Skip to main content
Journal of Neurophysiology logoLink to Journal of Neurophysiology
. 2020 May 13;123(6):2406–2425. doi: 10.1152/jn.00034.2020

Serotonergic modulation across sensory modalities

Tyler R Sizemore 1,, Laura M Hurley 2, Andrew M Dacks 1,3
PMCID: PMC7311732  PMID: 32401124

Abstract

The serotonergic system has been widely studied across animal taxa and different functional networks. This modulatory system is therefore well positioned to compare the consequences of neuromodulation for sensory processing across species and modalities at multiple levels of sensory organization. Serotonergic neurons that innervate sensory networks often bidirectionally exchange information with these networks but also receive input representative of motor events or motivational state. This convergence of information supports serotonin’s capacity for contextualizing sensory information according to the animal’s physiological state and external events. At the level of sensory circuitry, serotonin can have variable effects due to differential projections across specific sensory subregions, as well as differential serotonin receptor type expression within those subregions. Functionally, this infrastructure may gate or filter sensory inputs to emphasize specific stimulus features or select among different streams of information. The near-ubiquitous presence of serotonin and other neuromodulators within sensory regions, coupled with their strong effects on stimulus representation, suggests that these signaling pathways should be considered integral components of sensory systems.

Keywords: auditory system, comparative study, olfactory system, sensory processing, serotonin

NEUROMODULATION OF SENSORY PROCESSING

A fundamental concept explored by work in motor systems, such as the stomatogastric ganglion (STG) of decapod crustaceans, is that modulatory signaling chemicals represent a diverse tool set to alter network activity. Modulators can have immediate, or latent, effects on the biophysical or synaptic properties of a neuron across adjacent synapses and/or nonadjacent synapses, and these effects do not necessarily change the membrane potential (Bargmann 2012; Katz 1999; Marder 2012). The diverse effects of neuromodulation allow neural networks, even those with relatively few neurons, to produce a wide range of different functional outputs. These core principles from early work in STG and motor systems have emerged as essential features for every aspect of neural function, including sensory processing.

Sensory systems internalize and process information from the environment to form a neural representation of an animal’s surroundings. However, all animals experience fluctuations in their ecology and internal state. To appropriately adjust behavior according to these fluctuations, neuromodulation endows the nervous system with the capacity to alter neural function at every level (synaptic, circuit, network, etc.) without necessarily adding new neurons. Neuromodulation supports effective sensory processing in a variable environment, allowing tuning of existing neural circuitry for flexible network output and consequently dynamic behavioral output. The goal of this review is to discuss fundamental features of neuromodulation of sensory processing by one neuromodulator, serotonin. Sensory neuromodulation by serotonergic neurons has been studied extensively in diverse taxa and different sensory systems and therefore offers an excellent opportunity to highlight common principles. Although we focus on the influence of serotonin on sensory processing, these common principles will hold true for other neuromodulatory systems that have comparable effects at cellular, circuit, and functional levels (McBurney-Lin et al. 2019; Schofield and Hurley 2018). It should be noted that this review is by no means exhaustive, and we have limited ourselves to the first few processing stages from the periphery in most cases. We provide numerous examples across taxa and sensory modalities that highlight several key aspects of the serotonergic system. These aspects include the anatomical and physiological characteristics of serotonergic neurons, serotonergic activity relative to context, and the pattern of serotonin receptor expression within sensory circuitry. Ultimately, these components of the serotonergic system enable serotonin to modulate the neuronal representation of sensory events.

SEROTONIN SOURCES ACROSS SENSORY SYSTEMS

Serotonin is an ancient and pervasive signaling molecule that acts in nearly every sensory system across diverse taxa (Gaspar and Lillesaar 2012; Hay-Schmidt 2000; Peroutka and Howell 1994). Concordantly, this single molecule has been implicated in a variety of broad state descriptors such as arousal, mood, and motivation (Cools et al. 2008; Luo et al. 2016; Miyazaki et al. 2012; Monti 2011). In general, there are two major sources of serotonin in mammals: gut derived and brain derived. The majority of the serotonin in the mammalian body is produced in the gut, chiefly by enterochromaffin cells, then absorbed by platelets and circulated throughout the periphery (reviewed in Bertrand and Bertrand 2010; Gershon 2013; Matthes and Bader 2018; Ni and Watts 2006). However, gut-derived serotonin does not appear to cross the blood-brain barrier (reviewed in Berger et al. 2009; El-Merahbi et al. 2015); therefore this review chiefly focuses on the consequences of brain-derived serotonin on sensory system operations across taxa.

There are ~26,000 neurons in the mouse and rat brain that produce serotonin (Table 1), but the majority of these serotoninergic neurons (~17,000 neurons) are collectively referred to as the raphe nuclei (Ishimura et al. 1988; Steinbusch 1981; Steinbusch and Nieuwenhuys 1981; Vertes and Crane 1997). The raphe nuclei can be further divided into several subpopulations, such as the dorsal raphe nucleus (DRN). In mice, the DRN constitutes the majority of serotonergic neurons in the brain (~9,000 serotonergic neurons; Ishimura et al. 1988; Ren et al. 2018; also see Hornung 2010). These DRN neurons, together with neurons from the median raphe nucleus (MRN), innervate and modulate every sensory processing center (Azmitia and Segal 1978; Doty 1983; Hurley et al. 2004; Jacob and Nienborg 2018; Jacobs and Azmitia 1992; McLean and Shipley 1987; Muzerelle et al. 2016; Takeuchi et al. 1982; Törk 1990).

Table 1.

Serotonergic neuron/perikarya estimates across taxa

Order Species Serotonergic Neuron Estimates Method Reference
Anaspidea Aplysia californica 120 α (Ono and McCaman 1984)
Arthrotardigrada Batillipes pennaki 10 α (Schulze et al. 2014)
Actinarctus doryphorus ocellatus 5
Blattodea Periplaneta americana 125* α (Bishop and O’Shea 1983; Klemm et al. 1984)
Calanoida Calanus finmarchicus 24 α (Hartline and Christie 2010)
Decapoda Homarus americanus 100 α (Beltz and Kravitz 1983)
Hyas araneus (zoea 1) 55*§ α (Harzsch and Dawirs 1995)
Hyas araneus (zoea 2) 110*§
Diptera Calliphora erythrocephalus 182§ α and β (Cantera and Nässel 1987; Nässel 1988)
Calliphora erythrocephalus (larvae) 97 α and β (Cantera and Nässel 1987)
Drosophila melanogaster 106 α (Vallés and White 1988)
Drosophila melanogaster (larvae) 84
Haplotaxida Allonais paraguayensis 2 α (Zattara and Bely 2015)
Amphichaeta sp. 2
Chaetogaster sp. 0
Dero digitata 4
Dero furcata 6
Monopylephorus rubroniveus 2
Nais stolci 8
Paranais litoralis 10
Pristina aequiseta 2
Pristina leidyi 2
Stylaria lacustris 4
Tubifex 6
Hemiptera Rhodnius prolixus 150 α (Lange et al. 1988)
Triatoma infestans 286§ α (Settembrini and Villar 2004)
Hirudinida Hirudo medicinalis 214 γ (Lent et al. 1991)
Hymenoptera Apis mellifera 75 α (Schürmann and Klemm 1984)
Isopoda Armadillidium vulgare 80 α (Thompson et al. 1994b)
Asellus meridianus 80
Ligia oceanica 80
Oniscus asellus 60
Ixodida Amblyomma americanum 69 α (Hummel et al. 2007)
Dermacentor albipictus 76
Lepidoptera Helicoverpa armigera (larvae) 60 α (Tang et al. 2019)
Manduca sexta 60 α (Homberg and Hildebrand 1989a, 1989b)
Odonata Epitheca sp. (nymph) 32§ α (Longley and Longley 1986)
Pachydiplax longipennis (nymph) 32§
Opisthopora Lumbricus terrestris 1,729 α (Spörhase-Eichmann et al. 1987a, 1987b)
Petromyzontiformes Petromyzon marinus 1,225 α (Antri et al. 2006)
Primates Homo sapiens 250,000 α (Baker et al. 1990; Baker et al. 1991)
Rhabditida Caenorhabditis elegans 16 α and δ (Desai et al. 1988; Duerr et al. 1999; Horvitz et al. 1982; Loer and Kenyon 1993; Sze et al. 2000)
Rodentia Mus musculus 25,824 β (Ishimura et al. 1988)
Rattus norvegicus 25,861 α (Vertes and Crane 1997)
Xiphosura Limulus polyphemus 158 α (Battelle et al. 1999)

Method(s) used in each reference: α, immunohistochemistry; β, peroxidase-antiperoxidase fluorescence; γ, glyoxylic acid-induced histochemistry; δ, formaldehyde-induced fluorescence. The following symbols indicate where certain numerical estimates correspond to:

central nervous system;

central brain;

*

central brain without optic lobes;

§

ventral nerve cord/spine.

Invertebrate brains contain far fewer neurons than vertebrate brains. For instance, the central nervous system of Caenorhabditis elegans has 302 neurons and Drosophila has ~135,000 neurons relative to ~70 million neurons estimated in mice (Bates et al. 2019; Chiang et al. 2011; Cook et al. 2019; Emmons 2015; Herculano-Houzel et al. 2006; Kohl and Jefferis 2011; Meinertzhagen 2018; Schlegel et al. 2017; White et al. 1986; Zheng et al. 2018). Despite having orders of magnitude fewer neurons, invertebrate sensory systems must accomplish the same fundamental neural computations as those of vertebrates. This notion extends to invertebrate serotonergic modulatory networks, where there are typically far fewer serotonergic neurons (Table 1) but many of the mechanisms for how serotonin modulates sensory processing are conserved. The Drosophila central brain, for instance, contains only ~90 serotonergic neurons (Vallés and White 1988), and only 2 widely projecting neurons (the “CSDns”) provide synaptic serotonin to the primary olfactory center, the antennal lobe (Dacks et al. 2006a; Roy et al. 2007). Despite having only 2 serotonergic neurons, compared with >100 neurons that innervate the vertebrate olfactory bulb from the DRN (Ren et al. 2018) and MRN (Muzerelle et al. 2016), serotonin modulates similar aspects of olfactory encoding in these taxa (see below). Thus, because vertebrate and invertebrate sensory systems must solve similar problems, comparing across taxa can reveal fundamental motifs of neuromodulation of sensory processing.

Although much of this review details the consequences of synaptic release of serotonin, serotonergic neurons, like other modulatory neurons, do not have to form a synapse with a given cell to modulate the cell’s activity (Eid et al. 2013; Fuxe et al. 2015). Serotonergic neurons have long been noted to use volume or bulk transmission as a means to release serotonin over large distances (sometimes >100 μm) and extended epochs (on the order of seconds) (Agnati et al. 1995; Beaudet and Descarries 1981; Bunin and Wightman 1998, 1999; Chazal and Ralston 1987; Gaudry 2018; Hornung 2010). For instance, in the cat auditory cortex most of the serotonergic boutons lack conventional synapses (DeFelipe et al. 1991). This principle extends across taxa, such as in the visual system of the house fly, Calliphora. Here, serotonergic processes are separated from other neurons by glia, lack synaptic specializations, and are dense core vesicle rich (a hallmark of bulk transmission) (Nässel et al. 1985). In addition, blood-borne serotonin contributes to sensory processing. For instance, blood-borne serotonin activates nociceptors (Sommer 2004; Viguier et al. 2013), is implicated in enhanced olfactory gain control (Suzuki et al. 2020; Zhang and Gaudry 2016), and potentially activates insect peripheral sensory structures. The CSDns in Drosophila do not directly synapse onto the olfactory sensory neurons (Coates et al. 2017), yet these neurons express the excitatory 5-HT2B receptor (Sizemore and Dacks 2016). In experiments where the olfactory sensory neuron somata were removed for performing antennal nerve shock, bath application of serotonin did not directly affect activity measured at the axon terminals of these neurons (Dacks et al. 2009). Therefore, the 5-HT2B receptor may localize to olfactory afferent soma in the antennae where serotonin in the hemolymph can act on them. Indeed, there are serotonergic fibers in the periphery (Vallés and White 1988), and the antennal hearts of many insects constantly circulate hemolymph into these olfactory appendages (Miller 1950; Pass 2000; Zhukovskaya and Polyanovsky 2017). Endocrine release of serotonin could coordinate the activity of olfactory afferents according to the animal’s current needs or directly modulate the gain of olfactory afferents. Alternatively, these mechanisms may operate as redundancies for the similar effects of presynaptic serotonin on afferents in other sensory systems (see below). Overall, bulk transmission of serotonin, whether endocrine or paracrine, can modulate sensory processing over a relatively large distance and over potentially slower timescales.

HETEROGENEITY IN SEROTONERGIC NETWORKS

Animals constantly integrate information from different sensory modalities under different internal physiological states and ecological contexts. The nervous system therefore must be able to independently modulate computations performed by sensory networks, so that information pertinent to the animal’s current state evokes the appropriate behavior. The serotonergic system has the capacity to influence sensory processing within a wide swath of behavioral contexts in a complex and even stimulus-specific manner. Individual DRN neurons are incredibly diverse in terms of efferent projections, nonserotonin transmitter content, intrinsic biophysical properties, sources of input, and transcriptional profiles (reviewed in Okaty et al. 2019). Together, these heterogeneous features of serotonergic neurons enable serotonin to have a nonuniform and complex influence on sensory processing.

Anatomical Heterogeneity of Serotonergic Neurons

Serotonergic neurons have heterogeneous anatomical projections, which can reflect multiple functional domains within specific sensory systems. For instance, serotonergic neurons do not project uniformly throughout the insect visual system but rather differentially innervate distinct layers within individual visual neuropils (Hamanaka et al. 2012; Homberg and Hildebrand 1989b; Nässel et al. 1987; Paulk et al. 2008; Schürmann and Klemm 1984; Vallés and White 1988). Additionally, in the moth Manduca antennal lobe, the CSDns do not innervate the regions occupied by olfactory afferents (Lizbinski et al. 2016; Sun et al. 1993). This suggests that Manduca CSDns do not directly act on olfactory afferents. In vertebrates, the olfactory bulb is innervated by serotonergic processes from both the MRN and DRN, but these processes are most dense in distinct synaptic layers (Gracia-Llanes et al. 2010; McLean and Shipley 1987; Muzerelle et al. 2016; Suzuki et al. 2015). Here, processes from the MRN are most dense in the region occupied mostly by periglomerular cells (a subclass of local interneuron), whereas processes from the DRN are densest in regions occupied by mitral/tufted cells (output neurons) and granule cells (another subclass of interneuron) (Muzerelle et al. 2016). Similarly, the density of serotonergic innervation varies both within and between rodent auditory nuclei, as is the case for nuclei in the superior olivary complex and the inferior colliculus (Hurley and Thompson 2001; Keesom et al. 2018; Thompson et al. 1994a). Within the Drosophila antennal lobe, the CSDns innervate glomeruli to varying degrees (Coates et al. 2017; Singh et al. 2013) and differentially connect with the various principal antennal lobe neuron types from animal to animal (Coates et al. 2017). Together, these examples illustrate the heterogeneous nature of serotonergic innervation of sensory networks. This heterogeneity allows distinct targeting of processing layers or stimulus-specific subcircuits.

Cotransmission and Serotonergic Neurons

In addition to different projection fields, recent technological innovations have revealed the molecular and anatomical diversity of subsets of serotonergic neurons within the vertebrate nuclei (Calizo et al. 2011; Fernandez et al. 2016; Huang et al. 2019; Ren et al. 2019; Spaethling et al. 2014; Templin et al. 2012). For instance, subpopulations of serotonergic DRN neurons with distinct trajectories coexpress either glutamate or GABA (Huang et al. 2019; Liu et al. 2014; Ren et al. 2018; Sengupta et al. 2017). Similarly, serotonergic DRN neurons can express several neuropeptides and nitric oxide synthase (Fu et al. 2010; Huang et al. 2019; Sengupta et al. 2017). In Drosophila, the CSDns polysynaptically evoke excitation via acetylcholine (Zhang and Gaudry 2016), although they may not synthesize acetylcholine, as an intersectional approach suggests that the CSDns may not express choline acetyltransferase (ChAT; Fig. 1A). Regardless, serotonergic neurons often use these cotransmitters to impact sensory systems in different ways. For example, when DRN projections to the olfactory bulb are activated, DRN-derived serotonin and glutamate differentially act on both output neuron subtypes (mitral and tufted cells). In this instance, glutamate directly enhances the odor-evoked responses in both output neuron subtypes, whereas serotonin enhances decorrelation of only mitral cell odor responses (Kapoor et al. 2016). Moreover, this enhancement is increased by pharmacologically blocking serotonin receptors and nearly abolished when glutamate receptors are similarly blocked (Kapoor et al. 2016). Although serotonin may be acting through polysynaptic interactions in this particular case (Brill et al. 2016; Hardy et al. 2005; Liu et al. 2012), the overall consequence is that cotransmission allows raphe neurons to affect their targets on different timescales through ionotropic and metabotropic receptors, respectively. This allows raphe neurons to both quickly alter a given downstream target’s neuronal activity and also leave that target’s activity altered for extended epochs.

Fig. 1.

Fig. 1.

Serotonergic modulation of sensory processing. A: intersectional immunohistochemistry using a rabbit polyclonal antibody against serotonin (5-HT, yellow; 1:5,000 dilution; ImmunoStar no. 20080) reveals that serotonergic Drosophila CSDns do not colabel with green fluorescent protein (GFP) expression with a protein-trap transgenic LexA driver for choline acetyltransferase (ChAT, green; BDSC_60319). Neuropil (magenta) delineated with a rat monoclonal antibody against N-cadherin (1:50 dilution; DSHB no. DN-Ex#8). No antibody was used to increase GFP signal. B: intersectional immunohistochemistry reveals that Drosophila CSDns express the 5-HT1B receptor subtype. Here, a protein-trap transgenic GAL4 driver for the 5-HT1B receptor subtype driving the expression of GFP (cyan; the 5-HT1B transgenics were a kind gift from Dr. Herman Dierick, Baylor College of Medicine) colabels with a goat polyclonal antibody against serotonin (yellow; 1:5,000 dilution; ImmunoStar no. 20079). Neuropil (magenta) delineated with a mouse monoclonal antibody against Bruchpilot (1:50 dilution; DSHB no. nc82). A rabbit polyclonal GFP antibody was used to increase GFP signal (1:1,000 dilution; ThermoFisher no. A-11122). C: schematic illustrating general network targets of 5-HT (yellow) highlighted in this review including sensory afferents (purple), local interneurons (orange), in particular presynaptic inhibition, and output neurons (blue). D: 5-HT can alter stimulus intensity coding by shifting the slope of the input-output relationship, modulating response strength, or offsetting the threshold for activation. E: 5-HT can also alter the encoding of stimulus identity by altering tuning breadth or by decreasing spontaneous activity to increase the signal-to-noise ratio. For both A and B, the same immunohistochemistry techniques described in Sizemore and Dacks (2016) were used to collect data. Scale bars, 10 μm.

Molecular Heterogeneity of Serotonergic Neurons

In addition to the anatomical and transmitter diversity of serotonergic neurons highlighted above, neurons within the DRN are highly molecularly diverse. For instance, different neurons within the DRN have different electrical properties due to differential ion channel expression levels (Calizo et al. 2011; Templin et al. 2012). This suggests that two given DRN neurons receiving identical synaptic input can still differentially modulate the same sensory network. To the best of our knowledge, ion channel expression profiles of serotonergic neurons have not been compared in invertebrates. However, the intrinsic properties and region-specific synaptic inputs for the CSDns enable a single serotonergic neuron to perform multiple operations across different regions of a given sensory domain (Zhang et al. 2019). Altogether, these heterogeneous features of individual and groups of serotonergic neurons support serotonin’s capacity for nuanced modulation of sensory processing.

THE CONTEXT FOR SEROTONIN RELEASE

Understanding how serotonin affects sensory processing has been, and remains, a fruitful and career-long endeavor of many investigators. Perhaps just as critical is to understand the circumstances in which serotonin exerts these effects. Serotonin release depends on both sensory and nonsensory input, and thus corresponds to many aspects of external events, self-generated behavioral patterns, and internal state.

Serotonergic Interactions with Sensory Systems Are Bidirectional

In addition to modulating sensory circuitry, serotonergic neurons often receive input from sensory systems, allowing their influence to be tempered based on the sensory input that the animal is experiencing. This may allow sensory networks to adapt to varying stimulus regimes. For example, serotonin-evoked changes in the signal-to-noise ratio might allow for stable stimulus representation in the face of heterogeneous environmental background (Hurley and Pollak 1999; Waterhouse et al. 1986, 1990). Sensory input to serotonergic neurons could also provide an opportunity for stimulus-specific modification of serotonin release. In some cases, primary sensory afferents and serotonergic cells are one and the same, thus allowing the sensory field to directly drive serotonin release. This is the case for the chemo/mechanosensory ADF neuron in C. elegans (Iwanir et al. 2016; Liu et al. 2019; Shao et al. 2019), select mechanosensory neurons in the antennae of cockroaches (Watanabe et al. 2014), and potentially auditory type II spiral ganglion afferents, as they express the serotonin transporter (Nielsen et al. 2006; Vyas et al. 2019). In other instances, serotonergic neurons are completely constrained to a sensory network, and therefore their activity is likely predominantly driven by sensory processing. For instance, several insect species have serotonergic neurons that make local projections within the optic lobes (Hamanaka et al. 2012; Homberg and Hildebrand 1989b; Leitinger et al. 1999), and the vertebrate retina possesses serotonergic amacrine cells (reviewed in Masson 2019) that contact retinal ganglion cells and bipolar cells.

Sensory Stimulus-Driven Serotonin Release

In addition to projecting to sensory and nonsensory regions (Gaudry 2018; Huang et al. 2019; Ren et al. 2019), many serotonergic neurons also receive input from sensory systems. In some cases, serotonergic neurons can receive sensory input locally within the networks that they are directly modulating. In Drosophila and moths, CSDn activity is influenced by odors (Hill et al. 2002; Zhang et al. 2019; Zhang and Gaudry 2016; Zhao and Berg 2009), via direct synaptic input from antennal lobe principal neurons (Berck et al. 2016; Coates et al. 2017; Sun et al. 1993; Zhang and Gaudry 2016). The organization of local input can even vary across sensory networks for a single neuron, as the CSDns can be excited and inhibited by a single odor due to local synaptic inputs to different neuronal compartments (Zhang et al. 2019). The raphe nuclei also have bidirectional connectivity, and in some cases sensory inputs arrive from cells that are relatively proximal to sensory transduction. Some retinal ganglion cells send collaterals to both the visual system and the DRN, although whether these retinal ganglion cells synapse directly onto serotonergic neurons remains unclear (Pickard et al. 2015). Inputs to serotonergic neurons in the DRN and MRN also originate from cortical and subcortical sensory regions including the inferior and superior colliculi or brain stem sensory nuclei (Ogawa et al. 2014; Pollak Dorocic et al. 2014). Furthermore, there are neurons in DRN and other raphe groups that respond to sensory stimuli across several modalities (Fornal et al. 1996; Gao and Mason 2000; Moriya et al. 2019; Rasmussen et al. 1984, 1986). Although the strongest sensory responses may occur in nonserotonergic raphe neurons, sensory responsiveness has been confirmed for serotonergic neurons themselves (Ranade and Mainen 2009; Ren et al. 2018; Waterhouse et al. 2004) and nonserotonergic DRN neurons may provide indirect sensory input to serotonergic neurons. Moreover, some serotonergic DRN neuron sensory responses are extremely short in latency, suggesting input from relatively early stages of sensory processing (Ranade and Mainen 2009). Thus, whether they are intrinsic to a network or span several networks, by having intimate access to the history of network activity serotonergic neurons can update their modulatory influence based on the stimulus regime or circuit state (Lizbinski and Dacks 2018).

Modality-Specific Serotonin Release

Variations in serotonergic neuron responses, as well as in their efferent connectivity, make it difficult to predict the conditions that elevate serotonin. Direct neurochemical analyses within sensory regions during behaviorally important events suggest that serotonin release within sensory regions is modality specific. For instance, serotonin levels in the electrosensory lateral line lobe of weakly electric fish increase in response to a synthetic signal mimicking the presence of a conspecific, yet strong auditory stimuli have no effect (Fotowat et al. 2016). Likewise, in rat temporal and occipital cortices that correspond to auditory and visually responsive regions (respectively), serotonergic changes are modality specific. Here, auditory input only evokes changes in the temporal cortex, whereas visual input only evokes changes in the occipital cortex (Müller et al. 2007; Pum et al. 2008). Modality-specific influence on serotonin levels can also occur across several processing stages. For instance, auditory stimuli influence the levels of serotonin or its metabolites at multiple sites in the auditory system, including the brain stem, midbrain, and cortex (Cransac et al. 1998; Hall et al. 2010). In contrast, exposing a rodent to a stressful odor, a component of fox urine, does not influence serotonin in auditory regions (Hall et al. 2010). Together, these examples highlight the modality-specific nature of serotonin release.

Despite the canalization of serotonergic projections to sensory regions suggested by these studies, some DRN neurons that send projections into sensory regions also have projections into nonsensory regions, suggesting coregulation of functionally related brain regions. For example, single DRN neurons in the rat project to functionally related regions of sensory cortex and cerebellum (Waterhouse et al. 1986) or to sensory and motor areas related to whisker sensation and movement in the cortex, thalamus, or medulla (Lee et al. 2008). Coregulation of functionally related brain regions is also suggested by the projections of defined subgroups of dorsal raphe neurons. For example, block of serotonin release from select DRN neurons increases the expression of aggressive behaviors in mice and also alters nonsocial behaviors (Niederkofler et al. 2016). These neurons not only project to some nodes of social behavior networks but also provide prominent inputs to sites in the auditory brain stem and midbrain. Although serotonin release in response to multimodal stimuli has not been systematically examined across sensory regions, these studies suggest that DRN projections may be organized according to sensory modality.

Self-Regulation of Serotonin Release

In addition to the stimulus- and modality-dependent and -independent means of adjusting serotonin release discussed above, serotonin release can also be modified locally via autoreceptors or heteroreceptors (reviewed in Andrade et al. 2015; Belmer and Maroteaux 2019). In several rodent species, for instance, DRN and MRN neurons express inhibitory serotonin receptors (5-HTRs) (Adell et al. 2001; Moret and Briley 1997; Piñeyro et al. 1996; Starkey and Skingle 1994). Moreover, the activity of these serotonin autoreceptors and heteroreceptors is implicated in aggression (Nautiyal et al. 2015) and mood disorders (Donaldson et al. 2014; Nautiyal et al. 2016; Riad et al. 2017; You et al. 2016). At the cellular level, the activity of these 5-HTRs along a given serotonergic neuron’s terminals can provide the means for cell-autonomous regulation of local serotonin release, without necessarily recruiting additional neurons. A similar theme may occur in the sea lamprey retina, where serotonergic amacrine cells in the retina express the inhibitory 5-HT1A receptor (Cornide-Petronio et al. 2015). In Drosophila, the CSDns express the inhibitory 5-HT1B receptor (Fig. 1B) and have several compartment-specific features, including synaptic connectivity (Zhang et al. 2019) and serotonin transporter expression (Kasture et al. 2019). Together these features of Drosophila CSDns provide several potential means for differential influence across networks. The inhibitory influence of serotonin on serotonergic neurons also appears to be mediated in an activity-independent and nonexocytotic manner (Mlinar et al. 2015), suggesting that the serotonergic network may self-regulate in a manner distinct from the context in which it affects other neurons.

Serotonergic Neurons Respond to Behavioral State and Context

In addition to sensory stimulus-driven responses, serotonergic neurons respond to complex stimuli, behavioral output, reward contingencies, and internal state. Much like the heterogeneous characteristics noted earlier (see above), raphe neurons as a population are markedly diverse in the stimuli to which they respond. In mammals, the activity of DRN neurons varies greatly in their responses to both behavioral outputs (Fornal et al. 1996; Heym et al. 1982; Jacobs et al. 2002; Jacobs and Fornal 1991) and more subtle stimulus qualities like reward contingencies based on recent experience. Subpopulations of DRN serotonergic neurons respond to reward, punishment, or learned predictors (Cohen et al. 2015; Li et al. 2016; Liu et al. 2014; Luo et al. 2016; Matias et al. 2017; Miyazaki et al. 2011a, 2011b, 2012; Stark and Scheich 1997). Furthermore, different stimulus features such as salience or valence can be encoded as bursts of spikes or changes in DRN tonic firing rate (Cohen et al. 2015). This heterogeneity likely reflects subpopulations of DRN neurons that support different behavioral functions. Frontal cortex- and olfactory bulb-projecting DRN neurons and amygdala-projecting DRN neurons differ in the brain regions from which they receive input. Moreover, the former are activated by reward and inhibited by punishment, whereas the latter are excited by both (Ren et al. 2018). Furthermore, the activation of raphe neurons is dependent upon environmental context. Serotonergic DRN neuron activity decreases with onset of movement in assays with low perceived threat but increases under high-threat conditions (Seo et al. 2019). Similarly, confinement within a small area, a mild stressor, increases serotonergic activity in the mouse auditory midbrain (Hall et al. 2010).

Social context can also influence serotonin release. Interactions with conspecifics increases serotonin release in the mouse auditory midbrain, in correlation with nonvocal behaviors such as social investigation or overall activity of the subjects, rather than the number of vocalizations produced (Hall et al. 2011; Hanson and Hurley 2014; Keesom and Hurley 2016). In male mice interacting with female partners, serotonergic activity correlates with female vocalizations that may indicate rejection, but this is a negative correlation (Keesom and Hurley 2016). Additionally, serotonin-auditory interactions are sensitive to prior experience. For example, the dynamics of serotonin release during some types of social interaction are also slower for individuals housed in isolation in early life (Keesom et al. 2017). These results suggest that serotonergic signaling is tied to the salience and valence of individual social interactions rather than being positively correlated with the sensory stimuli associated with that interaction. Finally, serotonergic neurons across taxa are also regulated by physiological contexts that vary over relatively long timescales including hunger (Voigt and Fink 2015) and circadian rhythm (Cagampang and Inouye 1994; Corthell et al. 2013; Jacobs et al. 1981; Kloppenburg et al. 1999; Monti 2011; Trulson and Jacobs 1979). Thus, serotonergic neurons are influenced by both the animal’s environment and broad behavioral states, both of which alter distinct aspects of temporal firing patterns.

In summary, the contexts for serotonin release are multidimensional and include both stimulus-dependent and -independent conditions. Anatomical and functional studies suggest that serotonergic neurons are sensitive to sensory events but are also influenced by factors such as internal state, motor activity, and the salience of stimuli with regard to prior events. From all this, it is reasonable to conclude that serotonin is capable of conveying information into sensory systems on the external context and internal state in which sensory events occur.

RECEPTOR BASIS OF COMMON MODULATORY EFFECTS

While serotonin clearly affects fundamental aspects of sensory processing, there are several paths by which these overall circuit outcomes can be achieved. Just as serotonergic neurons are themselves diverse, there is a diversity of serotonin receptors (5-HTRs) that vary in their affinity for serotonin, time course of action, and the secondary messenger system to which they couple (Nichols and Nichols 2008). This receptor diversity allows serotonin to differentially target neuronal populations that support distinct sensory computations across modalities (Fig. 1C).

Comparing Serotonin Receptors across Taxa

An array of 5-HTRs are encoded in nearly every animal genome (Azmitia 2007; Ishita et al. 2020; Moroz et al. 2014; Peroutka and Howell 1994; Ribeiro et al. 2005; Vleugels et al. 2015). The first 5-HTR emerged ~700–800 million years ago (Peroutka and Howell 1994), and there are seven major 5-HTR families in vertebrates (5-HT1–7) and at least three across the invertebrates (5-HT1, 2, and 7) (Table 2). However, there are notable clade-specific exceptions such as the MOD-1 ionotropic 5-HTR in C. elegans (Ranganathan et al. 2000), the 5-HT8 receptor in Pieris rapae (Qi et al. 2014), the 5-HT4 and 6 receptors in some mollusks (Kim et al. 2019; Nagakura et al. 2010; Tamvacakis et al. 2015, 2018), the nonfunctional 5-HT5B receptor subtype in humans (Grailhe et al. 2001), the absence of these receptors in the Ctenophora genome (Moroz et al. 2014), and the 5-HT4 receptor in Apostichopus japonicus (Wang et al. 2017). Invertebrate 5-HTRs are typically named for the vertebrate 5-HTR family with which they share the most sequence homology, but the pharmacological properties of these counterparts can differ. Methysergide, for example, acts as a broad-spectrum 5-HTR antagonist in vertebrates but agonizes or has no effect on select invertebrate 5-HTRs (Blenau et al. 2017; Dacks et al. 2013; Röser et al. 2012). There are 14 subtypes of vertebrate 5-HTRs (e.g., within the 5-HT2 family there are 5-HT2A–C), some of which can have several isoforms as a result of posttranscriptional modifications to the nascent 5-HTR transcript (Bockaert et al. 2006; Burns et al. 1997; Hannon and Hoyer 2008; Tanaka and Watanabe 2020; Vleugels et al. 2015). Conversely, invertebrate 5-HTR subtypes are generally encoded at distinct genomic loci, and each 5-HTR has a single predicted isoform (e.g., Drosophila 5-HTRs: Colas et al. 1995; Gasque et al. 2013; Saudou et al. 1992; Witz et al. 1990). Of these different 5-HTR subtypes, several (if not all) are expressed in olfactory, auditory, mechanosensory, visual, and gustatory centers of both vertebrates and invertebrates (Table 3). Moreover, reoccurring themes have begun to emerge wherein serotonin acts through these diverse receptors within these sensory modalities to modulate fundamental computations including breadth of responsive range, detection thresholds, and discriminating between stimuli.

Table 2.

Serotonergic receptors that have been molecular cloned and their intracellular effects assayed across members of various taxa

Order Species Receptor Subtype Response Reference
Anaspideae Aplysia californica 5-HTap1 Decrease cAMP (Angers et al. 1998)
5-HTap2 Decrease cAMP (Barbas et al. 2002)
Aplysia kurodai 5-HTapAC1 Increase cAMP (Lee et al. 2009)
Basommatophora Lymnaea stagnalis 5-HT2 Increase IP3 (Gerhardt et al. 1996)
Blattodea Periplaneta americana 5-HT1 Decrease cAMP (Troppmann et al. 2010)
Coleoptera Tribolium castaneum 5-HT1 Decrease cAMP (Vleugels et al. 2013)
5-HT7 Increase cAMP (Vleugels et al. 2014)
Decapoda Panulirus interruptus 5-HT1A Decrease cAMP (Spitzer et al. 2008a)
5-HT2B Increase IP3 (Clark et al. 2004)
Procambarus clarkii 5-HT1A Decrease cAMP (Spitzer et al. 2008b)
5-HT2B Increase IP3
Diptera Aedes aegypti 5-HT7 Increase cAMP (Lee and Pietrantonio 2003; Pietrantonio et al. 2001)
Anopheles gambiae 5-HT2 (“AGAP002229”) Increase Ca2+ (Ngai et al. 2019)
5-HT2 (“AGAP002232”) Increase Ca2+
Calliphora vicina 5-HT7 Increase cAMP (Röser et al. 2012)
5-HT2A Increase Ca2+
Drosophila melanogaster 5-HT1A and 1B Decrease cAMP (Saudou et al. 1992)
5-HT2A Increase Ca2+ (Colas et al. 1995; Gasque et al. 2013)
5-HT2B Increase Ca2+ (Gasque et al. 2013)
5-HT7 Increase cAMP (Witz et al. 1990)
Hymenoptera Apis mellifera 5-HT2A and 2B Increase Ca2+ (Thamm et al. 2013)
5-HT7 Increase cAMP (Schlenstedt et al. 2006)
Ixodida Boophilus microplus 5-HT1 Decrease cAMP (Chen et al. 2004)
Lepetellida Haliotis discus hannai 5-HT1B Decrease cAMP (Kim et al. 2019)
5-HT4 Increase cAMP
Lepidoptera Bombyx mori 5-HT1A Decrease cAMP (Xiong et al. 2019)
Manduca sexta 5-HT2 Increase IP3 (Dacks et al. 2013)
5-HT7 Increase cAMP
Pieris rapae 5-HT1A and 1B Decrease cAMP (Qi et al. 2017)
5-HT7 Increase cAMP
5-HT8 Increase Ca2+ (Qi et al. 2014)
Primates Homo sapiens 5-HT1A, 1B, 1D, 1E, 1F Decrease cAMP (Adham et al. 1993; Hamblin et al. 1992; McAllister et al. 1992; Stam et al. 1992; Stam et al. 1994; Weinshank et al. 1992)
5-HT2A, 2B, 2C Increase IP3 (Kursar et al. 1994; Schmuck et al. 1994; Stam et al. 1992; Stam et al. 1994)
5-HT4 Increase cAMP (Blondel et al. 1998)
5-HT5A Inconclusive (Grailhe et al. 2001)
5-HT6 Increase cAMP (Kohen et al. 1996)
5-HT7 Increase cAMP (Stam et al. 1997)
Rhabditidae Caenorhabditis elegans SER-1 Increase Ca2+ (Hamdan et al. 1999)
SER-4 Decrease cAMP (Olde and McCombie 1997)
SER-7 Increase cAMP (Hobson et al. 2006)
Rodentia Mus musculus 5-HT1A, 1B, 1D, 1F Decrease cAMP (Amlaiky et al. 1992; Charest et al. 1993; Maroteaux et al. 1992; Weydert et al. 1992; Yu et al. 1991)
5-HT2A, 2B, 2C Increase IP3 (Foguet et al. 1992; Loric et al. 1992)
5-HT4 Increase cAMP (Claeysen et al. 1999)
5-HT5A and 5B Inconclusive (Matthes et al. 1993)
5-HT6 Increase cAMP (Kohen et al. 2001)
5-HT7 Increase cAMP (Plassat et al. 1993)
Rattus norvegicus 5-HT1A, 1B, 1D, 1F Decrease cAMP (Adham et al. 1993; Albert et al. 1990; Hamblin et al. 1992; Lovenberg et al. 1993)
5-HT2A, 2B, 2C Increase Ca2+/IP3 (Julius et al. 1988; Kursar et al. 1994; Pritchett et al. 1988)
5-HT4 Increase cAMP (Gerald et al. 1995)
5-HT5A Decrease cAMP (Erlander et al. 1993; Thomas et al. 2000)
5-HT5B Inconclusive (Erlander et al. 1993; Wisden et al. 1993)
5-HT6 Increase cAMP (Ruat et al. 1993a)
5-HT7 Increase cAMP (Ruat et al. 1993b; Shen et al. 1993)
Synallactida Apostichopus japonicus 5-HT4 Increase cAMP (Wang et al. 2017)

“Response” indicates the in vitro consequences of receptor activation. IP3, inositol trisphosphate.

Table 3.

