Next Article in Journal
ATP13A2 Regulates Cellular α-Synuclein Multimerization, Membrane Association, and Externalization
Previous Article in Journal
Targeting the Extracellular Matrix in Abdominal Aortic Aneurysms Using Molecular Imaging Insights
Previous Article in Special Issue
Copper Dyshomeostasis in Neurodegenerative Diseases—Therapeutic Implications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Migraine: Calcium Channels and Glia

1
Laboratory of Neurobiology, Department of Neurology, Poznan University of Medical Sciences, 49 Przybyszewskiego St., 60-355 Poznan, Poland
2
Chair and Department of Neurology, Poznan University of Medical Sciences, 49 Przybyszewskiego St., 60-355 Poznan, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(5), 2688; https://doi.org/10.3390/ijms22052688
Submission received: 23 January 2021 / Revised: 26 February 2021 / Accepted: 3 March 2021 / Published: 7 March 2021

Abstract

:
Migraine is a common neurological disease that affects about 11% of the adult population. The disease is divided into two main clinical subtypes: migraine with aura and migraine without aura. According to the neurovascular theory of migraine, the activation of the trigeminovascular system (TGVS) and the release of numerous neuropeptides, including calcitonin gene-related peptide (CGRP) are involved in headache pathogenesis. TGVS can be activated by cortical spreading depression (CSD), a phenomenon responsible for the aura. The mechanism of CSD, stemming in part from aberrant interactions between neurons and glia have been studied in models of familial hemiplegic migraine (FHM), a rare monogenic form of migraine with aura. The present review focuses on those interactions, especially as seen in FHM type 1, a variant of the disease caused by a mutation in CACNA1A, which encodes the α1A subunit of the P/Q-type voltage-gated calcium channel.

1. Review Criteria

Articles discussed in this Review were identified by PubMed searches for the years 1990 to the present, using the search terms “migraine and calcium”, “migraine and calcium signaling”, “CACNA1A mutations”, “migraine and CaV2.1”, “migraine and glia”, “glia and calcium in migraine” among others. The reference lists of identified papers were searched for further relevant articles, and related citations for identified papers as listed on the PubMed site were also evaluated.

2. Migraine

Migraine is a common primary headache disorder that affects 11% of adults worldwide. The prevalence of disease is three times higher in women (15–18%) than in men [1]. Two peaks of incidence have been observed among migraine sufferers: the first after puberty and the second in adults aged 35–40 years [2]. In 25% of cases, the migraine begins in childhood. According to World Health Organization (WHO) data 324 million people struggle with this disease and 3000 migraine attacks occur every day per one million people [3]. The Global Burden of Disease study listed migraine as the third of 289 of the most prevalent diseases worldwide [4].
The disease is divided into two main clinical subtypes: migraine with aura (MA) and migraine without aura (MO). Aura is defined as spreading neurological disturbances such as visual, sensory or motor symptoms that precede or accompany the headache. The most common clinical features of aura are visual changes including flashing scotoma, loss of vision, and visual hallucinations. More rarely, numbness, tingling, ataxia, aphasia, confusion, ringing in the ears, and dizziness occur. MO, also called common migraine, occurs in two-thirds of patients. The migraine pain lasts 4–72 h, is moderate or severe, unilateral, throbbing, worsens with physical activity, and is often accompanied by photophobia, phonophobia, and nausea/vomiting [5,6,7,8]. Migraine is also classified according to the frequency of attacks into chronic, lasting at least 15 days each month, and milder, episodic forms. Chronic migraine, sometimes called transformed, affects about 2% of the population and occurs after many years of a typical, episodic migraine. The tendency to transformation may be increased by coexisting depression, anxiety, panic attacks, social phobia and by other pain syndromes [4,9].

2.1. The Pathogenesis of Migraine

Migraine pathophysiology involves complex mechanisms in which the trigeminovascular system (TGVS) and cortical spreading depression (CSD) play an important role (Figure 1) [10]. The TGVS is a major afferent pathway for pain from cranial vessels and dura mater and consists of neurons whose bodies reside in the trigeminal ganglion (TG) and upper cervical dorsal root ganglia [11]. The TG consists mainly of primary afferent neurons and glial cells. As it is not protected by the blood–brain barrier (BBB), neuropeptides released in the TG, such as calcitonin gene-related peptide (CGRP), substance P (SP), and neurokinin A (NKA) thereby enter systemic circulation [12].
CSD consists of a slow wave of depolarization followed by longer-lasting suppression of neurons and glial cells. It is characterized by increased K+ and decreased Na+ extracellular levels and changes in the gradients of other ions, e.g., Mg2+, Zn2+, Cl [13]. Although the direct triggers for spontaneous CSD, especially in migraine, are unknown, elevated extracellular concentrations of K+ and glutamate are thought to cause dendritic depolarization in a non-synaptic manner. Animal model studies have suggested that CSD is the mechanism underlying visual aura in migraine [14], a conclusion corroborated by neuroimaging studies which have linked electrical changes in the visual cortex during aura (geometrical shapes, scintillating scotoma) with the pattern of CSD [15].
According to the neurovascular theory of migraine, headache is a result of TGVS activation by CSD. CSD triggers the pain pathway via activation of trigeminal afferents which transmit information to the TG, the caudal trigeminal nucleus (TNC) and ultimately to cortical and brainstem structures involved in pain processing (Figure 1). Activation of the TGVS leads to local release of vasodilators such as CGRP, SP, NKA, NO, and a transient increase in cortical blood flow followed by sustained flow decrease [16,17,18].

2.2. Migraine Disturbs Calcium Homeostasis

Currently, it is believed that the distribution of various ions between intracellular and extracellular compartments is altered in migraine. As aforementioned, increased extracellular K+ and decreased Na+ promote CSD [13]. Calcium levels are also altered in the course of migraine [19]. These findings have led scientists to argue that migraine is a channelopathy. It is believed that mutations in genes encoding channel subunits or proteins modulating channel function lead to ionic disturbances in synapses, thereby increasing susceptibility to CSD and migraines [20,21,22].
Dysregulated calcium currents as seen in the context of migraine derive largely from aberrant function of the high-voltage activated calcium channel CaV2.1 and the transient receptor potential ankyrin channel, encoded by the genes CACNA1A and TRPA1, respectively (Figure 2) [20,21,23]. Localized presynaptically, the CaV2.1 channel plays an important role in communication between neurons by controlling the release of neurotransmitters [24]. Certain mutations in the CACNA1A gene result in increased activation of CaV2.1, which in turn leads to increased intracellular Ca2+ [20,21,22,25]. TRPA1, a nonselective Ca2+-permeable ion channel, belongs to a family of transient potential receptors serving as modalities for the sensation of environmental stimuli. Expressed on Aδ and C afferent fibers, TRPA1 transduces pain from a broad array of irritants, both food (e.g., allyl isothiocyanate in mustard) and chemical (e.g., formaldehyde) and is thought to be implicated in the pathogenesis of headache [23,26]. Although the mechanisms whereby environmental irritants cause headache remain largely unknown, activation of TRPA1 by mechanical or chemical stimuli causes CGRP to be released and increases cerebral blood flow [26,27].
Beyond CaV2.1, other voltage-gated calcium channels (VGCCs) may play a role in the pathogenesis of migraine [28]. Interestingly, while presynaptic CaV2 channels might be expected to drive the release of CGRP associated with migraine, the high-voltage activated and canonically postsynaptic CaV1 channels and the low-voltage activated CaV3 channels [24,29] were both also found to regulate CGRP release in the trigeminal ganglion, as evidenced through pharmacological blockade experiments [30]. Further, by correlating the genetic codependency of Ca2+ levels with the risk of migraine headache, Yin et al. [31] showed that migraine can be linked to inherited hypercalcemia.
Several classes of drugs are currently used in the treatment and prevention of migraine, including angiotensin converting enzyme inhibitors, angiotensin receptor blockers, Ca2+ channel blockers, serotonin antagonists, alpha adrenergic agonists, and NMDA receptor antagonists [32]. Ca2+ channel blockers exert their effect by blocking Ca2+ influx into vascular smooth muscle and cardiac muscle cells during membrane depolarization, leading to a reduction in blood pressure. Their main application is in the treatment of arterial hypertension and angina and while used also for migraine pain relief therapy, they may themselves result in side effects such as headaches and dizziness. As Ca2+ channel blockers are specific for different VGCCs and vary broadly in their pharmacologic effects, the development of therapeutics that selectively target those channels which are centrally expressed and implicated in migraine pathogenesis is of high research value.

2.3. Experimental Evidence for the Role of CGRP in Migraine

Literature reports suggest that CGRP may be a principal mediator of migraine in the TGVS [33]. The CGRP neuropeptide is expressed in nearly half of TG neurons, mostly in those forming unmyelinated nociceptive C-fibers. Conversely, the CGRP receptor is expressed by the majority of neurons forming the second class of TG fibers, myelinated A-fibers, as well as in smooth muscle tissue of the dura’s vasculature [34]. There are two isoforms of CGRP, differing by three amino acids and the tissues where they are predominantly expressed. α-CGRP is most abundant in the CNS and peripheral nerves of the somatosensory system, while the β-isoform is mostly present in the enteric nervous system and motor neurons [35]. The connection between CGRP and migraine has been documented quite well, however a clear mechanistic explanation for its role in the disease is lacking. It has been shown that CGRP administered intravenously induces a delay in the onset of a migraine attack in individuals with a history of migraine, but not in healthy controls. Notably, CGRP is not allogenic, since it does not provoke any somatic effects beyond headache [36] or an erythemic flare in the skin [37]. Interestingly, while CGRP infusion was shown to induce delayed headache in patients suffering from migraines, the neuromodulator did not cause premonitory symptoms, such as aura, suggesting attacks may be periphery-derived [38]. Moreover, CGRP is likely released into the bloodstream during a migraine attack, especially seeing how it has been detected in samples drawn from the jugular vein of sufferers [16,36,37,38,39,40,41]. Increased CGRP levels have also been observed between attacks in migraine patients as compared to controls. As such, CGRP may be associated with the chronic form of migraine, particularly given that in this disease subtype, the level of the neuropeptide was significantly higher than in episodic migraine patients [41]. The level of CGRP was also shown to be elevated in a nitroglycerin model of migraine, which responded to the triptan family of drugs commonly used as abortive therapy for the disease [42]. Specifically, administration of sumatriptan significantly reduced CGRP release from the TG, and secondarily caused significant decline of blood CGRP levels, along with reduced headache intensity. Another study evaluating rizatriptan correlated higher salivary CGRP levels with better response to the drug, thereby potentially providing a way for identifying promising patient candidates for specific anti-migraine therapies [43]. Taken together, the above studies indicate a pivotal role of CGRP in TGVS regulation of migraine attacks and indeed, CGRP is becoming a target point for emerging therapies. Several clinical studies tackling the CGRP pathway for relieving migraine showed promising results and involve CGRP antagonists as well as antibodies against the CGRP receptor and CGRP itself. The CGRP antagonists (BI 44,370 TA, MK-3207, olcegepant, imegepant, telcagepant, and ubrogepant) were shown to be effective as migraine abortive therapy [44,45,46,47,48,49,50]. Further, various antibodies against CGRP: eptinezumab, fremanezumab and galcanezumab proved to be effective for various forms of migraine, while retaining a safe therapeutic profile [51,52,53,54,55,56]. Similarly promising results were noted for erenumab (AMG 334), a monoclonal antibody against the CGRP receptor [57,58].

2.4. The Role of CGRP and Glial Cells in Migraine Pathogenesis

Since the mid-1980s, it has been postulated that CGRP integrally influences the neuron-glial interactions associated with migraine pain propagation [33]. CGRP is likely released from TG neurons in response to migraine triggers e.g., stress and hypoxia communicated via afferents from the periphery [59]. Moreover, although a direct link between CGRP and CSD has not been established, increased extracellular K+, seen as a condition for the latter, may drive CGRP release (Figure 2) [25]. CGRP triggers NO production which in turn leads to increased expression of CGRP and neuronal nitric oxide synthase (nNOS) [60] creating a positive feedback loop that promotes sensitization of primary peripheral trigeminal fibers and activity-independence of central second-order neurons [61]. These mediators further stimulate the surrounding glial cells, termed satellite cells, to produce interleukin-1β (IL-1β) which promotes increased cyclooxygenase activity, such that is tied to the production of proinflammatory prostaglandin E2 (PGE2) [62,63,64]. Similarly, CGRP-stimulated satellite cells release TNFα, which has also been implicated in the positive feedback driving TG neuron sensitization, both directly and via the release of additional proinflammatory cytokines [65,66]. Indeed, the literature suggests that the secretion of CGRP from one type of TG neuron may induce cytokine secretion in other TG neurons as well as in adjacent satellite glial cells [62,67,68]. It is the TG satellite cell-mediated up-regulation of specific proalgogenic receptors combined with long-term sensitization of TG neurons that enhances pain [61]. Further corroborating this hypothesis, a study by Cady et al. [69] demonstrated that CGRP injection into the rat temporomandibular joint led to formation of the activated phenotype of both microglia and astrocytes in the TNC. This phenomenon is thought to be responsible for maintaining the central sensitization of neurons involved in pain perception and is likely implicated in the progression from episodic to chronic migraine [61]. Conversely, Cornelison et al. [70] showed that administering CGRP into the cisternae of the rat brain lead to activation of astrocytes but not microglia, as the authors observed no change in the level of the microglial marker ionized calcium-binding adapter molecule (Iba1), but did note enhanced expression of the astrocytic markers glial fibrillary acidic protein (GFAP) and protein kinase A (PKA).
Several inhibitors of glial cell activation including naltrexone, naloxone, minocycline and ibudilast have been proposed as prophylactics against migraine [71]. However, clinical trial results published by Kwok et al. [66] demonstrated similar attack frequency and intensity, as well as no changes in allodynia, quality of life, medication use, or the secondary measures of headache in chronic migraine patients who received ibudilast for eight weeks. Still, comparable studies have not been performed in patients with episodic migraine, leaving the question of whether ibdulast could prove effective in preventing the transformation of episodic migraine into a chronic disorder unanswered.
The interactions between activated trigeminal neurons and adjacent glial cells which are mediated by gap junctions and paracrine signaling are likely also connected to the development of peripheral sensitization within the TG and other elements of migraine pathogenesis [64]. It is hypothesized that elevated expression and activity of gap junctions and pannexin (Panx) channels at the level of the sensory ganglia and TG neurons in inflammatory and neuropathic models of pain may lead to augmented excitation of sensory neurons. It is also worth noting, the gap junctions and Panx in glial cells may contribute to development of migraine with aura, as they facilitate the spreading of signals between satellite glial cells, including Ca2+ waves [72].
Ca2+ spreading among glia and the aberrant transport of other ions may be associated in particular with a certain subtype of migraine with aura, called familial hemiplegic migraine (FHM). Despite the fact that neuron-glia co-sensitization occurs often in the course of migraine, in FHM the role of CGRP seems limited, as shown by the studies of Hansen et al. [73]. This fuels the suspicion that disturbances in ion transport within the brain arise from CGRP-independent processes [74].

2.5. Familial Hemiplegic Migraine

As the TGVS richly expresses ion channels [30], the hypothesis that migraine headache may be a result of dysregulated nerve excitation due to one or more channelopathies is an attractive one. Moreover, migraine shares clinical similarities with other channelopathies, e.g., myotonia or periodic paralysis including frequency and duration of attacks, paroxysmal character, triggers for attack, and gender-related predilection for attack [74]. Genetic studies of FHM have also substantiated the hypothesis that migraine, or at least aurae arise as a result of ionopathy.
FHM, the most severe subtype of MA, is characterized by the presence of temporary unilateral hemiparesis (numbness and/or motor weakness). Usually the symptoms of FHM start in the first or second decade of life. They may be accompanied by cerebellar atrophy and disturbances in cerebral blood flow. Less frequent are atypical attacks with cerebellar signs, encephalopathy, coma, prolonged hemiplegia, epileptic seizure, confusion, or fever, with full recovery or nystagmus and ataxia between attacks [75].
FHM is a rare, genetically heterogeneous disease, inherited in an autosomal dominant pattern with approximately 70–90% penetrance [76]. Three causative genes for FHM have been identified: CACNA1A (FHM type 1), ATP1A2 (FHM type 2) and SCN1A (FHM type 3) [77,78,79]. The CACNA1A gene encodes the α1A subunit of the P/Q-type high-voltage activated calcium channel. The ATP1A2 gene encodes the catalytic α2 subunit of Na+/K+ ATPase, which is exclusively expressed in astrocytes where it maintains the electrochemical gradient of Na+ and K+ ions essential for transport of Ca2+ and glutamate. The Na+/K+ ATPase also regulates the reuptake of K+ and glutamate. The elevated level of K+ due to FHM2 mutations triggers CSD [80]. The SCN1A gene encodes the α subunit of the neuronal voltage-gated sodium channel (Nav1.1). Nav1.1 is mainly expressed in the cerebral cortex and spinal cord where it is responsible for the generation and propagation of action potentials. Mutations in FHM genes occur also in epileptic patients [81].

3. Structure and Functions of CaV2.1

Voltage-dependent Ca2+ channels are multiprotein complexes consisting of α1, α2δ, β and γ subunits [24]. The structural and functional diversity of VGCCs results from the multiple isoforms of each subunit, especially α1, different gating kinetics, and the many proteins with which they interact, often via transient, low-affinity molecular interactions (Table 1). CaV1, CaV2 and CaV3 are paralogs which have arisen through gene duplication events unaccompanied by speciation. CaV1 channels control synaptic integration and modulate NMDA receptor-mediated plasticity at the post-synapse, regulate enzyme activity and gene expression, and initiate excitation-contraction coupling. CaV2 channels mediate neurotransmission at the pre-synaptic active zone, while the CaV3 subfamily modulates the depolarization threshold for action potential initiation in cardiomyocytes and thalamic neurons. Both CaV1 and CaV2 are classified as high voltage-activated, while CaV3s are low voltage-activated channels [24,29,82].
Of all of the culprits potentially implicated in migraine channelopathy, the P/Q-type CaV2.1 channel has received the most attention. CaV2.1 channels are present in presynaptic terminals and somatodendritic membranes in the brain and spinal cord [83]. Importantly, they are expressed in brain regions responsible for nociception or even strongly implicated in migraine pathogenesis e.g., TG and brainstem and control the release of vasoactive neuropeptides in the TGVS [84].
As presented in Figure 3, the α1A subunit of CaV2.1 is formed from four homologous domains (I–IV) consisting of six transmembrane regions (S1–S6). The S4 region constitutes the voltage sensor, while S5, the P-loop, and S6 form the pore region, which determines ion selectivity and conductance properties [24,29,82].
CaV2.1 channels control action-potential evoked neurotransmitter release, by triggering activation of the exocytotic machinery at the pre-synapse [25]. CaV2.1s are able to interact with numerous Ca2+-binding proteins [24] and maintain short-term plasticity via Ca2+-dependent inactivation and Ca2+-dependent facilitation. CaV2.1 channels are also involved in local excitability of neurons, Ca2+ signaling, cell survival or gene expression [85]. Remarkably, CaV2.1 channels facilitate both fast neurotransmission and modulation of neuromuscular transmission of acetylcholine mediated by muscarinic M1 and M2 receptors and protein kinases A and C [86,87].

3.1. Mutations in CACNA1A

The CACNA1A gene consists of 47 exons and is located on chromosome 19p13. Age and gender-dependent alternative splice variants of exon 37 were found, which correspond to differences in channel kinetics and confer subfunctionalities to isoforms expressed in different brain regions [88].
Mutations in CACNA1A are associated with a few neurological diseases, including FHM, episodic ataxia type 2 (EA2), spinocerebellar ataxia type 6 (SCA6) and nonprogressive congenital ataxia (NPCA), and epilepsy [89]. Patients with FHM may present with symptoms of ataxia. Mutations in FHM genes are not found in common forms of MA and MO. CACNA1A mutations are responsible for about half of FHM cases. 21 FHM1 mutations were identified, all of them missense mutations leading to substitutions of amino acids in functional regions of the CaV2.1 channel, with the majority occurring in transmembrane segments of the α1 subunit (Figure 3).
Generally, different mutations are associated with pure FHM1 and FHM1 with cerebellar symptoms. For example, the R192Q mutation is responsible for a mild form of FHM1, whereas the S218L mutation causes a severe, often lethal phenotype [90]. Moreover, individuals with the same mutation may differ in terms of clinical symptoms, which suggests that epigenetic and environmental factors may be involved in determining phenotype (Table 1) [75].
Mutations in CACNA1A may be divided into three groups: gain of function, loss of function, or biallelic mutations. Half of FHM1 cases are caused by gain of function mutations. 13 of FHM1 mutations (R192Q, S218L, R583Q, T666M, V714A, D715E, Y1246C, K1336E, V1457L, W1684R, V1696I, I1710T, I1811L) were investigated in heterologous expression systems expressing recombinant CaV2.1 channels [105,106,107,108,109,110,111,112,113,114,115,116,117]; some of them (R192Q, S218L, T666M, V714A, I1811L) were also studied in neurons from CaV2.1−/− mice expressing human CaV2.1 α1 subunits [108,111,118,119,120].
Studies on the HEK293 cell line [106] as well as on CaV2.1−/− neurons [108] evidenced that there is a decreased density of functional CaV2.1 in FHM1 mutant cells as compared to WT. This, however, has been rationalized as an artifact of overexpression post-transfection insofar as decreased numbers of functional channels were not observed when hCaV2.1 were expressed endogenously in knock-in mice [25]. Nonetheless, other studies on CaV2.1−/− neurons found no differences in synaptic strength [118,119]. Further, CaV2.1−/+ mice have a reduced response to neuroinflammatory and neuropathic pain due to having roughly half the number of CaV2.1 channels, albeit in an age-dependent manner. This is compatible with the idea that CaV2.1 channels may be pronociceptive in terms of inflammation and neuropathic pain and antinociceptive in response to acute non-injurious noxious thermal stimuli.
CACNA1A mutations result in a wide spectrum of consequences including increased channel open probability, a lower voltage of activation leading to enhanced Ca2+ influx, reduced channel inhibition by G-protein βγ heterodimers, altered synaptic morphology, excitation-inhibition imbalance and enhanced glutamate release. According to in vitro models of CSD, the increased glutamate release and interaction between glutamate and pre- and postsynaptic glutamate NMDA receptors may facilitate CSD. Importantly, CACNA1A mutations promoting increased Ca2+ influx seem to selectively capacitate glutamate release at pyramidal neurons, without altering fast-spiking inhibitory interneurons [117]. Further, the excitotoxic effects of glutamate on neuronal and glial cells have been shown to alter brain energy metabolism [117,120,121]. A study by Eikermann-Haerter et al. [115] showed that mice expressing the R192Q CACNA1 mutation were more sensitive to CSD than those bearing the S218L variant. Moreover, female mutant mice were more susceptible to CSD and neurological deficits than males. The R192Q mouse model pointed not only to a decreased threshold for CSD but also increased CaV2.1 current density [122]. Susceptibility to CSD can also be heightened by female hormones and allele dosage. In addition, FHM1 mice showed greater oxygen consumption leading to tissue anoxia, which may be responsible for prolonged aura [117].

3.2. Calcium-Related Therapeutics in Migraine

Migraine pharmacotherapy includes prophylactic agents taken every day (e.g., antidepressants and antiepileptic drugs) and agents taken at headache onset (e.g., triptans). Nonetheless, migraine patients typically migrate to analgesics, typically nonsteroidal anti-inflammatory drugs (NSAIDs) available without prescription, despite mixed reports over their efficacy. Unfortunately, triptans, selective 5-HT1B/1D receptor agonists, which make up 80% of prescribed medications, proved to be effective in only 60% of migraine patients not responding to NSAIDs [123,124,125].
Unfortunately, most drugs currently in use do not prevent the recurrence of migraine attacks. Thus, migraine may be the cause of numerous absences at school, or at work. It was calculated that the costs of absenteeism from work due to migraines are much higher than the costs of its treatment. Important also is the fact that 3–4 times more funds are spent for the treatment of chronic migraine in comparison with the episodic subtype. It is estimated that in the USA the annual total costs incurred for migraine, taking into account both the treatment, as well as losses due to absence from work add up to as much as 13 billion USD [126]. The lack of efficacious analgesics or means to prevent migraine recurrence in many patients has created a need for finding novel therapeutics.
Currently, there are no selective small molecule inhibitors of CaV2.1. Only two peptide toxins are selective for CaV2.1: ω-agatoxin IVA and ω-agatoxin IVB isolated from the venom of American funnel web spider A. aperta. Both ω-agatoxins bind on the outside of the pore region of α1a subunit of CaV2.1 but have different kinetics: blocking by ω-agatoxin IVB is eightfold slower than IVA [127]. However, there are drugs that are not selective for but act also at CaV2.1 and show clinical efficacy in patients with FHM1 including flunarizine and the non-dihydropyridine verapamil [128].
New targets for migraine therapy may focused on different Ca2+ channels. One class worth noting is acid-sensing ion channels (ASICs), among which the ASIC1 predominates in the CNS. ASIC1 channels are responsible for enhancing Ca2+ permeability and pain signaling [129] and their overexpression has been found in chronic inflammatory and neuropathic models [130]. How ASICs are implicated in migraine pathogenesis remains largely unknown. Still, ASICs are activated by decreases in pH regulated in part by serotonin (5-HT) and nerve growth factor (NGF), changes in both of which are observed in migraine patients [131,132]. Secondly, ASIC1a is overexpressed in hypothalamic orexinergic neurons thought to be involved in migraine pathophysiology. Finally, lower pH secondary to ASIC activation may initiate or propagate CSD [129,132]. Corroborating this hypothesis, a study using experimental models showed that both amiloride (a nonselective blocker of ASICs) and tarantula toxin PcTx1 (a selective blocker of ASIC1a) inhibited CSD [133].

4. Conclusions

Migraine is a multifactorial neurological disease whose pathogenesis has not been fully elucidated. Although numerous dysregulated processes likely converge to manifest as migraine, disturbances in the distribution of various ions between extracellular and intracellular compartments seem to be a common denominator in the pathophysiology of the disease. Migraine ”ionopathy” likely derives from one or multiple channelopathies which may be inherited e.g., mutations in CACNA1A and TRPA1, or result from the effects of toxic environmental factors. Certain parts of the brain appear particularly impacted by migraine-associated channelopathies. The TG, which is the major afferent pathway for pain signals from cranial vessels and dura mater, is implicated in migraine pathogenesis largely owing to its release of neuropeptides such as CGRP. CGRP is one of the most important molecules involved in neuron-glial communication, whose breakdown is related to the propagation of migraine pain and the sensitization of neurons which likely precedes the progression from episodic to chronic migraine. How calcium channelopathies impact CGRP pathways however, remains largely unknown. CSD, a purportedly non-synaptic phenomenon, and gap-junctional interactions also play a role in MA by driving aberrant neuron-neuron and neuron-glia signal propagation. As is the case with other neurological disorders, the complex and still poorly understood pathogenesis of migraine makes treatment difficult. Nonetheless, the pivotal role of CGRP in the disease has made it a prominent therapeutic target.

Author Contributions

M.K., M.P., T.P., J.D.; writing—original draft preparation, W.K.; investigation, J.D.; investigation. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the grant of NCBiR (European Union), No. POWR.03.01.00-00-T006/18-00.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bolay, H.; Ozge, A.; Saginc, P.; Orekici, G.; Uludüz, D.; Yalın, O.; Siva, A.; Bıçakçi, S.; Karakurum, B.; Öztürk, M. Gender influences headache characteristics with increasing age in migraine patients. Cephalalgia 2015, 35, 792–800. [Google Scholar] [CrossRef]
  2. Silberstein, S.D. Migraine. Lancet 2004, 363, 381–391. [Google Scholar] [CrossRef]
  3. World Health Organization. Atlas of Headache Disorders and Resources in the World 2011; World Health Organisation: Geneva, Switzerland, 2011.
  4. Vos, T.; Flaxman, A.D.; Naghavi, M.; Lozano, R.; Michaud, C.; Ezzati, M.; Shibuya, K.A.; Salomon, J.; Abdalla, S.; Aboyans, V.; et al. Years lived with disability (YLDs) for 1160 sequelae of 289 diseases and injuries 1990–2010: A systematic analysis for the Global Burden of Disease Study 2010. Lancet 2012, 380, 2163–2196. [Google Scholar] [CrossRef]
  5. Simon, R.P.; Aminoff, M.J.; Greenberg, D.A. Clinical Neurology, 7th ed.; McGraw-Hill Professional Publishing: New York, NY, USA, 2009; Available online: http://public.ebookcentral.proquest.com/choice/publicfullrecord.aspx?p=4667887 (accessed on 30 October 2020).
  6. Viana, M.; Sprenger, T.; Andelova, M.; Goadsby, P.J. The typical duration of migraine aura: A systematic review. Cephalalgia 2013, 33, 483–490. [Google Scholar] [CrossRef]
  7. Headache Classification Committee of the International Headache Society (IHS). The International Classification of Headache Disorders; (beta version), 3rd ed. Cephalalgia 2013, 33, 629–808. [Google Scholar] [CrossRef] [Green Version]
  8. Headache Classification Committee of the International Headache Society (IHS). The International Classification Of Headache Disorders. 3rd edition. Cephalalgia 2018, 38, 1–211. [Google Scholar] [CrossRef]
  9. Menken, M.; Munsat, T.L.; Toole, J.F. The global burden of disease study: Implications for neurology. Arch. Neurol. 2000, 57, 418. [Google Scholar] [CrossRef]
  10. Pietrobon, D. Calcium channels and migraine. Biochim. Biophys. Acta (BBA) Biomembr. 2013, 1828, 1655–1665. [Google Scholar] [CrossRef]
  11. May, A.; Goadsby, P.J. The Trigeminovascular System in Humans: Pathophysiologic Implications for Primary Headache Syndromes of the Neural Influences on the Cerebral Circulation. Br. J. Pharmacol. 1999, 19, 115–127. [Google Scholar] [CrossRef] [Green Version]
  12. Eftekhari, S.; Salvatore, C.A.; Johansson, S.; Chen, T.-B.; Zeng, Z.; Edvinsson, L. Localization of CGRP, CGRP receptor, PACAP and glutamate in trigeminal ganglion. Relation to the blood–brain barrier. Brain Res. 2015, 1600, 93–109. [Google Scholar] [CrossRef]
  13. Leao, A.A.P. Spreading depression of activity in the cerebral cortex. J. Neurophysiol. 1944, 7, 359–390. [Google Scholar] [CrossRef]
  14. Charles, A.C.; Baca, S.M. Cortical spreading depression and migraine. Nat. Rev. Neurol. 2013, 9, 637–644. [Google Scholar] [CrossRef] [PubMed]
  15. Hansen, J.M.; Baca, S.M.; VanValkenburgh, P.; Charles, A. Distinctive anatomical and physiological features of migraine aura revealed by 18 years of recording. Brain 2013, 136, 3589–3595. [Google Scholar] [CrossRef] [Green Version]
  16. Goadsby, P.J.; Edvinsson, L.; Ekman, R. Release of vasoactive peptides in the extracerebral circulation of hu-mans and the cat during activation of the trigeminovascular system. Ann. Neurol. 1988, 23, 193–196. [Google Scholar] [CrossRef] [PubMed]
  17. Bolay, H.; Reuter, U.; Dunn, A.K.; Huang, Z.; Boas, D.A.; Moskowitz, M.A. Intrinsic brain activity triggers trigeminal meningeal afferents in a migraine model. Nat. Med. 2002, 8, 136–142. [Google Scholar] [CrossRef]
  18. Zhang, X.; Levy, D.; Kainz, V.; Noseda, R.; Jakubowski, M.; Burstein, R. Activation of central trigeminovascular neurons by cortical spreading depression. Ann. Neurol. 2010, 69, 855–865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Gargus, J.J. Genetic Calcium Signaling Abnormalities in the Central Nervous System: Seizures, Migraine, and Autism. Ann. N. Y. Acad. Sci. 2008, 1151, 133–156. [Google Scholar] [CrossRef] [PubMed]
  20. Ducros, A.; Tournier-Lasserve, E.; Bousser, M.-G. The genetics of migraine. Lancet Neurol. 2002, 1, 285–293. [Google Scholar] [CrossRef]
  21. Montagna, P. Migraine: A genetic disease? Neurol. Sci. 2008, 29, 47–51. [Google Scholar] [CrossRef]
  22. Mulley, J.C.; Scheffer, I.E.; Petrou, S.; Dibbens, L.M.; Berkovic, S.F.; Harkin, L.A. SCN1A mutations and epilepsy. Hum. Mutat. 2005, 25, 535–542. [Google Scholar] [CrossRef]
  23. Talavera, K.; Startek, J.B.; Alvarez-Collazo, J.; Boonen, B.; Alpizar, Y.A.; Sanchez, A.; Naert, R.; Nilius, B. Mammalian transient receptor potential trpa1 channels: From structure to disease. Physiol. Rev. 2020, 100, 725–803. [Google Scholar] [CrossRef]
  24. Piekut, T.; Wong, Y.Y.; Senatore, A.; Walker, E.S.; Smith, C.L.; Gauberg, J.; Harracksingh, A.N.; Lowden, C.; Novogradac, B.B.; Cheng, H.-Y.M.; et al. Early metazoan origin and multiple losses of a novel clade of rim presynaptic calcium channel scaffolding protein homologs. Genome Biol. Evol. 2020, 12, 1217–1239. [Google Scholar] [CrossRef]
  25. Pietrobon, D. CaV2.1 channelopathies. Pflügers Archiv Eur. J. Physiol. 2010, 460, 375–393. [Google Scholar] [CrossRef]
  26. Wild, V.; Messlinger, K.; Fischer, M.J. Hydrogen sulfide determines HNO-induced stimulation of trigeminal afferents. Neurosci. Lett. 2015, 602, 104–109. [Google Scholar] [CrossRef]
  27. Kunkler, P.E.; Ballardl, C.J.; Oxford, G.S.; Hurley, J.H. TRPA1 receptors mediate environmental irritant-induced meningeal vasodilatation. Pain 2011, 152, 38–44. [Google Scholar] [CrossRef] [Green Version]
  28. Nanou, E.; Catterall, W.A. Calcium channels, synaptic plasticity, and neuropsychiatric disease. Neuron 2018, 98, 466–481. [Google Scholar] [CrossRef] [Green Version]
  29. Catterall, W.A. Voltage-gated calcium channels. Handb. Cell Sign. 2010, 3, 897–909. [Google Scholar] [CrossRef] [Green Version]
  30. Amrutkar, D.; Ploug, K.; Olesen, J.; Jansen-Olesen, I. Role for voltage gated calcium channels in calcitonin gene-related peptide release in the rat trigeminovascular system. Neuroscience 2011, 172, 510–517. [Google Scholar] [CrossRef] [PubMed]
  31. Yin, P.; Anttila, V.; Siewert, K.M.; Palotie, A.; Smith, G.D.; Voight, B.F. Serum calcium and risk of migraine: A Mendelian randomization study. Hum. Mol. Genet. 2016, 26, 820–828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Rau, J.C.; Dodick, D.W. Other preventive anti-migraine treatments: Ace inhibitors, arbs, calcium channel blockers, serotonin antagonists, and nmda receptor antagonists. Curr. Treat. Options Neurol. 2019, 21, 17. [Google Scholar] [CrossRef] [PubMed]
  33. Edvinsson, L. Functional role of perivascular peptides in the control of cerebral circulation. Trends Neurosci. 1985, 8, 126–131. [Google Scholar] [CrossRef]
  34. Eftekhari, S.; Salvatore, C.; Calamari, A.; Kane, S.; Tajti, J.; Edvinsson, L. Differential distribution of calcitonin gene-related peptide and its receptor components in the human trigeminal ganglion. Neuroscience 2010, 169, 683–696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Mulderry, P.K.; Ghatei, M.A.; Spokes, R.A.; Jones, P.M.; Pierson, A.M.; Hamid, Q.A.; Kanse, S.; Amara, S.G.; Burrin, J.M.; Legon, S.; et al. Differential expression of a-CGRP and b-CGRP by primary sensory neurons and enteric au-tonomic neurons of the rat. Neuroscience 1988, 25, 195–205. [Google Scholar]
  36. Thomsen, L.; Kruuse, C.; Iversen, H.K.; Olesen, J. A nitric oxide donor (nitroglycerin) triggers genuine migraine attacks. Eur. J. Neurol. 1994, 1, 73–80. [Google Scholar] [CrossRef]
  37. Weidner, C.; Klede, M.; Rukwied, R.; Lischetzki, G.; Neisius, U.; Schmelz, M.; Skov, P.S.; Petersen, L.J. Acute effects of substance p and calcitonin gene-related peptide in human skin—A microdialysis study. J. Investig. Dermatol. 2000, 115, 1015–1020. [Google Scholar] [CrossRef]
  38. Guo, S.; Vollesen, A.L.; Olesen, J.; Ashina, M. Premonitory and nonheadache symptoms induced by CGRP and PACAP38 in patients with migraine. Pain 2016, 157, 2773–2781. [Google Scholar] [CrossRef] [Green Version]
  39. Goadsby, P.J.; Edvinsson, L.; Ekman, R. Vasoactive peptide release in the extracerebral circulation of humans during migraine headache. Ann. Neurol. 1990, 28, 183–187. [Google Scholar] [CrossRef]
  40. Goadsby, P.J.; Edvinsson, L. The trigeminovascular system and migraine: Studies characterizing cerebrovascular and neuropeptide changes seen in humans and cats. Ann. Neurol. 1993, 33, 48–56. [Google Scholar] [CrossRef] [PubMed]
  41. Cernuda-Morollón, E.; Larrosa, D.; Ramón, C.; Vega, J.; Martínez-Camblor, P.; Pascual, J. Interictal increase of CGRP levels in peripheral blood as a biomarker for chronic migraine. Neurology 2013, 81, 1191–1196. [Google Scholar] [CrossRef]
  42. Juhasz, G.; Zsombok, T.; Jakab, B.; Nemeth, J.; Szolcsanyi, J.; Bagdy, G. Sumatriptan causes parallel decrease in plasma calcitonin gene-related peptide (Cgrp) Concentration and migraine headache during nitroglycerin induced migraine attack. Cephalalgia 2005, 25, 179–183. [Google Scholar] [CrossRef]
  43. Cady, R.K.; Vause, C.V.; Ho, T.W.; Bigal, M.E.; Durham, P.L. Elevated saliva calcitonin gene-related peptide levels during acute migraine predict therapeutic response to rizatriptan. Headache: J. Head Face Pain 2009, 49, 1258–1266. [Google Scholar] [CrossRef]
  44. Ho, T.W.; Olesen, J.; Dodick, D.W.; Kost, J.; Lines, C.; Ferrari, M.D. Antimigraine efficacy of telcagepant based on patient’s historical triptan response. Headache 2011, 51, 64–72. [Google Scholar] [CrossRef] [PubMed]
  45. Hoffmann, J.; Goadsby, P.J. New agents for acute treatment of migraine: CGRP receptor antagonists, iNOS inhibitors. Curr. Treat. Options Neurol. 2011, 14, 50–59. [Google Scholar] [CrossRef] [PubMed]
  46. Olesen, J.; Diener, H.-C.; Husstedt, I.W.; Goadsby, P.J.; Hall, D.; Meier, U.; Pollentier, S.; Lesko, L.M. Calcitonin gene–related peptide receptor antagonist bibn 4096 BS for the acute treatment of migraine. N. Engl. J. Med. 2004, 350, 1104–1110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Silberstein, S.D. Emerging target-based paradigms to prevent and treat migraine. Clin. Pharmacol. Ther. 2013, 93, 78–85. [Google Scholar] [CrossRef] [PubMed]
  48. Voss, T.; Lipton, R.B.; Dodick, D.W.; Dupre, N.; Ge, J.Y.; Bachman, R.; Assaid, C.; Aurora, S.K.; Michelson, D.A. phase IIb randomized, double-blind, placebo-controlled trial of ubrogepant for the acute treatment of mi-graine. Cephalalgia 2016, 36, 887–898. [Google Scholar] [CrossRef]
  49. Diener, H.-C.; Barbanti, P.; Dahlöf, C.; Reuter, U.; Habeck, J.; Podhorna, J. BI 44370 TA, an oral CGRP antagonist for the treatment of acute migraine attacks: Results from a phase II study. Cephalalgia 2011, 31, 573–584. [Google Scholar] [CrossRef]
  50. Marcus, R.; Goadsby, P.J.; Dodick, D.; Stock, D.; Manos, G.; Fischer, T.Z. BMS-927711 for the acute treatment of migraine: A double-blind, randomized, placebo controlled, dose-ranging trial. Cephalalgia 2014, 34, 114–125. [Google Scholar] [CrossRef]
  51. Stauffer, V.L.; Dodick, D.W.; Zhang, Q.; Carter, J.N.; Ailani, J.; Conley, R.R. Evaluation of galcanezumab for the prevention of episodic migraine: The EVOLVE-1 randomized clinical trial. JAMA Neurol. 2018, 75, 1080–1088. [Google Scholar] [CrossRef]
  52. Skljarevski, V.; Matharu, M.; Millen, A.B.; Ossipov, M.H.; Kim, B.-K.; Yang, J.Y. Efficacy and safety of galcanezumab for the prevention of episodic migraine: Results of the EVOLVE-2 Phase 3 randomized controlled clinical trial. Cephalalgia 2018, 38, 1442–1454. [Google Scholar] [CrossRef]
  53. Dodick, D.W.; Goadsby, P.J.; Spierings, E.L.H.; Scherer, J.C.; Sweeney, S.P.; Grayzel, D.S. Safety and efficacy of LY2951742, a monoclonal antibody to calcitonin gene-related peptide, for the prevention of migraine: A phase 2, randomised, double-blind, placebo-controlled study. Lancet Neurol. 2014, 13, 885–892. [Google Scholar] [CrossRef]
  54. Detke, H.C.; Goadsby, P.J.; Wang, S.; Friedman, D.I.; Selzler, K.J.; Aurora, S.K. Galcanezumab in chronic migraine: The randomized, double-blind, placebo-controlled REGAIN study. Neurology 2018, 91, e2211–e2221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Bigal, E.M.; Dodick, D.W.; Rapoport, A.M.; Silberstein, S.D.; Ma, Y.; Yang, R.; Loupe, P.S.; Burstein, R.; Newman, L.C.; Lipton, R.B. Safety, tolerability, and efficacy of TEV-48125 for preventive treatment of high-frequency episodic migraine: A multicentre, randomised, double-blind, placebo-controlled, phase 2b study. Lancet Neurol. 2015, 14, 1081–1090. [Google Scholar] [CrossRef]
  56. Dodick, D.W.; Goadsby, P.J.; Silberstein, S.D.; Lipton, R.B.; Olesen, J.; Ashina, M.; Wilks, K.; Kudrow, D.; Kroll, R.; Kohrman, B.; et al. Safety and efficacy of ALD403, an antibody to calcitonin gene-related peptide, for the prevention of frequent episodic migraine: A randomised, double-blind, placebo-controlled, exploratory phase 2 trial. Lancet Neurol. 2014, 13, 1100–1107. [Google Scholar] [CrossRef]
  57. Sun, H.; Dodick, D.W.; Silberstein, S.; Goadsby, P.J.; Reuter, U.; Ashina, M.; Saper, J.; Cady, R.; Chon, Y.; Dietrich, J.; et al. Safety and efficacy of AMG 334 for prevention of episodic migraine: A randomised, double-blind, placebo-controlled, phase 2 trial. Lancet Neurol. 2016, 15, 382–390. [Google Scholar] [CrossRef]
  58. Tepper, S.; Ashina, M.; Reuter, U.; Brandes, J.L.; Doležil, D.; Silberstein, S.; Winner, P.; Leonardi, D.; Mikol, D.; Lenz, R. Safety and efficacy of erenumab for preventive treatment of chronic migraine: A randomised, double-blind, placebo-controlled phase 2 trial. Lancet Neurol. 2017, 16, 425–434. [Google Scholar] [CrossRef]
  59. Messlinger, K.; Balcziak, L.K.; Russo, A.F. Cross-talk signaling in the trigeminal ganglion: Role of neuropeptides and other mediators. J. Neural Transm. 2020, 127, 431–444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Dieterle, A.; Fischer, M.J.; Link, A.S.; Neuhuber, W.L.; Messlinger, K. Increase in CGRP- and nNOS-immunoreactive neurons in the rat trigeminal ganglion after infusion of an NO donor. Cephalalgia 2011, 31, 31–42. [Google Scholar] [CrossRef]
  61. Iyengar, S.; Johnson, K.W.; Ossipov, M.H.; Aurora, S.K. CGRP and the trigeminal system in migraine. Headache J. Head Face Pain 2019, 59, 659–681. [Google Scholar] [CrossRef] [Green Version]
  62. Capuano, A.; De Corato, A.; Lisi, L.; Tringali, G.; Navarra, P.; Russo, C.D. Proinflammatory-activated trigeminal satellite cells promote neuronal sensitization: Relevance for migraine pathology. Mol. Pain 2009, 5, 43. [Google Scholar] [CrossRef] [Green Version]
  63. De Corato, A.; Lisi, L.; Capuano, A.; Tringali, G.; Tramutola, A.; Navarra, P.; Russo, C.D. Trigeminal satellite cells express functional calcitonin gene-related peptide receptors, whose activation enhances interleukin-1beta pro-inflammatory effects. J. Neuroimmunol. 2011, 237, 39–46. [Google Scholar] [CrossRef]
  64. Thalakoti, S.; Patil, V.V.; Damodaram, S.; Vause, C.V.; Langford, L.E.; Freeman, S.E.; Durham, P.L. Neuron-glia signaling in trigeminal ganglion: Implications for migraine pathology. Headache 2007, 47, 1008–1023. [Google Scholar] [CrossRef] [Green Version]
  65. Bowen, E.J.; Schmidt, T.W.; Firm, C.S.; Russo, A.F.; Durham, P.L. Tumor necrosis factor-alpha stimulation of calcitonin gene-related peptide expression and secretion from rat trigeminal ganglion neurons. J. Neurochem. 2005, 96, 65–77. [Google Scholar] [CrossRef] [Green Version]
  66. Kwok, Y.H.; Swift, J.E.; Gazerani, P.; Rolan, P. A double-blind, randomized, placebo-controlled pilot trial to determine the efficacy and safety of ibudilast, a potential glial attenuator, in chronic migraine. J. Pain Res. 2016, 9, 899–907. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Durham, Z.L.; Hawkins, J.L.; Durham, P.L. Tumor necrosis factor-Alpha stimulates cytokine expression and transient sensitization of trigeminal nociceptive neurons. Arch. Oral Biol. 2017, 75, 100–106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Durham, P.L. Diverse physiological roles of calcitonin gene-related peptide in migraine pathology: Modulation of neuronal-glial-immune cells to promote peripheral and central sensitization. Curr. Pain Headache Rep. 2016, 20, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Cady, R.J.; Glenn, J.R.; Smith, K.M.; Durham, P.L. Calcitonin gene-related peptide promotes cellular changes in trigeminal neurons and glia implicated in peripheral and central sensitization. Mol. Pain. 2011, 7. [Google Scholar] [CrossRef] [Green Version]
  70. Cornelison, L.E.; Hawkins, J.L.; Durham, P.L. Elevated levels of calcitonin gene-related peptide in upper spinal cord promotes sensitization of primary trigeminal nociceptive neurons. Neuroscience 2016, 339, 491–501. [Google Scholar] [CrossRef] [Green Version]
  71. Weir, A.G.; Cader, M.Z. New directions in migraine. BMC Med. 2011, 9, 116. [Google Scholar] [CrossRef] [Green Version]
  72. Spraya, D.C.; Hananib, M. Gap junctions, pannexins and pain. Neurosci. Lett. 2019, 695, 46–52. [Google Scholar] [CrossRef] [PubMed]
  73. Hansen, J.M.; Thomsen, L.L.; Olesen, J.; Ashina, M. Calcitonin gene-related peptide does not cause the familial hemiplegic migraine phenotype. Neurology 2008, 71, 841–847. [Google Scholar] [CrossRef]
  74. Ptáček, L.J. Channelopathies: Ion channel disorders of muscle as a paradigm for paroxysmal disorders of the nervous system. Neuromuscul. Disord. 1997, 7, 250–255. [Google Scholar] [CrossRef]
  75. Ducros, A.; Denier, C.; Joutel, A.; Cecillon, M.; Lescoat, C.; Vahedi, K.; Darcel, F.; Vicaut, E.; Bousser, M.-G.; Tournier-Lasserve, E. The clinical spectrum of familial hemiplegic migraine associated with mutations in a neuronal calcium channel. N. Engl. J. Med. 2001, 345, 17–24. [Google Scholar] [CrossRef]
  76. Goadsby, P.J.; Holland, P.R.; Martins-Oliveira, M.; Hoffmann, J.; Schankin, C.; Akerman, S. Pathophysiology of migraine: A disorder of sensory processing. Physiol. Rev. 2017, 97, 553–622. [Google Scholar] [CrossRef]
  77. Ophoff, A.R.; Terwindt, G.M.; Haan, J.; Lindhout, D.; van Ommen, G.-J.B.; Hofker, M.H.; Ferrari, M.D.; Frants, R.R.; Vergouwe, M.N.; van Eijk, R.; 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. [Google Scholar] [CrossRef] [Green Version]
  78. De Fusco, M.; Marconi, R.; Silvestri, L.; Atorino, L.; Rampoldi, L.; Morgante, L.; Ballabio, A.; Aridon, P.; Casari, G. Haploinsufficiency of ATP1A2 encoding the Na+/K+ pump α2 subunit associated with familial hemiplegic migraine type 2. Nat. Genet. 2003, 33, 192–196. [Google Scholar] [CrossRef] [PubMed]
  79. Dichgans, M.; Freilinger, T.; Eckstein, G.; Babini, E.; Lorenz-Depiereux, B.; Biskup, S.; Ferrari, M.D.; Herzog, J.; Maagdenberg, A.M.J.M.V.D.; Pusch, M.; et al. Mutation in the neuronal voltage-gated sodium channel SCN1A in familial hemiplegic migraine. Lancet 2005, 366, 371–377. [Google Scholar] [CrossRef]
  80. Leo, L.; Gherardini, L.; Barone, V.; De Fusco, M.; Pietrobon, D.; Pizzorusso, T.; Casari, G. Increased susceptibility to cortical spreading depression in the mouse model of familial hemiplegic migraine type 2. PLoS Genet. 2011, 7, e1002129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Huang, Y.; Xiao, H.; Qin, X.; Nong, Y.; Zou, D.; Wu, Y. The genetic relationship between epilepsy and hemiplegic migraine. Neuropsychiatr. Dis. Treat. 2017, 13, 1175–1179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Urbano, F.J.; Pagani, M.R.; Uchitel, O.D. Calcium channels, neuromuscular synaptic transmission and neurological diseases. J. Neuroimmunol. 2008, 201-202, 136–144. [Google Scholar] [CrossRef] [PubMed]
  83. Westenbroek, R.; Sakurai, T.; Elliott, E.; Hell, J.; Starr, T.; Snutch, T.; Catterall, W. Immunochemical identification and subcellular distribution of the alpha 1A subunits of brain calcium channels. J. Neurosci. 1995, 15, 6403–6418. [Google Scholar] [CrossRef] [Green Version]
  84. Akerman, S.; Williamson, D.J.; Goadsby, P.J. Voltage-dependent calcium channels are involved in neurogenic dural vasodilatation via a presynaptic transmitter release mechanism: Voltage-dependent calcium channels. Br. J. Pharmacol. 2003, 140, 558–566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Pietrobon, D. Function and dysfunction of synaptic calcium channels: Insights from mouse models. Curr. Opin. Neurobiol. 2005, 15, 257–265. [Google Scholar] [CrossRef]
  86. Santafé, M.M.; Salon, I.; Garcia, N.; Lanuza, M.A.; Uchitel, O.D.; Tomàs, J. Modulation of ACh release by presynaptic muscarinic autoreceptors in the neuromuscular junction of the newborn and adult rat: Muscarinic autoreceptors in neuromuscular transmission. Eur. J. Neurosci. 2003, 17, 119–127. [Google Scholar] [CrossRef] [PubMed]
  87. Santafé, M.M.; Lanuza, M.A.; Garcia, N.; Tomàs, J. Muscarinic autoreceptors modulate transmitter release through protein kinase C and protein kinase A in the rat motor nerve terminal. Eur. J. Neurosci. 2006, 23, 2048–2056. [Google Scholar] [CrossRef] [PubMed]
  88. Chang, S.Y.; Yong, T.F.; Yu, C.Y.; Liang, M.C.; Pletnikova, O.; Troncoso, J.C.; Burgunder, J.-M.; Soong, T.W. Age and gender-dependent alternative splicing of P/Q-type calcium channel EF-hand. Neuroscience 2007, 145, 1026–1036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Travaglini, L.; Nardella, M.; Bellacchio, E.; D’Amico, A.; Capuano, A.; Frusciante, R.; Di Capua, M.; Cusmai, R.; Barresi, S.; Morlino, S.; et al. Missense mutations of CACNA1A are a frequent cause of autosomal dominant nonprogressive congenital ataxia. Eur. J. Paediatr. Neurol. 2017, 21, 450–456. [Google Scholar] [CrossRef] [PubMed]
  90. Kors, E.E.; Terwindt, G.M.; Vermeulen, F.L.M.G.; Fitzsimons, R.B.; Jardine, P.E.; Heywood, P.; Love, S.; Maagdenberg, A.M.J.M.V.D.; Haan, J.; Frants, R.R.; et al. Delayed cerebral edema and fatal coma after minor head trauma: Role of the CACNA1A calcium channel subunit gene and relationship with familial hemiplegic migraine. Ann. Neurol. 2001, 49, 753–760. [Google Scholar] [CrossRef] [PubMed]
  91. Chan, Y.-C.; Burgunder, J.-M.; Wilder-Smith, E.; Chew, S.-E.; Lam-Mok-Sing, K.M.; Sharma, V.; Ong, B.K. Electroencephalographic changes and seizures in familial hemiplegic migraine patients with the CACNA1A gene S218L mutation. J. Clin. Neurosci. 2008, 15, 891–894. [Google Scholar] [CrossRef]
  92. Freilinger, T.; Ackl, N.; Ebert, A.; Schmidt, C.; Rautenstrauss, B.; Dichgans, M.; Danek, A. A novel mutation in CACNA1A associated with hemiplegic migraine, cerebellar dysfunction and late-onset cognitive decline. J. Neurol. Sci. 2011, 300, 160–163. [Google Scholar] [CrossRef] [PubMed]
  93. Battistini, S.; Stenirri, S.; Piatti, M.; Gelfi, C.; Righetti, P.G.; Rocchi, R.; Giannini, F.; Battistini, N.; Guazzi, G.C.; Ferrari, M.; et al. A new CACNA1A gene mutation in acetazolamide-responsive familial hemiplegic migraine and ataxia. Neurology 1999, 53, 38. [Google Scholar] [CrossRef] [PubMed]
  94. Terwindt, G.; Kors, E.; Haan, J.; Vermeulen, F.; Maagdenberg, A.V.D.; Frants, R.; Ferrari, M. Mutation analysis of the CACNA1A calcium channel subunit gene in 27 patients with sporadic hemiplegic migraine. Arch. Neurol. 2002, 59, 1016–1018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Alonso, I.; Barros, J.; Tuna, A.; Coelho, J.; Sequeiros, J.; Silveira, I.; Coutinho, P. Phenotypes of spinocerebellar ataxia type 6 and familial hemiplegic migraine caused by a unique CACNA1a missense mutation in patients from a large family. Arch. Neurol. 2003, 60, 610–614. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. de Vries, B.; Freilinger, T.; Vanmolkot, K.R.; Koenderink, J.B.; Stam, A.H.; Terwindt, G.M.; Babini, E.; van den Boogerd, E.H.; van den Heuvel, J.J.; Frants, R.R.; et al. Systematic analysis of three FHM genes in 39 sporadic patients with hemiplegic migraine. Neurology 2007, 69, 2170–2176. [Google Scholar] [CrossRef] [Green Version]
  97. Friend, K.L.; Crimmins, D.; Phan, T.G.; Sue, C.M.; Colley, A.; Fung, V.S.; Morris, J.G.; Sutherland, G.R.; Richards, R.I. Detection of a novel missense mutation and second recurrent mutation in the CACNA1A gene in individuals with EA-2 and FHM. Hum. Genet. 1999, 105, 261–265. [Google Scholar] [CrossRef] [PubMed]
  98. Ducros, A.; Denier, C.; Hirsch, E.; Chedru, F.; Bisgård, C.; Lucotte, G.; Després, P.; Billard, C.; Barthez, M.; Ponsot, G.; et al. Recurrence of the T666M calcium channel CACNA1a gene mutation in familial hemiplegic migraine with progressive cerebellar ataxia. Am. J. Hum. Genet. 1999, 64, 89–98. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Kors, E.E.; Haan, J.; Giffin, N.J.; Pazdera, L.; Schnittger, C.; Lennox, G.G.; Terwindt, G.M.; Vermeulen, F.L.; Van den Maagdenberg, A.M.; Frants, R.R.; et al. Expanding the phenotypic spectrum of the CACNA1A gene T666M mutation: A description of 5 families with familial hemiplegic migraine. Arch. Neurol. 2003, 60, 684–688. [Google Scholar] [CrossRef] [Green Version]
  100. Stam, A.; Vanmolkot, K.; Kremer, H.; Gärtner, J.; Brown, J.; Leshinsky-Silver, E.; Gilad, R.; Kors, E.; Frankhuizen, W.; Ginjaar, H.; et al. CACNA1A R1347Q: A frequent recurrent mutation in hemiplegic migraine. Clin. Genet. 2008, 74, 481–485. [Google Scholar] [CrossRef] [PubMed]
  101. Vahedi, K.; Denier, C.; Ducros, A.; Bousson, V.; Levy, C.; Chabriat, H.; Haguenau, M.; Tournier-Lasserve, E.; Bousser, M.G. CACNA1A gene de novo mutation causing hemiplegic migraine, coma, and cerebellar atrophy. Neurology 2000, 55, 1040–1042. [Google Scholar] [CrossRef]
  102. Carrera, P.; Piatti, M.; Stenirri, S.; Grimaldi, L.M.E.; Marchioni, E.; Curcio, M.; Righetti, P.G.; Ferrari, M.; Gelfi, C. Genetic heterogeneity in Italian families with familial hemiplegic migraine. Neurology 1999, 53, 26. [Google Scholar] [CrossRef]
  103. Kors, E.E.; Melberg, A.; Vanmolkot, K.R.J.; Kumlien, E.; Haan, J.; Raininko, R.; Flink, R.; Ginjaar, H.B.; Frants, R.R.; Ferrari, M.D.; et al. Childhood epilepsy, familial hemiplegic migraine, cerebellar ataxia, and a new CACNA1A mutation. Neurology 2004, 63, 1136–1137. [Google Scholar] [CrossRef]
  104. Labrum, R.W.; Rajakulendran, S.; Graves, T.D.; Eunson, L.H.; Bevan, R.; Sweeney, M.G.; Hammans, S.R.; Tubridy, N.; Britton, T.; Carr, L.J.; et al. Large scale calcium channel gene rearrangements in episodic ataxia and hemiplegic migraine: Implications for diagnostic testing. J. Med Genet. 2009, 46, 786–791. [Google Scholar] [CrossRef] [Green Version]
  105. Kraus, R.L.; Sinnegger, M.J.; Glossmann, H.; Hering, S.; Striessnig, J. Familial hemiplegic migraine mutations change α 1A Ca 2+ Channel Kinetics. J. Biol. Chem. 1998, 273, 5586–5590. [Google Scholar] [CrossRef] [Green Version]
  106. Hans, M.; Luvisetto, S.; Williams, M.E.; Spagnolo, M.; Urrutia, A.; Tottene, A.; Brust, P.F.; Johnson, E.C.; Harpold, M.M.; Stauderman, K.A.; et al. Functional consequences of mutations in the human α 1a calcium channel subunit linked to familial hemiplegic migraine. J. Neurosci. 1999, 19, 1610–1619. [Google Scholar] [CrossRef] [Green Version]
  107. Stenirri, S.; Carrera, P.; Kraus, R.L.; Sinnegger, M.J.; Koschak, A.; Glossmann, H.; Striessnig, J. Three new familial hemiplegic migraine mutants affect P/Q-type CA2+ channel kinetics. J. Biol. Chem. 2000, 275, 9239–9243. [Google Scholar] [CrossRef] [Green Version]
  108. Tottene, A.; Fellin, T.; Pagnutti, S.; Luvisetto, S.; Striessnig, J.; Fletcher, C.; Pietrobon, D. Familial hemiplegic mi-graine mutations increase Ca2+ influx through single human CaV2.1 channels and decrease maximal CaV2.1 current density in neurons. Proc. Natl. Acad. Sci. USA 2002, 99, 13284–13289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Melliti, K.; Grabner, M.; Seabrook, G.R. The familial hemiplegic migraine mutation R192q reduces G-protein-mediated inhibition of p/q-type (Ca v 2.1) calcium channels expressed in human embryonic kidney cells. J. Physiol. 2003, 546, 337–347. [Google Scholar] [CrossRef] [PubMed]
  110. Müllner, C.; Broos, L.A.; van den Maagdenberg, A.M.; Striessnig, J. Familial hemiplegic migraine type 1 mutations k1336e, w1684r, and v1696i alter CA V 2.1 CA 2+ channel gating: Evidence for β-subunit isoform-specific effects. J. Biol. Chem. 2004, 279, 51844–51850. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Tottene, A.; Pivotto, F.; Fellin, T.; Cesetti, T.; van den Maagdenberg, A.M.; Pietrobon, D. Specific kinetic alterations of human CA V 2.1 calcium channels produced by mutation s218l causing familial hemiplegic migraine and delayed cerebral edema and coma after minor head trauma. J. Biol. Chem. 2005, 280, 17678–17686. [Google Scholar] [CrossRef] [Green Version]
  112. Barrett, C.F.; Cao, Y.-Q.; Tsien, R.W. Gating deficiency in a familial hemiplegic migraine type 1 mutant P/Q-type calcium channel. J. Biol. Chem. 2005, 280, 24064–24071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Weiss, N.; Sandoval, A.; Felix, R.; Van den Maagdenberg, A.; De Waard, M. The S218L familial hemiplegic mi-graine mutation promotes deinhibition of Cav2.1 calcium channels during direct G-protein regulation. Pflüg Arch. 2008, 457, 315–326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Adams, P.J.; Garcia, E.; David, L.S.; Mulatz, K.J.; Spacey, S.D.; Snutch, T.P. Ca V 2.1 P/Q-type calcium channel alter-native splicing affects the functional impact of familial hemiplegic migraine mutations: Implications for calcium channelopathies. Channels 2009, 3, 110–121. [Google Scholar] [CrossRef] [Green Version]
  115. Eikermann-Haerter, K.; Dileköz, E.; Kudo, C.; Savitz, S.I.; Waeber, C.; Baum, M.J.; Ferrari, M.D.; van den Maagden-berg, A.M.; Moskowitz, M.A.; Ayata, C. Genetic and hormonal factors modulate spreading depression and tran-sient hemiparesis in mouse models of familial hemiplegic migraine type 1. J. Clin. Invest. 2009, 119, 99–109. [Google Scholar]
  116. Serra, S.A.; Fernàndez-Castillo, N.; Macaya, A.; Cormand, B.; Valverde, M.A.; Fernández-Fernández, J.M. The hem-iplegic migraine-associated Y1245C mutation in CACNA1A results in a gain of channel function due to its effect on the voltage sensor and G-protein-mediated inhibition. Pflüg Arch. 2009, 458, 489–502. [Google Scholar] [CrossRef] [PubMed]
  117. Khennouf, L.; Gesslein, B.; Lind, B.L.; van den Maagdenberg, A.M.; Lauritzen, M. Activity-dependent calcium, oxygen, and vascular responses in a mouse model of familial hemiplegic migraine type 1: Calcium, oxygen and vascular responses in migraine. Ann. Neurol. 2016, 80, 219–232. [Google Scholar] [CrossRef]
  118. Cao, Y.Q.; Piedras-Rentería, E.S.; Smith, G.B.; Chen, G.; Harata, N.C.; Tsien, R.W. Presynaptic Ca2+ Channels Com-pete for Channel Type-Preferring Slots in Altered Neurotransmission Arising from Ca2+ Channelopathy. Neuron 2004, 43, 387–400. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Cao, Y.-Q.; Tsien, R.W. Effects of familial hemiplegic migraine type 1 mutations on neuronal P/Q-type Ca2+ channel activity and inhibitory synaptic transmission. Proc. Natl. Acad. Sci. USA 2005, 102, 2590–2595. [Google Scholar] [CrossRef] [Green Version]
  120. Tottene, A.; Conti, R.; Fabbro, A.; Vecchia, D.; Shapovalova, M.; Santello, M.; Maagdenberg, A.M.V.D.; Ferrari, M.D.; Pietrobon, D. Enhanced excitatory transmission at cortical synapses as the basis for facilitated spreading depression in CAV2.1 knockin migraine mice. Neuron 2009, 61, 762–773. [Google Scholar] [CrossRef] [PubMed]
  121. Haan, J.; Kors, E.E.; Vanmolkot, K.R.; van den Maagdenberg, A.M.; Frants, R.R.; Ferrari, M.D. Migraine genetics: An up-date. Curr. Pain Headache Rep. 2005, 9, 213–220. [Google Scholar] [CrossRef]
  122. van den Maagdenberg, A.M.; Pietrobon, D.; Pizzorusso, T.; Kaja, S.; Broos, L.A.; Cesetti, T.; van de Ven, R.C.; Tottene, A.; van der Kaa, J. A cacna1a knockin migraine mouse model with increased susceptibil-ity to cortical spreading depression. Neuron 2004, 41, 701–710. [Google Scholar] [CrossRef]
  123. Smitherman, T.A.; Burch, R.; Sheikh, H.; Loder, E. The prevalence, impact, and treatment of migraine and SE-vere headaches in the United States: A review of statistics from national surveillance studies. Headache 2013, 53, 427–436. [Google Scholar] [CrossRef]
  124. Antonaci, F.; Ghiotto, N.; Wu, S.; Pucci, E.; Costa, A. Recent advances in migraine therapy. SpringerPlus 2016, 5, 637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Diener, H.-C.; Dodick, D.W.; Goadsby, P.J.; Lipton, R.B.; Olesen, J.; Silberstein, S.D. Chronic migraine—classification, characteristics and treatment. Nat. Rev. Neurol. 2012, 8, 162–171. [Google Scholar] [CrossRef]
  126. Burton, W.N.; Landy, S.H.; Downs, K.E.; Runken, M.C. The impact of migraine and the effect of migraine treat-ment on workplace productivity in the united states and suggestions for future research. Mayo Clin. Proc. 2009, 84, 436–445. [Google Scholar] [CrossRef] [Green Version]
  127. Nimmrich, V.; Gross, G. P/Q-type calcium channel modulators: P/Q-type calcium channel blockers. Br. J. Pharmacol. 2012, 167, 741–759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Beswick, P. Progress in the discovery of Ca channel blockers for the treatment of pain. In Comprehensive Medicinal Chemistry III; Elsevier BV: Amsterdam, The Netherlands, 2017; pp. 65–130. [Google Scholar]
  129. Dussor, G. ASICs as therapeutic targets for migraine. Neuropharmacology 2015, 94, 64–71. [Google Scholar] [CrossRef] [Green Version]
  130. Mazzuca, M.; Heurteaux, C.; Alloui, A.; Diochot, S.; Baron, A.; Voilley, N.; Blondeau, N.; Escoubas, P.; Gélot, A.; Cupo, A.; et al. A tarantula peptide against pain via ASIC1a channels and opioid mechanisms. Nat. Neurosci. 2007, 10, 943–945. [Google Scholar] [CrossRef] [PubMed]
  131. 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. [Google Scholar] [CrossRef]
  132. Kowalska, M.; Prendecki, M.; Kozubski, W.; Lianeri, M.; Dorszewska, J. Molecular factors in migraine. Oncotarget 2016, 7, 50708–50718. [Google Scholar] [CrossRef] [Green Version]
  133. Holland, P.R.; Akerman, S.; Andreou, A.P.; Karsan, N.; Wemmie, J.A.; Goadsby, P.J. Acid-sensing ion channel 1: A novel therapeutic target for migraine with aura. Ann. Neurol. 2012, 72, 559–563. [Google Scholar] [CrossRef]
Figure 1. Cortical spreading depression. Changes in central nervous system flow that may induce a migraine attack may be associated with the occurrence of spreading cortical spreading depression (CSD). Oligemia (reduced vascular flow) begins in the occipital and parieto-occipital areas, then moves through the cerebral cortex and stops at the medial and lateral sulcus. In some patients, CSD may extend to the frontal lobes. This phenomenon seems to be responsible for the aura formation during a migraine headache.
Figure 1. Cortical spreading depression. Changes in central nervous system flow that may induce a migraine attack may be associated with the occurrence of spreading cortical spreading depression (CSD). Oligemia (reduced vascular flow) begins in the occipital and parieto-occipital areas, then moves through the cerebral cortex and stops at the medial and lateral sulcus. In some patients, CSD may extend to the frontal lobes. This phenomenon seems to be responsible for the aura formation during a migraine headache.
Ijms 22 02688 g001
Figure 2. Induction of migraine attack due to sensitization of trigeminal ganglia and disturbed glutamatergic release. TRPA1 receptors in sensory neurons in trigeminal ganglia (TG) activate under environmental irritants taken by inhalation, ingestion or in an unknown mechanism. Activation of the TRPA1 receptor stimulates the release of calcitonin gene-related peptide (CGRP) that may enter bloodstream during migraine attack. CGRP receptor is expressed by majority on the neurons forming myelinated TG A-fibers, as well as on smooth muscles of dura’s vasculature. CGRP receptor stimulation increases blood flow. Besides CGRP, activated TG neurons secrete nitric oxide (NO). These mediators stimulate further the surrounding glial cells to produce interleukin-1β (IL-1β) that in turn leads to increased activity of cyclooxygenase (COX), associated with production of proinflammatory prostaglandin E2 (PGE2). This phenomenon may be one of foundations of the TG neurons sensitization. Subsequently, CGRP released by TG neurons may also promote release of tumor necrosis factor-α (TNF-α) in the glial satellite cells. That may cause a positive feedback loop of further TG-neuronal synthesis and secretion of CGRP. Furthermore, the released TNF-α may itself sensitize TG neurons and inflict overproduction of various other proinflammatory cytokines. Activation of TRPA channels increase the intracellular calcium ion levels (Ca2+), similarly to L-type of Voltage Gated Calcium Channel (CaV2.1). Mutations in CACNA1A gene lead to enhanced Ca2+ currents. The enhanced Ca2+ influx lead to excitation-inhibition imbalance and enhanced glutamate release. The interaction between glutamate and pre- and postsynaptic glutamate NMDA receptors (NMDAr) may facilitate the cortical spreading depression (CSD)—believed as one of the causes of migraine attack. Ca2+ channel blockers affect the CaV2.1 channels and reduce excessive Ca2+ influx to the cells, normalizing glutamate release and thus may be also used in the treatment of migraine headaches.
Figure 2. Induction of migraine attack due to sensitization of trigeminal ganglia and disturbed glutamatergic release. TRPA1 receptors in sensory neurons in trigeminal ganglia (TG) activate under environmental irritants taken by inhalation, ingestion or in an unknown mechanism. Activation of the TRPA1 receptor stimulates the release of calcitonin gene-related peptide (CGRP) that may enter bloodstream during migraine attack. CGRP receptor is expressed by majority on the neurons forming myelinated TG A-fibers, as well as on smooth muscles of dura’s vasculature. CGRP receptor stimulation increases blood flow. Besides CGRP, activated TG neurons secrete nitric oxide (NO). These mediators stimulate further the surrounding glial cells to produce interleukin-1β (IL-1β) that in turn leads to increased activity of cyclooxygenase (COX), associated with production of proinflammatory prostaglandin E2 (PGE2). This phenomenon may be one of foundations of the TG neurons sensitization. Subsequently, CGRP released by TG neurons may also promote release of tumor necrosis factor-α (TNF-α) in the glial satellite cells. That may cause a positive feedback loop of further TG-neuronal synthesis and secretion of CGRP. Furthermore, the released TNF-α may itself sensitize TG neurons and inflict overproduction of various other proinflammatory cytokines. Activation of TRPA channels increase the intracellular calcium ion levels (Ca2+), similarly to L-type of Voltage Gated Calcium Channel (CaV2.1). Mutations in CACNA1A gene lead to enhanced Ca2+ currents. The enhanced Ca2+ influx lead to excitation-inhibition imbalance and enhanced glutamate release. The interaction between glutamate and pre- and postsynaptic glutamate NMDA receptors (NMDAr) may facilitate the cortical spreading depression (CSD)—believed as one of the causes of migraine attack. Ca2+ channel blockers affect the CaV2.1 channels and reduce excessive Ca2+ influx to the cells, normalizing glutamate release and thus may be also used in the treatment of migraine headaches.
Ijms 22 02688 g002
Figure 3. Locations of FHM1 mutations in the secondary structure of the calcium channel α2.1 subunit. 1—R192Q, 2—R195K, 3—S218L, 4—V581M, 5—R583Q, 6—T666M, 7—V714A, 8—D715E, 9—Y1246C, 10—K1336E, 11—R1347Q, 12—C1370Y, 13—Y1385C, 14—V1457L, 15—C1535S, 16—R1668W, 17—L1682P, 18—W1684R, 19—V1696I, 20—I1710T, 21—I1811L. I-IV- number of extracellular loop.
Figure 3. Locations of FHM1 mutations in the secondary structure of the calcium channel α2.1 subunit. 1—R192Q, 2—R195K, 3—S218L, 4—V581M, 5—R583Q, 6—T666M, 7—V714A, 8—D715E, 9—Y1246C, 10—K1336E, 11—R1347Q, 12—C1370Y, 13—Y1385C, 14—V1457L, 15—C1535S, 16—R1668W, 17—L1682P, 18—W1684R, 19—V1696I, 20—I1710T, 21—I1811L. I-IV- number of extracellular loop.
Ijms 22 02688 g003
Table 1. Clinical presentation of CACNA1 mutation in FHM.
Table 1. Clinical presentation of CACNA1 mutation in FHM.
MutationLocationPopulation Phenotypic SpectrumReferences
R192Q
rs121908211
exon 4Italian familyNo data Ophoff et al. [77]
R195K
rs121908222
exon 4French familyFHM without cerebellar signsDucros et al. [75]
S218L
rs121908225
exon 5British and Australian families,
Malaysian family
Minor head trauma–triggered delayed severe
cerebral edema and coma; childhood seizures
Kors et al. [90]
Chan et al. [91]
V581Mexon 13German familyFHM with cerebellar
dysfunction and late-onset cognitive decline
Freilinger et al. [92]
R583Q
rs121908217
exon 13Italian family,
Dutch patient,
Portuguese family,
Dutch patient
FHM with consciousness and fever lasting several days, late-onset cerebellar ataxia and cerebellar atrophy; symptoms triggered by minor head trauma; sporadic hemiplegic migraine without cerebellar signs, age at onset 13 yearsBattistini et al. [93]
Terwindt et al. [94]
Alonso et al. [95]
de Vries et al. [96]
T666M
rs121908212
exon 16American family,
Australian family,
French families,
Dutch patient,
Dutch families
Age at onset between 2 and 22 years; FHM with progressive cerebellar ataxia; sporadic hemiplegic migraine; progressive cognitive dysfunctionOphoff et al. [77]
Friend et al. [97]
Ducros et al. [98]
Terwindt et al. [94]
Kors et al. [99]
V714A
rs121908213
exon 17British familyAge at onset between 10 and 21 yearsOphoff et al. [77]
D715E
rs121908218
exon 17French familyFHM with progressive cerebellar ataxiaDucros et al. [98]
K1336Eexon 25French familyFHM without cerebellar signsDucros et al. [75]
R1347Q
rs121908230
exon 25Dutch familiesWide clinical spectrum ranging from (trauma triggered) hemiplegic migraine with and without ataxia, loss of consciousness and epilepsy, early age at onset (usually before the age of 3)Stam et al. [100]
Y1385C
rs121908219
exon 26French patientsFHM with cerebellar signs, coma, hyperthermia, meningeal signs, and partial seizure, Vahedi et al. [101]
Ducros et al. [75]
V1457L
rs121908237
exon 27Italian familyMean age at onset 34 years, various degrees of aphasia congruent with the hemispheric dominance, without cerebellar ataxia or comaCarrera et al. [102]
R1668Wexon 32French familyFHM with or without cerebellar signsDucros et al. [75]
W1684Rexon 32French familyFHM with cerebellar signsDucros et al. [75]
V1696Iexon 33French familyFHM without cerebellar signsDucros et al. [75]
I1710T
rs121909326
exon 33Dutch familyFHM with childhood-onset of cerebellar ataxia (SCA6), childhood complex partial and generalized tonic-clonic seizures that occurred independently of the FHM attacks Kors et al. [103]
I1811L
rs121908214
exon 36Dutch and American familiesOnly one family displayed cerebellar atrophy and in that family only some members were affectedOphoff et al. [77]
18.2-kb deletionexons 39–47Irish, Indian, and Danish patientsFHM with or without ataxia, no seizures, no family history Labrum et al. [104]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kowalska, M.; Prendecki, M.; Piekut, T.; Kozubski, W.; Dorszewska, J. Migraine: Calcium Channels and Glia. Int. J. Mol. Sci. 2021, 22, 2688. https://doi.org/10.3390/ijms22052688

AMA Style

Kowalska M, Prendecki M, Piekut T, Kozubski W, Dorszewska J. Migraine: Calcium Channels and Glia. International Journal of Molecular Sciences. 2021; 22(5):2688. https://doi.org/10.3390/ijms22052688

Chicago/Turabian Style

Kowalska, Marta, Michał Prendecki, Thomas Piekut, Wojciech Kozubski, and Jolanta Dorszewska. 2021. "Migraine: Calcium Channels and Glia" International Journal of Molecular Sciences 22, no. 5: 2688. https://doi.org/10.3390/ijms22052688

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

Article Metrics

Back to TopTop