Open access peer-reviewed chapter

Role of Calcium in Vomiting

Written By

Weixia Zhong and Nissar A. Darmani

Submitted: 21 April 2018 Reviewed: 07 May 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.78370

From the Edited Volume

Calcium and Signal Transduction

Edited by John N. Buchholz and Erik J. Behringer

Chapter metrics overview

1,515 Chapter Downloads

View Full Metrics

Abstract

Cisplatin-like chemotherapeutics cause vomiting via calcium (Ca2+)-dependent release of multiple neurotransmitters/mediators (dopamine, serotonin, substance P, prostaglandins and leukotrienes) from the gastrointestinal enterochromaffin cells and/or the brainstem. Intracellular Ca2+ signaling is triggered by activation of diverse emetic receptors (including neurokininergic NK1, serotonergic 5-HT3, dopaminergic D2, cholinergic M1, or histaminergic H1), whose stimulation in vomit-competent species evokes emesis. Other emetogens such as cisplatin, rotavirus NSP4 protein, and bacterial toxins can also induce intracellular Ca2+ elevation. Our findings demonstrate that application of the L-type Ca2+ channel (LTCC) agonist FPL 64176 and the intracellular Ca2+ mobilizing agent thapsigargin (a sarco/endoplasmic reticulum Ca2+-ATPase inhibitor) cause vomiting in the least shrew. On the other hand, blockade of LTCCs by corresponding antagonists (nifedipine or amlodipine) not only provide broad-spectrum antiemetic efficacy against diverse agents that specifically activate emetogenic receptors such as 5-HT3, NK1, D2, and M1 receptors, but can also potentiate the antiemetic efficacy of palonosetron against the nonspecific emetogen, cisplatin. In this review, we will provide an overview of Ca2+ involvement in the emetic process; discuss the relationship between Ca2+ signaling and the prevailing therapeutics in control of vomiting; highlight the current evidence for Ca2+-signaling blockers/inhibitors in suppressing emetic behavior and also draw attention to the clinical benefits of Ca2+-signaling blockers/inhibitors for the treatment of nausea and vomiting.

Keywords

  • cisplatin
  • vomiting
  • antiemesis
  • Ca2+
  • L-type Ca2+ channel

1. Introduction

Acute (≤24 h) and delayed (>24 h) phases of chemotherapy-induced nausea and vomiting cause distressing side-effects which affect the well-being and quality of life of cancer patients receiving chemotherapy, especially cisplatin [1]. Major neurotransmitter mechanisms underlying chemotherapy-induced nausea and vomiting have been subject of considerable research over the past 45 years. As presented in brief in Figure 1, cancer chemotherapeutics such as cisplatin evoke vomiting via local release of a variety of emetic neurotransmitters/mediators (including dopamine, serotonin (5-HT), substance P, prostaglandins and leukotrienes) both from the enterochromaffin cells of the gastrointestinal tract and the brainstem emetic loci in the dorsal vagal complex containing the nucleus tractus solitarius, the dorsal motor nucleus of the vagus and the area postrema [2, 3, 4]. The area postrema and the nucleus tractus solitarius contain large numbers of fenestrated capillaries which lack blood-brain barrier and permit neurons in both areas access to blood-borne circulating factors including emetogens [5]. The chemoreceptor trigger zone, in the area postrema has high concentrations of emetic receptors for serotonin (5-HT3), dopamine (D2/3), neurokinin (NK1), and opioids (μ), among others [2]. Direct stimulation of these receptors in the chemoreceptor trigger zone by emetogens is one important mechanism by which vomiting can occur [6]. The nucleus tractus solitarius receives emesis-related information from the area postrema as well as the gastrointestinal tract conveyed by vagal afferents. The dorsal motor nucleus of the vagus receives axonal projections from nucleus tractus solitarius [7] and sends emetic signals via motor efferent pathways to the gastrointestinal tract and modulates vomiting behaviors [2, 5, 8, 9] (Figure 1). In addition, chemotherapeutic drugs may evoke release of emetic neurotransmitters/mediators from the gastrointestinal tract into the blood to be directly delivered to the area postrema via a blood-borne pathway which then triggers vomiting [2, 10], and/or the released emetic neurotransmitters/mediators stimulate their corresponding receptors present on vagal afferents in the gastrointestinal tract which indirectly activate brainstem emetic loci primarily in the nucleus tractus solitaries to trigger vomiting [6].

Figure 1.

Brief illustration of the mechanisms underlying vomiting induced by chemotherapeutic agent cisplatin. Mechanisms underlying cisplatin-induced vomiting can be simplified as: (1) cisplatin can increase cytoplasmic Ca2+ level to evoke Ca2+-dependent release of emetic neurotransmitters/mediators at the brainstem emetic loci, the dorsal vagal complex, and subsequently activates diverse receptors and their corresponding signaling pathways. These emetic signals are output to the gastrointestinal tract via efferents to trigger vomiting [2, 4, 5, 6, 7, 8, 9]; (2) cisplatin-induced peripheral release of neurotransmitters/mediators from the gastrointestinal tract into the blood can directly stimulate the dorsal vagal complex, activate receptors signaling pathways and trigger vomiting [2, 10]; and (3) the peripherally-released emetic neurotransmitters/mediators stimulate their corresponding receptors present on vagal afferents in the gastrointestinal tract which indirectly activate brainstem emetic nuclei and trigger vomiting [6].

Ca2+ is not only one of the most universal and versatile signaling molecules, it is also an extremely important factor in both the physiology and pathology of living organisms. At rest, diverse cells have strict and well-regulated mechanisms to maintain low nM cytosolic Ca2+ levels [11]. Cytoplasmic Ca2+ concentration is a dominant factor in determining the amount of transmitter released from nerve terminals [12]. Thus, Ca2+ mobilization can be an important aspect of vomit induction since it is involved in both triggering the quantity of neurotransmitter released coupled with receptor activation, as well as post-receptor excitation-transcription coupling mechanisms [13]. Studies using Ca2+ imaging performed in vitro in the brainstem slice preparation suggest that emetic agents evoke direct excitatory effects on cytosolic Ca2+ signals in vagal afferent terminals in the nucleus tractus solitarius which potentiate local neurotransmitter release [5, 14, 15]. Therefore, chemotherapeutics including cisplatin seem to activate emetic circuits through a number of neurotransmitters released in a Ca2+-dependent manner in specific vomit-associated neuroanatomical structures. In both the periphery and the brainstem, emetic neurotransmitters/mediators—such as acetylcholine, dopamine, 5-HT, substance P, prostaglandins, leukotrienes, and/or histamine—may act independently or in combination to evoke vomiting after cisplatin administration [16] (Figure 1). In this review, we focus on the current evidence supporting the significance of Ca2+ signaling in emesis generation and its relationship to antiemetic efficacy, as well as the corresponding development of potential novel antiemetic medications, as shown in brief in Figure 2.

Figure 2.

Overview of evidence for suppression of Ca2+ signaling involved in anti-vomiting actions of antiemetic agents. (1) Netupitant and palonosetron are highly selective respective antagonists of NK1Rs and 5-HT3Rs are approved to treat the acute- and delayed- phases of chemotherapy-induced nausea and vomiting (CINV) in cancer patients [79, 80, 81, 82]. Our studies [83, 84, 85, 86] indicate that suppression of Ca2+ signaling is involved in antiemetic efficacy of both palonosetron and netupitant. (2) Cannabinoids such as delta-9-tetrahydrocannabinol exert their antiemetic efficacy via direct activation of CB1 receptors (CB1R) [92, 94, 98, 99, 100]. The ability of CB1R agonists to suppress both extracellular Ca2+ influx [111, 112, 113, 114, 115] and intracellular Ca2+ release from the sarco/endoplasmic reticulum stores [15, 117], result in inhibition of Ca2+-dependent neurotransmitter release [108] and is probably the fundamental mechanisms underlying the antiemetic efficacy of CB1R cannabinoid agonists against CINV [95, 96, 97]. (3) Glucocorticoids such as dexamethasone reduce both acute and delayed CINV [6]. Glucocorticoids’ ability to decrease the abnormal elevation of cytosolic Ca2+ concentration [122], and subsequently control Ca2+-dependent neurotransmitter release [6, 121, 126] and inflammatory responses [6]. Increased release of endocannabinoids and subsequent CB1R activation may also be involved in antiemetic actions of glucocorticoids [123, 124, 125]. (4) The L-type Ca2+ channel (LTCC) antagonist flunarizine can reduce cyclic vomiting in patients [151, 152]. Gabapentin binds to the alpha-2/delta auxiliary subunits of LTCCs, and exerts inhibitory actions on trafficking and activation kinetics of LTCCs [153]. Gabapentin can be used as an anti-nausea and antiemetic agent in postoperative nausea and vomiting [154, 155] and in CINV [156, 157]. (5) LTCC antagonists (nifedipine and amlodipine) are broad-spectrum antiemetics when delivered systemically against diverse specific and nonselective emetogens. (6) Suppression of intracellular Ca2+ release from the sarco/endoplasmic reticulum through the inositol trisphosphate receptors (IP3Rs) and ryanodine receptors (RyRs) may be additional targets for the prevention of nausea and vomiting, since functional and physical linkages between Ca2+ channels on cell membrane and IP3Rs/RyRs play a role in Ca2+ signaling [160, 161, 162, 163, 164, 165, 166]. In the least shrew emesis model, the RyRs antagonist dantrolene can potentiate the antiemetic efficacy of amlodipine against 5-HT3R agonist 2-Methyl-5-HT-induced vomiting [25]; and dantrolene together with the IP3R antagonist 2-APB can potentiate the antiemetic efficacy of nifedipine against thapsigargin-induced vomiting [70].

Advertisement

2. Emerging roles of Ca2+ in emesis

2.1. Emetic receptor stimulation increases intracellular Ca2+ concentration

Excitatory receptor activation by corresponding agonists can increase cytosolic Ca2+ levels via both mobilization of intracellular Ca2+ stores (e.g., endoplasmic reticulum = ER) and influx from extracellular fluid [17]. The evoked cytoplasmic Ca2+ increase may result from direct activation of ion channels, or indirectly via signal transduction pathways following G protein-coupled receptor activation. The neurokinin NK1 receptor (NK1R) is a member of the tachykinin family of G-protein-coupled receptors. NK1R stimulation by substance P or corresponding selective agonists such as GR73632, can increase cytosolic Ca2+ concentration. In fact GR73632-induced activation of NK1Rs can evoke intracellular Ca2+ release from the sarco/endoplasmic reticulum stores via Gα/q-mediated phospholipase C pathway, which subsequently evokes extracellular Ca2+ influx through L-type Ca2+ channels (LTCCs) [17, 18, 19]. The serotonergic 5-HT3 receptor (5-HT3R) is a Ca2+-permeable ligand-gated ion channel [20]. Cell lines studies have demonstrated that activation of 5-HT3Rs by 5-HT or its analogs can evoke extracellular Ca2+ influx into cells in a manner sensitive to both 5-HT3R antagonists (tropisetron, MDL7222, metoclopramide) and LTCC blockers (verapamil, nimodipine, nitrendipine) [20, 21, 22, 23, 24]. These studies suggest that both L-type- and 5-HT3-receptor Ca2+-permeable ion channels are involved in extracellular Ca2+ influx evoked by 5-HT3R agonists. Moreover, 5-HT3R activation indirectly causes release of Ca2+ from ryanodine-sensitive intracellular Ca2+ stores subsequent to the evoked extracellular Ca2+ influx which greatly amplifies the cytoplasmic concentration of Ca2+ [23]. In fact, our findings from behavioral studies in the least shrew emesis model [25] further support the notion of Ca2+-induced Ca2+ release following 5-HT3R stimulation, which will be discussed in more detail in Section 3.4. Other emetogens such as agonists of dopamine D2- [26, 27], cholinergic M1- [28, 29], histaminergic H1- [30, 31], and opiate μ- [32, 33] receptors, as well as cisplatin [34], prostaglandins [35, 36], rotavirus NSP4 protein [37, 38] and bacterial toxins [39, 40] also possess the potential to mobilize Ca2+ which involve extracellular Ca2+ influx and/or Ca2+ release from intracellular Ca2+ pools. Much of the discussed evidence has been acquired from isolated cells.