Serotonin receptor subtype expression within auditory, mechanosensory, gustatory, olfactory, and visual processing centers across vertebrates and invertebrates

Sensory System Order Receptor Subtype Method Reference
Audition/mechanosensation Anaspidea 5-HT7 In situ hybridization (Lee et al. 2009)
Araneae 5-HT1 In situ hybridization (Sukumar et al. 2018)
Diptera All subtypes Transgenics (Howard et al. 2019)
Decapoda 5-HT1 Immunohistochemistry (Fickbohm et al. 2005)
Rodentia 5-HT1A, 1C, 2 In situ hybridization (Wright et al. 1995)
5-HT2A, 2C Immunohistochemistry (Li et al. 2003)
5-HT4A Immunohistochemistry (Suwa et al. 2014)
5-HT1B, 1D, 1F In situ hybridization (Bruinvels et al. 1994)
5-HT2A Immunohistochemistry (Cornea-Hébert et al. 1999)
5-HT2A Immunohistochemistry (Basura et al. 2008)
5-HT2B Immunohistochemistry (Tadros et al. 2007)
5-HT2 Autoradiography (Malgouris et al. 1993)
5-HT1A In situ hybridization and autoradiography (Chalmers and Watson 1991)
5-HT1A In situ hybridization (Pompeiano et al. 1992)
5-HT7 Autoradiography (To et al. 1995)
5-HT1A. 1B Immunohistochemistry (Peruzzi and Dut 2004)
Gustation Blattodea 5-HT1 Immunohistochemistry (Troppmann et al. 2010)
Decapoda 5-HT1, 2 Immunohistochemistry (Vázquez-Acevedo et al. 2009)
5-HT1 Immunohistochemistry (Spitzer et al. 2005)
Diptera 5-HT1A Transgenics (Luo et al. 2012)
5-HT1A, 1B, 2A, 7 Transgenics (Huser et al. 2017)
5-HT1B Transgenics (Liu et al. 2015)
Hymenoptera 5-HT7 In situ hybridization (Schlenstedt et al. 2006)
Orthoptera 5-HT1A, 1B Immunohistochemistry (Shao et al. 2010)
Rhabditida 5-HT7 Transgenics (Hobson et al. 2006)
Rodentia 5-HT2A Immunohistochemistry (Cornea-Hébert et al. 1999)
5-HT4A Immunohistochemistry (Suwa et al. 2014)
5-HT2 Autoradiography (Malgouris et al. 1993)
Olfaction Decapoda 5-HT1 Immunohistochemistry (Spitzer et al. 2005)
5-HT2 Immunohistochemistry (Vázquez-Acevedo et al. 2009)
Diptera All subtypes Transgenics (Sizemore and Dacks 2016)
5-HT1A, 1B, 2A, 7 Transgenics (Huser et al. 2017)
Hymenoptera 5-HT7 In situ hybridization (Schlenstedt et al. 2006)
5-HT1A Immunohistochemistry (Thamm et al. 2010)
Lepidoptera 5-HT1 Immunohistochemistry (Dacks et al. 2013)
Orthoptera 5-HT1A, 1B Immunohistochemistry (Shao et al. 2010)
Rhabditida SER-5 (5-HT6-like) Transgenics (Harris et al. 2010)
Rodentia 5-HT2 Autoradiography (Pazos and Palacios 1985)
5-HT3 In situ hybridization (Tecott et al. 1993)
5-HT1A, 1C, 2 In situ hybridization (Wright et al. 1995)
5-HT1E Immunohistochemistry (Klein and Teitler 2012)
5-HT4A Immunohistochemistry (Suwa et al. 2014)
5-HT1C In situ hybridization (Mengod et al. 1990)
5-HT2A Immunohistochemistry (Cornea-Hébert et al. 1999)
5-HT1B In situ hybridization (Voigt et al. 1991)
5-HT2A Immunohistochemistry (Hamada et al. 1998)
5-HT1B, 1D, 1F In situ hybridization (Bruinvels et al. 1994)
5-HT2 Autoradiography (Malgouris et al. 1993)
5-HT2A. 2C Immunohistochemistry (Li et al. 2003)
Vision Carnivora 5-HT1 Autoradiography (Mower 1991)
5-HT1 Autoradiography (Skangiel-Kramska and Kossut 1992)
5-HT1A, 1C, 2, 3 Autoradiography (Dyck and Cynader 1993)
Cypriniformes 5-HT1A, 1B In situ hybridization (Norton et al. 2008)
Decapoda 5-HT1, 2 Immunohistochemistry (Vázquez-Acevedo et al. 2009)
5-HT1 Immunohistochemistry (Spitzer et al. 2005)
Diptera All subtypes Transgenics (Sampson et al. 2019) (preprint)
Hymenoptera 5-HT1A Immunohistochemistry (Thamm et al. 2010)
5-HT7 In situ hybridization (Schlenstedt et al. 2006)
Lagomorpha 5-HT1A, 7 In situ hybridization (Chidlow et al. 1998)
Orthoptera 5-HT1A, 1B Immunohistochemistry (Shao et al. 2010)
Petromyzontiformes 5-HT1A In situ hybridization (Cornide-Petronio et al. 2015)
Primates 5-HT1, 2 Autoradiography (Rakic et al. 1988)
5-HT1, 2 Autoradiography (Rakic and Lidow 1995)
5-HT1A, 2 Autoradiography (Impieri et al. 2019)
5-HT1B, 2A In situ hybridization (Watakabe et al. 2009)
Rodentia 5-HT1A, 1C, 2 In situ hybridization (Wright et al. 1995)
5-HT1A Immunohistochemistry (Zhou et al. 2019)
5-HT2A, 2C Immunohistochemistry (Li et al. 2003)
5-HT4A Immunohistochemistry (Suwa et al. 2014)
5-HT2A Immunohistochemistry (Cornea-Hébert et al. 1999)
5-HT1B, 1D, 1F In situ hybridization (Bruinvels et al. 1994)
5-HT1F In situ hybridization (Adham et al. 1993)

Serotonergic Modulation of Sensory Computations: Gain Control

Sensory systems employ several strategies to dynamically adjust the range of individual stimulus features that they encode. Given the ubiquity of serotonergic systems, it is not surprising that serotonin either modulates or plays a direct role in these processes. For instance, animals experience fluctuations in stimulus intensity as they navigate the world. In cases where the animal experiences intense sensory input (i.e., bright lighting, high odor concentration, etc.), the neurons may fail to properly encode the given stimulus as a result of saturation. Conversely, animals may fail to detect ecologically important stimuli (i.e., the scent of a predator) if they are present at low intensities. To overcome these hurdles, sensory systems typically use a suite of computations, such as “gain control” (reviewed in Carandini and Heeger 2012), to adaptively adjust the sensory input-to-output ratio of a network (Fig. 1D). Across modalities, 5-HTRs expressed by sensory afferents can enable direct serotonergic modulation of the gain of sensory input. For instance, chronic activation of nociceptive afferents in Drosophila larvae induces direct serotonergic inhibitory feedback mediated by 5-HT1B receptor that causes desensitization of afferents over the course of development (Kaneko et al. 2017). Additionally, although the receptor basis remains unknown, serotonin decreases sensory afferent activity in proprioceptor and mechanosensory networks (Gaudry and Kristan 2009; Nagata et al. 2019). Conversely, serotonin can activate excitatory 5-HTRs in photoreceptors to directly increase their excitability (Cheng and Frye 2020; Han et al. 2007; Pootanakit et al. 1999), therefore increasing the sensitivity of these sensory afferents.

Serotonergic Modulation of Sensory Computations: Sharpening and Broadening

In addition to directly acting on sensory afferents to modulate gain control, serotonin can indirectly modulate sensory afferents via local interneurons (LNs). For example, serotonin stimulates 5-HT2C-expressing LNs (juxtaglomerular cells) in the olfactory bulb to increase the amount of presynaptic inhibition exerted upon olfactory afferents (Petzold et al. 2009). In doing so, serotonin reduces the gain of olfactory afferent responses and thus the amount of sensory input entering the olfactory bulb (Petzold et al. 2009). Moreover, serotonin can also indirectly enhance presynaptic inhibition by activating 5-HT2A receptors expressed by excitatory LNs (external tufted cells), which in turn provide excitatory drive to inhibitory short axon and periglomerular cells (Brill et al. 2016; Liu et al. 2012). In this way, serotonin can further drive inhibitory inputs to sensory afferents as a means of decreasing sensory input. Serotonin similarly indirectly decreases the strength of sensory afferent input in the Drosophila antennal lobe by enhancing presynaptic inhibition (Dacks et al. 2009; Gaudry 2018) and the strength of projection neuron (PN) responses by enhancing postsynaptic inhibition (Suzuki et al. 2020). Altogether, these examples highlight the different mechanisms by which serotonin regulates the resolution with which sensory networks encode stimulus intensity.

In addition to compensating for large variations in stimulus intensity, animals may also need to adjust their ability to resolve different stimuli. At the neuronal level, sensory systems use inhibition and lateral excitation to sharpen or broaden the resolution with which they encode stimulus identity (reviewed in Martin et al. 2011). Not surprisingly, serotonin also targets these aspects of sensory encoding (Fig. 1E). Such is the case in the inferior colliculus, where serotonin sharpens auditory neurons’ responses to primary sound frequencies by decreasing their responsiveness to frequencies both within and outside of the central range (Hurley and Pollak 2001). In the piriform cortex, serotonin sharpens neuronal representations of odors by decreasing their spontaneous activity but leaving their odor-evoked responses unaffected (Lottem et al. 2016).

In these instances, serotonin hones the neurons’ responses, whether by narrowing the receptive range or by decreasing spontaneous activity. However, serotonin can also broaden the receptive range of neurons in a given sensory system. For instance, in the vertebrate retina serotonin can decrease lateral inhibition by activating inhibitory 5-HTRs expressed by amacrine cells, therefore broadening the number of retinal ganglion cells that are responsive to a given stimulus (Trakhtenberg et al. 2017; Zhou et al. 2019). In the olfactory bulb, serotonin enhances feedforward excitation to mitral cells from interneurons via the 5-HT2A receptor (Brill et al. 2016; Huang et al. 2017; Liu et al. 2012). The combined actions of both serotonin and glutamate released by DRN neurons increase the sensitivity of tufted cells and decorrelate odor-evoked responses of mitral cells, presumably increasing the separation of representations of different odors (Kapoor et al. 2016). Within the Drosophila antennal lobe, serotonin could potentially affect the breadth of odor-evoked representations, as a specialized population of excitatory interneurons can broaden odor-tuning (Huang et al. 2010; Shang et al. 2007; Yaksi and Wilson 2010) and these neurons express excitatory 5-HTRs (Sizemore and Dacks 2016). Serotonin can also directly modulate the excitability of second-order neurons within a sensory system, thus potentially influencing tuning breadth. Exogenous application of serotonin increases the excitability of antennal lobe output neurons (Dacks et al. 2006b; Kloppenburg et al. 1999; Kloppenburg and Hildebrand 1995; Zhang and Gaudry 2016). This effect, however, is at least partially polysynaptic and depends on the method of delivery, as activation of Drosophila CSDns can have little to no effect on the odor-evoked responses of output neurons depending on the glomerulus (Zhang and Gaudry 2016). This discrepancy could arise from cell class-specific receptor expression in the antennal lobe (AL) (Sizemore and Dacks 2016), differences in binding affinities of the 5-HTRs for serotonin (Gasque et al. 2013), or differences in the time course of receptor activation and inactivation. Regardless, there appear to be a variety of means by which serotonin can affect the resolution with which stimulus identity is encoded.

Serotonergic Regulation between Sensory Information Streams

By its combined action on different neural types within a sensory circuit, serotonin can functionally regulate the balance among different streams of information within sensory regions. For example, in the superficial layers of the superior colliculus, serotonin acting via 5-HT1A receptors decreases the responses of single postsynaptic neurons to stimulation of both ascending and descending visual pathways (Mooney et al. 1996). In contrast, activation of 5-HT1B receptors largely decreases responses to stimulation of the ascending visual pathway, likely by decreasing transmitter release presynaptically. Endogenous serotonin release could therefore favor descending over ascending visual streams via these two inhibitory serotonin receptor types. In the dorsal cochlear nucleus (DCN), the high density of serotonergic fibers in cell groups that receive descending and multisensory input is proposed to regulate the convergence of multimodal information at the level of the principal cells (fusiform cells) in this nucleus (Klepper and Herbert 1991). As in the superior colliculus, serotonin acts on multiple types of excitatory and inhibitory DCN neurons through different classes of receptors (Tang and Trussell 2015, 2017). Serotonin postsynaptically increases the excitability of fusiform neurons through 5-HT2A and 5-HT7 receptors. At the same time, serotonin presynaptically decreases the amplitude of excitatory potentials from auditory nerve fibers through 5-HT1A receptors and increases the excitability of inhibitory interneurons through 5-HT2 receptors. This causes responses of the principal neurons of the DCN to stimulation of auditory-only pathways to be dampened, whereas responses to stimulation of multisensory pathways are enhanced. In multiple sensory cortices, serotonin also has neuron-specific, receptor-specific, or layer-specific effects (Foehring et al. 2002; García-Oscos et al. 2015; Jang et al. 2012; Lee et al. 2018; Torres-Escalante et al. 2004; Xiang and Prince 2003). These selective effects may alter the balance between different sources of information such as inter- and intracolumnar inputs or intralayer versus feedforward sources of input (Cervantes-Ramírez et al. 2019; Xiang and Prince 2003).

Serotonergic Metamodulation across Sensory Systems

In addition to acting on fast synaptic transmission within a given sensory system, serotonin can also act on other modulatory neurons that influence sensory processing. Indeed, all sensory systems are influenced by multiple neuromodulators released from intrinsic or extrinsic neurons (Berg et al. 2009; Carlsson et al. 2010; Chalasani et al. 2010; Hurley et al. 2004; Iwano and Kanzaki 2005; Jacob and Nienborg 2018; Lizbinski et al. 2018; Nässel 2018; Nässel and Zandawala 2019; Schofield and Hurley 2018). Neuromodulators released by a given modulatory neuron can also influence other modulatory neurons within the network. Collective changes in concentration over time of each neuromodulatory molecule reflect the overall “modulatory tone” of any given sensory network. These changes in modulatory tone might therefore reflect dramatic shifts in the animal’s behavior state as it relates to a given sensory experience. Individual modulators, such as serotonin, can therefore profoundly alter networkwide activity by simply adjusting existing modulatory circuitry present in a given sensory network.

Serotonin influences GABAergic modulation in both vertebrate and invertebrate primary olfactory systems (reviewed in Lizbinski and Dacks 2018), and peptidergic modulation by serotonin is likely present across taxa. For instance, interneurons of the vertebrate sensory cortex that release vasoactive intestinal peptide (VIP) also express the excitatory 5-HT3 ionotropic receptor (Cardin 2018; Lee et al. 2010; Rudy et al. 2011). Activating 5-HT3 receptors in VIP interneurons causes a hyperpolarization in 5-HT3-negative inhibitory interneurons, which subsequently disinhibits pyramidal neurons (Jiang et al. 2015; Pfeffer et al. 2013; Takesian et al. 2018). Moreover, serotonergic stimulation of VIP interneurons also produces a latent, GABAB receptor-mediated hyperpolarization in these same pyramidal cells (Takesian et al. 2018). Therefore, by acting through these interneurons serotonin can have a large impact on network dynamics and even modulate distinct aspects of sensory processing (for example see Pi et al. 2013). Moreover, the activity of the VIP interneurons appears to be at least one determinant for the changes observed in the activity of visual cortex circuitry according to the animal’s ongoing behavioral state (Batista-Brito et al. 2017; Bennett et al. 2013; Fu et al. 2014; Pakan et al. 2016; Polack et al. 2013). Collectively, these results suggest that there may be a serotonin-induced contingency switching module in visual cortex wherein the animal’s locomotor activity induces serotonergic activation of VIP interneurons. Then, perhaps after some epoch after behavior initiation, negative feedback terminates this serotonin-induced module.

CONCLUDING REMARKS

Here we have presented numerous examples of serotonin’s capacity for adjusting sensory processing at nearly every stage of signaling. Regardless of modality or species, serotonergic systems are heterogeneous at the level of individual neurons, as well as diverse at the level of whole populations. Moreover, the suite of serotonin receptors further expands the means with which serotonin affects select features, such as odor coding. These heterogeneous features of the serotonin system allow for widespread, nuanced effects of serotonin on sensory processing that vary in a context-dependent manner. Subsequently, these heterogeneous features also complicate assignments of a singular role for serotonin. However, serotonergic modulation is widespread throughout the animal kingdom, and currently the majority of our understanding regarding the cellular mechanisms underlying serotonergic modulation of sensory processing comes from a handful of organisms (i.e., rodents, fruit flies, etc.). By comparing across modalities and diverse taxa, we can reveal convergent adaptations that reveal fundamental molecular, cellular, and network mechanisms of sensory modulation. Similar approaches might also reveal divergent adaptations that reveal the selective pressures that sculpt neuromodulation.

Future directions for understanding the role of serotonin in sensory processing include the following:

  • • 

    How does the full extent of serotonergic neuron diversity vary from animal to animal, and what factors contribute to this variation?

  • • 

    How do the properties of serotonergic neurons and expression patterns of 5-HTRs change in response to different external and/or internal demands?

  • • 

    How much of the context-dependent effects of serotonin arises from the heterogeneous nature of serotonergic neurons, and how much arises from different 5-HTR expression motifs?

  • • 

    How does 5-HTR autoreceptor and heteroreceptor activity influence serotonergic modulation of sensory processing?

  • • 

    To what extent are 5-HTRs expressed in specific neuronal compartments, and what are the consequences of 5-HTR distribution patterns for sensory processing?

GRANTS

This work was supported by a Grant-In-Aid of Research (G20141015669888) from Sigma Xi, The Scientific Research Society to T.R.S.; National Science Foundation IOS Grants 1856436 and 1456298 to L.M.H.; and NIH Grant DC-016293 and US Air Force Office of Scientific Research Grant FA9550-17-1-0117 to A.M.D.

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the authors.

AUTHOR CONTRIBUTIONS

T.R.S. conceived and designed research; T.R.S. and A.D. prepared figures; T.R.S., L.M.H., and A.D. drafted manuscript; T.R.S., L.M.H., and A.D. edited and revised manuscript; T.R.S., L.M.H., and A.D. approved final version of manuscript.

ACKNOWLEDGMENTS

We are grateful for the feedback we received from Kristyn Lizbinski and Kevin Daly regarding this manuscript. We also acknowledge Dawn Blitz for promoting this collaborative review. We could not highlight every relevant study, and we apologize to everyone whose work was not discussed.

