Skip to main content

Genetic mouse models of migraine

Abstract

Mouse models of rare monogenic forms of migraine provide a unique experimental system to study the cellular and circuit mechanisms of the primary brain dysfunctions causing a migraine disorder. Here, we discuss the migraine-relevant phenotypes and the migraine-relevant functional alterations in the brain of five genetic mouse models of migraine, four of which carry mutations derived from patients with familial hemiplegic migraine (FHM) and the fifth carry a mutation from patients with both phenotypically normal MA and familial advanced sleep phase syndrome (FASPS). We focus on the latter mouse model, in which a ubiquitous serine-threonine kinase is mutated, and on two mouse models of pure FHM, in which a voltage-gated calcium channel controlling neurotransmitter release at most brain synapses and a Na/K ATPase that is expressed mainly in astrocytes in the adult brain are mutated, respectively. First, we describe the behavioral phenotypes of the genetic animal models and review the evidence that an increased susceptibility to experimentally induced cortical spreading depression (CSD) is a key migraine-relevant phenotype common to the five models. Second, we review the synaptic alterations in the cerebral cortex of the genetic models of migraine and discuss the mechanisms underlying their increased susceptibility to CSD. Third, we review the alterations in the trigeminovascular pain pathway and discuss possible implications for migraine pain mechanisms. Finally, we discuss the insights into migraine pathophysiology obtained from the genetic models of migraine, in particular regarding the mechanisms that make the brain of migraineurs susceptible to the ignition of “spontaneous” CSDs. Although the reviewed functional studies support the view of migraine as a disorder of the brain characterized by dysfunctional regulation of the excitatory/inhibitory balance in specific neuronal circuits, much work remains to be done in the genetic mouse models e.g. to identfy the relevant dysfunctional circuits and to establish whether and how the alterations in the function of specific circuits (in the cerebral cortex and/or other brain areas) are state-dependent and may, in certain conditions, favor CSD ignition and the migraine attack.

Introduction

Migraine is much more than an episodic headache and a pain disorder. It is a complex brain disorder primarily affecting the sensory nervous system and characterized by a global dysfunction in multisensory information processing and integration. Indeed, in most attacks, the typical throbbing unilateral headache is associated with amplification of percepts from multiple senses indicating amplification of sensory gain. Hypersensitivity to sensory stimuli may persists in the interictal period, during which the brain of migraineurs show several alterations in sensory physiology. Interestingly the magnitude of some of these alterations increases in the interictal period towards the next attack and becomes maximal the day before the attack in temporal coincidence with prodromal symptoms (such as difficulty with speech, reading, concentration, increased emotionality, irritability, sensory hypersensitivity) that in many migraineurs are highly predictive of the attack [1,2,3,4,5]. The neurobiological mechanisms of the causative brain dysfunctions underlying the onset of a migraine attack and the alterations in multisensory information processing remain largely unknown and are key unanswered questions in migraine neurobiology.

In about 30% of migraineurs the headache is preceded by transient sensory (most frequently visual) disturbances, the so called migraine aura, whose neurophysiological correlate is now recognized to be cortical spreading depression (CSD) [6, 7]. CSD is a self-sustaining, slowly propagating (2–5 mm/min) wave of nearly complete depolarization of a sizable population of brain cells that lasts about one minute and silences brain electrical activity for several minutes. CSD can be induced in healthy brain tissue by intense depolarizing stimuli that increase the extracellular concentration of K+ ions, [K]e, above a critical threshold and release glutamate and other neurotransmitters. Although already the first studies of CSD mechanisms in the 50s pointed to [K]e and glutamate as key players in the CSD initiation mechanisms, these mechanisms are still incompletely understood [6].

There is evidence from animal studies that CSD can activate and sensitize the trigeminovascular pain pathway and hence can initiate the headache mechanisms [2, 8,9,10,11,12,13]. It is generally believed that the migraine headache begins with the activation and sensitization of trigeminal sensory afferents, that innervate cranial tissues, in particular the meninges, and subsequent activation and sensitization of second order neurons in the trigeminocervical complex (comprising the trigeminal subnucleus caudalis and dorsal horn of the first cervical segments, indicted here for simplicity as TNC) and of higher order neurons in areas of the brainstem and forebrain to which the TNC projects directly or indirectly; these areas are involved in different aspects of pain and in the complex migraine symptomatology [1, 2, 4]. Whereas the properties of pial afferents remain largely unknown, the dural afferents are nociceptors with properties similar to those in the rest of the body [1, 14]. A sterile meningeal inflammation is considered to be a key mechanism that may underlie the sustained activation and sensitization of meningeal nociceptors during migraine attacks [1, 14].

It has been shown that a single experimental CSD can lead to delayed sustained increases in dural blood flow and in ongoing activity of rat dural nociceptors and TNC trigeminovascular neurons as well as delayed sensitization of these neurons [8,9,10, 12, 15]. It has been suggested that the delayed trigeminal activation may result from CSD-induced release of proinflammatory molecules in the meninges e.g. as a consequence of a parenchimal inflammation initiated by CSD-induced opening of pannexin1 channels and inflammasome activation [11] and/or as a consequence of CSD-induced pial and dural macrophage activation [16]. Activation of peptidergic meningeal nociceptors and consequent release of proinflammatory neuropeptides, including calcitonin gene-related peptide (CGRP), from their peripheral nerve endings can then further promote meningeal inflammation [1, 14]. Given the efficacy of monoclonal antibodies against CGRP in migraine treatment [17], it is interesting that intravenous administration of such antibodies inhibited the CSD-induced activation of A-δ meningeal nociceptors and the CSD-induced activation and sensitization of high threshold TNC neurons which receive input from A-δ fibers [12, 18]. Moreover, a CGRP receptor antagonist reversed CSD-induced behavioural alterations associated with pain perception in awake animals, such as freezing, grooming and reduced thresholds of tactile allodynia, without blocking CSD waves in the cortex [13]. Thus, understanding the neurobiological mechanisms that make the brain of migraineurs susceptible to ignition of spontaneous CSDs is another (or perhaps the) key unanswered question in migraine neurobiology.

Migraine is a complex polygenic genetic disorder, with heritability estimates as high as 50% [19, 20]. Although genome-wide association studies (GWAS) are providing increasing insights into the common genetic variants associated with migraine [21], the study of the functional consequences of GWAS hits is very difficult, if not impossible, given also the fact that they generally lie in intronic or intergenic regions and therefore they likely influence gene regulation rather than directly protein function. Thus, “common” migraine is not amenable to being instantiated in a mouse model and defies attempts at determining mechanism. In contrast, rare monogenic forms of migraine are caused by mutations that directly affect protein function, and the functional consequences of the disease-causing mutations can be studied in genetic mouse models of the disease. Thus far there are five monogenic migraine mutations that have knock-in (KI) mouse lines associated with them, allowing investigation of the underlying mechanisms. Four of these are derived from patients with familial hemiplegic migraine (FHM) and one from patients with both migraine with aura (MA) and familial advanced sleep phase syndrome (FASPS), a rare sleep condition in which individuals go to sleep unusually early in the evening and wake early in the morning.

Apart from the motor weakness or hemiplegia during aura and the possible longer duration of aura, typical FHM attacks resemble common MA attacks and both types of attacks may alternate in patients and co-occur within families [20, 22]. Thus, FHM and MA are considered to be part of the same spectrum and may share pathogenetic mechanisms, despite clinical observations that the response to infusion of CGRP and glycerylnitrate seem to differ [20] [23]. Some FHM patients can have “atypical” severe attacks and show additional ictal and/or permanent neurological features such as epilepsy, loss of consciousness, ataxia and cognitive impairment [20, 22].

Thus far, three FHM causative genes, all encoding ion channels or transporters, have been identified: CACNA1A (FHM1), ATP1A2 (FHM2) and SCNA1A (FHM3). CACNA1A and SCNA1A encode the pore-forming subunits of the voltage-gated ion channels CaV2.1 and NaV1.1, while ATP1A2 encodes the α2 Na/K ATPase (α2 NKA) [24,25,26]. CaV2.1 channels are widely expressed in the nervous system, including all structures implicated in the pathogenesis of migraine; being localized at the active zones of most brain synaptic terminals, they play a dominant role in initiating synaptic transmission, particularly at central synapses; their somatodendritic localization points to additional postsynaptic roles ( [27] and references therein). FHM1 mutations produce gain of function of recombinant human CaV2.1 channels, mainly due to increased channel open probability and channel activation at lower voltages; the gain-of-function effect may be dependent on the specific CaV2.1 splice variant and/or auxiliary subunit [28] ( [27] and references therein).

Being highly expressed in inhibitory interneurons in several brain areas and being mainly localized at the axon initial segment, NaV.1.1 channels play a key role in interneurons excitability, particularly in sustaining high-frequency firing [29,30,31]. Indeed loss-of-function mutations in NaV1.1 channels cause a spectrum of epilepsy syndromes [32]. Although the findings regarding the functional consequences of FHM3 mutations on recombinant human NaV1.1 channels are somewhat conflicting, overall they suggest that, most likely, FHM3 is associated with gain-of-function of NaV1.1 channels and consequent selective hyperexcitability of cortical interneurons [33].

The α2 NKAs are primarily expressed in neurons during development and at the time of birth and almost exclusively in astrocytes in the adult brain [34,35,36,37,38]. At cortical excitatory synapses the α2 NKAs are colocalized with GLAST and GLT-1 glutamate transporters at perisynaptic astrocytic process [35, 39], where a large fraction of GLT-1/α2 NKA couples exhibit a separation distance indicative of physical coupling [38], thus supporting a key role of α2 NKAs in glutamate clearance during synaptic activity [39]. In contrast, α2 NKAs are not present in the large majority of astrocytic processes surrounding inhibitory synapses [35, 38]. α2 NKAs play also an important role in K+ clearance during neuronal activity [39] [40] and, most likely, in astrocytic Na+ and Ca2+ homeostasis [41, 42]. FHM2 mutations cause the complete or partial loss-of-function of recombinant α2 NKAs [22, 43].

A mutation in the casein kinase 1δ (CK1δ) gene, which among many other functions serves as a circadian clock gene, was identified as a cause of FASPS in a family that presented for clinical evaluation of a debilitating MA and, in addition, exhibited circadian patterns consistent with FASPS [44, 45]. CKIδ is a ubiquitous serine-threonine kinase that phosphorylates the circadian clock protein Per2 and many other proteins involved in brain signaling [46]. The identified CK1δ mutation resulted in reduced enzyme function in vitro [45]. Later, a second family was identified, with a second mutation in the coding region of the CK1δ gene – in this family again both advanced sleep phase and migraine segregated with the mutation [47].

The FHM1, FHM2 and CK1δ mouse models of migraine, that are the subject of the present review, provide a unique experimental system to study the cellular and circuit mechanisms of the primary brain dysfunctions causing a migraine disorder, and thus to tackle the key unanswered questions in migraine neurobiology mentioned above.

Genetic mouse models of migraine

Four different FHM mouse models were generated by introducing the human FHM1 R192Q or S218 L and FHM2 W887R or G301R mutations into the orthologous genes [48,49,50,51]. Whereas mutations R192Q and W887R cause in humans typical FHM attacks without additional clinical features (pure FHM) [24, 25], mutations S218 L and G301R cause severe clinical syndromes with atypical attacks that may include, in addition to hemiplegic migraine, prolonged coma/torpor or confusional state, epileptic seizures, elevated temperature, cerebral edema, transient or permanent cerebellar signs such as ataxia, nystagmus and cerebellar atrophy [52] [53, 54]. In agreement with the gain-of-function of FHM1 mutant human CaV2.1 channels [27] and the loss-of-function of FHM2 mutant human α2 NKAs [43], an increased neuronal Ca2+ current was measured in FHM1 mice in different types of neurons [27, 48, 55] and the brain expression of the α2 NKA was about 50% reduced in heterozygous FHM2 mice [50, 51]. The more severe clinical phenotype caused by the S218 L FHM1 mutation correlates with the larger gain-of-function of recombinant human and native neuronal mouse CaV2.1 channels produced by the S218 L compared to the R192Q mutation [49, 56]. In contrast, such correlation is not evident for the FHM2 mutations, since both the W887R and the G301R mutations completely eliminate pump activity of recombinant human α2 NKAs [57, 58] and reduce to a similar extent the α2 expression in the brain of adult heterozygous W887R and G301R mice [50, 51], likely due to ER retention and degradation of the misfolded protein [50, 54] (but cf. [59] for unaltered expression of the α2 NKA in the hippocampus of G301R mice in contrast with [51]).

The CK1δ mouse model was generated by inserting the human T44A mutation into the mouse genome via bacterial artificial chromosome (BAC) technique [47]. The migraines in patients with the T44A CK1δ mutation are phenotypically normal (they do not involve hemiplegia) [47].

We will focus on reviewing the functional studies in FHM1 and FHM2 KI mice which carry mutations causing pure FHM and thus should model migraine as close as possible, and will discuss the insights into migraine pathophysiology obtained from these studies. We will only briefly discuss the different findings in FHM1 and FHM2 KI mice which carry the severe syndrome-causing mutations and might give insights into the additional clinical features associated with these mutations. For the CK1δ mouse, as the subjects’ migraines are phenotypically normal, all migraine relevant phenotypes may be useful to extrapolate more generally to the migraine population; however for this model we need to consider whether the association with a sleep disorder affects the insights generated.

Migraine-relevant phenotypes in the genetic mouse models

Behavioural phenotypes

Homozygous KI mice carrying the R192Q pure FHM1 mutation (FHM1 mice) and heterozygous KI mice carrying the W887R pure FHM2 mutation (FHM2 mice) do not show an overt phenotype [48, 50]. However, the FHM1 mice showed signs of photophobia in a modified elevated plus maze in which the safe closed arms were brightly illuminated. Moreover, behavioral changes suggestive of unilateral head pain, such as increased head grooming, abnormal eye blinking, sustained single-eye closures and full-body shuddering, were reported in FHM1 mice when subjected to novelty or restrain stress [60]. Interestingly, systemic administration of the antimigraine drug rizatriptan normalized these pain measures. Given the higher female prevalence in migraine, another interesting finding is that the increase in shudder/blink frequency was larger in female mice [60]. These kind of data are not yet available for the FHM2 and CK1δ mice. However, in a SHIRPA primary screening assessing sensory, motor and neuropsychiatric functions, the FHM2 mice showed an increased level of fear/anxiety as the only behavioural anomaly [50], a feature that does not appear to be shared by the FHM1 mice [60].

Heterozygous KI mice carrying the severe G301R FHM2 mutation (G301R FHM2 mice) revealed several behavioural alterations, which although not comprising altered level of anxiety, included increased startle response to aversive acoustic stimuli, stress-induced depression-like phenotypes, decreased sociability and increased compulsive behaviour (as shown by increased face and body grooming and increased marble-burying); the latter was female-specific and reverted by progestin-only contraceptive treatment [51]. However, these symptoms were not reported in the two families in which the G301R mutation was identified (except for obsessive compulsive disorder in one member) [53, 54] and, with the exception of the hypersensitivity to noxious stimuli and, possibly, the stress-induced depression-like phenotypes, they are not part of the core symptoms reported for either FHM or migraine.

Homozygous KI mice carrying the severe S218L FHM1 mutation (S218L FHM1 mice) exhibit the main features of the human S218 L clinical syndrome, including attacks of hemiparesis, attacks of generalized seizures, mild cerebellar ataxia and brain edema after mild head impact. These features were not observed in heterozygous S218L mice [49].

Cortical spreading depression

A key migraine-relevant phenotype that the four FHM animal models have in common is increased susceptibility to experimentally induced CSD, as revealed by a lower stimulation threshold for initiation of CSD in vivo and in vitro [39, 48,49,50, 55] or a higher frequency of CSDs elicited by prolonged epidural high KCl application in vivo [61, 62]. Moreover, the rate of CSD propagation was increased in all four FHM animal models.

The CK1δ model provides additional evidence for increased CSD susceptibility as a key phenotype of models of MA. CK1δ mice had a decreased CSD initiation threshold as well as an increased number of CSD to a constant stimulus. There was a trend toward faster CSD velocity in mutants compared to wild type littermates, but this was not significant. [47].

In agreement with the higher female prevalence in migraine, the velocity of propagation and the frequency of CSDs induced by prolonged KCl were larger in female than in male in both FHM1 mouse mutants; the sex difference was abrogated by ovariectomy and enhanced by orchiectomy, suggesting that female and male gonadal hormones exert reciprocal effects on CSD susceptibility [61, 63]. However, no gender differences in induction and propagation of experimental CSD were found in FHM2 KI mice [50, 62], although, interestingly, the CSD frequency induced by prolonged KCl in postmenopausal female G301R FHM2 mutants was lower than in aged males and younger females [62], suggesting that perhaps a gender-based difference might be present in a limited period of the female menstrual cycle. In the CK1δ mouse, there was an intermediate sex phenotype between the FHM1 and FHM2 models. There was a gradient of CSD thresholds, with increasing threshold in the order mutant female < mutant male < wild type female < wild type male. Taken overall, the difference in thresholds between the four groups was significant, but on post-hoc testing this difference appeared to be driven by the difference between the two most separated groups (mutant female and wild type male). These experiments were not powered to detect a sex difference, so future work may confirm or refute the phenotype more definitively [47].

In male FHM1 mice, the frequency of CSDs induced by prolonged KCl was also increased after administration of the stress hormone corticosterone, but not after acute restrain stress [64]. Neither the stress hormone nor restrain stress affected CSD frequency in male wild type mice [64]. Even chronic stress did not affect CSD frequency and velocity in male wild type mice [65, 66]; however, interestingly, the threshold for CSD induction was lower after either acute or chronic stress [66], pointing to the possibility that the difference in CSD threshold between FHM mutants and wild type mice might be larger in stressed compared to control animals. Measurements of CSD threshold in FHM mutants after stress remain to be done.

In contrast with wild type mice, a single CSD caused prolonged hemiplegia with leaning and circling in FHM1 mice, and CSD readily propagated into the striatum in the majority of FHM1 but not wild type mice, suggesting corticostriatal CSD propagation as a likely explanation for the more severe motor deficits in FHM1 mutants [61, 67, 68].

Moreover, the typical reduction of cerebral blood flow after CSD was more prolonged in FHM1 mice [69] and more severe in heterozygous S218L FHM1 mice [70]. In contrast, the cerebral blood flow in G301R FHM2 mice was similar to that in wild type mice, but the typical reduction of spontaneous activity after CSD was more prolonged [51].

During CSD, the increase of the intracellular concentration of Ca2+ ions [Ca2+]in in cortical neurons and neuropil was larger in both FHM1 mouse models compared to wild type mice [69, 70], which was expected given the gain-of-function of CaV2.1 channels in cortical pyramidal cells from both mutants [55, 71]. Consistent with this, the metabolic oxygen consumption was larger in FHM1 mice and resulted in a larger decrease of tissue oxygenation during CSD compared to wild type mice, due to a larger mismatch between oxygen consumption and oxygen supply from the cerebral blood flow (CBF), being CBF similar in the two genotypes during CSD [69] (but cf. [70]).

In good correlation with the larger CaV2.1 gain-of-function produced by the S218L compared to the R192Q mutation [28, 48, 49, 56], the strength of CSD facilitation as well as the severity of the post-CSD neurological motor deficits and the propensity of CSD to propagate into subcortical structures were larger in S218L compared to R192Q FHM1 mice [49, 61, 67, 68]. Moreover, unlike the R192Q mice, the S218 L mice frequently developed multiple CSDs after a single CSD-inducing stimulus, which were more frequent in homozygotes compared to heterozygotes [49]. Furthermore, unlike the R192Q mice (and the heterozygous S218L), the homozygous S218L mutants developed generalized seizure after about 1 h from a single CSD [61]. In awake G301R FHM2 mice, full tonic-clonic seizures were frequently observed after a certain number of CSDs elicited by continous KCl application [62]. These unique CSD features might contribute to the additional clinical symptoms in the severe human syndromes caused by the S218L and G301R mutations.

Migraine-relevant alterations in the brain of genetic mouse models of migraine

Synaptic alterations in the cerebral cortex and CSD mechanisms

Direct measurements of synaptic transmission at different cortical synapses in pure FHM1 mice invariably revealed enhanced glutamatergic transmission at the excitatory synapses, due to enhanced action-potential (AP) evoked Ca2+ influx through mutant presynaptic CaV2.1 channels and enhanced probability of glutamate release [55]. As expected from the enhanced release probability, short-term depression was increased at mutant excitatory synapses on both pyramidal cells and fast-spiking interneurons (athough to a different extent) [55]. In striking contrast with glutamatergic transmission, GABAergic transmission at different cortical inhibitory synapses was unaltered in FHM1 mice, despite being initiated by CaV2.1 channels [55, 72]. The lack of effect on inhibitory synaptic transmission in contrast with the gain-of-function effect on excitatory transmission appears to be a common feature of FHM1 mutations since it was shown also for the severe S218L mutation [71]. Expression of interneuron-specific CaV2.1 channels whose gating properties are barely affected by the FHM1 mutation likely underlies this unexpected finding [72].

Although inhibitory transmission was not investigated, evidence for increased glutamatergic neurotransmission in FHM1 mice was also obtained at several other excitatory synapses in different brain areas, including those onto dorsal suprachiamastic nucleus neurons [73], cerebellar parallel fibers-Purkinje cell synapses [74] and stratum radiatum CA1 synapses [75]. Interestingly, long-term potentiation (LTP) at these hippocampal synapses was enhanced in FHM1 mice, while long term depression was unaltered; paradoxically, spatial learning in contextual fear conditioning and Morris water maze tests was impaired [75].

In the cerebral cortex, specialized core microcircuits, which involve different types of inhibitory interneurons and mediate feedback inhibition, feedforward inhibition and disinhibition, regulate the relative strenght of excitatory and inhibitory synaptic conductances in cortical neurons and their temporal and spatial relationships and, thus, dynamically maintain the excitatory-inhibitory (E/I) balance necessary for the transfer of information while preventing runaway excitation [76]. These microcircuits are essential for the correct processing of sensory information (e.g. for sensory gain control, surround suppression, synchronization and generation of cortical rythms, formation of cell assemblies and information transfer to higher areas) [76, 77].

The differential effect of FHM1 mutations on excitatory and inhibitory synaptic transmission (and on short-term synaptic plasticity at different cortical synapses) implies that these core microcircuits are most likely dysfunctional in FHM1 [78] and points to impaired regulation of cortical E/I balance as a primary brain dysfunction and a key pathogenic mechanism in FHM1 [79]. On the basis of the available data on the effect of the FHM1 mutations on synaptic transmission at different cortical synapses, it is not straightforward to predict their effect on microcircuit and network function. In fact, while enhanced excitatory transmission at the synapses on cortical principal neurons would increase network excitation, enhanced transmission at the synapses onto inhibitory interneurons may lead to increased recruitment of the interneurons and hence to increased inhibition.. This could, at least in part, explain the unexpected reduction in neuronal calcium responses to prolonged repeated whisker stimulation reported in FHM1 mice [69].

As shown by CSD rescue experiments, there is a causative link between increased glutamatergic transmission at cortical synapses and facilitation of experimental CSD in FHM1 mice. In fact, the facilitation of initiation and propagation of CSD in mutant cortical slices was completely eliminated when AP-evoked glutamate release at pyramidal cell synapses was brought back to wild type values by partially inhibiting the CaV2.1 channels [55]. The finding that propagation of CSD to subcortical structures in FHM1 mice was eliminated by systemic treatment with pregabalin, a drug that reduced excitatory transmission in mutant hippocampal slices, suggests that the increased propensity of CSD to propagate into subcortical structures is also linked to increased excitatory neurotransmission [68].

The important role of excessive glutamatergic transmission in migraine mechanisms, in particular CSD susceptibility, is underscored and supported by the functional studies in pure FHM2 mice. These mutants show reduced rate of synaptic glutamate clearance by cortical astrocytes during neuronal activity and reduced density of glutamate transporters GLT-1 at perisynaptic astrocytic processes, which mirrors the reduced expression of the α2 NKA [39].

LTP induced by high frequency stimulation at the hippocampal perforant path synapses in the dentate gyrus was enhanced in the FHM2 mice; in contrast, LTP was unaltered at stratum radiatum-CA1 synapses [80]. Basal synaptic transmission was unaltered in both areas, judging from the similar input-output curves and paired pulse ratios in field recordings in wild type and mutant hippocampal slices [80]. These findings may be consistent with the larger impairment of glutamate clearance uncovered in FHM2 cortical slices after high-frequency stimulation of glutamate release compared to low frequency (single pulse) stimulation [39]. Also consistent with this is the finding that mice with a 60–80% reduction of GLT-1 expression (after conditional knockout in adolescents) show unaltered basal synaptic transmission at corticostriatal synapses, but decreased EPSC depression during prolonged stimulation [81]. Interestingly, these mice showed increased compulsive behaviour (as shown by increased self-grooming), which was rescued by treatment with the NMDA glutamate receptor (NMDAR) antagonist memantine.

Memantine treatment also rescued the female-specific compulsive behaviour in heterozygous G301R FHM2 mice [51], suggesting that, although not directly measured, also in these mice the expression of GLT-1 receptors is reduced (possibly more than in pure FHM2 mice which did not show compulsive behaviour) and results in impaired clearance of glutamate at excitatory synapses and increased NMDAR activation likely as a consequence of glutamate spillover. In apparent conflict with this, the rate of glutamate uptake measured in mixed neuron-astrocytes cultures from heterozygous G301R FHM2 embrios was not significantly decreased [51]. The unsuitability of astrocytic cultures for the study of NKA function [59] might underlie this finding.

Since, in the cortex, the α2 NKA pump is localized in astrocytic processes surrounding glutamatergic synapses and only in a small fraction, if any, of astrocytic processes surrounding GABAergic synapses [35, 38], FHM2 mutations likely affect excitatory but not inhibitory synaptic transmission, and thus may lead to altered circuit function and impaired regulation of cortical E/I balance, as in FHM1.

Interestingly, it has been shown that the defective glutamate clearance at cortical excitatory synapses in the FHM2 mice may largely account for the lower threshold for induction of experimental CSD in these mice [39]. The FHM2 mutants also showed reduced rate of K+ clearance during neuronal activity, and the defective clearance of both glutamate and K+ likely accounts for the increased rate of CSD propagation [39].

Overall the findings in the mouse models of pure FHM support the conclusion that their increased susceptibility to experimental CSD is largely due to excessive cortical glutamatergic transmission, arising from either increased glutamate release (FHM1) or impaired glutamate clearance (FHM2).

Together with pharmacological data in wild type mice providing strong support for a key role of glutamate NMDARs and CaV2.1 channels in initiation and/or propagation of experimental CSD [6, 82, 83], the findings in FHM mice support a model of CSD initiation in which i) CaV2.1-dependent glutamate release and consequent activation of NMDARs are key elements for generation of the net self-sustaining inward current necessary to initiate the positive feedback cycle that ignites a propagating CSD when the removal of K+ and glutamate from the interstitium does not keep pace with their release and ii) the α2 NKA pumps exert a dampening role owing to their key role in K+ and, in particular, glutamate clearance by astrocytes [6, 39, 55, 84]. Moreover, the findings are consistent with a model of CSD propagation in which interstitial K+ diffusion initiates the CSD positive-feedback cycle in contiguous dendritic regions [6, 39, 55, 85].

In FHM1 mice carrying the severe S218L mutation, gain-of-function of additional CaV2.1-dependent processes, besides enhanced glutamatergic synaptic transmission, likely underlie the particularly high susceptibility to CSD and high propensity of CSD to spread into subcortical structures as well as some of the unique features of CSD, including its recurrence, which were not observed in pure FHM1 mice [49, 61, 67, 68, 71]. A specific feature of mice carrying the S218L mutation is the presence of a fraction of mutant CaV2.1 channels that is open at resting potential, as revealed by the reduced frequency of miniature excitatory postsynaptic currents (mEPSCs) after block of CaV2.1 channels in cortical slices from both heterozygous and homozygous S218L mutants [71], and, by increased baseline [Ca2+]in in layer 2/3 axonal boutons and shafts in heterozygous S218L mice in vivo [70]. Both a reduction of mEPSCs frequency after blocking CaV2.1 channels and an increase in basal [Ca2+]in in synaptic terminals were measured at Calyx of Held synapses in brainstem slices from S218L mice [86]. In contrast, the mEPSCs frequency at cortical and brainstem synapses was not altered in pure FHM1 mice, indicating that presynaptic CaV2.1 channels carrying the R192Q mutation are closed at resting potential in brain slices [55, 87]. Probably as a consequence of the increase in baseline [Ca2+]in, the heterozygous S218L mice showed some alterations in axonal and dendritic morphology in the resting state, including slightly larger boutons [88]. It remains to be seen whether this is a specific functional consequence of severe FHM1 mutations, which contributes to some of the additional clinical features associated with them.

Alterations in the trigeminovascular pain pathway and pain mechanisms

The function of the trigeminovascular (TGV) pain pathway is expected to be altered in FHM1 mice because CaV2.1 channels are involved in controlling neurotransmitter release at different levels in the trigeminovascular system, e.g. CGRP release from capsaicin-sensitive perivascular terminals of meningeal nociceptors as well as release at central synapses onto TNC neurons including synapses of the descending inhibitory and facilitatory pathways that regulate TGV pain transmission ( [27] and references therein). However, relatively few studies investigated the function of the trigeminovascular pain network in FHM1 KI mice; moreover, most of these studies were in vitro and focused on the peripheral part of the network.

Investigation of CGRP release from dura mater in fluid-filled hemisected skulls revealed that neither basal nor K+-evoked CGRP release were significantly altered in adult FHM1 mice [89, 90]. Since a large fraction of peptidergic dural trigeminal ganglion (TG) afferents are capsaicin-sensitive [1], these findings are consistent with the fact that the CaV2.1 current was unaltered in small capsaicin-sensitive TG neurons from adult FHM1 mice, which, according to retrograde labelling from the dura, constitute the majority of small dural afferents [89]. This may also contribute to explain the finding that dural artery vasodilation induced in vivo by systemic capsaicin was not increased in FHM1 mice; actually, vasodilation induced by both systemic capsaicin and CGRP was decreased [90], suggesting downregulation and/or desensitization of blood vessels CGRP receptors, perhaps as a compensatory mechanism. The lower fraction of CGRP-expressing neurons uncovered in trigeminal ganglia of FHM1 mice [91] may be an additional compensatory mechanism, that might also contribute to the unaltered CGRP release from dura mater in the FHM1 mutants.

All together, these findings argue against the idea that increased CGRP release from perivascular TG fibers at the dura and consequent increased vasodilation and mast cell degranulation facilitate the development of neurogenic inflammation (following activation of meningeal nociceptors e.g. by CSD) in FHM1 compared to wild type mice. Perhaps this is a consequence of compensatory mechanisms that might be triggered by the occurrence of spontaneous CSDs in the FHM1 mutants.

Judging from the finding of unaltered c-fos expression in the TNC after in vivo electrical stimulation of the dura in FHM1 mice [92], also synaptic transmission at the central terminals of dural TG afferents might be unaffected by the FHM1 mutation. This would be consistent with unaltered CaV2.1 current in most dural TG afferents of FHM1 mice, as was shown in small capsaicin-sensitive dural afferents [89]. This further underscores the importance of neuron subtype-specific effects of FHM1 mutations in the FHM1 pathophysiology. Indeed, the finding of increased c-fos expression in several thalamic nuclei after in vivo electrical stimulation of the dura [92], is consistent with increased synaptic transmission at TNC-thalamus excitatory synapses as a consequence of gain-of-function of CaV2.1 channels located at thalamic synaptic terminals of TNC neurons. Although this remains to be demonstrated, it would contribute to increase the gain of the TGV pain pathway in FHM1.

Depending on the study, K+-evoked CGRP release from isolated trigeminal ganglia was either increased [89] or unaltered [90] in adult FHM1 mice; in the latter study, also CGRP release from TNC was unaltered in the mutants. Enhanced K+-evoked CGRP release from trigeminal ganglia implies gain-of-function of CaV2.1 channels in some TG neurons in FHM1 mice; this was indeed shown in a subpopulation of small capsaicin-insensitive neurons, which do not innervate the dura [89]. Given that in these neurons the action potential-evoked CaV2.1 current is larger in FHM1 mice [89], one predicts enhanced transmitter release upon their activation. However, the function, transmitters and possible involvement in migraine pain of this subpopulation of small TG neurons remain unknown.

In cultured TG neurons from FHM1 mice pups, also basal (besides K+-evoked) CGRP release was increased, suggesting opening of mutant CaV2.1 channels at resting potential [93]. Congruently, these cultured TG neurons show interesting CaV2.1-dependent alterations such as loss of constitutive inhibition of ATP-gated P2X3 receptors (P2X3Rs) by brain natriuretic peptide receptors, which leads to increased P2X3R current and enhanced excitability in response to ATP in FHM1 mice [94,95,96]. The neuronal upregulation of the P2X3R function (as well as the upregulation of P2X7 receptors function recently uncovered in satellite glial cells and macrophages) were eliminated after inhibition of the CGRP receptors [93, 97, 98]. This is consistent with the idea that the increased basal release of CGRP promotes sensitization of P2X3R-expressing TG neurons, cross-talk between neurons and satellite glial cells and macrophages, resulting in a local persistent inflammatory environment in the FHM1 TG [93, 97, 98]. However, basal release of CGRP was not increased in trigeminal ganglia from adult FHM1 mice [89, 90], suggesting caution in drawing conclusions regarding migraine pain mechanisms from findings in pups TG cultures. Whether the adult TG shows a basal inflammatory phenotype in FHM1 mutants remains unclear, since in FHM1 ganglia the number of active macrophages was increased (in all divisions), but the protein level of the pro-inflammatory cytokines IL1beta, IL6 and TNFalpha was unaltered [99]. Interestingly, a larger fraction of TG neurons was immunoreactive for active phosphorylated CaMKII in FHM1 compared to wild type ganglia; the difference in amount of phosphoprotein between the two genotypes was eliminated after blockade of CaV2.1 channels, suggesting facilitation of basal CaV2.1-dependent Ca signalling in FHM1 TG neurons [94].

The CK1δ mouse showed strong evidence for activation of migraine-relevant pain pathways. Nitroglycerin (NTG) infusion has been used as a trigger for migraine without aura in humans [100], and was later adapted for rat [101]. After adaptation of the methods for mouse [102], it was used to test both heat and mechanical withdrawal thresholds in CK1δ mice and wild type littermates. There was a significant reduction in both heat and mechanical withdrawal thresholds in CK1δ mutant mice [47]. As NTG-induced threshold changes were responsive to the migraine abortive sumatriptan in wild type mice [102] these data were taken as evidence for an enhanced algesic response to a migraine trigger in CK1δ mice [47]. There was also a significant increase in the number of c-fos-reactive cells in the TNC after NTG in CK1δ mutant compared to wild type mice, consistent with increased activation of craniofacial pain networks by this migraine trigger [47].

Insights into migraine pathophysiology

The genetic mouse models of migraine support the view of migraine as a disorder of the brain characterized by dysfunctional regulation of the E/I balance in specific neuronal circuits in the cerebral cortex and other brain structures. Moreover, they support a key role of CSD in the pathogenesis of migraine with aura and provide insights into how a “spontaneous” CSD can arise in the brain of migraineurs.

The induction of experimental CSD in healthy tissue requires intense depolarizing stimuli that, according to the model of CSD initiation proposed in the previous section, increase the extracellular [K+] above a critical value and release enough glutamate to overwhelm the binding capacity of astrocytic glutamate transporters, thus leading to cooperative activation of the high number of synaptic and extrasynaptic NMDARs necessary for initiation of the CSD positive feedback cycle. In migraineurs CSD is not induced by experimental depolarizing stimuli, but arises “spontaneously” in certain conditions. How can this occurs?

The findings in the FHM mouse models suggest that ignition of a “spontaneous” CSD could be favoured by conditions leading to excessive activation of synaptic and extrasynaptic NMDARs, i.e. conditions leading to membrane depolarization and to overwhelming of the transport capacity of astrocytic glutamate transporters. This would probably require high-frequency repetitive or synchronous activity of a sufficient number of excitatory synapses in which glutamatergic transmission is potentiated, e.g. as a consequence of increased glutamate release as in FHM1 or reduced expression of astrocytic α2 NKA and glutamate transporters, as in FHM2, or as a consequence of other mechanisms in common migraine. We hypothesize that this may occur in certain conditions as a consequence of dysfunctional regulation of the E/I balance in specific cortical circuits. Much work remains to be done in the FHM models to identify the relevant dysfunctional cortical circuits and establish whether indeed dysfunctional regulation of the E/I balance in these circuits may favour CSD ignition and identify the specific conditions (brain states) in which this may occur.

The behavioural phenotypes and the functional analysis of the genetic mouse models of migraine are consistent with the concept of migraine as a disorder of sensory network gain and plasticity [3]. Much work remains to be done to investigate possible alterations in sensory processing in awake animals, that may underlie some of the interictal alterations in sensory processing shown by migraineurs, and to investigate the underlying cellular and circuit mechanisms. It will be also important to investigate whether the alterations in the function of specific circuits (in the cortex and/or other brain structures) in the genetic models are modulated by state-dependent changes in plasticity, and may thus underlie some of the interictal cyclical changes in sensory physiology and/or some of the prodromal symptoms shown by migraineurs.

Translational relevance

An important point to make at the outset is that all models are what their name says – models. As such they are abstractions from disease reality that allow cleaner testing of hypotheses than the disease state allows; or asking of questions that cannot even be asked in the disease state. In return for the abstraction and simplification from the disease state, models are ‘expected’ to generate insight that is not otherwise possible. In this regard all of the models described have both elements of simplification and abstraction; but they also have delivered on their promise. Probably the most important insight coming from the diverse genetic models of migraine is the role of circuit excitability, with CSD as the primary circuit phenotype, underlain perhaps by excessive glutamatergic neurotransmission and/or excessive activation of glutamate NMDARs.

The question arises whether models derived from rare monogenic forms of migraine can yield insights for the rest of the disease. Most migraineurs do not have hemiplegic migraine or any monogenic form; indeed most migraineurs do not have migraine with aura, and all models thus far come from families with migraine with aura. A first reply is that across biology, the use of rare mutants has enabled major discoveries about disease mechanisms relevant to the larger population. Their value comes from their rarity; their monogenic nature allows for identification of specific mechanisms that would be impossible to detect for migraine in the larger population. As to whether the insights gained from monogenic models are generalizable to migraine at large, the jury is still out, but there is reason to be very optimistic. The fact that diverse mutations in neurons and non-neuronal cells converge on a single circuit phenotype – CSD – is quite promising from the point of view of generalization. This is especially true for migraine with aura of course. It is also worth noting that the addition of the CK1δ mouse, from a family with non-hemiplegic migraine, strongly suggests that the insights gained from monogenic models are not confined to humans with hemiplegic migraine.

For the clinician who desires a ‘bottom line’ on the relevance of these genetic models of migraine, the most important point is that by allowing the pursuit of precise mechanisms (what protein is acting, in what way, on what circuit?) they also allow the pursuit of precise solutions – drugs, biologics, stimulation paradigms, or other ways of addressing the disease that we have not yet considered. They are not the only way to derive new treatments – brute-force empiricism has worked for centuries – but they are certainly more targeted, and arguably much more satisfying, because they have the ability to actually tell us how the disease works.

Conclusions

Mouse models of rare monogenic forms of migraine provide a unique experimental system to study the cellular and circuit mechanisms of the primary brain dysfunctions causing a migraine disorder. A key migraine-relevant phenotype that these animal models have in common is increased susceptibility to experimentally induced CSD. In the FHM mouse models this is largely due to excessive cortical glutamatergic transmission, arising from either increased glutamate release (FHM1) or impaired glutamate clearance (FHM2). The genetic animal models provide insights into how a “spontaneous” CSD can arise in the brain of migraineurs and support the view of migraine as a disorder of the brain characterized by dysfunctional regulation of the E/I balance in specific neuronal circuits in the cerebral cortex and other brain structures. Much work remains to be done in these models to identify the relevant dysfunctional circuits and to establish whether and how the alterations in the function of specific circuits are state-dependent and may, in certain conditions, favour CSD ignition and the migraine attack.

Availability of data and materials

Not applicable.

Abbreviations

[Ca2+]in :

Intracellular Ca2+ concentration

[K]e :

Extracellular concentration of K+ ions

AP:

Action potential

1δ (CK1δ):

Casein kinase

CBF:

cerebral blood flow

CGRP:

Calcitonin gene-related peptide

CSD:

Cortical spreading depression

E/I:

Excitatory-inhibitory balance

FASPS:

Familial advanced sleep phase syndrome

FHM:

Familial hemiplegic migraine

FHM1 mice:

Homozygous knock-in mice carrying the R192Q pure FHM1 mutation

FHM2 mice:

Heterozygous knock-in mice carrying the W887R pure FHM2 mutation

GWAS:

Genome-wide association studies

KI:

Knock-in

mEPSC:

miniature excitatory postsynaptic current

NKA:

Na/K ATPase

NMDAR:

NMDA glutamate receptor

NTG:

Nitroglycerin

P2X3R:

P2X3 receptor

TG:

Trigeminal ganglion

TGV:

Trigeminovascular

TNC:

Trigeminocervical complex

References

  1. Pietrobon D, Moskowitz MA. Pathophysiology of migraine. Annu Rev Physiol 2013;75:365–391. PubMed PMID: 23190076.

  2. Burstein R, Noseda R, Borsook D (2015) Migraine: Multiple Processes. Complex Pathophysiology The Journal of Neuroscience 35(17):6619–6629.

    CAS  PubMed  Google Scholar 

  3. Brennan KC, Pietrobon D. A systems neuroscience approach to migraine. Neuron. 2018 Mar 7;97(5):1004–1021. PubMed PMID: 29518355. Pubmed Central PMCID: 6402597.

  4. Goadsby PJ, Holland PR, Martins-Oliveira M, Hoffmann J, Schankin C, Akerman S. Pathophysiology of Migraine: A Disorder of Sensory Processing. Physiological Reviews. 2017 2017/04/01;97(2):553–622.

  5. de Tommaso M, Ambrosini A, Brighina F, Coppola G, Perrotta A, Pierelli F, et al. Altered processing of sensory stimuli in patients with migraine. Nat Rev Neurol. 2014 03//print;10(3):144–55.

  6. Pietrobon D, Moskowitz MA. Chaos and commotion in the wake of cortical spreading depression and spreading depolarizations. Nat Rev Neurosci 2014 Jun;15(6):379–393. PubMed PMID: 24857965.

  7. Ayata C, Lauritzen M. Spreading Depression, Spreading Depolarizations, and the Cerebral Vasculature. Physiological Reviews. 2015 07/01;95(3):953–93. PubMed PMID: PMC4491545.

  8. 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 Jun 30;30(26):8807–8814. PubMed PMID: 20592202.

  9. Zhang X, Levy D, Kainz V, Noseda R, Jakubowski M, Burstein R. Activation of central trigeminovascular neurons by cortical spreading depression. Ann Neurol 2011 May;69(5):855–865. PubMed PMID: 21416489.

  10. Zhao J, Levy D. Modulation of intracranial meningeal nociceptor activity by cortical spreading depression: a reassessment. Journal of Neurophysiology. 2015 02/18 12/09/received 02/16/accepted;113(7):2778–85. PubMed PMID: PMC4416615.

  11. Karatas H, Erdener SE, Gursoy-Ozdemir Y, Lule S, Eren-Kocak E, Sen ZD, et al. Spreading depression triggers headache by activating neuronal Panx1 channels. Science. 2013 Mar 1;339(6123):1092–1095. PubMed PMID: 23449592.

    Article  CAS  Google Scholar 

  12. Melo-Carrillo A, Noseda R, Nir R-R, Schain AJ, Stratton J, Strassman AM et al (2017) Selective inhibition of Trigeminovascular neurons by Fremanezumab: a humanized monoclonal anti-CGRP antibody. J Neurosci 37(30):7149–7163.

    Article  CAS  Google Scholar 

  13. Filiz A, Tepe N, Eftekhari S, Boran HE, Dilekoz E, Edvinsson L, et al. CGRP receptor antagonist MK-8825 attenuates cortical spreading depression induced pain behavior. Cephalalgia. 2019 Mar;39(3):354–365. PubMed PMID: 28971699.

  14. Levy D (2010) Migraine pain and nociceptor activation—where do we stand? Headache: The Journal of Head and Face Pain 50(5):909–916.

    Article  Google Scholar 

  15. 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 Feb;8(2):136–142. PubMed PMID: 11821897.

  16. Schain AJ, Melo-Carrillo A, Borsook D, Grutzendler J, Strassman AM, Burstein R. Activation of pial and dural macrophages and dendritic cells by cortical spreading depression. Ann Neurol 2018 Mar;83(3):508–521. PubMed PMID: 29394508. Pubmed Central PMCID: 5965700.

  17. Russo AF. Calcitonin Gene-Related Peptide (CGRP): A New Target for Migraine. Annual Review of Pharmacology and Toxicology. 2015 2015/01/06;55(1):533–52.

  18. Melo-Carrillo A, Strassman AM, Nir R-R, Schain AJ, Noseda R, Stratton J et al (2017) Fremanezumab—a humanized monoclonal anti-CGRP antibody—inhibits thinly myelinated (Aδ) but not unmyelinated (C) meningeal nociceptors. J Neurosci 37(44):10587–10596.

    Article  CAS  Google Scholar 

  19. Sutherland HG, Griffiths LR (2017) Genetics of migraine: insights into the molecular basis of migraine disorders. Headache: The Journal of Head and Face Pain. 57(4):537–569.

    Article  Google Scholar 

  20. Ferrari MD, Klever RR, Terwindt GM, Ayata C, van den Maagdenberg AMJM. Migraine pathophysiology: lessons from mouse models and human genetics. The Lancet Neurology. 2015 1//;14(1):65–80.

  21. Gormley P, Anttila V, Winsvold BS, Palta P, Esko T, Pers TH, et al. Meta-analysis of 375,000 individuals identifies 38 susceptibility loci for migraine. Nature Genetics. 2016 06/20/online;48:856.

  22. Pietrobon D. Familial hemiplegic migraine. Neurotherapeutics. 2007 Apr;4(2):274–284. PubMed PMID: 17395138.

  23. Hansen JM, Thomsen LL, Olesen J, Ashina M. Coexisting typical migraine in familial hemiplegic migraine. Neurology. 2010 Feb 16;74(7):594–600. PubMed PMID: 20157162.

  24. Ophoff RA, Terwindt GM, Vergouwe MN, van Eijk R, Oefner PJ, Hoffman SMG et al (1996) Familial hemiplegic migraine and episodic ataxia type-2 are caused by mutations in the Ca2+ channel gene CACNL1A4. Cell. 87:543–552.

    Article  CAS  Google Scholar 

  25. De Fusco M, Marconi R, Silvestri L, Atorino L, Rampoldi L, Morgante L, et al. Haploinsufficiency of ATP1A2 encoding the Na+/K+ pump alpha2 subunit associated with familial hemiplegic migraine type 2. Nat Genet 2003 Feb;33(2):192–196. PubMed PMID: 12539047.

  26. Dichgans M, Freilinger T, Eckstein G, Babini E, Lorenz-Depiereux B, Biskup S, et al. Mutation in the neuronal voltage-gated sodium channel SCN1A in familial hemiplegic migraine. Lancet. 2005 Jul 30-Aug 5;366(9483):371–7. PubMed PMID: 16054936.

  27. Pietrobon D. Calcium channels and migraine. Biochimica et Biophysica Acta (BBA) - Biomembranes. 2013 2013/07/01/;1828(7):1655–65.

  28. Tottene A, Fellin T, Pagnutti S, Luvisetto S, Striessnig J, Fletcher C, et al. Familial hemiplegic migraine mutations increase ca (2+) influx through single human CaV2.1 channels and decrease maximal CaV2.1 current density in neurons. Proc Natl Acad Sci U S A 2002 Oct 1;99(20):13284–13289. PubMed PMID: 12235360.

  29. Yu FH, Mantegazza M, Westenbroek RE, Robbins CA, Kalume F, Burton KA, et al. Reduced sodium current in GABAergic interneurons in a mouse model of severe myoclonic epilepsy in infancy. Nat Neurosci 2006 Sep;9(9):1142–1149. PubMed PMID: 16921370.

  30. Ogiwara I, Miyamoto H, Morita N, Atapour N, Mazaki E, Inoue I, et al. Nav1.1 Localizes to Axons of Parvalbumin-Positive Inhibitory Interneurons: A Circuit Basis for Epileptic Seizures in Mice Carrying an Scn1a Gene Mutation. J. Neurosci. 2007 May 30, 2007;27(22):5903–5914.

  31. Hedrich UBS, Liautard C, Kirschenbaum D, Pofahl M, Lavigne J, Liu Y et al (2014) Impaired action potential initiation in GABAergic interneurons causes Hyperexcitable networks in an epileptic mouse model carrying a human Na<sub>V</sub>1.1 mutation. J Neurosci 34(45):14874–14889.

    Article  Google Scholar 

  32. Schutte SS, Schutte RJ, Barragan EV, O'Dowd DK. Model systems for studying cellular mechanisms of SCN1A-related epilepsy. J.Neurophysiology. 2016 02/03 08/24/received 01/22/accepted;115(4):1755–66. PubMed PMID: PMC4819066.

  33. Cestèle S, Schiavon E, Rusconi R, Franceschetti S, Mantegazza M. Nonfunctional NaV1.1 familial hemiplegic migraine mutant transformed into gain of function by partial rescue of folding defects. Proceedings of the National Academy of Sciences. 2013 October 22, 2013;110(43):17546–51.

  34. McGrail KM, Phillips JM, Sweadner KJ. Immunofluorescent localization of three Na,K-ATPase isozymes in the rat central nervous system: both neurons and glia can express more than one Na,K-ATPase J Neurosci 1991 Feb;11(2):381–391. PubMed PMID: 1846906.

  35. Cholet N, Pellerin L, Magistretti PJ, Hamel E. Similar perisynaptic glial localization for the Na+,K+-ATPase alpha 2 subunit and the glutamate transporters GLAST and GLT-1 in the rat somatosensory cortex. Cereb Cortex 2002 May;12(5):515–525. PubMed PMID: 11950769.

  36. Ikeda K, Onaka T, Yamakado M, Nakai J, Ishikawa TO, Taketo MM, et al. Degeneration of the amygdala/piriform cortex and enhanced fear/anxiety behaviors in sodium pump alpha2 subunit (Atp1a2)-deficient mice. J Neurosci 2003 Jun 1;23(11):4667–4676. PubMed PMID: 12805306.

  37. Moseley AE, Lieske SP, Wetzel RK, James PF, He S, Shelly DA, et al. The Na,K-ATPase alpha 2 isoform is expressed in neurons, and its absence disrupts neuronal activity in newborn mice. J Biol Chem 2003 Feb 14;278(7):5317–5324. PubMed PMID: 12458206.

  38. Melone M, Ciriachi C, Pietrobon D, Conti F. Heterogeneity of astrocytic and neuronal GLT-1 at cortical excitatory synapses, as revealed by its Colocalization with Na+/K+-ATPase alpha isoforms. Cereb Cortex 2018 27. PubMed PMID: 30260367.

  39. Capuani C, Melone M, Tottene A, Bragina L, Crivellaro G, Santello M et al (2016) Defective glutamate and K<sup>+</sup> clearance by cortical astrocytes in familial hemiplegic migraine type 2. EMBO Molecular Medicine 8(8):967–986.

    Article  CAS  Google Scholar 

  40. Larsen BR, Assentoft M, Cotrina ML, Hua SZ, Nedergaard M, Kaila K et al (2014) Contributions of the Na+/K+-ATPase, NKCC1, and Kir4.1 to hippocampal K+ clearance and volume responses. Glia. 62(4):608–622.

    Article  Google Scholar 

  41. Rose CR, Verkhratsky A. Glial ionic excitability: the role for sodium. Glia. 2016 Oct;64(10):1609–10. PubMed PMID: 27509366.

  42. Verkhratsky A, Nedergaard M. Physiology of Astroglia. Physiol Rev 2018 Jan 1;98(1):239–389. PubMed PMID: 29351512. Pubmed Central PMCID: 6050349.

  43. Bøttger P, Doğanlı C, Lykke-Hartmann K. Migraine- and dystonia-related disease-mutations of Na+/K+-ATPases: Relevance of behavioral studies in mice to disease symptoms and neurological manifestations in humans. Neuroscience & Biobehavioral Reviews. 2012 2//;36(2):855–71.

  44. Jones CR, Campbell SS, Zone SE, Cooper F, DeSano A, Murphy PJ, et al. Familial advanced sleep-phase syndrome: a short-period circadian rhythm variant in humans. Nat Med 1999 Sep;5(9):1062–1065. PubMed PMID: 10470086.

  45. Xu Y, Padiath QS, Shapiro RE, Jones CR, Wu SC, Saigoh N, et al. Functional consequences of a CKIdelta mutation causing familial advanced sleep phase syndrome. Nature. 2005 Mar 31;434(7033):640–644. PubMed PMID: 15800623.

  46. Knippschild U, Gocht A, Wolff S, Huber N, Lohler J, Stoter M. The casein kinase 1 family: participation in multiple cellular processes in eukaryotes. Cell Signal 2005 Jun;17(6):675–689. PubMed PMID: 15722192.

  47. Brennan KC, Bates EA, Shapiro RE, Zyuzin J, Hallows WC, Huang Y, et al. Casein kinase idelta mutations in familial migraine and advanced sleep phase. Science translational medicine. 2013 May 1;5(183):183ra56, 1–11. PubMed PMID: 23636092.

  48. van den Maagdenberg AM, Pietrobon D, Pizzorusso T, Kaja S, Broos LA, Cesetti T, et al. A Cacna1a knockin migraine mouse model with increased susceptibility to cortical spreading depression. Neuron. 2004 Mar 4;41(5):701–710. PubMed PMID: 15003170.

  49. van den Maagdenberg AM, Pizzorusso T, Kaja S, Terpolilli N, Shapovalova M, Hoebeek FE, et al. High cortical spreading depression susceptibility and migraine-associated symptoms in ca(v)2.1 S218L mice. Ann Neurol 2010 Jan;67(1):85–98. PubMed PMID: 20186955.

  50. Leo L, Gherardini L, Barone V, De Fusco M, Pietrobon D, Pizzorusso T, et al. Increased susceptibility to cortical spreading depression in the mouse model of familial hemiplegic migraine type 2. PLoS Genet 2011 Jun;7(6):e1002129. PubMed PMID: 21731499.

    Article  CAS  Google Scholar 

  51. Bøttger P, Glerup S, Gesslein B, Illarionova NB, Isaksen TJ, Heuck A, et al. Glutamate-system defects behind psychiatric manifestations in a familial hemiplegic migraine type 2 disease-mutation mouse model. Scientific Reports. 2016 02/25/online;6:22047.

  52. Kors EE, Terwindt GM, Vermeulen FL, Fitzsimons RB, Jardine PE, Heywood P, 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 Jun;49(6):753–760. PubMed PMID: 11409427.

  53. Spadaro M, Ursu S, Lehmann-Horn F, Liana V, Giovanni A, Paola G, et al. A G301R Na(+)/K(+)-ATPase mutation causes familial hemiplegic migraine type 2 with cerebellar signs. Neurogenetics. 2004 Sep;5(3):177–185. PubMed PMID: 15459825.

  54. Santoro L, Manganelli F, Fortunato MR, Soldovieri MV, Ambrosino P, Iodice R, et al. A new Italian FHM2 family: clinical aspects and functional analysis of the disease-associated mutation. Cephalalgia. 2011;31(7):808–819. PubMed PMID: 21398422.

  55. Tottene A, Conti R, Fabbro A, Vecchia D, Shapovalova M, Santello M, et al. Enhanced Excitatory Transmission at Cortical Synapses as the Basis for Facilitated Spreading Depression in Ca(v)2.1 Knockin Migraine Mice. Neuron. 2009 MAR 12;61(5):762–73. PubMed PMID: ISI:000264366900015.

  56. Tottene A, Pivotto F, Fellin T, Cesetti T, van den Maagdenberg AM, Pietrobon D. Specific kinetic alterations of human CaV2.1 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(18):17678–17686. PubMed PMID: 15743764.

    Article  CAS  Google Scholar 

  57. Koenderink JB, Zifarelli G, Qiu LY, Schwarz W, De Pont JJ, Bamberg E, et al. Na,K-ATPase mutations in familial hemiplegic migraine lead to functional inactivation. Biochim Biophys Acta 2005;1669(1):61–68. PubMed PMID: 15843000.

  58. Tavraz NN, Durr KL, Koenderink JB, Freilinger T, Bamberg E, Dichgans M, et al. Impaired plasma membrane targeting or protein stability by certain ATP1A2 mutations identified in sporadic or familial hemiplegic migraine. Channels (Austin). 2009;3(2):82–87. PubMed PMID: 19372756.

  59. Stoica A, Larsen BR, Assentoft M, Holm R, Holt LM, Vilhardt F, et al. The alpha2beta2 isoform combination dominates the astrocytic Na(+) /K(+) -ATPase activity and is rendered nonfunctional by the alpha2.G301R familial hemiplegic migraine type 2-associated mutation. Glia. 2017;65(11):1777–1793. PubMed PMID: 28787093.

  60. Chanda ML, Tuttle AH, Baran I, Atlin C, Guindi D, Hathaway G, et al. Behavioral evidence for photophobia and stress-related ipsilateral head pain in transgenic Cacna1a mutant mice. Pain. 2013;154(8):1254–1262. PubMed PMID: 23673147.

  61. Eikermann-Haerter K, Dilekoz E, Kudo C, Savitz SI, Waeber C, Baum MJ, et al. Genetic and hormonal factors modulate spreading depression and transient hemiparesis in mouse models of familial hemiplegic migraine type 1. Journal of Clinical Investigation. 2009 JAN;119(1):99–109. PubMed PMID: ISI:000262139900014.

  62. Kros L, Lykke-Hartmann K, Khodakhah K. Increased susceptibility to cortical spreading depression and epileptiform activity in a mouse model for FHM2. Sci Rep 2018 Nov 16;8(1):16959. PubMed PMID: 30446731. Pubmed Central PMCID: 6240030.

  63. Eikermann-Haerter K, Baum MJ, Ferrari MD, van den Maagdenberg AM, Moskowitz MA, Ayata C. Androgenic suppression of spreading depression in familial hemiplegic migraine type 1 mutant mice. Ann Neurol 2009;66(4):564–568. PubMed PMID: 19847904.

  64. Shyti R, Eikermann-Haerter K, van Heiningen SH, Meijer OC, Ayata C, Joëls M, et al. Stress hormone corticosterone enhances susceptibility to cortical spreading depression in familial hemiplegic migraine type 1 mutant mice. Experimental neurology. 2015 1//;263(0):214–20.

  65. Kaufmann D, Brennan KC. The effects of chronic stress on migraine relevant phenotypes in male mice. Front Cell Neurosci 2018;12:294. PubMed PMID: 30283302. Pubmed Central PMCID: 6156251.

  66. Yapici-Eser H, Donmez-Demir B, Kilic K, Eren-Kocak E, Dalkara T. Stress modulates cortical excitability via alpha-2 adrenergic and glucocorticoid receptors: as assessed by spreading depression. Exp Neurol 2018;307:45–51. PubMed PMID: 29856967.

  67. Eikermann-Haerter K, Yuzawa I, Qin T, Wang Y, Baek K, Kim YR, et al. Enhanced subcortical spreading depression in familial hemiplegic migraine type 1 mutant mice. J Neurosci 2011;31(15):5755–5763. PubMed PMID: 21490217.

  68. Cain SM, Bohnet B, LeDue J, Yung AC, Garcia E, Tyson JR, et al. In vivo imaging reveals that pregabalin inhibits cortical spreading depression and propagation to subcortical brain structures. Proceedings of the National Academy of Sciences. 2017 February 28, 2017;114(9):2401–6.

    Article  CAS  Google Scholar 

  69. Khennouf L, Gesslein B, Lind BL, van den Maagdenberg AMJM, Lauritzen M (2016) Activity-dependent calcium, oxygen, and vascular responses in a mouse model of familial hemiplegic migraine type 1. Ann Neurol 80(2):219–232.

    Article  CAS  Google Scholar 

  70. Eikermann-Haerter K, Arbel-Ornath M, Yalcin N, Yu ES, Kuchibhotla KV, Yuzawa I et al (2015) Abnormal synaptic Ca2+ homeostasis and morphology in cortical neurons of familial hemiplegic migraine type 1 mutant mice. Ann Neurol 78(2):193–210.

    Article  CAS  Google Scholar 

  71. Vecchia D, Tottene A, van den Maagdenberg AMJM, Pietrobon D. Abnormal cortical synaptic transmission in CaV2.1 knockin mice with the S218L missense mutation which causes a severe familial hemiplegic migraine syndrome in humans. Frontiers in Cellular Neuroscience. 2015 2015-February-17;9(8). English.

  72. Vecchia D, Tottene A, van den Maagdenberg AMJM, Pietrobon D. Mechanism underlying unaltered cortical inhibitory synaptic transmission in contrast with enhanced excitatory transmission in CaV2.1 knockin migraine mice. Neurobiology of Disease. 2014 9//;69(0):225–34.

  73. van Oosterhout F, Michel S, Deboer T, Houben T, van de Ven RCG, Albus H et al (2008) Enhanced circadian phase resetting in R192Q Cav2.1 calcium channel migraine mice. Ann Neurol 64(3):315–324.

    Article  Google Scholar 

  74. Adams PJ, Rungta RL, Garcia E, van den Maagdenberg AM, MacVicar BA, Snutch TP. Contribution of calcium-dependent facilitation to synaptic plasticity revealed by migraine mutations in the P/Q-type calcium channel. Proc Natl Acad Sci U S A 2010;107(43):18694–18699. PubMed PMID: 20937883.

  75. Dilekoz E, Houben T, Eikermann-Haerter K, Balkaya M, Lenselink AM, Whalen MJ et al (2015) Migraine mutations impair hippocampal learning despite enhanced long-term potentiation. J Neurosci 35(8):3397–3402.

    Article  CAS  Google Scholar 

  76. Tremblay R, Lee S, Rudy B. GABAergic Interneurons in the Neocortex: From Cellular Properties to Circuits. Neuron. 2016 2016/07/20/;91(2):260–92.

  77. Isaacson JS, Scanziani M. How inhibition shapes cortical activity. Neuron. 2011;72(2):231–243. PubMed PMID: 22017986.

  78. Pietrobon D (2018) Ion channels in migraine disorders. Current Opinion in Physiology 2:98–108.

    Article  Google Scholar 

  79. Vecchia D, Pietrobon D. Migraine: a disorder of brain excitatory–inhibitory balance? Trends in Neurosciences. 2012 8//;35(8):507–20.

  80. Iure A, Mazzocchetti P, Bastioli G, Picconi B, Costa C, Marchionni I, et al. Differential effect of FHM2 mutation on synaptic plasticity in distinct hippocampal regions. Cephalalgia. 2019 28:333102419839967. PubMed PMID: 30922082.

  81. Aida T, Yoshida J, Nomura M, Tanimura A, Iino Y, Soma M, et al. Astroglial glutamate transporter deficiency increases synaptic excitability and leads to pathological repetitive behaviors in mice. Neuropsychopharmacology : official publication of the American College of Neuropsychopharmacology 2015;40(7):1569–1579. PubMed PMID: 25662838. Pubmed Central PMCID: 4915262.

  82. Chung DY, Sugimoto K, Fischer P, Bohm M, Takizawa T, Sadeghian H, et al. Real-time non-invasive in vivo visible light detection of cortical spreading depolarizations in mice. J Neurosci Methods 2018;309:143–146. PubMed PMID: 30194041. Pubmed Central PMCID: 6214472.

  83. Tottene A, Urbani A, Pietrobon D. Role of different voltage-gated Ca2+ channels in cortical spreading depression: specific requirement of P/Q-type Ca2+ channels. Channels (Austin) 2011;5(2):110–114. PubMed PMID: 21150292.

  84. Kager H, Wadman WJ, Somjen GG. Conditions for the triggering of spreading depression studied with computer simulations. J Neurophysiol 2002;88(5):2700–2712. PubMed PMID: 12424305.

  85. Enger R, Tang W, Vindedal GF, Jensen V, Johannes Helm P, Sprengel R, et al. Dynamics of Ionic Shifts in Cortical Spreading Depression. Cereb Cortex. 2015. PubMed PMID: 25840424.

  86. Di Guilmi MN, Wang T, Inchauspe CG, Forsythe ID, Ferrari MD, van den Maagdenberg AM, et al. Synaptic gain-of-function effects of mutant Cav2.1 channels in a mouse model of familial hemiplegic migraine are due to increased basal [Ca2+]i. J Neurosci 2014 May 21;34(21):7047–7058. PubMed PMID: 24849341. Pubmed Central PMCID: 4028489.

  87. Inchauspe CG, Urbano FJ, Di Guilmi MN, Ferrari MD, van den Maagdenberg AM, Forsythe ID, et al. Presynaptic CaV2.1 calcium channels carrying familial hemiplegic migraine mutation R192Q allow faster recovery from synaptic depression in mouse calyx of held. J Neurophysiol 2012;108(11):2967–2976. PubMed PMID: 22956801. Pubmed Central PMCID: 3544873.

  88. Eikermann-Haerter K, Arbel-Ornath M, Yalcin N, Yu ES, Kuchibhotla KV, Yuzawa I, et al. Abnormal synaptic Ca2+ homeostasis and morphology in cortical neurons of familial hemiplegic migraine type 1 mutant mice. Annals of Neurology. 2015:n/a-n/a.

  89. Fioretti B, Catacuzzeno L, Sforna L, Gerke-Duncan MB, van den Maagdenberg AMJM, Franciolini F et al (2011) Trigeminal ganglion neuron subtype-specific alterations of CaV2.1 calcium current and excitability in a Cacna1a mouse model of migraine. J Physiol 589(23):5879–5895.

    Article  CAS  Google Scholar 

  90. Chan KY, Labastida-Ramirez A, Ramirez-Rosas MB, Labruijere S, Garrelds IM, Danser AH, et al. Trigeminovascular calcitonin gene-related peptide function in Cacna1a R192Q-mutated knock-in mice. J Cereb Blood Flow Metab 2019;39(4):718–729. PubMed PMID: 28792272. Pubmed Central PMCID: 6446415.

  91. Mathew R, Andreou AP, Chami L, Bergerot A, van den Maagdenberg AM, Ferrari MD, et al. Immunohistochemical characterization of calcitonin gene-related peptide in the trigeminal system of the familial hemiplegic migraine 1 knock-in mouse. Cephalalgia. 2011 October 1, 2011;31(13):1368–80.

    Article  Google Scholar 

  92. Park J, Moon H, Akerman S, Holland PR, Lasalandra MP, Andreou AP, et al. Differential trigeminovascular nociceptive responses in the thalamus in the familial hemiplegic migraine 1 knock-in mouse: A Fos protein study. Neurobiology of Disease. 2014 2014/04/01/;64(Supplement C):1–7.

  93. Ceruti S, Villa G, Fumagalli M, Colombo L, Magni G, Zanardelli M, et al. Calcitonin gene-related peptide-mediated enhancement of purinergic neuron/glia communication by the algogenic factor bradykinin in mouse trigeminal ganglia from wild-type and R192Q Cav2.1 Knock-in mice: implications for basic mechanisms of migraine pain. J Neurosci 2011;31(10):3638–3649. PubMed PMID: 21389219.

  94. Nair A, Simonetti M, Birsa N, Ferrari MD, van den Maagdenberg AM, Giniatullin R, et al. Familial hemiplegic migraine Ca(v)2.1 channel mutation R192Q enhances ATP-gated P2X3 receptor activity of mouse sensory ganglion neurons mediating trigeminal pain. Mol Pain. 2010;6:48. PubMed PMID: 20735819.

  95. Marchenkova A, van den Maagdenberg AMJM, Nistri A. Loss of inhibition by brain natriuretic peptide over P2X3 receptors contributes to enhanced spike firing of trigeminal ganglion neurons in a mouse model of familial hemiplegic migraine type-1. Neuroscience. 2016 2016/09/07/;331(Supplement C):197–205.

  96. Marchenkova A, Vilotti S, Ntamati N (2016) Maagdenberg AMvd, Nistri a. inefficient constitutive inhibition of P2X3 receptors by brain natriuretic peptide system contributes to sensitization of trigeminal sensory neurons in a genetic mouse model of familial hemiplegic migraine. Mol Pain 12:1744806916646110.

    Article  Google Scholar 

  97. Hullugundi SK, Ferrari MD, van den Maagdenberg AMJM, Nistri A (2013) The mechanism of functional up-regulation of P2X3 receptors of trigeminal sensory neurons in a genetic mouse model of familial hemiplegic migraine type 1 (FHM-1). PLoS One 8(4):e60677.

    Article  CAS  Google Scholar 

  98. Nowodworska A, van den Maagdenberg AMJM, Nistri A, Fabbretti E. In situ imaging reveals properties of purinergic signalling in trigeminal sensory ganglia in vitro. Purinergic Signalling. 2017 2017/12/01;13(4):511–20.

  99. Franceschini A, Vilotti S, Ferrari MD, van den Maagdenberg AM, Nistri A, Fabbretti E. TNFalpha levels and macrophages expression reflect an inflammatory potential of trigeminal ganglia in a mouse model of familial hemiplegic migraine. PLoS One 2013;8(1):e52394. PubMed PMID: 23326332. Pubmed Central PMCID: 3543418.

  100. Thomsen LL, Kruuse C, Iversen HK, Olesen J. A nitric oxide donor (nitroglycerin) triggers genuine migraine attacks. Eur J Neurol 1994 Sep;1(1):73–80. PubMed PMID: 24283432.

  101. Tassorelli C, Joseph SA. Systemic nitroglycerin induces Fos immunoreactivity in brainstem and forebrain structures of the rat. Brain Res 1995;682(1–2):167–181. PubMed PMID: 7552308.

  102. Bates EA, Nikai T, Brennan KC, Fu YH, Charles AC, Basbaum AI, et al. Sumatriptan alleviates nitroglycerin-induced mechanical and thermal allodynia in mice. Cephalalgia. 2010;30(2):170–178. PubMed PMID: 19489890. Pubmed Central PMCID: 4854191.

Download references

Acknowledgements

None.

Funding

Telethon Italy (GGP14234) to D.P.

NIH/NINDS (NS085413 and NS102978) to K.C.B.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization: D.P. and K.C.B. Writing-Review and editing: D.P. and K.C.B. Funding acquisition: D.P. and K.C.B. Both authors read and approved the final manuscript.

Corresponding authors

Correspondence to Daniela Pietrobon or K. C. Brennan.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

Not applicable.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pietrobon, D., Brennan, K.C. Genetic mouse models of migraine. J Headache Pain 20, 79 (2019). https://doi.org/10.1186/s10194-019-1029-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s10194-019-1029-5

Keywords