The least shrew is an emesis-competent mammal whose reactions to common emetogens are well-defined and correlate closely with human responses [2]. 2-Methyl-5-HT is a well-known selective emetic agonist targeting the emesis-prone 5-HT3Rs [4]. This vomit-competent species is an excellent animal model for studying the emetic activity of diverse agents [2]. In fact least shrews exhibit dose-dependent full emetic responses to intraperitoneal administration of both the peripherally-acting 5-HT, as well as to its central nervous system-penetrating analog, 2-Methyl-5-HT [4, 41, 42]. In our studies, incubation of least shrew brainstem slices containing the dorsal vagal complex emetic loci with 2-Methyl-5-HT, results in a rapid increase in intracellular Ca2+ concentration as reflected by an increase in fluo-4 AM fluorescence intensity in a palonosetron (a 5-HT3R antagonist)/nifedipine-sensitive manner [22, 25].

2.2. Emetic potential of Ca2+ channel activators: behavioral and immunohistochemical evidence

A variety of Ca2+-permeable ion-channels mediating extracellular Ca2+ influx are present in the plasma membrane. Among them are voltage-gated LTCCs, which can be activated by membrane depolarization, and serve as the principal route of Ca2+ entry in electrically excitable cells such as neurons and muscle [43, 44]. Recently we have acquired direct evidence for the proposal that Ca2+ mobilization is an important facet in the mediation of emesis. In fact we have identified the novel emetogen FPL64176 (Figure 2), a selective agonist of LTCCs, which causes vomiting in the least shrew in a dose-dependent manner [45, 46]. All tested shrews vomited at a 10 mg/kg dose of FPL64176 administered intraperitoneally (i.p.). LTCCs have been shown to be present in enterochromaffin cells of guinea pig and human small intestinal crypts [47]. Furthermore, in these cells FPL64176 not only can enhance cytosolic Ca2+ concentration, but also increases 5-HT release from enterochromaffin cells [47]. The latter findings may have underpinnings for the mechanisms underlying FPL64176-evoked vomiting observed in least shrew model of emesis. FPL64176 (10 mg/kg., i.p.) can cause Ca2+-dependent 5-HT release from shrew intestinal enterochromaffin cells which in turn could increase vagal afferent activity via stimulation of 5-HT3 receptors, thereby indirectly triggering emetic signals in the brainstem [2, 48].

Our most recent work has focused on the Ca2+-mobilizing agent thapsigargin (Figure 3), a specific and potent inhibitor of the sarco/endoplasmic reticulum Ca2+-ATPase (SERCA) pump which transports the free cytosolic Ca2+ into the lumen of the sarco/endoplasmic reticulum to counter-balance the cytosolic intracellular Ca2+ release from the sarco/endoplasmic reticulum into the cytoplasm via the inositol trisphosphate receptors (IP3Rs) and ryanodine receptors (RyRs) [49, 50, 51]. Thapsigargin also causes intracellular release of stored Ca2+ from the sarco/endoplasmic reticulum into the cytosol which subsequently evokes extracellular Ca2+ influx predominantly through store-operated Ca2+ entry (SOCE) in non-excitable cells [52, 53, 54]. In total, these events lead to a significant rise in the free concentration of cytosolic Ca2+ [55, 56, 57]. In addition, a partial involvement of LTCCs in thapsigargin-evoked contraction has also been demonstrated in rat stomach smooth muscle cells [58], rat gastric smooth muscle [59], and cat gastric smooth muscle [60]. On the other hand, the potential of thapsigargin as a Ca2+-modulating cancer chemotherapeutic agent has been evaluated in both cells and animal models [61]. Thapsigargin-evoked increases in cytosolic Ca2+ concentration can lead to cell apoptosis, which can result in eradication of cancer cells of the breast [62], prostate [63], colon [64] and kidneys [65]. Clinically, a prodrug form of thapsigargin, mipsagargin, is currently under clinical trial as a targeted cancer chemotherapeutic agent with selective toxicity against cancer cells in tumor sites with minimal side-effects to the host [66, 67, 68, 69]. In our studies, intraperitoneal administration of thapsigargin (0.1–10 mg/kg) caused vomiting in least shrews in a dose-dependent, but bell-shaped manner, with maximal efficacy at 0.5 mg/kg. An important consideration for the emetic potential of thapsigargin is that it augments the cytosolic levels of free Ca2+ in emetic loci as a result of SERCA inhibition as indicated in our latest discussed finding [70], which is the first study to reflect emesis as a major side-effect of thapsigargin when delivered systemically.

Figure 3.

A schematic representation of extracellular Ca2+ influx and intracellular Ca2+ release contributing to thapsigargin-elicited Ca2+ mobilization. Intracellular Ca2+ release from the sarco/endoplasmic reticulum (SER) Ca2+ stores through the inositol triphosphate receptors (IP3Rs) and ryanodine receptors (RyRs) is counter-balanced by continuous Ca2+ uptake from the cytoplasm into SER stores by the SER Ca2+-ATPase pump (SERCA). Thapsigargin is a specific inhibitor of SERCA and thus enhances cytosolic levels of Ca2+, a process involving SER Ca2+ release via IP3Rs and RyRs as well as extracellular Ca2+ entry through Ca2+ channels located in the plasma membrane including store-operated Ca2+ channels (SOCE) and L-type Ca2+ channels (LTCCs) [49, 50, 51, 52, 53, 54, 55, 56, 57, 58, 59, 60].

c-Fos induction has been used to evaluate differential neuronal activation [71]. Our lab has applied immunostaining and detected c-Fos induction in the brainstem emetic nuclei to demonstrate central responsiveness to peripheral administration of a variety of emetogens [4, 70, 72, 73]. The participation of the central emetic neurons in FPL64176-induced vomiting is further indicated by evoked c-Fos expression in brainstem emetic nuclei, the nucleus tractus solitarius and the dorsal motor nucleus of the vagus (unpublished data). Thus, the blood-brain barrier permeable agent FPL64176 [74, 75, 76] could excite emetic neurons directly in the nucleus tractus solitarius and the dorsal motor nucleus of the vagus. Thapsigargin (0.5 mg/kg) also causes increases in c-Fos immunoreactivity in the brainstem emetic nuclei including the area postrema, the nucleus tractus solitarius and the dorsal motor nucleus of the vagus [70].

Advertisement

3. Ca2+ intervention mechanisms relevant to antiemetic approaches

3.1. Receptor antagonist antiemetic regimens such as netupitant/palonosetron (NEPA)

The ultimate aim of prophylactic management of chemotherapy-induced nausea and vomiting is to abolish both the acute- and delayed phases of vomiting which will help to improve the well-being and quality of life of cancer patients receiving chemotherapy. Cisplatin-like chemotherapeutics cause release of multiple emetogenic neurotransmitters in both the central nervous system and the gastrointestinal tract and no available single antiemetic administered alone can provide complete efficacy. Significant initial work had suggested that while activation of 5-HT3Rs by serotonin in the gastrointestinal tract is involved in the mediation of acute phase of chemotherapy-induced nausea and vomiting, the delayed phase is due to stimulation of NK1Rs subsequent to release of substance P in the brainstem [77, 78]. However, our more recent findings suggest that 5-HT and substance P are concomitantly involved in both emetic phases in the gastrointestinal tract as well as in the brainstem [2, 16]. While netupitant is a highly selective and a longer-acting second generation NK1R antagonist, palonosetron is considered as a second generation 5-HT3R antagonist with a unique antiemetic profile in both humans [79, 80] and the least shrew model of emesis [45]. A successful regimen of an oral fixed combined dose of netupitant/palonosetron (NEPA) (Figure 2) has been formulated with over 85% clinical efficacy, good tolerability, and high central nervous system penetrance for the prophylactic treatment of acute and delayed chemotherapy-induced nausea and vomiting in cancer patients receiving chemotherapy [9, 81, 82].

Recent evidence accumulated from HEK293 cells stably transfected with 5-HT3Rs suggest that suppression of Ca2+ signaling is involved in antiemetic efficacy of both palonosetron and netupitant. Indeed, Rojas et al. [83, 84] have shown that palonosetron causes a persistent inhibition of 5-HT3R function as reflected by a near complete suppression of 5-HT-evoked extracellular Ca2+ influx. They have further demonstrated that palonosetron can prevent enhancement of substance P-induced intracellular Ca2+ release in response to serotonin in NG108–15 cells expressing both 5-HT3Rs and NK1Rs [85]. Our Ca2+ monitoring studies performed on acutely-prepared least shrew brainstem slices also demonstrate that palonosetron can abolish enhancement of intracellular Ca2+ levels in brainstem slices evoked by the selective 5-HT3R agonist 2-Methyl-5-HT [25]. The latter finding provides more relevant ex-vivo evidence for the Ca2+-modulating antiemetic effect of palonosetron in a vomit-competent species. The role of netupitant in suppression of substance P-evoked enhancement of intracellular Ca2+ levels has also been demonstrated via Ca2+ mobilization assays in vitro in CHO cells expressing the human NK1Rs. Moreover, pronetupitant, an intravenous alternative to the oral netupitant, appears to be more potent than netupitant in both in vitro Ca2+ measurement studies and in vivo animal behavioral evaluations of substance P in rats [86]. In addition, another clinically approved NK1R antagonist antiemetic rolapitant, has been shown to suppress the ability of the selective NK1R agonist GR73632 to evoke intracellular Ca2+ release [9, 87, 88, 89]. The discussed findings clearly suggest that Ca2+ is a major player in the initiation of vomiting evoked by diverse emetogens.

3.2. Cannabinoid CB1 receptor agonists

Before the introduction of first generation 5-HT3R antagonists, several phyto- and synthetic cannabinoids including dronabinol (delta-9-tetrahydrocannabinol, Δ9-THC (Figure 2)), levonantradol and nabilone, were evaluated in cancer patients for suppression of chemotherapy-induced nausea and vomiting that were not effectively controlled by other available antiemetics [2, 90]. Cannabinoids are increasingly being tested as antiemetics against cisplatin-induced emesis in animal experiments using house musk shrews [91], ferrets [92], or least shrews [73, 93]; nausea-related behavior in rats [91]; radiation-induced emesis in the least shrew [94]; as well as both phases of chemotherapy-induced nausea and vomiting in the clinic [95, 96, 97]. Cannabinoid agonists exert their antiemetic efficacy via direct activation of CB1 receptors (CB1R) since their antiemetic effects were reversed by CB1R antagonists [92, 94, 98, 99, 100]. Significant evidence for a role for CB2Rs in emesis is currently lacking [101]. The presence of CB1Rs in the brainstem nuclei involved in emesis has been confirmed, with a high density of CB1R immunoreactivity in the dorsal motor nucleus of the vagus and the medial subnucleus of the nucleus tractus solitarius, a moderate density in the commissural subnucleus of the nucleus tractus solitarius, and lower densities in the area postrema and dorsal subnucleus of the nucleus tractus solitarius [73, 92]. CB1R distribution has been also observed in the myenteric plexus of the stomach and duodenum [92]. Furthermore, CB1Rs have been localized in the myenteric plexus of the rat and guinea pig intestine in nearly all cholinergic neuron terminals [102, 103]. These as well as behavioral evidence [42] suggest that the antiemetic action of cannabinoids involve both the central dorsal vagal complex and intestinal emetic loci. In addition, primary cultures of guinea-pig myenteric neurons express CB1Rs and exogenously added cannabinoids suppress their neuronal activity, synaptic transmission and mitochondrial transport along axons [104]. Moreover, the CB1/2R agonist WIN55212-2 can suppress intestinal activity since it can attenuate the electrically-evoked contractions of the myenteric plexus-longitudinal muscle preparation of the guinea-pig small intestine in a Ca2+-dependent and CB1R-specific manner [105]. Thus, CB1R agonists in the in vivo setting can also suppress the gastrointestinal tract motility [104]. Using whole-cell patch-clamp recordings in brainstem slices, Derbenev et al. [106, 107] have shown that activation of presynaptic CB1Rs in the dorsal vagal complex inhibits synaptic transmission to the dorsal motor nucleus of the vagus neurons, which may explain suppression of visceral motor responses caused by cannabinoids.

