Next Article in Journal
IQGAP1 Regulates Actin Polymerization and Contributes to Bleomycin-Induced Lung Fibrosis
Previous Article in Journal
The CpG Island Methylator Phenotype Status in Synchronous and Solitary Primary Colorectal Cancers: Prognosis and Effective Therapeutic Drug Prediction
Previous Article in Special Issue
Serotonin Transporter Activity in Mouse Oocytes Is a Positive Indicator of Follicular Growth and Oocyte Maturity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Serotonin 4 Receptor Subtype: A Target of Particular Interest, Especially for Brain Disorders

by
Véronique Sgambato
1,2
1
Institut des Sciences Cognitives Marc Jeannerod (ISCMJ), Unité Mixte de Recherche 5229 du Centre National de la Recherche Scientifique (CNRS), 69675 Bron, France
2
UFR Biosciences, Université de Lyon 1, 69100 Villeurbanne, France
Int. J. Mol. Sci. 2024, 25(10), 5245; https://doi.org/10.3390/ijms25105245
Submission received: 15 March 2024 / Revised: 22 April 2024 / Accepted: 7 May 2024 / Published: 11 May 2024
(This article belongs to the Special Issue Serotonin in Health and Diseases)

Abstract

:
In recent years, particular attention has been paid to the serotonin 4 receptor, which is well expressed in the brain, but also peripherally in various organs. The cerebral distribution of this receptor is well conserved across species, with high densities in the basal ganglia, where they are expressed by GABAergic neurons. The 5-HT4 receptor is also present in the cerebral cortex, hippocampus, and amygdala, where they are carried by glutamatergic or cholinergic neurons. Outside the central nervous system, the 5-HT4 receptor is notably expressed in the gastrointestinal tract. The wide distribution of the 5-HT4 receptor undoubtedly contributes to its involvement in a plethora of functions. In addition, the modulation of this receptor influences the release of serotonin, but also the release of other neurotransmitters such as acetylcholine and dopamine. This is a considerable asset, as the modulation of the 5-HT4 receptor can therefore play a direct or indirect beneficial role in various disorders. One of the main advantages of this receptor is that it mediates a much faster antidepressant and anxiolytic action than classical selective serotonin reuptake inhibitors. Another major benefit of the 5-HT4 receptor is that its activation enhances cognitive performance, probably via the release of acetylcholine. The expression of the 5-HT4 receptor is also altered in various eating disorders, and its activation by the 5-HT4 agonist negatively regulates food intake. Additionally, although the cerebral expression of this receptor is modified in certain movement-related disorders, it is still yet to be determined whether this receptor plays a key role in their pathophysiology. Finally, there is no longer any need to demonstrate the value of 5-HT4 receptor agonists in the pharmacological management of gastrointestinal disorders.

1. Introduction

Serotonin (5-HT, 5-hydroxytryptamine) is a monoamine synthesized from an essential amino acid, L-tryptophan. L-tryptophan can cross the blood–brain barrier and be converted by the enzyme tryptophan hydroxylase 2 (TPH2) into L-5-hydroxytryptophan. The latter is then converted to 5-HT after decarboxylation by aromatic amino acid decarboxylase (AADC). TPH2 is specific to serotonergic neurons in the central nervous system and is responsible for 5-HT synthesis in the brain [1,2], while AADC is localized in many cell types. In the brain, the majority of the cell bodies of serotonin-containing neurons are located in the brain stem and grouped into nine nuclei (named B1 to B9) which are, in turn, divided into the rostral or superior group and the caudal or inferior group [3,4]. The rostral group is located in the pons and midbrain and consists of both the dorsal nucleus (DRN) and medial nucleus (MRN) of the raphe [5,6]. The caudal group is located in the medulla and consists of nuclei projecting mainly to the spinal cord [5]. Although, in many brain territories, projections from serotonergic neurons originate from both the DRN and MRN, in some regions, 5-HT innervation originates solely or preferentially from the DRN or MRN. This is the case, for example, of the hippocampus, the dorsal part of which receives serotonergic innervation from the MRN [7]. On the other hand, the 5-HT fibers of the striatum, globus pallidus, lateral septum, and amygdala, including the prefrontal cortex, come from the DRN [8]. Along with the MRN, the DRN innervates the ventral part of the hippocampus and the prefrontal cortex [8] and certain nuclei of the thalamus [7,9,10]. There are also dense serotonergic projections between the DRN and MRN [11]. The central nervous system therefore contains a small number of 5-HT neurons grouped together in nuclei, but these neurons release serotonin diffusely throughout the brain at chemical synapses or by volume transmission [4,12]. Serotonin can therefore exert a modulating action on a wide variety of neuronal circuits and therefore functions [13,14]. The action of serotonin is complicated by the identification of seven families of serotonin receptor subtypes (5-HT1 to 5-HT7 receptors) in the brain [15]. Six of these are metabotropic G-protein-coupled receptors and one (the 5-HT3 receptor) is a ligand-gated ion channel (ionotropic receptor) [16]. This leads to the existence of eighteen 5-HT receptor subtypes plus a number of isoforms resulting from alternative mRNA splicing [17]. While all 5-HT receptor subtypes have post-synaptic expression on target cells, some subtypes, such as 5-HT1A, 5-HT1B/D, and 5-HT2B, are also localized pre-synaptically on the 5-HT neurons themselves (autoreceptors). Finally, 5-HT receptors also differ in their brain expression profiles in the brain [18] and in their interactions with other neurotransmitter systems.
In recent years, particular attention has been focused on the serotonin 4 receptor subtype [19,20,21], originally identified and cloned in rodents [22,23]. This receptor is positively coupled to adenylate cyclase via a Gs-type G protein leading to the canonical activation of the cAMP/PKA pathway [24]. But activation of the 5-HT4 receptor can also lead to the activation of the ERK signaling pathway independently of G proteins [25]. In addition to the fact that the 5-HT4 receptor has a constitutive activity, which is expressed at low receptor levels, another important feature of the receptor is that it can establish homodimers or heterodimers with several other proteins. These dimers functionally impact the signaling pathways on which they act [26]. In the brain, the 5-HT4 receptor is mainly located on neurons and has a dual somatodendritic and axonal localization [27]. Its cerebral distribution is highly conserved across species [27,28] with the highest densities in basal ganglia, particularly in the striatum [29,30] and the substantia nigra [31]. The 5-HT4 receptor is expressed by GABAergic neurons, as neurotoxic lesions of rodent basal ganglia components lead to dramatic decreases in 5-HT4 receptor binding in these regions [27,32]. The post-synaptic expression of the 5-HT4 receptor seems independent of dopaminergic neurons, as a dopaminergic lesion induced by intranigral injection of the 6-hydroxydopamine in rodents does not induce decreases in its expression in the striatum or the substantia nigra [27,32]. Similarly, the lesion of serotonergic neurons by intra-raphe injection of the 5,7-dihydroxytryptamine upregulates the 5-HT4 receptor binding in the basal ganglia and hippocampal regions [32]. The 5-HT4 receptor is also highly expressed in the cerebral cortex, hippocampus, and amygdala, where it may be carried by glutamatergic or cholinergic neurons, depending on the region [33,34]. The distribution of this receptor has also been studied in the post-mortem human brain by in situ hybridization and ligand binding studies and has been found to be high in basal ganglia regions and moderate to low in hippocampal formation and the cortical mantle [28]. Substantial evidence from preclinical and clinical studies have indicated an inverse relationship between the central 5-HT tonus and the density of the 5-HT4 receptor. In animals, 5-HT depletion or 5-HT boosting by the use of antidepressants, or again serotonin transporter gene manipulation, all lead to alterations of the 5-HT4 receptor binding in the brain [35,36,37]. In humans, the polymorphism of the serotonin transporter also results in different 5-HT levels and to different expression levels of the 5-HT4 receptor [38]. And as it has been demonstrated that the density of the 5-HT4 receptor adapts to the variations in brain extracellular 5-HT levels induced by antidepressant treatment, the central 5-HT4 receptor binding is now used as an in vivo biomarker of serotonergic tonus [39]. Outside the central nervous system, the 5-HT4 receptor is known to be also expressed in the gastrointestinal tract, the cardiovascular system, the bladder, and the adrenal glands [40].
The 5-HT4 receptor is involved in several functions. It mediates a positive feedback effect on the 5-HT neurons themselves. Indeed, the 5-HT4 receptor in the prefrontal cortex has a positive influence on the rate of discharge of 5-HT neurons in the raphe and increases the probability of action potential emissions via the positive feedback loop from the prefrontal cortex to the dorsal raphe [41,42]. The administration of 5-HT4 agonists also activates raphe 5-HT neurons [41,43]. It also releases 5-HT in the hippocampus [44]. The action of 5-HT4 receptors in the hippocampus plays an important role in the regulation of synaptic plasticity and related memory mechanisms [45]. Apart from the serotonergic system, the modulation of the 5-HT4 receptor has effects on the homeostasis of other neurotransmitter systems such as acetylcholine or dopamine [46]. The stimulation of the 5-HT4 receptor facilitates the release of acetylcholine in the frontal cortex of awake rats [47]. Similarly, the 5-HT4 receptor has been shown to facilitate dopamine release in the rat striatum both in vitro and in vivo [48]. The fact that modulation of the 5-HT4 receptor can affect the release of different neurotransmitters highlights the importance of this receptor as a potential therapeutic player in a variety of pathological conditions.

