Skip to main content

Genetics of migraine: where are we now?

Abstract

Migraine is a complex brain disorder explained by the interaction of genetic and environmental factors. In monogenic migraines, including familial hemiplegic migraine and migraine with aura associated with hereditary small-vessel disorders, the identified genes code for proteins expressed in neurons, glial cells, or vessels, all of which increase susceptibility to cortical spreading depression. The study of monogenic migraines has shown that the neurovascular unit plays a prominent role in migraine. Genome-wide association studies have identified numerous susceptibility variants that each result in only a small increase in overall migraine risk. The more than 180 known variants belong to several complex networks of “pro-migraine” molecular abnormalities, which are mainly neuronal or vascular. Genetics has also highlighted the importance of shared genetic factors between migraine and its major co-morbidities, including depression and high blood pressure. Further studies are still needed to map all of the susceptibility loci for migraine and then to understand how these genomic variants lead to migraine cell phenotypes.

Peer Review reports

Migraine, a complex genetic condition

The goal of genetics is to identify key proteins in order to better understand the pathophysiology of a disease, to define new therapeutic targets and to find diagnostic biomarkers. Migraine is a highly disabling, complex brain disorder with a strong familial aggregation. Twin and family studies conducted in the 1990s demonstrated the existence of hereditary factors in migraine [1, 2]. In these studies, the estimated heritability of migraine ranged from 35% to 60%. In population-based studies, the relative risk of migraine for a first-degree relative of an index case was 1.5- to 4-fold compared with the general population [3]. The risk was higher for relatives of cases with higher pain scores and attack frequency, early age of onset, and migraine with aura (MwA).

More recent studies estimate the heritability of migraine to be about 42%. They also reinforce the idea that migraine is a complex disease resulting from interactions between genes and the environment, interactions between genes themselves, and as yet unknown factors [3]. Heritability is higher in MwA than in migraine without aura (MO) [4].

Migraine is predominantly polygenic, with multiple genetic variants, each with a minor-effect size, accumulating to lead to the disease. A portion of MwA cases could be explained by the conjunction of a small number of genetic variants with moderate effect size, or by a single variant with a major functional effect as in monogenic migraines [5] (Fig. 1). In these much rarer disorders, a pathogenic mutation in a single gene is sufficient to produce the disease with almost complete penetrance. The classical example of monogenic migraine is familial hemiplegic migraine (FHM), which is inherited in an autosomal dominant fashion [6]. Migraines can also be part of the clinical spectrum of other hereditary neurological conditions, such as cerebral autosomal dominant arteriopathy with subcortical infarcts and leukoencephalopathy (CADASIL). Other examples will be discussed further in this paper.

Fig. 1
figure 1

Overview of migraine and genetics. An overview of the complex genetic architecture of migraine, from polygenic model on the left, to monogenic model on the right. Created with BioRender.com

Familial hemiplegic migraine (FHM), a monogenic form of migraine

Genetic heterogeneity and clinical variability

Hemiplegic migraine (HM) is a rare disease with an estimated prevalence of 0.01% in the general population [6, 7]. Familial HM, diagnosed when at least one first- or second-degree relative also has HM, accounts for two-third of the cases. Sporadic HM (SHM), diagnosed in the absence of family history, accounts for one-third of cases. HM attacks begin during youth (mean age of onset 12–17 years old), and comprise motor weakness during the aura, always associated with at least one other symptom of typical aura (visual, sensory, speech and language) and often with brainstem aura symptoms (70%) [8, 9]. The frequency of attacks varies from more than one per week to a few over the course of a lifetime, with an average of 3 to 4 per year [10]. Duration of HM aura is often longer than that of typical aura (several days to weeks) [11,12,13]. Severe attacks with confusion, coma, fever, seizures and reversible brain edema may occur [14,15,16,17], sometimes triggered by mild head trauma [18,19,20]. HM can be pure or associated with a combination of early-onset epilepsy, cerebellar ataxia, learning disabilities, and/or mental retardation, which may begin before or after HM onset [21,22,23,24,25].

FHM is primarily a monogenic disorder, with an autosomal dominant pattern of inheritance and high penetrance; 70–90% of individuals with a pathogenic mutation clinically express the disease. FHM is genetically heterogeneous and is subdivided into FHM1, FHM2 and FHM3, based on the presence of mutations in the CACNA1A, ATP1A2 and SCN1A genes, respectively [6, 26,27,28] (Table 1, Fig. 2). The PRRT2 gene should be added to the main FHM genes because recent data have shown that PRRT2 is involved at least as frequently as SCN1A [29, 30]. For convenience, we will therefore refer to FHM4 for HM associated with PRRT2 mutations. Other genes have been reported in a small number of cases and families, and additional data are needed before they can be considered causal (Table 2).

Table 1 The four major genes implicated in familial hemiplegic migraine
Fig. 2
figure 2

Genetics of familial hemiplegic migraine (FHM). Glutamatergic synapse of the central nervous system with proteins encoded by genes involved in familial hemiplegic migraine and their functional roles. Created with BioRender.com

Table 2 Other genes potentially implicated in familial hemiplegic migraine (FHM)

Clinical studies of patients with mutations in the four major FHM genes have shown that attacks of HM are similar regardless of the gene involved, and that prolonged auras with confusion are possible in all FHM types. The association of HM with epilepsy is present in 7% of overall HM patients [7, 49], but seems more frequent in FHM2 [50].

Conversely, different mutations in the same gene can influence the phenotype. In FHM1, the two mutations most commonly involved in severe attacks with coma and fever are T666M and S218L [17]. Moreover, the nature of the mutated gene also influences the spectrum of manifestations associated with HM attacks [6]. Febrile comas are frequent in FHM1 (up to 30%), possible in FHM2 (up to 15%), and have not been described in FHM3 and FHM4 [17, 51]. Cerebellar ataxia is common in FHM1 [52,53,54,55,56]; a phenomenon of repeated transient blindness was observed only in FHM3 [57]; mental retardation has been described in FHM1, 2 and 4 [22,23,24]; and finally, the association of HM with paroxysmal dyskinesia or hypersomnia is suggestive of FHM4 [30]. Finally, there is great variability in HM attacks and associated manifestations between individuals who carry different mutations in the same gene, and even between affected family members who carry the same mutation. This variability suggests that other genetic or environmental factors can modulate the clinical phenotype [26].

FHM1 and CACNA1A mutations

CACNA1A, localized on 19p13, was the first identified HM gene [58]. It encodes the main α1 pore-forming subunit of the neuronal voltage-gated calcium channels CaV2.1 or P/Q. These channels are expressed in synaptic endings in the brain and the cerebellar, and play a role in controlling neurotransmitter release. More than 25 CACNA1A mutations have been identified in FHM1. Most are missense mutations resulting in a gain of function, which increases Ca2+ influx, glutamatergic neurotransmission and neuronal excitability [59].

There are two transgenic FHM1 knock-in (KI) mouse models [60, 61]. KI mice for the R192Q mutation, which causes pure FHM1, show no clinical abnormalities. KI mice for the S218L mutation, which causes very severe FHM1, show cerebellar ataxia, transient hemiparesis and epilepsy. FHM1-KI mice exhibit increased CaV2.1 currents and neurotransmitter release, loss of balance between excitatory and inhibitory cortical neurotransmissions, enhanced cortical excitatory transmission in visual cortex [62] and increased susceptibility to cortical spreading depression (CSD). These transgenic mice have also been shown to exhibit head pain [63], increased trigeminal activity, tissue anoxia during prolonged aura, increased sensitivity to cerebral ischemia, and altered trigeminal nociception mediated by CGRP [64].

The CACNA1A gene is also mutated in other neurological disorders. Episodic ataxia type 2 (EA2), characterized by paroxysmal ataxia, dizziness and nausea, is associated with CACNA1A mutations responsible for loss of function and decreased Ca2+ influx [65]. Spinocerebellar ataxia type 6 (SCA6), characterized by progressive cerebellar ataxia, is caused by an expansion of a CAG repeat in the terminal portion of CACNA1A, which results in toxic degeneration of cerebellar Purkinje cells [16].

FHM2 and ATP1A2 mutations

The ATP1A2 gene on 1q23.2 encodes the α2 isoform of the catalytic subunit of the A1A2 ATP-dependent transmembrane pump (α2 Na+/K+-ATPase) [66]. In the CNS of adults, this pump is primarily expressed in astrocytes, where it provides extracellular K+ clearance and produces a Na+ gradient necessary for glutamate reuptake from the synaptic cleft. More than 80 ATP1A2 mutations have been identified in FHM2. Missense mutations are the most common, but small deletions, a stop-codon altering mutation, and an exonic duplication have also been reported. These mutations result in a variable loss of function of the α2 Na+/K+-ATPase pump. The mutated pumps are reported to have lower glutamate uptake, slowing down recovery from neuronal excitation and promoting excitatory cortical transmission, thereby facilitating the initiation of CSD waves. There are several models of FHM2-KI transgenic mice. Heterozygous transgenic mice show no clinical abnormalities but have increased susceptibility to CSD [67, 68]. Mice with partially knock-out (KO) of ATP1A2 also show increased susceptibility to CSD [69]. Another mouse model with complete KO of ATP1A2 in astrocytes showed episodic paralysis and spontaneous waves of CSD with decreased EEG activity [70]. These animals had abnormalities in brain metabolism with increased serine and glycine. A serine- and glycine-free diet suppressed attacks of paralysis in these mutants.

FHM3 and SCN1A mutations

SCN1A on 2q24.3 encodes the α1 subunit forming the pore of NaV1.1 channels [71]. These voltage-dependent neuronal sodium channels are involved in the genesis and propagation of action potentials in cortical neurons, particularly in GABAergic inhibitory interneurons [72]. SCN1A was already known as an epilepsy gene with over 100 missense and nonsense mutations identified in various forms of childhood epilepsies. A dozen SCN1A mutations, mainly missense variants leading to a gain of function, have been identified in FHM3. Their functional consequences are complex [73, 74]. The mouse model carrying the L1649Q variant showed an increased susceptibility to CSD. The L1649Q mutation results in a defect in Na + channel inactivation with increased Na + currents and hyperactivity of inhibitory interneurons.

FHM4 and PRRT2 mutations

PRRT2 encodes the PRRT2 protein, which plays an important role in brain development, synapse formation, and neurotransmitter release. PRRT2 is expressed in presynaptic terminals and interacts with proteins of the exocytosis complex. Mutations in PRRT2 have now been identified in several dozen cases of FHM4, two-thirds of which have pure FHM, and one-third of which have FHM associated with epilepsy, mental retardation or dyskinesia [21, 75,76,77,78,79,80]. Mutations in PRRT2 are also associated with several other neurological diseases, including benign familial infantile epilepsy (BFIE), infantile seizure syndrome with choreoathetosis (ICCA) and paroxysmal kinesigenic dyskinesia (PKD) [81, 82].

The different mutations in PRRT2 (point duplication, small deletions, missense, total deletions) all induce a loss of function leading to haploinsufficiency. A given PRRT2 mutation can be associated with several diseases. Indeed, the c.649dupC mutation is common in FHM4, but is also the main causative mutation in PKD and BFIE.

PRRT2-KO mice exhibit paroxysmal abnormal movements upon acquisition of locomotion, develop abnormal audiogenic motor behaviors in adulthood, and have a lowered seizure threshold [83]. Their excitatory hippocampal neurons display increased excitability. Human and murine homozygous KO-PRRT2 neurons in culture express overactive NaV1.2 and NaV1.6 channels, indicating that PRRT2 inhibits voltage-gated sodium channels.

Further experiments are needed to understand the factors underlying the great phenotypical variability associated with PRRT2 mutations, and the potential influence of modifier genes or of the non-mutated allele.

Other potential FHM genes

Mutations in several other genes have been identified in HM (Table 2) [29]. All these genes were already known to be involved in other inherited diseases. In a large cohort of index HM cases from New-Zealand and Australia, analysis of potential new HM genes increased the diagnosis rate from 21% to 27,8% (PKND 0,4%; SLC4A4 3,9%; ATP1A3 0,9%; SLC1A3 0,4%; SLC2A1 1,3%) [29]. Analysis of other large cohorts of index cases, as well as functional studies assessing CSD in animal models would be important to confirm that these genes actually cause FHM.

Genetic architecture of hemiplegic migraine

Among index cases suspected of having HM and referred for genetic diagnosis, a minority has a mutation in one of the four major genes, 15% in a French cohort of 697 patients [78], and 21% in a New Zealand and Australian cohort of 230 patients [29]. These two independent studies yielded similar results, with the most frequent mutations found in the ATP1A2 gene (6,3–10%) followed by CACNA1A (3,7–7%), PRRT2 (2,2-3,5%) and finally, SCN1A (1,7 -2,1%). In contrast, a Dutch study of a cohort of 301 patients found higher rates of mutations in major genes: CACNA1A in 107/301 (35.5%), ATP1A2 in 75/301 (24.9%), SCN1A in 26/301 (8.6%), and PRRT2 in 1 /47 (2.1%) [21]. In addition, only three mutations were identified in a Finnish cohort of 293 HM patients: one in CACNA1A (0.34%) and two in ATP1A 2(0.68%) [31]. The PRRT2 gene was not screened in the Finnish study [31]. These differences could be due to different recruitment methods of the cohorts.

In typical HM cases in whom there are no mutations in the four main genes, additional single-gene variants may be identified by future systematic studies, such as exome studies, and full genome sequencing. New variants with large-effect sizes are expected to be involved in only a small proportion of familial and sporadic cases. In other cases of HM, the inheritance may be polygenic, involving multiple variants with each a small-effect size, or oligogenic, with a combination of one or few variants with a moderate-effect size with or without multiple pro-migraine variants of small-effect size.

SHM, diagnosed in the absence of any affected relative, can result from a de novo mutation of one of the FHM genes in a subject whose two parents do not have a mutation [47, 84,85,86,87,88]. These de novo mutations can be passed to offspring, transforming SHM into FHM. SHM can also result from mutations in known FHM genes with low penetrance, mosaicism in the transmitting parent, or pathogenic variants in as yet unknown genes. Other SHM types might have a different mode of inheritance, either recessive with compound heterozygotes or polygenic [26]. Finally, environmental and psychosocial factors including exposure to stress, psychological and physical trauma, abuse, or negative life events may also play an important role in SHM. According to the US military personnel HM cohort, the incidence of HM was zero from 1997 to 2007, and then steadily increased, with a 25-fold increase in new cases between 2008 and 2017 [89].

Links between FHM and the common varieties of migraine

FHM is a model of hereditary severe migraine with aura. Some authors have suggested that the mechanisms of FHM, namely increased sensitivity to CSD, may be involved in common migraine, with and without aura. Danish studies have shown that the risk of migraine with typical aura (eg without motor deficit) was increased in individuals with FHM compared with the general population, whereas the risk of MO was similar [90, 91]. Thus, FHM is a major model for migraine aura associated with cortical excitability, with subsequent headache triggered by CSD, and FHM genes do not play a major role in the genetics of the common varieties of migraine [92, 93]. A recent study has identified a polymorphism in the FHM1 CACNA1A gene as a susceptibility locus for common varieties of migraine, among 122 other loci [94].

Other monogenic varieties of migraine

Monogenic migraine with aura and TRESK mutations

TRESK is a two-pore K+ channel responsible for maintaining membrane excitability. By a free flow of K+ ions, it contributes to the formation of leakage currents in the trigeminal ganglion and dorsal root ganglia. It is therefore though to play a role in pain processing mechanisms. Mice with a functional knock-out of TRESK show a ‘painful’ behavioural phenotype, and exhibit hyperexcitability of the dorsal root and trigeminal ganglia. In addition, in the trigeminal ganglion, TRESK expression is restricted to nociceptive neurons [5, 95].

A frameshift mutation in KCNK18, which encodes the TRESK channel, was described in a large family with visual MwA following autosomal dominant inheritance [96]. All family members with migraine carried the p.(F139Wfs*24) mutation, which has been shown to exert a dominant negative effect resulting in complete loss of TRESK function and increased neuronal excitability [97].

The causal link between TRESK mutations and migraine has been called into question by the discovery of another mutation with a dominant negative effect, C110R, in individuals without migraine [98]. However, further research revealed that the p.(F139Wfs*24) variant introduces an alternative start codon that shortens the TRESK protein and damages its function resulting in hyperexcitability of nociceptors. Such effects were not observable for C110R [97, 99]. Another missense mutation in the KCNK18 gene (W101R) was identified in a 12-year-old male with migraine with brainstem aura and intellectual disability. This variant was inherited from his mother, who had migraine with aura [100]. Further investigations revealed impaired TRESK channel function associated with this variant [101]. In another study, pharmacological inhibition of TRESK influenced the response to capsaicin, a TRPV1 receptor agonist, and resulted in increased CGRP release and meningeal blood flow [102]. These data suggested that MwA can be caused by a rare genetic variant inducing a major functional effect.

A study of 200,000 exome-sequenced UK Biobank participants conducted in 2022 found the frameshift variant p.(F139Wfs*24) in 196 (0.10%) of the 193,433 participants classified as controls and in 10 (0.14%) of the 7194 migraine cases (p = 0.33) [103]. The authors concluded that KCNK18 should no longer be regarded as being involved in migraine etiology. A major limitation of this study is that the clinical status of healthy controls was not assessed in details. This may imply that a proportion of the 193,433 participants may in fact have been affected by migraine, as mis- and underdiagnosis of migraine and migraine aura is highly common [104]. Studies of cohorts of patients with a firm diagnosis of MwA and well characterized healthy controls are needed to further elucidate the role of KCNK18 and its product, TRESK in migraine.

Familial advanced sleep-phase syndrome (FASP), migraine and CSNK1D mutations

FASPS, which causes an extreme tendency to wake-up early in the morning, can be caused by mutation in a circadian clock gene, CSNK1D, which codes for casein kinase 1 delta (CKIδ). In the two large families with a CKIδ mutation, the sleep disorder was associated with migraine [105]. Transgenic mice expressing T44A variant of CSNK1D displayed a high propensity for nitroglycerin-induced mechanical hyperalgesia, and a reduced threshold for CSD. These findings suggest that migraine may be caused by a mutation in a gene that encodes neither an ion channel nor a protein involved in glutamate signaling. In addition, the link between migraine and FASPS is consistent with the known role of hypothalamus in migraine [106, 107].

ROSAH syndrome, migraine and ALPK1 mutations

ROSAH syndrome (retinal dystrophy, optic nerve edema, splenomegaly, anhidrosis, migraine headache) is an autosomal dominant condition caused by a missense mutation in the ALPK1 gene which was identified in five families. ALPK1 encodes Alpha Kinase 1, which plays a role in inflammation, cellular trafficking, and possibly also affects CGRP activity [26, 108]. Its role in migraine is unknown.

Monogenic cerebral vasculopathies and migraine

CADASIL (Cerebral autosomal dominant arteriopathy with subcortical infarcts and leukoencephalopathy) is the commonest type of familial cerebral small-vessel disease responsible for recurrent lacunar stroke, leading to dementia and premature death [27, 109]. CADASIL is caused by NOTCH3 mutations, which cause progressive destruction of vascular smooth muscle cells [110]. Migraine is very common in CADASIL (up to 75%), mainly migraine with typical aura, but also hemiplegic migraine or migraine without aura. Migraine is often the first manifestation of the disease, 15 to 20 years before cerebral infarctions [111, 112]. Mutant mice, either knock-in for a CADASIL mutation or NOTCH3-knock-out, have shown increased sensitivity to CSD [113].

Retinal vasculopathy with cerebral leukodystrophy (RVCL) caused by mutations in TREX1, and disorders due to COL4A1 and COL4A2 mutations are other small-vessel diseases that frequently involve migraines [27].

The study of these conditions, especially CADASIL, showed that gene expressed only in vessels could be implicated in migraine, which was later on confirmed by studies in polygenic migraines.

Susceptibility genes for migraine with aura and migraine with aura

Genome-wide association studies (GWAS)

Identification of gene variants involved in migraine has proven difficult, and there have been 30 years of studies without significant results. Because MO and MwA display strong familial aggregation, which may suggest Mendelian inheritance, the initial hope was that the techniques used successfully in FHM would identify the genes for the most common migraines. Initial studies showed that the FHM genes were not involved in the common varieties of MO and MwA. Linkage analysis studies identified dozens of loci that were presented as possibly containing genes involved in migraine, but these were never discovered [5, 114].

Researchers turned to genome-wide association study (GWAS), which examines millions of polymorphisms called SNPs (single nucleotide polymorphisms) in very large cohorts of patients and healthy controls. Each SNP is a variation in the genetic code at a single DNA base pair. More than 100 million SNPs exist in the human genome, and 4–5 million SNPs are distributed throughout an individual genome [115]. A GWAS identifies SNPs that are significantly associated with the disease of interest, by assessing differences in allele frequencies between large numbers of patients and controls. For each SNP, the level of significance is very difficult to reach (5 × 10− 8) because of the multiplicity of tests performed.

In 2010, the first GWAS identified a single migraine susceptibility locus [116] (Table 3) Over the past decade, the International Headache Genetics Consortium (IHGC; www.headachegenetics.org/) has conducted several migraine GWAS, and with increasing sample sizes, the number of associated genetic variants has progressively increased [117,118,119]. The 2016 migraine GWAS, including 59,674 migraine sufferers and 316,078 controls, identified 38 distinct genomic loci associated with migraine [120]. Tissue expression enrichment analyzes clearly demonstrated the enrichment in genes involved in arterial and smooth-muscle function [120, 121]. The other pathways identified were the neuronal pathway [122], and the pathway related to homeostasis of iron ions and other metals [120].

Table 3 Genome wide association studies in migraine from 2010 to 2022

In a third recent migraine GWAS from 2021, 79 independent loci were significantly associated with migraine [127]. Of note, this was an ethnically diverse study that included adult individuals (28,852 cases vs. 525,717 controls) from East Asian, African American, and Hispanic/Latino descent.

The most recent migraine GWAS published in 2022 by Hautakangas et al. [94] included 102,084 cases and 771,257 controls and identified 123 distinct loci associated with migraine, of which 86 were novel compared to the 2016 GWAS. Additional analyses even increased the number of independent SNPs to 167. Enrichment analyses in the 2022 migraine GWAS clearly pointed to both vascular and central nervous system tissues and cell types. The newly identified loci involve genes encoding known migraine drug targets, namely calcitonin gene-related peptide (CGRP, encoded by CALCA/CALCB),) and serotonin 1F receptor (HTR1F). The former is the target for CGRP antibodies, and the latter for ditans. In addition, an analysis of about 30,000 patients from the 2022 GWAS with a precise diagnosis of the type of migraine (eg, MO or MwA) showed that three risk variants were specific for MwA (including a SNP in CACNA1A the FHM1 gene), two were specific for MO and nine were associated with both types.

Given that some risk loci were found in the Hautakangas 2022 GWAS only, some in the Choquet 2021 GWAS only, and some in both studies, there are now about 180 migraine risk loci.

In addition to these large GWAS including mainly cases with European ancestries, other smaller GWAS conducted in Asia replicated some of the results obtained in European cases and yielded other new SNPs [128]. Another GWAS study conducted in Asian population identified eight novel susceptibility loci correlated with age of migraine onset [129].

Altogether, migraine GWAS have identified more than 180 low-effect-size genetic variants all across the genome, with enrichment in vascular and neuronal cells/tissues, confirming that migraine is a polygenic neurovascular disorder. Recent GWAS demonstrated that all migraine varieties share common molecular mechanisms, and that MO and MwA have specific genetic risk factors and distinct mechanisms. Due to the large size of the samples, clinical data in GWAS were limited and did not permit to study other migraine varieties, such as “pure” MwA (patients having only attacks of MwA and never MO) or chronic MO.

Future studies will have to determine which are the causal genes modified by the SNPs, which are mostly located in non-coding regions, and do not necessarily affect the closest gene. A first challenging step will be to select the list of most-likely causal genes based on GWAS results. For example, the intronic SNP rs9349379 near PHACTR1 is a proven risk loci for migraine, coronary artery disease, fibromuscular dysplasia, hypertension and cervical artery dissection [130]. A functional study showed that this SNP had no effect on PHACTR1 but on the gene encoding endothelin-1 (ET-1; EDN1), a strong vascular smooth muscle cells constrictor located 600 kB upstream of the risk SNP [130]. More recent data suggested that SNP rs9349379 may in fact regulate the expression of PHACTR1, and not EDN1, and that PHACTR1 could have a role in arterial compliance [131]. This debate on a single SNP shows that enormous amounts of experiments will be necessary to study the 180 migraine SNPs.

In addition, some variants not identified by the mean of GWAS could also be implicated in migraine susceptibility through gene-gene interactions. A case-control study suggested that synergetic effects between a variant in NRXN2, coding a component of the synaptic vesicle machinery, and two other genes, GABRE and CASK, were associated with migraine [132].

Polygenic risk score and genetic architecture of migraine

The almost 200 variants identified by the latest GWAS each explain only a small fraction of the genetic risk, and their sum do not explain the full heritability of migraine. The Polygenic Risk Score (PRS) assesses the individual genetic risk of migraine as the sum of all SNPs and alleles that increase the risk of migraine carried by an individual. The PRS may be used to analyze the genetic links between the different primary headaches and the different forms of migraine, and between migraine and its comorbidities. The PRS can also assess pharmacogenetic effects [5].

In a study of 1589 migraine sufferers from the Finnish population, familial cases had a significantly higher PRS than non-familial cases [133]. The genetic burden was higher in MwA and FHM compared to MO, and was associated with an earlier age of onset of migraine. These data show that migraine is primarily due to an accumulation of minor variants that produce a favorable (pro-migraine) genetic background, and not to highly deleterious single gene mutations. Noteworthy, a recent study using a PRS derived from 38,872 variants associated with migraine in 8602 subjects in Finland showed a correlation between PRS and migraine diagnoses according to ICHD3 criteria [134]. Non-headache, non-migraine headache, probable migraine, migraine headache, migraine with typical visual aura and hemiplegic migraine formed a continuum along the increasing PRS, which paves the way for the potential concept of genetic classifications [134].

The known small-effect-size pro-migraine variants do not explain the full heritability of migraine. The “missing heritability” may be explained by several hypothesis. First, there are probably hundreds of other small-effect-size variants, which increase the risk of migraine but fall below the required levels of significance in GWAS. Second, persons with multiple disease associated SNPs may have an additive effect conferring a greater overall risk. Third, technical limitations in short read sequencing and Sanger analyses may account for part of this “missing heritability”. Finally, variants with minor or moderate-effect-size probably explain a part of the missing heritability, and cannot be identified by GWAS. A study sequenced the genomic areas associated with migraine in a large cohort of patients and identified four rare variants altering regulatory areas close to four variants discovered by GWAS [135]. Another study analyzed RNA sequencing using a coexpression network of aorta, trigeminal ganglion and visual cortex, combined with a whole sequence genoming. The authors identified a ‘gene module’, a set of coexpressed genes, in the visual cortex that had increased mutations in migraine. Pathway analysis of this module revealed association with hormonal signaling, Alzheimer’s disease, serotonin receptors and heterotrimeric G protein signaling pathway. Noteworthy, mutations in two genes involved in glutamate signalling, CACNA1B and ATXN1, were found in several migraine families [136]. Using whole-exome sequencing in small populations, new SNPs have been associated with responsiveness to verapamil as a preventive therapy [137], and neurological outcome, including migraine, after head trauma [138]. Finally, “private” large-size-effect variants may be identified by chance, such as in the very rare families carrying mutations of KCNK18 or CSNK1D. In these families, the strong penetrance of the migraine phenotype could result from the cosegregation of the “private” large-size-effect mutation and a pro-migraine genetic background due to a high PRS, whereas most carriers of the same rare variant would not express a migraine phenotype thanks to a non-permissive genetic background. Some authors have even hypothesized that FHM would not be truly autosomal dominant but the result of a rare mutation on a pro-migraine background.

The genetics of migraine thus seems very complex, based on the interaction of hundreds of common small-effect-size variants with rare variants affecting regulatory areas, and with possible “private” moderate to large-effect-size variants.

Shared genetic background of migraine and its comorbid diseases and traits

The comorbidities of migraine are diseases whose prevalence is increased in migraine sufferers compared to controls and for which certain pathological mechanisms could be shared (Fig. 3). Genetics is a tool to explore some of the common mechanisms, firstly by identification of genes associated with both conditions (GWAS), secondly, by estimation of the genetic correlation, namely the proportion of variance the two conditions share due to genetics (genetic correlation studies), and finally, by evaluation of causal relationships between two conditions by using genetic variants as proxies of an exposure (Mendelian randomization [MR] studies) [5, 139].

Fig. 3
figure 3

Shared genetic background of migraine and its comorbid diseases. Genetic relation of migraine and some of its clinically most relevant comorbidities. Dotted arrow: Genetic association or correlation as demonstrated by GWAS or genetic correlation studies. Solid arrow: Causal association of genetic variants as demonstrated by Mendelian randomization studies. +, liability to one disease increases risk for the comorbidity; −, liability to one disease decreases risk for the comorbidity; ADHD, attention deficit hyperactivity disorder; BP, blood pressure; CAD, coronary artery disease; CeAD, cervical artery dissection. Created with BioRender.com

With regard to genetic studies of associations and correlations, several large studies based on the comparison of GWAS data have shown the existence of a shared genetic susceptibility between migraine and various disorders, including psychiatric disorders [140, 141], ischemic stroke [142], coronary artery disease [143], hypertension [144, 145], sleep disorders [146], and also endometriosis [147], fibromuscular dysplasia [148], type 2 diabetes, hyperlipidemia, autoimmune diseases, asthma, other respiratory conditions [144], restless legs syndrome [149] and hemostatic profile [150]. In contrast, one study found no causal link between genetic susceptibility to migraine and Alzheimer’s disease, intelligence, and brain size [151].

Mendelian randomization studies have provided evidence of causal associations between genetic variants predisposing for migraine and those predisposing for some of the above-mentioned conditions. Findings of Mendelian randomization studies on migraine and comorbid conditions are summarized in Table 4.

Table 4 Mendelian randomization studies for causal associations of genetic variants predisposing for migraine and its comorbid disorders and traits

Other recent genetic findings

Genetics of headache

A 2018 British GWAS studied 74,461 individuals who had had a headache interfering with daily activities in the previous month and 149,312 controls [170]. The majority of patients probably had tension headache and less often migraine. This study identified 28 headache susceptibility loci of which 14 had already been identified by GWAS in migraine, and 14 were new. The majority of the potential headache genes were neuronal and not vascular. This study also found a shared genetic background between the headache phenotype and many psychological traits associated with vulnerability to depression and negative emotions, highlighting the importance of links between psychiatric conditions and painful conditions.

Another recent GWAS including 2084 Taiwanese patients and 11,822 age- and sex-matched controls identified two loci, rs10493859 in TGFBR3 and rs13312779 in FGF23, both functionally relevant to vascular function and migraine, to be significantly associated with self-reported headache [171].

Until recently, studies on genetics of Cluster headache (CH) have been dominated by candidate gene studies with conflicting findings [172, 173]. A first Italian GWAS on 99 patients and 360 controls identified ADCYAP1R1 and MME gene variants as possibly associated with susceptibility for CH [174]. These findings were not replicated in a larger Swedish cohort [175]. In 2021, two GWAS out of which one included Dutch cases (n = 840) and controls (n = 1457) and Norwegian cases (n = 144) and controls (n = 1800) [176], and the second one UK cases (n = 852) and controls (n = 5614) and Swedish cases (n = 591) and controls (n = 1134) [177] independently identified four risk loci for CH on chromosome 1, chromosome 2 (two loci), and chromosome 6, respectively. Subsequently, a meta-analysis of both studies analyzing 8,039,373 variants confirmed a significant association of all 8 index variants (in the 4 loci) and identified three additional loci with genome-wide significance on chromosomes 7, 10 and 19. The nearest genes to the loci on chromosome 2 and 6 are MERTK and UFL1/FHL5, respectively. Interestingly, as stated in Table 3, UFL1/FHL5 has previously been identified as a migraine risk locus [120].

Genetics of chronic migraine

Chronic migraine is the most disabling form of migraine, and causes for migraine chronification remain incompletely understood. In order to identify genetic variants contributing to migraine chronification, a comparison of patients with chronic migraine and patients with episodic migraine is necessary, whereas most studies attempting to find genetic risk factors have compared chronic migraineurs to healthy controls. Recent studies comparing chronic and episodic migraine have found genetic variants in the TRPM8 gene [178], the TRPV1 gene [179], and HLA class I alleles [180] to be associated with chronic migraine.

Previous to these studies, a candidate gene-association study examined 144 SNPs from 48 candidate genes in patients with chronic or high-frequency episodic migraine compared to healthy controls, and did not reveal any significant findings [181].

The first study assessing whole-genome sequencing data in patients with chronic compared to episodic migraine did not show any significant difference [182].

Further studies are needed to determine the proportion of genetic and environmental factors in chronic migraine.

Genetics factors underlying treatment response

Genetic factors strongly influence the absorption, distribution, metabolism and excretion of drugs. Studies addressing genetic factors underlying treatment response to triptans have described GNB3 C825T gene polymorphism to be associated with a better response to triptans in CH patients, and polymorphisms in the PRDM16, SLC6A4 and DRD2 genes to be associated with a better, inconsistent and worse response to triptans in migraine patients, respectively [183]. Another study showed that a polygenic risk score doubling the risk of migraine was associated with a better response to triptans [184].

Recently, whole-exome sequencing in a discovery cohort of migraine patients treated by verapamil (definitive responders n = 21 and definitive non-responders n = 14), followed by genotyping in a confirmation cohort (n = 185), identified 13 SNP, which were highly correlated with the changes in the number of migraine days [137].

In the future, determining the genetic profile of an individual could allow the choice of treatments with the best profile of efficacy and tolerance [5].

Mitochondrial DNA and migraine

Mitochondrial dysfunction has been suspected to contribute to migraine pathophysiology, since migraine-like headache is a clinical feature of several mitochondrial diseases [185,186,187], presumable mitochondrial biomarkers have been found to be elevated in migraine patients [188], and several studies have reported mitochondrial DNA (mtDNA) candidate variants possibly associated with migraine [189, 190]. However, the first GWAS assessing 775 mitochondrial DNA variants in 4021 migraine sufferers and 14,288 controls found no migraine-associated variants, ruling out the mitochondrial hypothesis suggested by older studies [191]. Limitations discussed by the authors were the diagnosis of migraine based on a questionnaire covering symptoms during the past 12 months instead of a clinical interview, and the absence of consideration of heteroplasmic variation, copy-number variations and epigenetic changes.

Conclusion

Genetics of migraine have made significant progress over the past 15 years [5, 28, 114]. The study of monogenic migraines identified key proteins of the susceptibility to CSD and helped to better appreciate the links between migraine and vascular disorders. GWAS have identified multiple susceptibility genes revealing several complex networks of “pro-migraine” molecular abnormalities, mainly neuronal and vascular (Fig. 1). Genetics has also underscored the importance of genetic factors shared between migraine and its major co-morbidities including depression and high blood pressure. Very large-scale studies are still needed to map all of the susceptibility loci to migraine and then to understand how these genomic variants lead to migraine cell phenotypes. Ultimately, the main pathophysiological mechanism in a given patient, neuronal or vascular or otherwise, could be determined through its genetic risk profile. Pharmacogenetics could help predict the therapeutic response and thus help prescribe the treatment with the best safety and efficacy profile.

References

  1. Russell MB, Hilden J, Sørensen SA et al (1993) Familial occurrence of migraine without aura and migraine with aura. Neurology 43:1369–1373

    Article  CAS  PubMed  Google Scholar 

  2. Russell MB, Olesen J (1995) Increased familial risk and evidence of genetic factor in migraine. BMJ 311:541–544

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Polderman TJC, Benyamin B, de Leeuw CA et al (2015) Meta-analysis of the heritability of human traits based on fifty years of twin studies. Nat Genet 47:702–709

    Article  CAS  PubMed  Google Scholar 

  4. Russell MB, Ulrich V, Gervil M et al (2002) Migraine without aura and migraine with aura are distinct disorders. A population-based twin survey. Headache 42:332–336

    Article  PubMed  Google Scholar 

  5. Cader MZ (2020) The genetics of migraine and the path to precision medicine. Prog Brain Res 255:403–418

    Article  PubMed  Google Scholar 

  6. Russell MB, Ducros A (2011) Sporadic and familial hemiplegic migraine: pathophysiological mechanisms, clinical characteristics, diagnosis, and management. Lancet Neurol 10:457–470

    Article  PubMed  Google Scholar 

  7. Thomsen LL, Eriksen MK, Roemer SF et al (2002) A population-based study of familial hemiplegic migraine suggests revised diagnostic criteria. Brain J Neurol 125:1379–1391

    Article  CAS  Google Scholar 

  8. Headache Classification Committee of the International Hedache Society (IHS). (2018) The International Classification of Headache Disorders, 3rd edition. Cephalalgia 38:1–211

  9. Lykke Thomsen L, Kirchmann Eriksen M, Faerch Romer S et al (2002) An epidemiological survey of hemiplegic migraine. Cephalalgia Int J Headache 22:361–375

    Article  CAS  Google Scholar 

  10. Thomsen LL, Ostergaard E, Olesen J et al (2003) Evidence for a separate type of migraine with aura: sporadic hemiplegic migraine. Neurology 60:595–601

    Article  PubMed  Google Scholar 

  11. Lai T-H, Hong C-T (2012) Prolonged symptoms in sporadic hemiplegic migraine: aura or migrainous infarction? Acta Neurol Taiwanica 21:129–132

    Google Scholar 

  12. Kumar G, Topper L, Maytal J (2009) Familial hemiplegic migraine with prolonged aura and multimodality imaging: a case report. Headache 49:139–142

    Article  PubMed  Google Scholar 

  13. Saleh C, Pierquin G, Beyenburg S (2016) Hemiplegic migraine presenting with prolonged somnolence: a case report. Case Rep Neurol 8:204–210

    Article  PubMed  PubMed Central  Google Scholar 

  14. Pelzer N, Blom D, Stam A et al (2017) Recurrent coma and fever in familial hemiplegic migraine type 2. A prospective 15-year follow-up of a large family with a novel ATP1A2 mutation. Cephalalgia 37:737–755

    Article  CAS  PubMed  Google Scholar 

  15. Echenne B, Ducros A, Rivier F et al (1999) Recurrent episodes of coma: an unusual phenotype of familial hemiplegic migraine with linkage to chromosome 1. Neuropediatrics 30:214–217

    Article  CAS  PubMed  Google Scholar 

  16. Indelicato E, Boesch S (2021) From genotype to phenotype: expanding the clinical Spectrum of CACNA1A variants in the era of next generation sequencing. Front Neurol 12. Epub ahead of print March 2. https://doi.org/10.3389/fneur.2021.639994

  17. Ducros A, Denier C, Joutel A et al (2001) The clinical spectrum of familial hemiplegic migraine associated with mutations in a neuronal calcium channel. N Engl J Med 345:17–24

    Article  CAS  PubMed  Google Scholar 

  18. Hauge AW, Kirchmann M, Olesen J (2011) Characterization of consistent triggers of migraine with aura. Cephalalgia Int J Headache 31:416–438

    Article  Google Scholar 

  19. Toldo I, Brunello F, Morao V et al (2019) First Attack and Clinical Presentation of Hemiplegic Migraine in Pediatric Age: A Multicenter Retrospective Study and Literature Review. Front Neurol 10. Epub ahead of print October 15. https://doi.org/10.3389/fneur.2019.01079

  20. Hansen JM, Hauge AW, Ashina M et al (2011) Trigger factors for familial hemiplegic migraine. Cephalalgia Int J Headache 31:1274–1281

    Article  Google Scholar 

  21. Pelzer N, Haan J, Stam AH et al (2018) Clinical spectrum of hemiplegic migraine and chances of finding a pathogenic mutation. Neurology 90:e575–e582

    Article  PubMed  Google Scholar 

  22. Hommersom MP, van Prooije TH, Pennings M et al (2021. Epub ahead of print November 22) The complexities of CACNA1A in clinical neurogenetics. J Neurol. https://doi.org/10.1007/s00415-021-10897-9

  23. Humbertclaude V, Riant F, Krams B et al (2020) Cognitive impairment in children with CACNA1A mutations. Dev Med Child Neurol 62:330–337

    Article  PubMed  Google Scholar 

  24. Guerin AA, Feigenbaum A, Donner EJ et al (2008) Stepwise developmental regression associated with novel CACNA1A mutation. Pediatr Neurol 39:363–364

    Article  PubMed  Google Scholar 

  25. Vanmolkot KRJ, Stroink H, Koenderink JB et al (2006) Severe episodic neurological deficits and permanent mental retardation in a child with a novel FHM2 ATP1A2 mutation. Ann Neurol 59:310–314

    Article  CAS  PubMed  Google Scholar 

  26. Sutherland HG, Albury CL, Griffiths LR (2019) Advances in genetics of migraine. J Headache Pain 20:1–20

    Article  CAS  Google Scholar 

  27. de Boer I, Terwindt GM, van den Maagdenberg AM (2020) Genetics of migraine aura: an update. J Headache Pain 21:64

    Article  PubMed  PubMed Central  Google Scholar 

  28. Bron C, Sutherland HG, Griffiths LR (2021) Exploring the hereditary nature of migraine. Neuropsychiatr Dis Treat 17:1183–1194

    Article  PubMed  PubMed Central  Google Scholar 

  29. Sutherland HG, Maksemous N, Albury CL et al (2020) Comprehensive exonic sequencing of hemiplegic migraine-related genes in a cohort of suspected probands identifies known and potential pathogenic variants. Cells 9:2368

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Riant F, Roos C, Roubertie A et al (2022) Hemiplegic migraine associated with PRRT2 variations: a clinical and genetic study. Neurology 98:e51–e61

    Article  CAS  PubMed  Google Scholar 

  31. Hiekkala ME, Vuola P, Artto V et al (2018) The contribution of CACNA1A, ATP1A2 and SCN1A mutations in hemiplegic migraine: a clinical and genetic study in Finnish migraine families. Cephalalgia 38:1849–1863

    Article  PubMed  Google Scholar 

  32. Amberger J, Bocchini CA, Scott AF et al (2009) McKusick’s online Mendelian inheritance in man (OMIM). Nucleic Acids Res 37:D793–D796

    Article  CAS  PubMed  Google Scholar 

  33. Gardiner AR, Jaffer F, Dale RC et al (2015) The clinical and genetic heterogeneity of paroxysmal dyskinesias. Brain J Neurol 138:3567–3580

    Article  Google Scholar 

  34. Lee H-Y, Xu Y, Huang Y et al (2004) The gene for paroxysmal non-kinesigenic dyskinesia encodes an enzyme in a stress response pathway. Hum Mol Genet 13:3161–3170

    Article  CAS  PubMed  Google Scholar 

  35. Igarashi T, Inatomi J, Sekine T et al (1999) Mutations in SLC4A4 cause permanent isolated proximal renal tubular acidosis with ocular abnormalities. Nat Genet 23:264–266

    Article  CAS  PubMed  Google Scholar 

  36. Suzuki M, Van Paesschen W, Stalmans I et al (2010) Defective membrane expression of the Na(+)-HCO(3)(−) cotransporter NBCe1 is associated with familial migraine. Proc Natl Acad Sci U S A 107:15963–15968

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Gil-Perotín S, Jaijo T, Verdú AG et al (2021) Epilepsy, status epilepticus, and hemiplegic migraine coexisting with a novel SLC4A4 mutation. Neurol Sci Off J Ital Neurol Soc Ital Soc Clin Neurophysiol 42:3647–3654

    Google Scholar 

  38. Potic A, Nmezi B, Padiath QS (2015) CAPOS syndrome and hemiplegic migraine in a novel pedigree with the specific ATP1A3 mutation. J Neurol Sci 358:453–456

    Article  CAS  PubMed  Google Scholar 

  39. Heinzen EL, Swoboda KJ, Hitomi Y et al (2012) De novo mutations in ATP1A3 cause alternating hemiplegia of childhood. Nat Genet 44:1030–1034

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. de Vries B, Mamsa H, Stam AH et al (2009) Episodic ataxia associated with EAAT1 mutation C186S affecting glutamate reuptake. Arch Neurol 66:97–101

    Article  PubMed  Google Scholar 

  41. Jen JC, Wan J, Palos TP et al (2005) Mutation in the glutamate transporter EAAT1 causes episodic ataxia, hemiplegia, and seizures. Neurology 65:529–534

    Article  CAS  PubMed  Google Scholar 

  42. Kovermann P, Hessel M, Kortzak D et al (2017) Impaired K(+) binding to glial glutamate transporter EAAT1 in migraine. Sci Rep 7:13913

    Article  PubMed  PubMed Central  Google Scholar 

  43. Paucar M, Granberg T, Lagerstedt-Robinson K et al (2020) SLC1A3 variant associated with hemiplegic migraine and acetazolamide-responsive MRS changes. Neurol Genet 6:e474

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Gburek-Augustat J, Heinze A, Abou Jamra R et al (2020) Hemiplegic migraine in glut1 deficiency syndrome and paroxysmal dyskinesia at ketogenic diet induction: case report and literature review. Mov Disord Clin Pract 7:965–970

    Article  PubMed  PubMed Central  Google Scholar 

  45. Bourque DK, Cordeiro D, Nimmo GAM et al (2021) Phenotypic and genotypic Spectrum of glucose Transporter-1 deficiency syndrome. Can J Neurol Sci J Can Sci Neurol 48:826–830

    Google Scholar 

  46. Mohammad SS, Coman D, Calvert S (2014) Glucose transporter 1 deficiency syndrome and hemiplegic migraines as a dominant presenting clinical feature. J Paediatr Child Health 50:1025–1026

    PubMed  Google Scholar 

  47. Weller CM, Leen WG, Neville BGR et al (2015) A novel SLC2A1 mutation linking hemiplegic migraine with alternating hemiplegia of childhood. Cephalalgia Int J Headache 35:10–15

    Article  Google Scholar 

  48. Coppola G, Pastorino GMG, Vetri L et al (2020) Familial Hemiplegic Migraine with an ATP1A4 Mutation: Clinical Spectrum and Carbamazepine Efficacy. Brain Sci 10. Epub ahead of print June 15. https://doi.org/10.3390/brainsci10060372

  49. Hasırcı Bayır BR, Tutkavul K, Eser M et al (2021) Epilepsy in patients with familial hemiplegic migraine. Seizure 88:87–94

    Article  PubMed  Google Scholar 

  50. Friedrich T, Tavraz NN, Junghans C (2016) ATP1A2 mutations in migraine: seeing through the facets of an ion pump onto the neurobiology of disease. Front Physiol 7. Epub ahead of print June 21. https://doi.org/10.3389/fphys.2016.00239

  51. Jurkat-Rott K, Freilinger T, Dreier JP et al (2004) Variability of familial hemiplegic migraine with novel A1A2 Na+/K+-ATPase variants. Neurology 62:1857–1861

    Article  CAS  PubMed  Google Scholar 

  52. Suzuki M, Fujiwara K, Tsubuku T et al (2016) Time course of downbeat positioning nystagmus in familial hemiplegic migraine type 1 treated with acetazolamide. J Neurol Sci 368:206–208

    Article  PubMed  Google Scholar 

  53. Battistini S, Stenirri S, Piatti M et al (1999) A new CACNA1A gene mutation in acetazolamide-responsive familial hemiplegic migraine and ataxia. Neurology 53:38–43

    Article  CAS  PubMed  Google Scholar 

  54. Terwindt GM, Ophoff RA, van Eijk R et al (2001) Involvement of the CACNA1A gene containing region on 19p13 in migraine with and without aura. Neurology 56:1028–1032

    Article  CAS  PubMed  Google Scholar 

  55. Yu W, Horowitz SH (2001) Familial hemiplegic migraine and its abortive therapy with intravenous verapamil. Neurology 57:1732–1733

    Article  CAS  PubMed  Google Scholar 

  56. Elliott MA, Peroutka SJ, Welch S et al (1996) Familial hemiplegic migraine, nystagmus, and cerebellar atrophy. Ann Neurol 39:100–106

    Article  CAS  PubMed  Google Scholar 

  57. Le Fort D, Safran AB, Picard F et al (2004) Elicited repetitive daily blindness: a new familial disorder related to migraine and epilepsy. Neurology 63:348–350

    Article  PubMed  Google Scholar 

  58. Ophoff RA, Terwindt GM, Vergouwe MN 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  PubMed  Google Scholar 

  59. Tottene A, Fellin T, Pagnutti S et al (2002) 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 99:13284–13289

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. van den Maagdenberg AM, Pietrobon D, Pizzorusso T et al (2004) A Cacna1a knockin migraine mouse model with increased susceptibility to cortical spreading depression. Neuron 41:701–710

    Article  PubMed  Google Scholar 

  61. van den Maagdenberg AMJM, Pizzorusso T, Kaja S et al (2010) High cortical spreading depression susceptibility and migraine-associated symptoms in ca(v)2.1 S218L mice. Ann Neurol 67:85–98

    Article  PubMed  Google Scholar 

  62. Meneghetti N, Cerri C, Vannini E et al (2022) Synaptic alterations in visual cortex reshape contrast-dependent gamma oscillations and inhibition-excitation ratio in a genetic mouse model of migraine. J Headache Pain 23:125

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Chanda ML, Tuttle AH, Baran I et al (2013) Behavioral evidence for photophobia and stress-related ipsilateral head pain in transgenic Cacna1a mutant mice. Pain 154:1254–1262

    Article  PubMed  Google Scholar 

  64. Ferrari MD, Klever RR, Terwindt GM et al (2015) Migraine pathophysiology: lessons from mouse models and human genetics. Lancet Neurol 14:65–80

    Article  CAS  PubMed  Google Scholar 

  65. Sintas Vives C, Carreño O, Fernández Castillo N et al (2017) Mutation spectrum in the CACNA1A gene in 49 patients with episodic ataxia. Sci Rep 7:2514

    Article  Google Scholar 

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

    Article  PubMed  Google Scholar 

  67. Leo L, Gherardini L, Barone V et al (2011) Increased susceptibility to cortical spreading depression in the mouse model of familial hemiplegic migraine type 2. PLoS Genet 7:e1002129

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Capuani C, Melone M, Tottene A et al (2016) Defective glutamate and K+ clearance by cortical astrocytes in familial hemiplegic migraine type 2. EMBO Mol Med 8:967–986

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Unekawa M, Ikeda K, Tomita Y et al (2018) Enhanced susceptibility to cortical spreading depression in two types of Na(+),K(+)-ATPase α2 subunit-deficient mice as a model of familial hemiplegic migraine 2. Cephalalgia Int J Headache 38:1515–1524

    Article  Google Scholar 

  70. Smith SE, Chen X, Brier LM et al (2020) Astrocyte deletion of α2-Na/K ATPase triggers episodic motor paralysis in mice via a metabolic pathway. Nat Commun 11:6164

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Dichgans M, Freilinger T, Eckstein G et al (2005) Mutation in the neuronal voltage-gated sodium channel SCN1A in familial hemiplegic migraine. Lancet Lond Engl 366:371–377

    Article  CAS  Google Scholar 

  72. Martin MS, Dutt K, Papale LA et al (2010) Altered function of the SCN1A voltage-gated sodium channel leads to gamma-aminobutyric acid-ergic (GABAergic) interneuron abnormalities. J Biol Chem 285:9823–9834

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Cestele S, Scalmani P, Rusconi R et al (2008) Self-limited Hyperexcitability: functional effect of a familial hemiplegic migraine mutation of the Nav1.1 (SCN1A) Na+ channel. J Neurosci 28:7273–7283

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Bertelli S, Barbieri R, Pusch M et al (2019) Gain of function of sporadic/familial hemiplegic migraine-causing SCN1A mutations: use of an optimized cDNA. Cephalalgia Int J Headache 39:477–488

    Article  Google Scholar 

  75. Cloarec R, Bruneau N, Rudolf G et al (2012) PRRT2 links infantile convulsions and paroxysmal dyskinesia with migraine. Neurology 79:2097–2103

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Gardiner AR, Bhatia KP, Stamelou M et al (2012) PRRT2 gene mutations: from paroxysmal dyskinesia to episodic ataxia and hemiplegic migraine. Neurology 79:2115–2121

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Marini C, Conti V, Mei D et al (2012) PRRT2 mutations in familial infantile seizures, paroxysmal dyskinesia, and hemiplegic migraine. Neurology 79:2109–2114

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Riant F, Roze E, Barbance C et al (2012) PRRT2 mutations cause hemiplegic migraine. Neurology 79:2122–2124

    Article  CAS  PubMed  Google Scholar 

  79. Dale RC, Gardiner A, Antony J et al (2012) Familial PRRT2 mutation with heterogeneous paroxysmal disorders including paroxysmal torticollis and hemiplegic migraine. Dev Med Child Neurol 54:958–960

    Article  PubMed  Google Scholar 

  80. Suzuki-Muromoto S, Kosaki R, Kosaki K et al (2020) Familial hemiplegic migraine with a PRRT2 mutation: phenotypic variations and carbamazepine efficacy. Brain and Development 42:293–297

    Article  PubMed  Google Scholar 

  81. Ebrahimi-Fakhari D, Saffari A, Westenberger A et al (2015) The evolving spectrum of PRRT2-associated paroxysmal diseases. Brain J Neurol 138:3476–3495

    Article  Google Scholar 

  82. Yang L, You C, Qiu S et al (2020) Novel and de novo point and large microdeletion mutation in PRRT2-related epilepsy. Brain Behav 10:e01597

    Article  PubMed  PubMed Central  Google Scholar 

  83. Michetti C, Castroflorio E, Marchionni I et al (2017) The PRRT2 knockout mouse recapitulates the neurological diseases associated with PRRT2 mutations. Neurobiol Dis 99:66–83

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Riant F, Ducros A, Ploton C et al (2010) De novo mutations in ATP1A2 and CACNA1A are frequent in early-onset sporadic hemiplegic migraine. Neurology 75:967–972

    Article  CAS  PubMed  Google Scholar 

  85. Jen JC, Klein A, Boltshauser E et al (2007) Prolonged hemiplegic episodes in children due to mutations in ATP1A2. J Neurol Neurosurg Psychiatry 78:523–526

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. De Sanctis S, Grieco GS, Breda L et al (2011) Prolonged sporadic hemiplegic migraine associated with a novel de novo missense ATP1A2 gene mutation. Headache 51:447–450

    Article  PubMed  Google Scholar 

  87. Pelzer N, Hoogeveen ES, Ferrari MD et al (2018) Brain atrophy following hemiplegic migraine attacks. Cephalalgia 38:1199–1202

    Article  PubMed  Google Scholar 

  88. Stubberud A, O’Connor E, Tronvik E et al (2021) R1352Q CACNA1A variant in a patient with sporadic hemiplegic migraine, Ataxia, seizures and cerebral Oedema: a case report. Case Rep Neurol 13:123–130

    Article  PubMed  PubMed Central  Google Scholar 

  89. Moore BA, Hale WJ, Nabity PS et al (2019) A retrospective, epidemiological review of hemiplegic migraines in a military population. Mil Med 184:781–787

    Article  PubMed  Google Scholar 

  90. Thomsen LL, Ostergaard E, Romer SF et al (2003) Sporadic hemiplegic migraine is an aetiologically heterogeneous disorder. Cephalalgia Int J Headache 23:921–928

    Article  CAS  Google Scholar 

  91. Thomsen LL, Olesen J, Russell MB (2003) Increased risk of migraine with typical aura in probands with familial hemiplegic migraine and their relatives. Eur J Neurol 10:421–427

    Article  CAS  PubMed  Google Scholar 

  92. Wieser T, Mueller C, Evers S et al (2003) Absence of known familial hemiplegic migraine (FHM) mutations in the CACNA1A gene in patients with common migraine: implications for genetic testing. Clin Chem Lab Med 41:272–275

    Article  CAS  PubMed  Google Scholar 

  93. Meamar R, Ostadsharif M, Saadatnia M et al (2013) Mutation analysis of CACNA1A gene in Iranian migrainous and review literatures. J Res Med Sci Off J Isfahan Univ Med Sci 18:S6–S10

    Google Scholar 

  94. Hautakangas H, Winsvold BS, Ruotsalainen SE et al (2022) Genome-wide analysis of 102,084 migraine cases identifies 123 risk loci and subtype-specific risk alleles. Nat Genet 54:152–160

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Weir GA, Pettingill P, Wu Y et al (2019) The role of TRESK in discrete sensory neuron populations and somatosensory processing. Front Mol Neurosci 12:170

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Lafrenière RG, Cader MZ, Poulin J-F et al (2010) A dominant-negative mutation in the TRESK potassium channel is linked to familial migraine with aura. Nat Med 16:1157–1160

    Article  PubMed  Google Scholar 

  97. Pettingill P, Weir GA, Wei T et al (2019) A causal role for TRESK loss of function in migraine mechanisms. Brain J Neurol 142:3852–3867

    Article  Google Scholar 

  98. Andres-Enguix I, Shang L, Stansfeld PJ et al (2012) Functional analysis of missense variants in the TRESK (KCNK18) K channel. Sci Rep 2:237

    Article  PubMed  PubMed Central  Google Scholar 

  99. Royal P, Andres-Bilbe A, Ávalos Prado P et al (2019) Migraine-associated TRESK mutations increase neuronal excitability through alternative translation initiation and inhibition of TREK. Neuron 101:232–245.e6

    Article  CAS  PubMed  Google Scholar 

  100. Han JY, Jang JH, Park J et al (2018) Targeted next-generation sequencing of Korean patients with developmental delay and/or intellectual disability. Front Pediatr 6:391

    Article  PubMed  PubMed Central  Google Scholar 

  101. Imbrici P, Nematian-Ardestani E, Hasan S et al (2020) Altered functional properties of a missense variant in the TRESK K(+) channel (KCNK18) associated with migraine and intellectual disability. Pflugers Arch 472:923–930

    Article  CAS  PubMed  Google Scholar 

  102. Lengyel M, Hajdu D, Dobolyi A et al (2021) TRESK background potassium channel modifies the TRPV1-mediated nociceptor excitability in sensory neurons. Cephalalgia Int J Headache 41:827–838

    Article  Google Scholar 

  103. Markel KA, Curtis D (2022) Study of variants in genes implicated in rare familial migraine syndromes and their association with migraine in 200,000 exome-sequenced UK biobank participants. Ann Hum Genet 86:353–360

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Diamond ML (2002) The role of concomitant headache types and non-headache co-morbidities in the underdiagnosis of migraine. Neurology 58:S3–S9

    Article  PubMed  Google Scholar 

  105. Brennan KC, Bates EA, Shapiro RE et al (2013) Casein kinase iδ mutations in familial migraine and advanced sleep phase. Sci Transl Med 5:183ra56 1–11

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Akerman S, Holland PR, Goadsby PJ (2011) Diencephalic and brainstem mechanisms in migraine. Nat Rev Neurosci 12:570–584

    Article  CAS  PubMed  Google Scholar 

  107. Alstadhaug K, Salvesen R, Bekkelund S (2008) 24-hour distribution of migraine attacks. Headache 48:95–100

    PubMed  Google Scholar 

  108. Williams LB, Javed A, Sabri A et al (2019) ALPK1 missense pathogenic variant in five families leads to ROSAH syndrome, an ocular multisystem autosomal dominant disorder. Genet Med Off J Am Coll Med Genet 21:2103–2115

    Google Scholar 

  109. Wardlaw JM, Smith C, Dichgans M (2019) Small vessel disease: mechanisms and clinical implications. Lancet Neurol 18:684–696

    Article  PubMed  Google Scholar 

  110. Joutel A, Corpechot C, Ducros A et al (1996) Notch3 mutations in CADASIL, a hereditary adult-onset condition causing stroke and dementia. Nature 383:707–710

    Article  CAS  PubMed  Google Scholar 

  111. Tan RYY, Markus HS (2016) CADASIL: migraine, encephalopathy, Stroke and Their Inter-Relationships. PloS One 11:e0157613

    Article  PubMed  PubMed Central  Google Scholar 

  112. Adib-Samii P, Brice G, Martin RJ et al (2010) Clinical spectrum of CADASIL and the effect of cardiovascular risk factors on phenotype: study in 200 consecutively recruited individuals. Stroke 41:630–634

    Article  PubMed  Google Scholar 

  113. Eikermann-Haerter K, Yuzawa I, Dilekoz E et al (2011) Cerebral autosomal dominant arteriopathy with subcortical infarcts and leukoencephalopathy syndrome mutations increase susceptibility to spreading depression. Ann Neurol 69:413–418

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Ducros A (2021) Genetics of migraine. Rev Neurol (Paris) 177:801–808

    Article  CAS  PubMed  Google Scholar 

  115. Auton A, Brooks LD, Durbin RM et al (2015) A global reference for human genetic variation. Nature 526:68–74

    Article  PubMed  Google Scholar 

  116. Anttila V, Stefansson H, Kallela M et al (2010) Genome-wide association study of migraine implicates a common susceptibility variant on 8q22.1. Nat Genet 42:869–873

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Chasman DI, Schürks M, Anttila V et al (2011) Genome-wide association study reveals three susceptibility loci for common migraine in the general population. Nat Genet 43:695–698

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Anttila V, Winsvold BS, Gormley P et al (2013) Genome-wide meta-analysis identifies new susceptibility loci for migraine. Nat Genet 45:912–917

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Freilinger T, Bohe M, Wegener B et al (2008) Expansion of the phenotypic Spectrum of the CACNA1A T666M mutation: a family with familial hemiplegic migraine type 1, Cerebellar Atrophy and Mental Retardation. Cephalalgia 28:403–407

    Article  CAS  PubMed  Google Scholar 

  120. Gormley P, Anttila V, Winsvold BS et al (2016) Meta-analysis of 375,000 individuals identifies 38 susceptibility loci for migraine. Nat Genet 48:856–866

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Finucane HK, Reshef YA, Anttila V et al (2018) Heritability enrichment of specifically expressed genes identifies disease-relevant tissues and cell types. Nat Genet 50:621–629

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Eising E, Huisman SMH, Mahfouz A et al (2016) Gene co-expression analysis identifies brain regions and cell types involved in migraine pathophysiology: a GWAS-based study using the Allen human brain atlas. Hum Genet 135:425–439

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Ligthart L, de Vries B, Smith AV et al (2011) Meta-analysis of genome-wide association for migraine in six population-based European cohorts. Eur J Hum Genet EJHG 19:901–907

    Article  CAS  PubMed  Google Scholar 

  124. Freilinger T, Anttila V, de Vries B et al (2012) Genome-wide association analysis identifies susceptibility loci for migraine without aura. Nat Genet 44:777–782

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Chen S-P, Fuh J-L, Chung M-Y et al (2018) Genome-wide association study identifies novel susceptibility loci for migraine in Han Chinese resided in Taiwan. Cephalalgia Int J Headache 38:466–475

    Article  Google Scholar 

  126. Chang X, Pellegrino R, Garifallou J et al (2018) Common variants at 5q33.1 predispose to migraine in African-American children. J Med Genet 55:831–836

    Article  CAS  PubMed  Google Scholar 

  127. Choquet H, Yin J, Jacobson AS et al (2021) New and sex-specific migraine susceptibility loci identified from a multiethnic genome-wide meta-analysis. Commun Biol 4:864

    Article  PubMed  PubMed Central  Google Scholar 

  128. Jiang Z, Zhao L, Zhang X et al (2021) Common variants in KCNK5 and FHL5 genes contributed to the susceptibility of migraine without aura in Han Chinese population. Sci Rep 11:6807

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Tsai C-K, Liang C-S, Lin G-Y et al (2021) Identifying genetic variants for age of migraine onset in a Han Chinese population in Taiwan. J Headache Pain 22:89

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Gupta RM, Hadaya J, Trehan A et al (2017) A genetic variant associated with five vascular diseases is a distal regulator of Endothelin-1 gene expression. Cell 170:522–533.e15

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Codina-Fauteux V-A, Beaudoin M, Lalonde S et al (2018) PHACTR1 splicing isoforms and eQTLs in atherosclerosis-relevant human cells. BMC Med Genet 19:97

    Article  PubMed  PubMed Central  Google Scholar 

  132. Alves-Ferreira M, Quintas M, Sequeiros J et al (2021) A genetic interaction of NRXN2 with GABRE, SYT1 and CASK in migraine patients: a case-control study. J Headache Pain 22:57

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. Gormley P, Kurki MI, Hiekkala ME et al (2018) Common variant burden contributes to the familial aggregation of migraine in 1,589 families. Neuron 98:743–753.e4

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Häppölä P, Gormley P, Nuottamo ME et al (2022) Polygenic risk provides biological validity for the ICHD-3 criteria among Finnish migraine families. Cephalalgia Int J Headache 42:345–356

    Article  Google Scholar 

  135. Techlo TR, Rasmussen AH, Møller PL et al (2020) Familial analysis reveals rare risk variants for migraine in regulatory regions. Neurogenetics 21:149–157

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Rasmussen AH, Kogelman LJA, Kristensen DM et al (2020) Functional gene networks reveal distinct mechanisms segregating in migraine families. Brain J Neurol 143:2945–2956

    Article  Google Scholar 

  137. Cutrer FM, Moyer AM, Atkinson EJ et al (2021) Genetic variants related to successful migraine prophylaxis with verapamil. Mol Genet Genomic Med 9:e1680

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  138. Ibrahim O, Sutherland HG, Maksemous N et al (2020) Exploring neuronal vulnerability to head trauma using a whole exome approach. J Neurotrauma 37:1870–1879

    Article  PubMed  PubMed Central  Google Scholar 

  139. Kraft P, Chen H, Lindström S (2020) The use of genetic correlation and Mendelian randomization studies to increase our understanding of relationships between complex traits. Curr Epidemiol Rep 7:104–112

    Article  PubMed  PubMed Central  Google Scholar 

  140. Yang Y, Zhao H, Boomsma DI et al (2018) Molecular genetic overlap between migraine and major depressive disorder. Eur J Hum Genet EJHG 26:1202–1216

    Article  CAS  PubMed  Google Scholar 

  141. Bahrami S, Hindley G, Winsvold BS et al (2022) Dissecting the shared genetic basis of migraine and mental disorders using novel statistical tools. Brain J Neurol 145:142–153

    Article  Google Scholar 

  142. Malik R, Freilinger T, Winsvold BS et al (2015) Shared genetic basis for migraine and ischemic stroke: a genome-wide analysis of common variants. Neurology 84:2132–2145

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Daghlas I, Guo Y, Chasman DI (2020) Effect of genetic liability to migraine on coronary artery disease and atrial fibrillation: a Mendelian randomization study. Eur J Neurol 27:550–556

    Article  CAS  PubMed  Google Scholar 

  144. Siewert KM, Klarin D, Damrauer SM et al (2020) Cross-trait analyses with migraine reveal widespread pleiotropy and suggest a vascular component to migraine headache. Int J Epidemiol 49:1022–1031

    Article  PubMed  PubMed Central  Google Scholar 

  145. Guo Y, Rist PM, Daghlas I et al (2020) A genome-wide cross-phenotype meta-analysis of the association of blood pressure with migraine. Nat Commun 11:3368

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Daghlas I, Vgontzas A, Guo Y et al (2020) Habitual sleep disturbances and migraine: a Mendelian randomization study. Ann Clin Transl Neurol 7:2370–2380

    Article  PubMed  PubMed Central  Google Scholar 

  147. Adewuyi EO, Sapkota Y, Iec IEC et al (2020) Shared Molecular Genetic Mechanisms Underlie Endometriosis and Migraine Comorbidity. Genes 11. Epub ahead of print February 29. https://doi.org/10.3390/genes11030268

  148. Georges A, Yang M-L, Berrandou T-E et al (2021) Genetic investigation of fibromuscular dysplasia identifies risk loci and shared genetics with common cardiovascular diseases. Nat Commun 12:6031

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  149. Jiang Y-J, Fann CS-J, Fuh J-L et al (2022) Genome-wide analysis identified novel susceptible genes of restless legs syndrome in migraineurs. J Headache Pain 23:39

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Guo Y, Rist PM, Sabater-Lleal M et al (2021) Association between hemostatic profile and migraine: a Mendelian randomization analysis. Neurology 96:e2481–e2487

    Article  PubMed  PubMed Central  Google Scholar 

  151. Daghlas I, Rist PM, Chasman DI (2020) Effect of genetic liability to migraine on cognition and brain volume: a Mendelian randomization study. Cephalalgia Int J Headache 40:998–1002

    Article  Google Scholar 

  152. Shu M-J, Li J-R, Zhu Y-C et al (2022) Migraine and ischemic stroke: a Mendelian randomization study. Neurol Ther 11:237–246

    Article  PubMed  Google Scholar 

  153. Lee K-J, Lee SJ, Bae H-J et al (2022) Exploring the causal inference of migraine on stroke: a Mendelian randomization study. Eur J Neurol 29:335–338

    Article  PubMed  Google Scholar 

  154. Daghals I, Sargurupremraj M, Danning R et al (2022) Migraine, stroke, and cervical arterial dissection: shared genetics for a triad of brain disorders with vascular involvement. Neurol Genet 8:e653

    PubMed  PubMed Central  Google Scholar 

  155. Yuan S, Daghlas I, Larsson SC (2022) Alcohol, coffee consumption, and smoking in relation to migraine: a bidirectional Mendelian randomization study. Pain 163:e342–e348

    Article  PubMed  Google Scholar 

  156. Chen H, Zhang H, Zheng L (2022) No causal association between coffee consumption and risk of migraine: a Mendelian randomization study. Front Genet 13:792313

    Article  PubMed  PubMed Central  Google Scholar 

  157. Johnsen MB, Winsvold BS, Børte S et al (2018) The causal role of smoking on the risk of headache. A Mendelian randomization analysis in the HUNT study. Eur J Neurol 25:1148–e102

    Article  CAS  PubMed  Google Scholar 

  158. Mitchell BL, Diaz-Torres S, Bivol S et al (2022) Elucidating the relationship between migraine risk and brain structure using genetic data. Brain J Neurol 145:3214–3224

    Article  Google Scholar 

  159. Chu S, Wu Z, Wu Z et al (2021) Association between insomnia and migraine risk: a case-control and bidirectional Mendelian randomization study. Pharmacogenomics Pers Med 14:971–976

    Google Scholar 

  160. Chen J, Chen X, Xie Y et al (2021) Irritable bowel syndrome and migraine: evidence from Mendelian randomization analysis in the UK biobank. Expert Rev Gastroenterol Hepatol 15:1233–1239

    Article  CAS  PubMed  Google Scholar 

  161. Tanha HM, Martin NG, Whitfield JB et al (2021) Association and genetic overlap between clinical chemistry tests and migraine. Cephalalgia Int J Headache 41:1208–1221

    Article  Google Scholar 

  162. Tanha HM, Sathyanarayanan A, Nyholt DR (2021) Genetic overlap and causality between blood metabolites and migraine. Am J Hum Genet 108:2086–2098

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Guo Y, Daghlas I, Gormley P et al (2021) Phenotypic and genotypic associations between migraine and lipoprotein subfractions. Neurology 97:e2223–e2235

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Yin P, Anttila V, Siewert KM et al (2017) Serum calcium and risk of migraine: a Mendelian randomization study. Hum Mol Genet 26:820–828

    CAS  PubMed  Google Scholar 

  165. Niu P-P, Wang X, Xu Y-M (2022) Higher circulating vitamin D levels are associated with decreased migraine risk: a Mendelian randomization study. Front Nutr 9:907789

    Article  PubMed  PubMed Central  Google Scholar 

  166. Abuduxukuer R, Niu P-P, Guo Z-N et al (2022) Circulating insulin-like growth factor 1 levels and migraine risk: a Mendelian randomization study. Neurol Ther 11:1677–1689

    Article  PubMed  PubMed Central  Google Scholar 

  167. Hannon E, Weedon M, Bray N et al (2017) Pleiotropic effects of trait-associated genetic variation on DNA methylation: utility for refining GWAS loci. Am J Hum Genet 100:954–959

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Hughes A, Wade KH, Dickson M et al (2021) Common health conditions in childhood and adolescence, school absence, and educational attainment: Mendelian randomization study. NPJ Sci Learn 6:1

    Article  PubMed  PubMed Central  Google Scholar 

  169. Harrison S, Davies AR, Dickson M et al (2020) The causal effects of health conditions and risk factors on social and socioeconomic outcomes: Mendelian randomization in UK biobank. Int J Epidemiol 49:1661–1681

    Article  PubMed  PubMed Central  Google Scholar 

  170. Meng W, Adams MJ, Hebert HL et al (2018) A genome-wide association study finds genetic associations with broadly-defined headache in UK biobank (N=223,773). EBioMedicine 28:180–186

    Article  PubMed  PubMed Central  Google Scholar 

  171. Tsao Y-C, Wang S-J, Hsu C-L et al (2022) Genome-wide association study reveals susceptibility loci for self-reported headache in a large community-based Asian population. Cephalalgia Int J Headache 42:229–238

    Article  Google Scholar 

  172. Cargnin S, Sances G, Shin JI et al (2021) Gene polymorphism association studies in cluster headache: a field synopsis and systematic meta-analyses. Headache 61:1060–1076

    Article  PubMed  Google Scholar 

  173. Gibson KF, Santos AD, Lund N et al (2019) Genetics of cluster headache. Cephalalgia Int J Headache 39:1298–1312

    Article  Google Scholar 

  174. Bacchelli E, Cainazzo MM, Cameli C et al (2016) A genome-wide analysis in cluster headache points to neprilysin and PACAP receptor gene variants. J Headache Pain 17:114

    Article  PubMed  PubMed Central  Google Scholar 

  175. Ran C, Fourier C, Michalska JM et al (2017) Screening of genetic variants in ADCYAP1R1, MME and 14q21 in a Swedish cluster headache cohort. J Headache Pain 18:88

    Article  PubMed  PubMed Central  Google Scholar 

  176. Harder AVE, Winsvold BS, Noordam R et al (2021) Genetic susceptibility loci in Genomewide association study of cluster headache. Ann Neurol 90:203–216

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. O’Connor E, Fourier C, Ran C et al (2021) Genome-wide association study identifies risk loci for cluster headache. Ann Neurol 90:193–202

    Article  PubMed  Google Scholar 

  178. Ling Y-H, Chen S-P, Fann CS-J et al (2019) TRPM8 genetic variant is associated with chronic migraine and allodynia. J Headache Pain 20:115

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Yakubova A, Davidyuk Y, Tohka J et al (2021) Searching for predictors of migraine Chronification: a pilot study of 1911A>G polymorphism of TRPV1 gene in episodic versus chronic migraine. J Mol Neurosci MN 71:618–624

    Article  CAS  PubMed  Google Scholar 

  180. Huang C, Chen S-P, Huang Y-H et al (2020) HLA class I alleles are associated with clinic-based migraine and increased risks of chronic migraine and medication overuse. Cephalalgia Int J Headache 40:493–502

    Article  CAS  Google Scholar 

  181. Louter MA, Fernandez-Morales J, de Vries B et al (2015) Candidate-gene association study searching for genetic factors involved in migraine chronification. Cephalalgia Int J Headache 35:500–507

    Article  CAS  Google Scholar 

  182. Chalmer MA, Rasmussen AH, Kogelman LJA et al (2021) Chronic migraine: genetics or environment? Eur J Neurol 28:1726–1736

    Article  PubMed  PubMed Central  Google Scholar 

  183. Belyaeva II, Subbotina AG, Eremenko II et al (2021) Pharmacogenetics in primary headache disorders. Front Pharmacol 12:820214

    CAS  PubMed  Google Scholar 

  184. Kogelman LJA, Esserlind A-L, Francke Christensen A et al (2019) Migraine polygenic risk score associates with efficacy of migraine-specific drugs. Neurol Genet 5:e364

    Article  PubMed  PubMed Central  Google Scholar 

  185. Guo S, Esserlind A-L, Andersson Z et al (2016) Prevalence of migraine in persons with the 3243A>G mutation in mitochondrial DNA. Eur J Neurol 23:175–181

    Article  CAS  PubMed  Google Scholar 

  186. Terrin A, Bello L, Valentino ML et al (2022) The relevance of migraine in the clinical spectrum of mitochondrial disorders. Sci Rep 12:4222

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Kraya T, Deschauer M, Joshi PR et al (2018) Prevalence of headache in patients with mitochondrial disease: a cross-sectional study. Headache 58:45–52

    Article  PubMed  Google Scholar 

  188. Burow P, Haselier M, Naegel S et al (2021) The mitochondrial biomarkers FGF-21 and GDF-15 in patients with episodic and chronic migraine. Cells 10. Epub ahead of print September 18. https://doi.org/10.3390/cells10092471

  189. Zaki EA, Freilinger T, Klopstock T et al (2009) Two common mitochondrial DNA polymorphisms are highly associated with migraine headache and cyclic vomiting syndrome. Cephalalgia Int J Headache 29:719–728

    Article  CAS  Google Scholar 

  190. Finnilä S, Autere J, Lehtovirta M et al (2001) Increased risk of sensorineural hearing loss and migraine in patients with a rare mitochondrial DNA variant 4336A>G in tRNAGln. J Med Genet 38:400–405

    Article  PubMed  PubMed Central  Google Scholar 

  191. Børte S, Zwart J-A, Skogholt AH et al (2020) Mitochondrial genome-wide association study of migraine - the HUNT study. Cephalalgia Int J Headache 40:625–634

    Article  Google Scholar 

Download references

Disclosures

Lou Grangeon has received travel and accommodation expenses from Lundbeck, not related to the submitted work.

Kristin Sophie Lange has received travel and accommodation expenses from Acticor Biotech and TEVA GmbH, not related to the submitted work.

Wietse Wiels is a fundamental research scholar of the Flanders Scientific Research Fund (FWO).

Anne Ducros had received personal fees for board membership and/or speaking from Allergan/Abbvie, Lilly, Lundbeck, Novartis, Pfizer, Teva, is associate editor of Cephalalgia and past-president of the French Headache Society, and has received research grants for her institution (CHU de Montpellier) from the Programme Hospitalier de Recherche Clinique (PHRC) and from Pfizer.

Author information

Authors and Affiliations

Authors

Consortia

Contributions

LG, KSL,CD and AD designed the outline, wrote the manuscript and the tables, corrected, formatted and led the review. MWP, DO, KM, WW, PM, FF collaborated to the literature search, the drafting of the tables and the revision of the manuscript. All authors reviewed the manuscript. The authors read and approved the final manuscript.

Corresponding author

Correspondence to Anne Ducros.

Ethics declarations

Competing interests

The authors declare no competing interests.

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 licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Grangeon, L., Lange, K.S., Waliszewska-Prosół, M. et al. Genetics of migraine: where are we now?. J Headache Pain 24, 12 (2023). https://doi.org/10.1186/s10194-023-01547-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s10194-023-01547-8

Keywords