Furthermore, in the central nervous system CB1R stimulation can result in inhibition of Ca2+-dependent neurotransmitter release from presynaptic nerve terminals which consequently leads to inhibition of neurotransmission [108]. In chemotherapy-induced nausea and vomiting, the CB1R-mediated antiemetic action of cannabinoids appears to be directly related to presynaptic inhibition of release of emetic neurotransmitters from nerve terminals. Figure 4 may help to explain the antiemetic action of cannabinoid CB1R agonists from the Ca2+ perspective. Indeed, the adenylyl cyclase/cyclic AMP (cAMP)/protein kinase A (PKA) signal transduction system is a well-established emetic signaling pathway [109]. PKA activation is known to phosphorylate both Ca2+ ion channels on plasma membrane and intracellular endoplasmic IP3Rs, which respectively increase extracellular Ca2+ influx and internal Ca2+ release from the sarco/endoplasmic reticulum stores [110]. CB1Rs are known to be Gi/o-protein coupled receptors which mediates inhibition of adenylate cyclase. This inhibition has been proposed to be the fundamental reason for CB1R agonists attenuating Ca2+-dependent emetic neurotransmitter release which would ultimately reduce postsynaptic neuronal activation in both dorsal vagal complex and gastrointestinal tract [93, 103]. Moreover, dose-dependent inhibitory action of cannabinoid CB1R agonists on extracellular Ca2+ influx via a number of voltage-gated Ca2+ channels residing in the cell membrane including N-type, P/Q type and L-type have been demonstrated in multiple experimental systems [111, 112, 113, 114, 115]. Additionally, cannabinoid CB1R agonists also block 5-HT3Rs in a non-competitive manner and thus prevent extracellular Ca2+ influx [115, 116].

Figure 4.

A schematic explanation of the antiemetic action of cannabinoid CB1R agonists from the perspective of Ca2+ signaling. Activation of CB1R initiates a Gi/o mechanism leading to the downregulation of extracellular Ca2+ influx through voltage-gated Ca2+ channels (VGCCs) as well as endoplasmic reticulum (ER) Ca2+ release via ryanodine receptors (RyRs) which has the potential to be activated by extracellular Ca2+ entry through VGCCs. The reduction in cytosolic Ca2+ attenuates Ca2+-dependent emetic neurotransmitter release, which further results in a reduction in postsynaptic neuronal activation, and ultimately suppression of the vomiting behavior [93, 103, 117].

Furthermore, CB1R agonists appear to inhibit the intracellular Ca2+ release channels located on the sarco/endoplasmic reticulum membrane, RyRs. Ca2+-induced Ca2+ release is a well-established feature of Ca2+ signal amplification. During neuronal activation, Ca2+-induced Ca2+ release Ca2+ signaling involves increased concentration of cytoplasmic Ca2+ via extracellular Ca2+ influx through voltage-gated Ca2+ channels (e.g., LTCCs) present on the cell membrane, which then causes release of stored intracellular Ca2+ from the sarco/endoplasmic reticulum into the cytosol through RyRs [117]. In fact RyRs have a wide distribution in the central nervous system including the brainstem [118]. RyRs not only can regulate Ca2+ homeostasis, but also other critical brain functions including neurotransmitter release [117]. Increased serum levels of the pro-inflammatory cytokine, tumor necrosis factor alpha (TNF-α), is associated with chemotherapy-evoked vomiting [119]. TNF-α can excite vagal afferent terminals by augmenting Ca2+ release from sarco/endoplasmic reticulum stores via sensitization of RyRs which subsequently amplifies neurotransmission in the brainstem [15]. Cannabinoid CB1R agonists prevent the TNF-α-evoked sensitization of RyRs and therefore attenuate intracellular Ca2+ release from the sarco/endoplasmic reticulum stores [15]. Peripheral RyRs also play a critical role in agonist-evoked Ca2+ oscillations in gut epithelial cells [120]. Therefore, the ability of CB1R agonists in preventing both extracellular Ca2+ influx as well as intracellular Ca2+ release from the sarco/endoplasmic reticulum stores may be the fundamental mechanisms underlying the broad-spectrum antiemetic efficacy of CB1R cannabinoid agonists.

3.3. Glucocorticoids

Glucocorticoids, used primarily as anti-allergic and anti-inflammatory drugs. They are also effective, either alone or in combination with other antiemetics, for the suppression of nausea and vomiting. Indeed, dexamethasone (Figure 2), one of the clinically used glucocorticoids, is effective in reducing both acute and delayed chemotherapy-induced nausea and vomiting, and when combined with 5-HT3 or neurokinin NK1 antagonists, it is utilized in patients receiving high emetogenic chemotherapy [6]. Glucocorticoids’ antiemetic effect has been related to its inhibitory effects in the following facets: (i) glucocorticoids control the inflammatory response involved in mediating chemotherapy-induced nausea and vomiting by reducing the production of inflammatory mediators such as cytokines, chemokines, inducible nitric oxide synthase, and increasing the gene transcription of anti-inflammatory proteins [6]; (ii) glucocorticoids can inhibit 5-HT and substance P release, both of which can evoke emesis [6, 121], (iii) glucocorticoids can cross the blood-brain barrier and can exert direct central inhibitory effects on the nucleus tractus solitarius [6], which may be due to a decrease in abnormal elevation of cytosolic Ca2+ concentration as well as downstream Ca2+ signals and the maintenance of Ca2+ homeostasis within the cell [122], (iv) inhibitory actions of glucocorticoid could also be due to increased release of endocannabinoids, anandamide and 2-arachidonoylglycerol, evoked by glucocorticoid administration which will then be followed by subsequent CB1R activation as well as glucocorticoid facilitation of synaptic γ-aminobutyric acid (GABA) release and suppression of glutamate release [123, 124]. The endocannabinoid system is composed of CBRs, endocannabinoids and the enzymes involved in their synthesis. Anandamide and 2-arachidonoylglycerol are among the well-studied endocannabinoids and endogenous activators of CBRs [125]. The role of CB1R agonists as antiemetics was discussed in Section 2.2. It has been suggested that dexamethasone may decrease motion sickness through modulation of the endocannabinoid/CB1 receptor system on the terminals of the nucleus tractus solitarius neurons that project to the output neurons of the DMNV as well as by endocannabinoid/CB1 receptor system-mediated inhibition of transmitter release from interneurons of the nucleus tractus solitarius [99, 126]. Selective elevation of 2-arachidonoylglycerol by inhibition of its major metabolic enzyme monoacylglycerol lipase, have been shown to suppress lithium chloride evoked vomiting in the house musk shrew (Suncus murinus) [127]. However, intraperitoneal administration of the endocannabinoid 2-arachidonoylglycerol can evoke vomiting in the least shrew in a dose-dependent manner probably via its rapid metabolism to arachidonic acid which is also a potent emetogen in this species [128]. Moreover, the cancer chemotherapeutic agent cisplatin can increase 2-arachidonoylglycerol but not anandamide levels in the least shrew brain [129].

Advertisement

4. Perspective in developing new antiemetic candidates

4.1. Antiemetic efficacy of LTCC blockers in the least shrew model of emesis

Nifedipine along with amlodipine, are among the most studied of Ca2+ channels blockers, and both belong to the dihydropyridine subgroup of LTCC antagonists. Relative to nifedipine, a fast and short-acting LTCC antagonist with a plasma half-life of 1.2 h, amlodipine is slow and longer acting, more extensively bound to plasma protein, with a larger volume of distribution, more gradual elimination, with a half-life of over 30 h [130, 131, 132, 133, 134]. We have evaluated the antiemetic efficacy of both nifedipine and amlodipine (Figure 2) by assessing mean emesis frequency and the percentage of shrews vomiting, and demonstrated that both LTCC blockers [45, 46] behave as broad-spectrum antiemetics when delivered systemically against diverse specific emetogens, including FPL 64176 (10 mg/kg, i.p.), the peripherally-acting and non-selective 5-HT3R agonist 5-HT (5 mg/kg, i.p.), the peripherally/centrally-acting and more selective 5-HT3R agonist 2-Methyl-5-HT (5 mg/kg, i.p.), the dopamine D2R-preferring agonist quinpirole (2 mg/kg, i.p.), the non-selective dopamine D2R agonist apomorphine (2 mg/kg, i.p.), the nonselective cholinergic agonist pilocarpine (2 mg/kg, i.p.), the M1-preferring cholinergic agonist McN-A343 (2 mg/kg, i.p.), and the selective neurokinin NK1R agonist GR73632 (5 mg/kg, i.p.). The vomiting behavior was recorded for 30 min. Our results suggest that both amlodipine and nifedipine act by suppressing the influx of extracellular Ca2+, thereby delay the onset as well as protecting least shrews from vomiting, further supporting our proposed Ca2+ hypothesis of emesis. Nifedipine appears to be more potent than amlodipine against vomiting caused by FPL64176, 5-HT, 2-Methyl-5-HT, GR73632, quinpirole and McN-A343. These potency disparities could be explained in terms of their pharmacokinetic and pharmacodynamic differences [130, 131, 132, 133, 134, 135, 136, 137, 138, 139].

Unlike the above tested emetogens which can evoke vomiting within minutes of administration, cisplatin (10 mg, i.p.) requires more exposure time (30–45 min) to begin to induce emesis since only its metabolites are emetogenic. The relative efficacy of amlodipine (5 mg/kg., i.p.) in reducing the frequency of cisplatin-evoked early vomiting by 80% compared with the observed lack of antiemetic action of nifedipine up to 20 mg/kg [45, 46], could be explained in terms of positively charged amlodipine associating more slowly with LTCCs, requiring more exposure time not only to reach its sites of action, but also to compensate for its slower receptor binding kinetics, which can lead to a more gradual onset of antagonism [140]. In addition, intracerebroventricular microinjection of another LTCC antagonist, nitrendipine, has been shown to attenuate nicotine-induced vomiting in the cat [141], which further supports the discussed broad-spectrum antiemetic efficacy of nifedipine and amlodipine as observed in the least shrew model. Cisplatin-based chemotherapeutics induce both immediate and delayed vomiting in humans and in vomit-competent animals [16, 142, 143]. In the least shrew, cisplatin (10 mg/kg, i.p.) causes emesis over 40 h with respective peak early- and delayed-phases occurring at 1–2 and 32–34 h post-injection [144]. Amlodipine, due to its unique pharmacokinetics, may offer practical advantages over other calcium antagonists in cisplatin-evoked delayed emesis.

4.2. Potentiation of antiemetic efficacy of 5-HT3R antagonists when combined with LTCC blockers

In 1996 Hargreaves and co-workers [20] demonstrated that members of all three major classes of LTCC antagonists can prevent the ability of the 5-HT3 receptor-selective agonist 1-(m-chlorophenyl)-biguanide to increase intracellular Ca2+ concentration in cell lines that possess either one or both of these two different Ca2+-ion channels. The latter interaction is not competitive since the binding site for the different classes of LTCC antagonists appear not to be the same as the serotonin 5-HT3R binding site itself (i.e., the orthosteric site) but instead, is an allosteric site in the 5-HT3 receptor channel complex. Furthermore, 5-HT release from enterochromaffin cells can be prevented by antagonists of both 5-HT3Rs and LTCCs [145, 146]. These findings provide possible mechanisms via which antagonists of both LTCCs and 5-HT3Rs can mutually prevent the biochemical and behavioral effects of their corresponding selective agonists, including the vomiting behavior induced by their corresponding selective agonists FPL64176 and 2-Methyl-5-HT as we reported previously [45]. We have further demonstrated that when non-effective antiemetic doses of their selective antagonists (nifedipine and palonosetron, respectively) are combined [45], the combination significantly and in an additive manner attenuate both the frequency and the percentage of shrews vomiting in response to either FPL 64176 or 2-Methyl-5-HT. Furthermore, although nifedipine alone up to 20 mg/kg dose failed to protect shrews from acute cisplatin-induced vomiting, its 0.5 mg/kg dose, significantly potentiated the antiemetic efficacy of a non-effective (0.025 mg/kg) as well as a semi-effective (0.5 mg/kg) dose of palonosetron. In another study we also utilized a combination of non-effective doses of amlodipine (0.5 mg/kg or 1 mg/kg) with a non- or semi-effective dose of the 5-HT3R antagonist palonosetron (0.05 or 0.5 mg/kg) [46]. The combined antiemetic doses produced a similar additive efficacy against vomiting induced by either FPL 64176 or cisplatin. In fact relative to each antagonist alone, the combination was at least 4 times more potent in reducing the vomit frequency and provided more protection against FPL 64176-induced vomiting. The observed additive antiemetic efficacy of a combination of 5-HT3- (and/or possibly NK1-) with LTCC-antagonists in the least shrew suggests that such a combination should provide greater emesis protection in cancer patients receiving chemotherapy in a manner similar to that reported between 5-HT3- and NK1-receptor antagonists both in the laboratory [144, 147] and in the clinic [148]. Although in our investigation, the mechanism underlying the additive antiemetic efficacy of combined low doses of LTCC antagonists with 5-HT3R antagonists was not directly studied, the published literature points to their interaction at the signal transduction level involving Ca2+ [20, 149, 150].

4.3. Clinical use of LTCC blockers as anti-nausea/antiemetic medication

There are several published clinical case reports that demonstrate Ca2+ channel blockers may provide protection against several causes of nausea and vomiting. The LTCC antagonist flunarizine (Figure 2) was shown to reduce cyclic vomiting on acute basis in one patient [151] and prophylactically in 8 other patients [152]. Gabapentin is a gamma-aminobutyric acid (GABA) analog and is predominantly used in the clinic for the management of pain [3]. Gabapentin binds to the alpha-2/delta auxiliary subunits of voltage-gated Ca2+ channels (VGCCs) (i.e., LTCCs), and exerts inhibitory actions on trafficking and activation kinetics of VGCCs [153] (Figure 2). Moreover, several other reports indicate that gabapentin can also be used as a well-tolerated, less-expensive and promising anti-nausea and antiemetic agent in diverse conditions including: postoperative nausea and vomiting [154, 155], moderately or highly emetogenic chemotherapy-induced nausea and vomiting, particularly effective against delayed chemotherapy-induced nausea and vomiting [156], and both acute and delayed nausea induced by chemotherapy [157], as well as hyperemesis gravidarum [158]. When combined with dexamethasone, gabapentin can also significantly reduce the 24-h incidence of postoperative nausea and vomiting [159]. Alpha-2/delta subunits of VGCCs control transmitter release and further facilitate excitatory transmission [153]. Gabapentin’s interaction with neuronal alpha-2/delta subunits of VGCCs and subsequent downregulation of neuronal Ca2+ signaling in emesis relevant sites, such as the dorsal vagal complex, is postulated to play a critical role in its anti-nausea and anti-vomiting effects [3].

4.4. Intracellular Ca2+ release channels: possible targets for suppression of emesis

A functional and physical linkage between LTCCs and RyRs appears to exist and plays an important role in intracellular Ca2+ release following voltage-dependent Ca2+ entry through LTCCs during neuronal depolarization to generate a transient increase in cytosolic Ca2+ [160, 161, 162]. Physical attachment of IP3Rs to plasma membrane Ca2+ influx channels through conformational coupling has also been proposed as one of the mechanisms connecting depletion of internal Ca2+ stores with stimulation of extracellular Ca2+ influx [163]. For example, Ca2+ release from IP3Rs was shown to couple with extracellular Ca2+ influx through LTCCs in non-excitable cells such as Jurkat human T lymphocytes [164] and drosophila S2 cells [165], as well as in excitable cells such as submucosal neurons in the rat distal colon [166]. We have found that 5-HT3R-mediated vomiting triggered by 5 mg/kg 2-Methyl-5-HT is insensitive to the intracellular Ca2+ release channel IP3R antagonist 2-APB, but in contrast, was dose-dependently suppressed by the RyR antagonist, dantrolene [25]. Furthermore, a combination of the semi-effective doses of amlodipine and dantrolene was more potent than each antagonist being tested alone [25]. Significant reductions (70–85%) in the frequency of Ca2+ mobilizer thapsigargin-evoked vomiting (see Section 1.2) were observed when shrews were pretreated with antagonists of either IP3Rs (2-APB at 1 and 2.5 mg/kg, i.p.)- or RyRs (dantrolene at 2.5 and 5 mg/kg, i.p.)-ER luminal Ca2+ release channels. Moreover, while a mixture of 2-APB (1 mg/kg) and dantrolene (2.5 mg/kg) did not offer additional protection than what was afforded when each drug administered alone, a combination of the latter doses of 2-APB plus dantrolene with a partially effective dose of nifedipine (2.5 mg/kg), led to a complete elimination of thapsigargin-evoked vomiting [70]. In another set of experiments [167], we found that pretreatment with the IP3R inhibitor 2-APB causes a significant reduction in NK1R agonist GR73632-induced emesis, however the RyR inhibitor dantrolene did not. Thus, RyRs and IP3Rs can be differentially modulated by various emetogens and their antagonists provide further efficacy when combined with LTCC antagonists (Figure 2). Suppression of Ca2+ release from the sarco/endoplasmic reticulum stores through IP3Rs and RyRs may be additional targets for the prevention of nausea and vomiting.

4.5. Ca2+-related signaling pathways in emesis

4.5.1. The role of cAMP-PKA in vomiting

In mammals, cyclic AMP (cAMP) is synthesized by 10 adenylate cyclase isoforms [168]. One of the best-studied second messenger molecules downstream of selected G-protein coupled receptors is cAMP. It is an example of a transient and diffusible second messenger involved in signal propagation by integrating multiple intracellular signaling pathways [169]. cAMP activates protein kinase A (PKA) which results in phosphorylation of downstream intracellular signals. The adenylyl cyclase/cAMP/PKA signaling pathway can phosphorylate Ca2+ ion-channels found on the plasma membrane and intracellular IP3Rs [110]. These Ca2+ channels respectively increase extracellular Ca2+-influx and intracellular Ca2+-release [110]. The emetic role of cAMP has been well established (Figure 5), since microinjection of cAMP analogs (e.g., 8-bromocAMP) or forskolin (to enhance endogenous levels of cAMP) in the brainstem dorsal vagal complex emetic locus area postrema, not only can increase electrical activity of local neurons, but also induces vomiting in dogs [170]. Moreover, administration of 8-chlorocAMP as a potential chemotherapeutic in cancer patients can evoke nausea and vomiting [171]. Furthermore, phosphodiesterase inhibitors (PDEI) such as rolipram prevent cAMP metabolism and consequently increase cAMP tissue levels, which leads to excessive nausea and vomiting in humans [172]. In fact, one major side-effect of older PDEIs is excessive nausea and vomiting which often precludes their use in the clinical setting [173]. In addition, we have demonstrated that increased brain cAMP levels evoke vomiting which can be prevented by SQ22536 (Figure 5), an inhibitor of adenylyl cyclase [109]. Moreover, PKA-phosphorylation is associated with peak vomit frequency during both immediate- and delayed-phases of vomiting caused by either cisplatin or cyclophosphamide in the least shrew [109, 144, 149] (Figure 5).

Figure 5.

Summarized behavioral and biochemical evidence for intracellular signaling molecules (cAMP, PKA, CaMKII, ERK1/2, PKC) related to emesis based on the least shrew emesis model. First, cyclic AMP (cAMP) is synthesized by adenylate cyclase and cAMP activates protein kinase A (PKA) [110, 168]. The adenylyl cyclase/cAMP/PKA signaling pathway can mediate vomiting. Indeed, increased levels of endogenous cAMP can evoke vomiting in animal models [109, 170] as well as humans [171, 172, 173], which can be prevented by adenylate cyclase inhibitor SQ22536 [109]. Evoked PKA-phosphorylation is associated with peak vomit frequency during both immediate- and delayed-phases of vomiting caused by cancer chemotherapeutics including cisplatin and cyclophosphamide in the least shrew [109, 144, 149]. In addition, Ca2+/calmodulin kinase IIα (CaMKIIα) and extracellular signal-regulated protein kinases 1 and 2 (ERK1/2) phosphorylation in the least shrew brainstem were elevated in vomiting evoked by the 5-HT3R agonist 2-Methyl-5-HT [46], thapsigargin [70], or the selective NK1R agonist GR73632 [167]. Phosphorylation of protein kinase Cα/βII (PKCα/βII) and ERK1/2 in least shrew brainstem were also upregulated in the vomiting induced by cisplatin [144, 149].

4.5.2. Activation and inhibition of CaMKII, ERK1/2, PKC, and Akt are correspondingly linked to emesis induction and prevention

Vomit-associated Ca2+ mobilization as well as time-dependent Ca2+/calmodulin kinase IIα (CaMKIIα) and extracellular signal-regulated protein kinases 1 and 2 (ERK1/2) phosphorylation in the least shrew brainstem occurs: (i) following 5-HT3R-evoked vomiting caused by its selective agonist 2-Methyl-5-HT [46], (ii) thapsigargin-induced emesis in the least shrew [70], as well as (iii) NK1R-mediated vomiting evoked by the selective NK1R agonist GR73632 in the least shrew [167] (Figure 5). Our additional behavioral evidence that inhibitors of CaMKII or ERK1/2 attenuate the evoked emesis provides further credence for involvement of CaMKII and ERK1/2 downstream of the discussed emetic receptors/effectors. Furthermore, other published evidence support phosphorylation of protein kinase Cα/βII (PKCα/βII) and ERK1/2 in least shrew brainstem are associated with cisplatin-induced emesis [144, 149] (Figure 5). In fact significant upregulation of ERK1/2 phosphorylation occurs with peak vomit frequency during both the immediate and delayed phases of emesis caused by cisplatin in the least shrew [144, 149].

It has been suggested that glucocorticoids’ antiemetic efficacy could be due to their anti-inflammatory effects [174] probably via a reduction in the synthesis of prostaglandins and leukotrienes [175], both of which can be increased during chemotherapy [6]. Although not all, but several prostaglandins (e.g., PGE2 and PGF2a) and cysteinyl leukotrienes (e.g., LTC4and LTD4), appear to be potent emetogens [72, 149, 176, 177]. Our findings demonstrate that unlike other leukotrienes (e.g., LTA4, LTB4 and LTF4), the above discussed leukotrienes are effective emetogens with the following potency order: LTC4 = LTD4 > LTE4. Regarding LTC4, the evoked vomiting was shown to be suppressed in a dose-dependent manner in the least shrew by the antiasthmatic drug pranlukast, the corresponding cysteinyl leukotrienes receptor 1 (CysLT1R) antagonist [72]. Although not available in the USA, the cost of other members of this class of drugs (montelukast and Zafirlukast) that are sold in the USA is less than one dollar per pill. Based on pranlukast’s efficacy against LTC4-induced vomiting [72], we envisaged it may have potential utility against cisplatin-evoked emesis. Our most recent publication [178] shows the potential of pranlukast (currently used for the treatment of various respiratory disorders including asthma), as a new class of antiemetic for the suppression of the acute- and delayed- phases of cisplatin-evoked vomiting in the least shrew. An intraperitoneal (i.p.) dose of 10 mg/kg pranlukast by itself significantly reduced the mean frequency of vomits by 70% and fully protected 46% of least shrews during the delayed-phase of cisplatin (10 mg/kg, i.p.)-evoked vomiting. Although pranlukast tended to substantially reduce both the mean frequency of vomits and the number of shrews vomiting during the early-phase, these reductions failed to attain significance. When pranlukast was combined with a first (tropisetron)- or a second (palonosetron)-generation 5-HT3R antagonist, it potentiated their antiemetic efficacy during both acute- and delayed-phases of cisplatin-evoked vomiting. Moreover, pranlukast potentiated the antiemetic efficacy of serotonin 5-HT3 receptor antagonists, tropisetron and palonosetron, against chemotherapy-induced nausea and vomiting. In fact per hour efficacy antiemetic profile of pranlukast combined with palonosetron or tropisetron during both phases of chemotherapy-induced nausea and vomiting in the least shrew resembles those of: (i) the NK1 receptor antagonist netupitant (5 mg/kg) plus palonosetron (0.1 mg/kg) in the same species [144]; (ii) netupitant plus ondansetron in ferrets [179]; and (iii) ondansetron plus aprepitant in combination with dexamethasone in ferrets [179]; and (iv) palonosetron plus netupitant in combination with dexamethasone in ferrets [179]. If analogs of pranlukast such as montelukast and zafirlukast can also provide similar antiemetic potential, then clinical trials should be initiated since this class of drugs are relatively inexpensive than available effective antiemetic regimens against chemotherapy-induced nausea and vomiting. Our related biochemical data indicates the mechanisms of antiemetic action of pranlukast are linked to suppression of cisplatin-elicited PKCα/βII, ERK1/2 and PKA activation (phosphorylation) in the least shrew brainstem [178]. Moreover, suppression of these signaling molecules may be shared in the anti-inflammatory signaling pathway of pranlukast.

When antiemetic mechanism of action of pranlukast against LTC4-induced vomiting or cisplatin-induced responses is discussed, Ca2+ is also an essential element. Montelukast and pranlukast were found to inhibit nucleotide-induced Ca2+ mobilization in a human monocyte-macrophage-like cell line, DMSO-differentiated U937 [180]. CysLT1 receptors belonging to the rhodopsin family of the G protein-coupled receptor genes respond to LTD4 with a strong increase in cytosolic Ca2+ concentration partially sensitive to pertussis toxin, and with the activation of the Ras-MAPK cascade totally dependent upon Gi/o [144]. These signaling effects were totally inhibited by various specific CysLT1-receptor antagonists, and CysLT1 antagonists inhibit both the P2Y agonist-induced activation of phospholipase C and intracellular Ca2+ mobilization [144].

Advertisement

5. Conclusion

Chemotherapy-induced nausea and vomiting is a particularly distressing side-effect of chemotherapeutics for oncology patients both physically and psychologically. The use of 5-HT3R antagonists combined with NK1R antagonists, has enhanced physician’s ability to further suppress nausea, the rates of acute- and delayed-vomiting in cancer patients receiving chemotherapy. In addition to the commonly reported adverse effects of these agents (including headache, diarrhea, constipation, hiccups, and fatigue), many patients still experience nausea and delayed vomiting [181, 182, 183]. Furthermore, the use of second generation 5-HT3R and NK1R antagonists for the prevention of chemotherapy-induced nausea and vomiting is currently cost-prohibitive for most patients in the world. Mechanisms that cause nausea are only partially understood and probably in part overlap with those of vomiting. There are still unmet needs for newer and less expensive therapeutic options to improve the treatment across the entire spectrum of chemotherapy-induced nausea and vomiting. Additional studies should involve combinations of agents that inhibit other neurotransmitter systems involved in nausea and vomiting.

As concluded in Figure 2, this systematic review shows clear evidence that Ca2+ modulation is an important contributor to antiemetic and probably anti-nausea signaling pathways. LTCC blockers, antagonists of intracellular IP3Rs and RyRs Ca2+ release channels as well as CysLT1R antagonists have the potential to provide less expensive (e.g., nifedipine, amlodipine, dantrolene, and pranlukast) broad-spectrum antiemetic agents for the clinic against diverse causes of nausea and vomiting. The discussed findings from the least shrew should help to open new avenues of research in other established animal models of emesis as well as in patients, targeting not only the already discussed Ca2+ channels, but also other Ca2+ channels that exist on both the plasma membrane and the membranes of intracellular organs such as the sarco/endoplasmic reticulum and mitochondria.

Advertisement

Acknowledgments

This work was supported by NIH/NCI grant (RO1CA207287) and Western University of Health Science startup fund to NAD.