1.1. Mood and Anxiety Disorders

One of the major interests of this receptor is its potential involvement in mood and anxiety disorders (Table 1). It has already been shown that the 5-HT4 receptor expression is altered in various rodent models of depression. Some studies have reported a decrease in 5-HT4 receptor expression [35,49], while others have observed an increase [50]. It is highly probable that these divergent results are due to the use of different animal models of depression. On the contrary, similar results were obtained by studies modulating 5-HT4 receptor activation. Drugs that activate the 5-HT4 receptor play a fast-acting beneficial role on behavioral markers of depression or anxiety [51]. Electrophysiological, molecular, morphological, and behavioral changes that have previously been specifically linked to beneficial long-term treatment with selective serotonin reuptake inhibitors are already present after only 3 days of treatment when using 5-HT4 agonists in rodents [52]. The lack of 5-HT4 receptor modulates mood-like disorders [53] and further impacts the response to antidepressants in rodents [54]. Overexpression of the 5-HT4 receptor in the medial prefrontal cortex leads to antidepressant effects [55], whereas its functional ablation throughout the brain [54] or only in the hippocampus [53] causes anxiety and mood disorders. The pharmacological modulation of the 5-HT4 receptor located in the lateral habenula also produces antidepressant effects, via changes in monoamine levels in the prefrontal cortex and hippocampus in parkinsonian rats [56]. Yet, links have been shown between the antidepressant action of 5-HT4 receptor agonists and mechanisms of synaptic plasticity or gene regulation in the hippocampus. For example, the administration of 5-HT4 receptor agonists leads to the upregulation of mRNAs encoding BDNF (brain-derived neurotrophic factor) in the hippocampus. This increase results from activation of the CREB (cAMP-response element-binding protein) signaling pathway, which stimulates the expression of various neurotrophic factors in the hippocampus, including BDNF [57]. Furthermore, it has been shown that the stimulation of the 5-HT4 receptor drives neurogenesis in the hippocampus, while the lack of or the antagonism toward the 5-HT4 receptor reduces these effects [58,59,60,61]. Furthermore, it has been shown that the activation of the 5-HT4 receptor boosts the formation and the function of dendritic spines [62,63] and more generally acts as a unique modulator of synaptic plasticity mechanisms in the hippocampus [64].
In support of these preclinical data, several human studies have also shown that expression levels of the 5-HT4 receptor are altered in mood disorders and/or depression (Table 1). Autopsy results of brain tissue from antidepressant-free, violent suicide victims have shown a high density of the 5-HT4 receptor in the frontal cortex and caudate nucleus compared with control subjects [65]. In vivo PET (positron emission tomography) data have shown a lower cerebral 5-HT4 receptor binding in patients with major depressive disorder than in healthy subjects [66,67]. And concomitant anxiety in patients suffering from depression is also negatively associated with 5-HT4 receptor brain binding [68]. There is also an association between brain 5-HT4 receptor binding and reactivity to emotional faces in depressed and healthy individuals: the higher the concomitant anxiety, the lower the binding in the brain [69]. The global brain binding of this receptor has also been positively linked to high trait aggression in men [70]. Finally, a polymorphism in the 5-HT4 receptor in its C-terminal region has been linked to unipolar depression in humans [71], suggesting that genomic variations of this gene may confer susceptibility to mood disorders. A more recent study aimed to assess the association of the 5-HT4 receptor polymorphism (rs1345697) with depressive symptom improvement after 6 months of antidepressant treatment and found that there is an impact of this rs1345697 polymorphism on the improvement in depressive symptoms and remission of these symptoms after antidepressant treatment [72]. Taken together, these preclinical and clinical data tend to show that there is rather a drop in expression of the 5-HT4 receptor associated with mood and anxiety disorders, and that the use of agonists of this receptor could reduce them.
Table 1. List and main characteristics of preclinical and clinical studies involving the 5-HT4 receptor in mood and anxiety disorders.
Table 1. List and main characteristics of preclinical and clinical studies involving the 5-HT4 receptor in mood and anxiety disorders.
Goal of the Study Methods of 5-HT4
Receptor Investigation
Models Main ResultsReference
5-HT4 receptor polymorphismPolymorphismBipolar or depressed individualsC-terminal region linked to mood disorders *[71]
Major depressed individualsGenotype GG of rs1345697 polymorphism associated with lower remission rates[72]
5-HT4 receptor expressionBinding assaysViolent suicide human victims Ijms 25 05245 i001in frontal cortex *[65]
Ijms 25 05245 i001in caudate nucleus *
AutoradiographyRat depression modelIjms 25 05245 i002in the hippocampus[35]
Ijms 25 05245 i002in hippocampus and basal ganglia under chronic paroxetine
Ijms 25 05245 i002in hippocampus, hypothalamus, and lateral globus pallidus after 5-HT depletion
Mouse depression modelsIjms 25 05245 i001in ventral hippocampus[50]
Ijms 25 05245 i001in caudal caudate putamen
DeletionKO mouseAttenuation of “dematuration” of hippocampal granule cells[58]
Anxiety and anhedonia induced and responses to antidepressant impacted[54]
qRT-PCR, Western blotRat depression modelsIjms 25 05245 i002in hippocampus associated with anhedonia[49]
qRT-PCRCasein kinase 2 KO mouse Ijms 25 05245 i001in prefrontal cortex associated with anti-depressant-like phenotype[55]
In situ hybridizationKO mouseFluoxetine-induced neurogenesis abolished[61]
OverexpressionMouseAnti-depressant-like phenotype[55]
PET imagingHealthy individualsIjms 25 05245 i002in striatum and limbic regions negatively correlated with number of first-degree relatives treated for major depression[66]
Positive correlation with high trait aggression, only in males[70]
Anxious and depressed individualsIjms 25 05245 i002in global brain[68]
Depressed individualsNegative correlation with concomitant anxiety *[69]
Ijms 25 05245 i002in global brain *[67]
5-HT4 receptor modulation5-HT4 receptor agonistRat depression modelsIjms 25 05245 i002Depressive-like behavioral responses[52]
RatIjms 25 05245 i002Depressive- and anxiety-like behavioral responses[57]
Rat model of Parkinson’s diseaseIjms 25 05245 i002Depressive-like behavioral responses[56]
KO mouseAntidepressant behavioral responses[53]
* in comparison with healthy individuals. Up arrows indicate an increase, down arrows a decrease.

1.2. Cognitive Disorders

Another attraction of the 5-HT4 receptor subtype is that its activation can improve learning and memory (Table 2). Several preclinical experimental paradigms have evidenced such pro-cognitive properties [73,74]. Significant and reproducible improvements in a delayed matching to sample task accuracy are observed following an oral administration of a 5-HT4 agonist in both younger and older monkeys [73]. In rodents, activation of the 5-HT4 receptor leads to an increase in acetylcholine release, which counteracts memory deficits [21,46]. And it has been shown that the 5-HT4 receptor plays a specific and unique role in modulating the long-term potentiation of synapses in the hippocampus [75] and thus the memory mechanisms dependent on this structure [64]. It also promotes here the maturation of neuronal dendritic spines [63]. In humans, PET imaging studies have shown an inverse relationship between brain 5-HT4 receptor binding and memory performance in healthy subjects [76,77]. Even weaker brain binding is associated with memory dysfunction in untreated depressive patients compared with healthy individuals [67]. Finally, as in animals, stimulation of the 5-HT4 receptor by prucalopride improves cognitive performance in healthy subjects [78,79]. This beneficial effect is achieved by enhancing the resting-state functional connectivity between regions involved in cognitive networks while reducing that of the default mode network [80].
Stimulation of 5-HT4 receptors is also proving promising in various pathological situations. In a rat model of Alzheimer’s disease, the administration of a 5-HT4 receptor agonist improves cognitive performance via a reduction in apoptotic neurons and an improvement in synaptic function in the hippocampus [84]. In humans with early Alzheimer’s disease, a PET study elegantly showed that there is an increase in brain 5-HT4 receptor binding (parietal cortex, lateral prefrontal cortex, lateral temporal cortex, posterior cingulate gyrus, and hippocampus), and that this global cerebral upregulation is positively correlated with the cortical detection of beta-amyloid accumulation [81]. These preclinical and clinical data indicate that upregulation of the 5-HT4 receptor in the brain is beneficial, particularly in the hippocampus, where 5-HT4 receptor stimulation favors synaptic plasticity and memory mechanisms. This also suggests that increased expression of this receptor is a compensatory mechanism designed not only to counteract the accumulation of amyloid plaques [85], but also to promote acetylcholine release [47] and improve cognition.
Pro-cognitive effects are also mediated by 5-HT4 receptors in Parkinson’s neurodegenerative disease and may be underpinned by similar mechanisms. Indeed, the systemic administration of a 5-HT4 agonist improves the facilitation of contextual fear extinction, although without affecting hippocampal 5-HT4 receptors at the mRNA level, in the parkinsonian mice [83]. But another study on the parkinsonian rat has shown that lesioning the dopaminergic neurons of the median forebrain bundle produces cognitive deficits which are associated with a hippocampal increase in the receptor protein expression. Moreover, the local modulation of this receptor can counteract the memory deficits displayed by the animal, notably via increases in dopamine and serotonin in the hippocampus and the amygdala [82]. This suggests that, once again, the increased expression of the 5-HT4 receptor is a compensatory mechanism that stimulates the release of other neurotransmitters, thereby counteracting memory deficits.
Taken together, these data indicate that the stimulation of 5-HT4 receptors triggers a number of possible mechanisms (release of acetylcholine, serotonin, or dopamine; stimulation of maturation and synaptic plasticity in the hippocampus; induction of mechanisms able to counteract the accumulation of toxic aggregates and neuronal apoptosis), aimed at improving cognitive impairments. They also strengthen the view that 5-HT4 receptor agonists could be used in humans under different conditions: in subjects without a neurodegenerative pathology to improve memory deficits, and in elderly subjects affected by several evolutive diseases involving severe cognitive impairment [86].