REFERENCES

  1. Adell A, Celada P, Artigas F. The role of 5-HT1B receptors in the regulation of serotonin cell firing and release in the rat brain. J Neurochem 79: 172–182, 2001. doi: 10.1046/j.1471-4159.2001.00550.x. [DOI] [PubMed] [Google Scholar]
  2. Adham N, Kao HT, Schecter LE, Bard J, Olsen M, Urquhart D, Durkin M, Hartig PR, Weinshank RL, Branchek TA. Cloning of another human serotonin receptor (5-HT1F): a fifth 5-HT1 receptor subtype coupled to the inhibition of adenylate cyclase. Proc Natl Acad Sci USA 90: 408–412, 1993. doi: 10.1073/pnas.90.2.408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Agnati LF, Bjelke B, Fuxe K. Volume versus wiring transmission in the brain: a new theoretical frame for neuropsychopharmacology. Med Res Rev 15: 33–45, 1995. doi: 10.1002/med.2610150104. [DOI] [PubMed] [Google Scholar]
  4. Albert PR, Zhou QY, Van Tol HH, Bunzow JR, Civelli O. Cloning, functional expression, and mRNA tissue distribution of the rat 5-hydroxytryptamine1A receptor gene. J Biol Chem 265: 5825–5832, 1990. [PubMed] [Google Scholar]
  5. Amlaiky N, Ramboz S, Boschert U, Plassat JL, Hen R. Isolation of a mouse “5HT1E-like” serotonin receptor expressed predominantly in hippocampus. J Biol Chem 267: 19761–19764, 1992. [PubMed] [Google Scholar]
  6. Andrade R, Huereca D, Lyons JG, Andrade EM, McGregor KM. 5-HT1A receptor-mediated autoinhibition and the control of serotonergic cell firing. ACS Chem Neurosci 6: 1110–1115, 2015. doi: 10.1021/acschemneuro.5b00034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Angers A, Storozhuk MV, Duchaîne T, Castellucci VF, DesGroseillers L. Cloning and functional expression of an Aplysia 5-HT receptor negatively coupled to adenylate cyclase. J Neurosci 18: 5586–5593, 1998. doi: 10.1523/JNEUROSCI.18-15-05586.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Antri M, Cyr A, Auclair F, Dubuc R. Ontogeny of 5-HT neurons in the brainstem of the lamprey, Petromyzon marinus. J Comp Neurol 495: 788–800, 2006. doi: 10.1002/cne.20910. [DOI] [PubMed] [Google Scholar]
  9. Azmitia EC. Serotonin and brain: evolution, neuroplasticity, and homeostasis. Int Rev Neurobiol 77: 31–56, 2007. doi: 10.1016/S0074-7742(06)77002-7. [DOI] [PubMed] [Google Scholar]
  10. Azmitia EC, Segal M. An autoradiographic analysis of the differential ascending projections of the dorsal and median raphe nuclei in the rat. J Comp Neurol 179: 641–667, 1978. doi: 10.1002/cne.901790311. [DOI] [PubMed] [Google Scholar]
  11. Baker KG, Halliday GM, Halasz P, Hornung JP, Geffen LB, Cotton RG, Törk I. Cytoarchitecture of serotonin-synthesizing neurons in the pontine tegmentum of the human brain. Synapse 7: 301–320, 1991. doi: 10.1002/syn.890070407. [DOI] [PubMed] [Google Scholar]
  12. Baker KG, Halliday GM, Törk I. Cytoarchitecture of the human dorsal raphe nucleus. J Comp Neurol 301: 147–161, 1990. doi: 10.1002/cne.903010202. [DOI] [PubMed] [Google Scholar]
  13. Barbas D, Zappulla JP, Angers S, Bouvier M, Castellucci VF, DesGroseillers L. Functional characterization of a novel serotonin receptor (5-HTap2) expressed in the CNS of Aplysia californica. J Neurochem 80: 335–345, 2002. doi: 10.1046/j.0022-3042.2001.00703.x. [DOI] [PubMed] [Google Scholar]
  14. Bargmann CI. Beyond the connectome: how neuromodulators shape neural circuits. BioEssays 34: 458–465, 2012. doi: 10.1002/bies.201100185. [DOI] [PubMed] [Google Scholar]
  15. Basura GJ, Abbas AI, O’Donohue H, Lauder JM, Roth BL, Walker PD, Manis PB. Ontogeny of serotonin and serotonin2A receptors in rat auditory cortex. Hear Res 244: 45–50, 2008. doi: 10.1016/j.heares.2008.07.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Bates AS, Janssens J, Jefferis GS, Aerts S. Neuronal cell types in the fly: single-cell anatomy meets single-cell genomics. Curr Opin Neurobiol 56: 125–134, 2019. doi: 10.1016/j.conb.2018.12.012. [DOI] [PubMed] [Google Scholar]
  17. Batista-Brito R, Vinck M, Ferguson KA, Chang JT, Laubender D, Lur G, Mossner JM, Hernandez VG, Ramakrishnan C, Deisseroth K, Higley MJ, Cardin JA. Developmental dysfunction of VIP interneurons impairs cortical circuits. Neuron 95: 884–895.e9, 2017. doi: 10.1016/j.neuron.2017.07.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Battelle BA, Calman BG, Hart MK. Cellular distributions and functions of histamine, octopamine, and serotonin in the peripheral visual system, brain, and circumesophageal ring of the horseshoe crab Limulus polyphemus. Microsc Res Tech 44: 70–80, 1999. doi:. [DOI] [PubMed] [Google Scholar]
  19. Beaudet A, Descarries L. The fine structure of central serotonin neurons. J Physiol (Paris) 77: 193–203, 1981. [PubMed] [Google Scholar]
  20. Belmer A, Maroteaux L. Regulation of raphe serotonin neurons by serotonin 1A and 2B receptors. Neuropsychopharmacology 44: 218–219, 2019. doi: 10.1038/s41386-018-0214-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Beltz BS, Kravitz EA. Mapping of serotonin-like immunoreactivity in the lobster nervous system. J Neurosci 3: 585–602, 1983. doi: 10.1523/JNEUROSCI.03-03-00585.1983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Bennett C, Arroyo S, Hestrin S. Subthreshold mechanisms underlying state-dependent modulation of visual responses. Neuron 80: 350–357, 2013. doi: 10.1016/j.neuron.2013.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Berck ME, Khandelwal A, Claus L, Hernandez-Nunez L, Si G, Tabone CJ, Li F, Truman JW, Fetter RD, Louis M, Samuel AD, Cardona A. The wiring diagram of a glomerular olfactory system. eLife 5: e14859, 2016. doi: 10.7554/eLife.14859. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Berg BG, Schachtner J, Homberg U. Gamma-aminobutyric acid immunostaining in the antennal lobe of the moth Heliothis virescens and its colocalization with neuropeptides. Cell Tissue Res 335: 593–605, 2009. doi: 10.1007/s00441-008-0744-z. [DOI] [PubMed] [Google Scholar]
  25. Berger M, Gray JA, Roth BL. The expanded biology of serotonin. Annu Rev Med 60: 355–366, 2009. doi: 10.1146/annurev.med.60.042307.110802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Bertrand PP, Bertrand RL. Serotonin release and uptake in the gastrointestinal tract. Auton Neurosci 153: 47–57, 2010. doi: 10.1016/j.autneu.2009.08.002. [DOI] [PubMed] [Google Scholar]
  27. Bishop CA, O’Shea M. Serotonin immunoreactive neurons in the central nervous system of an insect (Periplaneta americana). J Neurobiol 14: 251–269, 1983. doi: 10.1002/neu.480140402. [DOI] [PubMed] [Google Scholar]
  28. Blenau W, Daniel S, Balfanz S, Thamm M, Baumann A. Dm5-HT2B: pharmacological characterization of the fifth serotonin receptor subtype of Drosophila melanogaster. Front Syst Neurosci 11: 28, 2017. doi: 10.3389/fnsys.2017.00028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Blondel O, Gastineau M, Dahmoune Y, Langlois M, Fischmeister R. Cloning, expression, and pharmacology of four human 5-hydroxytryptamine 4 receptor isoforms produced by alternative splicing in the carboxyl terminus. J Neurochem 70: 2252–2261, 1998. doi: 10.1046/j.1471-4159.1998.70062252.x. [DOI] [PubMed] [Google Scholar]
  30. Bockaert J, Claeysen S, Bécamel C, Dumuis A, Marin P. Neuronal 5-HT metabotropic receptors: fine-tuning of their structure, signaling, and roles in synaptic modulation. Cell Tissue Res 326: 553–572, 2006. doi: 10.1007/s00441-006-0286-1. [DOI] [PubMed] [Google Scholar]
  31. Brill J, Shao Z, Puche AC, Wachowiak M, Shipley MT. Serotonin increases synaptic activity in olfactory bulb glomeruli. J Neurophysiol 115: 1208–1219, 2016. doi: 10.1152/jn.00847.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Bruinvels AT, Landwehrmeyer B, Gustafson EL, Durkin MM, Mengod G, Branchek TA, Hoyer D, Palacios JM. Localization of 5-HT1B, 5-HT1D alpha, 5-HT1E and 5-HT1F receptor messenger RNA in rodent and primate brain. Neuropharmacology 33: 367–386, 1994. doi: 10.1016/0028-3908(94)90067-1. [DOI] [PubMed] [Google Scholar]
  33. Bunin MA, Wightman RM. Quantitative evaluation of 5-hydroxytryptamine (serotonin) neuronal release and uptake: an investigation of extrasynaptic transmission. J Neurosci 18: 4854–4860, 1998. doi: 10.1523/JNEUROSCI.18-13-04854.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Bunin MA, Wightman RM. Paracrine neurotransmission in the CNS: involvement of 5-HT. Trends Neurosci 22: 377–382, 1999. doi: 10.1016/S0166-2236(99)01410-1. [DOI] [PubMed] [Google Scholar]
  35. Burns CM, Chu H, Rueter SM, Hutchinson LK, Canton H, Sanders-Bush E, Emeson RB. Regulation of serotonin-2C receptor G-protein coupling by RNA editing. Nature 387: 303–308, 1997. doi: 10.1038/387303a0. [DOI] [PubMed] [Google Scholar]
  36. Cagampang FR, Inouye ST. Diurnal and circadian changes of serotonin in the suprachiasmatic nuclei: regulation by light and an endogenous pacemaker. Brain Res 639: 175–179, 1994. doi: 10.1016/0006-8993(94)91780-9. [DOI] [PubMed] [Google Scholar]
  37. Calizo LH, Akanwa A, Ma X, Pan YZ, Lemos JC, Craige C, Heemstra LA, Beck SG. Raphe serotonin neurons are not homogenous: electrophysiological, morphological and neurochemical evidence. Neuropharmacology 61: 524–543, 2011. doi: 10.1016/j.neuropharm.2011.04.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Cantera R, Nässel DR. Postembryonic development of serotonin-immunoreactive neurons in the central-nervous-system of the blowfly. 2. The thoracico-abdominal ganglia. Cell Tissue Res 250: 449–459, 1987. doi: 10.1007/BF00219091. [DOI] [Google Scholar]
  39. Carandini M, Heeger DJ. Normalization as a canonical neural computation. Nat Rev Neurosci 13: 51–62, 2012. [Erratum in Nat Rev Neurosci 14: 152, 2013.] doi: 10.1038/nrn3136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Cardin JA. Inhibitory interneurons regulate temporal precision and correlations in cortical circuits. Trends Neurosci 41: 689–700, 2018. doi: 10.1016/j.tins.2018.07.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Carlsson MA, Diesner M, Schachtner J, Nässel DR. Multiple neuropeptides in the Drosophila antennal lobe suggest complex modulatory circuits. J Comp Neurol 518: 3359–3380, 2010. doi: 10.1002/cne.22405. [DOI] [PubMed] [Google Scholar]
  42. Cervantes-Ramírez V, Canto-Bustos M, Aguilar-Magaña D, Pérez-Padilla EA, Góngora-Alfaro JL, Pineda JC, Atzori M, Salgado H. Citalopram reduces glutamatergic synaptic transmission in the auditory cortex via activation of 5-HT1A receptors. Neuroreport 30: 1316–1322, 2019. doi: 10.1097/WNR.0000000000001366. [DOI] [PubMed] [Google Scholar]
  43. Chalasani SH, Kato S, Albrecht DR, Nakagawa T, Abbott LF, Bargmann CI. Neuropeptide feedback modifies odor-evoked dynamics in Caenorhabditis elegans olfactory neurons. Nat Neurosci 13: 615–621, 2010. doi: 10.1038/nn.2526. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Chalmers DT, Watson SJ. Comparative anatomical distribution of 5-HT1A receptor mRNA and 5-HT1A binding in rat brain—a combined in situ hybridisation/in vitro receptor autoradiographic study. Brain Res 561: 51–60, 1991. doi: 10.1016/0006-8993(91)90748-K. [DOI] [PubMed] [Google Scholar]
  45. Charest A, Wainer BH, Albert PR. Cloning and differentiation-induced expression of a murine serotonin1A receptor in a septal cell line. J Neurosci 13: 5164–5171, 1993. doi: 10.1523/JNEUROSCI.13-12-05164.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Chazal G, Ralston HJ 3rd. Serotonin-containing structures in the nucleus raphe dorsalis of the cat: an ultrastructural analysis of dendrites, presynaptic dendrites, and axon terminals. J Comp Neurol 259: 317–329, 1987. doi: 10.1002/cne.902590302. [DOI] [PubMed] [Google Scholar]
  47. Chen A, Holmes SP, Pietrantonio PV. Molecular cloning and functional expression of a serotonin receptor from the Southern cattle tick, Boophilus microplus (Acari: Ixodidae). Insect Mol Biol 13: 45–54, 2004. doi: 10.1111/j.1365-2583.2004.00457.x. [DOI] [PubMed] [Google Scholar]
  48. Cheng KY, Frye MA. Neuromodulation of insect motion vision. J Comp Physiol A Neuroethol Sens Neural Behav Physiol 206: 125–137, 2020. doi: 10.1007/s00359-019-01383-9. [DOI] [PubMed] [Google Scholar]
  49. Chiang AS, Lin CY, Chuang CC, Chang HM, Hsieh CH, Yeh CW, Shih CT, Wu JJ, Wang GT, Chen YC, Wu CC, Chen GY, Ching YT, Lee PC, Lin CY, Lin HH, Wu CC, Hsu HW, Huang YA, Chen JY, Chiang HJ, Lu CF, Ni RF, Yeh CY, Hwang JK. Three-dimensional reconstruction of brain-wide wiring networks in Drosophila at single-cell resolution. Curr Biol 21: 1–11, 2011. doi: 10.1016/j.cub.2010.11.056. [DOI] [PubMed] [Google Scholar]
  50. Chidlow G, Le Corre S, Osborne NN. Localization of serotonin1A and serotonin7 receptors in rabbit ocular and brain tissues. Neuroscience 87: 675–689, 1998. doi: 10.1016/S0306-4522(98)00181-X. [DOI] [PubMed] [Google Scholar]
  51. Claeysen S, Sebben M, Becamel C, Bockaert J, Dumuis A. Novel brain-specific 5-HT4 receptor splice variants show marked constitutive activity: role of the C-terminal intracellular domain. Mol Pharmacol 55: 910–920, 1999. [PubMed] [Google Scholar]
  52. Clark MC, Dever TE, Dever JJ, Xu P, Rehder V, Sosa MA, Baro DJ. Arthropod 5-HT2 receptors: a neurohormonal receptor in decapod crustaceans that displays agonist independent activity resulting from an evolutionary alteration to the DRY motif. J Neurosci 24: 3421–3435, 2004. doi: 10.1523/JNEUROSCI.0062-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Coates KE, Majot AT, Zhang X, Michael CT, Spitzer SL, Gaudry Q, Dacks AM. Identified serotonergic modulatory neurons have heterogeneous synaptic connectivity within the olfactory system of Drosophila. J Neurosci 37: 7318–7331, 2017. doi: 10.1523/JNEUROSCI.0192-17.2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Cohen JY, Amoroso MW, Uchida N. Serotonergic neurons signal reward and punishment on multiple timescales. eLife 4: e06346, 2015. doi: 10.7554/eLife.06346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Colas JF, Launay JM, Kellermann O, Rosay P, Maroteaux L. Drosophila 5-HT2 serotonin receptor: coexpression with fushi-tarazu during segmentation. Proc Natl Acad Sci USA 92: 5441–5445, 1995. doi: 10.1073/pnas.92.12.5441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Cook SJ, Jarrell TA, Brittin CA, Wang Y, Bloniarz AE, Yakovlev MA, Nguyen KC, Tang LT, Bayer EA, Duerr JS, Bülow HE, Hobert O, Hall DH, Emmons SW. Whole-animal connectomes of both Caenorhabditis elegans sexes. Nature 571: 63–71, 2019. doi: 10.1038/s41586-019-1352-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Cools R, Roberts AC, Robbins TW. Serotoninergic regulation of emotional and behavioural control processes. Trends Cogn Sci 12: 31–40, 2008. doi: 10.1016/j.tics.2007.10.011. [DOI] [PubMed] [Google Scholar]
  58. Cornea-Hébert V, Riad M, Wu C, Singh SK, Descarries L. Cellular and subcellular distribution of the serotonin 5-HT2A receptor in the central nervous system of adult rat. J Comp Neurol 409: 187–209, 1999. doi:. [DOI] [PubMed] [Google Scholar]
  59. Cornide-Petronio ME, Anadón R, Barreiro-Iglesias A, Rodicio MC. Tryptophan hydroxylase and serotonin receptor 1A expression in the retina of the sea lamprey. Exp Eye Res 135: 81–87, 2015. doi: 10.1016/j.exer.2015.04.017. [DOI] [PubMed] [Google Scholar]
  60. Corthell JT, Stathopoulos AM, Watson CC, Bertram R, Trombley PQ. Olfactory bulb monoamine concentrations vary with time of day. Neuroscience 247: 234–241, 2013. doi: 10.1016/j.neuroscience.2013.05.040. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Cransac H, Cottet-Emard JM, Hellström S, Peyrin L. Specific sound-induced noradrenergic and serotonergic activation in central auditory structures. Hear Res 118: 151–156, 1998. doi: 10.1016/S0378-5955(98)00031-8. [DOI] [PubMed] [Google Scholar]
  62. Dacks AM, Christensen TA, Hildebrand JG. Phylogeny of a serotonin-immunoreactive neuron in the primary olfactory center of the insect brain. J Comp Neurol 498: 727–746, 2006a. doi: 10.1002/cne.21076. [DOI] [PubMed] [Google Scholar]
  63. Dacks AM, Dacks JB, Christensen TA, Nighorn AJ. The cloning of one putative octopamine receptor and two putative serotonin receptors from the tobacco hawkmoth, Manduca sexta. Insect Biochem Mol Biol 36: 741–747, 2006b. doi: 10.1016/j.ibmb.2006.07.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Dacks AM, Green DS, Root CM, Nighorn AJ, Wang JW. Serotonin modulates olfactory processing in the antennal lobe of Drosophila. J Neurogenet 23: 366–377, 2009. doi: 10.3109/01677060903085722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Dacks AM, Reale V, Pi Y, Zhang W, Dacks JB, Nighorn AJ, Evans PD. A characterization of the Manduca sexta serotonin receptors in the context of olfactory neuromodulation. PLoS One 8: e69422, 2013. doi: 10.1371/journal.pone.0069422. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. DeFelipe J, Hendry SH, Hashikawa T, Jones EG. Synaptic relationships of serotonin-immunoreactive terminal baskets on GABA neurons in the cat auditory cortex. Cereb Cortex 1: 117–133, 1991. doi: 10.1093/cercor/1.2.117. [DOI] [PubMed] [Google Scholar]
  67. Desai C, Garriga G, McIntire SL, Horvitz HR. A genetic pathway for the development of the Caenorhabditis elegans HSN motor neurons. Nature 336: 638–646, 1988. doi: 10.1038/336638a0. [DOI] [PubMed] [Google Scholar]
  68. Donaldson ZR, Piel DA, Santos TL, Richardson-Jones J, Leonardo ED, Beck SG, Champagne FA, Hen R. Developmental effects of serotonin 1A autoreceptors on anxiety and social behavior. Neuropsychopharmacology 39: 291–302, 2014. doi: 10.1038/npp.2013.185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Doty RW. Nongeniculate afferents to striate cortex in macaques. J Comp Neurol 218: 159–173, 1983. doi: 10.1002/cne.902180204. [DOI] [PubMed] [Google Scholar]
  70. Duerr JS, Frisby DL, Gaskin J, Duke A, Asermely K, Huddleston D, Eiden LE, Rand JB. The cat-1 gene of Caenorhabditis elegans encodes a vesicular monoamine transporter required for specific monoamine-dependent behaviors. J Neurosci 19: 72–84, 1999. doi: 10.1523/JNEUROSCI.19-01-00072.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Dyck RH, Cynader MS. Autoradiographic localization of serotonin receptor subtypes in cat visual cortex: transient regional, laminar, and columnar distributions during postnatal development. J Neurosci 13: 4316–4338, 1993. doi: 10.1523/JNEUROSCI.13-10-04316.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Eid L, Champigny MF, Parent A, Parent M. Quantitative and ultrastructural study of serotonin innervation of the globus pallidus in squirrel monkeys. Eur J Neurosci 37: 1659–1668, 2013. doi: 10.1111/ejn.12164. [DOI] [PubMed] [Google Scholar]
  73. El-Merahbi R, Löffler M, Mayer A, Sumara G. The roles of peripheral serotonin in metabolic homeostasis. FEBS Lett 589: 1728–1734, 2015. doi: 10.1016/j.febslet.2015.05.054. [DOI] [PubMed] [Google Scholar]
  74. Emmons SW. The beginning of connectomics: a commentary on White et al. (1986) ‘The structure of the nervous system of the nematode Caenorhabditis elegans’. Philos Trans R Soc Lond B Biol Sci 370: 20140309, 2015. doi: 10.1098/rstb.2014.0309. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Erlander MG, Lovenberg TW, Baron BM, de Lecea L, Danielson PE, Racke M, Slone AL, Siegel BW, Foye PE, Cannon K. Two members of a distinct subfamily of 5-hydroxytryptamine receptors differentially expressed in rat brain. Proc Natl Acad Sci USA 90: 3452–3456, 1993. doi: 10.1073/pnas.90.8.3452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Fernandez SP, Cauli B, Cabezas C, Muzerelle A, Poncer JC, Gaspar P. Multiscale single-cell analysis reveals unique phenotypes of raphe 5-HT neurons projecting to the forebrain. Brain Struct Funct 221: 4007–4025, 2016. doi: 10.1007/s00429-015-1142-4. [DOI] [PubMed] [Google Scholar]
  77. Fickbohm DJ, Spitzer N, Katz PS. Pharmacological manipulation of serotonin levels in the nervous system of the opisthobranch mollusc Tritonia diomedea. Biol Bull 209: 67–74, 2005. doi: 10.2307/3593142. [DOI] [PubMed] [Google Scholar]
  78. Foehring RC, van Brederode JF, Kinney GA, Spain WJ. Serotonergic modulation of supragranular neurons in rat sensorimotor cortex. J Neurosci 22: 8238–8250, 2002. doi: 10.1523/JNEUROSCI.22-18-08238.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Foguet M, Hoyer D, Pardo LA, Parekh A, Kluxen FW, Kalkman HO, Stühmer W, Lübbert H. Cloning and functional characterization of the rat stomach fundus serotonin receptor. EMBO J 11: 3481–3487, 1992. doi: 10.1002/j.1460-2075.1992.tb05427.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Fornal CA, Metzler CW, Marrosu F, Ribiero-do-Valle LE, Jacobs BL. A subgroup of dorsal raphe serotonergic neurons in the cat is strongly activated during oral-buccal movements. Brain Res 716: 123–133, 1996. doi: 10.1016/0006-8993(96)00006-6. [DOI] [PubMed] [Google Scholar]
  81. Fotowat H, Harvey-Girard E, Cheer JF, Krahe R, Maler L. Subsecond sensory modulation of serotonin levels in a primary sensory area and its relation to ongoing communication behavior in a weakly electric fish. eNeuro 3: ENEURO.0115-16.2016, 2016. doi: 10.1523/ENEURO.0115-16.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Fu W, Le Maître E, Fabre V, Bernard JF, David Xu ZQ, Hökfelt T. Chemical neuroanatomy of the dorsal raphe nucleus and adjacent structures of the mouse brain. J Comp Neurol 518: 3464–3494, 2010. doi: 10.1002/cne.22407. [DOI] [PubMed] [Google Scholar]
  83. Fu Y, Tucciarone JM, Espinosa JS, Sheng N, Darcy DP, Nicoll RA, Huang ZJ, Stryker MP. A cortical circuit for gain control by behavioral state. Cell 156: 1139–1152, 2014. doi: 10.1016/j.cell.2014.01.050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Fuxe K, Agnati LF, Marcoli M, Borroto-Escuela DO. Volume transmission in central dopamine and noradrenaline neurons and its astroglial targets. Neurochem Res 40: 2600–2614, 2015. doi: 10.1007/s11064-015-1574-5. [DOI] [PubMed] [Google Scholar]
  85. Gao K, Mason P. Serotonergic raphe magnus cells that respond to noxious tail heat are not ON or OFF cells. J Neurophysiol 84: 1719–1725, 2000. doi: 10.1152/jn.2000.84.4.1719. [DOI] [PubMed] [Google Scholar]
  86. García-Oscos F, Torres-Ramírez O, Dinh L, Galindo-Charles L, Pérez Padilla EA, Pineda JC, Atzori M, Salgado H. Activation of 5-HT receptors inhibits GABAergic transmission by pre-and post-synaptic mechanisms in layer II/III of the juvenile rat auditory cortex. Synapse 69: 115–127, 2015. doi: 10.1002/syn.21794. [DOI] [PubMed] [Google Scholar]
  87. Gaspar P, Lillesaar C. Probing the diversity of serotonin neurons. Philos Trans R Soc Lond B Biol Sci 367: 2382–2394, 2012. doi: 10.1098/rstb.2011.0378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Gasque G, Conway S, Huang J, Rao Y, Vosshall LB. Small molecule drug screening in Drosophila identifies the 5HT2A receptor as a feeding modulation target. Sci Rep 3: srep02120, 2013. doi: 10.1038/srep02120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Gaudry Q. Serotonergic modulation of olfaction in rodents and insects. Yale J Biol Med 91: 23–32, 2018. [PMC free article] [PubMed] [Google Scholar]
  90. Gaudry Q, Kristan WB Jr. Behavioral choice by presynaptic inhibition of tactile sensory terminals. Nat Neurosci 12: 1450–1457, 2009. doi: 10.1038/nn.2400. [DOI] [PubMed] [Google Scholar]
  91. Gerald C, Adham N, Kao HT, Olsen MA, Laz TM, Schechter LE, Bard JA, Vaysse PJ, Hartig PR, Branchek TA. The 5-HT4 receptor: molecular cloning and pharmacological characterization of two splice variants. EMBO J 14: 2806–2815, 1995. doi: 10.1002/j.1460-2075.1995.tb07280.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Gerhardt CC, Leysen JE, Planta RJ, Vreugdenhil E, Van Heerikhuizen H. Functional characterisation of a 5-HT2 receptor cDNA cloned from Lymnaea stagnalis. Eur J Pharmacol 311: 249–258, 1996. doi: 10.1016/0014-2999(96)00410-4. [DOI] [PubMed] [Google Scholar]
  93. Gershon MD. 5-Hydroxytryptamine (serotonin) in the gastrointestinal tract. Curr Opin Endocrinol Diabetes Obes 20: 14–21, 2013. doi: 10.1097/MED.0b013e32835bc703. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Gracia-Llanes FJ, Blasco-Ibáñez JM, Nácher J, Varea E, Liberia T, Martínez P, Martínez-Guijarro FJ, Crespo C. Synaptic connectivity of serotonergic axons in the olfactory glomeruli of the rat olfactory bulb. Neuroscience 169: 770–780, 2010. doi: 10.1016/j.neuroscience.2010.05.034. [DOI] [PubMed] [Google Scholar]
  95. Grailhe R, Grabtree GW, Hen R. Human 5-HT5 receptors: the 5-HT5A receptor is functional but the 5-HT5B receptor was lost during mammalian evolution. Eur J Pharmacol 418: 157–167, 2001. doi: 10.1016/S0014-2999(01)00933-5. [DOI] [PubMed] [Google Scholar]
  96. Hall IC, Rebec GV, Hurley LM. Serotonin in the inferior colliculus fluctuates with behavioral state and environmental stimuli. J Exp Biol 213: 1009–1017, 2010. doi: 10.1242/jeb.035956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Hall IC, Sell GL, Hurley LM. Social regulation of serotonin in the auditory midbrain. Behav Neurosci 125: 501–511, 2011. doi: 10.1037/a0024426. [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Hamada S, Senzaki K, Hamaguchi-Hamada K, Tabuchi K, Yamamoto H, Yamamoto T, Yoshikawa S, Okano H, Okado N. Localization of 5-HT2A receptor in rat cerebral cortex and olfactory system revealed by immunohistochemistry using two antibodies raised in rabbit and chicken. Brain Res Mol Brain Res 54: 199–211, 1998. doi: 10.1016/S0169-328X(97)00322-7. [DOI] [PubMed] [Google Scholar]
  99. Hamanaka Y, Kinoshita M, Homberg U, Arikawa K. Immunocytochemical localization of amines and GABA in the optic lobe of the butterfly, Papilio xuthus. PLoS One 7: e41109, 2012. doi: 10.1371/journal.pone.0041109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Hamblin MW, Metcalf MA, McGuffin RW, Karpells S. Molecular cloning and functional characterization of a human 5-HT1B serotonin receptor: a homologue of the rat 5-HT1B receptor with 5-HT1D-like pharmacological specificity. Biochem Biophys Res Commun 184: 752–759, 1992. doi: 10.1016/0006-291X(92)90654-4. [DOI] [PubMed] [Google Scholar]
  101. Hamdan FF, Ungrin MD, Abramovitz M, Ribeiro P. Characterization of a novel serotonin receptor from Caenorhabditis elegans: cloning and expression of two splice variants. J Neurochem 72: 1372–1383, 1999. doi: 10.1046/j.1471-4159.1999.721372.x. [DOI] [PubMed] [Google Scholar]
  102. Han L, Zhong YM, Yang XL. 5-HT2A receptors are differentially expressed in bullfrog and rat retinas: a comparative study. Brain Res Bull 73: 273–277, 2007. doi: 10.1016/j.brainresbull.2007.04.005. [DOI] [PubMed] [Google Scholar]
  103. Hannon J, Hoyer D. Molecular biology of 5-HT receptors. Behav Brain Res 195: 198–213, 2008. doi: 10.1016/j.bbr.2008.03.020. [DOI] [PubMed] [Google Scholar]
  104. Hanson JL, Hurley LM. Context-dependent fluctuation of serotonin in the auditory midbrain: the influence of sex, reproductive state and experience. J Exp Biol 217: 526–535, 2014. doi: 10.1242/jeb.087627. [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Hardy A, Palouzier-Paulignan B, Duchamp A, Royet JP, Duchamp-Viret P. 5-Hydroxytryptamine action in the rat olfactory bulb: in vitro electrophysiological patch-clamp recordings of juxtaglomerular and mitral cells. Neuroscience 131: 717–731, 2005. doi: 10.1016/j.neuroscience.2004.10.034. [DOI] [PubMed] [Google Scholar]
  106. Harris G, Mills H, Wragg R, Hapiak V, Castelletto M, Korchnak A, Komuniecki RW. The monoaminergic modulation of sensory-mediated aversive responses in Caenorhabditis elegans requires glutamatergic/peptidergic cotransmission. J Neurosci 30: 7889–7899, 2010. doi: 10.1523/JNEUROSCI.0497-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Hartline DK, Christie AE. Immunohistochemical mapping of histamine, dopamine, and serotonin in the central nervous system of the copepod Calanus finmarchicus (Crustacea; Maxillopoda; Copepoda). Cell Tissue Res 341: 49–71, 2010. doi: 10.1007/s00441-010-0974-8. [DOI] [PubMed] [Google Scholar]
  108. Harzsch S, Dawirs RR. A developmental study of serotonin-immunoreactive neurons in the larval central nervous system of the spider crab Hyas araneus (Decapoda, Brachyura). Invert Neurosci 1: 53–65, 1995. doi: 10.1007/BF02331832. [DOI] [PubMed] [Google Scholar]
  109. Hay-Schmidt A. The evolution of the serotonergic nervous system. Proc Biol Sci 267: 1071–1079, 2000. doi: 10.1098/rspb.2000.1111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Herculano-Houzel S, Mota B, Lent R. Cellular scaling rules for rodent brains. Proc Natl Acad Sci USA 103: 12138–12143, 2006. doi: 10.1073/pnas.0604911103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Heym J, Steinfels GF, Jacobs BL. Activity of serotonin-containing neurons in the nucleus raphe pallidus of freely moving cats. Brain Res 251: 259–276, 1982. doi: 10.1016/0006-8993(82)90743-0. [DOI] [PubMed] [Google Scholar]
  112. Hill ES, Iwano M, Gatellier L, Kanzaki R. Morphology and physiology of the serotonin-immunoreactive putative antennal lobe feedback neuron in the male silkmoth Bombyx mori. Chem Senses 27: 475–483, 2002. doi: 10.1093/chemse/27.5.475. [DOI] [PubMed] [Google Scholar]
  113. Hobson RJ, Hapiak VM, Xiao H, Buehrer KL, Komuniecki PR, Komuniecki RW. SER-7, a Caenorhabditis elegans 5-HT7-like receptor, is essential for the 5-HT stimulation of pharyngeal pumping and egg laying. Genetics 172: 159–169, 2006. doi: 10.1534/genetics.105.044495. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Homberg U, Hildebrand JG. Serotonin-immunoreactive neurons in the median protocerebrum and suboesophageal ganglion of the sphinx moth Manduca sexta. Cell Tissue Res 258: 1–24, 1989a. doi: 10.1007/BF00223139. [DOI] [PubMed] [Google Scholar]
  115. Homberg U, Hildebrand JG. Serotonin immunoreactivity in the optic lobes of the sphinx moth Manduca sexta and colocalization with FMRFamide and SCPB immunoreactivity. J Comp Neurol 288: 243–253, 1989b. doi: 10.1002/cne.902880204. [DOI] [PubMed] [Google Scholar]
  116. Hornung JP. The neuroanatomy of the serotonergic system. Handb Behav Neurosci 21: 51–64, 2010. doi: 10.1016/S1569-7339(10)70071-0. [DOI] [Google Scholar]
  117. Horvitz HR, Chalfie M, Trent C, Sulston JE, Evans PD. Serotonin and octopamine in the nematode Caenorhabditis elegans. Science 216: 1012–1014, 1982. doi: 10.1126/science.6805073. [DOI] [PubMed] [Google Scholar]
  118. Howard CE, Chen CL, Tabachnik T, Hormigo R, Ramdya P, Mann RS. Serotonergic modulation of walking in Drosophila. Curr Biol 29: 4218–4230.e8, 2019. doi: 10.1016/j.cub.2019.10.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  119. Huang J, Zhang W, Qiao W, Hu A, Wang Z. Functional connectivity and selective odor responses of excitatory local interneurons in Drosophila antennal lobe. Neuron 67: 1021–1033, 2010. doi: 10.1016/j.neuron.2010.08.025. [DOI] [PubMed] [Google Scholar]
  120. Huang KW, Ochandarena NE, Philson AC, Hyun M, Birnbaum JE, Cicconet M, Sabatini BL. Molecular and anatomical organization of the dorsal raphe nucleus. eLife 8: e46464, 2019. doi: 10.7554/eLife.46464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Huang Z, Thiebaud N, Fadool DA. Differential serotonergic modulation across the main and accessory olfactory bulbs. J Physiol 595: 3515–3533, 2017. doi: 10.1113/JP273945. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Hummel NA, Li AY, Witt CM. Serotonin-like immunoreactivity in the central nervous system of two ixodid tick species. Exp Appl Acarol 43: 265–278, 2007. doi: 10.1007/s10493-007-9120-z. [DOI] [PubMed] [Google Scholar]
  123. Hurley LM, Devilbiss DM, Waterhouse BD. A matter of focus: monoaminergic modulation of stimulus coding in mammalian sensory networks. Curr Opin Neurobiol 14: 488–495, 2004. doi: 10.1016/j.conb.2004.06.007. [DOI] [PubMed] [Google Scholar]
  124. Hurley LM, Pollak GD. Serotonin differentially modulates responses to tones and frequency-modulated sweeps in the inferior colliculus. J Neurosci 19: 8071–8082, 1999. doi: 10.1523/JNEUROSCI.19-18-08071.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  125. Hurley LM, Pollak GD. Serotonin effects on frequency tuning of inferior colliculus neurons. J Neurophysiol 85: 828–842, 2001. doi: 10.1152/jn.2001.85.2.828. [DOI] [PubMed] [Google Scholar]
  126. Hurley LM, Thompson AM. Serotonergic innervation of the auditory brainstem of the Mexican free-tailed bat, Tadarida brasiliensis. J Comp Neurol 435: 78–88, 2001. doi: 10.1002/cne.1194. [DOI] [PubMed] [Google Scholar]
  127. Huser A, Eschment M, Güllü N, Collins KA, Böpple K, Pankevych L, Rolsing E, Thum AS. Anatomy and behavioral function of serotonin receptors in Drosophila melanogaster larvae. PLoS One 12: e0181865, 2017. doi: 10.1371/journal.pone.0181865. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Impieri D, Zilles K, Niu M, Rapan L, Schubert N, Galletti C, Palomero-Gallagher N. Receptor density pattern confirms and enhances the anatomic-functional features of the macaque superior parietal lobule areas. Brain Struct Funct 224: 2733–2756, 2019. doi: 10.1007/s00429-019-01930-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Ishimura K, Takeuchi Y, Fujiwara K, Tominaga M, Yoshioka H, Sawada T. Quantitative analysis of the distribution of serotonin-immunoreactive cell bodies in the mouse brain. Neurosci Lett 91: 265–270, 1988. doi: 10.1016/0304-3940(88)90691-X. [DOI] [PubMed] [Google Scholar]
  130. Ishita Y, Chihara T, Okumura M. Serotonergic modulation of feeding behavior in Caenorhabditis elegans and other related nematodes. Neurosci Res 154: 9–19, 2020. doi: 10.1016/j.neures.2019.04.006. [DOI] [PubMed] [Google Scholar]
  131. Iwanir S, Brown AS, Nagy S, Najjar D, Kazakov A, Lee KS, Zaslaver A, Levine E, Biron D. Serotonin promotes exploitation in complex environments by accelerating decision-making. BMC Biol 14: 9, 2016. doi: 10.1186/s12915-016-0232-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Iwano M, Kanzaki R. Immunocytochemical identification of neuroactive substances in the antennal lobe of the male silkworm moth Bombyx mori. Zool Sci 22: 199–211, 2005. doi: 10.2108/zsj.22.199. [DOI] [PubMed] [Google Scholar]
  133. Jacob SN, Nienborg H. Monoaminergic neuromodulation of sensory processing. Front Neural Circuits 12: 51, 2018. doi: 10.3389/fncir.2018.00051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  134. Jacobs BL, Azmitia EC. Structure and function of the brain serotonin system. Physiol Rev 72: 165–229, 1992. doi: 10.1152/physrev.1992.72.1.165. [DOI] [PubMed] [Google Scholar]
  135. Jacobs BL, Fornal CA. Activity of brain serotonergic neurons in the behaving animal. Pharmacol Rev 43: 563–578, 1991. [PubMed] [Google Scholar]
  136. Jacobs BL, Heym J, Trulson ME. Behavioral and physiological correlates of brain serotoninergic unit activity. J Physiol (Paris) 77: 431–436, 1981. [PubMed] [Google Scholar]
  137. Jacobs BL, Martín-Cora FJ, Fornal CA. Activity of medullary serotonergic neurons in freely moving animals. Brain Res Brain Res Rev 40: 45–52, 2002. doi: 10.1016/S0165-0173(02)00187-X. [DOI] [PubMed] [Google Scholar]
  138. Jang HJ, Cho KH, Park SW, Kim MJ, Yoon SH, Rhie DJ. Layer-specific serotonergic facilitation of IPSC in layer 2/3 pyramidal neurons of the visual cortex. J Neurophysiol 107: 407–416, 2012. doi: 10.1152/jn.00535.2011. [DOI] [PubMed] [Google Scholar]
  139. Jiang X, Shen S, Cadwell CR, Berens P, Sinz F, Ecker AS, Patel S, Tolias AS. Principles of connectivity among morphologically defined cell types in adult neocortex. Science 350: aac9462, 2015. doi: 10.1126/science.aac9462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Julius D, MacDermott AB, Axel R, Jessell TM. Molecular characterization of a functional cDNA encoding the serotonin 1c receptor. Science 241: 558–564, 1988. doi: 10.1126/science.3399891. [DOI] [PubMed] [Google Scholar]
  141. Kaneko T, Macara AM, Li R, Hu Y, Iwasaki K, Dunnings Z, Firestone E, Horvatic S, Guntur A, Shafer OT, Yang CH, Zhou J, Ye B. Serotonergic modulation enables pathway-specific plasticity in a developing sensory circuit in Drosophila. Neuron 95: 623–638, 2017. [Erratum in Neuron 95: 722, 2017.] doi: 10.1016/j.neuron.2017.06.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Kapoor V, Provost AC, Agarwal P, Murthy VN. Activation of raphe nuclei triggers rapid and distinct effects on parallel olfactory bulb output channels. Nat Neurosci 19: 271–282, 2016. doi: 10.1038/nn.4219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Kasture AS, Bartel D, Steinkellner T, Sucic S, Hummel T, Freissmuth M. Distinct contribution of axonal and somatodendritic serotonin transporters in drosophila olfaction. Neuropharmacology 161: 107564, 2019. doi: 10.1016/j.neuropharm.2019.03.007. [DOI] [PubMed] [Google Scholar]
  144. Katz PS. Beyond Neurotransmission: Neuromodulation and Its Importance for Information Processing. Oxford, UK: Oxford Univ. Press, 1999. [Google Scholar]
  145. Keesom SM, Hurley LM. Socially induced serotonergic fluctuations in the male auditory midbrain correlate with female behavior during courtship. J Neurophysiol 115: 1786–1796, 2016. doi: 10.1152/jn.00742.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Keesom SM, Morningstar MD, Sandlain R, Wise BM, Hurley LM. Social isolation reduces serotonergic fiber density in the inferior colliculus of female, but not male, mice. Brain Res 1694: 94–103, 2018. doi: 10.1016/j.brainres.2018.05.010. [DOI] [PubMed] [Google Scholar]