References

  1. 1. Kottschade L, Novotny P, Lyss A, Mazurczak M, Loprinzi C, Barton D. Chemotherapy-induced nausea and vomiting: Incidence and characteristics of persistent symptoms and future directions NCCTG N08C3 (Alliance). Supportive Care in Cancer. 2016;24:2661-2667
  2. 2. Darmani NA, Ray AP. Evidence for a re-evaluation of the neurochemical and anatomical bases of chemotherapy-induced vomiting. Chemical Reviews. 2009;109:3158-3199
  3. 3. Guttuso T Jr. Gabapentin's anti-nausea and anti-emetic effects: A review. Experimental Brain Research. 2014;232:2535-2539
  4. 4. Ray AP, Chebolu S, Darmani NA. Receptor-selective agonists induce emesis and Fos expression in the brain and enteric nervous system of the least shrew (Cryptotis parva). Pharmacology, Biochemistry, and Behavior. 2009;94:211-218
  5. 5. Rogers RC, Nasse JS, Hermann GE. Live-cell imaging methods for the study of vagal afferents within the nucleus of the solitary tract. Journal of Neuroscience Methods. 2006;150:47-58
  6. 6. Chu CC, Hsing CH, Shieh JP, Chien CC, Ho CM, Wang JJ. The cellular mechanisms of the antiemetic action of dexamethasone and related glucocorticoids against vomiting. European Journal of Pharmacology. 2014;722:48-54
  7. 7. Travagli RA, Anselmi L. Vagal neurocircuitry and its influence on gastric motility. Nature Reviews. Gastroenterology & Hepatology. 2016;13:389-401
  8. 8. Babic T, Browning KN. The role of vagal neurocircuits in the regulation of nausea and vomiting. European Journal of Pharmacology. 2014;722:38-47
  9. 9. Rojas C, Slusher BS. Mechanisms and latest clinical studies of new NK1 receptor antagonists for chemotherapy-induced nausea andvomiting: Rolapitant and NEPA (netupitant/palonosetron). Cancer Treatment Reviews. 2015;41:904-913
  10. 10. Hesketh PJ. Chemotherapy-induced nausea and vomiting. The New England Journal of Medicine. 2008;358:2482-2494
  11. 11. Seaton G, Hogg EL, Jo J, Whitcomb DJ, Cho K. Sensing change: The emerging role of calcium sensors in neuronal disease. Seminars in Cell & Developmental Biology. 2011;22:530-535
  12. 12. Katz B, Miledi R. A study of synaptic transmission in the absence of nerve impulses. The Journal of Physiology. 1967;192:407-436
  13. 13. Zuccotti A, Clementi S, Reinbothe T, Torrente A, Vandael DH, Pirone A. Structural and functional differences between l-type calcium channels: Crucial issues for future selective targeting. TIPS. 2011;32:366-375
  14. 14. Rogers RC, Van Meter MJ, Hermann GE. Tumor necrosis factor potentiates central vagal afferent signaling by modulating ryanodine channels. The Journal of Neuroscience. 2006;26:12642-12646
  15. 15. Rogers RC, Hermann GE. Tumor necrosis factor activation of vagal afferent terminal calcium is blocked by cannabinoids. The Journal of Neuroscience. 2012;32:5237-5241
  16. 16. Darmani NA, Crim JL, Janoyan JJ, Abad J, Ramirez J. A re-evaluation of the neurotransmitter basis of chemotherapy-induced immediate and delayed vomiting: Evidence from the least shrew. Brain Research. 2009;1248:40-58
  17. 17. Suzuki Y, Inoue T, Ra C. L-type Ca2+ channels: A new player in the regulation of Ca2+ signaling, cell activation and cell survival in immune cells. Molecular Immunology. 2010;47:640-648
  18. 18. Lin YR, Kao PC, Chan MH. Involvement of Ca2+ signaling in tachykinin-mediated contractile responses in swine trachea. Journal of Biomedical Science. 2005;12:547-558
  19. 19. Miyano K, Morioka N, Sugimoto T, Shiraishi S, Uezono Y, Nakata Y. Activation of the neurokinin-1 receptor in the rat spinal astrocytes induced Ca2+ release from IP3-sensitive Ca2+ stores and extracellular Ca2+ influx through TRPC3. Neurochemistry International. 2010;57:923-934
  20. 20. Hargreaves AC, Gunthorpe MJ, Taylor CW, Lumis SC. Direct inhibition of 5-hydroxytryptamine3 receptors by antagonists of L-type Ca2+ channels. Molecular Pharmacology. 1996;50:1284-1294
  21. 21. Homma K, Kitamura Y, Ogawa h, Oka K. Serotonin induces the increase in intracellular Ca2+ that enhances neurite outgrowth in PC12 cells via activation of 5-HT3 receptors and voltage gated channels. Journal of Neuroscience Research. 2006;84:316-325
  22. 22. Hutchinson TE, Zhong W, Chebolu S, Wilson SM, Darmani NA. L-type calcium channels contribute to 5-HT3-receptor-evoked CaMKIIalpha and ERK activation and induction of emesis in the least shrew (Cryptotis parva). European Journal of Pharmacology. 2015;755:110-118
  23. 23. Ronde P, Nichols RA. 5-HT3 receptors induce rises in cytosolic and nuclear calcium in NG108-15 via calcium-induced calcium release. Cell Calcium. 1997;22:357-365
  24. 24. Takenouchi T, Munekata E. Serotonin increases Ca2+ concentration in PC12h cells: Effect of tachikinin peptides. Neuroscience Letters. 1998;24:141-144
  25. 25. Zhong W, Hutchinson TE, Chebolu S, Darmani NA. Serotonin 5-HT3 receptor-mediated vomiting occurs via the activation of Ca2+/CaMKII-dependent ERK1/2 signaling in the least shrew (Cryptotis parva). PLoS One. 2014;9:e104718
  26. 26. Aman TK, Shen RY, Haj-Dahmane S. D2-like dopamine receptors depolarize dorsal raphe serotonin neurons through the activation of nonselective cationic conductance. Journal of Pharmacology and Experimental Therapeutics. 2007;320:376-385
  27. 27. Wu J, Dougherty JJ, Nichols RA. Dopamine receptor regulation of Ca2+ levels in individual isolated nerve terminals from rat striatum: Comparison of presynaptic D1-like and D2-like receptors. J. Neuroscience. 2006;98:481-494
  28. 28. Oliveira L, Correia-de-Sa P. Protein kinase A and cav1 (L-type) channels are common targets to facilitatory adenosine and muscarinic m1 receptors on rat motoneurons. Neuro-Signals. 2005;14:262-272
  29. 29. Sculptoreano A, Yoshimura N, de Goroat WC, Somogyi GT. Proteinkinase C is involved in M1-muscarinic receptor-mediated facilitation of l-type Ca2+ channels in neurons of the major pelvic ganglion of the adult male rat. Neurochemical Research. 2001;26:933-942
  30. 30. Barajas M, Andrade A, Hernandez-Hernandez O, Felix R, Arias-Montano J-A. Histamine-induced Ca2+ entry in human astrocytoma U373 MG cells: Evidence for involvement of store-operated channels. Journal of Neuroscience Research. 2008;86:3456-3468
  31. 31. Yoshimoto K, Hattori Y, Houzen H, Kanno M, Yasuda K. Histamine H1-receptor-mediated increase in the Ca2+ transient without a change in the Ca2+ current in electrically stimulated Guinea-pig atrial myocytes. British Journal of Pharmacology. 1998;124:1744-1750
  32. 32. Ono T, Inoue M, Rashid MH, Sumikawa K, Ueda H. Stimulation of peripheral nociceptor endings by low dose morphine and its signaling mechanism. Neurochemistry International. 2002;41:399-407
  33. 33. Smart D, Hirst RA, Hirota HK, Grandy DK, Lambert DG. The effects of recombinant rat u-opioid receptor activation in CHO cells on phospholipase C, [Ca2+]i and adenylyl cyclase. British Journal of Pharmacology. 1997;120:1165-1171
  34. 34. Splettstoesser F, Florea A-M, Busselberg D. IP3 receptor antagonist, 2-APB, attenuates cisplatin induced Ca2+-influx in Hela-S3 cells and prevents activation of calpain and induction of apoptosis. British Journal of Pharmacology. 2007;151:1176-1186
  35. 35. Almirza WHM, Peters PHJ, van Zoelen EJJ, Theuvenet APR. Role of TRPC channels, Stim1 and Orai1 in PGF2a-induced calcium signaling in NRK fibroblasts. Calcium Cell. 2012;51:12-21
  36. 36. Rodríguez-Lagunas MJ, Martín-Venegas R, Moreno JJ, Ferrer R. PGE2 promotes Ca2+-mediated epithelial barrier disruption through EP1 and EP4 receptors in Caco-2 cell monolayers. American Journal of Physiology. Cell Physiology. 2010;299:C324-C334
  37. 37. Hagbom M, Sharma S, Lundgren O, Svensson L. Towards a human rotavirus disease model. Current Opinion in Virology. 2012;2:408-418
  38. 38. Hyser JM, Collinson-Pautz MR, Utama B, Estes MK. Rotavirus Disrupts Calcium Homeostasis by NSP4 viroporin Activity. Mbio.asm.org. e00265. 2010
  39. 39. Poppoff MR, Poulain B. Bacterial toxins and the nervous system: Neurotoxins and multipotential toxins interacting with neuronal cells. Toxins. 2010;2:683-737
  40. 40. Timar Peregrin T, Svensson M, Ahlman H, Jodal M, Lundgren O. The effects on net fluid transport of noxious stimulation of jejunal mucosa in anesthetized rats. Acta Physiologica Scandinavica. 1999;166:55-64
  41. 41. Darmani NA. Serotonin 5-HT3 receptor antagonists prevent cisplatin-induced emesis in Cryptotis parva: A new experimental model of emesis. Journal of Neural Transmission. 1998;105:1143-1154
  42. 42. Darmani NA, Johnson JC. Central and peripheral mechanisms contribute to the antiemetic actions of delta-9-tetrahydrocannabinol against 5-hydroxytryptophan-induced emesis. European Journal of Pharmacology. 2004;488:201-212
  43. 43. Suzuki Y, Yoshimaru T, Inoue T, Ra C. Ca v 1.2 L-type Ca2+ channel protects mast cells against activation-induced cell death by preventing mitochondrial integrity disruption. Molecular Immunology. 2009;46:2370-2380
  44. 44. Yoshimaru T, Suzuki Y, Inoue T, Ra C. L-type Ca2+ channels in mast cells: Activation by membrane depolarization and distinct roles in regulating mediator release from store-operated Ca2+ channels. Molecular Immunology. 2009;46:1267-1277
  45. 45. Darmani NA, Zhong W, Chebolu S, Vaezi M, Alkam T. Broad-spectrum antiemetic potential of the L-type calcium channel antagonist nifedipine and evidence for its additive antiemetic interaction with the 5-HT(3) receptor antagonist palonosetron in the least shrew (Cryptotis parva). European Journal of Pharmacology. 2014;722:2-12
  46. 46. Zhong W, Chebolu S, Darmani NA. Broad-spectrum antiemetic efficacy of the L-type calcium channel blocker amlodipine in the least shrew (Cryptotis parva). Pharmacology, Biochemistry, and Behavior. 2014;120:124-132
  47. 47. Lomax RB, Gallego S, Novalbos J, García AG, Warhurst G. L-type calcium channels in enterochromaffin cells from Guinea pig and human duodenal crypts: An in situ study. Gastroenterology. 1999;117:1363-1369
  48. 48. Minami M, Endo T, Yokota H, Ogawa T, Nemoto M, Hamaue N, Hirafuji M, Yoshioka M, Nagahisa A, Andrews PL. Effects of CP-99, 994, a tachykinin NK(1) receptor antagonist,on abdominal afferent vagal activity in ferrets: Evidence for involvement of NK(1) and 5-HT(3) receptors. European Journal of Pharmacology. 2001;428:215-220
  49. 49. Garaschuk O, Yaari Y, Konnerth A. Release and sequestration of calcium by ryanodine-sensitive stores in rat hippocampal neurons. Journal de Physique. 1997;502:13-30
  50. 50. Gómez-Viquez L, Guerrero-Serna G, García U, Guerrero-Hernández A. SERCA pump optimizes Ca2+ release by a mechanism independent of store filling in smooth muscle cells. Biophysical Journal. 2003;85:370-380
  51. 51. Gómez-Viquez NL, Guerrero-Serna G, Arvizu F, García U, Guerrero-Hernández A. Inhibition of SERCA pumps induces desynchronized RyR activation inoverloaded internal Ca2+ stores in smooth muscle cells. American Journal of Physiology. Cell Physiology. 2010;298:C1038-C1046
  52. 52. Cheng KT, Ong HL, Liu X, Ambudkar IS. Contribution and regulation of TRPC channels in store-operated Ca2+ entry. Current Topics in Membranes. 2013;71:149-179
  53. 53. Feske S. Calcium signaling in lymphocyte activation and disease. Nature Reviews. Immunology. 2007;7:690-702
  54. 54. Parekh AB, Putney JW Jr. Store-operated calcium channels. Physiological Reviews. 2005;85:757-810
  55. 55. Beltran-Parrazal L, Fernandez-Ruiz J, Toledo R, Manzo J, Morgado-Valle C. Inhibition of endoplasmic reticulum Ca2+ ATPase in preBötzinger complex of neonatal rat does not affect respiratory rhythm generation. Neuroscience. 2012;224:116-124
  56. 56. Michelangeli F, East JM. A diversity of SERCA Ca2+ pump inhibitors. Biochemical Society Transactions. 2011;39:789-797
  57. 57. Solovyova N, Verkhratsky A. Neuronal endoplasmic reticulum acts as a single functional Ca2+ store shared by ryanodine and inositol-1,4,5-trisphosphate receptors as revealed by intra-ER [Ca2+] recordings in single rat sensory neurones. Pflügers Archiv. 2003;446:447-454
  58. 58. Smaili SS, Cavalcanti PM, Oshiro ME, Ferreira AT, Jurkiewicz A. Ca2+ release-activated channels in rat stomach smooth muscle cells. European Journal of Pharmacology. 1998;342:119-122
  59. 59. Van Geldre LA, Lefebvre RA. Nitrergic relaxation in rat gastric fundus: Influence of mechanism of induced tone and possible role of sarcoplasmic/endoplasmic reticulum Ca2+ ATPase. Life Sciences. 2004;74:3259-3274
  60. 60. Petkov GV, Boev KK. The role of sarcoplasmic reticulum and sarcoplasmic reticulum Ca2+-ATPase in the smooth muscle tone of the cat gastric fundus. Pflügers Archiv. 1996;431:928-935
  61. 61. Denmeade SR, Jakobsen CM, Janssen S, Khan SR, Garrett ES, Lilja H, Christensen SB, Isaacs JT. Prostate-specific antigen-activated thapsigargin prodrug as targeted therapy for prostate cancer. Journal of the National Cancer Institute. 2003;95:990-1000
  62. 62. Jackisch C, Hahm HA, Tombal B, McCloskey D, Butash K, Davidson NE, Denmeade SR. Delayed micromolar elevation in intracellular calcium precedes induction of apoptosis in thapsigargin-treated breast cancer cells. Clinical Cancer Research. 2000;6:2844-2850
  63. 63. Dubois C, Vanden Abeele F, Sehgal P, Olesen C, Junker S, Christensen SB, Prevarskaya N, Møller JV. Differential effects of thapsigargin analogues on apoptosis of prostate cancer cells: Complex regulation by intracellular calcium. The FEBS Journal. 2013;280:5430-5440