1.3. Movement Disorders

Few studies have addressed the question of whether the 5-HT4 receptor subtype is in any way involved in movement disorders (Table 3). However, some animal models exist for inducing motor deficits or complications, in which either the impact of the pharmacological modulation of this receptor or its potential alteration of expression has been analyzed. In a rat model of morphine-driven stereotyped behavior and haloperidol-induced catalepsy, the systemic pharmacological blockade of the 5-HT4 receptor has no impact, suggesting that 5-HT4 receptor antagonism does not modulate motor behaviors related to the change in dopamine transmission [87]. Similarly, the turning behavior induced by amphetamine in unilateral dopamine-denervated rats is not affected after antagonizing the 5-HT4 receptor [88]. Divergent results were obtained in the parkinsonian mice. One study showed that the administration of 5-HT4 receptor agonists did not improve motor impairment [83]. On the contrary, a 7-day administration of prucalopride, a 5-HT4 agonist, induced a beneficial effect on motor deficits (bradykinesia, exploration, balance, and muscle strength) in the parkinsonian mice, by attenuating the striatal loss of dopamine [89]. Studies that have simply analyzed variations in receptor expression without using pharmacological ligands show divergent results. Changes in the striatal expression of the 5-HT4 receptor are not observed after dopaminergic lesion induction, nor after a chronic levodopa treatment, leading to abnormal involuntary movements (i.e., dyskinesias) in a rat model of Parkinson’s disease [90]. By contrast, using in vivo PET imaging and immunohistochemistry on post-mortem tissue, we found that the 5-HT4 receptor is upregulated in the striatum of parkinsonian macaques and rats, and further increased following chronic levodopa treatment, inducing dyskinesias in both animal species [91]. In humans, binding studies have not revealed any significant differences in 5-HT4 receptor expression in the putamen or the substantia nigra between control subjects and Parkinson’s sufferers, although there is a tendency for this receptor to increase in the putamen of patients [92,93]. By contrast, in Huntington’s patients, who exhibit involuntary movements, a marked putaminal decrease in the 5-HT4 receptor density is found [92]. However, a 3-week treatment with mosapride, a 5-HT4 receptor agonist, improved motor functions in a small number of Parkinsonian patients [94]. Taken together, these results show that the question of possible links between altered 5-HT4 receptor expression in the brain and movement disorders remains a largely open question. Further studies are sorely needed to elucidate whether or not this 5-HT4 receptor plays a role in the pathophysiology of movement disorders.

1.4. Food Disorders

Several studies, first in rodents and non-human primates, then in humans, have shown that the 5-HT4 receptor is associated with the control of food intake (Table 4). In normal or food-deprived mice, the local stimulation of the 5-HT4 receptor in the nucleus accumbens, a region involved in the reward system, inhibits food intake, an effect that is not obtained in 5-HT4-receptor-deficient mice [95]. The same group has also shown that deletion or overexpression of the 5-HT4 receptor in the medial prefrontal cortex induces overeating or hypophagia, respectively [96]. In obese rats, there is a higher level of the accumbal 5-HT4 receptor, whose stimulation with an agonist reduces eating [97]. In normal macaque monkeys, here again, the modulation of this receptor reduces or enhances the motivation for food intake, following systemic administration of a 5-HT4 receptor agonist or antagonist, respectively (personal communications). Similarly, in dopamine-lesioned macaques, the local injection of a 5-HT4 agonist into the ventral striatum (the rodent equivalent of the nucleus accumbens) triggers a reduced motivation for food intake (personal communications). In humans, a PET imaging study found that obesity is associated with the greater availability of this receptor in the nucleus accumbens and ventral pallidum of the brain reward circuit [98]. All these preclinical and clinical data point in the same direction and show that the high expression of 5-HT4 receptors in the nucleus accumbens is associated with a high body mass index, and that the stimulation of 5-HT4 receptors helps reduce the food intake.

1.5. Other Brain Disorders

Finally, one clinical study has reported an association between the methylation of the 5-HT4 receptor promoter and autism spectrum disorder [99]. Another has linked hippocampal structural changes in psychosis to the distribution of the 5-HT4 receptor [100], opening new avenues of research on this receptor. More recently, a study investigating the possible involvement of the serotonergic system in the pathophysiology of amyotrophic lateral sclerosis showed, using a transgenic mouse model, that systemic administration of a 5-HT4 receptor antagonist worsened the disease’s phenotype. It degraded motor function, without affecting disease progression. However, the 5-HT4 receptor antagonist also induced an increase in the expression of proteins forming abnormal aggregates, as well as an increase in the number of glial cells at the expense of neurons, in these mice [101], suggesting that using 5-HT4 receptors agonists might be of particular interest.

1.6. Gastrointestinal Disorders

Interestingly also, the 5-HT4 receptor agonists have been evaluated as prokinetics in the gastrointestinal tract (Table 5). The gavage of prucalopride causes faster whole gut transit and colonic motility and these effects are blocked by a 5-HT4 receptor antagonist and are not detected in 5-HT4 receptor knockout mice [102]. Sacral nerve electrical stimulation can improve defecation and accelerate the recovery of colonic transmission functions in a rat model of constipation, and these effects involve an upregulation of the 5-HT4 receptor in the colon [103]. The 5-HT4 receptor agonists prucalopride and tegaserod have both been approved by the FDA (Food and Drug Administration) and are used to treat various gastrointestinal-related conditions, including nausea, diarrhea, gastroparesis, constipation, and pain in irritable bowel syndrome [104]. In patients suffering from chronic idiopathic constipation [104], prucalopride is highly efficient, and other highly selective 5-HT4 receptor agonists are still in advanced development [105]. In other pathologies, such as Parkinson’s disease, 5-HT4 receptor agonists can also improve non-motor symptoms such as constipation. This has been demonstrated both preclinically and clinically. Rats with bilateral lesions of nigral dopaminergic neurons exhibit a reduced 5-HT4 receptor expression in the colonic muscular layer and manifest constipation, which is counteracted by a 5-HT4 receptor agonist (cisapride) [106]. In the MPTP (1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine)-treated mice, a beneficial effect is obtained on the integrity of the intestinal barrier after intraperitoneal administration of prucalopride, the 5-HT4 receptor agonist, for 7 days [89]. Clinical studies have also tested the effects of the pharmacological stimulation of the 5-HT4 receptor on constipation in Parkinson’s disease (see [107]) (Table 5). It has been shown that administration of the 5-HT4 receptor agonist mosapride for 3 months improves the intestinal transit in patients without worsening their parkinsonian symptoms or causing serious side-effects [108]. Another study involving a small number of parkinsonian patients showed an increasing gastric motility during mosapride treatment [94]. These data are interesting insofar as constipation is a non-motor symptom that may precede the onset of motor symptoms in Parkinson’s disease. We could therefore imagine that the therapeutic targeting of the 5-HT4 receptor could be modulated according to the symptom (motor or non-motor), but also according to the stage of the pathology (early, advanced, or late).

2. Conclusions

All these studies, both preclinical and clinical, highlight this 5-HT4 receptor as a potential therapeutic target, especially for various brain disorders. While the benefits of activating this receptor have been well demonstrated at the preclinical level for certain conditions, such as depressive symptoms, memory deficits, food consumption, and constipation, there is still some way to go in terms of the interest in acting pharmacologically on this receptor in the field of movement disorders. In humans, the activation of this receptor to counteract certain disorders is promising but remains to be validated, since only gastrointestinal disorders can be improved so far by using compounds acting toward this receptor. Further studies are sorely needed to unravel all the mysteries surrounding this receptor, both in terms of its functions and its potential therapeutic use.

Funding

Writing this review article was supported by the Fondation de France and the Agence Nationale de la Recherche. V.S. is supported by INSERM (Institut National de la Santé et de la Recherche Médicale).