  147. Keesom SM, Sloss BG, Erbowor-Becksen Z, Hurley LM. Social experience alters socially induced serotonergic fluctuations in the inferior colliculus. J Neurophysiol 118: 3230–3241, 2017. doi: 10.1152/jn.00431.2017. [DOI] [PubMed] [Google Scholar]
  148. Kim KS, Kim MA, Sohn YC. Molecular characterization, expression analysis, and functional properties of multiple 5-hydroxytryptamine receptors in Pacific abalone (Haliotis discus hannai). Gen Comp Endocrinol 276: 52–59, 2019. doi: 10.1016/j.ygcen.2019.03.001. [DOI] [PubMed] [Google Scholar]
  149. Klein MT, Teitler M. Distribution of 5-ht1E receptors in the mammalian brain and cerebral vasculature: an immunohistochemical and pharmacological study. Br J Pharmacol 166: 1290–1302, 2012. doi: 10.1111/j.1476-5381.2012.01868.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Klemm N, Steinbusch HW, Sundler F. Distribution of serotonin-containing neurons and their pathways in the supraoesophageal ganglion of the cockroach Periplaneta americana (L.) as revealed by immunocytochemistry. J Comp Neurol 225: 387–395, 1984. doi: 10.1002/cne.902250306. [DOI] [PubMed] [Google Scholar]
  151. Klepper A, Herbert H. Distribution and origin of noradrenergic and serotonergic fibers in the cochlear nucleus and inferior colliculus of the rat. Brain Res 557: 190–201, 1991. doi: 10.1016/0006-8993(91)90134-H. [DOI] [PubMed] [Google Scholar]
  152. Kloppenburg P, Ferns D, Mercer AR. Serotonin enhances central olfactory neuron responses to female sex pheromone in the male sphinx moth Manduca sexta. J Neurosci 19: 8172–8181, 1999. doi: 10.1523/JNEUROSCI.19-19-08172.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Kloppenburg P, Hildebrand JG. Neuromodulation by 5-hydroxytryptamine in the antennal lobe of the sphinx moth Manduca sexta. J Exp Biol 198: 603–611, 1995. [DOI] [PubMed] [Google Scholar]
  154. Kohen R, Fashingbauer LA, Heidmann DE, Guthrie CR, Hamblin MW. Cloning of the mouse 5-HT6 serotonin receptor and mutagenesis studies of the third cytoplasmic loop. Brain Res Mol Brain Res 90: 110–117, 2001. doi: 10.1016/S0169-328X(01)00090-0. [DOI] [PubMed] [Google Scholar]
  155. Kohen R, Metcalf MA, Khan N, Druck T, Huebner K, Lachowicz JE, Meltzer HY, Sibley DR, Roth BL, Hamblin MW. Cloning, characterization, and chromosomal localization of a human 5-HT6 serotonin receptor. J Neurochem 66: 47–56, 1996. doi: 10.1046/j.1471-4159.1996.66010047.x. [DOI] [PubMed] [Google Scholar]
  156. Kohl J, Jefferis GS. Neuroanatomy: decoding the fly brain. Curr Biol 21: R19–R20, 2011. doi: 10.1016/j.cub.2010.11.067. [DOI] [PubMed] [Google Scholar]
  157. Kursar JD, Nelson DL, Wainscott DB, Baez M. Molecular cloning, functional expression, and mRNA tissue distribution of the human 5-hydroxytryptamine2B receptor. Mol Pharmacol 46: 227–234, 1994. [PubMed] [Google Scholar]
  158. Lange AB, Orchard I, Lloyd RJ. Immunohistochemical and electrochemical detection of serotonin in the nervous system of the blood-feeding bug, Rhodnius prolixus. Arch Insect Biochem Physiol 8: 187–201, 1988. doi: 10.1002/arch.940080305. [DOI] [Google Scholar]
  159. Lee DW, Pietrantonio PV. In vitro expression and pharmacology of the 5-HT7-like receptor present in the mosquito Aedes aegypti tracheolar cells and hindgut-associated nerves. Insect Mol Biol 12: 561–569, 2003. doi: 10.1046/j.1365-2583.2003.00441.x. [DOI] [PubMed] [Google Scholar]
  160. Lee KK, Soutar CN, Dringenberg HC. Gating of long-term potentiation (LTP) in the thalamocortical auditory system of rats by serotonergic (5-HT) receptors. Brain Res 1683: 1–11, 2018. doi: 10.1016/j.brainres.2018.01.004. [DOI] [PubMed] [Google Scholar]
  161. Lee S, Hjerling-Leffler J, Zagha E, Fishell G, Rudy B. The largest group of superficial neocortical GABAergic interneurons expresses ionotropic serotonin receptors. J Neurosci 30: 16796–16808, 2010. doi: 10.1523/JNEUROSCI.1869-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  162. Lee SB, Lee HS, Waterhouse BD. The collateral projection from the dorsal raphe nucleus to whisker-related, trigeminal sensory and facial motor systems in the rat. Brain Res 1214: 11–22, 2008. doi: 10.1016/j.brainres.2008.04.003. [DOI] [PubMed] [Google Scholar]
  163. Lee YS, Choi SL, Lee SH, Kim H, Park H, Lee N, Lee SH, Chae YS, Jang DJ, Kandel ER, Kaang BK. Identification of a serotonin receptor coupled to adenylyl cyclase involved in learning-related heterosynaptic facilitation in Aplysia. Proc Natl Acad Sci USA 106: 14634–14639, 2009. doi: 10.1073/pnas.0907502106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Leitinger G, Pabst MA, Kral K. Serotonin-immunoreactive neurones in the visual system of the praying mantis: an immunohistochemical, confocal laser scanning and electron microscopic study. Brain Res 823: 11–23, 1999. doi: 10.1016/S0006-8993(98)01339-0. [DOI] [PubMed] [Google Scholar]
  165. Lent CM, Zundel D, Freedman E, Groome JR. Serotonin in the leech central nervous system: anatomical correlates and behavioral effects. J Comp Physiol A Neuroethol Sens Neural Behav Physiol 168: 191–200, 1991. doi: 10.1007/BF00218411. [DOI] [PubMed] [Google Scholar]
  166. Li Q, Wichems CH, Ma L, Van de Kar LD, Garcia F, Murphy DL. Brain region-specific alterations of 5-HT2A and 5-HT2C receptors in serotonin transporter knockout mice. J Neurochem 84: 1256–1265, 2003. doi: 10.1046/j.1471-4159.2003.01607.x. [DOI] [PubMed] [Google Scholar]
  167. Li Y, Zhong W, Wang D, Feng Q, Liu Z, Zhou J, Jia C, Hu F, Zeng J, Guo Q, Fu L, Luo M. Serotonin neurons in the dorsal raphe nucleus encode reward signals. Nat Commun 7: 10503, 2016. doi: 10.1038/ncomms10503. [DOI] [PMC free article] [PubMed] [Google Scholar]
  168. Liu H, Qin LW, Li R, Zhang C, Al-Sheikh U, Wu ZX. Reciprocal modulation of 5-HT and octopamine regulates pumping via feedforward and feedback circuits in C. elegans. Proc Natl Acad Sci USA 116: 7107–7112, 2019. [Erratum in Proc Natl Acad Sci USA 116: 10598, 2019.] doi: 10.1073/pnas.1819261116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Liu S, Aungst JL, Puche AC, Shipley MT. Serotonin modulates the population activity profile of olfactory bulb external tufted cells. J Neurophysiol 107: 473–483, 2012. doi: 10.1152/jn.00741.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. Liu Y, Luo J, Carlsson MA, Nässel DR. Serotonin and insulin-like peptides modulate leucokinin-producing neurons that affect feeding and water homeostasis in Drosophila. J Comp Neurol 523: 1840–1863, 2015. doi: 10.1002/cne.23768. [DOI] [PubMed] [Google Scholar]
  171. Liu Z, Zhou J, Li Y, Hu F, Lu Y, Ma M, Feng Q, Zhang JE, Wang D, Zeng J, Bao J, Kim JY, Chen ZF, El Mestikawy S, Luo M. Dorsal raphe neurons signal reward through 5-HT and glutamate. Neuron 81: 1360–1374, 2014. doi: 10.1016/j.neuron.2014.02.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  172. Lizbinski KM, Dacks AM. Intrinsic and extrinsic neuromodulation of olfactory processing. Front Cell Neurosci 11: 424, 2018. doi: 10.3389/fncel.2017.00424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Lizbinski KM, Marsat G, Dacks AM. Systematic analysis of transmitter coexpression reveals organizing principles of local interneuron heterogeneity. eNeuro 5: ENEURO.0212-18.2018, 2018. doi: 10.1523/ENEURO.0212-18.2018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Lizbinski KM, Metheny JD, Bradley SP, Kesari A, Dacks AM. The anatomical basis for modulatory convergence in the antennal lobe of Manduca sexta. J Comp Neurol 524: 1859–1875, 2016. doi: 10.1002/cne.23926. [DOI] [PMC free article] [PubMed] [Google Scholar]
  175. Loer CM, Kenyon CJ. Serotonin-deficient mutants and male mating behavior in the nematode Caenorhabditis elegans. J Neurosci 13: 5407–5417, 1993. doi: 10.1523/JNEUROSCI.13-12-05407.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Longley AJ, Longley RD. Serotonin immunoreactivity in the nervous system of the dragonfly nymph. J Neurobiol 17: 329–338, 1986. doi: 10.1002/neu.480170407. [DOI] [PubMed] [Google Scholar]
  177. Loric S, Launay JM, Colas JF, Maroteaux L. New mouse 5-HT2-like receptor. Expression in brain, heart and intestine. FEBS Lett 312: 203–207, 1992. doi: 10.1016/0014-5793(92)80936-B. [DOI] [PubMed] [Google Scholar]
  178. Lottem E, Lörincz ML, Mainen ZF. Optogenetic activation of dorsal raphe serotonin neurons rapidly inhibits spontaneous but not odor-evoked activity in olfactory cortex. J Neurosci 36: 7–18, 2016. doi: 10.1523/JNEUROSCI.3008-15.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Lovenberg TW, Erlander MG, Baron BM, Racke M, Slone AL, Siegel BW, Craft CM, Burns JE, Danielson PE, Sutcliffe JG. Molecular cloning and functional expression of 5-HT1E-like rat and human 5-hydroxytryptamine receptor genes. Proc Natl Acad Sci USA 90: 2184–2188, 1993. doi: 10.1073/pnas.90.6.2184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Luo J, Becnel J, Nichols CD, Nässel DR. Insulin-producing cells in the brain of adult Drosophila are regulated by the serotonin 5-HT1A receptor. Cell Mol Life Sci 69: 471–484, 2012. doi: 10.1007/s00018-011-0789-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Luo M, Li Y, Zhong W. Do dorsal raphe 5-HT neurons encode “beneficialness”? Neurobiol Learn Mem 135: 40–49, 2016. doi: 10.1016/j.nlm.2016.08.008. [DOI] [PubMed] [Google Scholar]
  182. Malgouris C, Flamand F, Doble A. Autoradiographic studies of RP 62203, a potent 5-HT2 receptor antagonist. Pharmacological characterization of [3H]RP 62203 binding in the rat brain. Eur J Pharmacol 233: 37–45, 1993. doi: 10.1016/0014-2999(93)90346-J. [DOI] [PubMed] [Google Scholar]
  183. Marder E. Neuromodulation of neuronal circuits: back to the future. Neuron 76: 1–11, 2012. doi: 10.1016/j.neuron.2012.09.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  184. Maroteaux L, Saudou F, Amlaiky N, Boschert U, Plassat JL, Hen R. Mouse 5HT1B serotonin receptor: cloning, functional expression, and localization in motor control centers. Proc Natl Acad Sci USA 89: 3020–3024, 1992. doi: 10.1073/pnas.89.7.3020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Martin JP, Beyerlein A, Dacks AM, Reisenman CE, Riffell JA, Lei H, Hildebrand JG. The neurobiology of insect olfaction: sensory processing in a comparative context. Prog Neurobiol 95: 427–447, 2011. doi: 10.1016/j.pneurobio.2011.09.007. [DOI] [PubMed] [Google Scholar]
  186. Masson J. Serotonin in retina. Biochimie 161: 51–55, 2019. doi: 10.1016/j.biochi.2018.11.006. [DOI] [PubMed] [Google Scholar]
  187. Matias S, Lottem E, Dugué GP, Mainen ZF. Activity patterns of serotonin neurons underlying cognitive flexibility. eLife 6: e20552, 2017. doi: 10.7554/eLife.20552. [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Matthes H, Boschert U, Amlaiky N, Grailhe R, Plassat JL, Muscatelli F, Mattei MG, Hen R. Mouse 5-hydroxytryptamine5A and 5-hydroxytryptamine5B receptors define a new family of serotonin receptors: cloning, functional expression, and chromosomal localization. Mol Pharmacol 43: 313–319, 1993. [PubMed] [Google Scholar]
  189. Matthes S, Bader M. Peripheral serotonin synthesis as a new drug target. Trends Pharmacol Sci 39: 560–572, 2018. doi: 10.1016/j.tips.2018.03.004. [DOI] [PubMed] [Google Scholar]
  190. McAllister G, Charlesworth A, Snodin C, Beer MS, Noble AJ, Middlemiss DN, Iversen LL, Whiting P. Molecular cloning of a serotonin receptor from human brain (5HT1E): a fifth 5HT1-like subtype. Proc Natl Acad Sci USA 89: 5517–5521, 1992. doi: 10.1073/pnas.89.12.5517. [DOI] [PMC free article] [PubMed] [Google Scholar]
  191. McBurney-Lin J, Lu J, Zuo Y, Yang H. Locus coeruleus-norepinephrine modulation of sensory processing and perception: a focused review. Neurosci Biobehav Rev 105: 190–199, 2019. doi: 10.1016/j.neubiorev.2019.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. McLean JH, Shipley MT. Serotonergic afferents to the rat olfactory bulb: I. Origins and laminar specificity of serotonergic inputs in the adult rat. J Neurosci 7: 3016–3028, 1987. doi: 10.1523/JNEUROSCI.07-10-03016.1987. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Meinertzhagen IA. Of what use is connectomics? A personal perspective on the Drosophila connectome. J Exp Biol 221: jeb164954, 2018. doi: 10.1242/jeb.164954. [DOI] [PubMed] [Google Scholar]
  194. Mengod G, Nguyen H, Le H, Waeber C, Lübbert H, Palacios JM. The distribution and cellular localization of the serotonin 1C receptor mRNA in the rodent brain examined by in situ hybridization histochemistry. Comparison with receptor binding distribution. Neuroscience 35: 577–591, 1990. doi: 10.1016/0306-4522(90)90330-7. [DOI] [PubMed] [Google Scholar]
  195. Miller A. The internal anatomy and histology of the imago of Drosophila melanogaster. In: Biology of Drosophila, edited by Demerec M. New York: John Wiley and Sons, 1950, p. 420–534. [Google Scholar]
  196. Miyazaki K, Miyazaki KW, Doya K. Activation of dorsal raphe serotonin neurons underlies waiting for delayed rewards. J Neurosci 31: 469–479, 2011a. doi: 10.1523/JNEUROSCI.3714-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  197. Miyazaki K, Miyazaki KW, Doya K. The role of serotonin in the regulation of patience and impulsivity. Mol Neurobiol 45: 213–224, 2012. doi: 10.1007/s12035-012-8232-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Miyazaki KW, Miyazaki K, Doya K. Activation of the central serotonergic system in response to delayed but not omitted rewards. Eur J Neurosci 33: 153–160, 2011b. doi: 10.1111/j.1460-9568.2010.07480.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Mlinar B, Montalbano A, Baccini G, Tatini F, Berlinguer Palmini R, Corradetti R. Nonexocytotic serotonin release tonically suppresses serotonergic neuron activity. J Gen Physiol 145: 225–251, 2015. doi: 10.1085/jgp.201411330. [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Monti JM. Serotonin control of sleep-wake behavior. Sleep Med Rev 15: 269–281, 2011. doi: 10.1016/j.smrv.2010.11.003. [DOI] [PubMed] [Google Scholar]
  201. Mooney RD, Huang X, Shi MY, Bennett-Clarke CA, Rhoades RW. Serotonin modulates retinotectal and corticotectal convergence in the superior colliculus. Prog Brain Res 112: 57–69, 1996. doi: 10.1016/S0079-6123(08)63320-8. [DOI] [PubMed] [Google Scholar]
  202. Moret C, Briley M. 5-HT autoreceptors in the regulation of 5-HT release from guinea pig raphe nucleus and hypothalamus. Neuropharmacology 36: 1713–1723, 1997. doi: 10.1016/S0028-3908(97)00145-7. [DOI] [PubMed] [Google Scholar]
  203. Moriya S, Yamashita A, Nishi R, Ikoma Y, Yamanaka A, Kuwaki T. Acute nociceptive stimuli rapidly induce the activity of serotonin and noradrenalin neurons in the brain stem of awake mice. IBRO Rep 7: 1–9, 2019. doi: 10.1016/j.ibror.2019.05.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Moroz LL, Kocot KM, Citarella MR, Dosung S, Norekian TP, Povolotskaya IS, Grigorenko AP, Dailey C, Berezikov E, Buckley KM, Ptitsyn A, Reshetov D, Mukherjee K, Moroz TP, Bobkova Y, Yu F, Kapitonov VV, Jurka J, Bobkov YV, Swore JJ, Girardo DO, Fodor A, Gusev F, Sanford R, Bruders R, Kittler E, Mills CE, Rast JP, Derelle R, Solovyev VV, Kondrashov FA, Swalla BJ, Sweedler JV, Rogaev EI, Halanych KM, Kohn AB. The ctenophore genome and the evolutionary origins of neural systems. Nature 510: 109–114, 2014. doi: 10.1038/nature13400. [DOI] [PMC free article] [PubMed] [Google Scholar]
  205. Mower GD. Comparison of serotonin 5-HT1 receptors and innervation in the visual cortex of normal and dark-reared cats. J Comp Neurol 312: 223–230, 1991. doi: 10.1002/cne.903120205. [DOI] [PubMed] [Google Scholar]
  206. Müller CP, De Souza Silva MA, Huston JP. Double dissociating effects of sensory stimulation and cocaine on serotonin activity in the occipital and temporal cortices. Neuropharmacology 52: 854–862, 2007. doi: 10.1016/j.neuropharm.2006.10.002. [DOI] [PubMed] [Google Scholar]
  207. Muzerelle A, Scotto-Lomassese S, Bernard JF, Soiza-Reilly M, Gaspar P. Conditional anterograde tracing reveals distinct targeting of individual serotonin cell groups (B5-B9) to the forebrain and brainstem. Brain Struct Funct 221: 535–561, 2016. doi: 10.1007/s00429-014-0924-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Nagakura I, Dunn TW, Farah CA, Heppner A, Li FF, Sossin WS. Regulation of protein kinase C Apl II by serotonin receptors in Aplysia. J Neurochem 115: 994–1006, 2010. doi: 10.1111/j.1471-4159.2010.06986.x. [DOI] [PubMed] [Google Scholar]
  209. Nagata A, Nakayama K, Nakamura S, Mochizuki A, Gemba C, Aoki R, Dantsuji M, Maki K, Inoue T. Serotonin1B receptor-mediated presynaptic inhibition of proprioceptive sensory inputs to jaw-closing motoneurons. Brain Res Bull 149: 260–267, 2019. doi: 10.1016/j.brainresbull.2019.05.001. [DOI] [PubMed] [Google Scholar]
  210. Nässel DR. Serotonin and serotonin-immunoreactive neurons in the nervous system of insects. Prog Neurobiol 30: 1–85, 1988. doi: 10.1016/0301-0082(88)90002-0. [DOI] [PubMed] [Google Scholar]
  211. Nässel DR. Substrates for neuronal cotransmission with neuropeptides and small molecule neurotransmitters in Drosophila. Front Cell Neurosci 12: 83, 2018. doi: 10.3389/fncel.2018.00083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  212. Nässel DR, Meyer EP, Klemm N. Mapping and ultrastructure of serotonin-immunoreactive neurons in the optic lobes of three insect species. J Comp Neurol 232: 190–204, 1985. doi: 10.1002/cne.902320205. [DOI] [PubMed] [Google Scholar]
  213. Nässel DR, Ohlsson L, Sivasubramanian P. Postembryonic differentiation of serotonin-immunoreactive neurons in fleshfly optic lobes developing in situ or cultured in vivo without eye discs. J Comp Neurol 255: 327–340, 1987. doi: 10.1002/cne.902550302. [DOI] [PubMed] [Google Scholar]
  214. Nässel DR, Zandawala M. Recent advances in neuropeptide signaling in Drosophila, from genes to physiology and behavior. Prog Neurobiol 179: 101607, 2019. doi: 10.1016/j.pneurobio.2019.02.003. [DOI] [PubMed] [Google Scholar]
  215. Nautiyal KM, Tanaka KF, Barr MM, Tritschler L, Le Dantec Y, David DJ, Gardier AM, Blanco C, Hen R, Ahmari SE. Distinct circuits underlie the effects of 5-HT1B receptors on aggression and impulsivity. Neuron 86: 813–826, 2015. doi: 10.1016/j.neuron.2015.03.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Nautiyal KM, Tritschler L, Ahmari SE, David DJ, Gardier AM, Hen R. A lack of serotonin 1B autoreceptors results in decreased anxiety and depression-related behaviors. Neuropsychopharmacology 41: 2941–2950, 2016. doi: 10.1038/npp.2016.109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  217. Ngai M, Shoue DA, Loh Z, McDowell MA. The pharmacological and functional characterization of the serotonergic system in Anopheles gambiae and Aedes aegypti: influences on flight and blood-feeding behavior. Sci Rep 9: 4421, 2019. doi: 10.1038/s41598-019-38806-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  218. Ni W, Watts SW. 5-hydroxytryptamine in the cardiovascular system: focus on the serotonin transporter (SERT). Clin Exp Pharmacol Physiol 33: 575–583, 2006. doi: 10.1111/j.1440-1681.2006.04410.x. [DOI] [PubMed] [Google Scholar]
  219. Nichols DE, Nichols CD. Serotonin receptors. Chem Rev 108: 1614–1641, 2008. doi: 10.1021/cr078224o. [DOI] [PubMed] [Google Scholar]
  220. Niederkofler V, Asher TE, Okaty BW, Rood BD, Narayan A, Hwa LS, Beck SG, Miczek KA, Dymecki SM. Identification of serotonergic neuronal modules that affect aggressive behavior. Cell Rep 17: 1934–1949, 2016. doi: 10.1016/j.celrep.2016.10.063. [DOI] [PMC free article] [PubMed] [Google Scholar]
  221. Nielsen K, Brask D, Knudsen GM, Aznar S. Immunodetection of the serotonin transporter protein is a more valid marker for serotonergic fibers than serotonin. Synapse 59: 270–276, 2006. doi: 10.1002/syn.20240. [DOI] [PubMed] [Google Scholar]
  222. Norton WH, Folchert A, Bally-Cuif L. Comparative analysis of serotonin receptor (HTR1A/HTR1B families) and transporter (slc6a4a/b) gene expression in the zebrafish brain. J Comp Neurol 511: 521–542, 2008. doi: 10.1002/cne.21831. [DOI] [PubMed] [Google Scholar]
  223. Ogawa SK, Cohen JY, Hwang D, Uchida N, Watabe-Uchida M. Organization of monosynaptic inputs to the serotonin and dopamine neuromodulatory systems. Cell Rep 8: 1105–1118, 2014. doi: 10.1016/j.celrep.2014.06.042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  224. Okaty BW, Commons KG, Dymecki SM. Embracing diversity in the 5-HT neuronal system. Nat Rev Neurosci 20: 397–424, 2019. doi: 10.1038/s41583-019-0151-3. [DOI] [PubMed] [Google Scholar]
  225. Olde B, McCombie WR. Molecular cloning and functional expression of a serotonin receptor from Caenorhabditis elegans. J Mol Neurosci 8: 53–62, 1997. doi: 10.1007/BF02736863. [DOI] [PubMed] [Google Scholar]
  226. Ono JK, McCaman RE. Immunocytochemical localization and direct assays of serotonin-containing neurons in Aplysia. Neuroscience 11: 549–560, 1984. doi: 10.1016/0306-4522(84)90044-7. [DOI] [PubMed] [Google Scholar]
  227. Pakan JM, Lowe SC, Dylda E, Keemink SW, Currie SP, Coutts CA, Rochefort NL. Behavioral-state modulation of inhibition is context-dependent and cell type specific in mouse visual cortex. eLife 5: e14985, 2016. doi: 10.7554/eLife.14985. [DOI] [PMC free article] [PubMed] [Google Scholar]
  228. Pass G. Accessory pulsatile organs: evolutionary innovations in insects. Annu Rev Entomol 45: 495–518, 2000. doi: 10.1146/annurev.ento.45.1.495. [DOI] [PubMed] [Google Scholar]
  229. Paulk AC, Phillips-Portillo J, Dacks AM, Fellous JM, Gronenberg W. The processing of color, motion, and stimulus timing are anatomically segregated in the bumblebee brain. J Neurosci 28: 6319–6332, 2008. doi: 10.1523/JNEUROSCI.1196-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  230. Pazos A, Palacios JM. Quantitative autoradiographic mapping of serotonin receptors in the rat brain. I. Serotonin-1 receptors. Brain Res 346: 205–230, 1985. doi: 10.1016/0006-8993(85)90856-X. [DOI] [PubMed] [Google Scholar]
  231. Peroutka SJ, Howell TA. The molecular evolution of G protein-coupled receptors: focus on 5-hydroxytryptamine receptors. Neuropharmacology 33: 319–324, 1994. doi: 10.1016/0028-3908(94)90060-4. [DOI] [PubMed] [Google Scholar]
  232. Peruzzi D, Dut A. GABA, serotonin and serotonin receptors in the rat inferior colliculus. Brain Res 998: 247–250, 2004. doi: 10.1016/j.brainres.2003.10.059. [DOI] [PubMed] [Google Scholar]
  233. Petzold GC, Hagiwara A, Murthy VN. Serotonergic modulation of odor input to the mammalian olfactory bulb. Nat Neurosci 12: 784–791, 2009. doi: 10.1038/nn.2335. [DOI] [PubMed] [Google Scholar]
  234. Pfeffer CK, Xue M, He M, Huang ZJ, Scanziani M. Inhibition of inhibition in visual cortex: the logic of connections between molecularly distinct interneurons. Nat Neurosci 16: 1068–1076, 2013. doi: 10.1038/nn.3446. [DOI] [PMC free article] [PubMed] [Google Scholar]
  235. Pi HJ, Hangya B, Kvitsiani D, Sanders JI, Huang ZJ, Kepecs A. Cortical interneurons that specialize in disinhibitory control. Nature 503: 521–524, 2013. doi: 10.1038/nature12676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  236. Pickard GE, So KF, Pu M. Dorsal raphe nucleus projecting retinal ganglion cells: why Y cells? Neurosci Biobehav Rev 57: 118–131, 2015. doi: 10.1016/j.neubiorev.2015.08.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  237. Pietrantonio PV, Jagge C, McDowell C. Cloning and expression analysis of a 5HT7-like serotonin receptor cDNA from mosquito Aedes aegypti female excretory and respiratory systems. Insect Mol Biol 10: 357–369, 2001. doi: 10.1046/j.0962-1075.2001.00274.x. [DOI] [PubMed] [Google Scholar]
  238. Piñeyro G, de Montigny C, Weiss M, Blier P. Autoregulatory properties of dorsal raphe 5-HT neurons: possible role of electrotonic coupling and 5-HT1D receptors in the rat brain. Synapse 22: 54–62, 1996. doi:. [DOI] [PubMed] [Google Scholar]
  239. Plassat JL, Amlaiky N, Hen R. Molecular cloning of a mammalian serotonin receptor that activates adenylate cyclase. Mol Pharmacol 44: 229–236, 1993. [PubMed] [Google Scholar]
  240. Polack PO, Friedman J, Golshani P. Cellular mechanisms of brain state-dependent gain modulation in visual cortex. Nat Neurosci 16: 1331–1339, 2013. doi: 10.1038/nn.3464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  241. Pollak Dorocic I, Fürth D, Xuan Y, Johansson Y, Pozzi L, Silberberg G, Carlén M, Meletis K. A whole-brain atlas of inputs to serotonergic neurons of the dorsal and median raphe nuclei. Neuron 83: 663–678, 2014. doi: 10.1016/j.neuron.2014.07.002. [DOI] [PubMed] [Google Scholar]
  242. Pompeiano M, Palacios JM, Mengod G. Distribution and cellular localization of mRNA coding for 5-HT1A receptor in the rat brain: correlation with receptor binding. J Neurosci 12: 440–453, 1992. doi: 10.1523/JNEUROSCI.12-02-00440.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  243. Pootanakit K, Prior KJ, Hunter DD, Brunken WJ. 5-HT2a receptors in the rabbit retina: potential presynaptic modulators. Vis Neurosci 16: 221–230, 1999. doi: 10.1017/S0952523899162035. [DOI] [PubMed] [Google Scholar]
  244. Pritchett DB, Bach AW, Wozny M, Taleb O, Dal Toso R, Shih JC, Seeburg PH. Structure and functional expression of cloned rat serotonin 5HT-2 receptor. EMBO J 7: 4135–4140, 1988. doi: 10.1002/j.1460-2075.1988.tb03308.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  245. Pum ME, Huston JP, De Souza Silva MA, Müller CP. Visual sensory-motor gating by serotonin activation in the medial prefrontal and occipital, but not in the rhinal, cortices in rats. Neuroscience 153: 361–372, 2008. doi: 10.1016/j.neuroscience.2008.02.029. [DOI] [PubMed] [Google Scholar]
  246. Qi YX, Jin M, Ni XY, Ye GY, Lee Y, Huang J. Characterization of three serotonin receptors from the small white butterfly, Pieris rapae. Insect Biochem Mol Biol 87: 107–116, 2017. doi: 10.1016/j.ibmb.2017.06.011. [DOI] [PubMed] [Google Scholar]
  247. Qi YX, Xia RY, Wu YS, Stanley D, Huang J, Ye GY. Larvae of the small white butterfly, Pieris rapae, express a novel serotonin receptor. J Neurochem 131: 767–777, 2014. doi: 10.1111/jnc.12940. [DOI] [PubMed] [Google Scholar]
  248. Rakic P, Goldman-Rakic PS, Gallager D. Quantitative autoradiography of major neurotransmitter receptors in the monkey striate and extrastriate cortex. J Neurosci 8: 3670–3690, 1988. doi: 10.1523/JNEUROSCI.08-10-03670.1988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  249. Rakic P, Lidow MS. Distribution and density of monoamine receptors in the primate visual cortex devoid of retinal input from early embryonic stages. J Neurosci 15: 2561–2574, 1995. doi: 10.1523/JNEUROSCI.15-03-02561.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  250. Ranade SP, Mainen ZF. Transient firing of dorsal raphe neurons encodes diverse and specific sensory, motor, and reward events. J Neurophysiol 102: 3026–3037, 2009. doi: 10.1152/jn.00507.2009. [DOI] [PubMed] [Google Scholar]
  251. Ranganathan R, Cannon SC, Horvitz HR. MOD-1 is a serotonin-gated chloride channel that modulates locomotory behaviour in C. elegans. Nature 408: 470–475, 2000. doi: 10.1038/35044083. [DOI] [PubMed] [Google Scholar]
  252. Rasmussen K, Heym J, Jacobs BL. Activity of serotonin-containing neurons in nucleus centralis superior of freely moving cats. Exp Neurol 83: 302–317, 1984. doi: 10.1016/S0014-4886(84)90100-6. [DOI] [PubMed] [Google Scholar]
  253. Rasmussen K, Strecker RE, Jacobs BL. Single unit response of noradrenergic, serotonergic and dopaminergic neurons in freely moving cats to simple sensory stimuli. Brain Res 369: 336–340, 1986. doi: 10.1016/0006-8993(86)90546-9. [DOI] [PubMed] [Google Scholar]
  254. Ren J, Friedmann D, Xiong J, Liu CD, Ferguson BR, Weerakkody T, DeLoach KE, Ran C, Pun A, Sun Y, Weissbourd B, Neve RL, Huguenard J, Horowitz MA, Luo L. Anatomically defined and functionally distinct dorsal raphe serotonin sub-systems. Cell 175: 472–487.e20, 2018. doi: 10.1016/j.cell.2018.07.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  255. Ren J, Isakova A, Friedmann D, Zeng J, Grutzner SM, Pun A, Zhao GQ, Kolluru SS, Wang R, Lin R, Li P, Li A, Raymond JL, Luo Q, Luo M, Quake SR, Luo L. Single-cell transcriptomes and whole-brain projections of serotonin neurons in the mouse dorsal and median raphe nuclei. eLife 8: e49424, 2019. doi: 10.7554/eLife.49424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  256. Riad M, Kobert A, Descarries L, Boye S, Rompré PP, Lacaille JC. Chronic fluoxetine rescues changes in plasma membrane density of 5-HT1A autoreceptors and serotonin transporters in the olfactory bulbectomy rodent model of depression. Neuroscience 356: 78–88, 2017. doi: 10.1016/j.neuroscience.2017.05.021. [DOI] [PubMed] [Google Scholar]
  257. Ribeiro P, El-Shehabi F, Patocka N. Classical transmitters and their receptors in flatworms. Parasitology 131, Suppl: S19–S40, 2005. doi: 10.1017/S0031182005008565. [DOI] [PubMed] [Google Scholar]
  258. Röser C, Jordan N, Balfanz S, Baumann A, Walz B, Baumann O, Blenau W. Molecular and pharmacological characterization of serotonin 5-HT2α and 5-HT7 receptors in the salivary glands of the blowfly Calliphora vicina. PLoS One 7: e49459, 2012. doi: 10.1371/journal.pone.0049459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  259. Roy B, Singh AP, Shetty C, Chaudhary V, North A, Landgraf M, Vijayraghavan K, Rodrigues V. Metamorphosis of an identified serotonergic neuron in the Drosophila olfactory system. Neural Dev 2: 20, 2007. doi: 10.1186/1749-8104-2-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  260. Ruat M, Traiffort E, Arrang JM, Tardivel-Lacombe J, Diaz J, Leurs R, Schwartz JC. A novel rat serotonin (5-HT6) receptor: molecular cloning, localization and stimulation of cAMP accumulation. Biochem Biophys Res Commun 193: 268–276, 1993a. doi: 10.1006/bbrc.1993.1619. [DOI] [PubMed] [Google Scholar]
  261. Ruat M, Traiffort E, Leurs R, Tardivel-Lacombe J, Diaz J, Arrang JM, Schwartz JC. Molecular cloning, characterization, and localization of a high-affinity serotonin receptor (5-HT7) activating cAMP formation. Proc Natl Acad Sci USA 90: 8547–8551, 1993b. doi: 10.1073/pnas.90.18.8547. [DOI] [PMC free article] [PubMed] [Google Scholar]
  262. Rudy B, Fishell G, Lee S, Hjerling-Leffler J. Three groups of interneurons account for nearly 100% of neocortical GABAergic neurons. Dev Neurobiol 71: 45–61, 2011. doi: 10.1002/dneu.20853. [DOI] [PMC free article] [PubMed] [Google Scholar]
  263. Sampson MM, Myers-Gschweng KM, Hardcastle BJ, Bonanno SL, Sizemore TR, Arnold RC, Gao F, Dacks AM, Frye MA, Krantz DE. Serotonergic modulation of a visual microcircuit in Drosophila melanogaster. bioRxiv 619759, 2019. doi: 10.1101/619759 [DOI] [PMC free article] [PubMed]
  264. Saudou F, Boschert U, Amlaiky N, Plassat JL, Hen R. A family of Drosophila serotonin receptors with distinct intracellular signalling properties and expression patterns. EMBO J 11: 7–17, 1992. doi: 10.1002/j.1460-2075.1992.tb05021.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  265. Schlegel P, Costa M, Jefferis GS. Learning from connectomics on the fly. Curr Opin Insect Sci 24: 96–105, 2017. doi: 10.1016/j.cois.2017.09.011. [DOI] [PubMed] [Google Scholar]
  266. Schlenstedt J, Balfanz S, Baumann A, Blenau W. Am5-HT7: molecular and pharmacological characterization of the first serotonin receptor of the honeybee (Apis mellifera). J Neurochem 98: 1985–1998, 2006. doi: 10.1111/j.1471-4159.2006.04012.x. [DOI] [PubMed] [Google Scholar]
  267. Schmuck K, Ullmer C, Engels P, Lübbert H. Cloning and functional characterization of the human 5-HT2B serotonin receptor. FEBS Lett 342: 85–90, 1994. doi: 10.1016/0014-5793(94)80590-3. [DOI] [PubMed] [Google Scholar]
  268. Schofield BR, Hurley LM. Circuits for modulation of auditory function. In: The Mammalian Auditory Pathways: Synaptic Organization and Microcircuits, edited by Oliver DL, Cant NB, Fay RR, Popper AN. Cham, Switzerland: Springer, 2018. p. 235–267. [Google Scholar]
  269. Schulze C, Neves RC, Schmidt-Rhaesa A. Comparative immunohistochemical investigation on the nervous system of two species of Arthrotardigrada (Heterotardigrada, Tardigrada). Zool Anz 253: 225–235, 2014. doi: 10.1016/j.jcz.2013.11.001. [DOI] [Google Scholar]
  270. Schürmann FW, Klemm N. Serotonin-immunoreactive neurons in the brain of the honeybee. J Comp Neurol 225: 570–580, 1984. doi: 10.1002/cne.902250407. [DOI] [PubMed] [Google Scholar]
  271. Sengupta A, Bocchio M, Bannerman DM, Sharp T, Capogna M. Control of amygdala circuits by 5-HT neurons via 5-HT and glutamate cotransmission. J Neurosci 37: 1785–1796, 2017. doi: 10.1523/JNEUROSCI.2238-16.2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  272. Seo C, Guru A, Jin M, Ito B, Sleezer BJ, Ho YY, Wang E, Boada C, Krupa NA, Kullakanda DS, Shen CX, Warden MR. Intense threat switches dorsal raphe serotonin neurons to a paradoxical operational mode. Science 363: 538–542, 2019. doi: 10.1126/science.aau8722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  273. Settembrini BP, Villar MJ. Distribution of serotonin in the central nervous system of the blood-feeding heteropteran, Triatoma infestans (Heteroptera: Reduviidae). J Morphol 260: 21–32, 2004. doi: 10.1002/jmor.10211. [DOI] [PubMed] [Google Scholar]
  274. Shang Y, Claridge-Chang A, Sjulson L, Pypaert M, Miesenböck G. Excitatory local circuits and their implications for olfactory processing in the fly antennal lobe. Cell 128: 601–612, 2007. doi: 10.1016/j.cell.2006.12.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
  275. Shao J, Zhang X, Cheng H, Yue X, Zou W, Kang L. Serotonergic neuron ADF modulates avoidance behaviors by inhibiting sensory neurons in C. elegans. Pflugers Arch 471: 357–363, 2019. doi: 10.1007/s00424-018-2202-4. [DOI] [PubMed] [Google Scholar]
  276. Shao QM, Fouda MM, Takeda M. Serotonin- and two putative serotonin receptors-like immunohistochemical reactivities in the ground crickets Dianemobius nigrofasciatus and Allonemobius allardi. J Insect Physiol 56: 1576–1586, 2010. doi: 10.1016/j.jinsphys.2010.05.015. [DOI] [PubMed] [Google Scholar]
  277. Shen Y, Monsma FJ Jr, Metcalf MA, Jose PA, Hamblin MW, Sibley DR. Molecular cloning and expression of a 5-hydroxytryptamine7 serotonin receptor subtype. J Biol Chem 268: 18200–18204, 1993. [PubMed] [Google Scholar]
  278. Singh AP, Das RN, Rao G, Aggarwal A, Diegelmann S, Evers JF, Karandikar H, Landgraf M, Rodrigues V, Vijayraghavan K. Sensory neuron-derived eph regulates glomerular arbors and modulatory function of a central serotonergic neuron. PLoS Genet 9: e1003452, 2013. [Erratum in PLoS Genet 13: e1007083, 2017.] doi: 10.1371/journal.pgen.1003452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  279. Sizemore TR, Dacks AM. Serotonergic modulation differentially targets distinct network elements within the antennal lobe of Drosophila melanogaster. Sci Rep 6: 37119, 2016. doi: 10.1038/srep37119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  280. Skangiel-Kramska J, Kossut M. 5-HT1 receptors in the structures of visual pathway of normal and monocularly deprived kittens. Acta Neurobiol Exp (Wars) 52: 71–81, 1992. [PubMed] [Google Scholar]
  281. Sommer C. Serotonin in pain and analgesia: actions in the periphery. Mol Neurobiol 30: 117–125, 2004. doi: 10.1385/MN:30:2:117. [DOI] [PubMed] [Google Scholar]
  282. Spaethling JM, Piel D, Dueck H, Buckley PT, Morris JF, Fisher SA, Lee J, Sul JY, Kim J, Bartfai T, Beck SG, Eberwine JH. Serotonergic neuron regulation informed by in vivo single-cell transcriptomics. FASEB J 28: 771–780, 2014. doi: 10.1096/fj.13-240267. [DOI] [PMC free article] [PubMed] [Google Scholar]
  283. Spitzer N, Antonsen BL, Edwards DH. Immunocytochemical mapping and quantification of expression of a putative type 1 serotonin receptor in the crayfish nervous system. J Comp Neurol 484: 261–282, 2005. doi: 10.1002/cne.20456. [DOI] [PubMed] [Google Scholar]
  284. Spitzer N, Cymbalyuk G, Zhang H, Edwards DH, Baro DJ. Serotonin transduction cascades mediate variable changes in pyloric network cycle frequency in response to the same modulatory challenge. J Neurophysiol 99: 2844–2863, 2008a. doi: 10.1152/jn.00986.2007. [DOI] [PubMed] [Google Scholar]
  285. Spitzer N, Edwards DH, Baro DJ. Conservation of structure, signaling and pharmacology between two serotonin receptor subtypes from decapod crustaceans, Panulirus interruptus and Procambarus clarkii. J Exp Biol 211: 92–105, 2008b. doi: 10.1242/jeb.012450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  286. Spörhase-Eichmann U, Gras H, Schurmann FW. Patterns of serotonin-immunoreactive neurons in the central nervous system of the earthworm Lumbricus terrestris L. 1. Ganglia of the ventral nerve cord. Cell Tissue Res 249: 601–614, 1987a. doi: 10.1007/BF00217332. [DOI] [Google Scholar]
  287. Spörhase-Eichmann U, Gras H, Schurmann FW. Patterns of serotonin immunoreactive neurons in the central nervous system of the earthworm Lumbricus terrestris L. 2. Rostral and caudal ganglia. Cell Tissue Res 249: 625–632, 1987b. doi: 10.1007/BF00217334. [DOI] [Google Scholar]
  288. Stam NJ, Roesink C, Dijcks F, Garritsen A, van Herpen A, Olijve W. Human serotonin 5-HT7 receptor: cloning and pharmacological characterisation of two receptor variants. FEBS Lett 413: 489–494, 1997. doi: 10.1016/S0014-5793(97)00964-2. [DOI] [PubMed] [Google Scholar]