  64. 64. Yamaguchi H, Bhalla K, Wang HG. Bax plays a pivotal role in thapsigargin-induced apoptosis of human colon cancer HCT116 cells by controlling Smac/Diablo and Omi/HtrA2 release from mitochondria. Cancer Research. 2003;63:1483-1489
  65. 65. Lee TJ, Kim SH, Choi YH, Song KS, Park JW, Kwon TK. Overexpression of Par-4 enhances thapsigargin-induced apoptosis via down-regulation of XIAP and inactivation of Akt in human renal cancer cells. Journal of Cellular Biochemistry. 2008;103:358-368
  66. 66. Denmeade SR, Mhaka AM, Rosen DM, Brennen WN, Dalrymple S, Dach I, Olesen C, Gurel B, Demarzo AM, Wilding G, Carducci MA, Dionne CA, Møller JV, Nissen P, Christensen SB, Isaacs JT. Engineering a prostate-specific membrane antigen-activated tumor endothelial cell prodrug for cancer therapy. Science Translational Medicine. 2012;4:140ra86
  67. 67. Denmeade SR, Isaacs JT. The SERCA pump as a therapeutic target: Making a “smart bomb” for prostate cancer. Cancer Biology & Therapy. 2005;4:14-22
  68. 68. Denmeade SR, Isaacs JT. Engineering enzymatically activated “molecular grenades” for cancer. Oncotarget. 2012;3:666-667
  69. 69. Doan NT, Paulsen ES, Sehgal P, Møller JV, Nissen P, Denmeade SR, Isaacs JT, Dionne CA, Christensen SB. Targeting thapsigargin towards tumors. Steroids. 2015;97:2-7
  70. 70. Zhong W, Chebolu S, Darmani NA. Thapsigargin-induced activation of Ca(2+)-CaMKII-ERK in brainstem contributes to substance P release and induction of emesis in the least shrew. Neuropharmacology. 2016;103:195-210
  71. 71. Seoane A, Massey PV, Keen H, Bashir ZI, Brown MW. L-type voltage-dependent calcium channel antagonists impair perirhinal long-term recognition memory and plasticity processes. The Journal of Neuroscience. 2009;29:9534-9544
  72. 72. Chebolu S, Wang Y, Ray AP, Darmani NA. Pranlukast prevents cysteinyl leukotriene-induced emesis in the least shrew (Cryptotis parva). European Journal of Pharmacology. 2010;628:195-201
  73. 73. Ray AP, Griggs L, Darmani NA. Delta 9-tetrahydrocannabinol suppresses vomiting behavior and Fos expression in both acute and delayed phases of cisplatin-induced emesis in the least shrew. Behavioural Brain Research. 2009;196:30-36
  74. 74. Jinnah HA, Yitta S, Drew T, Kim BS, Visser JE, Rothstein JD. Calcium channel activation and self-biting in mice. Proceedings of the National Academy of Sciences of the United States of America. 1999;96:15228-15232
  75. 75. Jinnah HA, Sepkuty JP, Ho T, Yitta S, Drew T, Rothstein JD, Hess EJ. Calcium channel agonists and dystonia in the mouse. Movement Disorders. 2000;15:542-551
  76. 76. Jinnah HA, Egami K, Rao L, Shin M, Kasim S, Hess EJ. Expression of c-FOS in the brain after activation of L-type calcium channels. Developmental Neuroscience. 2003;25:403-411
  77. 77. Andrews PLR, Rudd JA. The role of Tachykinins and the Tachykinin NK1 receptor in nausea and emesis. In: Holzer P, editor. Tachykinins, Springer: Handbook of Experimental Pharmacology; 2004. pp. 359-440
  78. 78. Hesketh PJ, Grunberg SM, Gralla RJ, Warr DG, Roila F, de Wit R, Chawla SP, Carides AD, Ianus J, Elmer ME, Evans JK, Beck K, Reines S, Horgan KJ. Aprepitant Protocol 052 Study Group. The oral neurokinin-1 antagonist aprepitant for the prevention of chemotherapy-induced nausea and vomiting: A multinational, randomized, double-blind, placebo-controlled trial in patients receiving high-dose cisplatin--the Aprepitant Protocol 052 Study Group. Journal of Clinical Oncology. 2003;21:4112-4119
  79. 79. Janicki PK. Management of acute and delayed chemotherapy-induced nausea and vomiting: Role of netupitant-palonosetron combination. Therapeutics and Clinical Risk Management. 2016;12:693-699
  80. 80. Gralla RJ, Bosnjak SM, Hontsa A, Balser C, Rizzi G, Rossi G, Borroni ME, Jordan K. A phase III study evaluating the safety and efficacy of NEPA, a fixed-dose combination of netupitant and palonosetron, for prevention of chemotherapy-induced nausea and vomiting over repeated cycles of chemotherapy. Annals of Oncology. 2014;25:1333-1339
  81. 81. Hesketh PJ, Aapro M, Jordan K, Schwartzberg L, Bosnjak S, Rugo H. A review of NEPA, a novel fixed antiemetic combination with the potential for enhancing guideline adherence and improving control of chemotherapy-induced nausea and vomiting. BioMed Research International. 2015;2015:651879
  82. 82. Keating GM. Netupitant/palonosetron: A review in the prevention of chemotherapy-induced nausea and vomiting. Drugs. 2015;75:2131-2141
  83. 83. Rojas C, Stathis M, Thomas AG, Massuda EB, Alt J, Zhang J, Rubenstein E, Sebastiani S, Cantoreggi S, Snyder SH, Slusher B. Palonosetron exhibits unique molecular interactions with the 5-HT3 receptor. Anesthesia and Analgesia. 2008;107:469-478
  84. 84. Rojas C, Thomas AG, Alt J, Stathis M, Zhang J, Rubenstein EB, Sebastiani S, Cantoreggi S, Slusher BS. Palonosetron triggers 5-HT(3) receptor internalization and causes prolonged inhibition of receptor function. European Journal of Pharmacology. 2010;626:193-199
  85. 85. Rojas C, Li Y, Zhang J, Stathis M, Alt J, Thomas AG, Cantoreggi S, Sebastiani S, Pietra C, Slusher BS. The antiemetic 5-HT3 receptor antagonist Palonosetron inhibits substance P-mediated responses in vitro and in vivo. The Journal of Pharmacology and Experimental Therapeutics. 2010;335:362-368
  86. 86. Ruzza C, Rizzi A, Malfacini D, Molinari S, Giuliano C, Lovati E, Pietra C, Calo' G. In vitro and in vivo pharmacological characterization of Pronetupitant, a prodrug of the neurokinin 1 receptor antagonist Netupitant. Peptides. 2015;69:26-32
  87. 87. Chasen MR, Rapoport BL. Rolapitant for the treatment of chemotherapy-induced nausea and vomiting: A review of the clinical evidence. Future Oncology. 2016;12:763-778
  88. 88. Duffy RA, Morgan C, Naylor R, Higgins GA, Varty GB, Lachowicz JE, Parker EM.Rolapitant (SCH 619734): A potent, selective and orally active neurokinin NK1 receptor antagonist with centrally-mediated antiemetic effects in ferrets. Pharmacology, Biochemistry, and Behavior. 2012;102:95-100
  89. 89. Rapoport B, Chua D, Poma A, Arora S, Wang Y, Fein LE. Study of rolapitant, a novel, long-acting, NK-1 receptor antagonist, for the prevention of chemotherapy-induced nausea and vomiting (CINV) due to highly emetogenic chemotherapy (HEC). Supportive Care in Cancer. 2015;23:3281-3288
  90. 90. Todaro B. Cannabinoids in the treatment of chemotherapy-induced nausea and vomiting. Journal of the National Comprehensive Cancer Network. 2012;10:487-492
  91. 91. Bolognini D, Rock EM, Cluny NL, Cascio MG, Limebeer CL, Duncan M, Stott CG, Javid FA, Parker LA, Pertwee RG. Cannabidiolic acid prevents vomiting in Suncus murinus and nausea-induced behaviour in rats by enhancing 5-HT1A receptor activation. British Journal of Pharmacology. 2013;168:1456-1470
  92. 92. Van Sickle MD, Oland LD, Ho W, Hillard CJ, Mackie K, Davison JS, Sharkey KA. Cannabinoids inhibit emesis through CB1 receptors in the brainstem of the ferret. Gastroenterology. 2001;121:767-774
  93. 93. Wang Y, Ray AP, McClanahan BA, Darmani NA. The antiemetic interaction of Delta9-tetrahydrocannabinol when combined with tropisetron or dexamethasone in the least shrew. Pharmacology, Biochemistry, and Behavior. 2009;91:367-373
  94. 94. Darmani NA, Janoyan JJ, Crim J, Ramirez J. Receptor mechanism and antiemetic activity of structurally-diverse cannabinoids against radiation-induced emesis in the least shrew. European Journal of Pharmacology. 2007;563:187-196
  95. 95. Pertwee RG. Targeting the endocannabinoid system with cannabinoid receptor agonists: Pharmacological strategies and therapeutic possibilities. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences. 2012;367:3353-3363
  96. 96. Punyamurthula NS, Hingorani T, Adelli G, Gul W, ElSohly MA, Repka MA, Majumdar S. Controlled release tablet formulation containing natural Δ9-tetrahydrocannabinol. Drug Development and Industrial Pharmacy. 2015;7:1-7
  97. 97. Tafelski S, Häuser W, Schäfer M. Efficacy, tolerability, and safety of cannabinoids for chemotherapy-induced nausea and vomiting—A systematic review of systematic reviews. Schmerz. 2016;30:14-24
  98. 98. Parker LA, Rock EM, Limebeer CL. Regulation of nausea and vomiting by cannabinoids. British Journal of Pharmacology. 2011;163:1411-1422
  99. 99. Van Sickle MD, Oland LD, Mackie K, Davison JS, Sharkey KA. Delta9-tetrahydrocannabinol selectively acts on CB1 receptors in specific regions of dorsal vagal complex to inhibit emesis in ferrets. American Journal of Physiology. Gastrointestinal and Liver Physiology. 2003;285:G566-G576
  100. 100. Ware MA, Daeninck P, Maida V. A review of nabilone in the treatment of chemotherapy-induced nausea and vomiting. Therapeutics and Clinical Risk Management. 2008;4:99-107
  101. 101. Sharkey KA, Darmani NA, Parker LA. Regulation of nausea and vomiting by cannabinoids and the endocannabinoid system. European Journal of Pharmacology. 2014;722:134-146
  102. 102. Darmani NA. Methods evaluating cannabinoid and endocannabinoid effects on gastrointestinal functions. Methods in Molecular Medicine. 2006;123:169-189
  103. 103. Galligan JJ. Cannabinoid signalling in the enteric nervous system. Neurogastroenterology and Motility. 2009;21:899-902
  104. 104. Boesmans W, Ameloot K, van den Abbeel V, Tack J, Vanden Berghe P. Cannabinoid receptor 1 signalling dampens activity and mitochondrial transport in networks of enteric neurons. Neurogastroenterology and Motility. 2009;21:958-e77
  105. 105. Coutts AA, Pertwee RG. Evidence that cannabinoid-induced inhibition of electrically evoked contractions of the myenteric plexus—longitudinal muscle preparation of Guinea-pig small intestine can be modulated by Ca2+ and cAMP. Can. Journal of Physiology and Pharmacology. 1998;76:340-346
  106. 106. Derbenev AV, Stuart TC, Smith BN. Cannabinoids suppress synaptic input to neurones of the rat dorsal motor nucleus of the vagus nerve. The Journal of Physiology. 2004;559(Pt 3):923-938
  107. 107. Derbenev AV, Monroe MJ, Glatzer NR, Smith BN. Vanilloid-mediated heterosynaptic facilitation of inhibitory synaptic input to neurons of the rat dorsal motor nucleus of the vagus. The Journal of Neuroscience. 2006;26:9666-9672
  108. 108. Szabo B, Schlicker E. Effects of cannabinoids on neurotransmission. Handbook of Experimental Pharmacology. 2005;168:327-365
  109. 109. Alkam T, Chebolu S, Darmani NA. Cyclophosphamide causes activation of protein kinase (PKA) in the brainstem of vomiting least shrews (Cryptotis parva). European Journal of Pharmacology. 2014;722:156-164
  110. 110. Yao L, Fan P, Jiang Z, Gordon A, Mochly-Rosen D, Diamond I. Dopamine and ethanol cause translocation of ePKC associated with eRACK: Cross-talk between cAMP-dependent protein kinase a and protein kinase c signaling pathways. Journal of Pharmacology and Experimental Therapeutics. 2008;73:1105-1112
  111. 111. Lalonde MR, Jollimore CA, Stevens K, Barnes S, Kelly ME. Cannabinoid receptor-mediated inhibition of calcium signaling in rat retinal ganglion cells. Molecular Vision. 2006;12:1160-1166
  112. 112. Lozovaya N, Min R, Tsintsadze V, Burnashev N. Dual modulation of CNS voltage-gated calcium channels by cannabinoids: Focus on CB1 receptor-independent effects. Cell Calcium. 2009;46:154-162
  113. 113. Straiker A, Stella N, Piomelli D, Mackie K, Karten HJ, Maguire G. Cannabinoid CB1 receptors and ligands in vertebrate retina: Localization and function of an endogenous signaling system. Proceedings of the National Academy of Sciences of the United States of America. 1999;96:14565-14570
  114. 114. Straiker A, Sullivan JM. Cannabinoid receptor activation differentially modulates ion channels in photoreceptors of the tiger salamander. Journal of Neurophysiology. 2003;89:2647-2654
  115. 115. Yang W, Li Q, Wang SY, Gao F, Qian WJ, Li F, Ji M, Sun XH, Miao Y, Wang Z. Cannabinoid receptor agonists modulate calcium channels in rat retinal müller cells. Neuroscience. 2016;313:213-224
  116. 116. Shi B, Yang R, Wang X, Liu H, Zou L, Hu X, Wu J, Zou A, Liu L. Inhibition of 5-HT(3) receptors-activated currents by cannabinoids in rat trigeminal ganglion neurons. Journal of Huazhong University of Science and Technology. Medical Sciences. 2012;32:265-271
  117. 117. Galeotti N, Quattrone A, Vivoli E, Norcini M, Bartolini A, Ghelardini C. Different involvement of type 1, 2, and 3 ryanodine receptors in memory processes. Learning & Memory. 2008;15:315-323
  118. 118. Ledbetter MW, Preiner JK, Louis CF, Mickelson JR. Tissue distribution of ryanodine receptor isoforms and alleles determined by reverse transcription polymerase chain reaction. The Journal of Biological Chemistry. 1994;269:31544-31551
  119. 119. Martin J, Howard SC, Pillai A, Vogel P, Naren AP, Davis S, Ringwald-Smith K, Buddington K, Buddington RK. The weaned pig as a model for doxorubicin-induced mucositis. Chemotherapy. 2014;60:24-36
  120. 120. Verma V, Carter C, Keable S, Bennett D, Thorn P. Identification and function of type-2 and type-3 ryanodine receptors in gut epithelial cells. The Biochemical Journal. 1996;319:449-454
  121. 121. Higa GM, Auber ML, Altaha R, Piktel D, Kurian S, Hobbs G, Landreth K. 5-Hydroxyindoleacetic acid and substance P profiles in patients receiving emetogenic chemotherapy. Journal of Oncology Pharmacy Practice. 2006;12:201-209