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Gutknecht, L.; Kriegebaum, C.; Waider, J.; Schmitt, A.; Lesch, K.-P. Spatiotemporal expression of tryptophan hydroxylase isoforms in murine and human brain: Convergent data from Tph2 knockout mice. Eur. Neuropsychopharmacol. 2009, 19, 266–282. [Google Scholar] [CrossRef] [PubMed]
  2. Walther, D.J.; Bader, M. A unique central tryptophan hydroxylase isoform. Biochem. Pharmacol. 2003, 66, 1673–1680. [Google Scholar] [CrossRef]
  3. Dahlström, A.; Fuxe, K. Localization of monoamines in the lower brain stem. Experientia 1964, 20, 398–399. [Google Scholar] [CrossRef] [PubMed]
  4. Hornung, J.P. The human raphe nuclei and the serotonergic system. J. Chem. Neuroanat. 2003, 26, 331–343. [Google Scholar] [CrossRef]
  5. Bjarkam, C.R.; Sørensen, J.C.; Geneser, F.A. Distribution and morphology of serotonin-immunoreactive neurons in the brainstem of the New Zealand white rabbit. J. Comp. Neurol. 1997, 380, 507–519. [Google Scholar] [CrossRef]
  6. Soares, J.G.; Cavalcanti, J.R.L.P.; Oliveira, F.G.; Pontes, A.L.B.; Sousa, T.B.; Freitas, L.M.; Cavalcante, J.S.; Nascimento, E.S.; Cavalcante, J.C.; Costa, M.S.M.O. Nuclear organization of the serotonergic system in the brain of the rock cavy (Kerodon rupestris). J. Chem. Neuroanat. 2012, 43, 112–119. [Google Scholar] [CrossRef]
  7. Azmitia, E.C.; Segal, M. An autoradiographic analysis of the differential ascending projections of the dorsal and median raphe nuclei in the rat. J. Comp. Neurol. 1978, 179, 641–667. [Google Scholar] [CrossRef]
  8. Michelsen, K.A.; Schmitz, C.; Steinbusch, H.W.M. The dorsal raphe nucleus- from silver stainings to a role in depression. Brain Res. Rev. 2007, 55, 329–342. [Google Scholar] [CrossRef] [PubMed]
  9. McQuade, R.; Sharp, T. Functional mapping of dorsal and median raphe 5-hydroxytryptamine pathways in forebrain of the rat using microdialysis. J. Neurochem. 1997, 69, 791–796. [Google Scholar] [CrossRef]
  10. McQuade, R.; Sharp, T. Release of cerebral 5-hydroxytryptamine evoked by electrical stimulation of the dorsal and median raphe nuclei: Effect of a neurotoxic amphetamine. Neuroscience 1995, 68, 1079–1088. [Google Scholar] [CrossRef]
  11. Jacobs, B.L.; Azmitia, E.C. Structure and function of the brain serotonin system. Physiol. Rev. 1992, 72, 165–229. [Google Scholar] [CrossRef] [PubMed]
  12. Abrams, J.K.; Johnson, P.L.; Hollis, J.H.; Lowry, C.A. Anatomic and functional topography of the dorsal raphe nucleus. Ann. N. Y. Acad. Sci. 2004, 1018, 46–57. [Google Scholar] [CrossRef] [PubMed]
  13. Berger, M.; Gray, J.A.; Roth, B.L. The expanded biology of serotonin. Annu. Rev. Med. 2009, 60, 355–366. [Google Scholar] [CrossRef] [PubMed]
  14. Fidalgo, S.; Ivanov, D.K.; Wood, S.H. Serotonin: From top to bottom. Biogerontology 2013, 14, 21–45. [Google Scholar] [CrossRef] [PubMed]
  15. Kroeze, W.K.; Kristiansen, K.; Roth, B.L. Molecular biology of serotonin receptors structure and function at the molecular level. Curr. Top. Med. Chem. 2002, 2, 507–528. [Google Scholar] [CrossRef] [PubMed]
  16. Nichols, D.E.; Nichols, C.D. Serotonin receptors. Chem. Rev. 2008, 108, 1614–1641. [Google Scholar] [CrossRef] [PubMed]
  17. Compan, V.; Laurent, L.; Jean, A.; Macary, C.; Bockaert, J.; Dumuis, A. Serotonin signaling in eating disorders. Wiley Interdiscip. Rev. Membr. Transp. Signal 2012, 1, 715–729. [Google Scholar] [CrossRef]
  18. Mengod, G.; Vilaró, M.T.; Raurich, A.; López-Giménez, J.F.; Cortés, R.; Palacios, J.M. 5-HT receptors in mammalian brain: Receptor autoradiography and in situ hybridization studies of new ligands and newly identified receptors. Histochem. J. 1996, 28, 747–758. [Google Scholar] [CrossRef] [PubMed]
  19. Bockaert, J.; Claeysen, S.; Compan, V.; Dumuis, A. 5-HT(4) receptors: History, molecular pharmacology and brain functions. Neuropharmacology 2008, 55, 922–931. [Google Scholar] [CrossRef]
  20. Rebholz, H.; Friedman, E.; Castello, J. Alterations of Expression of the Serotonin 5-HT4 Receptor in Brain Disorders. Int. J. Mol. Sci. 2018, 19, 3581. [Google Scholar] [CrossRef]
  21. Roux, C.M.; Leger, M.; Freret, T. Memory Disorders Related to Hippocampal Function: The Interest of 5-HT4Rs Targeting. Int. J. Mol. Sci. 2021, 22, 12082. [Google Scholar] [CrossRef] [PubMed]
  22. Dumuis, A.; Bouhelal, R.; Sebben, M.; Cory, R.; Bockaert, J. A nonclassical 5-hydroxytryptamine receptor positively coupled with adenylate cyclase in the central nervous system. Mol. Pharmacol. 1988, 34, 880–887. [Google Scholar] [PubMed]
  23. Gerald, C.; Adham, N.; Kao, H.T.; Olsen, M.A.; Laz, T.M.; Schechter, L.E.; Bard, J.A.; Vaysse, P.J.; Hartig, P.R.; Branchek, T.A.; et al. The 5-ht4 receptor: Molecular cloning and pharmacological characterization of two splice variants. EMBO J. 1995, 14, 2806–2815. [Google Scholar] [CrossRef] [PubMed]
  24. Barthet, G.; Gaven, F.; Framery, B.; Shinjo, K.; Nakamura, T.; Claeysen, S.; Bockaert, J.; Dumuis, A. Uncoupling and endocytosis of 5-hydroxytryptamine 4 receptors. Distinct molecular events with different grk2 requirements. J. Biol. Chem. 2005, 280, 27924–27934. [Google Scholar] [CrossRef] [PubMed]
  25. Barthet, G.; Framery, B.; Gaven, F.; Pellissier, L.; Reiter, E.; Claeysen, S.; Bockaert, J.; Dumuis, A. 5-hydroxytryptamine 4 receptor activation of the extracellular signal-regulated kinase pathway depends on src activation but not on g protein or beta-arrestin signaling. Mol. Biol. Cell 2007, 18, 1979–1991. [Google Scholar] [CrossRef] [PubMed]
  26. Mitroshina, E.V.; Marasanova, E.A.; Vedunova, M.V. Functional Dimerization of Serotonin Receptors: Role in Health and Depressive Disorders. Int. J. Mol. Sci. 2023, 24, 16416. [Google Scholar] [CrossRef]
  27. Vilaró, M.T.; Cortés, R.; Mengod, G. Serotonin 5-HT4 receptors and their mRNAs in rat and guinea pig brain: Distribution and effects of neurotoxic lesions. J. Comp. Neurol. 2005, 484, 418–439. [Google Scholar] [CrossRef] [PubMed]
  28. Bonaventure, P.; Hall, H.; Gommeren, W.; Cras, P.; Langlois, X.; Jurzak, M.; Leysen, J.E. Mapping of serotonin 5-HT(4) receptor mRNA and ligand binding sites in the post-mortem human brain. Synapse 2000, 36, 35–46. [Google Scholar] [CrossRef]
  29. Beliveau, V.; Ganz, M.; Feng, L.; Ozenne, B.; Højgaard, L.; Fisher, P.M.; Svarer, C.; Greve, D.N.; Knudsen, G.M. A High-Resolution In Vivo Atlas of the Human Brain’s Serotonin System. J. Neurosci. 2017, 37, 120–128. [Google Scholar] [CrossRef]
  30. Sgambato-Faure, V.; Tremblay, L. Dopamine and serotonin modulation of motor and non-motor functions of the non-human primate striato-pallidal circuits in normal and pathological states. J. Neural Transm. 2018, 125, 485–500. [Google Scholar] [CrossRef]
  31. Waeber, C.; Sebben, M.; Nieoullon, A.; Bockaert, J.; Dumuis, A. Regional distribution and ontogeny of 5-HT4 binding sites in rodent brain. Neuropharmacology 1994, 33, 527–541. [Google Scholar] [CrossRef] [PubMed]
