Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Apr 1.
Published in final edited form as: Headache. 2014 Apr;54(4):619–639. doi: 10.1111/head.12323

Ion channels and migraine

Jin Yan 1, Gregory Dussor 2
PMCID: PMC4005924  NIHMSID: NIHMS572747  PMID: 24697223

Abstract

Migraine is one of the most common neurological disorders. Despite its prevalence, the basic physiology of the molecules and mechanisms that contribute to migraine headache is still poorly understood, making the discovery of more effective treatments extremely difficult. The consistent presence of head-specific pain during migraine suggests an important role for activation of the peripheral nociceptors localized to the head. Accordingly, this review will cover the current understanding of the biological mechanisms leading to episodic activation and sensitization of the trigeminovascular pain pathway, focusing on recent advances regarding activation and modulation of ion channels.

Keywords: dural afferents, TRESK, ASICs, TRPA1, TRPV1, TRPV4, TRPM8, P2X3

Introduction

Migraine is a complex neurologic disorder characterized by episodic headache and several associated manifestations, including nausea, vomiting, phonophobia and/or photophobia.1 In one third of patients, attacks of disabling headaches are preceded by a focal, transient neurological phenomenon termed ‘aura’ (migraine with aura).2 According to the American Migraine Prevalence and Prevention study, the cumulative incidence is 43% in women and 18% in men.3 Given that the prevalence of migraine is highest during the ages of 25 to 55,4 often the peak ages of productivity, the disorder imposes a huge burden on patients, their families, employers and society. Despite its prevalence, the basic physiology of the molecules and mechanisms that contribute to migraine headache is still poorly understood, making the discovery of more effective treatments extremely difficult.

The pain of migraine likely requires activation of primary afferent neurons that innervate cranial tissues,5 although this is still a matter of debate within the field. Migraine aura, however, is more widely accepted to arise from cortical spreading depression (CSD) in cerebral sensory cortices.6 Migraine is known to be trigged by specific factors in approximately 75% of patients.7 These factors include stress, menstruation, noise, odors, heat, head/neck movement, neck pain, and certain foods.8, 9 It is not clear how these factors bring on migraine attacks. Psychophysical, evoked potentials, transcranial magnetic stimulation (TMS), and magnetoencephalographic studies indicate enhanced cortical responsiveness to diverse stimuli in migraineurs.1012 Most studies observe a lowered threshold for neuronal activation in migraineurs compared with controls, which suggest that migraine may result from an altered state of neuronal excitability. Whether this occurs within central as well as peripheral neurons is unclear. A contribution from peripheral neurons innervating the cranial region provides the simplest physiological mechanism for the head-specific pain of migraine. But one interesting and unresolved question is whether there is increased peripheral input during migraine or whether increased neuronal excitability within the CNS allows pain to develop in response to tonic activity in peripheral neurons. This review will focus on ion channel-based mechanisms that activate trigeminal sensory afferents under normal and sensitized conditions. Although a simple interpretation for a role of these channels is that they mediate increased neuronal activity from peripheral tissues, their activation may also be important for tonic activity.

Ion channels are pore-forming transmembrane proteins that allow the passage of ions into and out of a cell. They mediate cell excitability and are essential for proper signaling and cell function. Therefore, dysfunction or aberrant regulation of ion channels could result in disruption of excitation-inhibition balance, neuronal excitability and peripheral or central sensitization. In support of this view, several prophylactic drugs for migraine work (albeit non-selectively) by inhibiting the function of ion channels such as voltage-gated sodium channels or voltage-gated calcium channels.13 Although the site of action of these drugs is not clear, modulation of ion channels leading to amelioration of neuronal hyperexcitability nonetheless represents a currently utilized therapeutic approach and an attractive strategy for migraine prophylactic drug discovery. In the peripheral nervous system, considerable progress has been made in understanding the role of ion channels that may participate in the pathophysiology of migraine. Fluorescent retrograde labeling methods have allowed the study of the properties of primary afferent neurons innervating the dura (i.e. dural afferents), which are likely to be key players in transmitting peripheral input contributing to the pain of migraine. This review will focus on (1) ion channels expressed on dural afferents and (2) recent advances regarding activation and modulation of the ion channels on dural afferents. We have also included the TRESK potassium channel, which has recently been linked with inherited migraine. Although the site of action of TRESK channels in migraine pathophysiology is not clear, recent studies indicate that this channel might be important for regulating excitability of primary afferent neurons.

The Properties of Dural Afferents

Advances in histological, electrophysiological and behavioral techniques have allowed a better understanding of the signaling cascades that may underlie migraine pathophysiology. The trigeminal sensory system is likely a key component in pain initiation and transmission in migraine.14, 15 Trigeminal nerves (primary afferent neurons) have cell bodies located in the trigeminal ganglion and they detect mechanical, chemical and temperature changes from intracranial and extracranial tissues, and send the information into the CNS. Primary afferent neurons mainly project to the upper cervical and medullary dorsal horn,16, 17 and sensory information then travels to higher brain regions. Anatomical studies demonstrate that only the meninges and large cranial vessels, but not the brain itself or its intrinsic blood vessels, are densely innervated by sensory fibers originating primarily from trigeminal ganglion.14, 18, 19 These data indicate that primary afferent neurons have the potential to detect and transmit noxious stimuli from cranial tissues. In vivo electrophysiological recording studies have demonstrated the ability of primary afferent neurons to respond to mechanical stimuli, such as punctate probing and stroking,20 thermal stimuli21 and chemical stimuli, such as KCl, capsaicin, acidic buffers and a mixture of inflammatory mediators (histamine, bradykinin, serotonin, and prostaglandin E2 at pH 5).20, 21 In addition, the only sensation that is evoked following stimulation of large cranial vessels and the meninges in humans is pain,22 indicating that the sensory innervation of the meninges is purely nociceptive. Therefore, similar to afferents innervating other tissues, meningeal sensory afferents are able to transmit and encode noxious stimuli, pointing to their important roles in headache pain processing.

One major question is where the noxious stimuli that can activate primary afferent trigeminal neurons originate. Preclinical studies show that primary afferent neurons can be sensitized by inflammatory mediators, allowing previously innocuous stimuli to become noxious or allowing spontaneous activity to develop. This hypothesis is supported by clinical observations that intracranial and circulating levels of various inflammatory mediators are significantly higher during migraine attacks.23, 24 Preclinical studies have also identified dural mast cells as a potential source for proinflammatory mediators. In addition to sensory fibers, the meninges are highly populated with mast cells.25, 26 These granulated cells reside within the dura and are in close proximity to blood vessels and nerve terminals.27 Stress (via the release of corticotrophin releasing factor or CRF),28 CGRP release,29 nitroglycerin infusion30 and increased estrogen levels,27 all factors associated with migraine in humans are known to trigger mast cell degranulation.31 Activation/degranulation of dural mast cells could therefore be a key event which links several seemingly unrelated triggers to the emergence of episodic headache. Preclinical studies also provide direct evidence demonstrating that mast cell degranulation by itself can activate meningeal nociceptors.32 Mast cell mediators, including serotonin (5-HT), prostaglandin I2 (PGI2), and to a lesser extent histamine can promote a robust sensitization and activation of meningeal nociceptors.33 Exposure of the dura to mast cell mediators in vivo leads to headache-like responses to normally subthreshold pH changes in a rat behavioral model of headache.34 These studies implicate dural mast cells as both “sensor” and “effector” cells that participate in detecting changes in the meninges induced by migraine triggers and promote the development of neurogenic inflammation. There could be a cycle occurring within the meninges where substances such as serotonin, histamine, and PGI2 (among others) activate/sensitize the afferents,33 initiate neurogenic inflammation which could then lead to more afferent activation causing further release of inflammatory mediators. The net outcome of this cycle is an amplified inflammatory cascade, greater activation/sensitization of dural afferents and a larger magnitude response.

Changes in the number of dural mast cell might contribute to headache attacks in migraine patients. In support of this hypothesis, a high prevalence of headache is observed in patients with mastocytosis, a disorder characterized by an increased number of tissue mast cells.35 Estrogen has been shown to promote an increase in dural mast cell density through recruitment of mast cells from spleen, which might contribute to the higher incidence of migraine in women.36 Taken together, clinical observations, preclinical studies, and anatomical and physiological characteristics of dural mast cells all point to an important role of these cells in migraine pathophysiology.

Another endogenous process proposed to promote meningeal inflammation is cortical spreading depression (CSD), thought to be the underlying mechanism for migraine aura. CSD is associated with a massive efflux of potassium,37 hydrogen ions,38 and neurotransmitters such as glutamate39 into extracellular space, which can cause a multitude of changes in expressions of growth factors and proinflammatory mediators, such as tumor necrosis factor-α (TNF-α) and IL-1β.40 These mediators have the potential to activate or sensitize meningeal nociceptors. Consistent with this hypothesis, preclinical studies confirm that CSD is able to induce a long lasting blood flow increase in the middle meningeal artery, cause plasma protein extravasation in dura mater, increase c-Fos staining in the ipsilateral trigeminal nucleus caudalis41 and cause delayed and long-term activation of meningeal nociceptors.42 One recent study demonstrated that following CSD, high-mobility group box 1 protein is released from cortical astrocytes through pannexin 1 channels, leading to increased blood flow in the middle meningeal artery, dural mast cell degranulation, and pain behaviors in mice.43 Although it is known that meningeal nociceptors can be activated or sensitized by proinflammatory mediators20, 33, 44, 45 or following CSD events42, 43 and mast cell degranulation,32 the mechanisms by which these events produce excitation of dural afferents are still not entirely clear. Multiple channels can contribute to activation of dural afferents (as with any population of nociceptor) in response to factors released during mast cell degranulation or CSD. The next section will focus on the current understanding of ionic mechanisms underlying activation and sensitization of dural afferents.

Ion Channels in Dural Afferents

Retrograde labeling techniques have been recently used in migraine research to study properties of afferents innervating the dura.44, 46, 47 Briefly, retrograde tracers, either 1,1-dioctadecyl-3,3,3,3-tetramethylindocarbocyanine perchlorate (DiI) or Fluorogold (FG), are applied to a section of dura surrounding the superior sagittal sinus bilaterally to label the nociceptors innervating the dura (i.e. dural afferents). This method provides specific labeling46 and the labeling process itself does not change the properties of the neurons.47 Utilizing this method, ion channels expressed on dural afferents have been identified and their properties have been studied (Table 1.).

Table 1.

Dural Afferent Transducers

Dural Afferent Transducers Evidence Potential Stimulating Events within the Dura Ref
ASICs
  1. 80% of dural afferents respond to acidic solution with ASIC3-like currents.

  2. Dural application of acidic solution induces migraine-related pain behavior through activation of ASICs.

mast cell degranulation secondary to stress, CGRP release, nitroglycerin infusion, increased estrogen level and CSD events 47
TRPA1
  1. 5.7% of identified dural afferents express TRPA1.

  2. TRPA1 agonists, mustard oil and the umbellulone evoke TRPA1-like currents in approximately 42% and 38% of dural afferents, respectively.

  3. Dural application of TRPA1 agonists produces migraine-related behaviors in a TRPA1-dependent manner.

exposure to a series of chemically diverse and highly reactive environmental agents and various electrophilic natural products or mechanical stimuli 46, 91, 92
TRPV1
  1. Nerve fibers in the dura mater exhibit TRPV1-immunoreactivity.

  2. 23.7% of identified dural afferents generate TRPV1 current.

Endogenous or exogenous mediators of inflammation have been shown to activate TRPV1 directly or lowered the activation threshold of TRPV1. 46, 101
TRPV4
  1. 56% and 49% of dural afferents generate currents in response to hypotonic solutions and the TRPV4 activator 4α-PDD.

  2. Dural application of hypotonic solution or 4α-PDD results in migraine-related pain behavior through activation of TRPV4.

sudden intracranial pressure change due to head jolts or rotation, breath-holding, sneezing or coughing 133
P2X 52% of dural afferents express P2X2 or P2X3 or both receptors. During CSD event, ATP is released into the extracellular space at levels exceeding 100 μM. 152

1. Acid Sensing Ion Channels (ASICs)

ASICs are voltage-insensitive cationic channels activated by increases in the concentration of extracellular protons.48 The ASIC family consists of 4 members, ASIC1-4, with several splice variants.49 Functional ASIC channels are assembled as heteromeric or homomultimeric channels. Different subunit composition of these channels gives rise to different pharmacology, current kinetics and sensitivity to pH changes.49 ASICs are widely expressed in the nervous system. ASIC1a and ASIC2 subunits are expressed in the central nervous system,5053 while almost all ASIC subunits are found in peripheral sensory neurons.52, 54, 55

Among all subunits, the ASIC3 subunit is of particular interest for several reasons. ASIC3 exhibits a sensitivity that allows it to respond to pH values at or near neutral pH. In addition to transient peak current, ASIC3 is also able to generate a non-desensitizing current in response to a physiologically relevant pH,56 making it suitable to detect prolonged and slow acidification. ASIC3 is highly expressed in sensory neurons and largely restricted to the periphery,49, 57 consistent with its role in detecting pH changes in peripheral tissues. Peripheral sensory neurons expressing ASIC3 innervate visceral organs including the colon and heart as well as skeletal muscles.5861 ASIC3 channels have been proposed to modulate painful conditions associated with these tissues, including angina, postoperative pain, various GI disorders and muscle pain.56, 58, 6063

With respect to migraine, ASICs on dural afferents have been proposed as a sensor of decreased extracellular pH within the dura.64 ASIC3 expression is found in most trigeminal neurons and approximately 80% of dural afferents express immunolabeling for ASIC3.34, 65 A recent electrophysiology study in dural afferents demonstrated that in response to acidic solution, 80% of dural afferents generate rapidly activating and rapidly desensitizing currents through activation of ASICs.47 These currents led to action potentials in dural afferents at physiologically relevant pH ranging from pH 6.8 to pH 7.0.47 Moreover, biophysical characterization indicated that ASIC-like currents in dural afferents are mediated by ASIC3 homomeric or heteromeric channels.47 In awake animals, dural application of acidic solution elicited migraine-related pain behavior through the activation of ASICs, providing support for their potential role in initiating migraine headache.47 This study demonstrated for the first time that small changes in extracellular pH can directly excite dural afferents via the opening of ASICs. Following pH stimulation, CGRP release is also increased in trigeminal ganglion neurons via activation of ASIC3,65 which can result in neurogenic inflammation and headache progression. A contribution of other ASIC subunits in migraine has also been reported. Amiloride was shown to block cortical spreading depression and inhibit trigeminal activation through an ASIC1-dependent mechanism in a preclinical study.66 Importantly, in a small clinical study of intractable migraine patients, amiloride reduced both aura and headache severity in 4 of 7 patients.66 This finding suggests the need for a larger-scale trial in a more general population of migraine patients. Further, since amiloride is a non-specific blocker of ASICs, the development of more selective pharmacological tools is necessary to clarify the role of ASIC subunits in human migraine pathophysiology.

One logical question surrounding an ASIC mechanism for migraine is where the pH drop would originate in the dura. The intragranular pH of isolated mast cells is approximately 5.5.67 Dural mast cells located in close proximity to nerve endings, and therefore mast cell degranulation secondary to migraine-related triggers, could acidify the surrounding local environment and directly activate dural afferents. Alternatively, CSD is shown to be accompanied by ischemia,68 which could induce a pH drop in the dura. Given the fact that ASICs are extremely sensitive to pH change and dural afferents express a high density of ASICs, it is possible that even small pH changes can activate these neurons and initiate headache. Prior studies show that CSD events and mast cell degranulation can activate dural afferents in anesthetized animals.32, 42 Further studies are essential to elucidate whether there is a role of ASICs in these pathophysiological conditions by utilizing either specific ASIC blockers or ASIC knockout animals.

2. Transient receptor potential channels (TRP channels)

TRP channels are a group of ion channels expressed on the cellular membrane as well as the endoplasmic reticulum that respond to a variety of stimuli, including heat, osmolarity changes, protons and various natural products.6971 The 28 mammalian TRPs can be divided into 6 subgroups: TRPC, TRPM, TRPV, TRPA, TRPP, TRPML based on their primary amino acid sequences.72 Activation of TRPs allows the influx of Ca2+ and Na+ and depolarization of the cell.69 TRP channels have been demonstrated to participate in sensing visual, gustatory, olfactory, auditory, mechanical, thermal, osmotic stimuli and environmental irritants.73 Therefore, TRP channels seem to be particularly well-suited candidates for transducing and encoding noxious stimuli in migraine pathophysiology.

2.1 Transient Receptor Potential Cation Channel A1 (TRPA1)

Expressed on sensory neurons74 as well as neurons in the airways,75, 76 TRPA1 is thought to mediate neuronal responses to a series of chemically diverse and highly reactive environmental agents such as formaldehyde,77 acrolein,78 chlorine,79 and cigarette smoke extract,80 various electrophilic natural products, including pungent plant derivatives like isothiocyanates,74 cinnamaldehyde,81 and allicin82 and by-products of oxidative and nitrative stress, such as nitrooleic acid,83 4-hydroxynonenal84 and reactive prostaglandins.85 A substantial amount of preclinical work has established a role for TRPA1 in pain transduction.86 Activation of the TRPA1 channel produces nociceptive behavior in rodents78 and a gain-of-function mutation of TRPA1 is linked with a heritable episodic pain syndrome in humans.87 TRPA1 is also thought to be involved in modulation of mechanotransduction in sensory neurons.8890

Recent evidence also proposes a role for TRPA1 in migraine pathophysiology, among which are two separate studies demonstrating TRPA1 expression on dural afferents. In mice, 5.7% of identified dural afferents expressed TRPA1,46 while in rats, two TRPA1 agonists, mustard oil (MO) and the environmental irritant umbellulone, evoked TRPA1-like currents in approximately 42% and 38% of dural afferents, respectively.91 The difference between the sensitivity of the detection methods, immunohistochemistry vs. electrophysiology, or species may account for the discrepancy in expression here. Umbellulone is the major volatile constituent of U. californica, a tree known to trigger violent headache crises.92 Recent studies demonstrated that exposure to umbellulone evokes nociceptive behavior, meningeal vasodilation and calcitonin gene-related peptide (CGRP) release in rodents through activation of TRPA1.92 Furthermore, dural application of mustard oil and umbellulone produced robust time-related tactile facial and hind paw allodynia and reduce exploratory activity in a TRPA1-dependent manner.91 Preclinical studies also found that intranasal delivery of the TRPA1 agonists mustard oil or acrolein results in meningeal vasodilatation via a TRPA1-dependent mechanism.78 It is not clear whether this is due to direct access to dural afferents through nasal delivery or as speculated, cross-excitation between nasal afferents and dural afferents through intraganglionic neurotransmission following nasal application of TRPA1 agoinsts.46 In accordance with this hypothesis, TRPA1+ neurons are found to be clustered around some dural afferents, indicating the possibilities of cross-excitation in the trigeminal ganglion.46 Activation/desensitization of TRPA1 was also shown to modulate the function of trigeminal ganglion neurons. Parthenolide is a major constituent of feverfew, which has long been used to treat headache.93 A recent study found that parthenolide behaves as a partial agonist for TRPA1.93 Following its initial activation, parthenolide causes desensitization of CGRP release from trigeminal neurons and renders TRPA1-expressing nerve terminals unresponsive to other stimuli.93 This study opens up new opportunities to develop migraine treatments targeting TRPA1, possibly via desensitization.

Many identified TRPA1 agonists, including formaldehyde, ammonium chloride, cigarette smoke, and umbellulone are chemical irritants and they have long been known to trigger migraine headache in susceptible individuals.7, 92, 94, 95 However, patients with gain-of-function mutations in TRPA1 have spontaneous limb pain with premonitory symptoms but no migraine,87 and inhalation of many TRPA1 agonists can cause severe headaches but not always migraine. The channel may be hyperactive with the mutation, but still in need of local activators (just lower levels of activators). Those activators may be present in sufficient concentrations in some people in the meninges, but in others they may be present in other tissues. Further investigation is required to evaluate how TRPA1 can contribute to headache in some conditions and migraine in others.

2.2 Transient Receptor Potential Cation Channel V1 (TRPV1)

Located on small and medium-sized neurons, either unmyelinated C fibers or thinly myelinated A fibers, in trigeminal and dorsal root ganglion neurons,96, 97 TRPV1 is activated directly by capsaicin and noxious temperatures (above 42 °C).98 Additionally, extracellular protons dramatically potentiate the response of TPRV1 to capsaicin and noxious temperature during tissue acidosis, such as that associated with arthritis, inflammation, and other forms of injury.99 Numerous studies focus on the role of TRPV1 in the generation and pathophysiology of migraine. A two-stage genetic association study showed that single nucleotide polymorphisms in the TRPV1 gene contribute to the genetic susceptibility to migraine in a Spanish population.100 An immunohistological study found that nerve fibers in the dura mater express TRPV1.101 In line with this finding, approximately 24% of identified dural afferents express TRPV1.46

Functional studies also elucidate the interplay between TRPV1 and the release of CGRP. CGRP, released at nerve terminals upon afferent activation, plays an important role in the generation of migraine headache. Infusion of CGRP can trigger a migraine attack and CGRP is also a potent vasodilator.102 Bolus injections of capsaicin induce dural vessel dilation through TRPV1-mediated release of CGRP.103 Ethanol, a well-known inducer of headache, is able to stimulate TRPV1 on primary afferent neurons, promoting neurogenic inflammation and CGRP-mediated coronary dilation.104 However, it is not clear how TRPV1 is activated within the dura. A growing list of endogenous or exogenous mediators of inflammation105 have been shown to either activate TRPV1 directly such as anandamide (an endogenous cannabinoid neurotransmitter),106 N-arachidonoyl-dopamine (NADA) (endovanilloids),107 lipoxygenase products,108 polyamines,109 and venoms from jellyfish110 and spiders111, or lower its activation threshold, including ethanol,112 protons,99 bradykinin (an endogenous inflammatory peptide),113 nerve-growth factor,113 prostaglandins,114 ATP,115 and prokineticins.116 Therefore, numerous mechanisms exist by which dural TRPV1 may be activated and sensitized following meningeal inflammation but which of these occurs before, or during migraine is not clear.

Multiple TRPV1-targeted therapies have been developed to potentially treat migraine. An intranasal TRPV1 agonist, Civamide has been used to deplete neuropeptides such as CGRP within the trigeminal system, thereby preventing migraine headache.117 Clinical studies demonstrate that Civamide Nasal Solution can also substantially reduce the frequency of another headache disorder, cluster headache, following a single one-week course of treatment.117 Another TRPV1 agonist under study is capsaicin. Capsaicin is known to desensitize TRPV1-expressing nerve endings not only to capsaicin but to other stimuli as well.118 Repeated intranasal capsaicin applications produce significant amelioration of migraine attacks.119 However, the initial activation of TRPV1 and subsequent peripheral neuropeptide release related with this group of drugs cause rather uncomfortable side effects.117 The mechanism by which TRPV1 agonists treat headache attacks is not clear and surprisingly headache is not reported as a side effect after initial administration of these compounds.117 The side effects of nasal TRPV1 agonists, including nasal burning, lacrimation and rhinorrhoea, indicating activation of nasal afferents.117 It may be speculated that the thresholds for activation, desensitization, and neuropeptide depletion between nasal afferents and dural afferents are different or the local concentration of TRPV1 agonists at the nerve endings are not equal and thus response properties and side effects are not identical in the territories supplied by these neuronal populations.

Despite a seemingly suggestive role for TRPV1 in migraine based on the studies described above, the ultimate efficacy of a TRPV1-based therapy for migraine is questionable. TRPV1 antagonists have been developed for various types of pain but side effects have so far prevented their progress beyond early stage clinical trials. AMG-517, a TRPV1 antagonist, was discontinued in clinical trials due to the incidence of long-lasting hyperthermia in susceptible humans with a maximal body temperature surpassing 40 °C,120 and similar reports of increased body temperature were reported with other TRPV1 antagonists. Further, the effects of TRPV1 antagonists are inconsistent in various studies.103, 120123 Capsazepine was able to inhibit the capsaicin-induced vasodilation103 and administration of SB-705498, a specific TRPV1 antagonist, was able to suppress the responses of second order neurons in nucleus caudalis due to the electric stimulation of the dura in cats.122 However in another study, A993610, a specific TRPV1 antagonist, had no effect on capsaicin-induced vasodilation and it did not affect the firing of second order neurons following electric stimulation of the middle meningeal artery in rats.121 Therefore, the clinical relevance of TRPV1 antagonists in the treatment of migraine remains to be established.

2.3 Transient Receptor Potential Cation Channel V4 (TRPV4)

Prior work has found that dural afferent nociceptors are mechanically sensitive.20, 22, 124, 125 Trigeminal afferent nociceptors can also be activated by dural application of solutions with either increased or decreased osmolarity.20 The most likely channel mediating these effects is the mechano-sensitive TRPV4. TRPV4 is activated by hypotonic challenge as well as by mechanical stimulation126129 and mice lacking TRPV4 exhibit reduced pressure and osmotic sensitivity.130, 131 Thus, TRPV4 represents a possible candidate for directly mediating the mechanosensitivity of dural afferent nociceptors. Expression of mRNA for TRPV4 has been found in the trigeminal ganglion132 and both hypotonic solutions and the TRPV4 activator 4α-PDD were shown to excite 56% and 49% of identified dural afferents respectively.133 In awake animals, dural application of hypotonic solutions or 4α-PDD resulted in migraine-related pain behavior through the activation of TRPV4.133 Similar to TRPV1, the endogenous mechanism for activation of TRPV4 is not clear. Currently there is no evidence for a decrease in plasma osmolarity before or during a migraine attack. However, there could be mechanical stimulation of the meninges following sudden intracranial pressure changes due to head jolts or rotation, breath-holding, sneezing or coughing etc. It has been reported that mechano-stimulation such as rapid head movements and coughing can worsen headache pain in migraine patients,134 suggesting increased mechanosensitivity of meningeal afferents. One possible basis for such symptoms could be sensitization of TRPV4. TRPV4 can be sensitized by pro-inflammatory mediators, including those released following mast cell degranulation e.g. PGE2, PAR2 etc.135138 Following sensitization, the threshold for TRPV4 activation in response to mechano-stimuli may decrease for dural afferents, contributing to intracranial mechanical hypersensitivity and this may also contribute to the throbbing pain of migraine. Therefore, TRPV4 is a promising candidate for transducing mechano-stimuli within the dura but this possibility merits further investigation.

2.4 Transient receptor potential melastatin 8 (TRPM8)

TRPM8 channels are activated by chemical cooling agents, including menthol and icilin and temperatures below 26 °C and therefore it has been proposed to sense cold stimuli. In support of this hypothesis, several studies have confirmed that TRPM8 knockout mice display severe behavioral deficits in response to cold stimuli, including diminished responses to acetone application to the hindpaw,139141 an impaired ability to discriminate warm and cold surfaces,139, 141 and a reduction in icilin-induced jumping.140 Cold allodynia is a symptom reported by patients during migraine attacks,142 which could suggest the involvement of TRPM8, but activation of TRPM8 on dural afferents is unlikely to mediate this effect as channel activity would seemingly be required at the location of allodynia. The strongest evidence thus far for a role of TRPM8 in migraine comes from three separate genome-wide association analyses performed on 3 different groups of migraine patients, all of which identified that a TRPM8 gene variant (2q37.1, rs10166942) is associated with increased susceptibility to common migraine.143145 Studies have yet to identify how this gene variant alters channel function or expression if at all. And although intriguing, it is important to confirm that the identified gene variant is causally related to migraine and to explore how this channel contributes to pathophysiology. This is particularly important in light of potentially limited expression of TRPM8 in dural afferents as one study found a lack of TRPM8 expression in dural afferents,46 while another study found sparse innervation of the dura with not all regions of the meninges containing TRPM8-positive afferents.146

3. P2X channels

P2X receptors are ligand-gated ion channels that activate in response to binding of extracellular ATP.147 Seven P2X subtypes have been identified and functional channels are assembled as homomeric and heteromultimeric channels.147 Among the 7 subtypes, P2X3 has been widely implicated in various pain conditions148 since it is highly expressed by peripheral sensory afferents.149151 Approximately 50% of dural afferents express P2X2 or P2X3 or both receptors152 and 40% of the trigeminal sensory neurons expressing P2X3 receptors also express P2X2 receptors and vice versa, indicating the existence of P2X2/3 heteromultimeric channels.152 P2X3-expressing trigeminal sensory neurons exhibit some unique neurochemical properties compared with those in dorsal root ganglion neurons. In contrast to almost complete co-localization between P2X3 and IB4 in dorsal root ganglia (98–99%),153, 154 only 64% of P2X3-expressing neurons are IB4 positive in trigeminal ganglion.152, 155 This suggests that P2X3 receptors are also expressed on peptidergic afferents in trigeminal ganglion and may contribute to increased CGRP release.

Increased expression of P2X3 channels may also be linked to migraine as P2X3 receptors can be up-regulated by several migraine-related mediators. The neurotrophin nerve growth factor (NGF) is an important mediator of inflammatory pain conditions. A clinical study found higher levels of NGF in the cerebrospinal fluid of patients with chronic headache, supporting its involvement in chronic head pain.24 NGF is able to enhance P2X3 currents in trigeminal neurons156 and neutralization of endogenous NGF decreases P2X3 receptor-mediated currents and delays their recovery from desensitization via a PKC-dependent pathway.157 CGRP has been demonstrated to up-regulate the membrane expression and activity of P2X3 receptors158 through enhancing gene transcription.159 And acidosis may further enhance ATP response through interacting with P2X2/3 heteromultimeric channels as P2X2 is strongly pH-sensitive.160, 161 P2X2/3 activity may be enhanced during acidosis conditions, such as those following mast cell degranulation or meningeal inflammation. Interestingly, P2X3 receptors might also be involved in the pathophysiology of familial hemiplegic migraine-1 with the R192Q missense mutation in the α1 subunit of CaV2.1 channels.162 Neurons expressing this mutation possess more abundant lipid raft compartments containing a larger fraction of P2X3 receptors at the membrane level.162 This pathological change could lead to enhanced functional responses of P2X3 receptors and contribute to migraine pathophysiology by enhancing ATP-mediated responses.162 Together, these studies link P2X3 to a variety of conditions implicated in migraine and suggest that the channel may play a role in migraine pathophysiology.

The source of ATP required to activate meningeal afferents is not clear. It has been shown that during a CSD event, ATP is released into the extracellular space at levels exceeding 100 mM,163 but given the instability of ATP in the extracellular space, the likelihood of ATP migrating to the meninges before being degraded is low. There may also be ATP release in the meninges following sympathetic outflow as ATP is present with norepinephrine in vesicles released by sympathetic neurons. As yet, there are no conclusive studies demonstrating an increase or a source of ATP within the meninges so this remains a matter of speculation. Ultimately, efficacy of P2X antagonists in preclinical headache models or in humans with migraine, particularly if these compounds are peripherally restricted, would lend support for a purinergic mechanism in migraine. Although specific antagonists for P2X3 receptors have not yet been tested for migraine, a recent study found that dihydroergotamine (DHE), an ergot alkaloid derivative used extensively in the acute treatment of migraine, represses ATP-mediated sensitization of trigeminal neurons via down regulation of P2X3 receptors.164 It is difficult to speculate that treatment antagonizing P2X3 receptor could be useful for migraine prevention based on this finding given the non-selectivity of DHE, but some of the efficacy of DHE could be due to purinergic actions. As mentioned above, this hypothesis awaits testing with more selective compounds.

4. Calcium-activated Potassium Channel (BKCa or MaxiK)

In addition to ion channels expressed on dural afferent nerve endings which could directly initiate the activation of nociceptive signaling, other channels expressed in the trigeminal signaling pathway may also contribute to the hyperactivity of the trigeminocervical complex. One example is large conductance calcium-activated potassium channels, also known as BKCa or MaxiK. The α subunit of MaxiK is encoded by the slowpoke (Slo) gene.165 Splice variants of Slo and co-association of Slo with its β subunit could give rise to a diversity of MaxiK channels.165 The MaxiK channel is widely expressed in the brain, both in soma, dendrites and axonal terminals.165 The highest expression level is observed in axonal terminals, suggesting that the neuronal MaxiK protein is most abundant in synaptic terminals.165 Activation of neuronal cell body MaxiK channels induces K+ efflux, hyperpolarization of neurons, and therefore reduces neuronal excitability.165 In addition, MaxiK located at axonal terminals has been found to provide finely tuned modulation of neurotransmitter release through modulation of Ca2+ entry.165 Located in close proximity to presynaptic voltage-gated calcium channels, MaxiK channels are thought to narrow the presynaptic action potential upon activation and therefore decrease presynaptic Ca2+ entry, resulting in reduced neurotransmitter release.166 Thus, MaxiK might play a key role in modulation of pain transmission from the peripheral to the higher centers of central nervous system.

With respect to migraine, activation of MaxiK is hypothesized to work on both synaptic terminals and cell bodies to decrease neuronal firing by preventing transmitter release and subsequent activation of dorsal horn neurons.167 Intravenous application of NS1619, an activator and opener of neuronal MaxiK channels, dose-dependently inhibited neurogenic dural vasodilation in a model of trigeminovascular nociception.167 This suggests that activation of MaxiK channels is able to reduce neuronal firing and potentially the release of neurotransmitters from trigeminal neurons that innervate the dura.167 In the CNS, MaxiK is expressed throughout the trigeminal nucleus caudalis.167 Intravenous application of NS1619 had no effect on the firing of caudalis neurons,167 whereas application of NS1619 directly onto the caudalis was able to inhibit neuronal firing induced by current injection or L-glutamate application,167 indicating that NS1619 is not able to cross the blood-brain barrier. Thus, although MaxiK is expressed within the CNS, effects of systemic administration of NS1619 on dural vasodilation is likely occurring via peripheral action. The expression of the MaxiK channel is also found in sensory neurons, mainly in DRG neurons with small and medium diameters.168 Mice with sensory neuron-specific knockout of MaxiK channels exhibited increased inflammatory pain, while acute nociceptive responses and neuropathic pain conditions were unaffected.168 These date suggest that MaxiK channels expressed on sensory neurons could play an important role in modulation of inflammatory pain conditions. Further investigation into the effects of MaxiK openers on trigeminal neurons, trigeminal nucleus caudalis neurons, and neurotransmitter release will help further clarify the role of MaxiK channels in migraine pathophysiology. The studies described above nonetheless suggest that manipulation of MaxiK channels may represent a novel therapeutic target for migraine treatment by affecting trigeminovascular neuronal firing in response to dural stimulation.

Sensitization of Dural Afferents

In vivo electrophysiological recording studies have found that dural afferents exhibit other properties of nociceptors, including sensitization.20, 22, 125, 169 Transient and recurrent release of inflammatory mediators within the dura and resultant sensitization of dural sensory nerve endings might be relevant to symptoms that are characteristic of migraine, such as the presence of exaggerated intracranial mechanosensitivity. Therefore, a key question is whether dural afferents are similar to nociceptors innervating other tissues or alternatively, whether they have unique sensitization properties that might be of significance for headache pathogenesis.

A combination of inflammatory mediators (IM), consisting of PGE2, bradykinin, and histamine, was used to compare the differences in sensitization between dural afferents and afferents innervating temporalis muscle (TM). Interestingly, 100% of dural afferents were sensitized by IM application, compared to only 28.5% of TM afferents.44 IM also depolarized the resting membrane potential in dural afferents, but not TM afferents.44 To understand the unique sensitization properties of dural afferents, the ionic mechanisms underlying the sensitization were investigated. First, IM was able to increase tetrodotoxin-resistant voltage-gated sodium currents in dural afferents, which most likely underlies the IM-induced decreases in action potential threshold.170 Most interestingly, IM application induced a depolarizing Cl current, which is most likely responsible for IM-induced membrane depolarization in dural afferents.170 This IM-sensitive Cl current has never been reported in other sensory afferents, indicating that it might be unique to dural afferents.44, 171 The ion channel mediating the IM-sensitive Cl current is yet to be determined but this channel could represent a significant new target for migraine therapeutics, particularly if there is a selective expression pattern within dural afferents.

In addition to above-mentioned inflammatory mediators, sensitizing effects of Interleukin-6 (IL-6) on dural afferents were also studied. IL-6 levels are elevated during migraine attacks172, 173 and IL-6 levels are strongly correlated with stress, a reliable migraine trigger. Stress is capable of evoking IL-6 release in a mast cell dependent manner.174 Thus, accumulating evidence points to IL-6 as a contributing factor in migraine. Following acute IL-6 application, trigeminal ganglion neurons were found to display phosphorylation of extracellular signal-regulated protein kinase (ERK), a neuronal activation marker,175 indicating that these neurons respond to IL-6 through the Mitogen-Activated Protein Kinase (MAPK) signaling pathway. In awake animals, direct application of IL-6 to the dura produced dose-dependent facial and hindpaw allodynia via activation of the ERK signaling pathway.176 In dural afferents, IL-6 application decreased the current threshold for action potential firing, indicating an increased excitability,176 and phosphorylation of the voltage-gated sodium channel Nav1.7 is most likely to contribute to these effects.176 These data not only provide a cellular mechanism by which inflammatory mediators in the meninges cause sensitization of dural afferents, but also suggest that the ionic channels, such as the IM-sensitive Cl current or the voltage-gated sodium channel Nav1.7 might represent novel targets for developing antimigraine therapies.

Apart from the difference of sensitization in subpopulations of dural afferents, there is also a gender difference in IM-induced sensitization of dural afferents. The American Migraine study reports that migraine incidence is approximately three times higher in women than in men.177, 178 Reasons for gender difference in migraine prevalence are not clear. Excitability and sensitization were compared between male and female dural afferents in rats and there was no difference between male and female dural afferents with respect to baseline excitability.179 However, a significantly higher proportion of dural afferent from female rats (100%) than from male rats (50%) were sensitized,179 which could contribute to higher susceptibility to migraine in women. The mechanisms for these differences are not yet clear but future studies should examine expression levels of IL-6 activated pathways, Nav1.7, and the IM-sensitive Cl current. Identification of sex differences in dural-afferent signaling will open up the possibility of developing different approaches for the treatment of migraine tailored specifically to men and women.

Inherited Migraine as a Channelopathy

Familial hemiplegic migraine is a rare inherited type of migraine with aura. Mutations in the genes CACNA1A at chromosome 19p13, encoding the α1 subunit of P/Q voltage-gated calcium channel Cav2. 1,180, 181 ATP1A2 at 1q23, encoding the α2 subunit of the Na+, K+ adenosinetriphosphatase (ATPase),182 and SCNA1A at 2q24, encoding α1 subunit of the votage-gated sodium channel Nav1.1183 have been linked to FHM1,2,3, respectively (reviewed 181, 184, 185). Recently, another susceptibility gene KCNK18, which encodes a two-pore domain (K2P) potassium channel, TWIK-related spinal cord potassium channel or TRESK, was linked to inherited migraine with aura.186

A negative resting membrane potential is critical for proper electrical signaling in excitable tissues. The resting membrane is mainly permeable to potassium ions due to the presence of background (leak) K+ currents.187 Only recently have several molecular entities responsible for leak K+ currents been identified.188 The family of cloned mammalian background potassium channels, named ‘two pore domain’ potassium channels (K2P), includes 15 related members.188 Based on their sequence similarities, the K2P channel family was further divided into six subfamilies, including TWIK, TREK, TASK, TALK, THIK and TRESK,188 all of which are characterized by a similar molecular architecture where each channel subunit contains two pore loop forming domain instead of one. K2P channel subunits dimerize to form functional channels. The electrophysiological features of K2P channels are also unique. Compared with other K+ channel families, the opening of K2P channels is less dependent on voltage, making them ideal candidates for generating background K+ currents and regulating the resting membrane potential.189

Among the K2P families, TRESK (TWIK-related spinal cord potassium channel) is the first member linked to a human disorder.190 Genetic screening studies have demonstrated a link between a frameshift mutation in the KCNK18 gene, denoted F139wfsX24, with migraine with aura.186 The mutation was discovered in a patient suffering from migraine with aura and 7 of the patient’s relatives who are also migraine sufferers but was absent from 8 relatives who do not have migraine.186 The F139wfsX24 mutation leads to a prematurely truncated and non-functional TRESK channel that can also suppress wild-type channel function through a dominant-negative fashion. However, it is unknown whether the aura or the migraine attacks in these patients are directly related to dysfunction of TRESK or due to other factors.

Several lines of preclinical evidence describe how TRESK may contribute to migraine pathophysiology. The expression of KCNK18 appears (in mice) in the trigeminal ganglion (TG) and dorsal root ganglion (DRG) at embryonic day 15.5 and reaches a peak in newborns.186, 190 In adult mice, KCNK18 expression is highest in trigeminal ganglion.186 Similarly, strong expression of TRESK is also found in human trigeminal ganglion neurons.186 Trigeminal ganglion neurons expressing mutant TRESK subunits exhibit a lower threshold for activation and a higher spike frequency in response to supra-threshold stimuli.191 Functional knockout of TRESK in mice results in a lowered threshold for activation, reduced action potential duration and slightly higher amplitudes of after-hyperpolarization in DRG neurons.192 Taken together, these studies suggest that TRESK channel is important for regulating excitability of primary afferent neurons. However, at this point, the site of action of this channel in migraine pathophysiology is not clear and this is a question that merits further investigation. The presence of aura in affected family members suggests that decreasing TRESK activity contributes to aura, possibly through lowering cortical spreading depression (CSD) threshold,186 which would suggest a CNS role for these channels in migraine. Regardless of the site of expression, these studies suggest that increasing TRESK activity will theoretically protect against migraine. Although not used for migraine, volatile anesthetics, such as isoflurane and halothane increase TRESK currents up to 3-fold at their clinical concentrations.193 Since the TRESK channel is distantly related to other K2P channels,188 it may be possible to develop novel molecules selectively targeting the TRESK channel that can be used as abortive or prophylactic drugs for migraine treatment.

Despite the interesting potential role for TRESK described above, a direct causative link between non-functional TRESK channels and migraine is questioned by a recent study showing that a loss of function mutation (C110R variant) of TRESK is found in both migraine and control cohorts.194 This study argues that non-functional TRESK channels alone are not sufficient to cause typical migraine. Several reasons might account for this discrepancy. Among K2P channel families, TREK, TALK, and TASK channels exhibit several splice variants188 and the expression pattern of the splice variants shows significant tissue specificity.195, 196 The splice variants of TRESK channels have never been examined. It is possible that the C110R variant and F139wfsX24 mutation are differentially expressed, with the F139wfsX24 mutation expressed in migraine-related tissues. Furthermore, migraine has now generally been recognized as a polygenic disorder and thus it is dependent upon gene-gene interaction as well as environmental triggers instead of a single channel mutation.197 Therefore, the TRESK channel is a novel and interesting new component in the migraine pathophysiology pathway but its role in neuronal excitability and migraine signaling requires more study.

Conclusions

Taken together, these studies indicate that ion channels can play an important role in migraine pathophysiology. Mutations of certain ion channels, such as TRESK and TRPM8, are linked with inherited migraine. Activation of primary afferent neurons is likely a critical step for the pain of migraine and multiple ion channels expressed on dural afferents can contribute to afferent input by sensing environmental changes in the meninges following CSD, ischemic, or inflammatory events etc. Future studies should more closely examine the role of ASICs in migraine by expanding the patient size of amiloride clinical trials for migraine and the development of more selective tools to better probe for a role of ASICs in migraine. It is also important to explore the endogenous mechanisms leading to the activation of TRP channels within the meninges. Knowledge of which and how specific TRP channels contribute to migraine triggered by specific factors will be critical in the development of effective therapy tailored to the individual migraine patient. Although genome-wide association studies have yielded plausible susceptibility genes in migraine, further work is required to determine whether the association is causal and functional studies are necessary to dissect the exact underlying molecular pathways of TRPM8 and TRESK channels.

Current therapies for the treatment of migraine have been restricted to drug classes such as NSAIDs, triptans, tricyclic antidepressants and anticonvulsants. However, such compounds provide limited benefit to many migraine patients. The studies described here provide several targets for therapeutics that are not manipulated by currently available drugs. A greater understanding of the role of ion channels in the mechanisms contributing to headache initiation can ultimately make it possible to develop novel ion-channel based migraine therapies with increased efficacy over current therapeutics.

Acknowledgments

Financial support: This work was supported by funds from The National Headache Foundation (GD) and the NIH NS072204 (GD).

Abbreviations

CSD

cortical spreading depression

FHM

familial hemiplegic migraine

K2P

a two-pore domain potassium channel

TRESK

TWIK-related spinal cord potassium channel

TG

trigeminal ganglion

ASICs

acid sensing ion channels

TRP

transient receptor potential channels

TRPA1

transient receptor potential cation channel A1

TRPV1

transient receptor potential cation channel V1

TRPV4

transient receptor potential cation channel V4

TRPM8

transient receptor potential melastatin 8

BKCa or MaxiK

calcium-activated potassium channel

TM

temporalis muscle

IM

inflammatory mediators

IL-6

Interleukin 6

Footnotes

Conflict of interests: No

References

  • 1.Goadsby PJ, Lipton RB, Ferrari MD. Migraine--current understanding and treatment. N Engl J Med. 2002;346:257–270. doi: 10.1056/NEJMra010917. [DOI] [PubMed] [Google Scholar]
  • 2.Lipton RB, Bigal ME, Steiner TJ, Silberstein SD, Olesen J. Classification of primary headaches. Neurology. 2004;63:427–435. doi: 10.1212/01.wnl.0000133301.66364.9b. [DOI] [PubMed] [Google Scholar]
  • 3.Stewart WF, Wood C, Reed ML, Roy J, Lipton RB. Cumulative lifetime migraine incidence in women and men. Cephalalgia. 2008;28:1170–1178. doi: 10.1111/j.1468-2982.2008.01666.x. [DOI] [PubMed] [Google Scholar]
  • 4.Lipton RB, Stewart WF, von Korff M. Burden of migraine: societal costs and therapeutic opportunities. Neurology. 1997;48:S4–9. doi: 10.1212/wnl.48.3_suppl_3.4s. [DOI] [PubMed] [Google Scholar]
  • 5.Levy D. Migraine pain and nociceptor activation--where do we stand? Headache. 2010;50:909–916. doi: 10.1111/j.1526-4610.2010.01670.x. [DOI] [PubMed] [Google Scholar]
  • 6.Pietrobon D, Striessnig J. Neurobiology of migraine. Nat Rev Neurosci. 2003;4:386–398. doi: 10.1038/nrn1102. [DOI] [PubMed] [Google Scholar]
  • 7.Kelman L. The triggers or precipitants of the acute migraine attack. Cephalalgia. 2007;27:394–402. doi: 10.1111/j.1468-2982.2007.01303.x. [DOI] [PubMed] [Google Scholar]
  • 8.Martin PR. Behavioral management of migraine headache triggers: learning to cope with triggers. Curr Pain Headache Rep. 2010;14:221–227. doi: 10.1007/s11916-010-0112-z. [DOI] [PubMed] [Google Scholar]
  • 9.Aukerman G, Knutson D, Miser WF. Management of the acute migraine headache. Am Fam Physician. 2002;66:2123–2130. [PubMed] [Google Scholar]
  • 10.Legg N, Khalil N. Visual dysfunction in migraine with aura. Cephalalgia. 1995;15:536–537. doi: 10.1046/j.1468-2982.1995.1506536.x. [DOI] [PubMed] [Google Scholar]
  • 11.Chronicle EP, Wilkins AJ, Coleston DM. Thresholds for detection of a target against a background grating suggest visual dysfunction in migraine with aura but not migraine without aura. Cephalalgia. 1995;15:117–122. doi: 10.1046/j.1468-2982.1995.015002117.x. [DOI] [PubMed] [Google Scholar]
  • 12.Mulleners WM, Chronicle EP, Palmer JE, Koehler PJ, Vredeveld JW. Visual cortex excitability in migraine with and without aura. Headache. 2001;41:565–572. doi: 10.1046/j.1526-4610.2001.041006565.x. [DOI] [PubMed] [Google Scholar]
  • 13.Cohen GL. Migraine prophylactic drugs work via ion channels. Med Hypotheses. 2005;65:114–122. doi: 10.1016/j.mehy.2005.01.027. [DOI] [PubMed] [Google Scholar]
  • 14.Goadsby PJ, Charbit AR, Andreou AP, Akerman S, Holland PR. Neurobiology of migraine. Neuroscience. 2009;161:327–341. doi: 10.1016/j.neuroscience.2009.03.019. [DOI] [PubMed] [Google Scholar]
  • 15.Messlinger K. Migraine: where and how does the pain originate? Exp Brain Res. 2009;196:179–193. doi: 10.1007/s00221-009-1756-y. [DOI] [PubMed] [Google Scholar]
  • 16.Hoskin KL, Zagami AS, Goadsby PJ. Stimulation of the middle meningeal artery leads to Fos expression in the trigeminocervical nucleus: a comparative study of monkey and cat. J Anat. 1999;194 (Pt 4):579–588. doi: 10.1046/j.1469-7580.1999.19440579.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Strassman AM, Mineta Y, Vos BP. Distribution of fos-like immunoreactivity in the medullary and upper cervical dorsal horn produced by stimulation of dural blood vessels in the rat. J Neurosci. 1994;14:3725–3735. doi: 10.1523/JNEUROSCI.14-06-03725.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Mayberg MR, Zervas NT, Moskowitz MA. Trigeminal projections to supratentorial pial and dural blood vessels in cats demonstrated by horseradish peroxidase histochemistry. J Comp Neurol. 1984;223:46–56. doi: 10.1002/cne.902230105. [DOI] [PubMed] [Google Scholar]
  • 19.Strassman AM, Levy D. Response properties of dural nociceptors in relation to headache. J Neurophysiol. 2006;95:1298–1306. doi: 10.1152/jn.01293.2005. [DOI] [PubMed] [Google Scholar]
  • 20.Strassman AM, Raymond SA, Burstein R. Sensitization of meningeal sensory neurons and the origin of headaches. Nature. 1996;384:560–564. doi: 10.1038/384560a0. [DOI] [PubMed] [Google Scholar]
  • 21.Bove GM, Moskowitz MA. Primary afferent neurons innervating guinea pig dura. J Neurophysiol. 1997;77:299–308. doi: 10.1152/jn.1997.77.1.299. [DOI] [PubMed] [Google Scholar]
  • 22.Ray BS, Wolff HG. Experimental studies on headache: Pain sensitive structures of the head and their significance in headache. Arch Surg. 1940;41:813–856. [Google Scholar]
  • 23.Perini F, D’Andrea G, Galloni E, et al. Plasma cytokine levels in migraineurs and controls. Headache. 2005;45:926–931. doi: 10.1111/j.1526-4610.2005.05135.x. [DOI] [PubMed] [Google Scholar]
  • 24.Sarchielli P, Alberti A, Floridi A, Gallai V. Levels of nerve growth factor in cerebrospinal fluid of chronic daily headache patients. Neurology. 2001;57:132–134. doi: 10.1212/wnl.57.1.132. [DOI] [PubMed] [Google Scholar]
  • 25.Dimlich RV, Keller JT, Strauss TA, Fritts MJ. Linear arrays of homogeneous mast cells in the dura mater of the rat. J Neurocytol. 1991;20:485–503. doi: 10.1007/BF01252276. [DOI] [PubMed] [Google Scholar]
  • 26.Strassman AM, Weissner W, Williams M, Ali S, Levy D. Axon diameters and intradural trajectories of the dural innervation in the rat. J Comp Neurol. 2004;473:364–376. doi: 10.1002/cne.20106. [DOI] [PubMed] [Google Scholar]
  • 27.Rozniecki JJ, Dimitriadou V, Lambracht-Hall M, Pang X, Theoharides TC. Morphological and functional demonstration of rat dura mater mast cell-neuron interactions in vitro and in vivo. Brain Res. 1999;849:1–15. doi: 10.1016/s0006-8993(99)01855-7. [DOI] [PubMed] [Google Scholar]
  • 28.Theoharides TC, Spanos C, Pang X, et al. Stress-induced intracranial mast cell degranulation: a corticotropin-releasing hormone-mediated effect. Endocrinology. 1995;136:5745–5750. doi: 10.1210/endo.136.12.7588332. [DOI] [PubMed] [Google Scholar]
  • 29.Ottosson A, Edvinsson L. Release of histamine from dural mast cells by substance P and calcitonin gene-related peptide. Cephalalgia. 1997;17:166–174. doi: 10.1046/j.1468-2982.1997.1703166.x. [DOI] [PubMed] [Google Scholar]
  • 30.Reuter U, Bolay H, Jansen-Olesen I, et al. Delayed inflammation in rat meninges: implications for migraine pathophysiology. Brain. 2001;124:2490–2502. doi: 10.1093/brain/124.12.2490. [DOI] [PubMed] [Google Scholar]
  • 31.Levy D. Migraine pain, meningeal inflammation, and mast cells. Curr Pain Headache Rep. 2009;13:237–240. doi: 10.1007/s11916-009-0040-y. [DOI] [PubMed] [Google Scholar]
  • 32.Levy D, Burstein R, Kainz V, Jakubowski M, Strassman AM. Mast cell degranulation activates a pain pathway underlying migraine headache. Pain. 2007;130:166–176. doi: 10.1016/j.pain.2007.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Zhang XC, Strassman AM, Burstein R, Levy D. Sensitization and activation of intracranial meningeal nociceptors by mast cell mediators. J Pharmacol Exp Ther. 2007;322:806–812. doi: 10.1124/jpet.107.123745. [DOI] [PubMed] [Google Scholar]
  • 34.Yan J, Wei X, Bischoff C, Edelmayer RM, Dussor G. pH-Evoked Dural Afferent Signaling Is Mediated by ASIC3 and Is Sensitized by Mast Cell Mediators. Headache. 2013;53:1250–1261. doi: 10.1111/head.12152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Smith JH, Butterfield JH, Pardanani A, Deluca GC, Cutrer FM. Neurologic symptoms and diagnosis in adults with mast cell disease. Clin Neurol Neurosurg. 113:570–574. doi: 10.1016/j.clineuro.2011.05.002. [DOI] [PubMed] [Google Scholar]
  • 36.Boes T, Levy D. Influence of sex, estrous cycle, and estrogen on intracranial dural mast cells. Cephalalgia. 2012;32:924–931. doi: 10.1177/0333102412454947. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Mayevsky A, Zeuthen T, Chance B. Measurements of extracellular potassium, ECoG and pyridine nucleotide levels during cortical spreading depression in rats. Brain Res. 1974;76:347–349. doi: 10.1016/0006-8993(74)90467-3. [DOI] [PubMed] [Google Scholar]
  • 38.Csiba L, Paschen W, Mies G. Regional changes in tissue pH and glucose content during cortical spreading depression in rat brain. Brain Res. 1985;336:167–170. doi: 10.1016/0006-8993(85)90430-5. [DOI] [PubMed] [Google Scholar]
  • 39.Van Harreveld A. Compounds in brain extracts causing spreading depression of cerebral cortical activity and contraction of crustacean muscle. J Neurochem. 1959;3:300–315. doi: 10.1111/j.1471-4159.1959.tb12636.x. [DOI] [PubMed] [Google Scholar]
  • 40.Jander S, Schroeter M, Peters O, Witte OW, Stoll G. Cortical spreading depression induces proinflammatory cytokine gene expression in the rat brain. J Cereb Blood Flow Metab. 2001;21:218–225. doi: 10.1097/00004647-200103000-00005. [DOI] [PubMed] [Google Scholar]
  • 41.Bolay H, Reuter U, Dunn AK, Huang Z, Boas DA, Moskowitz MA. Intrinsic brain activity triggers trigeminal meningeal afferents in a migraine model. Nat Med. 2002;8:136–142. doi: 10.1038/nm0202-136. [DOI] [PubMed] [Google Scholar]
  • 42.Zhang X, Levy D, Noseda R, Kainz V, Jakubowski M, Burstein R. Activation of meningeal nociceptors by cortical spreading depression: implications for migraine with aura. J Neurosci. 2010;30:8807–8814. doi: 10.1523/JNEUROSCI.0511-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Karatas H, Erdener SE, Gursoy-Ozdemir Y, et al. Spreading depression triggers headache by activating neuronal Panx1 channels. Science. 2013;339:1092–1095. doi: 10.1126/science.1231897. [DOI] [PubMed] [Google Scholar]
  • 44.Harriott AM, Gold MS. Electrophysiological properties of dural afferents in the absence and presence of inflammatory mediators. J Neurophysiol. 2009;101:3126–3134. doi: 10.1152/jn.91339.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Schepelmann K, Ebersberger A, Pawlak M, Oppmann M, Messlinger K. Response properties of trigeminal brain stem neurons with input from dura mater encephali in the rat. Neuroscience. 1999;90:543–554. doi: 10.1016/s0306-4522(98)00423-0. [DOI] [PubMed] [Google Scholar]
  • 46.Huang D, Li S, Dhaka A, Story GM, Cao YQ. Expression of the transient receptor potential channels TRPV1, TRPA1 and TRPM8 in mouse trigeminal primary afferent neurons innervating the dura. Mol Pain. 2012;8:66. doi: 10.1186/1744-8069-8-66. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Yan J, Edelmayer RM, Wei X, De Felice M, Porreca F, Dussor G. Dural afferents express acid-sensing ion channels: a role for decreased meningeal pH in migraine headache. Pain. 2011;152:106–113. doi: 10.1016/j.pain.2010.09.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Wemmie JA, Price MP, Welsh MJ. Acid-sensing ion channels: advances, questions and therapeutic opportunities. Trends Neurosci. 2006;29:578–586. doi: 10.1016/j.tins.2006.06.014. [DOI] [PubMed] [Google Scholar]
  • 49.Lingueglia E. Acid-sensing ion channels in sensory perception. J Biol Chem. 2007;282:17325–17329. doi: 10.1074/jbc.R700011200. [DOI] [PubMed] [Google Scholar]
  • 50.Baron A, Voilley N, Lazdunski M, Lingueglia E. Acid sensing ion channels in dorsal spinal cord neurons. J Neurosci. 2008;28:1498–1508. doi: 10.1523/JNEUROSCI.4975-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Price MP, Snyder PM, Welsh MJ. Cloning and expression of a novel human brain Na+ channel. J Biol Chem. 1996;271:7879–7882. doi: 10.1074/jbc.271.14.7879. [DOI] [PubMed] [Google Scholar]
  • 52.Waldmann R, Champigny G, Voilley N, Lauritzen I, Lazdunski M. The mammalian degenerin MDEG, an amiloride-sensitive cation channel activated by mutations causing neurodegeneration in Caenorhabditis elegans. J Biol Chem. 1996;271:10433–10436. doi: 10.1074/jbc.271.18.10433. [DOI] [PubMed] [Google Scholar]
  • 53.Wu L-J, Duan B, Mei Y-D, et al. Characterization of Acid-sensing Ion Channels in Dorsal Horn Neurons of Rat Spinal Cord. J Biol Chem. 2004;279:43716–43724. doi: 10.1074/jbc.M403557200. [DOI] [PubMed] [Google Scholar]
  • 54.Mamet J, Baron A, Lazdunski M, Voilley N. ProInflammatory Mediators, Stimulators of Sensory Neuron Excitability via the Expression of Acid-Sensing Ion Channels. J Neurosci. 2002;22:10662–10670. doi: 10.1523/JNEUROSCI.22-24-10662.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Voilley N, de Weille J, Mamet J, Lazdunski M. Nonsteroid anti-inflammatory drugs inhibit both the activity and the inflammation-induced expression of acid-sensing ion channels in nociceptors. J Neurosci. 2001;21:8026–8033. doi: 10.1523/JNEUROSCI.21-20-08026.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Yagi J, Wenk HN, Naves LA, McCleskey EW. Sustained currents through ASIC3 ion channels at the modest pH changes that occur during myocardial ischemia. Circ Res. 2006;99:501–509. doi: 10.1161/01.RES.0000238388.79295.4c. [DOI] [PubMed] [Google Scholar]
  • 57.Waldmann R, Bassilana F, de Weille J, Champigny G, Heurteaux C, Lazdunski M. Molecular cloning of a non-inactivating proton-gated Na+ channel specific for sensory neurons. J Biol Chem. 1997;272:20975–20978. doi: 10.1074/jbc.272.34.20975. [DOI] [PubMed] [Google Scholar]
  • 58.Jones RC, 3rd, Xu L, Gebhart GF. The mechanosensitivity of mouse colon afferent fibers and their sensitization by inflammatory mediators require transient receptor potential vanilloid 1 and acid-sensing ion channel 3. J Neurosci. 2005;25:10981–10989. doi: 10.1523/JNEUROSCI.0703-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Molliver DC, Immke DC, Fierro L, Pare M, Rice FL, McCleskey EW. ASIC3, an acid-sensing ion channel, is expressed in metaboreceptive sensory neurons. Mol Pain. 2005;1:35. doi: 10.1186/1744-8069-1-35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Page AJ, Brierley SM, Martin CM, et al. Different contributions of ASIC channels 1a, 2, and 3 in gastrointestinal mechanosensory function. Gut. 2005;54:1408–1415. doi: 10.1136/gut.2005.071084. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Sutherland SP, Benson CJ, Adelman JP, McCleskey EW. Acid-sensing ion channel 3 matches the acid-gated current in cardiac ischemia-sensing neurons. Proc Natl Acad Sci U S A. 2001;98:711–716. doi: 10.1073/pnas.011404498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Deval E, Noel J, Gasull X, et al. Acid-sensing ion channels in postoperative pain. J Neurosci. 31:6059–6066. doi: 10.1523/JNEUROSCI.5266-10.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Sluka KA, Radhakrishnan R, Benson CJ, et al. ASIC3 in muscle mediates mechanical, but not heat, hyperalgesia associated with muscle inflammation. Pain. 2007;129:102–112. doi: 10.1016/j.pain.2006.09.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Burstein R. Deconstructing migraine headache into peripheral and central sensitization. Pain. 2001;89:107–110. doi: 10.1016/s0304-3959(00)00478-4. [DOI] [PubMed] [Google Scholar]
  • 65.Durham PL, Masterson CG. Two mechanisms involved in trigeminal CGRP release: implications for migraine treatment. Headache. 2013;53:67–80. doi: 10.1111/j.1526-4610.2012.02262.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Holland PR, Akerman S, Andreou AP, Karsan N, Wemmie JA, Goadsby PJ. Acid-sensing ion channel 1: a novel therapeutic target for migraine with aura. Ann Neurol. 2012;72:559–563. doi: 10.1002/ana.23653. [DOI] [PubMed] [Google Scholar]
  • 67.De Young MB, Nemeth EF, Scarpa A. Measurement of the internal pH of mast cell granules using microvolumetric fluorescence and isotopic techniques. Arch Biochem Biophys. 1987;254:222–233. doi: 10.1016/0003-9861(87)90098-1. [DOI] [PubMed] [Google Scholar]
  • 68.Lambert GA, Michalicek J. Cortical spreading depression reduces dural blood flow--a possible mechanism for migraine pain? Cephalalgia. 1994;14:430–436. doi: 10.1046/j.1468-2982.1994.1406430.x. discussion 393–434. [DOI] [PubMed] [Google Scholar]
  • 69.Ramsey IS, Delling M, Clapham DE. An introduction to TRP channels. Annu Rev Physiol. 2006;68:619–647. doi: 10.1146/annurev.physiol.68.040204.100431. [DOI] [PubMed] [Google Scholar]
  • 70.Liu M, Liu MC, Magoulas C, Priestley JV, Willmott NJ. Versatile regulation of cytosolic Ca2+ by vanilloid receptor I in rat dorsal root ganglion neurons. J Biol Chem. 2003;278:5462–5472. doi: 10.1074/jbc.M209111200. [DOI] [PubMed] [Google Scholar]
  • 71.Karai LJ, Russell JT, Iadarola MJ, Olah Z. Vanilloid receptor 1 regulates multiple calcium compartments and contributes to Ca2+-induced Ca2+ release in sensory neurons. J Biol Chem. 2004;279:16377–16387. doi: 10.1074/jbc.M310891200. [DOI] [PubMed] [Google Scholar]
  • 72.Vriens J, Appendino G, Nilius B. Pharmacology of vanilloid transient receptor potential cation channels. Mol Pharmacol. 2009;75:1262–1279. doi: 10.1124/mol.109.055624. [DOI] [PubMed] [Google Scholar]
  • 73.Numazaki M, Tominaga M. Nociception and TRP Channels. Curr Drug Targets CNS Neurol Disord. 2004;3:479–485. doi: 10.2174/1568007043336789. [DOI] [PubMed] [Google Scholar]
  • 74.Jordt SE, Bautista DM, Chuang HH, et al. Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1. Nature. 2004;427:260–265. doi: 10.1038/nature02282. [DOI] [PubMed] [Google Scholar]
  • 75.Fajardo O, Meseguer V, Belmonte C, Viana F. TRPA1 channels mediate cold temperature sensing in mammalian vagal sensory neurons: pharmacological and genetic evidence. J Neurosci. 2008;28:7863–7875. doi: 10.1523/JNEUROSCI.1696-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Nassini R, Materazzi S, De Siena G, De Cesaris F, Geppetti P. Transient receptor potential channels as novel drug targets in respiratory diseases. Curr Opin Investig Drugs. 2010;11:535–542. [PubMed] [Google Scholar]
  • 77.McNamara CR, Mandel-Brehm J, Bautista DM, et al. TRPA1 mediates formalin-induced pain. Proc Natl Acad Sci U S A. 2007;104:13525–13530. doi: 10.1073/pnas.0705924104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Bautista DM, Jordt SE, Nikai T, et al. TRPA1 mediates the inflammatory actions of environmental irritants and proalgesic agents. Cell. 2006;124:1269–1282. doi: 10.1016/j.cell.2006.02.023. [DOI] [PubMed] [Google Scholar]
  • 79.Bessac BF, Jordt SE. Breathtaking TRP channels: TRPA1 and TRPV1 in airway chemosensation and reflex control. Physiology (Bethesda) 2008;23:360–370. doi: 10.1152/physiol.00026.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Andre E, Campi B, Materazzi S, et al. Cigarette smoke-induced neurogenic inflammation is mediated by alpha, beta-unsaturated aldehydes and the TRPA1 receptor in rodents. J Clin Invest. 2008;118:2574–2582. doi: 10.1172/JCI34886. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Bandell M, Story GM, Hwang SW, et al. Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin. Neuron. 2004;41:849–857. doi: 10.1016/s0896-6273(04)00150-3. [DOI] [PubMed] [Google Scholar]
  • 82.Bautista DM, Movahed P, Hinman A, et al. Pungent products from garlic activate the sensory ion channel TRPA1. Proc Natl Acad Sci U S A. 2005;102:12248–12252. doi: 10.1073/pnas.0505356102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Taylor-Clark TE, Ghatta S, Bettner W, Undem BJ. Nitrooleic acid, an endogenous product of nitrative stress, activates nociceptive sensory nerves via the direct activation of TRPA1. Mol Pharmacol. 2009;75:820–829. doi: 10.1124/mol.108.054445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Trevisani M, Siemens J, Materazzi S, et al. 4-Hydroxynonenal, an endogenous aldehyde, causes pain and neurogenic inflammation through activation of the irritant receptor TRPA1. Proc Natl Acad Sci U S A. 2007;104:13519–13524. doi: 10.1073/pnas.0705923104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Materazzi S, Nassini R, Andre E, et al. Cox-dependent fatty acid metabolites cause pain through activation of the irritant receptor TRPA1. Proc Natl Acad Sci U S A. 2008;105:12045–12050. doi: 10.1073/pnas.0802354105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Basbaum AI, Bautista DM, Scherrer G, Julius D. Cellular and molecular mechanisms of pain. Cell. 2009;139:267–284. doi: 10.1016/j.cell.2009.09.028. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Kremeyer B, Lopera F, Cox JJ, et al. A gain-of-function mutation in TRPA1 causes familial episodic pain syndrome. Neuron. 2010;66:671–680. doi: 10.1016/j.neuron.2010.04.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Corey DP, Garcia-Anoveros J, Holt JR, et al. TRPA1 is a candidate for the mechanosensitive transduction channel of vertebrate hair cells. Nature. 2004;432:723–730. doi: 10.1038/nature03066. [DOI] [PubMed] [Google Scholar]
  • 89.Kwan KY, Glazer JM, Corey DP, Rice FL, Stucky CL. TRPA1 modulates mechanotransduction in cutaneous sensory neurons. J Neurosci. 2009;29:4808–4819. doi: 10.1523/JNEUROSCI.5380-08.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Kerstein PC, del Camino D, Moran MM, Stucky CL. Pharmacological blockade of TRPA1 inhibits mechanical firing in nociceptors. Mol Pain. 2009;5:19. doi: 10.1186/1744-8069-5-19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Edelmayer RM, Le LN, Yan J, et al. Activation of TRPA1 on dural afferents: a potential mechanism of headache pain. Pain. 2012;153:1949–1958. doi: 10.1016/j.pain.2012.06.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Nassini R, Materazzi S, Vriens J, et al. The ‘headache tree’ via umbellulone and TRPA1 activates the trigeminovascular system. Brain. 2012;135:376–390. doi: 10.1093/brain/awr272. [DOI] [PubMed] [Google Scholar]
  • 93.Materazzi S, Benemei S, Fusi C, et al. Parthenolide inhibits nociception and neurogenic vasodilatation in the trigeminovascular system by targeting the TRPA1 channel. Pain. 2013;154:2750–2758. doi: 10.1016/j.pain.2013.08.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Wantke F, Focke M, Hemmer W, et al. Exposure to formaldehyde and phenol during an anatomy dissecting course: sensitizing potency of formaldehyde in medical students. Allergy. 2000;55:84–87. doi: 10.1034/j.1398-9995.2000.00307.x. [DOI] [PubMed] [Google Scholar]
  • 95.Irlbacher K, Meyer BU. Nasally triggered headache. Neurology. 2002;58:294. doi: 10.1212/wnl.58.2.294. [DOI] [PubMed] [Google Scholar]
  • 96.Guo A, Vulchanova L, Wang J, Li X, Elde R. Immunocytochemical localization of the vanilloid receptor 1 (VR1): relationship to neuropeptides, the P2X3 purinoceptor and IB4 binding sites. Eur J Neurosci. 1999;11:946–958. doi: 10.1046/j.1460-9568.1999.00503.x. [DOI] [PubMed] [Google Scholar]
  • 97.Ichikawa H, Sugimoto T. VR1-immunoreactive primary sensory neurons in the rat trigeminal ganglion. Brain Res. 2001;890:184–188. doi: 10.1016/s0006-8993(00)03253-4. [DOI] [PubMed] [Google Scholar]
  • 98.Caterina MJ, Julius D. The vanilloid receptor: a molecular gateway to the pain pathway. Annu Rev Neurosci. 2001;24:487–517. doi: 10.1146/annurev.neuro.24.1.487. [DOI] [PubMed] [Google Scholar]
  • 99.Jordt SE, Tominaga M, Julius D. Acid potentiation of the capsaicin receptor determined by a key extracellular site. Proc Natl Acad Sci U S A. 2000;97:8134–8139. doi: 10.1073/pnas.100129497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Carreno O, Corominas R, Fernandez-Morales J, et al. SNP variants within the vanilloid TRPV1 and TRPV3 receptor genes are associated with migraine in the Spanish population. Am J Med Genet B Neuropsychiatr Genet. 2012;159B:94–103. doi: 10.1002/ajmg.b.32007. [DOI] [PubMed] [Google Scholar]
  • 101.Shimizu T, Toriumi H, Sato H, et al. Distribution and origin of TRPV1 receptor-containing nerve fibers in the dura mater of rat. Brain Res. 2007;1173:84–91. doi: 10.1016/j.brainres.2007.07.068. [DOI] [PubMed] [Google Scholar]
  • 102.Doods H, Arndt K, Rudolf K, Just S. CGRP antagonists: unravelling the role of CGRP in migraine. Trends Pharmacol Sci. 2007;28:580–587. doi: 10.1016/j.tips.2007.10.005. [DOI] [PubMed] [Google Scholar]
  • 103.Akerman S, Kaube H, Goadsby PJ. Vanilloid type 1 receptors (VR1) on trigeminal sensory nerve fibres play a minor role in neurogenic dural vasodilatation, and are involved in capsaicin-induced dural dilation. Br J Pharmacol. 2003;140:718–724. doi: 10.1038/sj.bjp.0705486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Nicoletti P, Trevisani M, Manconi M, et al. Ethanol causes neurogenic vasodilation by TRPV1 activation and CGRP release in the trigeminovascular system of the guinea pig. Cephalalgia. 2008;28:9–17. doi: 10.1111/j.1468-2982.2007.01448.x. [DOI] [PubMed] [Google Scholar]
  • 105.Szallasi A, Cortright DN, Blum CA, Eid SR. The vanilloid receptor TRPV1: 10 years from channel cloning to antagonist proof-of-concept. Nat Rev Drug Discov. 2007;6:357–372. doi: 10.1038/nrd2280. [DOI] [PubMed] [Google Scholar]
  • 106.Zygmunt PM, Petersson J, Andersson DA, et al. Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide. Nature. 1999;400:452–457. doi: 10.1038/22761. [DOI] [PubMed] [Google Scholar]
  • 107.Huang SM, Bisogno T, Trevisani M, et al. An endogenous capsaicin-like substance with high potency at recombinant and native vanilloid VR1 receptors. Proc Natl Acad Sci U S A. 2002;99:8400–8405. doi: 10.1073/pnas.122196999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Hwang SW, Cho H, Kwak J, et al. Direct activation of capsaicin receptors by products of lipoxygenases: endogenous capsaicin-like substances. Proc Natl Acad Sci U S A. 2000;97:6155–6160. doi: 10.1073/pnas.97.11.6155. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Ahern GP, Wang X, Miyares RL. Polyamines are potent ligands for the capsaicin receptor TRPV1. J Biol Chem. 2006;281:8991–8995. doi: 10.1074/jbc.M513429200. [DOI] [PubMed] [Google Scholar]
  • 110.Cuypers E, Yanagihara A, Karlsson E, Tytgat J. Jellyfish and other cnidarian envenomations cause pain by affecting TRPV1 channels. FEBS Lett. 2006;580:5728–5732. doi: 10.1016/j.febslet.2006.09.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Siemens J, Zhou S, Piskorowski R, et al. Spider toxins activate the capsaicin receptor to produce inflammatory pain. Nature. 2006;444:208–212. doi: 10.1038/nature05285. [DOI] [PubMed] [Google Scholar]
  • 112.Trevisani M, Smart D, Gunthorpe MJ, et al. Ethanol elicits and potentiates nociceptor responses via the vanilloid receptor-1. Nat Neurosci. 2002;5:546–551. doi: 10.1038/nn0602-852. [DOI] [PubMed] [Google Scholar]
  • 113.Chuang HH, Prescott ED, Kong H, et al. Bradykinin and nerve growth factor release the capsaicin receptor from PtdIns(4,5)P2-mediated inhibition. Nature. 2001;411:957–962. doi: 10.1038/35082088. [DOI] [PubMed] [Google Scholar]
  • 114.Moriyama T, Higashi T, Togashi K, et al. Sensitization of TRPV1 by EP1 and IP reveals peripheral nociceptive mechanism of prostaglandins. Mol Pain. 2005;1:3. doi: 10.1186/1744-8069-1-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Tominaga M, Wada M, Masu M. Potentiation of capsaicin receptor activity by metabotropic ATP receptors as a possible mechanism for ATP-evoked pain and hyperalgesia. Proc Natl Acad Sci U S A. 2001;98:6951–6956. doi: 10.1073/pnas.111025298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Negri L, Lattanzi R, Giannini E, et al. Impaired nociception and inflammatory pain sensation in mice lacking the prokineticin receptor PKR1: focus on interaction between PKR1 and the capsaicin receptor TRPV1 in pain behavior. J Neurosci. 2006;26:6716–6727. doi: 10.1523/JNEUROSCI.5403-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Diamond S, Freitag F, Phillips SB, Bernstein JE, Saper JR. Intranasal civamide for the acute treatment of migraine headache. Cephalalgia. 2000;20:597–602. doi: 10.1046/j.1468-2982.2000.00088.x. [DOI] [PubMed] [Google Scholar]
  • 118.Anand P, Bley K. Topical capsaicin for pain management: therapeutic potential and mechanisms of action of the new high-concentration capsaicin 8% patch. Br J Anaesth. 2011;107:490–502. doi: 10.1093/bja/aer260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Fusco BM, Barzoi G, Agro F. Repeated intranasal capsaicin applications to treat chronic migraine. Br J Anaesth. 2003;90:812. doi: 10.1093/bja/aeg572. [DOI] [PubMed] [Google Scholar]
  • 120.Gavva NR, Treanor JJ, Garami A, et al. Pharmacological blockade of the vanilloid receptor TRPV1 elicits marked hyperthermia in humans. Pain. 2008;136:202–210. doi: 10.1016/j.pain.2008.01.024. [DOI] [PubMed] [Google Scholar]
  • 121.Summ O, Holland PR, Akerman S, Goadsby PJ. TRPV1 receptor blockade is ineffective in different in vivo models of migraine. Cephalalgia. 2011;31:172–180. doi: 10.1177/0333102410375626. [DOI] [PubMed] [Google Scholar]
  • 122.Lambert GA, Davis JB, Appleby JM, Chizh BA, Hoskin KL, Zagami AS. The effects of the TRPV1 receptor antagonist SB-705498 on trigeminovascular sensitisation and neurotransmission. Naunyn Schmiedebergs Arch Pharmacol. 2009;380:311–325. doi: 10.1007/s00210-009-0437-5. [DOI] [PubMed] [Google Scholar]
  • 123.Chizh BA, O’Donnell MB, Napolitano A, et al. The effects of the TRPV1 antagonist SB-705498 on TRPV1 receptor-mediated activity and inflammatory hyperalgesia in humans. Pain. 2007;132:132–141. doi: 10.1016/j.pain.2007.06.006. [DOI] [PubMed] [Google Scholar]
  • 124.Kaube H, Hoskin KL, Goadsby PJ. Activation of the trigeminovascular system by mechanical distension of the superior sagittal sinus in the cat. Cephalalgia. 1992;12:133–136. doi: 10.1046/j.1468-2982.1992.1203133.x. [DOI] [PubMed] [Google Scholar]
  • 125.Levy D, Strassman AM. Mechanical response properties of A and C primary afferent neurons innervating the rat intracranial dura. J Neurophysiol. 2002;88:3021–3031. doi: 10.1152/jn.00029.2002. [DOI] [PubMed] [Google Scholar]
  • 126.Liedtke W, Tobin DM, Bargmann CI, Friedman JM. Mammalian TRPV4 (VR-OAC) directs behavioral responses to osmotic and mechanical stimuli in Caenorhabditis elegans. Proc Natl Acad Sci U S A. 2003;100 (Suppl 2):14531–14536. doi: 10.1073/pnas.2235619100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Vriens J, Watanabe H, Janssens A, Droogmans G, Voets T, Nilius B. Cell swelling, heat, and chemical agonists use distinct pathways for the activation of the cation channel TRPV4. Proc Natl Acad Sci U S A. 2004;101:396–401. doi: 10.1073/pnas.0303329101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Watanabe H, Davis JB, Smart D, et al. Activation of TRPV4 channels (hVRL-2/mTRP12) by phorbol derivatives. J Biol Chem. 2002;277:13569–13577. doi: 10.1074/jbc.M200062200. [DOI] [PubMed] [Google Scholar]
  • 129.Liedtke W, Choe Y, Marti-Renom MA, et al. Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor. Cell. 2000;103:525–535. doi: 10.1016/s0092-8674(00)00143-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Mizuno A, Matsumoto N, Imai M, Suzuki M. Impaired osmotic sensation in mice lacking TRPV4. Am J Physiol Cell Physiol. 2003;285:C96–101. doi: 10.1152/ajpcell.00559.2002. [DOI] [PubMed] [Google Scholar]
  • 131.Suzuki M, Mizuno A, Kodaira K, Imai M. Impaired pressure sensation in mice lacking TRPV4. J Biol Chem. 2003;278:22664–22668. doi: 10.1074/jbc.M302561200. [DOI] [PubMed] [Google Scholar]
  • 132.Kitahara T, Li HS, Balaban CD. Changes in transient receptor potential cation channel superfamily V (TRPV) mRNA expression in the mouse inner ear ganglia after kanamycin challenge. Hear Res. 2005;201:132–144. doi: 10.1016/j.heares.2004.09.007. [DOI] [PubMed] [Google Scholar]
  • 133.Wei X, Edelmayer RM, Yan J, Dussor G. Activation of TRPV4 on dural afferents produces headache-related behavior in a preclinical rat model. Cephalalgia. 2011;31:1595–1600. doi: 10.1177/0333102411427600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Burstein R, Cutrer MF, Yarnitsky D. The development of cutaneous allodynia during a migraine attack clinical evidence for the sequential recruitment of spinal and supraspinal nociceptive neurons in migraine. Brain. 2000;123 (Pt 8):1703–1709. doi: 10.1093/brain/123.8.1703. [DOI] [PubMed] [Google Scholar]
  • 135.Grant AD, Cottrell GS, Amadesi S, et al. Protease-activated receptor 2 sensitizes the transient receptor potential vanilloid 4 ion channel to cause mechanical hyperalgesia in mice. J Physiol. 2007;578:715–733. doi: 10.1113/jphysiol.2006.121111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Cenac N, Altier C, Chapman K, Liedtke W, Zamponi G, Vergnolle N. Transient receptor potential vanilloid-4 has a major role in visceral hypersensitivity symptoms. Gastroenterology. 2008;135:937–946. 946 e931–932. doi: 10.1053/j.gastro.2008.05.024. [DOI] [PubMed] [Google Scholar]
  • 137.Sipe WE, Brierley SM, Martin CM, et al. Transient receptor potential vanilloid 4 mediates protease activated receptor 2-induced sensitization of colonic afferent nerves and visceral hyperalgesia. Am J Physiol Gastrointest Liver Physiol. 2008;294:G1288–1298. doi: 10.1152/ajpgi.00002.2008. [DOI] [PubMed] [Google Scholar]
  • 138.Alessandri-Haber N, Joseph E, Dina OA, Liedtke W, Levine JD. TRPV4 mediates pain-related behavior induced by mild hypertonic stimuli in the presence of inflammatory mediator. Pain. 2005;118:70–79. doi: 10.1016/j.pain.2005.07.016. [DOI] [PubMed] [Google Scholar]
  • 139.Bautista DM, Siemens J, Glazer JM, et al. The menthol receptor TRPM8 is the principal detector of environmental cold. Nature. 2007;448:204–208. doi: 10.1038/nature05910. [DOI] [PubMed] [Google Scholar]
  • 140.Colburn RW, Lubin ML, Stone DJ, Jr, et al. Attenuated cold sensitivity in TRPM8 null mice. Neuron. 2007;54:379–386. doi: 10.1016/j.neuron.2007.04.017. [DOI] [PubMed] [Google Scholar]
  • 141.Dhaka A, Murray AN, Mathur J, Earley TJ, Petrus MJ, Patapoutian A. TRPM8 is required for cold sensation in mice. Neuron. 2007;54:371–378. doi: 10.1016/j.neuron.2007.02.024. [DOI] [PubMed] [Google Scholar]
  • 142.Burstein R, Yarnitsky D, Goor-Aryeh I, Ransil BJ, Bajwa ZH. An association between migraine and cutaneous allodynia. Ann Neurol. 2000;47:614–624. [PubMed] [Google Scholar]
  • 143.Chasman DI, Schurks M, Anttila V, et al. Genome-wide association study reveals three susceptibility loci for common migraine in the general population. Nat Genet. 2011;43:695–698. doi: 10.1038/ng.856. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Freilinger T, Anttila V, de Vries B, et al. Genome-wide association analysis identifies susceptibility loci for migraine without aura. Nat Genet. 2012;44:777–782. doi: 10.1038/ng.2307. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Ghosh J, Pradhan S, Mittal B. Genome-wide-associated variants in migraine susceptibility: a replication study from north India. Headache. 2013;53:1583–1594. doi: 10.1111/head.12240. [DOI] [PubMed] [Google Scholar]
  • 146.Newsom J, Holt JL, Neubert JK, Caudle R, Ahn AH. A high density of TRPM8 expressing sensory neurons in specialized structures of the head. Abstract retrieved from Society for Neuroscience. 2012 [Google Scholar]
  • 147.Burnstock G. P2X receptors in sensory neurones. Br J Anaesth. 2000;84:476–488. doi: 10.1093/oxfordjournals.bja.a013473. [DOI] [PubMed] [Google Scholar]
  • 148.Ford AP. In pursuit of P2X3 antagonists: novel therapeutics for chronic pain and afferent sensitization. Purinergic Signal. 2012;8:3–26. doi: 10.1007/s11302-011-9271-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Burnstock G. Purinergic P2 receptors as targets for novel analgesics. Pharmacol Ther. 2006;110:433–454. doi: 10.1016/j.pharmthera.2005.08.013. [DOI] [PubMed] [Google Scholar]
  • 150.Chen CC, Akopian AN, Sivilotti L, Colquhoun D, Burnstock G, Wood JN. A P2X purinoceptor expressed by a subset of sensory neurons. Nature. 1995;377:428–431. doi: 10.1038/377428a0. [DOI] [PubMed] [Google Scholar]
  • 151.North RA. Molecular physiology of P2X receptors. Physiol Rev. 2002;82:1013–1067. doi: 10.1152/physrev.00015.2002. [DOI] [PubMed] [Google Scholar]
  • 152.Staikopoulos V, Sessle BJ, Furness JB, Jennings EA. Localization of P2X2 and P2X3 receptors in rat trigeminal ganglion neurons. Neuroscience. 2007;144:208–216. doi: 10.1016/j.neuroscience.2006.09.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Bradbury EJ, Burnstock G, McMahon SB. The expression of P2X3 purinoreceptors in sensory neurons: effects of axotomy and glial-derived neurotrophic factor. Mol Cell Neurosci. 1998;12:256–268. doi: 10.1006/mcne.1998.0719. [DOI] [PubMed] [Google Scholar]
  • 154.Ruan HZ, Moules E, Burnstock G. Changes in P2X3 purinoceptors in sensory ganglia of the mouse during embryonic and postnatal development. Histochem Cell Biol. 2004;122:539–551. doi: 10.1007/s00418-004-0714-9. [DOI] [PubMed] [Google Scholar]
  • 155.Ambalavanar R, Moritani M, Dessem D. Trigeminal P2X3 receptor expression differs from dorsal root ganglion and is modulated by deep tissue inflammation. Pain. 2005;117:280–291. doi: 10.1016/j.pain.2005.06.029. [DOI] [PubMed] [Google Scholar]
  • 156.Simonetti M, Fabbro A, D’Arco M, et al. Comparison of P2X and TRPV1 receptors in ganglia or primary culture of trigeminal neurons and their modulation by NGF or serotonin. Mol Pain. 2006;2:11. doi: 10.1186/1744-8069-2-11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.D’Arco M, Giniatullin R, Simonetti M, et al. Neutralization of nerve growth factor induces plasticity of ATP-sensitive P2X3 receptors of nociceptive trigeminal ganglion neurons. J Neurosci. 2007;27:8190–8201. doi: 10.1523/JNEUROSCI.0713-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Fabbretti E, D’Arco M, Fabbro A, Simonetti M, Nistri A, Giniatullin R. Delayed upregulation of ATP P2X3 receptors of trigeminal sensory neurons by calcitonin gene-related peptide. J Neurosci. 2006;26:6163–6171. doi: 10.1523/JNEUROSCI.0647-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Simonetti M, Giniatullin R, Fabbretti E. Mechanisms mediating the enhanced gene transcription of P2X3 receptor by calcitonin gene-related peptide in trigeminal sensory neurons. J Biol Chem. 2008;283:18743–18752. doi: 10.1074/jbc.M800296200. [DOI] [PubMed] [Google Scholar]
  • 160.King BF, Ziganshina LE, Pintor J, Burnstock G. Full sensitivity of P2X2 purinoceptor to ATP revealed by changing extracellular pH. Br J Pharmacol. 1996;117:1371–1373. doi: 10.1111/j.1476-5381.1996.tb15293.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161.Wildman SS, King BF, Burnstock G. Zn2+ modulation of ATP-responses at recombinant P2X2 receptors and its dependence on extracellular pH. Br J Pharmacol. 1998;123:1214–1220. doi: 10.1038/sj.bjp.0701717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Gnanasekaran A, Sundukova M, van den Maagdenberg AM, Fabbretti E, Nistri A. Lipid rafts control P2X3 receptor distribution and function in trigeminal sensory neurons of a transgenic migraine mouse model. Mol Pain. 2011;7:77. doi: 10.1186/1744-8069-7-77. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 163.Schock SC, Munyao N, Yakubchyk Y, et al. Cortical spreading depression releases ATP into the extracellular space and purinergic receptor activation contributes to the induction of ischemic tolerance. Brain Res. 2007;1168:129–138. doi: 10.1016/j.brainres.2007.06.070. [DOI] [PubMed] [Google Scholar]
  • 164.Masterson CG, Durham PL. DHE repression of ATP-mediated sensitization of trigeminal ganglion neurons. Headache. 2010;50:1424–1439. doi: 10.1111/j.1526-4610.2010.01714.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 165.Gribkoff VK, Starrett JE, Jr, Dworetzky SI. Maxi-K potassium channels: form, function, and modulation of a class of endogenous regulators of intracellular calcium. Neuroscientist. 2001;7:166–177. doi: 10.1177/107385840100700211. [DOI] [PubMed] [Google Scholar]
  • 166.Robitaille R, Charlton MP. Presynaptic calcium signals and transmitter release are modulated by calcium-activated potassium channels. J Neurosci. 1992;12:297–305. doi: 10.1523/JNEUROSCI.12-01-00297.1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.Akerman S, Holland PR, Lasalandra MP, Goadsby PJ. Inhibition of trigeminovascular dural nociceptive afferents by Ca(2+)-activated K(+) (MaxiK/BK(Ca)) channel opening. Pain. 2010;151:128–136. doi: 10.1016/j.pain.2010.06.028. [DOI] [PubMed] [Google Scholar]
  • 168.Lu R, Lukowski R, Sausbier M, et al. BK channels expressed in sensory neurons modulate inflammatory pain in mice. Pain. 2013 doi: 10.1016/j.pain.2013.12.005. [DOI] [PubMed] [Google Scholar]
  • 169.Levy D, Strassman AM. Distinct sensitizing effects of the cAMP-PKA second messenger cascade on rat dural mechanonociceptors. J Physiol. 2002;538:483–493. doi: 10.1113/jphysiol.2001.013175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Vaughn AH, Gold MS. Ionic mechanisms underlying inflammatory mediator-induced sensitization of dural afferents. J Neurosci. 2010;30:7878–7888. doi: 10.1523/JNEUROSCI.6053-09.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Gold MS, Traub RJ. Cutaneous and colonic rat DRG neurons differ with respect to both baseline and PGE2-induced changes in passive and active electrophysiological properties. J Neurophysiol. 2004;91:2524–2531. doi: 10.1152/jn.00866.2003. [DOI] [PubMed] [Google Scholar]
  • 172.Fidan I, Yuksel S, Ymir T, Irkec C, Aksakal FN. The importance of cytokines, chemokines and nitric oxide in pathophysiology of migraine. J Neuroimmunol. 2006;171:184–188. doi: 10.1016/j.jneuroim.2005.10.005. [DOI] [PubMed] [Google Scholar]
  • 173.Sarchielli P, Alberti A, Baldi A, et al. Proinflammatory cytokines, adhesion molecules, and lymphocyte integrin expression in the internal jugular blood of migraine patients without aura assessed ictally. Headache. 2006;46:200–207. doi: 10.1111/j.1526-4610.2006.00337.x. [DOI] [PubMed] [Google Scholar]
  • 174.Huang M, Pang X, Karalis K, Theoharides TC. Stress-induced interleukin-6 release in mice is mast cell-dependent and more pronounced in Apolipoprotein E knockout mice. Cardiovasc Res. 2003;59:241–249. doi: 10.1016/s0008-6363(03)00340-7. [DOI] [PubMed] [Google Scholar]
  • 175.Melemedjian OK, Asiedu MN, Tillu DV, et al. IL-6- and NGF-induced rapid control of protein synthesis and nociceptive plasticity via convergent signaling to the eIF4F complex. J Neurosci. 2010;30:15113–15123. doi: 10.1523/JNEUROSCI.3947-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Yan J, Melemedjian OK, Price TJ, Dussor G. Sensitization of dural afferents underlies migraine-related behavior following meningeal application of interleukin-6 (IL-6) Mol Pain. 2012;8:6. doi: 10.1186/1744-8069-8-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Stewart WF, Lipton RB, Celentano DD, Reed ML. Prevalence of migraine headache in the United States. Relation to age, income, race, and other sociodemographic factors. JAMA. 1992;267:64–69. [PubMed] [Google Scholar]
  • 178.Lipton RB, Stewart WF. Migraine in the United States: a review of epidemiology and health care use. Neurology. 1993;43:S6–10. [PubMed] [Google Scholar]
  • 179.Scheff NN, Gold MS. Sex differences in the inflammatory mediator-induced sensitization of dural afferents. J Neurophysiol. 2011;106:1662–1668. doi: 10.1152/jn.00196.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Ophoff RA, Terwindt GM, Vergouwe MN, et al. Familial hemiplegic migraine and episodic ataxia type-2 are caused by mutations in the Ca2+ channel gene CACNL1A4. Cell. 1996;87:543–552. doi: 10.1016/s0092-8674(00)81373-2. [DOI] [PubMed] [Google Scholar]
  • 181.Pietrobon D. Calcium channels and migraine. Biochim Biophys Acta. 2012 doi: 10.1016/j.bbamem.2012.11.012. [DOI] [PubMed] [Google Scholar]
  • 182.De Fusco M, Marconi R, Silvestri L, et al. Haploinsufficiency of ATP1A2 encoding the Na+/K+ pump alpha2 subunit associated with familial hemiplegic migraine type 2. Nat Genet. 2003;33:192–196. doi: 10.1038/ng1081. [DOI] [PubMed] [Google Scholar]
  • 183.Dichgans M, Freilinger T, Eckstein G, et al. Mutation in the neuronal voltage-gated sodium channel SCN1A in familial hemiplegic migraine. Lancet. 2005;366:371–377. doi: 10.1016/S0140-6736(05)66786-4. [DOI] [PubMed] [Google Scholar]
  • 184.Pietrobon D. Familial hemiplegic migraine. Neurotherapeutics. 2007;4:274–284. doi: 10.1016/j.nurt.2007.01.008. [DOI] [PubMed] [Google Scholar]
  • 185.de Vries B, Frants RR, Ferrari MD, van den Maagdenberg AM. Molecular genetics of migraine. Hum Genet. 2009;126:115–132. doi: 10.1007/s00439-009-0684-z. [DOI] [PubMed] [Google Scholar]
  • 186.Lafreniere RG, Cader MZ, Poulin JF, et al. A dominant-negative mutation in the TRESK potassium channel is linked to familial migraine with aura. Nat Med. 2010;16:1157–1160. doi: 10.1038/nm.2216. [DOI] [PubMed] [Google Scholar]
  • 187.Hodgkin AL, Huxley AF. Potassium leakage from an active nerve fibre. J Physiol. 1947;106:341–367. doi: 10.1113/jphysiol.1947.sp004216. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Enyedi P, Czirjak G. Molecular background of leak K+ currents: two-pore domain potassium channels. Physiol Rev. 2010;90:559–605. doi: 10.1152/physrev.00029.2009. [DOI] [PubMed] [Google Scholar]
  • 189.Duprat F, Lesage F, Fink M, Reyes R, Heurteaux C, Lazdunski M. TASK, a human background K+ channel to sense external pH variations near physiological pH. EMBO J. 1997;16:5464–5471. doi: 10.1093/emboj/16.17.5464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Lafreniere RG, Rouleau GA. Migraine: Role of the TRESK two-pore potassium channel. Int J Biochem Cell Biol. 2011;43:1533–1536. doi: 10.1016/j.biocel.2011.08.002. [DOI] [PubMed] [Google Scholar]
  • 191.Liu P, Xiao Z, Ren F, et al. Functional Analysis of a Migraine-Associated TRESK K+ Channel Mutation. J Neurosci. 2013;33:12810–12824. doi: 10.1523/JNEUROSCI.1237-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Dobler T, Springauf A, Tovornik S, et al. TRESK two-pore-domain K+ channels constitute a significant component of background potassium currents in murine dorsal root ganglion neurones. J Physiol. 2007;585:867–879. doi: 10.1113/jphysiol.2007.145649. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 193.Liu C, Au JD, Zou HL, Cotten JF, Yost CS. Potent activation of the human tandem pore domain K channel TRESK with clinical concentrations of volatile anesthetics. Anesth Analg. 2004;99:1715–1722. doi: 10.1213/01.ANE.0000136849.07384.44. table of contents. [DOI] [PubMed] [Google Scholar]
  • 194.Andres-Enguix I, Shang L, Stansfeld PJ, et al. Functional analysis of missense variants in the TRESK (KCNK18) K channel. Sci Rep. 2012;2:237. doi: 10.1038/srep00237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Gu W, Schlichthorl G, Hirsch JR, et al. Expression pattern and functional characteristics of two novel splice variants of the two-pore-domain potassium channel TREK-2. J Physiol. 2002;539:657–668. doi: 10.1113/jphysiol.2001.013432. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Xian Tao L, Dyachenko V, Zuzarte M, et al. The stretch-activated potassium channel TREK-1 in rat cardiac ventricular muscle. Cardiovasc Res. 2006;69:86–97. doi: 10.1016/j.cardiores.2005.08.018. [DOI] [PubMed] [Google Scholar]
  • 197.Maher BH, Griffiths LR. Identification of molecular genetic factors that influence migraine. Mol Genet Genomics. 2011;285:433–446. doi: 10.1007/s00438-011-0622-3. [DOI] [PubMed] [Google Scholar]

RESOURCES