  122. 122. Suwanjang W, Holmström KM, Chetsawang B, Abramov AY. Glucocorticoids reduce intracellular calcium concentration and protects neurons against glutamate toxicity. Cell Calcium. 2013;53:256-263
  123. 123. Di S, Malcher-Lopes R, Halmos KC, Tasker JG. Nongenomic glucocorticoid inhibition via endocannabinoid release in the hypothalamus: A fast feedback mechanism. The Journal of Neuroscience. 2003;23:4850-4857
  124. 124. Di S, Malcher-Lopes R, Marcheselli VL, Bazan NG, Tasker JG. Rapid glucocorticoid-mediated endocannabinoid release and opposing regulation of glutamate and gamma-aminobutyric acid inputs to hypothalamic magnocellular neurons. Endocrinology. 2005;146:4292-4301
  125. 125. Darmani NA, Zhong W, Endo- phyto- and synthetic-cannabinoids and the cannabinoid-induced hyperemesis syndrome. Gastro Open Journal. 2017;SE(1):S1-S8
  126. 126. Zheng Y, Wang XL, Mo FF, Li M. Dexamethasone alleviates motion sickness in rats in part by enhancing the endocannabinoidsystem. European Journal of Pharmacology. 2014;727:99-105
  127. 127. Parker LA, Niphakis MJ, Downey R, et al. Effect of selective inhibition of monoacylglycerol lipase (MAGL) on acute nausea, anticipatory nausea, and vomiting in rats and Suncus murinus. Psychopharmacology. 2015;232:583-593
  128. 128. Darmani NA. The potent emetogenic effects of the endocannabinoid, 2-AG (2-arachidonoyl-glycerol) are blocked by delta (9)-tetrahydrocannabinol and other cannnabinoids. The Journal of Pharmacology and Experimental Therapeutics. 2002;300:34-42
  129. 129. Darmani NA, McClanahan BA, Trinh C, et al. Cisplatin increases brain 2-arachidonoylglycerol (2-AG) and concomitantly reduces intestinal 2-AG and anandamide levels in the least shrew. Neuropharmacology. 2005;49:502-513
  130. 130. Burges R, Moisey D. Unique pharmacologic properties of amlodipine. The American Journal of Cardiology. 1994;73:2A-9A
  131. 131. Abernethy DR. Amlodipine: Pharmacokinetic profile of a low-clearance calcium antagonist. Journal of Cardiovascular Pharmacology. 1991;17(Suppl 1):S4-S7
  132. 132. Burges RA. The pharmacological profile of amlodipine in relation to ischaemic heart disease. Postgraduate Medical Journal. 1991;67(Suppl 3):S9-S15
  133. 133. Reid JL, Meredith PA, Donnelly R, Elliott HL. Pharmacokinetics of calcium antagonists. Journal of Cardiovascular Pharmacology. 1988;12(Suppl 7):S22-S26
  134. 134. Toal CB, Meredith PA, Elliott HL. Long-acting dihydropyridine calcium-channel blockers and sympathetic nervous system activity in hypertension: A literature review comparing amlodipine and nifedipine GITS. Blood Pressure. 2012;21(Suppl 1):S3-S10
  135. 135. Burges RA, Dodd MG, Gardiner DG. Pharmacologic profile of amlodipine. The American Journal of Cardiology. 1989;64:10I-18I; discussion 18I-20I
  136. 136. Croom KF, Wellington K. Modified-release nifedipine: A review of the use of modified-release formulations in the treatment of hypertension and angina pectoris. Drugs. 2006;66:497-528
  137. 137. Meredith PA, Reid JL. Differences between calcium antagonists: Duration of action and trough to peak ratio. Journal of Hypertension. 1993;11(Suppl 1):S21-S26
  138. 138. Burges RA, Dodd MG. Amlodipine. Cardiovascular Drug Reviews. 1990;8:25-44
  139. 139. Nayler WG, Gu XH. The unique binding properties of amlodipine: A long-acting calcium antagonist. Journal of Human Hypertension. 1991;5(Suppl 1):S55-S59
  140. 140. Qu Y-L, Sugiyama K, Hattori K, Yamamoto A, Watanabe K, Nagatoma T. Slow association pf positively charged Ca2+ channel antagonist amlodipine to dihhydropyridine receptor sites in the rat brain membranes. General Pharmacology. 1996;27:137-140
  141. 141. Samardzic R, Bajcetic M, Beleslin DB. Opposite effects of ethanol and nitrendipine on nicotine-induced emesis and convulsions. Alcohol. 1999;18:215-219
  142. 142. Hesketh PJ, Van Belle S, Aapro M, Tattersall FD, Naylor RJ, Hargreaves R, Carides AD, Evans JK, Horgan KJ. Differential involvement of neurotransmitters through the time course of cisplatin-induced emesis as revealedby therapy with specific receptor antagonists. European Journal of Cancer. 2003;39:1074-1080
  143. 143. Rudd JA, Andrews PLR. Mechanisms of acute, delayed, and anticipatory emesis induced by anticancer therapies. In: Hesketh PJ, editor. Management of Nausea and Vomiting in Cancer and Cancer Treatment. Sudbury, MA: Jones and Bartlett; 2005. pp. 15-65
  144. 144. Darmani NA, Zhong W, Chebolu S, Mercadante F. Differential and additive suppressive effects of 5-HT3 (palonosetron)- and NK1 (netupitant)-receptor antagonists on cisplatin-induced vomiting and ERK1/2, PKA and PKC activation. Pharmacology, Biochemistry and Behavior. 2015;131:104-111
  145. 145. Minami M, Endo T, Hirafugi M, Hamaue N, Liu Y, Hiroshige T, Nemoto M, Saito H, Yoshioka M. Pharmacological aspects of anticancer drug-induced emesis with emphasis on serotonin release and vagal nerve activity. Pharmacology & Therapeutics. 2003;99:149-165
  146. 146. Minami M, Taquchi S, Kikuchi T, Endo T, Hamaue N, Hiroshige T, Liu Y, Yue W, Hirafuji M. Effects of fluvoxamine, a selective serotonin re-uptake inhibitor, on serotonin release from the mouse isolated ileum. Research Communications in Molecular Pathology and Pharmacology. 2003:113-114, 115-131
  147. 147. Darmani NA, Chebolu S, Amos B, Alkam T. Synergistic antiemetic interactions between serotonergic 5-HT3- and tachykininergic NK1-receptor antagonists in the least shrew (Cryptotis parva). Pharmacology, Biochemistry, and Behavior. 2011;99:573-579
  148. 148. Warr D. Management of highly emetogenic chemotherapy. Current Opinion in Oncology. 2012;24:371-375
  149. 149. Darmani NA, Dey D, Chebolu S, Amos B, Kandpal R, Alkam T. Cisplatin causes over-expression of tachykinin NK(1) receptors and increases ERK1/2- and PKA-phosphorylation during peak immediate- and delayed-phase emesis in the least shrew (Cryptotis parva) brainstem. European Journal of Pharmacology. 2013;698:161-169
  150. 150. Stathis M, Pietra C, Rojas C, Slusher BS. Inhibition of substance P-mediated responses in NG108-15 cells by netupitant and palonosetron exhibit synergistic effects. European Journal of Pharmacology. 2012;689:25-30
  151. 151. Van Driessche A, Sermigin E, Paemeleire K, van Coster R, Vogelaers D. Cyclic vomiting syndrome: Case report and short review of the literature. Acta Clinica Belgica. 2012;67:123-126
  152. 152. Kothare SV. Efficacy of flunarizine in the prophylaxis of cyclical vomiting syndrome and abdominal migraine. European Journal of Paediatric Neurology. 2005;9:23-26
  153. 153. Patel R, Dickenson AH. Mechanisms of the gabapentinoids and α 2 δ-1 calcium channel subunit in neuropathic pain. Pharmacological Research Perspect. 4, e00205. 2016
  154. 154. Achuthan S, Singh I, Varthya SB, Srinivasan A, Chakrabarti A, Hota D. Gabapentin prophylaxis for postoperative nausea and vomiting in abdominal surgeries: A quantitative analysis of evidence from randomized controlled clinical trials. British Journal of Anaesthesia. 2015;114:588-597
  155. 155. Memari F, Jadidi R, Noroozi A, Mohammadbeigi A, Falahati J. Protecting effect of gabapentin for nausea and vomiting in the surgery of cesarean after spinal anesthesia. Anesthesia, Essays and Researches. 2015;9:401-404
  156. 156. Cruz FM, de Iracema Gomes Cubero D, Taranto P, Lerner T, Lera AT, da Costa Miranda M, da Cunha Vieira M, de Souza Fêde AB, Schindler F, Carrasco MM, de Afonseca SO, Pinczowski H, del Giglio A. Gabapentin for the prevention of chemotherapy- induced nausea and vomiting: A pilot study. Supportive Care in Cancer. 2012;20:601-606
  157. 157. Guttuso T Jr, Roscoe J, Griggs J. Effect of gabapentin on nausea induced by chemotherapy in patients with breast cancer. Lancet. 2003;361:1703-1705
  158. 158. Guttuso T Jr, Robinson LK, Amankwah KS. Gabapentin use in hyperemesis gravidarum: A pilot study. Early Human Development. 2010;86:65-66
  159. 159. Misra S, Parthasarathi G, Vilanilam GC. The effect of gabapentin premedication on postoperative nausea, vomiting, and pain in patients on preoperative dexamethasone undergoing craniotomy for intracranial tumors. Journal of Neurosurgical Anesthesiology. 2013;25:386-391
  160. 160. Berrout J, Isokawa M. Homeostatic and stimulus-induced coupling of the L-type Ca2+ channel to the ryanodine receptor in the hippocampal neuron in slices. Cell Calcium. 2009;46:30-38
  161. 161. Katoh H, Schlotthauer K, Bers DM. Transmission of information from cardiac dihydropyridine receptor to ryanodine receptor: Evidence from BayK 8644 effects on resting Ca(2+) sparks. Circulation Research. 2000;87:106-111
  162. 162. Resende RR, da CJL, Kihara AH, Adhikari A, Lorencon E. Intracellular Ca2+ regulation during neuronal differentiation of murine embryonal carcinoma and mesenchymal stem cells. Stem Cells and Development. 2010;19:379-394
  163. 163. Li N, Sul JY, Haydon PG. A calcium-induced calcium influx factor, nitric oxide, modulates the refilling of calcium stores in astrocytes. The Journal of Neuroscience. 2003;23:10302-10310
  164. 164. Wang Q, Wu YJ. Lysophosphatidylcholine induces Ca(2+) mobilization in Jurkat human T lymphocytes and CTLL-2 mouse T lymphocytes by different pathways. European Journal of Pharmaceutical Sciences. 2011;44:602-609
  165. 165. Wang P, Wang Q, Yang L, Qin QL, Wu YJ. Characterization of lysophosphatidylcholine-induced changes of intracellular calcium in Drosophila S2 cells. Life Sciences. 2015;131:57-62
  166. 166. Rehn M, Bader S, Bell A, Diener M. Distribution of voltage-dependent and intracellular Ca2+ channels in submucosal neurons from rat distal colon. Cell and Tissue Research. 2013;353:355-366
  167. 167. Zhong W, Chebolu S, Darmani NA. Intracellular signaling involved in neurokinin NK1 receptor-mediated emesis in the least shrew. Abstract and poster presentation (416.06). 2017 Neuroscience Conference
  168. 168. Halls ML, Cooper DM. Adenylyl cyclase signalling complexes—Pharmacological challenges and opportunities. Pharmacology & Therapeutics. 2017;172:171-180
  169. 169. Gancedo JM. Biological roles of cAMP: Variations on a theme in the different kingdoms of life. Biological Reviews of the Cambridge Philosophical Society. 2013;88:645-668
  170. 170. Carpenter DO, Briggs DB, Knox AP, Strominger N. Excitation of area postrema neurons by transmitters, peptides and cyclic nucleotides. Journal of Neurophysiology. 1988;59:358-369
  171. 171. Propper DJ, Saunders MP, Salisbury AJ, Long L, O'Byrne KJ, Braybrooke JP, et al. Phase I study of the novel cyclic AMP (cAMP) analogue 8-chloro-cAMP in patients with cancer: Toxicity, hormonal, and immunological effects. Clinical Cancer Research. 1999;5:1682-1689
  172. 172. Mori F, Perez-Torres S, De Caro R, Porzionato A, Macchi V, Belata J, et al. The human area postrema and other nuclei related to the emetic reflex express cAMP phosphor diesterases4B and 4D. Journal of Chemical Neuroanatomy. 2010;40:36-42
  173. 173. Vanmierlo T, Creemers P, Akkerman S, van Duinen M, Sambeth A, De Vry J, et al. The PDE4 inhibitor roflumilast improves memory in rodents at non-emetic doses. Behavioural Brain Research. 2016;303:26-33
  174. 174. Sam TS, Chan SW, Rudd JA, Yeung JH. Action of glucocorticoids to antagonise cisplatin-induced acute and delayed emesis in the ferret. European Journal of Pharmacology. 2001;417:231-237
  175. 175. Vane JR, Botting RM. Mechanism of action of anti-inflammatory drugs. Scandinavian Journal of Rheumatology. Supplement. 1996;102:9-21
  176. 176. Kan KK, Jones RL, Ngan MP, Rudd JA. Actions of prostanoids to induce emesis and defecation in the ferret. European Journal of Pharmacology. 2002;453:299-308
  177. 177. Kan KK, Jones RL, Ngan MP, Rudd JA. Action of prostanoids on the emetic reflex of Suncus murinus (the house musk shrew). European Journal of Pharmacology. 2003;477:247-251
  178. 178. Darmani NA, Chebolu S, Zhong W, Kim WD, Narlesky M, Adams J, et al. The anti-asthmatic drug pranlukast suppresses the delayed-phase vomiting and reverses intracellular indices of emesis evoked by cisplatin in the least shrew (Cryptotis parva). European Journal of Pharmacology. 2017;809:20-31
  179. 179. Rudd JA, Ngan MP, Lu Z, Higgins GA, Giuliano C, Lovati E, et al. Profile of antiemetic activity of Netupitant alone or in combination with palonosetron and dexamethasone in ferrets and Suncus murinus (house musk shrew). Frontiers in Pharmacology. 2016;7:263
  180. 180. Mamedova L, Capra V, Accomazzo MR, Gao ZG, Ferrario S, Fumagalli M, Abbracchio MP, Rovati GE, Jacobson KA. CysLT1 leukotriene receptor antagonists inhibit the effects of nucleotides acting at P2Y receptors. Biochemical Pharmacology. 2005;71:115-125
  181. 181. Navari RM. The safety of antiemetic medications for the prevention of chemotherapy-induced nausea andvomiting. Expert Opinion on Drug Safety. 2016;15:343-356
  182. 182. Slatkin NE. Cannabinoids in the treatment of chemotherapy-induced nausea and vomiting: Beyond prevention of acute emesis. The Journal of Supportive Oncology. 2007;5:1-9
  183. 183. Sommariva S, Pongiglione B, Tarricone R. Impact of chemotherapy-induced nausea and vomiting on health-related quality of life and resource utilization: A systematic review. Critical Reviews in Oncology/Hematology. 2016;99:13-36

Written By

Weixia Zhong and Nissar A. Darmani

Submitted: 21 April 2018 Reviewed: 07 May 2018 Published: 05 November 2018