Submit manuscript...
Journal of
eISSN: 2373-6410

Neurology & Stroke

Correspondence:

Received: January 01, 1970 | Published: ,

Citation: DOI:

Download PDF

Abstract

Nowadays migraine ranks 9th in the list of leading causes of disability among population. In Russia migraine prevalence is two times higher than the world index and inflicts a considerable damage on the state economy. Despite almost one-century history of studying migraine, science until now cannot explain many cases of attack occurrence. It causes difficulties both for diagnosis and treatment – the therapy of patients with migraine is not sufficiently effective. Today one of the investigation directions is searching of migraine biomarkers confirming diagnosis. In this review we attempted to generalize the results of available works targeted at searching genetic markers of migraine.

Keywords: Migraine; Gene; Polymorphism

Introduction

Migraine is now one of the leading causes of disability (ranks 9th according to the WHO), comparable to such diseases as cancer, diabetes, cardiovascular diseases and others. In the female population, migraine-related disability ratio promotes this disease to 3rd place. According to epidemiological studies, migraine prevalence in the world for 1 year in the adult population ranges on average from 10.2% [1] to 14.7 % [2]. In Russia, migraine prevalence exceeds world figures almost 1.5-2 fold, being 20.3%, and annual indirect costs (days lost due to disability) related to primary headaches total US $22.8 billion (1.75% of Russia’s gross domestic product) [3]. Thus, migraine is not only a medical, but also a significant economic problem.

Until now, the diagnosis of "migraine" is exclusively clinical, and any diagnostic tests are aimed only at excluding other causes of headache [4]. There are also problems with migraine treatment and although both traditional analgesics and specific anti-migraine products are available in the market, treatment of migraine patients is still not sufficiently effective. For example, specific anti-migraine agents (triptans) help control only two out of three attacks, and migraine prevention products are considered effective, if they reduce the frequency of attacks by 50% or more. The chronification of migraine attacks and the development of chronic daily headaches, occurring in 1% of patients per year [5], are a significant clinical problem. However, about 10% of migraine patients in the population and 40-60% of patients visiting specialized headache centers are resistant to standard therapy [6]. Treatment of such patients is the most expensive.

Thus, searching for migraine biomarkers that confirm such diagnosis, instead of refuting other diagnoses, is the principal vector in this scientific field. In this review, we attempted to summarize the available information about studies aimed at searching for genetic markers of migraine.

Discussion

Inheritance of migraine

Hereditary factors play an important role in the development of migraine [7]. Relatives of such patients have migraine much more often than the population in general; if both parents have migraine, the risk that their offsprings will have this disease reaches 60-90% (vs. 11% in the control group), and the leading role belongs to the mother: in this case the risk of disease in children is 72%.

Long-term studies have demonstrated familial aggregation of migraine symptoms, and in some cases a positive family history (presence of the disease in family history) is a diagnostic criterion for migraine. Studies of monozygotic and dizygotic twins also demonstrated the presence of a significant genetic component in the development of migraine: in monozygotic twins with migraine, concordance value is 1.5-2 times higher than in dizygotic twins (for MWOA and MWA) [8,9]. A large study involving about 30,000 pairs of twins showed that genetics and environmental factors contribute almost equally to the development of migraine [10]. Studies of twins who grew up together or separately showed that general environmental factors play a secondary role [11,12].

Differences in migraine prevalence between populations also may serve as an indirect evidence of the genetic basis of migraine pathogenesis; such differences may be due to the differences in allele frequencies between populations. According to foreign researchers [13,14], the genetic component in migraine with aura is stronger than that in migraine without aura. Some authors define migraine as a polygenic multiple-factor disease [15,16]. Currently, there is a belief that it is not the disease itself that is inherited, but rather a predisposition to respond to external stimuli of the nervous and vascular systems.

Monogenic migraine syndromes

This section presents rare neurological disorders, in which migraine attacks are a part of a broader clinical spectrum and can be regarded as a monogenic subtype of migraine. These subtypes may help identify and understand the pathophysiological mechanisms of migraine.

CADASIL-syndrome:(Cerebral Autosomal Dominant Arteriopathy with Subcortical Infarcts and Leukoencephalopathy) - a "cerebral autosomal dominant arteriopathy with subcortical infarcts and leukoencephalopathy”, characterized by recurrent subcortical ischemic strokes with severe white matter hyperintensity, seizures, cognitive decline, depression and other psychoneurological symptoms. Migraine, in particular migraine with aura, is a characteristic peculiarity of more than a third of patients, which occurs at least one decade prior to other symptoms [17]. CADASIL is caused by mutations in the NOTCH3 gene that encodes the NOTCH3 receptor and plays a key role in the functioning of smooth muscle cells that make up small arteries and arterioles in the brain [18]. Mutations lead to dysfunction of the signaling pathway that regulates the development of vessels during embryogenesis and supports the structural/functional stability of blood vessels in adults [19,20]. A specific feature of CADASIL is the accumulation of NOTCH3 receptor due to its slow elimination, which leads to the formation of granular osmiophilic deposits, and this affects small blood vessels and results in reduced cell adhesion and cell death, as well as in the transformation of smooth muscle cells in the middle layer and in fibrosis [21]. Thus, CADASIL may be caused by vascular dysfunction, which results in the death of smooth muscle cells in the vessels and in the degeneration of the structure of vessels.

Mitochondrial encephalopathy, lactic acidosis, and stroke-like episodes (MELAS): This disease is caused by mutations in several mitochondrial genes, most frequently in the MTTL1 gene encoding the mitochondrial tRNA for leucine (nucleotide A to nucleotide G transition in position 3243), and is characterized by seizures, stroke-like episodes and lactic acidosis [22]. A typical picture of MELAS includes seizures with neurovisual manifestations of cortical infarcts, which are often combined with migraine-like headaches; as well as hemiparesis, hemianopsia, cortical blindness, episodic vomiting, and short stature. Systemic manifestations may include cardiac, renal, endocrine or gastrointestinal disorders [23].

Cerebral hereditary angiopathy with vascular retinopathy and internal organ dysfunction (CHARIOT):A progressive systemic disease of small blood vessels, which is caused by mutations in the TREX1 gene [24,25]. The TREX1 gene is located in chromosome 3p21 and encodes human DNAase III (3' repair exonuclease) - an autonomous, non-processive 3'-5' DNA specific exonuclease [26]. This enzyme is localized in the perinuclear area of the cell, which plays a fundamental role in granzyme A-mediated cell death and, when mutated, indirectly activates the autoimmune reaction against the undigested double-stranded DNA from dying cells [27]. The main peculiarities of this disease include a progressive blindness due to vascular retinopathy; focal and cerebral neurological symptoms associated with cerebral edema and white matter lesions; and premature death. Additional symptoms, such as migraine and Raynaud's syndrome, are observed in more than a half of patients and occur almost ten years before other symptoms [25-28].

Patients with the familial advanced sleep-phase syndrome (FASPS) have serious disturbances of the sleep-wake cycle and other circadian rhythms. The disease is caused by missense mutations in the CSNK1D gene encoding Iδ (CK1δ) casein kinase that is involved in the phosphorylation of Per2 circadian rhythm protein [29-31]. In two independent families, CSNK1D mutations were observed in 9 of 11 patients with the familial advanced sleep-phase syndrome and migraine with aura [29]. Screening of two families with migraine with aura and FASPS identified two missense mutations (c.44T> A and c.46H> R) in the CSNK1D gene, which lead decreased enzyme levels [30]. Mice with T44A (Csnk1d) mutation a have lower threshold for cortical spreading depression, accompanied by increased spontaneous and induced activation of the calcium signaling pathway in astrocytes [29].

COL4A1-related syndromes:The COL4A1 gene encodes alpha-1 subunit of type IV collagen. Mutations in this gene may lead to several autosomal dominant disorders with overlapping characteristics, including perinatal hemorrhage with porencephalia [32-35], and small vessel disease, which result in hemorrhage and hemiparesis in childhood or adulthood [36]. The association of COLA4A1 mutations with migraine is not quite reliable and may be a random discovery, despite the fact that 10 out of 52 COLA4A1 mutation carriers have confirmed migraines (with or without aura) [35].

Familial and sporadic hemiplegic migraine (FHM) ischaracterized by migraine attacks combined with transient unilateral motor weakness. Aura, headaches and associated symptoms are identical, and attacks can be caused by similar triggers; the same medicinal products are used for treatment and prevention. In 75% of FHM patients, hemiplegic episodes may alternate with migraine episodes without motor weakness. FHM and migraine are more common in women, and migraine rates increase among first-degree relatives. FHM patients may also have additional transient and persistent neurological disorders, such as ataxia, epilepsy, cognitive disorders or loss of consciousness [37].

FHMs are genetically heterogeneous. 5 types of FHM are distinguished:
1) Type 1 FHM - missense mutations in the CACNA1Agene (50-75% of families) [15,38].
2) Type 2 FHM -mainly deletions and frameshift in the ATP1A2 gene (20% to 30% of cases) [39].
3) Type 3 FHM - mutations in the SCN1Agene on 2q24 [40].
4) Type 4 FHM - mutations in the CACNA1Egene on 1q25-q31 [41].
5) FHM induced by mutations in other genes: SLC1A3 [42], SLC4A4 [43], PRR2 [44].

Association studies

Approaches to studying candidate genes are widely used to study the genetics of migraine. Repeated studies were conducted for a significant number of genes, and those studies either confirmed or refuted the association. However, studies of candidate genes are interesting, as they can reveal the contribution of common genetic variants to the complex phenotype of specific ethnic groups, particularly genetic isolates. Candidate genes were previously grouped into four functional families of genes, namely, neurological, cardiovascular, hormonal and inflammatory genes [45].

  1. Genes involved in the nervous system functioning:  This category includes mainly candidate genes, the products of which are needed for the functioning of the nervous system:

1) Ion channels. For example, genes encoding calcium (CACNA1A, CACNB2, CACNB4) or potassium (KCNAB3, KCNB2, KCNG4, KCNJ10, KCNK18, KCNN3) channels.
2) Subunits of Na+/K+ -ATPase,
3) Molecules involved in the synthesis, release and binding of neuropeptides (calcitonin gene-related peptide) or neurotransmitters (glutamate, GABA, dopamine, serotonin) connected with neuronal excitation and/or nociception.

Some case-control association studies gave positive results for the DBH, DDC, DRD2, DRD3, DRD4, GRIA1, GRIA3, HTR2, 5-HTTLPR, MAOA, SLC6A3, SLC6A4 and BDNFgenes, although the results of most studies were negative, especially for the first two gene families [46-54]. Nevertheless, careful screening of 150 genes expressed in the brain and involved in ion homeostasis (channels, transporters, antiporters and auxiliary subunits) made it possible to identify three potassium channel encoding genes associated with migraine, namely KCNK18, KCNG4 and KCNAB3 [55]. KCNK18 is particularly interesting in terms of its expression in the trigeminal and dorsal root ganglia, and its relationship with the MWA was detected also by analyzing the linkage groups.

  1. Vascular genes: Association studies of the genes involved in the regulation of blood pressure, endothelial cell function, vasoconstriction (narrowing of the blood vessels) and vasodilation (widening of the blood vessels) provided more consistent positive results. Many vascular genes associated with migraine also pose a risk of stroke and heart diseases [56-58]. Common functional variants in several vascular genes can predispose the person to migraine and at the same time influence the type and frequency of attacks [58]:
  1. Angiotensin converting enzyme (ACE) plays a key role in maintaining blood pressure and vessel wall pressure. Homozygous deletion (DD) in the human ACE gene increases the enzymatic activity of ACE and is associated with the frequency and duration of MWA attacks [59-61].
  2. A number of studies revealed association between variants of the 5-10-methylenetetrahydrofolate reductase (MTHFR) gene and migraine. MTHFR is a key component of remethylation of homocysteine to methionine and catalyzes the conversion of 5,10-methylenetetrahydrofolate to 5-methyltetrahydrofolate. Mutations in the MTHFR gene may lead to hyperhomocysteinemia due to lower enzymatic activity. Several studies involving different ethnic groups [62-68] and several recent meta-analyzes [65-71] have confirmed the contribution of the T677 allele in the MTHFR gene (rs1801133) to the pathogenesis of migraine. However, it was reported that the absence of any connection with MTHFR gene variants can be due to age and selective survival [72].
  3. NOTCH3 encodes a transmembrane receptor that regulates the development of vessels and differentiation during embryogenesis, and also contributes to the integrity of vessels in adults (19-20). In addition to rare NOTCH3mutations leading to MWA in the context of CADASIL, other variants are significantly associated with migraine [73-77]. Therefore, NOTCH3 may play a more active role in the pathogenesis of common migraine without aura.
  4. Other endothelial genes assessed for association with migraine encode endothelin-1 (EDN1), endothelin receptor type A and B (EDNRA and EDNRB), inducible NO-synthase (NOS2), endothelial NO-synthase (NOS3), and vascular endothelial growth factor (VEGF) [56,78-84]. Several studies have found association between EDNRA alleles and migraine, and one study involving Finnish and German patients with migraine showed association between MWA and the rs2048894 (EDNRA) substitution, especially with the age of diseases onset <20 years [78].
  1. Hormones and genes:Genes controlling the metabolism of estrogen and progesterone theoretically may be associated with migraine and at least partially may explain gender distribution, as well as menstrual migraine [85]. However, the results of studies of genetic association are controversial, although there are some positive results [86-93]. In a recent study, three estrogen receptor 1 (ESR1) haplotypes were associated with the disease. In addition to the ESR1 gene, other hormonal genes were studied: estrogen receptor 2 (ESR2), progesterone receptor (PGR), androgen receptor (AR), follicle-stimulating hormone receptor (FSHR), nuclear receptor interacting protein 1 (NRIP1) and cytochrome P450, family 19, subfamily A, polypeptide 1 (CYP19A1). However, a meta-analysis of these genes showed association only with polymorphic variants c.594G>A and c.325C>G of the ESR1 gene, with no difference between MWA and MWOA found [93].
  1. Inflammation and genes: In animal and human studies, it is supposed that inflammation and immune system components may play a role in the pathogenesis of migraine. In case of CSD (cortical spreading depression), for example, this process causes a local neurogenic inflammation with the activation of mast cells and macrophages, accompanied by the release of pro-inflammatory cytokines, ultimately leading to sensitization of meningeal nociceptive nerve endings [94]. For the COX-2, HLA-DRB1, LTA, TNFA, TNFBandTNFRSF1B genes, positive associations with migraines were detected [95-102].

Genome-wide association studies (GWAS)

To date, six GWAS have been conducted, which studied migraine:

  1. The first GWAS was prepared by the International Headache Genetics Consortium (IHGC) in 2010. Anttila et al. [103] conducted a two-stage association study on six clinical and one European population samples [103]. When comparing 2,748 patients with migraine from three European headache clinics (Finland, Germany and Denmark) and the control sample (n=10,747), the minor allele of rs1835740 on chromosome 8q22.1 was identified, which is associated with migraine. This association was obtained on 3,202 patients and 40,062 controls, and the combined results showed that the risk of migraine increases by 18% in the presence of the minor allele of rs1835740, with a stronger effect in MWA than in MWOA. This substitution is between the MTDH and PGCP genes.

Interestingly, these genes may simultaneously participate in glutamate homeostasis. In cultured astrocytes, MTDH (metadherin) suppresses the transcription of the EAAT2 gene that is main transporter of glutamate in astrocytes: this, in turn, causes an increase in glutamate concentrations in the synaptic cleft due to a delay in glutamate removal, thereby reducing the threshold for cortical spreading depression (CSD) that plays an important role in the pathophysiology of migraine [104,105]. However, subsequent studies failed to confirm the association between the MTDH gene and migraine [106-108], although its role in the development of clinical characteristics of migraine [107] and in the pathogenesis of other types of headaches [108] was demonstrated.

  1. The sample in a subsequent population GWAS consisted of women only and included 5,122 female patients and 18,108 controls [109]. Substitutions of rs2651899 (locus 1p36.32, gene PRDM16), rs2078371 (locus 1p13.2), rs10166942 (locus 2q37.1, gene TRPM8), rs17172526 (locus 7p14.2, gene SEPT7), rs2203834 (locus 8p22, gene C8orf79), rs13290757 (locus 9q33.3), rs11172113 (locus 12q13.3, gene LRP1) showed association with migraine at the first stage of the study. The association was confirmed for three out of seven substitutions in repeated analysis of three samples and with a combination of samples from the initial stage and repeated analysis (rs2651899, rs10166942 and rs11172113). For none of the three substitutions, association with migraine type (MWA and MWOA) or clinical features of migraine was demonstrated. TRPM8 is expressed in sensory neurons and neurons of the spinal ganglia. This gene encodes cold- and menthol-activated ion channels and participates in the generation of cold-induced pain [110]. Studying the role of TRPM8 in animal models with neuropathic pain also confirms the functional association with migraine [111]. LRP1 expresses in the brain tissues and in many other tissues [31], modulates synaptic transmission and interacts with the NMDA glutamatergic receptors. PRDM16 is a pleiotropic gene that is important for craniofacial development, fat determination and the proliferation of cardiac myocytes, neural and leukocyte neural progenitor cells [112]. It was shown that the Prdm16 mouse homologue acts as a negative TGF-β regulator (the TGFBR2 gene is also a candidate gene of migraine) [113]. Association with PRDM16, but not with LRP1 and TRPM8, has been recently reproduced in a Chinese Han sample [114]. At the same time, association with LRP1 and TRPM8 was found in Denmark and Iceland in a sample consisting of 2,523 patients and 38,170 controls, and meta-analysis confirmed the association for all the three loci [115].
  2. Lighthart et al. [116] conducted meta-analysis of migraine GWAS; in their work, they studied six European samples from the Dutch Icelandic migraine genetics consortium, involving 2,446 patients and 8,534 controls [116]. 32 SNP demonstrated a weak association with migraine. The best result was obtained for rs9908234 localized in the nerve growth factor receptor (NGFR) gene. However, this association was not repeated in three samples from the Netherlands and Australia. In repeated analysis of 18 SNP in two samples, the association was not reproduced. This study confirmed the association between migraine and the metadherine gene (MTDH), identified in the first GWAS.
  3. Freilinger et al. [117] tried to find SNP associated with MWOA and conducted a GWAS that included 2,326 patients with MWOA and 4,580 controls from the population of Germany and the Netherlands [117]. The association was verified additionally in four independent repeated European samples that included 2,508 patients with MWOA and 2,652 controls. The 1q22 locus contains 6 SNP, a considerable genome-wide association for which was received at the first stage of the study. Association for rs1050316 and rs3790455 was reproduced in repeated samples. All associated SNP were localized within the myocyte enhancer factor 2D gene (MEF2D). The MEF2D protein is a transcription factor and is highly expressed in the brain, where it regulates the differentiation of neurons and limits synaptic excitation [118,119]. Taking into account the involvement of glutamatergic neurotransmission in CSD and in the pathogenesis of migraine [104], and elevated plasma levels of glutamate in patients with migraine [120], MEF2D can be considered as a candidate gene of migraine. The 3p24 locus contains rs7640543 that showed association at an early stage in repeated samples and genome-wide significance in meta-analysis of pooled samples. This polymorphic variant is located in the gene encoding the transforming growth factor-beta receptor (TGFBR2). The TGFBR2 is involved in the regulation of cell proliferation, differentiation and extracellular matrix production [121]. The missense mutation p.Arg460His is associated with migraine headaches in 11 out of 14 carriers of the mutation in a large pedigree [122]. Locus 6p24, rs9349379 reached genome-wide significance in meta-analysis of pooled samples. This SNP is localized in the gene encoding phosphatase and actin regulator 1 (PHACTR1). The gene product controls synaptic activity and synaptic morphology by regulating binding of phosphatase 1 and actin proteins and is involved in the functioning of endothelial cells [123-125]. Locus 9p33, rs6478241 also reached genome-wide significance in meta-analysis of pooled samples. The rs6478241 substitution is localized in the ASTN2 gene that is a member of the astroactin gene family. This gene plays an important role in glial-directed migration that is necessary for the development of laminar architecture of the cortical regions in the brain [126]. For two SNP in the 2q37 locus, genome-wide significance was shown in meta-analysis that pooled all the samples (rs10166942 and rs17862920). The substitutions are localized in the TRPM8 gene. The 12q13 locus contains rs11172113, genome-wide significance of which was shown in meta-analysis.
  4. Cox et al. [127] conducted a GWAS based on pedigrees of an isolated population of Norfolk Island with high migraine prevalence (25.5%) [127]. During their work, they identified association with the substitution of rs4807347 localized in the ZNF555 gene encoding the "zinc finger 555" protein, which was confirmed in an independent sample (Women’s Genome Health Study, WGHS). This study also demonstrated association between the ADARB2 (rs883248, rs2271275, rs1046914, rs10903399), GRM7 (rs1391950 and rs11713183) and HTR7(rs2800143) genes and migraine phenotype in the population of Norfolk Island. The HTR7 and GRM7 genes related to the serotonergic system. These genes are expressed mainly in the brain, function in the presence of positively activated adenylate cyclase in a cell and may play a role in regulating the circadian rhythms and the neuroendocrine function, as well as in the development of affective mood disorders [128,129].
  5. The sample in a large meta-analysis was 23,285 patients with migraine and 95,425 controls [130]. The study identified 12 loci associated with susceptibility to migraine. Five loci were not previously associated with migraine (near to AJAP1 - 1p36, near to TSPAN2 - 1p13, inside FHL5 - 6q16, inside C7orf10 - 7p14 and near to MMP16 - 8q21). The remaining loci confirmed previous associations with migraine (PRDM16, MED2D, TRPM8, TGFBR2, PHATCTR1, ASTN2 and LRP1). The FHL5 gene encodes a transcription factor that regulates cAMP dependent elements CREM and CREB6, which play a role in synaptic plasticity and memory formation [131,132]. The c7orf10 (or SUGCT) gene encodes succinyl-CoA-glutarate-CoA transferase. Mutations in this gene are associated with phenotypically mild or even clinically asymptomatic glutaraciduria type III, a rare metabolism abnormality leading to a constant excretion (elimination) of glutaric acid [133]. AJAP1 is expressed in the brain and is associated with tumor invasion and the regulation of metalloproteinase activity [134]. TSPAN2, a member of the tetraspanin family, encodes a cell surface protein that mediates signal transduction in the regulation of cell development, activation, growth and motility. It was shown that TSPAN2acts as a metalloproteinase activity regulator [135]. The protein encoded by the MMP16 gene belongs to the metalloproteinase family, the members of which are widely expressed in human tissues and participate in the destruction of extracellular matrix in normal physiological processes. This protein cleaves the LRP1 protein encoded by another candidate gene of migraine [109].

Thus, migraine-related GWAS and subsequent meta-analyses identified associated polymorphic variants of susceptibility genes, which can be grouped into five pathways:

  1. Glutamatergic neurotransmission (rs1835740 - MTDH, rs11172113 - LRP1, rs3790455 - MEF2D);
  2. Synapse development and neuroplasticity (rs6478241 - ASTN2, rs13208321 - FHL5);
  3. Pain sensitivity (rs10166942 - TRPM8);
  4. Metalloproteinases (rs10504861 - near to MMP16, rs10915437 - near to AJAP1, rs12134493 - near to TSPAN2);
  5. Vascular system and metabolism (rs4379368 - C7orf10, rs2651899 - PRDM16, rs9349379 - PHACTR1, rs7640543 - near to TGFBR2).

Common variants demonstrated in several GWAS proved to be very valuable and underlined the glutamatergic role in the pathogenesis of migraine, with such role probably underlying cortical spreading depression and sensitization of nociceptive nerve endings [27]. Despite the fact that GWAS have identified new candidate genes responsible for the pathogenesis of migraine, the results of these studies have not brought us closer to understanding its molecular and genetic bases.

Conclusion

Due to the fact that the polymorphic variants of genes apparently have no significant effect on the pathogenesis of migraine individually, but rather there is an integrated effect of a complex genotype on pathogenesis, it is difficult to determine the contribution of polymorphic variants of individual genes. For example, the protein encoded by the LRP1 gene associated with migraine is cleaved by metalloproteinase that is encoded by another candidate gene, MMP16 [136]. Also, for most genes their role in the disease development processes remains unclear, as their cellular processes are not linked with the currently available data on the pathogenesis of migraine: TGFBR2, PHACTR1, С7orf10, ADARB2, ZNF555, etc.

Conflict of Interest

No conflict.