  289. Stam NJ, Van Huizen F, Van Alebeek C, Dijkema R, Tonnaer JA, Olijve W. Genomic organization, coding sequence and functional expression of human 5-HT2 and 5-HT1A receptor genes. Eur J Pharmacol 227: 153–162, 1992. doi: 10.1016/0922-4106(92)90123-D. [DOI] [PubMed] [Google Scholar]
  290. Stam NJ, Vanderheyden P, van Alebeek C, Klomp J, de Boer T, van Delft AM, Olijve W. Genomic organisation and functional expression of the gene encoding the human serotonin 5-HT2C receptor. Eur J Pharmacol 269: 339–348, 1994. doi: 10.1016/0922-4106(94)90042-6. [DOI] [PubMed] [Google Scholar]
  291. Stark H, Scheich H. Dopaminergic and serotonergic neurotransmission systems are differentially involved in auditory cortex learning: a long-term microdialysis study of metabolites. J Neurochem 68: 691–697, 1997. doi: 10.1046/j.1471-4159.1997.68020691.x. [DOI] [PubMed] [Google Scholar]
  292. Starkey SJ, Skingle M. 5-HT1D as well as 5-HT1A autoreceptors modulate 5-HT release in the guinea-pig dorsal raphé nucleus. Neuropharmacology 33: 393–402, 1994. doi: 10.1016/0028-3908(94)90069-8. [DOI] [PubMed] [Google Scholar]
  293. Steinbusch HW. Distribution of serotonin-immunoreactivity in the central nervous system of the rat-cell bodies and terminals. Neuroscience 6: 557–618, 1981. doi: 10.1016/0306-4522(81)90146-9. [DOI] [PubMed] [Google Scholar]
  294. Steinbusch HW, Nieuwenhuys R. Localization of serotonin-like immunoreactivity in the central nervous system and pituitary of the rat, with special references to the innervation of the hypothalamus. Adv Exp Med Biol 133: 7–35, 1981. doi: 10.1007/978-1-4684-3860-4_1. [DOI] [PubMed] [Google Scholar]
  295. Sukumar V, Liu H, Meisner S, French AS, Torkkeli PH. Multiple biogenic amine receptor types modulate spider, Cupiennius salei, mechanosensory neurons. Front Physiol 9: 857, 2018. doi: 10.3389/fphys.2018.00857. [DOI] [PMC free article] [PubMed] [Google Scholar]
  296. Sun XJ, Tolbert LP, Hildebrand JG. Ramification pattern and ultrastructural characteristics of the serotonin-immunoreactive neuron in the antennal lobe of the moth Manduca sexta: a laser scanning confocal and electron microscopic study. J Comp Neurol 338: 5–16, 1993. doi: 10.1002/cne.903380103. [DOI] [PubMed] [Google Scholar]
  297. Suwa B, Bock N, Preusse S, Rothenberger A, Manzke T. Distribution of serotonin 4(a) receptors in the juvenile rat brain and spinal cord. J Chem Neuroanat 55: 67–77, 2014. doi: 10.1016/j.jchemneu.2013.12.004. [DOI] [PubMed] [Google Scholar]
  298. Suzuki Y, Kiyokage E, Sohn J, Hioki H, Toida K. Structural basis for serotonergic regulation of neural circuits in the mouse olfactory bulb. J Comp Neurol 523: 262–280, 2015. doi: 10.1002/cne.23680. [DOI] [PubMed] [Google Scholar]
  299. Suzuki Y, Schenk JE, Tan H, Gaudry Q. A population of interneurons signals changes in the basal concentration of serotonin and mediates gain control in the Drosophila antennal lobe. Curr Biol 30: 1110–1118.e4, 2020. doi: 10.1016/j.cub.2020.01.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  300. Sze JY, Victor M, Loer C, Shi Y, Ruvkun G. Food and metabolic signalling defects in a Caenorhabditis elegans serotonin-synthesis mutant. Nature 403: 560–564, 2000. doi: 10.1038/35000609. [DOI] [PubMed] [Google Scholar]
  301. Tadros SF, D’Souza M, Zettel ML, Zhu X, Lynch-Erhardt M, Frisina RD. Serotonin 2B receptor: upregulated with age and hearing loss in mouse auditory system. Neurobiol Aging 28: 1112–1123, 2007. doi: 10.1016/j.neurobiolaging.2006.05.021. [DOI] [PubMed] [Google Scholar]
  302. Takesian AE, Bogart LJ, Lichtman JW, Hensch TK. Inhibitory circuit gating of auditory critical-period plasticity. Nat Neurosci 21: 218–227, 2018. [Erratum in Nat Neurosci 21: 1495, 2018.] doi: 10.1038/s41593-017-0064-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  303. Takeuchi Y, McLean JH, Hopkins DA. Reciprocal connections between the amygdala and parabrachial nuclei: ultrastructural demonstration by degeneration and axonal transport of horseradish peroxidase in the cat. Brain Res 239: 583–588, 1982. doi: 10.1016/0006-8993(82)90532-7. [DOI] [PubMed] [Google Scholar]
  304. Tamvacakis AN, Senatore A, Katz PS. Identification of genes related to learning and memory in the brain transcriptome of the mollusc, Hermissenda crassicornis. Learn Mem 22: 617–621, 2015. doi: 10.1101/lm.038158.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  305. Tamvacakis AN, Senatore A, Katz PS. Single neuron serotonin receptor subtype gene expression correlates with behaviour within and across three molluscan species. Proc Biol Sci 285: 20180791, 2018. doi: 10.1098/rspb.2018.0791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  306. Tanaka M, Watanabe Y. RNA editing of serotonin 2C receptor and alcohol intake. Front Neurosci 13: 1390, 2020. doi: 10.3389/fnins.2019.01390. [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Tang QB, Song WW, Chang YJ, Xie GY, Chen WB, Zhao XC. Distribution of serotonin-immunoreactive neurons in the brain and gnathal ganglion of caterpillar Helicoverpa armigera. Front Neuroanat 13: 56, 2019. doi: 10.3389/fnana.2019.00056. [DOI] [PMC free article] [PubMed] [Google Scholar]
  308. Tang ZQ, Trussell LO. Serotonergic regulation of excitability of principal cells of the dorsal cochlear nucleus. J Neurosci 35: 4540–4551, 2015. doi: 10.1523/JNEUROSCI.4825-14.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  309. Tang ZQ, Trussell LO. Serotonergic modulation of sensory representation in a central multisensory circuit is pathway specific. Cell Rep 20: 1844–1854, 2017. doi: 10.1016/j.celrep.2017.07.079. [DOI] [PMC free article] [PubMed] [Google Scholar]
  310. Tecott LH, Maricq AV, Julius D. Nervous system distribution of the serotonin 5-HT3 receptor mRNA. Proc Natl Acad Sci USA 90: 1430–1434, 1993. doi: 10.1073/pnas.90.4.1430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  311. Templin JS, Bang SJ, Soiza-Reilly M, Berde CB, Commons KG. Patterned expression of ion channel genes in mouse dorsal raphe nucleus determined with the Allen Mouse Brain Atlas. Brain Res 1457: 1–12, 2012. doi: 10.1016/j.brainres.2012.03.066. [DOI] [PMC free article] [PubMed] [Google Scholar]
  312. Thamm M, Balfanz S, Scheiner R, Baumann A, Blenau W. Characterization of the 5-HT1A receptor of the honeybee (Apis mellifera) and involvement of serotonin in phototactic behavior. Cell Mol Life Sci 67: 2467–2479, 2010. doi: 10.1007/s00018-010-0350-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  313. Thamm M, Rolke D, Jordan N, Balfanz S, Schiffer C, Baumann A, Blenau W. Function and distribution of 5-HT2 receptors in the honeybee (Apis mellifera). PLoS One 8: e82407, 2013. doi: 10.1371/journal.pone.0082407. [DOI] [PMC free article] [PubMed] [Google Scholar]
  314. Thomas EA, Matli JR, Hu JL, Carson MJ, Sutcliffe JG. Pertussis toxin treatment prevents 5-HT5a receptor-mediated inhibition of cyclic AMP accumulation in rat C6 glioma cells. J Neurosci Res 61: 75–81, 2000. doi:. [DOI] [PubMed] [Google Scholar]
  315. Thompson GC, Thompson AM, Garrett KM, Britton BH. Serotonin and serotonin receptors in the central auditory system. Otolaryngol Head Neck Surg 110: 93–102, 1994a. doi: 10.1177/019459989411000111. [DOI] [PubMed] [Google Scholar]
  316. Thompson KS, Zeidler MP, Bacon JP. Comparative anatomy of serotonin-like immunoreactive neurons in isopods: putative homologues in several species. J Comp Neurol 347: 553–569, 1994b. doi: 10.1002/cne.903470407. [DOI] [PubMed] [Google Scholar]
  317. To ZP, Bonhaus DW, Eglen RM, Jakeman LB. Characterization and distribution of putative 5-ht7 receptors in guinea-pig brain. Br J Pharmacol 115: 107–116, 1995. doi: 10.1111/j.1476-5381.1995.tb16327.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  318. Törk I. Anatomy of the serotonergic system. Ann NY Acad Sci 600: 9–34, 1990. doi: 10.1111/j.1749-6632.1990.tb16870.x. [DOI] [PubMed] [Google Scholar]
  319. Torres-Escalante JL, Barral JA, Ibarra-Villa MD, Pérez-Burgos A, Góngora-Alfaro JL, Pineda JC. 5-HT1A, 5-HT2, and GABAB receptors interact to modulate neurotransmitter release probability in layer 2/3 somatosensory rat cortex as evaluated by the paired pulse protocol. J Neurosci Res 78: 268–278, 2004. doi: 10.1002/jnr.20247. [DOI] [PubMed] [Google Scholar]
  320. Trakhtenberg EF, Pita-Thomas W, Fernandez SG, Patel KH, Venugopalan P, Shechter JM, Morkin MI, Galvao J, Liu X, Dombrowski SM, Goldberg JL. Serotonin receptor 2C regulates neurite growth and is necessary for normal retinal processing of visual information. Dev Neurobiol 77: 419–437, 2017. doi: 10.1002/dneu.22391. [DOI] [PubMed] [Google Scholar]
  321. Troppmann B, Balfanz S, Baumann A, Blenau W. Inverse agonist and neutral antagonist actions of synthetic compounds at an insect 5-HT1 receptor. Br J Pharmacol 159: 1450–1462, 2010. doi: 10.1111/j.1476-5381.2010.00638.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  322. Trulson ME, Jacobs BL. Raphe unit activity in freely moving cats: correlation with level of behavioral arousal. Brain Res 163: 135–150, 1979. doi: 10.1016/0006-8993(79)90157-4. [DOI] [PubMed] [Google Scholar]
  323. Vallés AM, White K. Serotonin-containing neurons in Drosophila melanogaster: development and distribution. J Comp Neurol 268: 414–428, 1988. doi: 10.1002/cne.902680310. [DOI] [PubMed] [Google Scholar]
  324. Vázquez-Acevedo N, Reyes-Colón D, Ruíz-Rodríguez EA, Rivera NM, Rosenthal J, Kohn AB, Moroz LL, Sosa MA. Cloning and immunoreactivity of the 5-HT 1Mac and 5-HT 2Mac receptors in the central nervous system of the freshwater prawn Macrobrachium rosenbergii. J Comp Neurol 513: 399–416, 2009. doi: 10.1002/cne.21979. [DOI] [PMC free article] [PubMed] [Google Scholar]
  325. Vertes RP, Crane AM. Distribution, quantification, and morphological characteristics of serotonin-immunoreactive cells of the supralemniscal nucleus (B9) and pontomesencephalic reticular formation in the rat. J Comp Neurol 378: 411–424, 1997. doi:. [DOI] [PubMed] [Google Scholar]
  326. Viguier F, Michot B, Hamon M, Bourgoin S. Multiple roles of serotonin in pain control mechanisms–implications of 5-HT7 and other 5-HT receptor types. Eur J Pharmacol 716: 8–16, 2013. doi: 10.1016/j.ejphar.2013.01.074. [DOI] [PubMed] [Google Scholar]
  327. Vleugels R, Lenaerts C, Baumann A, Vanden Broeck J, Verlinden H. Pharmacological characterization of a 5-HT1-type serotonin receptor in the red flour beetle, Tribolium castaneum. PLoS One 8: e65052, 2013. doi: 10.1371/journal.pone.0065052. [DOI] [PMC free article] [PubMed] [Google Scholar]
  328. Vleugels R, Lenaerts C, Vanden Broeck J, Verlinden H. Signalling properties and pharmacology of a 5-HT7 -type serotonin receptor from Tribolium castaneum. Insect Mol Biol 23: 230–243, 2014. doi: 10.1111/imb.12076. [DOI] [PubMed] [Google Scholar]
  329. Vleugels R, Verlinden H, Vanden Broeck J. Serotonin, serotonin receptors and their actions in insects. Neurotransmitter (Houst) 2: e314, 2015. doi: 10.14800/nt.314. [DOI] [Google Scholar]
  330. Voigt JP, Fink H. Serotonin controlling feeding and satiety. Behav Brain Res 277: 14–31, 2015. doi: 10.1016/j.bbr.2014.08.065. [DOI] [PubMed] [Google Scholar]
  331. Voigt MM, Laurie DJ, Seeburg PH, Bach A. Molecular cloning and characterization of a rat brain cDNA encoding a 5-hydroxytryptamine1B receptor. EMBO J 10: 4017–4023, 1991. doi: 10.1002/j.1460-2075.1991.tb04977.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  332. Vyas P, Wu JS, Jimenez A, Glowatzki E, Fuchs PA. Characterization of transgenic mouse lines for labeling type I and type II afferent neurons in the cochlea. Sci Rep 9: 5549, 2019. doi: 10.1038/s41598-019-41770-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  333. Wang T, Yang Z, Zhou N, Sun L, Lv Z, Wu C. Identification and functional characterisation of 5-HT4 receptor in sea cucumber Apostichopus japonicus (Selenka). Sci Rep 7: 40247, 2017. doi: 10.1038/srep40247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  334. Watakabe A, Komatsu Y, Sadakane O, Shimegi S, Takahata T, Higo N, Tochitani S, Hashikawa T, Naito T, Osaki H, Sakamoto H, Okamoto M, Ishikawa A, Hara S, Akasaki T, Sato H, Yamamori T. Enriched expression of serotonin 1B and 2A receptor genes in macaque visual cortex and their bidirectional modulatory effects on neuronal responses. Cereb Cortex 19: 1915–1928, 2009. doi: 10.1093/cercor/bhn219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  335. Watanabe H, Shimohigashi M, Yokohari F. Serotonin-immunoreactive sensory neurons in the antenna of the cockroach Periplaneta americana. J Comp Neurol 522: 414–434, 2014. doi: 10.1002/cne.23419. [DOI] [PubMed] [Google Scholar]
  336. Waterhouse BD, Azizi SA, Burne RA, Woodward DJ. Modulation of rat cortical area 17 neuronal responses to moving visual stimuli during norepinephrine and serotonin microiontophoresis. Brain Res 514: 276–292, 1990. doi: 10.1016/0006-8993(90)91422-D. [DOI] [PubMed] [Google Scholar]
  337. Waterhouse BD, Devilbiss D, Seiple S, Markowitz R. Sensorimotor-related discharge of simultaneously recorded, single neurons in the dorsal raphe nucleus of the awake, unrestrained rat. Brain Res 1000: 183–191, 2004. doi: 10.1016/j.brainres.2003.11.030. [DOI] [PubMed] [Google Scholar]
  338. Waterhouse BD, Moises HC, Woodward DJ. Interaction of serotonin with somatosensory cortical neuronal responses to afferent synaptic inputs and putative neurotransmitters. Brain Res Bull 17: 507–518, 1986. doi: 10.1016/0361-9230(86)90218-2. [DOI] [PubMed] [Google Scholar]
  339. Weinshank RL, Zgombick JM, Macchi MJ, Branchek TA, Hartig PR. Human serotonin 1D receptor is encoded by a subfamily of two distinct genes: 5-HT1D alpha and 5-HT1D beta. Proc Natl Acad Sci USA 89: 3630–3634, 1992. doi: 10.1073/pnas.89.8.3630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  340. Weydert A, Cloez-Tayarani I, Fillion MP, Simon-Chazottes D, Guenet JL, Fillion G. Molecular cloning of two partial serotonin 5-HT1D receptor sequences in mouse and one in guinea pig. C R Acad Sci III 314: 429–435, 1992. [PubMed] [Google Scholar]
  341. White JG, Southgate E, Thomson JN, Brenner S. The structure of the nervous system of the nematode Caenorhabditis elegans. Philos Trans R Soc Lond B Biol Sci 314: 1–340, 1986. doi: 10.1098/rstb.1986.0056. [DOI] [PubMed] [Google Scholar]
  342. Wisden W, Parker EM, Mahle CD, Grisel DA, Nowak HP, Yocca FD, Felder CC, Seeburg PH, Voigt MM. Cloning and characterization of the rat 5-HT5B receptor. Evidence that the 5-HT5B receptor couples to a G protein in mammalian cell membranes. FEBS Lett 333: 25–31, 1993. doi: 10.1016/0014-5793(93)80368-5. [DOI] [PubMed] [Google Scholar]
  343. Witz P, Amlaiky N, Plassat JL, Maroteaux L, Borrelli E, Hen R. Cloning and characterization of a Drosophila serotonin receptor that activates adenylate cyclase. Proc Natl Acad Sci USA 87: 8940–8944, 1990. doi: 10.1073/pnas.87.22.8940. [DOI] [PMC free article] [PubMed] [Google Scholar]
  344. Wright DE, Seroogy KB, Lundgren KH, Davis BM, Jennes L. Comparative localization of serotonin1A, 1C, and 2 receptor subtype mRNAs in rat brain. J Comp Neurol 351: 357–373, 1995. doi: 10.1002/cne.903510304. [DOI] [PubMed] [Google Scholar]
  345. Xiang Z, Prince DA. Heterogeneous actions of serotonin on interneurons in rat visual cortex. J Neurophysiol 89: 1278–1287, 2003. doi: 10.1152/jn.00533.2002. [DOI] [PubMed] [Google Scholar]
  346. Xiong R, Zhao W, Chen X, Li T, Li H, Li Y, Shen W, Chen P. Pharmacological characterization of the 5-HT1A receptor of Bombyx mori and its role in locomotion. Comp Biochem Physiol A Mol Integr Physiol 231: 56–65, 2019. doi: 10.1016/j.cbpa.2019.01.020. [DOI] [PubMed] [Google Scholar]
  347. Yaksi E, Wilson RI. Electrical coupling between olfactory glomeruli. Neuron 67: 1034–1047, 2010. doi: 10.1016/j.neuron.2010.08.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  348. You IJ, Wright SR, Garcia-Garcia AL, Tapper AR, Gardner PD, Koob GF, David Leonardo E, Bohn LM, Wee S. 5-HT1A autoreceptors in the dorsal raphe nucleus convey vulnerability to compulsive cocaine seeking. Neuropsychopharmacology 41: 1210–1222, 2016. doi: 10.1038/npp.2015.268. [DOI] [PMC free article] [PubMed] [Google Scholar]
  349. Yu L, Nguyen H, Le H, Bloem LJ, Kozak CA, Hoffman BJ, Snutch TP, Lester HA, Davidson N, Lübbert H. The mouse 5-HT1C receptor contains eight hydrophobic domains and is X-linked. Brain Res Mol Brain Res 11: 143–149, 1991. doi: 10.1016/0169-328X(91)90116-F. [DOI] [PubMed] [Google Scholar]
  350. Zattara EE, Bely AE. Fine taxonomic sampling of nervous systems within Naididae (Annelida: Clitellata) reveals evolutionary lability and revised homologies of annelid neural components. Front Zool 12: 8, 2015. doi: 10.1186/s12983-015-0100-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  351. Zhang X, Coates KE, Dacks AM, Gunay C, Lauritzen JS, Li F, Calle-Schuler SA, Bock D, Gaudry Q. Regulation of modulatory cell activity across olfactory structures in Drosophila melanogaster. bioRxiv 522177, 2019. doi: 10.1101/522177 [DOI]
  352. Zhang X, Gaudry Q. Functional integration of a serotonergic neuron in the Drosophila antennal lobe. eLife 5: e16836, 2016. doi: 10.7554/eLife.16836. [DOI] [PMC free article] [PubMed] [Google Scholar]
  353. Zhao XC, Berg BG. Morphological and physiological characteristics of the serotonin-immunoreactive neuron in the antennal lobe of the male oriental tobacco budworm, Helicoverpa assulta. Chem Senses 34: 363–372, 2009. doi: 10.1093/chemse/bjp013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  354. Zheng Z, Lauritzen JS, Perlman E, Robinson CG, Nichols M, Milkie D, Torrens O, Price J, Fisher CB, Sharifi N, Calle-Schuler SA, Kmecova L, Ali IJ, Karsh B, Trautman ET, Bogovic JA, Hanslovsky P, Jefferis GS, Kazhdan M, Khairy K, Saalfeld S, Fetter RD, Bock DD. A complete electron microscopy volume of the brain of adult Drosophila melanogaster. Cell 174: 730–743.e22, 2018. doi: 10.1016/j.cell.2018.06.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  355. Zhou X, Li G, Zhang S, Wu J. 5-HT1A receptor agonist promotes retinal ganglion cell function by inhibiting OFF-type presynaptic glutamatergic activity in a chronic glaucoma model. Front Cell Neurosci 13: 167, 2019. doi: 10.3389/fncel.2019.00167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  356. Zhukovskaya MI, Polyanovsky AD. Biogenic amines in insect antennae. Front Syst Neurosci 11: 45, 2017. doi: 10.3389/fnsys.2017.00045. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Journal of Neurophysiology are provided here courtesy of American Physiological Society

RESOURCES