Skip to main content
UKPMC Funders Author Manuscripts logoLink to UKPMC Funders Author Manuscripts
. Author manuscript; available in PMC: 2007 Oct 1.
Published in final edited form as: Expert Opin Ther Targets. 2007 Apr;11(4):527–540. doi: 10.1517/14728222.11.4.527

The 5-HT3 receptor as a therapeutic target

Andrew J Thompson , Sarah CR Lummis †,
PMCID: PMC1994432  EMSID: UKMS960  PMID: 17373882

Abstract

The 5-HT3 receptor is a neurotransmitter-gated ion channel. It is a member of the Cys-loop family of receptors, which also includes nicotinic acetylcholine, glycine and GABAA receptors. Each member of the family consists of an arrangement of five subunits surrounding a central ion-conducting pore. The 5-HT3 receptor binding site is composed of six loops from two adjacent subunits, and the critical ligand binding residues within these loops are well documented. There are a range of 5-HT3 receptor agonists and competitive antagonists, but it is the antagonists that dominate their clinical use. Studies have proposed a range of disease symptoms that might be amenable to 5-HT3 receptor selective compounds; however, so far only the treatment of emesis and irritable bowel syndrome have been fully realised. In this review, the authors look at the structure, function and distribution of 5-HT3 receptors and how this may influence their role in disease. The authors also describe the existing clinical applications of 5-HT3 antagonists and the future potential of these drugs.

Keywords: 5-HT3, alosetron, Cys-loop receptor, dolasetron, emesis, granisetron, irritable bowel syndrome, ondansetron, palonosetron, ramosetron, tropisetron

1. Introduction

The 5-HT3 receptor is an ionotropic ligand-gated ion channel (LGIC) and thereby differs from other serotonin receptors (5-HT1 to 5-HT7) whose actions are mediated via G-proteins. The structure and function of 5-HT3 receptors shows they are members of the Cys-loop family of LGICs, which includes glycine, GABAA and nicotinic acetylcholine (nACh) receptors. Members of this family share a structure that is composed of five pseudo-symmetrically arranged subunits surrounding a central ion-conducting pore. Each subunit is composed of an extracellular, a transmembrane and an intracellular domain (Figure 1). The extracellular domain contains the binding site for agonists and competitive antagonists; it is the major therapeutic target in 5-HT3 receptors and the site of action of all of the drugs discussed in this review. The binding site is formed at the interface of two adjacent subunits by the convergence of three amino acid loops (A – C) from one (the principal) subunit and three β-strands (D – F) from the adjacent (or complementary) subunit. The transmembrane region contains four membrane-spanning α-helices (M1 – M4) and a short C-terminus. M2 from each subunit lines the pore and contains regions responsible for channel gating and ion selectivity. In 5-HT3 receptors, this pore is predominantly sodium and potassium selective, and its opening results in a rapidly activating and then desensitising inward current [1,2]. A large loop between M3 and M4 forms the intracellular domain and is involved in channel conductance and modulation. There is evidence that parts of the transmembrane and intracellular regions may be responsible for the aetiology of some 5-HT3 related pathologies (e.g., alcohol [3]), but, so far, there have been no therapeutic developments that specifically target these regions.

Figure 1. A schematic representation of a typical Cys-loop receptor subunit.

Figure 1

The diagram at the lower left is a cross-section of the transmembrane region shown from above and demonstrates how five subunits associate to form a central ion-conducting pore that is lined by M2 α-helices. Attention is drawn to six loops that form the binding ligand binding site (A – F), regions associated with ion-selectivity (dark lines either side of M2) and the region that has been shown to influence ion conductivity (R-R-R).

2. Distribution of 5-HT3 receptors

So far, genes for five 5-HT3 subunits have been identified (A – E). 5-HT3A receptor subunit mRNA is widely distributed in the adult human brain and internal organs, and has also been found in extraneuronal cells such as monocytes, T cells, synovial tissue and primary chondrocytes, which suggests a role in inflammation [4,5]. The distribution of 5-HT3B receptor subunit mRNA is not as widespread but it is still detectable across a range of adult brain regions and kidney [6]. Immunochemical studies suggest that 5-HT3B receptor subunits are either restricted to the peripheral nervous system (PNS), or exist in the CNS in low abundance or discretely localised cell populations [7-9]. A more recent study that used quantitative real-time PCR to analyse the tissue distribution of 5-HT3B described similar results, but also reported two new 5-HT3B splice variants that were almost exclusively found in the brain [10]. Interestingly, 5-HT3B appears to be expressed in anatomical structures that are involved in drug-induced emesis, although there has been no direct link between heteromeric receptors and the effects of antiemetic drugs. 5-HT3C receptor subunit mRNA also has a relatively wide distribution within adult brain, colon, intestine, lung, muscle and stomach, whereas 5-HT3D mRNA is restricted to kidney, colon and liver and 5-HT3E mRNA is restricted to the colon, intestine and stomach [9,11]. As yet, however, there is no published evidence that the genes for 5-HT3C, 5-HT3D and 5-HT3E receptor subunits are transcribed, and thus it may be that 5-HT3 receptors are homomeric A-only or heteromeric A+B receptors,. There is also some evidence that 5-HT3 receptor subunits may co-express with subunits from other ligand-gated ion channels, such as the nACh α4 subunit [7,12].

5-HT3 receptors are located in many brain areas including the hippocampus, entorhinal cortex, frontal cortex, cingulate cortex, amygdala, nucleus accumbens, substantia nigra and ventral tegmental area, with highest levels in the brain stem, especially areas involved in the vomiting reflex such as the area postrema and the nucleus tractus solitarius. These brain regions are protected by the blood–brain barrier with the exception of the area postrema and the nucleus tractus solitarius. The area postrema is one of four structures in the ventricular system and, like the spinal cord, is surrounded by an ependymal layer. This layer lacks the tight junctions of the blood–brain barrier and allows the area postrema to fulfil a chemosensory role. 5-HT3 receptors have also been detected in the dorsal horn and dorsal root ganglia of the spine and in combination with the area postrema are responsible for the vomiting reflex [13-15]. Interestingly, although the levels of 5-HT3 receptors are highest in these regions, they are still low when compared with the densities of other serotonin receptors. 5-HT3 receptors are found pre- and postsynaptically and activation can modulate the release of a variety of neurotransmitters, including dopamine, cholecystokinin, GABA, substance P and acetylcholine. There also appears to be differential cellular localisation of presynaptic and postsynaptic 5-HT3 receptors within different central regions, depending on the nature of the neurons expressing them [13,16]. Consistent with their role in emesis, 5-HT3 receptors are also involved in information transfer in the gastrointestinal tract, and in the enteric nervous system they regulate gut motility and peristalsis [17]. They also play an important role in the urinary tract, and expression of hypersensitive and constitutively active 5-HT3 receptors in mice lead to excitotoxic neuronal cell death, resulting in their early death due to uropathy [18].

3. Structure of the 5-HT3 receptor

Structural details of the 5-HT3 receptor at the molecular level are unresolved, but a wealth of convergent evidence shows that the structure of these receptors is closely related to the structure of the nACh receptor (see [19,20] for reviews). Consequently, the 5-HT3 receptor is thought to be well represented by cryo-electron microscope images of the nACh receptor and by crystal structures of the acetylcholine-binding protein (AChBP), a protein that is homologous with the extracellular domain of the nACh receptor [21] (Figures 1 and 2). Chimaeric receptors that combine AChBP with the transmembrane domain of the 5-HT3 receptor can be activated by acetylcholine and further demonstrate the structural and functional similarity between these proteins [22].

Figure 2. A homology model of the extracellular, transmembrane and intracellular domains of the 5-HT3 receptor.

Figure 2

A. The receptor is shown from the side and the position of the membrane is shown as a grey box. So far, the only resolved structure within the intracellular domain of each subunit is an α-helix. B. The receptor is shown from above, looking down towards the membrane and through the central ion-conducting pore. A. and B. are homology models based on cryo-electron microscopy images of the nACh receptor at 4 Å resolution (PDB ID; 2bg9). C. A homology model of the extracellular domains of two adjacent subunits (principal and complementary). This model was based on the crystal structure of AChBP at 2.7 Å (PDB ID; 1i9b) and highlights the six loops that converge to form the ligand binding site. Only two of the five subunits have been shown for ease of viewing. 5-HT (green) and granisetron (red) are docked into the binding site. The positions of these ligands is based upon the most likely orientations taken from [23] and [24].

AChBP: Acetylcholine binding protein; 5-HT: 5-Hydroxytryptamine; nACh: Nicotinic acetylcholine receptor; PDB: Protein DataBank.

The 5-HT3 receptor is composed of five subunits that surround a central ion-conducting pore (Figures 1 and 2). The extracellular, N-terminal, domain contains the ligand binding site and crystal structures of AChBP have been used to create 5-HT3 receptor homology models of this region [23-26]. These models indicate that the ligand binding site lies at the interface of two adjacent subunits and is formed by three loops (A – C) from the ‘principal’ subunit and three β-strands (D – F) from the adjacent or ‘complementary’ subunit. A number of studies have identified key residues that are involved in both agonist and antagonist binding. As many of the 5-HT3 therapeutics are competitive inhibitors, these studies have been important in understanding the mechanisms of ligand binding. Comprehensive reviews of the 5-HT3 ligand binding site can be found in Thompson et al. [27,28].

The transmembrane domain of each 5-HT3 receptor subunit is primarily composed of four (M1 – M4) transmembrane α-helices (Figure 2) [2,29]. M2 α-helices from each subunit form an inner ring that is in direct contact with the permeating ions, and an outer ring consists of M1, M3 and M4. M2 residues that lie along one side of an α-helix line the water-accessible pore [30,31], and a kink at the centre of the M2 helices forms a hydrophobic constriction that represents the channel gate. Binding of 5-HT to its receptor causes movements within the extracellular domain that are translated to the M2 helices and open this gate. Studies of a conserved proline residue in the M2 – M3 loop of the 5-HT3 receptor show that a transition between the trans and cis configuration of this residue may provide the molecular switch that is responsible for channel opening [32]. Compounds such as anaesthetics and n-alcohols may directly affect this region and alter the frequency of open time events (see below). Residues within M2 are also responsible for ion selectivity, as a ring of amino acids at the intracellular side of the M2 helices has been shown to influence selectivity properties in both 5-HT3 receptor and other Cys-loop members [19]. In the 5-HT3 receptor, mutations at the extracellular side of the channel have also been implicated in charge selectivity, and it is likely that charged amino acids at both the intracellular and extracellular sides of the pore concentrate the relevant ions before they pass through the channel (Figure 1) [33]. A possible therapeutic role for these charged residues is discussed in Section 5 of this review.

An increasing number of compounds are being identified which may act via the transmembrane domain of 5-HT3 receptors. Picrotoxin, for example, a classic GABAA receptor antagonist, blocks the channel, and binding has been shown to be affected by mutations at a site close to the channel gate [34]. The hypertensive drug, diltiazem, which blocks voltage-gated calcium channels, is also known to block the 5-HT3 receptor channel, highlighting the common mechanisms that many of these drugs share, and also the promiscuity that many of these compounds display. A wide range of substances, including alcohols, steroids and anaesthetics have also been reported to modulate 5-HT3 receptors in a non-competitive fashion. Given their hydrophobicity, it is likely that the actions of these compounds are at binding sites located within the membrane, although their mechanisms of action are largely unknown. Volatile anaesthetics and n-alcohols with small carbon chain lengths enhance the function of 5-HT3 receptors and become more inhibitory with increasing carbon chain length. This dependency on molecular volume indicates that there is a binding pocket of limited size, and the similarities in their behaviour suggests that these agents act at the same site [35]. A study of residues in the M2 – M3 loop has attempted to identify this binding site. Although it was shown that the modulatory effects of n-alcohols and anaesthetics can be altered by mutations here, the authors concluded that they did not represent a binding site for these agents [36,37]. Interestingly, the effects of n-alcohols and anaesthetics are reduced in heteromeric 5-HT3AB receptors [38].

The intracellular domain is formed by a loop of ∼ 110 residues between M3 and M4. The structure of this domain remains uncertain, but functionally it has a role in channel conductance and receptor modulation. Homomeric 5-HT3 receptors composed of A-subunits alone form functional channels with a conductance that is so small (sub-pS) that it cannot be resolved directly. Although the B-subunit cannot form homomeric channels, it can be combined with A-subunits to generate functional heteromeric receptors that display a much larger conductance (9 – 17 pS) [6,39]. This difference is the consequence of three arginine residues that lie within an α-helix in the M3 – M4 loop (Figure 1) [40]. The intracellular domain is also known to modulate 5-HT3 receptor function as a result of post-translational modifications. The effects of these modifications have little therapeutic significance and are discussed in Thompson et al. [19]. Post-translational modification of the extracellular domain has been shown to be responsible for cell surface expression and calcium permeability, but it is unlikely to have an impact on future clinical developments [41,42].

4. Therapeutic uses of 5-HT3 antagonists

Five 5-HT3 antagonists are available for clinical use in Europe at present. These are tropisetron, ondansetron, granisetron, dolasetron and palonosetron. Others include azasetron and ramosetron, which are available in the Far East and alosetron, which has been approved by the FDA for the treatment of irritable bowel syndrome in the US (Table 1 and Figure 3). Owing to the unfavourable effects of 5-HT3 agonists (e.g., nausea and anxiety) no clinical use of these is likely in the near future.

Table 1.

Kd, Ki and IC50 values for a range of 5-HT3 receptor antagonists

Antagonist Kd, Ki or IC50 Species Reference
In present clinical use:
Granisetron 230 pM N1E-115 [127]
Granisetron 2.01 nM* Mouse [128]
Granisetron 2.43 nM NG 108-15 [128]
Granisetron 410 pM N1E-115 [129]
Granisetron 1.44 nM* Human [130]
Granisetron 5.13 nM Rat brain homogenate [131]
Tropesetron 3.80 nM* Mouse [128]
Tropisetron 3.85 nM NG 108-15 [128]
Tropisetron 46.0 pM§ Rabbit nodose ganglion [132]
Tropisetron 4.90 nM Rat brain homogenate [131]
Tropisetron 11.0 nM§* Human [53]
Ondansetron 34.3 pM NG 108-15 [128]
Ondansetron 4.03 nM N1E-115 [129]
Ondansetron 4.90 nM* Human [130]
Ondansetron 57.0 pM§ Rabbit nodose ganglion [132]
Ondansetron 7.40 nM N1E-115 [127]
Ondansetron 46.8 nM Rat brain homogenate [131]
Palonosetron 31.6pM NG 108-15 [133]
Palonosetron 31.6 pM Rat brain [133]
Palonosetron 5.01 nM Guinea-pig ileum [133]
Dolasetron 20.0 nM NG 108-15 [134]
Azasetron 0.33 nM Rat small intestine [135]
Alosetron 3.16 nM Rat brain homogenate [131]
Alosetron 398 pM* Human [136]
Alosetron 158 pM Rat brain homogenate [136]
Ramosetron 0.15 nM* Human [137]
Not in present clinical use:
Bemesetron 328 pM§ Rabbit nodose ganglion [132]
LY-278,584 5.00 nM§* Human [53]
Y-25130 36.0 nM§* Human [53]
MDL-72222 16.0 nM N1E-115 [127]
MDL-72222 30.2 nM Rat brain homogenate [131]
BRL-46470 150 pM NG 108-15 [128]
BRL-46470 1.58 nM Rat brain homogenate [131]
ICS-205-930 640 pM N1E-115 [127]
Quipazine 510 pM N1E-115 [129]
Quipazine 1.00 nM N1E-115 [127]
Quipazine 1.10 nM Rat brain homogenate [131]
GR-65630 2.50 nM N1E-115 [127]
SDZ 206-830 871 pM Rat brain homogenate [131]
(S)-zacopride 955 pM Rat brain homogenate [131]
(R)-zacopride 11.0 nM Rat brain homogenate [131]
Renzapride 67.6 nM Rat brain homogenate [131]
Clozapine 269 nM Rat brain homogenate [131]
2-(4-methyl-1-piperazine)cyclohexa[c]
quinoline
230 pM Rat brain homogenate [138]
Indisetron 1.70 nM Rat brain homogenate [139]
Lerisetron 0.80 nM* Mouse [140]
Cilansetron 0.19 nM Rat brain homogenate [141]
*

Recombinantly expressed in cells.

Note that quipazine has been classified as both an agonist and antagonist.

§

IC50 values, calculated using electrophysiological techniques.

Figure 3. Molecular structures of commercially available 5-HT3 receptor antagonists.

Figure 3

As a consequence of their potentially different subunit combinations and their varied tissue-specific distribution, it might be anticipated that 5-HT3 receptors would provide a wide scope for novel therapeutic targets. Indeed, studies have revealed a diversity of potential disease targets that might be amenable to alleviation by 5-HT3 receptor-selective compounds, the majority of which also have the advantage of being able to cross the blood–brain barrier [43,44]. Such disease targets include addiction, pruritis, emesis, fibromyalgia, migraine, rheumatic diseases and neurological phenomena such as anxiety, psychosis, nociception and cognitive function. Other possible targets are chronic heart pain and bulimia. Fortunately, despite a range of actions, 5-HT3 receptor antagonists do not appear to alter normal behaviour in animal models, and the only typical physiological changes in clinical volunteers are mild effects on gut transit, constipation, headache, dizziness and clinically insignificant asymptomatic changes in cardiovascular behaviour [45]. All of these effects are reversible after termination of the drug. For further reading on a number of these therapeutic applications, a series of reviews can be found in [46]. Although these reviews were first published in 1994, many of the discussions still apply today.

4.1 Emesis

At present, 5-HT3 antagonists are primarily used for controlling chemotherapy- and radiotherapy-induced nausea and vomiting (CINV) and in postoperative nausea and vomiting (PONV). In combination with substances such as corticosteroids (e.g., dexamethasone), they are important for treating acute and delayed symptoms of these therapies. The introduction of new, more potent, 5-HT3 antagonists such as palonosetron, has further improved the treatment of these symptoms, and in combination with corticosteroids has been shown to have an improved long-term benefit compared with some of the established 5-HT3 antagonists [47]. There is also clinical evidence that 5-HT3 receptor antagonists could be useful for the alleviation of vomiting during pregnancy and following caesarean section [48,49]. It is believed that vomiting occurs because of the release of serotonin from enterochromaffin cells of the intestinal mucosa, which results in the stimulation of peripheral 5-HT3 receptors in the adjacent vagal afferent neurons [50]. This effect is coincidental with a local release of 5-HT in the area postrema, located on the dorsal surface of the medulla elongata, and the actions at both locations triggers the vomiting reflex. The therapeutic effects result from inhibition of this vomiting reflex. Interestingly, as the area postrema lacks a blood–brain diffusion barrier, it is able to detect emetic toxins in the blood, as well as in the cerebrospinal fluid. However, circulating substances have not been shown to directly trigger the emetic response, which appears to be due to depolarisation of the vagal afferent nerves that terminate in this brainstem region [50]. For this reason, the use of 5-HT3 antagonists for relieving vomiting caused by intoxication has not been pursued to any great extent. It has been suggested that the 5-HT3B receptor subunit may play an important contribution to the effectiveness of these compounds and a study of polymorphisms has shown a positive link between a mutation in the promoter region of the 5-HT3B gene and the frequency of vomiting [51]. However, it must be stressed that the pharmacology of homomeric and heteromeric receptors is not hugely different and other studies have found no link between polymorphisms and pharmacological responses [52-55]. It is also noteworthy that another potential postoperative use of 5-HT3 receptor antagonists is for the prevention of pain during the injection of anaesthetics and for postoperative shivering. Studies have shown that dolasetron is as effective as the local anaesthetic lidocaine at preventing pain, but may not be as effective in preventing shivering [56,57]. Further information regarding the clinical application of 5-HT3 antagonist in CINV and PONV can be found in [45], and guidelines for the clinical use of these compounds can be found in [58-60].

4.2 Irritable bowel syndrome and intestinal effects

The 5-HT3 receptor antagonist alosetron is an effective treatment for irritable bowel syndrome as it decreases gut transit [61], increases fluid absorption [62] and reduces pain in irritable bowel syndrome patients [63-65]. Although desirable for irritable bowel syndrome patients, these effects probably contribute to the constipation experienced by some individuals undertaking 5-HT3 antagonist-based therapy for other diseases. During clinical trials the use of alosetron was shown to be an efficient compound for the effective treatment of female patients with irritable bowel syndrome but post-marketing surveillance revealed several adverse reactions including reports of severe constipation or ischaemic colitis and even death. Consequently, the drug was voluntarily withdrawn in November 2000, before being reintroduced in 2002 for patients for whom the benefit-to-risk balance was favourable, and who did not respond adequately to conventional treatment [66]. This incident was unusual for 5-HT3 antagonists which are usually well tolerated but highlights that if a symptom is not life threatening, the side effects must be resolved if the ameliorative capacity of these drugs is required for long-term use [67,68]. Future treatment of irritable bowel syndrome with 5-HT3 antagonists may depend on combination with other therapies (e.g., tegaserod, a 5-HT4 antagonist) in order to target multiple sites of action [69-71].

4.3 Schizophrenia, anxiety and other neurological disorders

Serotonergic neurons have a regional distribution in brain areas implicated in a range of neurological phenomenon and there has been much interest in the therapeutic potential of 5-HT3 receptor antagonists for antipsychotic, antinociceptive and other psychiatric disorders. As 5-HT3 antagonists freely pass the blood–brain barrier, these compounds appear to be ideal therapeutic candidates but, so far, this potential has not been realised [72]. The theory of 5-HT involvement in schizophrenia and bipolar disorder was first suggested in the mid-1950s, and proposed that there was a serotonergic deficiency in schizophrenic individuals [73,74]. Serotonin receptors have been implicated in many of the symptoms of schizophrenia and are prime candidates because of their functional diversity and their ability to modulate the release of other neurotransmitters such as dopamine, GABA, substance P and acetylcholine. So far, the focus of 5-HT and its impact on schizophrenia has largely been on the G-protein-coupled 5-HT2 receptors, but the administration of selective 5-HT3 antagonists such as ondansetron and tropisetron has been shown to improve P50 auditory gating in schizophrenic patients, and may also have a therapeutic use for the neurocognitive deficits of this disorder [75,76]. Further evidence comes from a recent study that identified two 5-HT3 sequence variations (R344H and P391R) in a small group of patients with bipolar disorder and schizophrenia [77]. The two mutations were located in a functionally important region, the M3 – M4 loop, which contains a number of potential phosphorylation sites and also has an important role in channel conductance and ion selectivity. However, the rarity of these mutations (single mutations in 2 individuals from a study of 428) and emerging electrophysiological evidence indicates that they are unlikely to be a major contributor to schizophrenia [78,79]. Association analysis of other polymorphisms has also revealed no link between 5-HT3 receptor genes and polymorphisms [54,55].

So far, the use of 5-HT3 antagonists for the clinical treatment of other psychological disorders has also met with little success, although there is experimental evidence that the 5-HT3 receptor plays a role, suggesting that further studies are worthwhile. For example, the deletion of the 5-HT3 receptor gene creates knockout mice that exhibit anxiolytic behaviour [80,81] and the use of 5-HT3 receptor antagonists has shown a range of anxiolytic effects [82]. Studies on 5-HT3 receptors have shown that the actions of antagonists such as mirtazapine and clozapine are competitive, and it has been shown that the enrichment of non-competitive antidepressants in cell surface lipid microdomains may be crucial for their effects [83]. Some antipsychotic drugs, such as chlorpromazine and related phenothiazines, also appear to act directly at the 5-HT3 receptor binding site [84]. However, many antipsychotic drugs are known to have a broad spectrum of activity, only some of which can be attributed to affects on the 5-HT3 receptor. For example, clozapine affects dopaminergic transmission but its actions at the 5-HT3 receptor may account for this drugs unique antipsychotic efficiency [85,86]. However, 5-HT2, dopamine and serotonin-selective re-uptake inhibitors are among the most widely prescribed drugs for these disorders and it is likely that these antipsychotic drugs will continue to have a more significant impact. It is also possible that many of the ameliorative effects of the 5-HT3 antagonists result from the serotonin-mediated dopamine, cholecystokinin or GABA responses, or through actions on other 5-HT receptor types. Indeed, a review by Olivier et al. [87] concluded that 5-HT3 antagonists are active in a number of animal models, are well tolerated in the long-term and appear to have no appreciable side effects, but due to the large body of often contradicting results, it is often difficult to interpret their effectiveness. It may be that the benefit of these chemicals may only be realised when used as a combination therapy. Interestingly, a number of antidepressant drugs have also been used for gastrointestinal disorders and other uses may eventually come to light [88].

4.4 Cognitive function

Cognition describes aspects of behaviour such as awareness, perception, reasoning and memory. The cortex and dorsal hippocampus are both important memory-related structures and antagonism of the 5-HT3 receptor at these locations inhibits the 5-HT modulated-release of acetylcholine without affecting the steady-state release. 5-HT3 receptor over-expressing mice have been shown to have enhanced learning, memory and attention, and ondansetron has been found to improve memory in patients > 50 years of age [89,90]. Interestingly, a polymorphism (C178T) in the regulatory region of the 5-HT3A receptor subunit has been linked to reduced activity in the amygdale and dorsal and medial prefrontal cortices, and was associated with a reduced reaction time at face recognition [91]. The same mutation has also been associated with increased susceptibility to bipolar disorder [92]. However, memory appears to be multifactoral and, like other neurological disorders, is likely to involve a range of receptors that may require a cocktail of drugs to alleviate symptoms. For example, the administration of 5-HT2A/2C or 5-HT4 receptor agonists or 5-HT1A or 5-HT3 and 5-HT1B receptor antagonists retards memory impairment and promotes learning in tasks that require a high cognitive demand [93].

4.5 Substance abuse and addiction

Using 5-HT3 antagonists for the alleviation of substance abuse has had some success. Antagonists are particularly effective at reducing ethanol and morphine self-administration but are less effective at reducing the self-administration of psychostimulants such as cocaine [94-96]. It has been shown that with the administration of ondansetron, alcohol craving is significantly reduced in early onset alcoholics but increases craving in late onset alcoholics [97]. It is believed that this effect may be the result of altered 5-HT3 modulation of dopamine release. Interestingly, substance abuse is particularly high among patients suffering from schizophrenia, suggesting a possible link between the systems that modulate these responses [98]. Co-expression of the B-subunit has been shown to reduce alcohol sensitivity in recombinant expressed 5-HT3 receptors [99].

4.6 Bulimia

It has been shown that increased vagal afferent nerve activity is associated with binge-eating and vomiting and can be suppressed by the use of the 5-HT3 antagonist ondansetron [100]. The depressive symptoms of these patients were also reduced. These findings are supported by evidence that the 5-HT3 antagonist m-chlorophenylpiperazine has also been shown to improve mood and patient perception of body image, although these affects may be compounded by actions at other 5-HT receptors [101].

4.7 Pruritis

Pruritis is the medical term for itching and can be the result of rashes caused by burns, infection and other local irritations, or can display systemic symptoms as a consequence of renal, hepatic, hematopoietic or endocrine pathologies. It has been suggested that 5-HT3 antagonists may have an antipruritic effect but current research reveals mixed reports in this area, and the effectiveness of treatment may vary according to the type of pruritus studied [102]. For example, in patients with cholestatic itch, either some or no benefit has been reported [103,104], whilst only marginal or no relief has been reported for haemodialysis-related pruritus [105,106]. However, the underlying mechanisms of this disorder are still poorly understood and will need further work if a therapeutic potential is to be realised [107].

4.8 Analgesics and anti-inflammatory actions

Pain results from the activation of sensory nociceptors (sensory pain), or as the consequence of damage to peripheral and central nerves (neuropathic pain). 5-HT3 receptors are located in pain-related regions and research has shown their involvement in pain processing and inflammation. Following tissue injury, the mechanism of pain and inflammation appears to be complex. Symptoms have been attributed to the 5-HT-mediated release of neuropeptides such as substance P, and 5-HT3 receptors are expressed in the immune system where activation leads to T cell activation and the secretion of cytokines and prostaglandins [5,108].

Fibromyalgia is a chronic pain illness characterised by widespread aches, pain, stiffness, tenderness, general fatigue and sleep disturbances. There is clinical evidence that the 5-HT3 antagonists granisetron, ondansetron and tropisetron can significantly reduce the effects of fibromyalgia when administered systemically [109-113]. Studies have also been performed on symptoms such as chronic lower back pain and arthritis [114,115]. A localised injection of tropesetron has also been shown to reduce pain, and the effect was longer lasting than a comparable injection of local anaesthetic [116]. The use of the 5-HT3 antagonist dolasetron has also been used for the reduction of local pain during the injection of anaesthetics [57]. The mechanisms for these local and systemic effects have been reviewed in Riering et al. [117] and Giordano et al. [118].

As well as its use in chemotherapy, methotrexate is used to treat several different types of rheumatic disease. However, as the effects of this drug can only be seen 3 – 12 weeks after first use, the emergence of nausea in some patients is of importance. Suppression of this side effect could potentially be accomplished using 5-HT3 receptor antagonist in the same way as they are used for CINV and PONV [119]. The effects of 5-HT3 antagonists on the pain relieving properties of acetylsylic acid (aspirin or acetosal), acetaminophen (paracetamol) may also be important. For example, co-administration of tropisetron or granisetron with acetaminophen completely blocks the analgesic effect of acetaminophen but ondansetron does not affect the actions of acetylsylic acid [120-122].

5. Expert opinion

So far, 5-HT3 receptor-based therapy has depended entirely on high-affinity competitive antagonists. The two main therapeutic applications for these have included their use as antemetics and for relieving the symptoms of irritable bowel syndrome. Other applications have been considered and a number of clinical trials have been conducted to assess their potential. However, the complex nature of some of the pathological symptoms, the difficulty in assessing patient benefit and the presence of established alternative drugs has limited their use in the clinic.

An interesting and potentially widespread application for 5-HT3 receptor antagonists in the future is their capacity to reduce pain. It has been shown that the systemic administration of the compounds has beneficial affects for patients suffering from fibromyalgia and the side effects of these compounds are few and often inconsequential. However, their effect at both central and peripheral 5-HT3 receptors introduces complex pharmacokinetic variability and may limit their clinical use. A more exciting development is the local administration of these drugs by injection or cream, both of which have been shown to have a measurable impact on pain reduction. This may include applications as diverse as alleviating the pain-related symptoms of tissue injury or arthritis. Whether or not these applications are successful will largely depend on further research to prove their effectiveness and the cost savings that these drugs can provide.

Hopefully, future studies will give us a better understanding of the promiscuous nature of some of the existing 5-HT3 antagonists, as their targeting of multiple receptors can produce complex behaviours, the effects of which can be counterproductive. The development of more specific ligands may also allow a more directed approach, while further improvements in drug half-life should enhance their long-term effectiveness. At present, little is known about the physiological role of the five 5-HT3 receptor subunits, and research in this area may lead to novel therapeutic interventions, particularly if subunit-specific antagonists can be found. It also seems increasingly likely that developments, particularly in the treatment of psychological disorders, will include combination therapies in which 5-HT3 antagonists are only a part of the overall treatment.

So far, there has not been any development of compounds that modulate receptor function by intracellular modulation and it is hard to imagine how this approach could be accomplished without having detrimental affects on other cellular responses. However, as the development costs of new therapeutics continue to rise, we may find new uses for existing compounds. For example, there is evidence that antimalarial compounds such as quinine, chloroquine and mefloquine are antagonists at a number of Cys-loop receptors [123,124]. These drugs have a good clinical record and they may turn out to have measurable benefits in the treatment of 5-HT3 receptor-related disorders. Future therapeutic applications might also use parts of the 5-HT3 receptor, although this approach has not been pursued as yet. For example, Broughman et al. [125,126] has found that M2 segments of the glycine receptor spontaneously self assemble to form chloride-permeable channels when exposed to the cell surface and could be used an effective means of channel replacement therapy by restoring chloride permeability in cystic fibrosis patients. Charged residues close to M2 are known to be responsible for ion selectivity in the 5-HT3 receptor and manipulation of these, or similar peptides, may allow the development of novel therapeutics tailored for specific clinical uses.

5-HT3 receptors are well understood in terms of their distribution, structure, function and pharmacology. However, there is still some way to go in order to understand their roles in both the CNS and the PNS, and a better knowledge of this might lead to more areas for therapeutic intervention by 5-HT3 receptor agonists, antagonists and modulators. At present, 5-HT3 receptor antagonists are proving to be useful agents for controlling chemotherapy-induced emesis and in irritable bowel syndrome, but as studies suggest there is considerable potential for therapeutic intervention in other areas, the authors anticipate that there will be further developments in their clinical use.

Acknowledgements

Work carried out in the authors' laboratory is supported by the Wellcome Trust. SCR Lummis is a Wellcome Trust Senior Research Fellow in Basic Biomedical Science.

Bibliography

Papers of special note have been highlighted as either of interest (•) or of considerable interest (••) to readers.

  • 1.DERKACH V, SURPRENANT A, NORTH RA. 5-HT3 receptors are membrane ion channels. Nature. 1989;339(6227):706–709. doi: 10.1038/339706a0. [DOI] [PubMed] [Google Scholar]
  • 2.MARICQ AV, PETERSON AS, BRAKE AJ, MYERS RM, JULIUS D. Primary structure and functional expression of the 5HT3 receptor, a serotonin-gated ion channel. Science. 1991;254(5030):432–437. doi: 10.1126/science.1718042. • First cloned 5-HT3 receptor subunit. [DOI] [PubMed] [Google Scholar]
  • 3.DAVIES DL, ASATRYAN L, KUO ST, et al. Effects of ethanol on adenosine 5′-triphosphate-gated purinergic and 5-hydroxytryptamine receptors. Alcohol. Clin. Exp. Res. 2006;30(2):349–358. doi: 10.1111/j.1530-0277.2006.00023.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.MIYAKE A, MOCHIZUKI S, TAKEMOTO Y, AKUZAWA S. Molecular cloning of human 5-hydroxytryptamine3 receptor: heterogeneity in distribution and function among species. Mol. Pharmacol. 1995;48(3):407–416. [PubMed] [Google Scholar]
  • 5.FIEBICH BL, AKUNDI RS, SEIDEL M, et al. Expression of 5-HT3A receptors in cells of the immune system. Scand. J. Rheumatol. 2004;33(Suppl. 119):9–11. [PubMed] [Google Scholar]
  • 6.DAVIES PA, PISTIS M, HANNA MC, et al. The 5-HT3B subunit is a major determinant of serotonin-receptor function. Nature. 1999;397(6717):359–363. doi: 10.1038/16941. [DOI] [PubMed] [Google Scholar]
  • 7.SUDWEEKS SN, HOOFT JA, YAKEL JL. Serotonin 5-HT3 receptors in rat CA1 hippocampal interneurons: functional and molecular characterization. J. Physiol. 2002;544(Pt. 3):715–726. doi: 10.1113/jphysiol.2002.029736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.MORALES M, WANG SD. Differential composition of 5-hydroxytryptamine3 receptors synthesized in the rat CNS and peripheral nervous system. J. Neurosci. 2002;22(15):6732–6741. doi: 10.1523/JNEUROSCI.22-15-06732.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.NIESLER B, FRANK B, KAPELLER J, RAPPOLD GA. Cloning, physical mapping and expression analysis of the human 5-HT3 serotonin receptor-like genes HTR3C, HTR3D and HTR3E. Gene. 2003;310:101–111. doi: 10.1016/s0378-1119(03)00503-1. [DOI] [PubMed] [Google Scholar]
  • 10.TZVETKOV MV, MEINEKE C, OETJEN E, HIRSCH-ERNST K, BROCKMOLLER J. Tissue-specific alternative promoters of the serotonin receptor gene HTR3B in human brain and intestine. Gene. 2007;386(1-2):52–62. doi: 10.1016/j.gene.2006.08.002. [DOI] [PubMed] [Google Scholar]
  • 11.KARNOVSKY AM, GOTOW LF, MCKINLEY DD, et al. A cluster of novel serotonin receptor 3-like genes on human chromosome 3. Gene. 2003;319:137–148. doi: 10.1016/s0378-1119(03)00803-5. [DOI] [PubMed] [Google Scholar]
  • 12.VAN HOOFT JA, SPIER AD, YAKEL JL, LUMMIS SC, VIJVERBERG HP. Promiscuous coassembly of serotonin 5-HT3 and nicotinic α4 receptor subunits into Ca2+-permeable ion channels. Proc. Natl. Acad. Sci. USA. 1998;95(19):11456–11461. doi: 10.1073/pnas.95.19.11456. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.MIQUEL MC, EMERIT MB, NOSJEAN A, et al. Differential subcellular localization of the 5-HT3-AS receptor subunit in the rat central nervous system. Eur. J. Neurosci. 2002;15(3):449–457. doi: 10.1046/j.0953-816x.2001.01872.x. [DOI] [PubMed] [Google Scholar]
  • 14.TECOTT LH, MARICQ AV, JULIUS D. Nervous system distribution of the serotonin 5-HT3 receptor mRNA. Proc. Natl. Acad. Sci. USA. 1993;90(4):1430–1434. doi: 10.1073/pnas.90.4.1430. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.KIA HK, MIQUEL MC, MCKERNAN RM, et al. Localization of 5-HT3 receptors in the rat spinal cord: immunohistochemistry and in situ hybridization. Neuroreport. 1995;6(2):257–261. doi: 10.1097/00001756-199501000-00008. [DOI] [PubMed] [Google Scholar]
  • 16.HUANG J, SPIER AD, PICKEL VM. 5-HT3A receptor subunits in the rat medial nucleus of the solitary tract: subcellular distribution and relation to the serotonin transporter. Brain Res. 2004;1028(2):156–169. doi: 10.1016/j.brainres.2004.09.009. [DOI] [PubMed] [Google Scholar]
  • 17.GALLIGAN JJ. Ligand-gated ion channels in the enteric nervous system. Neurogastroenterol. Motil. 2002;14:611–623. doi: 10.1046/j.1365-2982.2002.00363.x. [DOI] [PubMed] [Google Scholar]
  • 18.BHATTACHARYA A, DANG H, ZHU QM, et al. Uropathic observations in mice expressing a constitutively active point mutation in the 5-HT3A receptor subunit. J. Neurosci. 2004;24:5537–5548. doi: 10.1523/JNEUROSCI.5658-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.THOMPSON AJ, LUMMIS SCR. The relationship between structure and function in the 5-HT3 receptor: the transmembrane domain. In: Arias HR, editor. Biological and Biophysical Aspects of Ligand-Gated Ion Channel Receptor Superfamilies. Research Signpost; Kerala, India: 2006. pp. 155–170. • Useful review of 5-HT3 receptors. [Google Scholar]
  • 20.REEVES DC, LUMMIS SC. The molecular basis of the structure and function of the 5-HT3 receptor: a model ligand-gated ion channel (review) Mol. Membr. Biol. 2002;19(1):11–26. doi: 10.1080/09687680110110048. • Good detailed structure–function review. [DOI] [PubMed] [Google Scholar]
  • 21.BREJC K, VAN DIJK WJ, KLAASSEN RV, et al. Crystal structure of an ACh-binding protein reveals the ligand-binding domain of nicotinic receptors. Nature. 2001;411(6835):269–276. doi: 10.1038/35077011. [DOI] [PubMed] [Google Scholar]
  • 22.BOUZAT C, GUMILAR F, SPITZMAUL G, et al. Coupling of agonist binding to channel gating in an ACh-binding protein linked to an ion channel. Nature. 2004;430(7002):896–900. doi: 10.1038/nature02753. [DOI] [PubMed] [Google Scholar]
  • 23.THOMPSON AJ, PRICE KL, REEVES DC, et al. Locating an antagonist in the 5-HT3 receptor binding site using modeling and radioligand binding. J. Biol. Chem. 2005;280(21):20476–20482. doi: 10.1074/jbc.M413610200. [DOI] [PubMed] [Google Scholar]
  • 24.REEVES DC, SAYED MF, CHAU PL, PRICE KL, LUMMIS SC. Prediction of 5-HT3 receptor agonist-binding residues using homology modeling. Biophys. J. 2003;84(4):2338–2344. doi: 10.1016/S0006-3495(03)75039-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.YAN D, WHITE M. Spatial orientation of the antagonist granisetron in the ligand-binding site of the 5-HT3 receptor. Mol. Pharmacol. 2005;68(2):365–371. doi: 10.1124/mol.105.011957. [DOI] [PubMed] [Google Scholar]
  • 26.MAKSAY G, BIKADI Z, SIMONYI M. Binding interactions of antagonists with 5-hydroxytryptamine3A receptor models. J. Recept. Signal Transduct. Res. 2003;23(2-3):255–270. doi: 10.1081/rrs-120025568. [DOI] [PubMed] [Google Scholar]
  • 27.THOMPSON AJ, ZHANG L, LUMMIS SCRL. The 5-HT3 receptor. In: Roth BL, editor. The serotonin receptors: from molecular pharmacology to human therapeutics. Humana Press; New Jersey, USA: 2006. pp. 439–457. • Useful review of 5-HT3 receptors. [Google Scholar]
  • 28.THOMPSON AJ, LUMMIS SCR. 5-HT3 receptors. Curr. Pharm. Des. 2006;12:3615–3630. doi: 10.2174/138161206778522029. • Useful review of 5-HT3 receptors. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.MIYAZAWA A, FUJIYOSHI Y, UNWIN N. Structure and gating mechanism of the acetylcholine receptor pore. Nature. 2003;424(6943):949–955. doi: 10.1038/nature01748. [DOI] [PubMed] [Google Scholar]
  • 30.REEVES DC, GOREN EN, AKABAS MH, LUMMIS SC. Structural and electrostatic properties of the 5-HT3 receptor pore revealed by substituted cysteine accessibility mutagenesis. J. Biol. Chem. 2001;276(45):42035–42042. doi: 10.1074/jbc.M106066200. [DOI] [PubMed] [Google Scholar]
  • 31.KANEEZ FS, WHITE M. Patch clamp study of serotonin-gated currents via 5-HT Type 3 receptors by using a novel approach SHAM for receptor channel scanning. J. Biomed. Biotechnol. 2004;2004(1):10–15. doi: 10.1155/S1110724304302020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.LUMMIS SC, BEENE DL, LEE LW, et al. Cis-trans isomerization at a proline opens the pore of a neurotransmitter-gated ion channel. Nature. 2005;438(7065):248–252. doi: 10.1038/nature04130. [DOI] [PubMed] [Google Scholar]
  • 33.THOMPSON AJ, LUMMIS SC. A single ring of charged amino acids at one end of the pore can control ion selectivity in the 5-HT3 receptor. Br. J. Pharmacol. 2003;140(2):359–365. doi: 10.1038/sj.bjp.0705424. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.DAS P, DILLON GH. Molecular determinants of picrotoxin inhibition of 5-hydroxytryptamine Type 3 receptors. J. Pharmacol. Exp. Ther. 2005;314(1):320–328. doi: 10.1124/jpet.104.080325. [DOI] [PubMed] [Google Scholar]
  • 35.STEVENS R, RUSCH D, SOLT K, RAINES DE, DAVIES PA. Modulation of human 5-hydroxytryptamine Type 3AB receptors by volatile anesthetics and n-alcohols. J. Pharmacol. Exp. Ther. 2005;314(1):338–345. doi: 10.1124/jpet.105.085076. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.SESSOMS-SIKES JS, HAMILTON ME, LIU LX, LOVINGER DM, MACHU TK. A mutation in transmembrane domain II of the 5-hydroxytryptamine3A receptor stabilizes channel opening and alters alcohol modulatory actions. J. Pharmacol. Exp. Ther. 2003;306(2):595–604. doi: 10.1124/jpet.103.050542. [DOI] [PubMed] [Google Scholar]
  • 37.HU XQ, HAYRAPETYAN V, GADHIYA JJ, et al. Mutations of L293 in transmembrane two of the mouse 5-hydroxytryptamine3A receptor alter gating and alcohol modulatory actions. Br. J. Pharmacol. 2006;148(1):88–101. doi: 10.1038/sj.bjp.0706685. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.SOLT K, STEVENS RJ, DAVIES PA, RAINES DE. General anesthetic-induced channel gating enhancement of 5-hydroxytryptamine Type 3 receptors depends on receptor subunit composition. J. Pharmacol. Exp. Ther. 2005;315(2):771–776. doi: 10.1124/jpet.105.090621. [DOI] [PubMed] [Google Scholar]
  • 39.HUSSY N, LUKAS W, JONES KA. Functional properties of a cloned 5-hydroxytryptamine ionotropic receptor subunit: comparison with native mouse receptors. J. Physiol. (Lond) 1994;481(Pt. 2):311–323. doi: 10.1113/jphysiol.1994.sp020441. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.KELLEY SP, DUNLOP JI, KIRKNESS EF, LAMBERT JJ, PETERS JA. A cytoplasmic region determines single-channel conductance in 5-HT3 receptors. Nature. 2003;424(6946):321–324. doi: 10.1038/nature01788. [DOI] [PubMed] [Google Scholar]
  • 41.QUIRK PL, RAO S, ROTH BL, SIEGEL RE. Three putative N-glycosylation sites within the murine 5-HT3A receptor sequence affect plasma membrane targeting, ligand binding, and calcium influx in heterologous mammalian cells. J. Neurosci. Res. 2004;77(4):498–506. doi: 10.1002/jnr.20185. [DOI] [PubMed] [Google Scholar]
  • 42.MONK SA, WILLIAMS JM, HOPE AG, BARNES NM. Identification and importance of N-glycosylation of the human 5-hydroxytryptamine3A receptor subunit. Biochem. Pharmacol. 2004;68(9):1787–1796. doi: 10.1016/j.bcp.2004.06.034. [DOI] [PubMed] [Google Scholar]
  • 43.CAMPIANI G, CAPPELLI A, NACCI V, et al. Novel and highly potent 5-HT3 receptor agonists based on a pyrroloquinoxaline structure. J. Med. Chem. 1997;40(22):3670–3678. doi: 10.1021/jm970376w. [DOI] [PubMed] [Google Scholar]
  • 44.GLENNON RA, YOUNG R, DUKAT M. 5-HT3 agonist 2-methylserotonin as a training drug in discrimination studies. Pharmacol. Biochem. Behav. 1992;41(2):361–364. doi: 10.1016/0091-3057(92)90111-r. [DOI] [PubMed] [Google Scholar]
  • 45.HAUS U, SPATH M, FARBER L. Spectrum of use and tolerability of 5-HT3 receptor antagonists. Scand. J. Rheumatol. 2004;33(Suppl. 119):12–18. [PubMed] [Google Scholar]
  • 46.KING FD, JONES BJ, SANGER GJ. In: 5-Hydroxytryptamine-3 Receptor Antagonist. King FD, Jones BJ, Sanger GJ, editors. CRC Press; Florida, USA: 1994. pp. 155–181. • Old, but still useful review of 5-HT3 receptors. [Google Scholar]
  • 47.AAPRO MS, GRUNBERG SM, MANIKHAS GM, et al. A Phase III, double-blind, randomized trial of palonosetron compared with ondansetron in preventing chemotherapy-induced nausea and vomiting following highly emetogenic chemotherapy. Ann. Oncol. 2006;17(9):1441–1449. doi: 10.1093/annonc/mdl137. [DOI] [PubMed] [Google Scholar]
  • 48.FUJII Y, TANAKA H, TOYOOKA H. Granisetron prevents nausea and vomiting during spinal anaesthesia for caesarean section. Acta Anaesthesiol. Scand. 1998;42(3):312–315. doi: 10.1111/j.1399-6576.1998.tb04922.x. [DOI] [PubMed] [Google Scholar]
  • 49.EINARSON A, MALTEPE C, NAVIOZ Y, et al. The safety of ondansetron for nausea and vomiting of pregnancy: a prospective comparative study. Br. J. Gynecol. 2004;111(9):940–943. doi: 10.1111/j.1471-0528.2004.00236.x. [DOI] [PubMed] [Google Scholar]
  • 50.MINAMI M, ENDO T, HIRAFUJI M, et al. Pharmacological aspects of anticancer drug-induced emesis with emphasis on serotonin release and vagal nerve activity. Pharmacol. Ther. 2003;99(2):149–165. doi: 10.1016/s0163-7258(03)00057-3. [DOI] [PubMed] [Google Scholar]
  • 51.TREMBLAY PB, KAISER R, SEZER O, et al. Variations in the 5-hydroxytryptamine Type 3B receptor gene as predictors of the efficacy of antiemetic treatment in cancer patients. J. Clin. Oncol. 2003;21(11):2147–2155. doi: 10.1200/JCO.2003.05.164. [DOI] [PubMed] [Google Scholar]
  • 52.BRADY CA, STANFORD IM, ALI I, et al. Pharmacological comparison of human homomeric 5-HT3A receptors versus heteromeric 5-HT3A/3B receptors. Neuropharmacology. 2001;41(2):282–284. doi: 10.1016/s0028-3908(01)00074-0. [DOI] [PubMed] [Google Scholar]
  • 53.DUBIN AE, HUVAR RD, ANDREA MR, et al. The pharmacological and functional characteristics of the serotonin 5-HT3A receptor are specifically modified by a 5-HT3B receptor subunit. J. Biol. Chem. 1999;274(43):30799–30810. doi: 10.1074/jbc.274.43.30799. [DOI] [PubMed] [Google Scholar]
  • 54.KAISER R, TREMBLAY PB, SEZER O, et al. Investigation of the association between 5-HT3A receptor gene polymorphisms and efficiency of antiemetic treatment with 5-HT3 receptor antagonists. Pharmacogenetics. 2004;14(5):271–278. doi: 10.1097/00008571-200405000-00001. [DOI] [PubMed] [Google Scholar]
  • 55.GUTIERREZ B, ARRANZ MJ, HUEZO-DIAZ P, et al. Novel mutations in 5-HT3A and 5-HT3B receptor genes not associated with clozapine response. Schizophr. Res. 2002;58(1):93–97. doi: 10.1016/s0920-9964(02)00205-0. [DOI] [PubMed] [Google Scholar]
  • 56.PIPER SN, ROHM KD, MALECK WH, et al. Dolasetron for preventing postanesthetic shivering. Anesth. Analg. 2002;94(1):106–111. doi: 10.1097/00000539-200201000-00020. [DOI] [PubMed] [Google Scholar]
  • 57.PIPER SN, ROHM KD, PAPSDORF M, et al. Dolasetron reduces pain on injection of propofol. Anasthesiol. Intensivmed. Notfallmed. Schmerzther. 2002;37(9):528–531. doi: 10.1055/s-2002-33767. [DOI] [PubMed] [Google Scholar]
  • 58.GRALLA RJ, OSOBA D, KRIS MG, et al. Recommendations for the use of antiemetics: evidence-based, clinical practice guidelines. American Society of Clinical Oncology. J. Clin. Oncol. 1999;17(9):2971–2994. doi: 10.1200/JCO.1999.17.9.2971. [DOI] [PubMed] [Google Scholar]
  • 59.KRIS MG, HESKETH PJ, SOMERFIELD MR, et al. American Society of Clinical Oncology guideline for antiemetics in oncology: update 2006. J. Clin. Oncol. 2006;24(18):2932–2947. doi: 10.1200/JCO.2006.06.9591. [DOI] [PubMed] [Google Scholar]
  • 60.CARLISLE JB, STEVENSON CA. Drugs for preventing postoperative nausea and vomiting. Cochrane Database Syst. Rev. 2006;3 doi: 10.1002/14651858.CD004125.pub2. CD004125. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.JOHNSON BA, ROACHE JD, JAVORS MA, et al. Ondansetron for reduction of drinking among biologically predisposed alcoholic patients: a randomized controlled trial. JAMA. 2000;284(8):963–971. doi: 10.1001/jama.284.8.963. [DOI] [PubMed] [Google Scholar]
  • 62.MERTZ H, MORGAN V, TANNER G, et al. Regional cerebral activation in irritable bowel syndrome and control subjects with painful and nonpainful rectal distention. Gastroenterology. 2000;118(5):842–848. doi: 10.1016/s0016-5085(00)70170-3. [DOI] [PubMed] [Google Scholar]
  • 63.DELVAUX M, LOUVEL D, MAMET JP, CAMPOS-ORIOLA R, FREXINOS J. Effect of alosetron on responses to colonic distension in patients with irritable bowel syndrome. Aliment. Pharmacol. Ther. 1998;12(9):849–855. doi: 10.1046/j.1365-2036.1998.00375.x. [DOI] [PubMed] [Google Scholar]
  • 64.CROWELL MD. The role of serotonin in the pathophysiology of irritable bowel syndrome. Am. J. Manag. Care. 2001;7(8 Suppl.):S252–S260. [PubMed] [Google Scholar]
  • 65.JONES RH, HOLTMANN G, RODRIGO L, et al. Alosetron relieves pain and improves bowel function compared with mebeverine in female nonconstipated irritable bowel syndrome patients. Aliment. Pharmacol. Ther. 1999;13(11):1419–1427. doi: 10.1046/j.1365-2036.1999.00678.x. [DOI] [PubMed] [Google Scholar]
  • 66.GLAXOSMITHKLINE . Lotronex (alosetron hydrochloride) prescribing information. GlaxoSmithKline; Research Triangle Park, NC, USA: 2002. [Google Scholar]
  • 67.ANDRESEN V, HOLLERBACH S. Reassessing the benefits and risks of alosetron: what is its place in the treatment of irritable bowel syndrome? Drug Saf. 2004;27(5):283–292. doi: 10.2165/00002018-200427050-00001. [DOI] [PubMed] [Google Scholar]
  • 68.MAYER EA, BRADESI S. Alosetron and irritable bowel syndrome. Expert Opin. Pharmacother. 2003;4(11):2089–2098. doi: 10.1517/14656566.4.11.2089. [DOI] [PubMed] [Google Scholar]
  • 69.BERARDI RR. Safety and tolerability of tegaserod in irritable bowel syndrome management. J. Am. Pharm. Assoc (Wash DC) 2004;44(1):41–51. doi: 10.1331/154434504322713228. [DOI] [PubMed] [Google Scholar]
  • 70.TALLEY NJ. New and emerging treatments for irritable bowel syndrome and functional dyspepsia. Expert Opin. Emerg. Drugs. 2002;7(1):91–98. doi: 10.1517/14728214.7.1.91. [DOI] [PubMed] [Google Scholar]
  • 71.HADLEY SK, GAARDER SM. Treatment of irritable bowel syndrome. Am. Fam. Physician. 2005;72(12):2501–2506. [PubMed] [Google Scholar]
  • 72.GREENSHAW AJ, SILVERSTONE PH. The non-antiemetic uses of serotonin 5-HT3 receptor antagonists. Clinical pharmacology and therapeutic applications. Drugs. 1997;53(1):20–39. doi: 10.2165/00003495-199753010-00003. [DOI] [PubMed] [Google Scholar]
  • 73.GADDUM JH, HAMEED KA. Drugs which antagonize 5-hydrpxytryptamine. Br. J. Pharmacol. 1954;9:240–248. doi: 10.1111/j.1476-5381.1954.tb00848.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.WOOLEY DW, SHAW E. A biochemical and pharmacological suggestion about certain mental disorders. Proc. Natl. Acad. Sci. USA. 1954;40:228–231. doi: 10.1073/pnas.40.4.228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.ADLER LE, CAWTHRA EM, DONOVAN KA, et al. Improved p50 auditory gating with ondansetron in medicated schizophrenia patients. Am. J. Psychiatry. 2005;162(2):386–388. doi: 10.1176/appi.ajp.162.2.386. [DOI] [PubMed] [Google Scholar]
  • 76.KOIKE K, HASHIMOTO K, TAKAI N, et al. Tropisetron improves deficits in auditory P50 suppression in schizophrenia. Schizophr. Res. 2005;76(1):67–72. doi: 10.1016/j.schres.2004.12.016. [DOI] [PubMed] [Google Scholar]
  • 77.NIESLER B, WEISS B, FISCHER C, et al. Serotonin receptor gene HTR3A variants in schizophrenic and bipolar affective patients. Pharmacogenetics. 2001;11(1):21–27. doi: 10.1097/00008571-200102000-00003. [DOI] [PubMed] [Google Scholar]
  • 78.THOMPSON AJ, SULLIVAN NL, LUMMIS SCR. Characterisation of 5-HT3 receptor mutations identified in schizophrenic patients and expressed in HEK293 cells. J. Mol. Neurosci. 2007 doi: 10.1385/JMN:30:3:273. In Press. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.KURZWELLY D, BARANN M, KOSTANIAN A, et al. Pharmacological and electrophysiological properties of the naturally occurring Pro391Arg variant of the human 5-HT3A receptor. Pharmacogenetics. 2004;14(3):165–172. doi: 10.1097/00008571-200403000-00004. [DOI] [PubMed] [Google Scholar]
  • 80.KELLEY SP, BRATT AM, HODGE CW. Targeted gene deletion of the 5-HT3A receptor subunit produces an anxiolytic phenotype in mice. Eur. J. Pharmacol. 2003;461(1):19–25. doi: 10.1016/s0014-2999(02)02960-6. [DOI] [PubMed] [Google Scholar]
  • 81.BHATNAGAR S, NOWAK N, BABICH L, BOK L. Deletion of the 5-HT3 receptor differentially affects behavior of males and females in the Porsolt forced swim and defensive withdrawal tests. Behav. Brain Res. 2004;153(2):527–535. doi: 10.1016/j.bbr.2004.01.018. [DOI] [PubMed] [Google Scholar]
  • 82.BARNES NM, COSTALL B, GE J, KELLY ME, NAYLOR RJ. The interaction of R(+)- and S(−)-zacopride with PCPA to modify rodent aversive behaviour. Eur. J. Pharmacol. 1992;218(1):15–25. doi: 10.1016/0014-2999(92)90142-q. [DOI] [PubMed] [Google Scholar]
  • 83.EISENSAMER B, UHR M, MEYR S, et al. Antidepressants and antipsychotic drugs colocalize with 5-HT3 receptors in raft-like domains. J. Neurosci. 2005;25(44):10198–10206. doi: 10.1523/JNEUROSCI.2460-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.LUMMIS SC, BAKER J. Radioligand binding and photoaffinity labelling studies show a direct interaction of phenothiazines at 5-HT3 receptors. Neuropharmacology. 1997;36(4-5):665–670. doi: 10.1016/s0028-3908(97)00054-3. [DOI] [PubMed] [Google Scholar]
  • 85.RAMMES G, EISENSAMER B, FERRARI U, et al. Antipsychotic drugs antagonize human serotonin Type 3 receptor currents in a noncompetitive manner. Mol. Psychiatry. 2004;9(9):846–858. doi: 10.1038/sj.mp.4001490. [DOI] [PubMed] [Google Scholar]
  • 86.EISENSAMER B, RAMMES G, GIMPL G, et al. Antidepressants are functional antagonists at the serotonin Type 3 (5-HT3) receptor. Mol. Psychiatry. 2003;8(12):994–1007. doi: 10.1038/sj.mp.4001314. [DOI] [PubMed] [Google Scholar]
  • 87.OLIVIER B, VAN WIJNGAARDEN I, SOUDIJN W. 5-HT3 receptor antagonists and anxiety; a preclinical and clinical review. Eur. Neuropsychopharmacol. 2000;10(2):77–95. doi: 10.1016/s0924-977x(99)00065-6. [DOI] [PubMed] [Google Scholar]
  • 88.OLDEN KW. The use of antidepressants in functional gastrointestinal disorders: new uses for old drugs. CNS Spectr. 2005;10(11):891–896. doi: 10.1017/s1092852900019866. [DOI] [PubMed] [Google Scholar]
  • 89.HARRELL AV, ALLAN AM. Improvements in hippocampal-dependent learning and decremental attention in 5-HT3 receptor overexpressing mice. Learn. Mem. 2003;10(5):410–419. doi: 10.1101/lm.56103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.PRESTON GC. 5-HT3 Antagonists and dosorders of cognition. In: Racagni G, Brunello N, Langer SZ, editors. Recent Advances in the treatment of neurodegenerative diorders and cognitive dysfunction. Vol. 7. International Academy of Biomedical Drug Research; Karger, Basel, Switzerland: 1994. pp. 89–93. [Google Scholar]
  • 91.IIDAKA T, OZAKI N, MATSUMOTO A, et al. A variant C178T in the regulatory region of the serotonin receptor gene HTR3A modulates neural activation in the human amygdala. J. Neurosci. 2005;25(27):6460–6466. doi: 10.1523/JNEUROSCI.5261-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.NIESLER B, FLOHR T, NOTHEN MM, et al. Association between the 5′ UTR variant C178T of the serotonin receptor gene HTR3A and bipolar affective disorder. Pharmacogenetics. 2001;11(6):471–475. doi: 10.1097/00008571-200108000-00002. [DOI] [PubMed] [Google Scholar]
  • 93.BUHOT MC, MARTIN S, SEGU L. Role of serotonin in memory impairment. Ann. Med. 2000;32(3):210–221. doi: 10.3109/07853890008998828. [DOI] [PubMed] [Google Scholar]
  • 94.HODGE CW, SAMSON HH, LEWIS RS, ERICKSON HL. Specific decreases in ethanol- but not water-reinforced responding produced by the 5-HT3 antagonist ICS 205-930. Alcohol. 1993;10(3):191–196. doi: 10.1016/0741-8329(93)90034-l. [DOI] [PubMed] [Google Scholar]
  • 95.DAWES MA, JOHNSON BA, MA JZ, et al. Reductions in and relations between ‘craving’ and drinking in a prospective, open-label trial of ondansetron in adolescents with alcohol dependence. Addict. Behav. 2005;30(9):1630–1637. doi: 10.1016/j.addbeh.2005.07.004. [DOI] [PubMed] [Google Scholar]
  • 96.PELTIER R, SCHENK S. GR38032F, a serotonin 5-HT3 antagonist, fails to alter cocaine self-administration in rats. Pharmacol. Biochem. Behav. 1991;39(1):133–136. doi: 10.1016/0091-3057(91)90410-4. [DOI] [PubMed] [Google Scholar]
  • 97.JOHNSON BA, ROACHE JD, AIT-DAOUD N, ZANCA NA, VELAZQUEZ M. Ondansetron reduces the craving of biologically predisposed alcoholics. Psychopharmacology (Berl) 2002;160(4):408–413. doi: 10.1007/s00213-002-1002-9. [DOI] [PubMed] [Google Scholar]
  • 98.POTVIN S, STIP E, ROY JY. Clozapine, quetiapine and olanzapine among addicted schizophrenic patients: towards testable hypotheses. Int. Clin. Psychopharmacol. 2003;18(3):121–132. doi: 10.1097/01.yic.0000063501.97247.38. [DOI] [PubMed] [Google Scholar]
  • 99.HAYRAPETYAN V, JENSCHKE M, DILLON GH, MACHU TK. Co-expression of the 5-HT3B subunit with the 5-HT3A receptor reduces alcohol sensitivity. Brain Res. Mol. Brain Res. 2005;142(2):146–150. doi: 10.1016/j.molbrainres.2005.09.011. [DOI] [PubMed] [Google Scholar]
  • 100.FARIS PL, ECKERT ED, KIM SW, et al. Evidence for a vagal pathophysiology for bulimia nervosa and the accompanying depressive symptoms. J. Affect. Disord. 2006;92(1):79–90. doi: 10.1016/j.jad.2005.12.047. [DOI] [PubMed] [Google Scholar]
  • 101.FRANK GK, KAYE WH, WELTZIN TE, et al. Altered response to meta-chlorophenylpiperazine in anorexia nervosa: support for a persistent alteration of serotonin activity after short-term weight restoration. Int. J. Eat. Disord. 2001;30(1):57–68. doi: 10.1002/eat.1054. [DOI] [PubMed] [Google Scholar]
  • 102.SCHWORER H, RAMADORI G. Treatment of pruritus: a new indication for serotonin Type 3 receptor antagonists. Clin. Investig. 1993;71(8):659–662. doi: 10.1007/BF00184497. [DOI] [PubMed] [Google Scholar]
  • 103.SCHWORER H, HARTMANN H, RAMADORI G. Relief of cholestatic pruritus by a novel class of drugs: 5-hydroxytryptamine Type 3 (5-HT3) receptor antagonists: effectiveness of ondansetron. Pain. 1995;61(1):33–37. doi: 10.1016/0304-3959(94)00145-5. [DOI] [PubMed] [Google Scholar]
  • 104.O'DONOHUE JW, PEREIRA SP, ASHDOWN AC, et al. A controlled trial of ondansetron in the pruritus of cholestasis. Aliment. Pharmacol. Ther. 2005;21(8):1041–1045. doi: 10.1111/j.1365-2036.2005.02430.x. [DOI] [PubMed] [Google Scholar]
  • 105.WEISSHAAR E, DUNKER N, DOMROSE U, NEUMANN KH, GOLLNICK H. Plasma serotonin and histamine levels in hemodialysis-related pruritus are not significantly influenced by 5-HT3 receptor blocker and antihistaminic therapy. Clin. Nephrol. 2003;59(2):124–129. doi: 10.5414/cnp59124. [DOI] [PubMed] [Google Scholar]
  • 106.WEISSHAAR E, DUNKER N, ROHL FW, GOLLNICK H. Antipruritic effects of two different 5-HT receptor antagonists and an antihistamine in haemodialysis patients. Exp. Dermatol. 2004;13(5):298–304. doi: 10.1111/j.0906-6705.2004.00184.x. [DOI] [PubMed] [Google Scholar]
  • 107.WEISSHAAR E, KUCENIC MJ, FLEISCHER AB., Jr Pruritus: a review. Acta Derm. Venereol. Suppl. (Stockh) 2003;213:5–32. [PubMed] [Google Scholar]
  • 108.KHAN NA, HICHAMI A. Ionotrophic 5-hydroxytryptamine Type 3 receptor activates the protein kinase C-dependent phospholipase D pathway in human T-cells. Biochem. J. 1999;344(Pt. 1):199–204. [PMC free article] [PubMed] [Google Scholar]
  • 109.PICHE T, VANBIERVLIET G, CHERIKH F, et al. Effect of ondansetron, a 5-HT3 receptor antagonist, on fatigue in chronic hepatitis C: a randomised, double blind, placebo controlled study. Gut. 2005;54(8):1169–1173. doi: 10.1136/gut.2004.055251. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.THE GK, PRINS J, BLEIJENBERG G, VAN DER MEER JW. The effect of granisetron, a 5-HT3 receptor antagonist, in the treatment of chronic fatigue syndrome patients-a pilot study. Neth J. Med. 2003;61(9):285–289. [PubMed] [Google Scholar]
  • 111.STRATZ T, FARBER L, VARGA B, et al. Fibromyalgia treatment with intravenous tropisetron administration. Drugs Exp. Clin. Res. 2001;27(3):113–118. [PubMed] [Google Scholar]
  • 112.SPATH M, STRATZ T, NEECK G, et al. Efficacy and tolerability of intravenous tropisetron in the treatment of fibromyalgia. Scand. J. Rheumatol. 2004;33(4):267–270. doi: 10.1080/03009740410005818. [DOI] [PubMed] [Google Scholar]
  • 113.MULLER W, STRATZ T. Local treatment of tendinopathies and myofascial pain syndromes with the 5-HT3 receptor antagonist tropisetron. Scand. J. Rheumatol. 2004;33(Suppl. 119):44–48. [PubMed] [Google Scholar]
  • 114.STRATZ T, MULLER W. Treatment of chronic low back pain with tropisetron. Scand. J. Rheumatol. 2004;33(Suppl. 119):76–78. doi: 10.1080/03009740410007113. [DOI] [PubMed] [Google Scholar]
  • 115.SAMBORSKI W, STRATZ T, MACKIEWICZ S, MULLER W. Intra-articular treatment of arthritides and activated osteoarthritis with the 5-HT3 receptor antagonist tropisetron. A double-blind study compared with methylprednisolone. Scand. J. Rheumatol. 2004;33(Suppl. 119):51–54. [PubMed] [Google Scholar]
  • 116.STRATZ T, MULLER W. Local treatment of rheumatic diseases with the 5-HT3 receptor antagonist tropisetron. Schmerz. 2003;17(3):200–203. doi: 10.1007/s00482-003-0219-9. [DOI] [PubMed] [Google Scholar]
  • 117.RIERING K, REWERTS C, ZIEGLGANSBERGER W. Analgesic effects of 5-HT3 receptor antagonists. Scand. J. Rheumatol. 2004;33(Suppl. 119):19–23. [PubMed] [Google Scholar]
  • 118.GIORDANO J, SCHULTEA T. Serotonin 5-HT3 Receptor mediation of pain and anti-nociception: implications for clinical therapeutics. Pain Physician. 2004;7(1):141–147. [PubMed] [Google Scholar]
  • 119.DEVLIN J, WAGSTAFF K, ARTHUR V, EMERY P. Granisetron (Kytril) suppresses methotrexate-induced nausea and vomiting among patients with inflammatory arthritis and is superior to prochlorperazine (Stemetil) Rheumatology (Oxford) 1999;38(3):280–282. doi: 10.1093/rheumatology/38.3.280. [DOI] [PubMed] [Google Scholar]
  • 120.SANDRINI M, VITALE G, PINI LA. Central antinociceptive activity of acetylsalicylic acid is modulated by brain serotonin receptor subtypes. Pharmacology. 2002;65(4):193–197. doi: 10.1159/000064343. [DOI] [PubMed] [Google Scholar]
  • 121.SANDRINI M, PINI LA, VITALE G. Differential involvement of central 5-HT1B and 5-HT3 receptor subtypes in the antinociceptive effect of paracetamol. Inflamm. Res. 2003;52(8):347–352. doi: 10.1007/s00011-003-1185-5. [DOI] [PubMed] [Google Scholar]
  • 122.PICKERING G, LORIOT MA, LIBERT F, et al. Analgesic effect of acetaminophen in humans: first evidence of a central serotonergic mechanism. Clin. Pharmacol. Ther. 2006;79(4):371–378. doi: 10.1016/j.clpt.2005.12.307. [DOI] [PubMed] [Google Scholar]
  • 123.BALLESTERO JA, PLAZAS PV, KRACUN S, et al. Effects of quinine, quinidine and chloroquine on α9α10 nicotinic cholinergic receptors. Mol. Pharmacol. 2005;68:822–829. doi: 10.1124/mol.105.014431. [DOI] [PubMed] [Google Scholar]
  • 124.SIEB JP, MILONE M, ENGEL AG. Effects of the quinoline derivatives quinine, quinidine, and chloroquine on neuromuscular transmission. Brain Res. 1996;712(2):179–189. doi: 10.1016/0006-8993(95)01349-0. [DOI] [PubMed] [Google Scholar]
  • 125.BROUGHMAN JR, BRANDT RM, HASTINGS C, et al. Channel-forming peptide modulates transepithelial electrical conductance and solute permeability. Am. J. Physiol. Cell Physiol. 2004;286(6):C1312–C1323. doi: 10.1152/ajpcell.00426.2002. [DOI] [PubMed] [Google Scholar]
  • 126.BROUGHMAN JR, SHANK LP, PRAKASH O, et al. Structural implications of placing cationic residues at either the NH2- or COOH-terminus in a pore-forming synthetic peptide. J. Membr. Biol. 2002;190(2):93–103. doi: 10.1007/s00232-002-1027-3. [DOI] [PubMed] [Google Scholar]
  • 127.LUMMIS SC, KILPATRICK GJ, MARTIN IL. Characterization of 5-HT3 receptors in intact N1E-115 neuroblastoma cells. Eur. J. Pharmacol. 1990;189(2-3):223–227. doi: 10.1016/0922-4106(90)90026-t. [DOI] [PubMed] [Google Scholar]
  • 128.DOWNIE DL, HOPE AG, LAMBERT JJ, et al. Pharmacological characterization of the apparent splice variants of the murine 5-HT3 R-A subunit expressed in Xenopus laevis oocytes. Neuropharmacology. 1994;33(3-4):473–482. doi: 10.1016/0028-3908(94)90078-7. [DOI] [PubMed] [Google Scholar]
  • 129.LUMMIS SC, SEPULVEDA MI, KILPATRICK GJ, BAKER J. Characterization of [3H]meta-chlorophenylbiguanide binding to 5-HT3 receptors in N1E-115 neuroblastoma cells. Eur. J. Pharmacol. 1993;243(1):7–11. doi: 10.1016/0014-2999(93)90160-j. [DOI] [PubMed] [Google Scholar]
  • 130.HOPE AG, PETERS JA, BROWN AM, LAMBERT JJ, BLACKBURN TP. Characterization of a human 5-hydroxytryptamine3 receptor type A (h5- HT3R-AS) subunit stably expressed in HEK 293 cells. Br. J. Pharmacol. 1996;118(5):1237–1245. doi: 10.1111/j.1476-5381.1996.tb15529.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.STEWARD LJ, GE J, BENTLEY KR, et al. Evidence that the atypical 5-HT3 receptor ligand, [3H]-BRL46470, labels additional 5-HT3 binding sites compared to [3H]-granisetron. Br. J. Pharmacol. 1995;116(2):1781–1788. doi: 10.1111/j.1476-5381.1995.tb16663.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.PETERS JA, MALONE HM, LAMBERT JJ. An electrophysiological investigation of the properties of 5-HT3 receptors of rabbit nodose ganglion neurones in culture. Br. J. Pharmacol. 1993;110(2):665–676. doi: 10.1111/j.1476-5381.1993.tb13863.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.WONG EH, CLARK R, LEUNG E, et al. The interaction of RS 25259-197, a potent and selective antagonist, with 5-HT3 receptors, in vitro. Br. J. Pharmacol. 1995;114(4):851–859. doi: 10.1111/j.1476-5381.1995.tb13282.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.BOEIJINGA PH, GALVAN M, BARON BM, et al. Characterization of the novel 5-HT3 antagonists MDL 73147EF (dolasetron mesilate) and MDL 74156 in NG108-15 neuroblastoma × glioma cells. Eur. J. Pharmacol. 1992;219(1):9–13. doi: 10.1016/0014-2999(92)90573-m. [DOI] [PubMed] [Google Scholar]
  • 135.KATAYAMA K, ASANO K, HAGA K, FUKUDA T. High affinity binding of azasetron hydrochloride to 5-hydroxytryptamine3 receptors in the small intestine of rats. Jpn. J. Pharmacol. 1997;73(4):357–360. doi: 10.1254/jjp.73.357. [DOI] [PubMed] [Google Scholar]
  • 136.CLAYTON NM, SARGENT R, BUTLER A, et al. The pharmacological properties of the novel selective 5-HT3 receptor antagonist, alosetron, and its effects on normal and perturbed small intestinal transit in the fasted rat. Neurogastroenterol. Motil. 1999;11(3):207–217. doi: 10.1046/j.1365-2982.1999.00148.x. [DOI] [PubMed] [Google Scholar]
  • 137.AKUZAWA S, ITO H, YAMAGUCHI T. Comparative study of [3H]ramosetron and [3H]granisetron binding in the cloned human 5-hydroxytryptamine3 receptors. Jpn. J. Pharmacol. 1998;78(3):381–384. doi: 10.1254/jjp.78.381. [DOI] [PubMed] [Google Scholar]
  • 138.CAPPELLI A, GALLELLI A, MANINI M, et al. Further studies on the interaction of the 5-hydroxytryptamine3 (5-HT3) receptor with arylpiperazine ligands. Development of a new 5-HT3 receptor ligand showing potent acetylcholinesterase inhibitory properties. J. Med. Chem. 2005;48(10):3564–3575. doi: 10.1021/jm0493461. [DOI] [PubMed] [Google Scholar]
  • 139.HAGIHARA K, HAYAKAWA T, ARAI T, et al. Antagonistic activities of N-3389, a newly synthesized diazabicyclo derivative, at 5-HT3 and 5-HT4 receptors. Eur. J. Pharmacol. 1994;271(1):159–166. doi: 10.1016/0014-2999(94)90276-3. [DOI] [PubMed] [Google Scholar]
  • 140.VENKATARAMAN P, JOSHI P, VENKATACHALAN SP, et al. Functional group interactions of a 5-HT3R antagonist. BMC Biochem. 2002;3:16. doi: 10.1186/1471-2091-3-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.VAN WIJNGAARDEN I, HAMMINGA D, VAN HES R, et al. Development of high-affinity 5-HT3 receptor antagonists. Structure–affinity relationships of novel 1,7-annelated indole derivatives. J. Med. Chem. 1993;36(23):3693–3699. doi: 10.1021/jm00075a026. [DOI] [PubMed] [Google Scholar]

RESOURCES