References

  1. Stovner Lj, Hagen K, Jensen R, Katsarava Z, Lipton R, et al. (2007) The global burden of headache: a documentation of headache prevalence and disability worldwide. Cephalalgia 27(3): 193-210.
  2. Steiner TJ, Stovner LJ, Birbeck GL (2013) Migraine: the seventh disabler. J Headache Pain 14: 1.
  3. Ayzenberg I, Katsarava Z, Sborowski A, Chernysh M, Osipova V, et al. (2014) Headache-attributed burden and its impact on productivity and quality of life in Russia: structured healthcare for headache is urgently needed. Eur Neurol J 21(5): 758-765.
  4. Osipova VV (2010) Current approaches to diagnosis and treatment of migraine (article in russian). Journal of Family Practice 2: 19-24.
  5. Katsarava Z, Limmroth V (2006) Is a combination of tramadol and acetaminophen effective for the treatment of acute migraine pain? Nat Clin Pract Neurol 2(7): 360-361.
  6. Loder E (2009) Migraine with aura and increased risk of ischaemic stroke. BMJ 27(339): b4380.
  7. Russell M, Iselius L, Olesen J (1995) Inheritance of migraine investigated by complex segregation analysis. Hum Genet 96(6): 726-730.
  8. Ulrich V, Gervil M, Kyvik KO, Olesen J, Russell MB (1999) Evidence of a genetic factor in migraine with aura: a population-based Danish twin study. Ann Neurol 45(2): 242-246.
  9. Gervil M, Ulrich V, Kyvik KO, Olesen J, Russell MB (1999) Migraine without aura: a population based twin study. Ann Neurol 46: 606-611.
  10. Mulder EJ, Van Baal C, Gaist D, Kallela M, Kaprio J, et al. (2003) Genetic and environmental influences on migraine: a twin study across six countries. Twin Res 6(5): 422-431.      
  11. Ziegler DK, Hur YM, Bouchard TJ Jr, Hassanein RS, Barter R (1998) Migraine in twins raised together and apart. Headache 38(6): 417-422.
  12. Svensson DA, Larsson B, Waldenlind E, Pedersen NL (2003) Shared rearing environment in migraine: results from twins reared apart and twins reared together. Headache 43(3): 235-244.
  13. Piterobon D and Striessing J (2003) Neurobiology of migraine. Nat Rev Neurosci 4(5): 386-398.
  14. de Vries B, Freilinger T, Vanmolkot KR, Koenderink JB, Stam AH, et al. (2007) Systematic analysis of three FHM genes in 39 sporadic patients with hemiplegic migraine. Neurology 69(23): 2170-2176.
  15.  De Vries B, Haan J, Frants RR, Van den Maagdenberg AM, Ferrari MD (2006) Genetic Biomarkers for Migraine. Headache. 46(7): 1059-1068.
  16. Lee H, Jen JC, Cha YH, Nelson SF, Baloh RW (2008) Phenotypic and Genetic Analysis of a Large Family With Migraine-Associated Vertigo. Headache 48(10): 1460-1467.
  17. Chabriat H, Vahedi K, Iba-Zizen MT, Joutel A, Nibbio A, et al. (1995) Clinical spectrum of CADASIL: a study of 7 families. Lancet 346(8980): 934-939.
  18. Joutel A, Corpechot C, Ducros A, Vahedi K, Chabriat H, et al. (1996) Notch3 mutations in CADASIL, a hereditary adult-onset condition causing stroke and dementia. Nature 383(6602): 707-710.
  19. Iso T, Hamamori Y, Kedes L (2003) Notch signaling in vascular development. Arterioscler Thromb Vasc Biol 23(4): 543-553.
  20. Alva JA, Iruela-Arispe ML (2004) Notch signaling in vascular morphogenesis. Curr Opin Hematol 11(4): 278-283.
  21. Ishiko A, Shimizu A, Nagata E, Takahashi K, Tabira T, et al. (2006) Notch3 ectodomain is a major component of granular osmiophilic material (GOM) in CADASIL. Acta Neuropathol 112(3): 333-339.
  22. Kaufmann P, Engelstad K, Wei Y, Kulikova R, Oskoui M, Sproule DM, et al. (2011) Natural history of MELAS associated with mitochondrial DNA m.3243A>G genotype. Neurology 77(22): 1965-1971.
  23. Finsterer J (2012) Inherited mitochondrial disorders. Adv Exp Med Biol 942: 187-213.
  24. Richards A, van den Maagdenberg AM, Jen JC, Kavanagh D, Bertram P, et al. (2007) C-terminal truncations in human 3’-5’ DNA exonuclease TREX1 cause autosomal dominant retinal vasculopathy with cerebral leukodystrophy. Nat Genet 39(9): 1068-1070.
  25. Terwindt GM, Haan J, Ophoff RA, Groenen SM, Storimans CW, et al. (1998) Clinical and genetic analysis of a large Dutch family with autosomal dominant vascular retinopathy, migraine and Raynaud’s phenomenon. Brain 121(Pt 2): 303-316.
  26. Ophoff RA, DeYoung J, Service SK, Joosse M, Caffo NA, et al. (2001) Hereditary vascular retinopathy, cerebroretinal vasculopathy, and hereditary endotheliopathy with retinopathy, nephropathy, and stroke map to a single locus on chromosome Am J Hum Genet 69(2): 447-453.
  27. Persico AM, Verdecchia M, Pinzone V, Guidetti V (2015) Migraine genetics: current findings and future lines of research. Neurogenetics 16(2): 77-95.
  28. Stam AH, Haan J, van den Maagdenberg AM, Ferrari MD, Terwindt GM (2009) Migraine and genetic and acquired vasculopathies. Cephalalgia 29(9): 1006-1017.
  29. Xu Y, Padiath QS, Shapiro RE, Jones CR, Wu SC, et al. (2005) Functional consequences of a CKIdelta mutation causing familial advanced sleep phase syndrome. Nature 434(7033): 640-644.
  30. Brennan KC, Bates EA, Shapiro RE, Zyuzin J, Hallows WC, et al. (2013) Casein kinase iδ mutations in familial migraine and advanced sleep phase. Sci Transl Med 5(183): 1-11.
  31. Lillis AP, Van Duyn LB, Murphy-Ullrich JE, Strickland DK (2008) LDL receptor-related protein 1: unique tissue-specifi c functions revealed by selective gene knockout studies. Physiol Rev 88(3): 887-918.
  32. Gould DB, Phalan FC, Breedveld GJ, van Mil SE, Smith RS, et al. (2005) Mutations in COL4A1 cause perinatal cerebral hemorrhage and porencephaly. Science 308(5725): 1167-1171.
  33. Breedveld G, de Coo IF, Lequin MH, Arts WF, Heutink P, et al. (2006) Novel mutations in three families confirms a major role of COL4A1 in hereditary porencephaly. J Med Genet 43(6): 490-495.
  34. van der Knaap MS1, Smit LM, Barkhof F, Pijnenburg YA, Zweegman S, et al. (2006) Neonatal porencephaly and adult stroke related to mutations in collagen IV A1. Ann Neurol 59(3): 504-511.
  35. Lanfranconi S, Markus HS (2010) COL4A1 mutations as a monogenic cause of cerebral small vessel disease: a systematic review. Stroke 41: e513-e518.
  36. Vahedi K, Massin P, Guichard JP, Miocque S, Polivka M, et al. (2003) Hereditary infantile hemiparesis, retinal arteriolar tortuosity, and leukoencephalopathy. Neurology 60(1): 57-63.
  37. Russell MB, Ducros A (2011) Sporadic and familial hemiplegic migraine: pathophysiological mechanisms, clinical characteristics, diagnosis, and management. Lancet Neurol 10(5): 457-470.
  38. Ophoff RA, Terwindt GM, Vergouwe MN, van Eijk R, Oefner PJ, et al. (1996) Familial hemiplegic migraine and episodic ataxia type-2 are caused by mutations in the Ca2+ channel gene CACNL1A4. Cell 87(3): 543-552.
  39. De Fusco M, Marconi R, Silvestri L, Atorino L, Rampoldi L, et al. (2003) Haploinsufficiency of ATP1A2 encoding the Na+/K+ pump alpha2 subunit associated with familial hemiplegic migraine type 2. Nat Genet 33(2): 192-196.
  40. Dichgans M, Freilinger T, Eckstein G, Babini E, Lorenz-Depiereux B, et al. (2005) Mutations in the neuronal voltage-gated sodium channel SCN1A in familial hemiplegic migraine. Lancet 336(9483): 371-377.
  41. Lea RA, Shepherd AG, Curtain RP, Nyholt DR, Quinlan S, et al. (2002) A typical migraine susceptibility region localizes to chromosome 1q31. Neurogenetics 4(1): 17-22.
  42. Freilinger T, Koch J, Dichgans M (2010) A novel mutation in SLC1A3 associated with pure hemiplegic migraine. J Headache Pain 11: 90.
  43. de Vries B, Mamsa H, Stam AH, Wan J, Bakker SL, et al. (2009) Episodic ataxia associated with EAAT1 mutation C186S affecting glutamate reuptake. Arch Neurol 66(1): 97-101.
  44. Méneret A, Gaudebout C, Riant F, Vidailhet M, Depienne C, et al. (2013) PRRT2 mutations and paroxysmal disorders. Eur J Neurol 20(6): 872-878.
  45. Maher BH, Griffiths LR (2011) Identification of molecular genetic factors that influence migraine. Mol Genet Genomics 285(6): 433-446.
  46. Corominas R, Ribases M, Camiña M, Cuenca-León E, Pardo J, et al. (2009) Two-stage case control association study of dopamine-related genes and migraine. BMC Med Genet 10: 95.
  47. Corominas R, Sobrido MJ, Ribasés M, Leon CE, Arias PB, et al. (2010) Association study of the serotoninergic system in migraine in the Spanish population. Am J Med Genet B Neuropsychiatr Genet 153: 177-184.
  48. Ishii M, Shimizu S, Sakairi Y, Nagamine A, Naito Y, et al. (20102) MAOA, MTHFR, and TNF-β genes polymorphisms and personality traits in the pathogenesis of migraine. Mol Cell Biochem 363: 357-366.
  49. Bayerer B, Engelbergs J, Savidou I, Boes T, Küper M, et al. (2010) Single nucleotide polymorphisms of the serotonin transporter gene in migraine – an association study. Headache 50: 319-322.
  50. Fernandez F, Colson N, Quinlan S, MacMillan J, Lea RA, et al. (2009) Association between migraine and a functional polymorphism at the dopamine beta-hydroxylase locus. Neurogenetics 10(3): 199-208.
  51. Todt U, Netzer C, Toliat M, Heinze A, Goebel I, et al. (2009) New genetic evidence for involvement of the dopamine system in migraine with aura. Hum Genet 125(3): 265-279.
  52. Mochi M, Cevoli S, Cortelli P, Pierangeli G, Soriani S, et al. (2003) A genetic association study of migraine with dopamine receptor 4, dopamine transporter and dopamine-beta-hydroxylase genes. Neurol Sci 23(6): 301-305.
  53. Formicola D, Aloia A, Sampaolo S, Farina O, Diodato D, et al. (2010) Common variants in the regulative regions of GRIA1 and GRIA3 receptor genes are associated with migraine susceptibility. BMC Med Genet 25: 103.
  54. Azimova J, Kondratieva N, Sergeev A, Skorobogatykh K, Kokaeva Z, et al. (2016) The Role of BDNF Gene Polymorphism in Formation of Clinical Characteristics of Migraine. J Neurol Stroke 4(2): 00123.
  55. Lafrenière RG, Rouleau GA (2012) Identification of novel genes involved in migraine. Headache 52(Suppl 2): 107-110.
  56. MacClellan LR, Howard TD, Cole JW, Stine OC, Giles WH, et al. (2009) Relation of candidate genes that encode for endothelial function to migraine and stroke : the Stroke Prevention in Young Women study. Stroke 40(10): e550-e557.
  57. Pizza V, Bisogno A, Lamaida E, Agresta A, Bandieramonte G, et al. (2010) Migraine and coronary artery disease:an open study on the genetic polymorphism of the 5, 10 methylenetetrahydrofolate (MTHFR) and angiotensin I-converting enzyme (ACE) genes. Cent Nerv Syst Agents Med Chem 10: 91-96.
  58. Colson NJ, Lea RA, Quinlan S, Griffiths LR (2006) The role of vascular and hormonal genes in migraine susceptibility. Mol Genet Metab 88(2): 107-113.
  59. Paterna S, Di Pasquale P, Cottone C, Seidita G, Cardinale A, et al. (1997) Migraine without aura and ACE-gene deletion polymorphism: is there a correlation? Preliminary findings. Cardiovasc Drugs Ther 11(4): 603-604.
  60. Kowa H, Fusayasu E, Ijiri T, Ishizaki K, Yasui K, et al. (2005) Association of the insertion/deletion polymorphism of the angiotensin I-converting enzyme gene in patients of migraine with aura. Neurosci Lett 374(2): 129-131.
  61. Joshi G, Pradhan S, Mittal B (2009) Role of the ACE ID and MTHFR C677T polymorphisms in genetic susceptibility of migraine in a north Indian population. J Neurol Sci 277(1-2): 133-137.
  62. Kowa H, Yasui K, Takeshima T, Urakami K, Sakai F, Nakashima K (2000) The homozygous C677T mutation in the methylenetetrahydrofolate reductase gene is a genetic risk factor for migraine. Am J Med Genet 96(6): 762-764.
  63. Oterino A, Valle N, Bravo Y, Muñoz P, Sánchez-Velasco P, et al. (2004) MTHFR T677 homozygosis influences the presence of aura in migraineurs. Cephalalgia 24(6): 491-494.
  64. Lea RA, Ovcaric M, Sundholm J, Sloyom L, MacMillan J, et al. (2005) Genetic variants of angiotensin converting enzyme and methylenetetrahydrofolate reductase may act in combination to increase migraine susceptibility. Brain Res Mol Brain Res 136(1-2): 112-117.
  65. Samaan Z, Gaysina D, Cohen-Woods S, Craddock N, Jones L, et al. (2011) Farmer A. Methylenetetrahydrofolate reductase gene variant (MTHFR C677T) and migraine: a case control study and meta-analysis. BMC Neurol 11: 66.
  66. An XK, Lu CX, Ma QL, Zhang XR, Burgunder JM, et al. (2013) Association of MTHFR C677T polymorphism with susceptibility to migraine in the Chinese population. Neurosci Lett 549: 78-81.
  67. Bahadir A, Eroz R, Dikici S (2013) Investigation of MTHFR C677T gene polymorphism, biochemical and clinical parameters in Turkish migraine patients: association with allodynia and fatigue. Cell Mol Neurobiol 33(8): 1055-1063.
  68. Azimova JE, Sergeev AV, Korobeynikova LA, Kondratieva NS, Kokaeva ZG, et al. (2013) Effects of MTHFR gene polymorphism on the clinical and electrophysiological characteristics of migraine. BMC Neurology 13-103.
  69. Liu R, Geng P2, Ma M3, Yu S4, Yang M, et al. (2014) MTHFR C677T polymorphism and migraine risk: a meta-analysis. J Neurol Sci 336(1-2): 68-73.
  70. Rubino E, Ferrero M, Rainero I, Binello E, Vaula G, et al. (2009) Association of the C677T polymorphism in the MTHFR gene with migraine: a meta-analysis. Cephalalgia 29(8) :818-825.
  71. Schürks M, Rist PM, Kurth T (2010) MTHFR 677C>T and ACE D/I polymorphisms in migraine: a systematic review and meta-analysis. Headache 50(4): 588-599.
  72. Scher AI, Eiriksdottir G, Garcia M (2013) Lack of association between the MTHFR C677T variant and migraine with aura in an older population: could selective survival play a role? Cephalalgia 33(5): 308-315.
  73. Federico A, Bianchi S, Dotti MT (2005) The spectrum of mutations for CADASIL diagnosis. Neurol Sci 26(2): 117-124.
  74. Ungaro C, Mazzei R, Conforti FL, Sprovieri T, Servillo P, et al. (2009) CADASIL: extended polymorphisms and mutational analysis of the NOTCH3 gene. J Neurosci Res 87(5): 1162-1167.
  75. Mosca L, Marazzi R, Ciccone A, Santilli I, Bersano A, et al. (2011) NOTCH3 gene mutations in subjects clinically suspected of CADASIL. J Neurol Sci 307(1-2): 144-148.
  76. Schwaag S, Evers S, Schirmacher A, Stögbauer F, Ringelstein EB, et al. (2006) Genetic variants of the NOTCH3 gene in migraine - a mutation analysis and association study. Cephalalgia 26(2): 158-161.
  77. Menon S, Cox HC, Kuwahata M, Quinlan S, MacMillan JC, et al. (2011) Association of a Notch 3 gene polymorphism with migraine susceptibility. Cephalalgia 31(3): 264-270.
  78. Tikka-Kleemola P, Kaunisto MA, Hämäläinen E, Todt U, Göbel H, et al. (2009) Genetic association study of endothelin-1 and its receptors EDNRA and EDNRB in migraine with aura. Cephalalgia 29(11): 1224-1231.
  79. Joshi G, Pradhan S, Mittal B (2011) Vascular gene polymorphisms (EDNRA −231 G>A and APOE HhaI) and risk for migraine. DNA Cell Biol 30: 577-584.
  80. Lemos C, Neto JL, Pereira-Monteiro J, Mendonça D, Barros J, et al. (2011) A role for endothelin receptor type A in migraine without aura susceptibility? A study in Portuguese patients. Eur J Neurol 18(4): 649-655.
  81. Tzourio C, Amrani M, Poirier O (2001) Association between migraine and endothelin type A receptor (ETA −231 A/G) gene polymorphism. Neurology 56: 1273-1277.
  82. Jia S, Ni J, Chen S, Jiang Y, Dong W, Gao Y (2011) Association of the pentanucleotide repeat polymorphism in NOS2 promoter region with susceptibility to migraine in a Chinese population. DNA Cell Biol 30(2): 117-122.
  83. de O S Mansur T, Gonçalves FM, Martins-Oliveira A, Speciali JG, Dach F (2012) Inducible nitric oxide synthase haplotype associated with migraine and aura. Mol Cell Biochem 364(1-2): 303-308.
  84. Borroni B, Rao R, Liberini P (2006) Endothelial nitric oxide synthase (Glu298Asp) polymorphism is an independent risk factor for migraine with aura. Headache 46(10): 1575-1579.
  85. Colson N, Fernandez F, Griffiths L (2010) Genetics of menstrual migraine: the molecular evidence. Curr Pain Headache Rep 14(5): 389-395.
  86. Colson NJ, Lea RA, Quinlan S, MacMillan J, Griffiths LR (2004) The estrogen receptor 1 G594A polymorphism is associated with migraine susceptibility in two independent case/control groups. Neurogenetics 5(2): 129-133.
  87. Colson NJ, Lea RA, Quinlan S, MacMillan J, Griffiths LR (2005) Investigation of hormone receptor genes in migraine. Neurogenetics 6(1): 17-23.
  88. Oterino A, Pascual J, Ruiz de Alegría C, Valle N, Castillo J, et al. (2006) Association of migraine and ESR1 G325C polymorphism. Neuroreport 17(1): 61-64.
  89. Oterino A, Toriello M, Cayón A, Castillo J, Colas R, et al. (2008) Multilocus analyses reveal involvement of the ESR1, ESR2, and FSHR genes in migraine. Headache 48(10): 1438-1450.
  90. Joshi G, Pradhan S, Mittal B (2010) Role of the oestrogen receptor (ESR1 PvuII and ESR1 325 C->G) and progesterone receptor (PROGINS) polymorphisms in genetic susceptibility to migraine in a North Indian population. Cephalalgia 30(3): 311-320.
  91. Ghosh J, Joshi G, Pradhan S, Mittal B (2012) Potential role of aromatase over estrogen receptor gene polymorphisms in migraine susceptibility: a case control study from North India. PLoS One 7(4): e34828.
  92. Rodriguez-Acevedo AJ, Maher BH, Lea RA, Benton M, Griffiths LR (2013) Association of oestrogen-receptor gene (ESR1) polymorphisms with migraine in the large Norfolk Island pedigree. Cephalalgia 33(14): 1139-1147.
  93. Schürks M, Rist PM, Shapiro RE, Kurth T (2010) Sex hormone receptor gene polymorphisms and migraine: a systematic review and meta-analysis. Cephalalgia 31(12): 1301-1314.
  94. Levy D (2012) Endogenous mechanisms underlying the activation and sensitization of meningeal nociceptors: the role of immunovascular interactions and cortical spreading depression. Curr Pain Headache Rep 16(3): 270-277.
  95. Rainero I, Grimaldi LM, Salani G (2004) Association between the tumor necrosis factor-alpha-308 G/A gene polymorphism and migraine. Neurology 62: 141-143.
  96. Mazaheri S, Hajilooi M, Rafiei A (2006) The G-308A promoter variant of the tumor necrosis factor-alpha gene is associated with migraine without aura. J Neurol 253(12): 1589-1593.
  97. Trabace S, Brioli G, Lulli P, Morellini M, Giacovazzo M, et al. (2002) Tumor necrosis factor gene polymorphism in migraine. Headache 42(5): 341-345.
  98. Lee KA, Jang SY, Sohn KM, Won HH, Kim MJ, et al. (2007) Association between a polymorphism in the lymphotoxin-a promoter region and migraine. Headache 47(7): 1056-1062.
  99. Dong W1, Jia S, Ye X, Ni J (2012) Association analysis of TNFRSF1B polymorphism with susceptibility for migraine in the Chinese Han population. J Clin Neurosci 19(5): 750-752.
  100. Rainero I, Fasano E, Rubino E, Rivoiro C, Valfrè W, et al. (2005) Association between migraine and HLA-DRB1 gene polymorphisms. J Headache Pain 6(4): 185-187.
  101. Dasdemir S, Cetinkaya Y, Gencer M, Ozkok E, Aydin M, et al. (2013) Cox-2 gene variants in migraine. Gene 518(2): 292-295.
  102. Yilmaz IA, Ozge A, Erdal ME, Edgünlü TG, Cakmak SE, et al. (2010) Cytokine polymorphism in patients with migraine: some suggestive clues of migraine and inflammation. Pain Med 11(4): 492-497.
  103. Anttila V, Stefansson H, Kallela M, Todt U, Terwindt GM, et al. (2010) International Headache Genetics Consortium. Genome-wide association study of migraine implicates a common susceptibility variant on 8q22.1. Nat Genet 42(10): 869-873.
  104. Kang DC, Su ZZ, Sarkar D, Emdad L, Volsky DJ, et al. (2005) Cloning and characterization of HIV-1-inducible astrocyte elevated gene-1, AEG-1. Gene 353(1): 8-15.
  105. Gasparini CF, Griffiths LR (2013) The biology of the glutamatergic system and potential role in migraine. Int J Biomed Sci 9(1): 1-8.
  106. Esserlind AL, Kirchmann M, Hauge AW, Le H, Olesen J (2012) A genotype-phenotype analysis of the 8q22.1 variant in migraine with aura. Eur J Neurol 19(4): 603-609.
  107. Christensen AF, Le H, Kirchmann M, Olesen J (2012) Genotypephenotype correlation in migraine without aura focusing on the rs1835740 variant on 8q22.1. J Headache Pain 13(1): 21-27.
  108. Azimova J, Kondratieva N, Sergeev A (2015) The Role of Polymorphism of Regulatory Region of MTDH Gene (Rs1835740) in Migraine and Other Forms of Primary Headaches. J Neurol Stroke 3(4): 00101.
  109. Chasman D, Schürks M, Anttila V, de Vries B, Schminke U, et al. (2011) Genome-wide association study reveals three susceptibility loci for common migraine in the general population. Nat Genet 43(7): 695-698.
  110. Proudfoot CJ, Garry EM, Cottrell DF, Rosie R, Anderson H, et al. (2006) Analgesia mediated by the TRPM8 cold receptor in chronic neuropathic pain. Curr Biol 16(16): 1591-1605.
  111. Biondi DM (2006) Is migraine a neuropathic pain syndrome?. Curr Pain Headache Rep 10(3): 167-178.
  112. Arndt AK, Schafer S, Drenckhahn JD, Sabeh MK, Plovie ER, et al. (2013) Fine mapping of the 1p36 deletion syndrome identifies mutation of PRDM16 as a cause of cardiomyopathy. Am J Hum Genet 93(1): 67-77.
  113. Bryan C. Bjork, Annick Turbe-Doan, Mary Prysak, Bruce J Herron, David R Beier, et al. (2010) Prdm16 is required for normal palatogenesis in mice. Hum Mol Genet 19(5): 774-789.
  114. Fan X, Wang J, Fan W, Chen L, Gui B, et al. (2014) Replication of migraine GWAS susceptibility loci in Chinese Han population. Headache 54(4): 709-715.
  115. Esserlind AL, Christensen AF, Le H, Kirchmann M, Hauge AW, et al. (2013) Replication and meta-analysis of common variants identifies a genome-wide significant locus in migraine. Eur J Neurol 20(5): 765-772.
  116. Lannie Ligthart, Boukje de Vries, Albert V Smith, M Arfan Ikram, Najaf Amin, et al. (2011) Meta-analysis of genome-wide association for migraine in six population-based European cohorts. Eur J Hum Genet 19(8): 901- 907.
  117. Freilinger T, Anttila V, de Vries B, Malik R, Kallela M, et al. (2012) International Headache Genetics Consortium . Genome-wide association analysis identifies susceptibility loci for migraine without aura. Nat Genet 44(7): 777-782.
  118. Shalizi A, Gaudillière B, Yuan Z, Stegmüller J, Shirogane T, et al. (2006) A calcium-regulated MEF2 sumoylation switch controls postsynaptic differentiation. Science 311(5763): 1012-1017.
  119. Flavell SW, Cowan CW, Kim TK, Greer PL, Lin Y, et al. (2006) Activity-dependent regulation of MEF2 transcription factors suppresses excitatory synapse number. Science 311(5763): 1008-1012.
  120. Ferrari MD, Odink J, Bos KD, Malessy MJ, Bruyn GW, et al. (1990) Neuroexcitatory plasma amino acids are elevated in migraine. Neurology 40(10): 1582-1586.
  121. Lin HY, Wang XF, Ng-Eaton E, Weinberg RA, Lodish HF (1992) Expression cloning of the TGF-beta type II receptor, a functional transmembrane serine/threonine kinase. Cell 68(4): 775-785.
  122. Law C, Bunyan D, Castle B, Day L, Simpson I, et al. (2006) Clinical features in a family with an R460H mutation in transforming growth factor beta receptor 2 gene. J Med Genet 43(12): 908-916.
  123. Allen PB, Greenfield AT, Svenningsson P, Haspeslagh DC, Greengard P (2004) Phactrs 1-4: a family of protein phosphatise 1 and actin regulatory proteins. Proc. Nat. Acad. Sci 101(18): 7187-7192.
  124. Greengard P, Allen PB, Nairn AC (1999) Beyond the Dopamine Receptor: the DARPP-32/Protein Phosphatase-1 Cascade. Neuron 23(3): 435-447.
  125. Jarray R, Allain B, Borriello L, Biard D, Loukaci A, et al. (2011) Depletion of the novel protein PHACTR-1 from human endothelial cells abolishes tube formation and induces cell death receptor apoptosis. Biochemie 93(10): 1668-1675.
  126. Wilson PM, Fryer RH, Fang Y, Hatten ME (2010) Astn2, a novel member of the astrotactin gene family, regulates the trafficking of ASTN1 during glial-guided neuronal migration. J. Neurosci 30(25): 8529-8540.
  127. Cox HC, Lea RA, Bellis C, Carless M, Dyer TD, et al. (2012) A genome-wide analysis of 'Bounty’ descendants implicates several novel variants in migraine susceptibility. Neurogenetics 13(3): 261-266.
  128. Bard JA, Zgombick J, Adham N, Vaysse P, Branchek TA, et al. (1993) Cloning of a novel human serotonin receptor (r5-HT7) positively linked to adenylate cyclase. J Biol Chem 268(31): 23422-23426.
  129. Vanhoenacker P1, Haegeman G, Leysen JE (2000) 5-HT7 receptors: current knowledge and future prospects. Trends Pharmacol Sci 21(2): 70-77.
  130. Anttila V, Winsvold B, Gormley P, Kurth T, Bettella F, et al. (2013) Genome-wide meta-analysis identifies new susceptibility loci for migraine. Nat Genet 45(8): 912-917.
  131. Dash PK, Hochner B, Kandel ER (1990) Injection of the cAMP-responsive element into the nucleus of Aplysia sensory neurons blocks long-term facilitation. Nature 345(6277): 718-721.
  132. Lee YS, Silva AJ (2009) The molecular and cellular biology of enhanced cognition. Nat Rev Neurosci 10(2): 126-140.
  133. Sherman EA, Strauss KA, Tortorelli S, Bennett MJ, Knerr I, et al. (2008) Genetic mapping of glutaric aciduria, type 3, to chromosome 7 and identification of mutations in c7orf10. Am J Hum Genet 83(5): 604-609.
  134. Schreiner A, Ruonala M, Jakob V, Suthaus J, Boles E, et al. (2007) Junction protein shrew-1 influences cell invasion and interacts with invasion-promoting protein CD147. Mol Biol Cell 18(4): 1272-1281.
  135. Marc A, Lafleur, Daosong Xu, Martin E. Hemler (2009) Tetraspanin proteins regulate membrane type-1 matrix metalloproteinase-dependent pericellular proteolysis. Mol Biol Cell 20(7): 2030-2040.
  136. Ferrari MD, Klever RR, Terwindt GM, Ayata C, van den Maagdenberg AM, et al. (2015) Migraine pathophysiology: lessons from mouse models and human genetics. Lancet Neurol 14(1): 65-80.
Creative Commons Attribution License

© . This is an open access article distributed under the terms of the, which permits unrestricted use, distribution, and build upon your work non-commercially.