  32. Compan, V.; Daszuta, A.; Salin, P.; Sebben, M.; Bockaert, J.; Dumuis, A. Lesion study of the distribution of serotonin 5-HT4 receptors in rat basal ganglia and hippocampus. Eur. J. Neurosci. 1996, 8, 2591–2598. [Google Scholar] [CrossRef] [PubMed]
  33. Egeland, M.; Warner-Schmidt, J.; Greengard, P.; Svenningsson, P. Co-expression of serotonin 5-HT(1B) and 5-HT(4) receptors in p11 containing cells in cerebral cortex, hippocampus, caudate-putamen and cerebellum. Neuropharmacology 2011, 61, 442–450. [Google Scholar] [CrossRef] [PubMed]
  34. Peñas-Cazorla, R.; Vilaró, M.T. Serotonin 5-HT4 receptors and forebrain cholinergic system: Receptor expression in identified cell populations. Brain Struct. Funct. 2015, 220, 3413–3434. [Google Scholar] [CrossRef]
  35. Licht, C.L.; Marcussen, A.B.; Wegener, G.; Overstreet, D.H.; Aznar, S.; Knudsen, G.M. The brain 5-HT4 receptor binding is down-regulated in the Flinders Sensitive Line depression model and in response to paroxetine administration. J. Neurochem. 2009, 109, 1363–1374. [Google Scholar] [CrossRef] [PubMed]
  36. Vidal, R.; Valdizan, E.M.; Mostany, R.; Pazos, A.; Castro, E. Long-term treatment with fluoxetine induces desensitization of 5-HT4 receptor-dependent signalling and functionality in rat brain. J. Neurochem. 2009, 110, 1120–1127. [Google Scholar] [CrossRef] [PubMed]
  37. Jennings, K.A.; Licht, C.L.; Bruce, A.; Lesch, K.P.; Knudsen, G.M.; Sharp, T. Genetic variation in 5-hydroxytryptamine transporter expression causes adaptive changes in 5-HT4 receptor levels. Int. J. Neuropsychopharmacol. 2012, 15, 1099–1107. [Google Scholar] [CrossRef] [PubMed]
  38. Fisher, P.M.; Holst, K.K.; Mc Mahon, B.; Haahr, M.E.; Madsen, K.; Gillings, N.; Baaré, W.F.; Jensen, P.S.; Knudsen, G.M. 5-HTTLPR status predictive of neocortical 5-HT4 binding assessed with [(11)C]SB207145 PET in humans. Neuroimage 2012, 62, 130–136. [Google Scholar] [CrossRef] [PubMed]
  39. Haahr, M.E.; Fisher, P.M.; Jensen, C.G.; Frokjaer, V.G.; Mahon, B.M.; Madsen, K.; Baaré, W.F.; Lehel, S.; Norremolle, A.; Rabiner, E.A.; et al. Central 5-HT4 receptor binding as biomarker of serotonergic tonus in humans: A [11C]SB207145 PET study. Mol. Psychiatry 2014, 19, 427–432. [Google Scholar] [CrossRef]
  40. Hegde, S.S.; Eglen, R.M. Peripheral 5-HT4 receptors. FASEB J. 1996, 10, 1398–1407. [Google Scholar] [CrossRef]
  41. Lucas, G.; Compan, V.; Charnay, Y.; Neve, R.L.; Nestler, E.J.; Bockaert, J.; Barrot, M.; Debonnel, G. Frontocortical 5-HT4 Receptors Exert Positive Feedback on Serotonergic Activity: Viral Transfections, Subacute and Chronic Treatments with 5-HT4 Agonists. Biol. Psychiatry 2005, 57, 918–925. [Google Scholar] [CrossRef] [PubMed]
  42. Conductier, G.; Dusticier, N.; Lucas, G.; Côté, F.; Debonnel, G.; Daszuta, A.; Dumuis, A.; Nieoullon, A.; Hen, R.; Bockaert, J.; et al. Adaptive Changes in Serotonin Neurons of the Raphe Nuclei in 5-HT4 Receptor Knock-out Mouse. Eur. J. Neurosci. 2006, 24, 1053–1062. [Google Scholar] [CrossRef] [PubMed]
  43. Moha ou Maati, H.; Bourcier-Lucas, C.; Veyssiere, J.; Kanzari, A.; Heurteaux, C.; Borsotto, M.; Haddjeri, N.; Lucas, G. The Peptidic Antidepressant Spadin Interacts with Prefrontal 5-HT4 and MGluR2 Receptors in the Control of Serotonergic Function. Brain Struct. Funct. 2016, 221, 21–37. [Google Scholar] [CrossRef] [PubMed]
  44. Ge, J.; Barnes, N.M. 5-HT4 receptor-mediated modulation of 5-HT release in the rat hippocampus in vivo. Br. J. Pharmacol. 1996, 117, 1475–1480. [Google Scholar] [CrossRef] [PubMed]
  45. Twarkowski, H.; Hagena, H.; Manahan-Vaughan, D. The 5-hydroxytryptamine 4 Receptor Enables Differentiation of Informational Content and Encoding in the Hippocampus. Hippocampus 2016, 26, 875–891. [Google Scholar] [CrossRef] [PubMed]
  46. Hatat, B.; Yahiaoui, S.; Lecoutey, C.; Davis, A.; Freret, T.; Boulouard, M.; Claeysen, S.; Rochais, C.; Dallemagne, P. A Novel invivo Anti-amnesic Agent, Specially Designed to Express Both Acetylcholinesterase (AChE) Inhibitory, Serotonergic Subtype 4 Receptor (5-HT4R) Agonist and Serotonergic Subtype 6 Receptor (5-HT6R) Inverse Agonist Activities, with a Potential Interest Against Alzheimer’s Disease. Front. Aging Neurosci. 2019, 11, 148. [Google Scholar] [PubMed]
  47. Consolo, S.; Arnaboldi, S.; Giorgi, S.; Russi, G.; Ladinsky, H. 5-HT4 receptor stimulation facilitates acetylcholine release in rat frontal cortex. Neuroreport 1994, 5, 1230–1232. [Google Scholar] [CrossRef] [PubMed]
  48. Steward, L.J.; Ge, J.; Stowe, R.L.; Brown, D.C.; Bruton, R.K.; Stokes, P.R.; Barnes, N.M. Ability of 5-HT4 receptor ligands to modulate rat striatal dopamine release in vitro and in vivo. Br. J. Pharmacol. 1996, 117, 55–62. [Google Scholar] [CrossRef] [PubMed]
  49. Bai, M.; Zhu, X.-Z.; Zhang, Y.; Zhang, S.; Zhang, L.; Xue, L.; Zhong, M.; Zhang, X. Anhedonia Was Associated with the Dysregulation of Hippocampal HTR4 and MicroRNA Let-7a in Rats. Physiol. Behav. 2014, 129, 135–141. [Google Scholar] [CrossRef]
  50. Licht, C.L.; Kirkegaard, L.; Zueger, M.; Chourbaji, S.; Gass, P.; Aznar, S.; Knudsen, G.M. Changes in 5-HT4 Receptor and 5-HT Transporter Binding in Olfactory Bulbectomized and Glucocorticoid Receptor Heterozygous Mice. Neurochem. Int. 2010, 56, 603–610. [Google Scholar] [CrossRef]
  51. Vidal, R.; Castro, E.; Pilar-Cuéllar, F.; Pascual-Brazo, J.; Díaz, A.; Rojo, M.L.; Linge, R.; Martín, A.; Valdizán, E.M.; Pazos, A. Serotonin 5-HT4 Receptors: A New Strategy for Developing Fast Acting Antidepressants? Curr. Pharm. Des. 2014, 20, 3751–3762. [Google Scholar] [CrossRef] [PubMed]
  52. Lucas, G.; Rymar, V.V.; Du, J.; Mnie-Filali, O.; Bisgaard, C.; Manta, S.; Lambas-Senas, L.; Wiborg, O.; Haddjeri, N.; Piñeyro, G.; et al. Serotonin(4) (5-HT(4)) receptor agonists are putative antidepressants with a rapid onset of action. Neuron 2007, 55, 712–725. [Google Scholar] [CrossRef] [PubMed]
  53. Karayol, R.; Medrihan, L.; Warner-Schmidt, J.L.; Fait, B.W.; Rao, M.N.; Holzner, E.B.; Greengard, P.; Heintz, N.; Schmidt, E.F. Serotonin receptor 4 in the hippocampus modulates mood and anxiety. Mol. Psychiatry 2021, 26, 2334–2349. [Google Scholar] [CrossRef] [PubMed]
  54. Amigó, J.; Díaz, A.; Pilar-Cuéllar, F.; Vidal, R.; Martín, A.; Compan, V.; Pazos, A.; Castro, E. The absence of 5-HT4 receptors modulates depression- and anxiety-like responses and influences the response of fluoxetine in olfactory bulbectomised mice: Adaptive changes in hippocampal neuroplasticity markers and 5-HT1A autoreceptor. Neuropharmacology 2016, 111, 47–58. [Google Scholar] [CrossRef] [PubMed]
  55. Castello, J.; LeFrancois, B.; Flajolet, M.; Greengard, P.; E Friedman, E.; Rebholz, H. CK2 regulates 5-HT4 receptor signaling and modulates depressive-like behavior. Mol. Psychiatry 2018, 23, 872–882. [Google Scholar] [CrossRef] [PubMed]
  56. Guo, Y.; Zhang, L.; Zhang, J.; Lv, S.-X.; Du, C.-X.; Wang, T.; Wang, H.-S.; Xie, W.; Liu, J. Activation and Blockade of Serotonin-4 Receptors in the Lateral Habenula Produce Antidepressant Effects in the Hemiparkinsonian Rat. Neuropsychobiology 2021, 80, 52–63. [Google Scholar] [CrossRef]
  57. Pascual-Brazo, J.; Castro, E.; Díaz, Á.; Valdizán, E.M.; Pilar-Cuéllar, F.; Vidal, R.; Treceño, B.; Pazos, Á. Modulation of Neuroplasticity Pathways and Antidepressant-like Behavioural Responses Following the Short-Term (3 and 7 Days) Administration of the 5-HT4 Receptor Agonist RS67333. Int. J. Neuropsychopharmacol. 2012, 15, 631–643. [Google Scholar] [CrossRef] [PubMed]
  58. Kobayashi, K.; Ikeda, Y.; Sakai, A.; Yamasaki, N.; Haneda, E.; Miyakawa, T.; Suzuki, H. Reversal of Hippocampal Neuronal Maturation by Serotonergic Antidepressants. Proc. Natl. Acad. Sci. USA 2010, 107, 8434–8439. [Google Scholar] [CrossRef] [PubMed]
  59. Mendez-David, I.; David, D.J.; Darcet, F.; Wu, M.V.; Kerdine-Römer, S.; Gardier, A.M.; Hen, R. Rapid Anxiolytic Effects of a 5-HT4 Receptor Agonist Are Mediated by a Neurogenesis-Independent Mechanism. Neuropsychopharmacology 2014, 39, 1366–1378. [Google Scholar] [CrossRef]
  60. Ishizuka, T.; Goshima, H.; Ozawa, A.; Watanabe, Y. Stimulation of 5-HT4 Receptor Enhances Differentiation of Mouse Induced Pluripotent Stem Cells into Neural Progenitor Cells. Clin. Exp. Pharmacol. Physiol. 2014, 41, 345–350. [Google Scholar] [CrossRef]
  61. Imoto, Y.; Kira, T.; Sukeno, M.; Nishitani, N.; Nagayasu, K.; Nakagawa, T.; Kaneko, S.; Kobayashi, K.; Segi-Nishida, E. Role of the 5-HT4 Receptor in Chronic Fluoxetine Treatment-Induced Neurogenic Activity and Granule Cell Dematuration in the Dentate Gyrus. Mol. Brain 2015, 8, 29. [Google Scholar] [CrossRef] [PubMed]
  62. Segi-Nishida, E. The Effect of Serotonin-Targeting Antidepressants on Neurogenesis and Neuronal Maturation of the Hippocampus Mediated via 5-HT1A and 5-HT4 Receptors. Front. Cell. Neurosci. 2017, 11, 142. [Google Scholar] [CrossRef] [PubMed]
  63. Schill, Y.; Bijata, M.; Kopach, O.; Cherkas, V.; Abdel-Galil, D.; Böhm, K.; Schwab, M.H.; Matsuda, M.; Compan, V.; Basu, S.; et al. Serotonin 5-HT4 receptor boosts functional maturation of dendritic spines via RhoA-dependent control of F-actin. Commun. Biol. 2020, 3, 76. [Google Scholar] [CrossRef]
  64. Hagena, H.; Manahan-Vaughan, D. The serotonergic 5-HT4 receptor: A unique modulator of hippocampal synaptic information processing and cognition. Neurobiol. Learn. Mem. 2017, 138, 145–153. [Google Scholar] [CrossRef] [PubMed]
  65. Rosel, P.; Arranz, B.; Urretavizcaya, M.; Oros, M.; San, L.; Navarro, M.A. Altered 5-HT2A and 5-HT4 postsynaptic receptors and their intracellular signalling systems IP3 and cAMP in brains from depressed violent suicide victims. Neuropsychobiology 2004, 49, 189–195. [Google Scholar] [CrossRef] [PubMed]
  66. Madsen, K.; Torstensen, E.; Holst, K.K.; Haahr, M.E.; Knorr, U.; Frokjaer, V.G.; Brandt-Larsen, M.; Iversen, P.; Fisher, P.M.; Knudsen, G.M. Familial risk for major depression is associated with lower striatal 5-HT₄ receptor binding. Int. J. Neuropsychopharmacol. 2014, 18, pyu034. [Google Scholar] [CrossRef]
  67. Köhler-Forsberg, K.; Dam, V.H.; Ozenne, B.; Sankar, A.; Beliveau, V.; Landman, E.B.; Larsen, S.V.; Poulsen, A.S.; Ip, C.-T.; Jørgensen, A.; et al. Serotonin 4 Receptor Brain Binding in Major Depressive Disorder and Association With Memory Dysfunction. JAMA Psychiatry 2023, 80, 296–304. [Google Scholar] [CrossRef] [PubMed]
  68. Köhler-Forsberg, K.; Ozenne, B.; Larsen, S.V.; Poulsen, A.S.; Landman, E.B.; Dam, V.H.; Ip, C.-T.; Jørgensen, A.; Svarer, C.; Knudsen, G.M.; et al. Concurrent anxiety in patients with major depression and cerebral serotonin 4 receptor binding. A NeuroPharm-1 study. Transl. Psychiatry 2022, 12, 273. [Google Scholar] [CrossRef] [PubMed]
  69. Sankar, A.; Ozenne, B.; Dam, V.H.; Svarer, C.; Jørgensen, M.B.; Miskowiak, K.W.; Frokjaer, V.G.; Knudsen GM and Fisher, P.M. Association between brain serotonin 4 receptor binding and reactivity to emotional faces in depressed and healthy individuals. Transl. Psychiatry 2023, 13, 165. [Google Scholar] [CrossRef]
  70. da Cunha-Bang, S.; Mc Mahon, B.; Fisher, P.M.; Jensen, P.S.; Svarer, C.; Knudsen, G.M. High trait aggression in men is associated with low 5-HT levels, as indexed by 5-HT4 receptor binding. Soc. Cogn. Affect. Neurosci. 2016, 11, 548–555. [Google Scholar] [CrossRef]
  71. Ohtsuki, T.; Ishiguro, H.; Detera-Wadleigh, S.D.; Toyota, T.; Shimizu, H.; Yamada, K.; Yoshitsugu, K.; Hattori, E.; Yoshikawa, T.; Arinami, T. Association between Serotonin 4 Receptor Gene Polymorphisms and Bipolar Disorder in Japanese Case-Control Samples and the NIMH Genetics Initiative Bipolar Pedigrees. Mol. Psychiatry 2002, 7, 954–961. [Google Scholar] [CrossRef]
  72. Poinsignon, V.; Colle, R.; Asmar, K.E.; Mendez-David, I.; David, D.J.; Ait Tayeb, A.E.K.; Chappell, K.; Gressier, F.; Herrero, H.; Fève, B.; et al. The GG genotype of the serotonin 4 receptor genetic polymorphism, rs1345697, is associated with lower remission rates after antidepressant treatment: Findings from the METADAP cohort. J. Affect. Disord. 2022, 299, 335–343. [Google Scholar] [CrossRef] [PubMed]
  73. Terry, A.V., Jr.; Buccafusco, J.J.; Jackson, W.J.; Prendergast, M.A.; Fontana, D.J.; Wong, E.H.; Bonhaus, D.W.; Weller, P.; Eglen, R.M. Enhanced delayed matching performance in younger and older macaques administered the 5-HT4 receptor agonist, RS 17017. Psychopharmacology 1998, 135, 407–415. [Google Scholar] [CrossRef] [PubMed]
  74. Quiedeville, A.; Boulouard, M.; Hamidouche, K.; Da Silva Costa-Aze, V.; Nee, G.; Rochais, C.; Dallemagne, P.; Fabis, F.; Freret, T.; Bouet, V. Chronic activation of 5-HT4 receptors or blockade of 5-HT6 receptors improve memory performances. Behav. Brain Res. 2015, 293, 10–17. [Google Scholar] [CrossRef] [PubMed]
  75. Teixeira, C.M.; Rosen, Z.B.; Suri, D.; Sun, Q.; Hersh, M.; Sargin, D.; Dincheva, I.; Morgan, A.A.; Spivack, S.; Krok, A.C.; et al. Hippocampal 5-HT Input Regulates Memory Formation and Schaffer Collateral Excitation. Neuron 2018, 98, 992–1004. [Google Scholar] [CrossRef] [PubMed]
  76. Haahr, M.E.; Fisher, P.; Holst, K.; Madsen, K.; Jensen, C.G.; Marner, L.; Lehel, S.; Baaré, W.; Knudsen, G.; Hasselbalch, S. The 5-HT4 receptor levels in hippocampus correlates inversely with memory test performance in humans. Hum. Brain Mapp. 2013, 34, 3066–3074. [Google Scholar] [CrossRef] [PubMed]
  77. Stenbæk, D.S.; Fisher, P.M.; Ozenne, B.; Andersen, E.; Hjordt, L.V.; McMahon, B.; Hasselbalch, S.G.; Frokjaer, V.G.; Knudsen, G.M. Brain serotonin 4 receptor binding is inversely associated with verbal memory recall. Brain Behav. 2017, 7, e00674. [Google Scholar] [CrossRef] [PubMed]
  78. Murphy, S.E.; Wright, L.C.; Browning, M.; Cowen, P.J.; Harmer, C.J. A role for 5-HT4 receptors in human learning and memory. Psychol. Med. 2020, 50, 2722–2730. [Google Scholar] [CrossRef] [PubMed]
  79. de Cates, A.N.; Wright, L.C.; Martens, M.A.G.; Gibson, D.; Türkmen, C.; Filippini, N.; Cowen, P.J.; Harmer, C.J.; Murphy, S.E. Déjà-vu? Neural and behavioural effects of the 5-HT4 receptor agonist, prucalopride, in a hippocampal-dependent memory task. Transl. Psychiatry 2021, 11, 497. [Google Scholar] [CrossRef]
  80. de Cates, A.N.; Martens, M.A.G.; Wright, L.C.; Gibson, D.; Spitz, G.; Gould van Praag, C.D.; Suri, S.; Cowen, P.J.; Murphy, S.E.; Harmer, C.J. 5-HT4 Receptor Agonist Effects on Functional Connectivity in the Human Brain: Implications for Procognitive Action. Biol. Psychiatry Cogn. Neurosci. Neuroimaging 2023, 8, 1124–1134. [Google Scholar] [CrossRef]
  81. Madsen, K.; Neumann, W.J.; Holst, K.; Marner, L.; Haahr, M.T.; Lehel, S.; Knudsen, G.M.; Hasselbalch, S.G. Cerebral serotonin 4 receptors and amyloid-β in early Alzheimer’s disease. J. Alzheimers Dis. 2011, 26, 457–466. [Google Scholar] [CrossRef]
  82. Wang, J.-W.; Liu, J.; Wang, Z.-L.; Gao, F.; Yang, J.; Wang, X.-C.; Guo, Y.; Wang, Y.; Ma, B.-R.; Wang, H.-S.; et al. Activation and blockade of 5-HT4 receptors in the dorsal hippocampus enhance working and hippocampus-dependent memories in the unilateral 6-hydroxydopamine lesioned rats. Behav. Brain Res. 2022, 431, 113952. [Google Scholar] [CrossRef] [PubMed]
  83. Ishii, T.; Kinoshita, K.-I.; Muroi, Y. Serotonin 5-HT4 Receptor Agonists Improve Facilitation of Contextual Fear Extinction in an MPTP-Induced Mouse Model of Parkinson’s Disease. Int. J. Mol. Sci. 2019, 20, 5340. [Google Scholar] [CrossRef] [PubMed]
  84. Hashemi-Firouzi, N.; Shahidi, S.; Soleimani Asl, S. Chronic stimulation of the serotonergic 5-HT4 receptor modulates amyloid-beta-related impairments in synaptic plasticity and memory deficits in male rats. Brain Res. 2021, 1773, 147701. [Google Scholar] [CrossRef] [PubMed]
  85. Cho, S.; Hu, Y. Activation of 5-HT4 receptors inhibits secretion of beta-amyloid peptides and increases neuronal survival. Exp. Neurol. 2007, 203, 274–278. [Google Scholar] [CrossRef] [PubMed]
  86. Barrett, M.J.; Cloud, L.J.; Shah, H.; Holloway, K.L. Therapeutic approaches to cholinergic deficiency in Lewy body diseases. Expert Rev. Neurother. 2020, 20, 41–53. [Google Scholar] [CrossRef] [PubMed]
  87. De Deurwaerdère, P.; Cervo, L.; Stinus, L.; Spampinato, U. Central 5-HT(4) receptors and dopamine-dependent motor behaviors: Searching for a functional role. Pharmacol. Biochem. Behav. 2002, 71, 627–633. [Google Scholar] [CrossRef] [PubMed]
  88. Reavill, C.; Hatcher, J.P.; Lewis, V.A.; Sanger, G.J.; Hagan, J. 5-HT4 receptors antagonism does not affect motor and reward mechanisms in the rat. Eur. J. Pharmacol. 1998, 357, 115–120. [Google Scholar] [CrossRef] [PubMed]
  89. Shi, Y.; Qiao, C.-M.; Zhou, Y.; Wu, J.; Cui, C.; Hong, H.; Jia, X.-B.; Huang, S.-B.; Yao, L.; Zhao, W.-J.; et al. Protective effects of prucalopride in MPTP-induced Parkinson’s disease mice: Neurochemistry, motor function and gut barrier. Biochem. Biophys. Res. Commun. 2021, 556, 16–22. [Google Scholar] [CrossRef]
  90. Padovan-Neto, F.E.; Patterson, S.; Voelkner, N.M.F.; Altwal, F.; Beverley, J.A.; West, A.R.; Steiner, H. Selective Regulation of 5-HT1B Serotonin Receptor Expression in the Striatum by Dopamine Depletion and Repeated L-DOPA Treatment: Relationship to L-DOPA-Induced Dyskinesias. Mol. Neurobiol. 2020, 57, 736–751. [Google Scholar] [CrossRef]
  91. Cirillo, R.; Duperrier, S.; Parekh, P.; Millot, M.; Li, Q.; Thiolat, M.L.; Morelli, M.; Xie, J.; Le Bars, D.; Redouté, J.; et al. Striatal serotonin 4 receptor is increased in experimental parkinsonism and dyskinesia. J. Park. Dis. 2024, 14, 261–267. [Google Scholar] [CrossRef]
  92. Reynolds, G.P.; Mason, S.L.; Meldrum, A.; De Keczer, S.; Parnes, H.; Eglen, R.M.; Wong, E.H. 5-Hydroxytryptamine (5-HT)4 receptors in post mortem human brain tissue: Distribution, pharmacology and effects of neurodegenerative diseases. Br. J. Pharmacol. 1995, 114, 993–998. [Google Scholar] [CrossRef]
  93. Wong, E.H.F.; Reynolds, G.P.; Bonhaus, D.W.; Hsu, S.; Eglen, R.M. Characterization of [3H]GR 113808 binding to 5-HT 4 receptors in brain tissues from patients with neurodegenerative disorders. Behav. Brain Res. 1996, 73, 249–252. [Google Scholar] [CrossRef]
  94. Asai, H.; Udaka, F.; Hirano, M.; Minami, T.; Oda, M.; Kubori, T.; Nishinaka, K.; Kameyama, M.; Ueno, S. Increased gastric motility during 5-HT4 agonist therapy reduces response fluctuations in Parkinson’s disease. Park. Relat. Disord. 2005, 11, 499–502. [Google Scholar] [CrossRef]
  95. Jean, A.; Conductier, G.; Manrique, C.; Bouras, C.; Berta, P.; Hen, R.; Charnay, Y.; Bockaert, J.; Compan, V. Anorexia induced by activation of serotonin 5-HT4 receptors is mediated by increases in CART in the nucleus accumbens. Proc. Natl. Acad. Sci. USA 2007, 104, 16335–16340. [Google Scholar] [CrossRef] [PubMed]
  96. Jean, A.; Laurent, L.; Delaunay, S.; Doly, S.; Dusticier, N.; Linden, D.; Neve, R.; Maroteaux, L.; Nieoullon, A.; Compan, V. Adaptive Control of Dorsal Raphe by 5-HT4 in the Prefrontal Cortex Prevents Persistent Hypophagia following Stress. Cell Rep. 2017, 21, 901–909. [Google Scholar] [CrossRef]
  97. Ratner, C.; Ettrup, A.; Bueter, M.; Haahr, M.E.; Compan, V.; le Roux, C.W.; Levin, B.; Hansen, H.H.; Knudsen, G.M. Cerebral markers of the serotonergic system in rat models of obesity and after Roux-en-Y gastric bypass. Obesity 2012, 20, 2133–2141. [Google Scholar] [CrossRef] [PubMed]
  98. Haahr, M.E.; Rasmussen, P.M.; Madsen, K.; Marner, L.; Ratner, C.; Gillings, N.; Baaré, W.F.; Knudsen, G.M. Obesity is associated with high serotonin 4 receptor availability in the brain reward circuitry. Neuroimage 2012, 61, 884–888. [Google Scholar] [CrossRef]
  99. Hu, Z.; Ying, X.; Huang, L.; Zhao, Y.; Zhou, D.; Liu, J.; Zhong, J.; Huang, T.; Zhang, W.; Cheng, F.; et al. Association of human serotonin receptor 4 promoter methylation with autism spectrum disorder. Medicine 2020, 99, e18838. [Google Scholar] [CrossRef] [PubMed]
  100. Park, M.T.M.; Jeon, P.; Khan, A.R.; Dempster, K.; Mallar Chakravarty, M.; Lerch, J.P.; MacKinley, M.; Théberge, J.; Palaniyappan, L. Hippocampal neuroanatomy in first episode psychosis: A putative role for glutamate and serotonin receptors. Prog. Neuropsychopharmacol. Biol. Psychiatry 2021, 110, 110297. [Google Scholar] [CrossRef]
  101. Jiang, S.S.; Gong, M.N.; Rao, W.; Chai, W.; Chen, W.Z.; Zhang, X.; Nie, H.B.; Xu, R.S. 5-Hydroxytryptamine: A potential therapeutic target in amyotrophic lateral sclerosis. Neural Regen. Res. 2023, 18, 2047–2055. [Google Scholar] [PubMed]
  102. Konen, J.R.; Haag, M.M.; Guseva, D.; Hurd, M.; Linton, A.A.; Lavoie, B.; Kerrigan, C.B.; Joyce, E.; Bischoff, S.; Swann, S.; et al. Prokinetic actions of luminally acting 5-HT4 receptor agonists. Neurogastroenterol. Motil. 2021, 33, e14026. [Google Scholar] [CrossRef] [PubMed]
  103. Zhu, Y.; Cheng, J.; Yin, J.; Yang, Y.; Guo, J.; Zhang, W.; Xie, B.; Lu, H.; Hao, D. Effects of sacral nerve electrical stimulation on 5 HT and 5 HT3AR/5 HT4R levels in the colon and sacral cord of acute spinal cord injury rat models. Mol. Med. Rep. 2020, 22, 763–773. [Google Scholar] [CrossRef] [PubMed]
  104. Mahajan, R. Prucalopride: A Recently Approved Drug by the Food and Drug Administration for Chronic Idiopathic Constipation. Int. J. Appl. Basic Med. Res. 2019, 9, 1–2. [Google Scholar] [CrossRef] [PubMed]
  105. Bassotti, G.; Satta, P.U.; Berti, G.; Lai, M.; Villanacci, V.; Bellini, M. Pharmacotherapeutic advances for chronic idiopathic constipation in adults. Expert Opin. Pharmacother. 2022, 23, 2053–2078. [Google Scholar] [CrossRef] [PubMed]
  106. Zhang, X.; Li, Y.; Liu, C.; Fan, R.; Wang, P.; Zheng, L.; Hong, F.; Feng, X.; Zhang, Y.; Li, L.; et al. Alteration of enteric monoamines with monoamine receptors and colonic dysmotility in 6-hydroxydopamine-induced Parkinson’s disease rats. Transl. Res. 2015, 166, 152–162. [Google Scholar] [CrossRef] [PubMed]
  107. Mozaffari, S.; Nikfar, S.; Daniali, M.; Abdollahi, M. The pharmacological management of constipation in patients with Parkinson’s disease: A much-needed relief. Expert Opin. Pharmacother. 2020, 21, 701–707. [Google Scholar] [CrossRef]
  108. Liu, Z.; Sakakibara, R.; Odaka, T.; Uchiyama, T.; Uchiyama, T.; Yamamoto, T.; Ito, T.; Asahina, M.; Yamaguchi, K.; Yamaguchi, T.; et al. Mosapride citrate, a novel 5-HT4 agonist and partial 5-HT3 antagonist, ameliorates constipation in parkinsonian patients. Mov. Disord. 2005, 20, 680–686. [Google Scholar] [CrossRef]
Table 2. List and main characteristics of preclinical and clinical studies involving the 5-HT4 receptor in cognitive disorders.
Table 2. List and main characteristics of preclinical and clinical studies involving the 5-HT4 receptor in cognitive disorders.
Goal of the Study Methods of 5-HT4 Receptor InvestigationModels Main ResultsReference
5-HT4 receptor expressionPET imagingSubjects with early Alzheimer’s diseaseIjms 25 05245 i001In global brain and positively correlated with beta-amyloid accumulation[81]
Healthy young adult individualsInverse relationship with memory performance [76]
Inverse relationship with positive and neutral word recall performance [77]
Depressed patientsIjms 25 05245 i002in global brain * associated with memory dysfunction[67]
Western blot Rat model of Parkinson’s disease Ijms 25 05245 i001in dorsal hippocampus[82]
5-HT4 receptor modulation5-HT4 receptor agonistMacaque young and old)Learning improved [73]
MouseMemory improved[74]
Mouse model of scopolamine-induced deficitMemory deficits counteracted[46]
Mouse model of Parkinson’s disease Restored facilitation of contextual fear extinction[83]
Rat model of Parkinson’s diseaseWorking memory improved[82]
Rat model of Alzheimer’s diseaseMemory deficits and associated mechanisms counteracted[84]
Healthy young adult individualsPro-cognitive effects induced[78]
Mnesic performance improved associated with increased hippocampal activity[79]
5-HT4 receptor agonist and antagonistePet1-cre; Ai32 mouseMemory formation modulated[75]
* in comparison with healthy individuals. Up arrows indicate an increase, down arrow indicates a decrease.
Table 3. List and main characteristics of preclinical and clinical studies involving the 5-HT4 receptor in movement disorders.
Table 3. List and main characteristics of preclinical and clinical studies involving the 5-HT4 receptor in movement disorders.
Goal of the Study Methods of 5-HT4 Receptor InvestigationModels Main ResultsReference
5-HT4 receptor expressionBinding assaysPatients affected by Huntington’s diseaseIjms 25 05245 i002in putamen *[92,93]
Patients affected by Parkinson’s diseaseIjms 25 05245 i003in putamen and nigra *[92,93]
In situ hybridizationRat model of Parkinson’s disease Ijms 25 05245 i003in striatum[90]
PET imaging immunohistochemistryIjms 25 05245 i001in motor striatum[91]
5-HT4 receptor modulation5-HT4 receptor agonist Patients affected by Parkinson’s diseaseMotor functions improved[94]
Mouse model of Parkinson’s diseaseNo motor improvement[83]
Motor deficits improved associated with rescued striatal dopamine levels[89]
5-HT4 receptor antagonistRat model of amphetamine-induced turning behaviorNo impact[88]
* in comparison with healthy individuals. Up arrow indicates an increase, down arrow a decrease. Horizontal arrows indicate no change.
Table 4. List and main characteristics of preclinical and clinical studies involving the 5-HT4 receptor in food disorders.
Table 4. List and main characteristics of preclinical and clinical studies involving the 5-HT4 receptor in food disorders.
Goal of the Study Methods of 5-HT4 Receptor InvestigationModels Main ResultsReference
5-HT4 receptor expressionAutoradiographyRat model of obesityIjms 25 05245 i001In nucleus accumbens[97]
PET imagingNormal to high body mass index individualsIjms 25 05245 i001In nucleus accumbens, ventral pallidum, left hippocampus, and orbitofrontal cortex, associated with obesity[98]
Deletion MouseIjms 25 05245 i001Food intake[96]
OverexpressionMouseIjms 25 05245 i002Food intake[96]
Food intake modulatedKO mouseIjms 25 05245 i002Food intake[96]
5-HT4 receptor modulation5-HT4 receptor agonist or antagonistMouseFood intake modulated[95]
MacaqueModulated food intake personal communications
5-HT4 receptor agonistMacaque model of Parkinson’s diseaseIjms 25 05245 i002Food intakepersonal communications
Up arrows indicate an increase, down arrows a decrease.
Table 5. List and main characteristics of preclinical and clinical studies involving the 5-HT4 receptor in gastrointestinal disorders.
Table 5. List and main characteristics of preclinical and clinical studies involving the 5-HT4 receptor in gastrointestinal disorders.
Goal of the Study Methods of
5-HT4 Receptor
Investigation
Models Main ResultsReference
5-HT4 receptor expressionELISA, qRT-PCR, immunohistochemistry, Western blotRat model of constipation Ijms 25 05245 i001In colon after sacral nerve electrical stimulation[103]
Western blotRat model of Parkinson’s disease Ijms 25 05245 i002In colon, associated with constipation[106]
5-HT4 receptor modulation5-HT4 receptor agonistPatients affected by Parkinson’s diseaseIjms 25 05245 i001Intestinal transit[108]
Ijms 25 05245 i001Gastric motility[94]
Rat model of Parkinson’s disease Ijms 25 05245 i001In colon[106]
Ijms 25 05245 i002Constipation
Mouse model of Parkinson’s diseaseBeneficial effect on intestinal barrier integrity[89]
5-HT4 receptor agonist and antagonistNaïve or KO mouseModulated gut transit and colonic motility [102]
Up arrows indicate an increase, down arrows a decrease.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sgambato, V. The Serotonin 4 Receptor Subtype: A Target of Particular Interest, Especially for Brain Disorders. Int. J. Mol. Sci. 2024, 25, 5245. https://doi.org/10.3390/ijms25105245

AMA Style

Sgambato V. The Serotonin 4 Receptor Subtype: A Target of Particular Interest, Especially for Brain Disorders. International Journal of Molecular Sciences. 2024; 25(10):5245. https://doi.org/10.3390/ijms25105245

Chicago/Turabian Style

Sgambato, Véronique. 2024. "The Serotonin 4 Receptor Subtype: A Target of Particular Interest, Especially for Brain Disorders" International Journal of Molecular Sciences 25, no. 10: 5245. https://doi.org/10.3390/ijms25105245

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop