–635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  301. White KA, Zhang Y-F, Zhang Z, Bhattarai JP, Moberly AH, in ‘t Zandt E, Peña-Bravo JI, et al. (2019) Glutamatergic neurons in the piriform cortex influence the activity of D1 and D2-type receptor expressing olfactory tubercle neurons. J. Neurosci, 1444–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  302. White LE (1965) Olfactory bulb projections of the rat. Anat Rec 152, 465–479. [Google Scholar]
  303. Whiteaker P, Cooper JF, Salminen O, Marks MJ, Mcclure‐Begley TD, Brown RWB, Collins AC, Lindstrom JM (2006) Immunolabeling demonstrates the interdependence of mouse brain α4 and β2 nicotinic acetylcholine receptor subunit expression. J. Comp. Neurol 499, 1016–1038. [DOI] [PubMed] [Google Scholar]
  304. Whittaker VP (1988) Model cholinergic systems: an overview, in Cholinergic Synap, pp. 3–22. Springer. [Google Scholar]
  305. Wieland S, Du D, Oswald MJ, Parlato R, Kohr G, Kelsch W (2014) Phasic Dopaminergic Activity Exerts Fast Control of Cholinergic Interneuron Firing via Sequential NMDA, D2, and D1 Receptor Activation. J. Neurosci 34, 11549–11559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  306. Wieland S, Schindler S, Huber C, Köhr G, Oswald MJ, Kelsch W (2015) Phasic Dopamine Modifies Sensory-Driven Output of Striatal Neurons through Synaptic Plasticity. J. Neurosci 35, 9946–9956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Wirtshafter D, Osborn CV (2004) The distribution of m4 muscarinic acetylcholine receptors in the islands of Calleja and striatum of rats and cynomolgus monkeys. J. Chem. Neuroanat 28, 107–116. [DOI] [PubMed] [Google Scholar]
  308. Wise RA (2004) Dopamine, learning and motivation. Nat Rev Neurosci 5, 483–494. [DOI] [PubMed] [Google Scholar]
  309. Woodhams PL, Allen YS, McGovern J, Allen JM, Bloom SR, Balazs R, Polak JM (1985) Immunohistochemical analysis of the early ontogeny of the neuropeptide Y system in rat brain. Neuroscience 15, 173–202. [DOI] [PubMed] [Google Scholar]
  310. Wright DE, Seroogy KB, Lundgren KH, Davis BM, Jennes L (1995) Comparative localization of serotonin1A, 1C, and2 receptor subtype mRNAs in rat brain. J. Comp. Neurol 351, 357–373. [DOI] [PubMed] [Google Scholar]
  311. Xiong A, Wesson DW (2016) Illustrated Review of the Ventral Striatum’s Olfactory Tubercle. Chem. Senses 41, 549–555. [DOI] [PMC free article] [PubMed] [Google Scholar]
  312. Yan Z, Flores-Hernandez J, Surmeier DJ (2001) Coordinated expression of muscarinic receptor messenger RNAs in striatal medium spiny neurons. Neuroscience 103, 1017–24. [DOI] [PubMed] [Google Scholar]
  313. Young LJ, Wang Z, Donadlson R, Rissman EF (1998) Estrogen receptor α is essential for induction of oxytocin receptor by estrogen. Neuroendocrinology 9, 933–36. [DOI] [PubMed] [Google Scholar]
  314. Záborszky L, Alheid GF, Beinfeld MC, Eiden LE, Heimer L, Palkovits M (1985) Cholecystokinin innervation of the ventral striatum: A morphological and radioimmunological study. Neuroscience 14, 427–453. [DOI] [PubMed] [Google Scholar]
  315. Zahm DS, Grosu S, Irving JC, Williams EA (2003) Discrimination of striatopallidum and extended amygdala in the rat: a role for parvalbumin immunoreactive neurons? Brain Res. 978, 141–154. [DOI] [PubMed] [Google Scholar]
  316. Zelano C, Bensafi M, Porter J, Mainland J, Johnson B, Bremner E, Telles C, Khan R, Sobel N (2005) Attentional modulation in human primary olfactory cortex. Nat Neurosci 8, 114–120. [DOI] [PubMed] [Google Scholar]
  317. Zhang J‐H, Sato M, Tohyama M (1991) Different postnatal development profiles of neurons containing distinct GABAA receptor β subunit mRNAs (β1, β2, and β3) in the rat forebrain. J. Comp. Neurol 308, 586–613. [DOI] [PubMed] [Google Scholar]
  318. Zhang JH, Sato M, Noguchi K, Tohyama M (1990) The differential expression patterns of the mRNAs encoding β subunits (β1, β2 and β3) of GABAA receptor in the olfactory bulb and its related areas in the rat brain. Neurosci. Lett 119, 257–60. [DOI] [PubMed] [Google Scholar]
  319. Zhang Z, Liu Q, Wen P, Zhang J, Rao X, Zhou Z, Zhang H, et al. (2017a) Activation of the dopaminergic pathway from VTA to the medial olfactory tubercle generates odor-preference and reward. Elife 6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  320. Zhang Z, Zhang H, Wen P, Zhu X, Wang L, Liu Q, Wang J, He X, Wang H, Xu F (2017b) Whole-Brain Mapping of the Inputs and Outputs of the Medial Part of the Olfactory Tubercle. Front. Neural Circuits 11, 52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  321. Zhou L, Furuta T, Kaneko T (2003) Chemical organization of projection neurons in the rat accumbens nucleus and olfactory tubercle. Neuroscience 120, 783–798. [DOI] [PubMed] [Google Scholar]
  322. Zhou L, Zhu DY (2009) Neuronal nitric oxide synthase: Structure, subcellular localization, regulation, and clinical implications. Nitric Oxide - Biol. Chem 20, 223–230.sss [DOI] [PubMed] [Google Scholar]
  323. Zhou Y, Danbolt NC (2013) GABA and Glutamate Transporters in Brain. Front. Endocrinol. (Lausanne). 4, 165. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES