Next Article in Journal
Molecular Insights into the Genetic Variability of ORF Virus in a Mediterranean Region (Sardinia, Italy)
Next Article in Special Issue
Idiopathic Intracranial Hypertension: Evaluation of Admissions and Emergency Readmissions through the Hospital Episode Statistic Dataset between 2002–2020
Previous Article in Journal
Association of SDF1 and MMP12 with Atherosclerosis and Inflammation: Clinical and Experimental Study
Previous Article in Special Issue
The Health Economic Evaluation of Bariatric Surgery Versus a Community Weight Management Intervention Analysis from the Idiopathic Intracranial Hypertension Weight Trial (IIH:WT)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Role of Metabolism in Migraine Pathophysiology and Susceptibility

by
Olivia Grech
1,2,
Susan P. Mollan
3,
Benjamin R. Wakerley
1,4,
Daniel Fulton
5,
Gareth G. Lavery
1,2 and
Alexandra J. Sinclair
1,2,4,*
1
Metabolic Neurology, Institute of Metabolism and Systems Research, College of Medical and Dental Sciences, University of Birmingham, Birmingham B15 2TT, UK
2
Centre for Endocrinology, Diabetes and Metabolism, Birmingham Health Partners, Birmingham B15 2TH, UK
3
Birmingham Neuro-Ophthalmology Unit, University Hospitals Birmingham NHS Foundation Trust, Birmingham B15 2TH, UK
4
Department of Neurology, Queen Elizabeth Hospital, University Hospitals Birmingham NHS Trust, Birmingham B15 2TH, UK
5
Institute of Inflammation and Ageing, University of Birmingham, Birmingham B15 2TT, UK
*
Author to whom correspondence should be addressed.
Life 2021, 11(5), 415; https://doi.org/10.3390/life11050415
Submission received: 1 April 2021 / Revised: 22 April 2021 / Accepted: 29 April 2021 / Published: 1 May 2021
(This article belongs to the Special Issue Idiopathic Intracranial Hypertension)

Abstract

:
Migraine is a highly prevalent and disabling primary headache disorder, however its pathophysiology remains unclear, hindering successful treatment. A number of key secondary headache disorders have headaches that mimic migraine. Evidence has suggested a role of mitochondrial dysfunction and an imbalance between energetic supply and demand that may contribute towards migraine susceptibility. Targeting these deficits with nutraceutical supplementation may provide an additional adjunctive therapy. Neuroimaging techniques have demonstrated a metabolic phenotype in migraine similar to mitochondrial cytopathies, featuring reduced free energy availability and increased metabolic rate. This is reciprocated in vivo when modelling a fundamental mechanism of migraine aura, cortical spreading depression. Trials assessing nutraceuticals successful in the treatment of mitochondrial cytopathies including magnesium, coenzyme q10 and riboflavin have also been conducted in migraine. Although promising results have emerged from nutraceutical trials in patients with levels of minerals or vitamins below a critical threshold, they are confounded by lacking control groups or cohorts that are not large enough to be representative. Energetic imbalance in migraine may be relevant in driving the tissue towards maximum metabolic capacity, leaving the brain lacking in free energy. Personalised medicine considering an individual’s deficiencies may provide an approach to ameliorate migraine.

1. Introduction

1.1. Background

Headache disorders are ranked the second most prevalent disease worldwide [1], with migraine in particular affecting 1 billion people [2]. Migraine is also the leading cause of disability amongst neurological disorders [3], and significantly reduces quality of life [4]. It severely disrupts sufferers personal [5] and work-related functionality [6] and is estimated to cost €93 billion in Europe, due to both health costs and lost productivity [7]. However, despite its significant prevalence and disability, the importance of migraine as a public health issue has only recently been recognised. Moreover, migraine remains underdiagnosed and often untreated [8], with the most commonly recommended treatment triptans only effective in ~60% of patients [9]. The relationship between episodic (<15 days with migraine per month) and chronic (≥15 days per month) migraine [10] is complex and there remains a crucial gap regarding the pathophysiology of headache, preventing the development of effective targeted drugs.

1.2. Headache Mechanisms

Research has since evolved from early concepts that migraine is solely vasculature driven [11], instead considering complex nociceptive structures and tissue excitability in its pathogenesis. Activation of the trigeminovascular system, consisting of sensory trigeminal nerve fibers which innervate cerebral blood vessels and dura mater, has long been hypothesised to underlie headache pain [12]. Prolonged central sensitisation of the system leads to hyperexcitability of trigeminal neurons and decrease in nociceptive threshold, which is hypothesised to drive the transition from episodic to chronic migraine [13]. Of importance is the neuropeptide calcitonin gene-related peptide (CGRP), which is both released and has receptors throughout the trigeminovascular system [14]. Clinical studies have demonstrated the causative role of CGRP in migraine by measuring increased levels in the circulation during attacks [15], and recording headache onset following CGRP infusion [16].
Migraine with aura is a subtype which features visual, sensory or central nervous system symptoms which often precede headache attack [10]. These symptoms are accompanied by a decrease in regional blood flow in the cortex and cortical spreading depression (CSD) [10,17]. CSD is a wave of depolarisation which propagates slowly (2 to 5 mm/min) across the cerebral cortex [18]. This leads to neurovascular changes, influx of Na+ and efflux of K+ and release of neuropeptides, further propagating the spread of depression [18,19]. These vast changes in tissue excitability are followed by a period of suppressed stimulated and spontaneous synaptic activity, termed neuronal silencing [20]. In addition to aura, clinical evidence has also associated CSD with traumatic brain injury, stroke and subarachnoid haemorrhage [17,21,22]. Animal studies have demonstrated the ability of CSD to activate the trigeminovascular system, particularly by its ability to release CGRP [23], implicating CSD in headache generation [24]. Emerging migraine therapeutics aim to target these mechanisms and include monoclonal antibodies and receptor antagonists. These aim to block either CGRP itself or its receptor and have provided effective pain relief in migraine [25,26,27,28,29].
There are a number of secondary headache conditions whose headache phenotype mimics migraine, such as post traumatic headache [30] and persistent post idiopathic intracranial hypertension (IIH) headache [31]. Headache in IIH is driven by raised intracranial pressure and reduction of intracranial pressure has been reported in some studies to reduce headache [10]. Despite resolution of papilledema and normalization of raised intracranial pressure, many suffer persistent post-IIH headache that imitates migraine [32]. A recent open label study of persistent post-IIH headaches evaluated the use of Erenumab, a monoclonal antibody that binds the CGRP ligand. CGRP therapy reduced frequency of monthly moderate/severe headache days by 71% and all headache days by 45% from baseline to 12 months [32]. This study brought new insights to the association of CGRP in IIH headache. Seven of the patients had a recurrence of their raised intracranial pressure (evidenced by papilloedema), yet no return of their headache, suggesting CGRP may have a role in headache attributed to IIH [33]. Similarly, CGRP was found to induce headache exacerbation with migraine-like features in patients with persistent post-traumatic headache attributed to mild traumatic brain injury in a recent randomized control trial [34]. The understanding of the biological underpinnings for conditions that mimic migraine are intriguing, where some have suggested shared biological foundations [35].

1.3. Energy Metabolism in Headache

A mismatch between brain energy supply and demand has also been hypothesised to contribute toward headache pathology, with many studies suggesting fundamental dysfunction of mitochondria [36]. Despite multiple clinical trials assessing the use of nutraceuticals to target this and support metabolic processes, these often lack suitable control groups, and the results remain contradictory [37,38]. Identifying specific dysfunctional metabolic pathways and targets has hampered the progress of therapeutic development. Although a multitude of metabolic and respiratory pathways are associated with headache disorders, this review will focus specifically on aspects which are identified as dysfunctional in migraine. This review will discuss the clinical and pre-clinical evidence evaluating metabolic perturbations with potential therapeutic value for migraine, and secondary headaches which mimic migraine, such as IIH.

2. Patient Studies

2.1. ATP-PCR System

ATP is the universal energy storage molecule and is mostly generated in the brain by oxidative phosphorylation of ADP in mitochondria. This is coupled to creatine kinase reactions which donate phosphate (Pi) from phosphocreatine (PCr) to ADP to re-synthesise ATP from ADP in the ATP-PCR system (Figure 1). This system allows rapid mobilization of limited high energy phosphates to regenerate ATP during metabolic stress. Phosphorus-31 nuclear magnetic resonance spectroscopy (31P-MRS) is a non-invasive tool capable of quantifying phosphorus-containing compounds including ATP and PCr in vivo in the brain [39]. It is also possible to calculate additional parameters such as ADP concentration, phosphorylation potential and V/Vmax (actual velocity of oxidative metabolism/the maximum oxidative capacity) [40].
31P-MRS studies have overwhelmingly highlighted aspects of mitochondrial dysfunction in migraine [40]. PCr content, indicative of free cellular energy, is significantly decreased at rest in the brain of familial hemiplegic migraine (FHM) and both migraine with and without aura patients [36,41,42]. Decreased PCr was also exhibited during migraine attacks in patients with aura [43] and in muscle following exercise in FHM patients [36], suggesting a systemic mitochondrial dysfunction. Increase in Pi is also demonstrated both between and during migraine attacks [36,41,43], resulting in a decreased PCr/Pi ratio [43,44,45]. Increased ADP [36,41,44] and percentage of V/Vmax during interictal phases is another consistent finding in migraine [36,46], indicating an increased metabolic rate and oxidative capacity, resulting in a lower energy reserve. In most studies, ATP concentration remained similar between patients and controls, except for Reyngoudt et al., who identified a significant decrease in ATP in migraine without aura patients [42]. To summarise, 31P-MRS findings in migraine patients indicate an imbalance between increased brain metabolism and decrease in free cellular energy availability, which has been hypothesised as a biochemical substrate for headache attack [42,47]. This pattern is typical of defective mitochondrial respiration with low PCr, high Pi and high ADP found in mitochondrial cytopathies [48], which hints that migraine may also share similar aspects of pathology.

2.2. Magnesium Availability

31P-MRS is also able to measure free cytosolic magnesium (Mg2+), an important coenzyme in the creatine kinase reaction. Mg2+ may have multiple targets in headache; low concentrations have been associated with spontaneous CSD [49], it is able to induce changes in vascular tone [50], and influence neurotransmitter release [51]. Reduced Mg2+ content has been exhibited both during headache attacks and interictal periods in migraine [52,53]. Decreased Mg2+ was also associated with reduced free energy released from ATP hydrolysis in several migraine subtypes [52]. Notably, reductions showed a trend in line with severity of clinical phenotype, with the lowest measurements in patients with migraine associated with stroke and highest in migraine without aura [52].
In addition to neuroimaging, direct measurements of serum Mg2+ have exhibited significant reductions during interictal periods in migraine patients compared to controls [54,55]. Total serum Mg2+ exhibited a negative linear relationship with migraine attack frequency [54], and a further reduction during attack [55]. Although this may suggest systemic Mg2+ deficiency, accurate testing is difficult to achieve due its compartmentalisation and absorption in the body, and there remains debate as to which form of Mg2+ to measure. In a study of chronic daily headache, ionized Mg2+ was significantly decreased in serum compared to controls, although there were no differences in total Mg2+ [56].
Studies attempting to assess the efficacy of Mg2+ for the treatment of headache have demonstrated conflicting results. Some have exhibited a prophylactic effect of oral Mg2+ (486–600 mg) at reducing migraine attack frequency and duration [57,58,59] whilst intravenous magnesium sulfate (1 g) demonstrated reduction in pain [37,60]. However, some trials proving its efficacy have been confounded by the lack of a control group, are not large enough to be conclusive [60,61], or have focused on a sub-cohorts of patients with lower serum Mg2+ [61]. Additionally a double-blind placebo-controlled study of oral Mg2+ observed no effect on migraine after 12 weeks treatment [59], and emergency department use of intravenous Mg2+ has demonstrated, in multiple studies, no significant pain relief [38,62]. Discrepancies in the effectiveness of Mg2+ treatment between trials could be due to variance in the serum levels of Mg2+ of participants, or route of administration [37,61]. Therefore, preliminary screening could identify patients who will better benefit from nutraceutical treatment, as those with ionised serum Mg2+ > 0.54 mmol/L did not respond to therapy. [61] Overall the current evidence does not convincingly demonstrate that Mg2+ in patients interacts with the pathophysiology of migraine.

2.3. Glycolysis and Glucose Metabolism

Metabolically challenging events such as fasting and exercise are established triggers of migraine attacks, further implicating the role of metabolism [63,64]. Migraineurs have demonstrated both impaired insulin sensitivity and higher fasting plasma insulin during interictal periods [65,66], with reduced insulin release during attacks [67]. Therefore, the metabolism of glucose has been of significant interest in migraine pathology.
18F-Fluorodeoxyglucose PET (18F-FDG PET) imaging allows the measurement of localised cerebral glucose metabolism with use of a radiotracer-labelled glucose analogue. Interictal periods in episodic migraineurs exhibited significant glucose hypometabolism in several regions involved in central pain processing, in comparison to controls [68]. A negative correlation was found between disease duration and lifetime headache frequency with glucose metabolism in the insula and anterior cingulate cortex. These results suggest repeated migraine attacks over time lead to progressive decline in glucose metabolism of central pain processes [68]. This is a similar finding in medication overuse headache (MOH) patients who exhibit hypometabolism in pain processing regions [69].
These perturbations, however, are not fixed and appear reversible with treatment. Hypometabolism in fronto-temporal areas in episodic migraineurs improved following three months of external trigeminal nerve stimulation, which also significantly decreased frequency of migraine attacks [70]. Brain regions recovered to almost normal glucose uptake following withdrawal from analgesics in MOH in all regions except the orbitofrontal cortex [69]. Although the underlying cause of hypometabolism in relation to migraine pathology remains unknown, these studies suggest that improving glucose utilisation may increase the threshold for sensitization in pain processing structures or prevent induction of migraine generating mechanisms.

2.4. Ketogenesis

In the absence of glucose and glycogen stores, ketone bodies are produced from fatty acids by astrocytes to serve as energetic substrate in a metabolic pathway known as ketogenesis. [71]. The ketogenic diet (low carbohydrates <50 mg/day [72]) is able to induce this pathway and has been effective for the treatment of severe epilepsy [73]. Given that tissue excitability is also altered in migraine, it has been hypothesised that this diet may be favourable at restoring brain metabolism and excitability in migraine pathophysiology [72].
Small (1–2 patients) case reports have noted migraine improvement in those on a ketogenic diet [74,75] and a proof-of-concept study demonstrated a significant reduction in attack frequency and number of days with headaches during the first month of ketogenesis [76]. A recent study observed alterations in cortical excitability as measured by visual and somatosensory-evoked potentials, in addition to reduced attack frequency and duration following one month of dieting [77]. However, the efficacy of this diet remains under scrutiny [78], and a pilot study of the modified Atkins diet (high-fat low-carbohydrates) for chronic daily headache in adolescents demonstrated no protective effect [79]. Moreover, although studies have demonstrated initial positive effects, clinical variables appear to worsen following the initial month of dieting, after which patients are out of the ketogenic phase [76]. There has been no strong evidence to outline the most conclusive diet for migraine prophylaxis and it is yet to be determined if it is the weight loss which is effective rather than the components of the diet.

2.5. Anaerobic Metabolism

Hypoxia contributes towards headache in disorders such as high-altitude headache and obstructive sleep apnoea [10]; however, it has also been hypothesised to trigger migraine attack [80,81]. Therefore, measuring anaerobic products of glycolysis including lactic acid (Figure 1) can provide insights into the role of hypoxia in migraine triggers and pathology.
Proton MRS (1H-MRS) is also a non-invasive imaging method which can provide measurements of neurotransmitters and metabolites including lactate [40]. This method has demonstrated elevated lactate in the brain during interictal periods in FHM [82] and migraine with aura [83,84]. It is thought that transient increases in lactate at rest and in the absence of hypoxia, may indicate a subtle mitochondrial dysfunction and aberrant glycolysis. These results, however, are controversial as lactate is endogenously present in low concentrations in healthy brain tissue, making subtle variations difficult to measure. Higher field strength (3T) 1H-MRS studies, measuring absolute rather than relative concentrations of metabolites, did not detect a significant increase in lactate in migraine without aura patients, in comparison to controls [85]. Furthermore, in a provocation study, normobaric hypoxia was able to induce migraine in aura patients and although lactate did increase, it was not significantly higher than in controls [80].
Direct measurements of lactate in blood or cerebrospinal fluid (CSF) have also demonstrated conflicting results in migraine. There have been key associations between lactate and CSD, with a marked increase in lactate/pyruvate ratio following 12 temporally grouped CSDs in a migraine-associated stroke patient [86]. However, there are discrepancies between lactate measurements from different sample types. For instance, significantly higher lactic acid concentrations have been exhibited in blood, serum and plasma for migraine [87,88], but only serum for tension-type headache [88,89]. Increased lactate measurements may be indicative of cerebral hypoxia and shifts to anaerobic respiratory processes. Therefore, it may be anticipated that pH in the brain tissue would decrease as a result of lactic-acidosis, however, this is not found in NMR studies [90].

2.6. Prophylactic Supplementation of Patients

In response to the evidence of metabolic dysfunction, there have been a plethora of trials assessing the efficacy of dietary supplements and nutraceuticals for headache prophylaxis. Coenzyme Q10 (CoQ10) is a key co-factor of the electron transport chain which has mostly been used in mitochondrial cytopathies to improve dysfunctional respiratory metabolism. Supplementation was able to reduce CSF lactate and pyruvate levels in mitochondrial encephalopathy, lactic acidosis, and stroke-like episodes (MELAS) [91] and 31P-MRS studies have demonstrated improvement in mitochondrial respiration. [92] Several emerging trials have also validated CoQ10′s efficacy in migraine, with oral supplementation (150–300 mg) capable of reducing migraine-attack frequency [93] with 50% reduction after 3 months [94,95] and reducing severity [94]. In addition to its role in mitochondrial respiration, CoQ10 also has an antioxidant and anti-inflammatory role, which would be beneficial in its treatment of headache. Supplementation in episodic migraine patients reduced inflammatory mediator TNF-α and CGRP, both thought to be involved in trigeminal sensitisation [96]. Although deficiency has been exhibited in juvenile cases of migraine [97], it is yet to be well documented in adults, but could illustrate a dysfunctional component of the mitochondrial respiratory chain.
Riboflavin (vitamin B2) is the precursor of flavin mononucleotide (FMN), a component of complex I, and flavin adenine dinucleotide (FAD) an electron donor in the mitochondrial respiratory chain. Riboflavin is able to reduce headache pain in MELAS, in which it is hypothesised to support complex I and II activity in the electron transport chain. [98] In migraine, treatment with oral riboflavin (400 mg) was able to reduce attack frequency [99] and severity [99,100], hinting towards shared pathology with MELAS.
There has been accumulating interest for the role of vitamin D in migraine, with several studies reporting low serum levels in migraine patients [101,102,103,104]. Moreover, the incidence of aura, phono/photophobia and resistance to medication was found to be significantly higher in migraine patients with a deficiency compared to those with normal levels [105]. Supplementation with vitamin D in migraine has demonstrated the ability to reduce headache frequency, headache diary result and migraine disability score [101,106,107]. Although the mechanism of vitamin D has remained elusive, a recent study revealed a significant reduction in serum CGRP concentration following 16 weeks of supplementation [107]. Future larger randomized control trials are needed to further investigate the role of vitamin D, as few studies have contradicted the findings of deficiency in migraine [108].
Palmitoylethanolamide (PEA) is an endogenous fatty acid amide abundant throughout the body including the central nervous system. It has analgesic and anti-inflammatory properties mediated by its binding to the peroxisome proliferator-activated receptor α (PPAR- α) [109,110,111]. PPAR- α also acts as a ligand binding transcription factor and stimulates fatty acid oxidation, [112] and PEA supplementation in animal models is able to modulate hepatic mitochondrial oxidative capacity [113] and glycolytic potential [114]. PEA has been effective at reducing pain in conditions of chronic pain and inflammation [115]; therefore it has become a nutraceutical of interest in migraine. In migraine with aura patients, PEA (1200 mg) and NSAIDs were able to significantly reduce pain intensity after 60 days in comparison to a group that received NSAIDs alone [116]. Similarly PEA was able to reduce headache frequency and intensity in a paediatric population with migraine without aura [117]. Use of nutraceuticals may be a useful additional headache therapy with a low side effect profile. Considering studies which have highlighted metabolite deficiencies, nutraceuticals may also pose as a personalised medicine dependant on an individual’s insufficiency.

3. Animal Data

Preclinical animal models have been invaluable in clarifying the pathophysiology of headache in addition to identifying therapeutic targets [118,119]. These models include mice expressing familial migraine mutations [120,121], models which sensitise the trigeminal system or directly induce CSD [122]. Studying metabolism in vivo is advantageous as respiratory processes can be easily manipulated and challenged, whilst more invasive biochemical methods can be utilised. Although direct measurements of pain cannot be achieved in rodents, appropriate behavioural studies can be used to provide a readout of this important clinical feature of headache [123].

3.1. Imaging and Labelling Studies

3.1.1. NADH and Oxidative Metabolism

Techniques using animal models benefit from mitochondrial respiration measurements in real-time and in vivo. Two-photon fluorescence imaging utilises the autofluoroscent reduced form of NADH, which donates an electron in the mitochondrial respiratory chain to become the oxidised and non-fluorescent NAD+ (Figure 1). Measuring NADH fluorescence in vivo allows the assessment of changes in mitochondrial redox potential, in addition to the occurrence of metabolically limiting hypoxia [124,125].
Following the widespread depolarisation of neurons and glia involved in CSD, there is a surge in ATP demand in cortical regions as the affected cells attempt to restore ionic gradients via ATPase Na+/K+ pumps [18]. This has established CSD as a metabolically demanding event. Multiple studies utilising NADH fluorescence measurements have characterised a typical response to CSD, in which there is a brief drop in fluorescence followed by prolonged overshoot [126,127,128]. This relates to biphasic neurovasculature changes in mice, including an initial increased cerebral blood flow followed by a prolonged period of vasoconstriction [129,130,131]. This can account for fluorescence patterns in which there is an initial surge in NADH oxidation and ATP production, corresponding with 31P-MRS studies that exhibit increased V/Vmax [36,46]. This is followed by a prolonged period during which NADH is not utilised and has been hypothesised to result from hypoxia that occurs, when oxidative demand exceeds the oxygen supply [127]. The amplitude of change in NADH fluorescence during this pattern was one order of magnitude higher than the typical 2–4% changes resulting from physiological activity [127]. Suggesting that these oxidative metabolic transitions are pathophysiological during CSD. Additionally, repetitive CSD induction led to gradual decrease in fluorescence, indicating tissue compromise and suggesting that chronic innervation may lead to long-term damage [128].
These studies have suggested that hypoxia could be a substrate for CSD in brain tissue. Specifically in mouse, hypoxia is capable of lowering the threshold [127] and increasing the duration of CSD [132]. Interestingly in FHM mice, tissue oxygenation as a result of CSD reached anoxia levels (1.16 ± 0.78 mmHg), in comparison to wildtype mice which reached hypoxic levels (13.69 ± 3.45 mmHg) [133]. The authors suggested FHM mice used more energy to restore ionic gradients, hinting that those with hereditary migraine are more susceptible to metabolic perturbations in response to CSD. NADH measurements support findings that CSD is a metabolically challenging event and suggest that individuals with pre-existing mitochondrial deficits may have a lowered threshold for headache generating mechanisms.

3.1.2. Glycolysis and Glucose Metabolism

Considering the significant evidence supporting the role of glucose metabolism and headache triggers, altering the glycaemic state in vivo can provide insights to the pathology underlying this susceptibility. Hyperglycaemia and hypoglycaemia were induced in rats by dextrose or insulin infusion [134]. Hyperglycaemia increased the threshold and reduced frequency of CSD events, whereas hypoglycaemia prolonged CSD duration and reduced KCl-induced CSD threshold by over 50% [134,135]. Another hypoglycaemic rat model induced by food restriction and insulin demonstrated faster CSD velocities than controls, which was reversed by injecting glucose [136]. Finally, CSD occurred spontaneously in insulin-treated rats (at blood glucose levels of 22 and 28 mg/dL) [137]. Apparent resistance to CSD in hyperglycaemia could be due to increased glucose availability which aids glycolysis and ATP production to maintain stimulus-induced rises in extracellular K+ [134]. The opposite therefore can be postulated for hypoglycaemia, which mirrors the glucose hypometabolism demonstrated in patients [68,69,138].
18FDG-PET tracing has also been utilised in rats in combination with microdialysis. This is able to provide an indication of extra-cellular availability of metabolites at the cortical surface during CSD. Unlike human studies, which exhibit hypometabolism in ictal periods, extracellular concentrations of glucose promptly decrease following CSD induction in rats, as it is rapidly utilized as an energy substrate (Figure 1) [139]. Cortical tissue exhibits a prolonged decrease in glucose content following CSD compared to contralateral unstimulated cortex, in conjunction with a rapid increase in lactate. Contradicting clinical studies, these changes in metabolites were accompanied by reduction in pH, indicating anaerobic glycolysis in response to CSD [140]. Moreover, 18FDG-PET confirmed an almost three-fold increase in lactate which was accompanied by extracellular acidification [141]. These studies further suggest an imbalance between ATP demand and oxygen availability as the tissue resorts to anaerobic processes. Utilisation of these alternative metabolic pathways may support ATP availability, and thus explain why no differences in ATP concentration were exhibited in 31P-MRS studies. Lactate was found to have a neuroprotective effect, reducing lesion size following excitotoxicity in ischemic stroke models [142]. Although it is a less energy efficient substrate, increased lactate in brain tissue may be a neuroprotective response to CSD.

3.2. Mitochondrial Studies in Animals

Whilst animal models allow the direct interrogation of mitochondrial integrity and function, few studies have been published utilising these methods. For instance, consumption of oxygen and extracellular acidification rates can be measured in vivo using the Seahorse XF analyser. This allows the quantification of mitochondrial and cytoplasmic respiratory pathways and has demonstrated alterations in a rat model of chronic migraine [143]. Using dural infusion of an inflammatory soup composed of histamine, serotonin, bradykinin and prostaglandin to induce trigeminal hypersensitivity, the model demonstrated a reduced spare respiratory capacity in trigeminal nucleus caudalis, a major component of the pain processing pathway [143]. Similarly, Clark-style oxygen electrodes able to measure oxygen consumption in mitochondria exhibited a significantly reduced mitochondrial membrane potential (the driving force for ATP synthesis) following CSD [144]. Although these findings mirror 31P-MRS studies indicating reduced mitochondrial energy reserves in migraine patients [17,45], it does not explain increased V/Vmax in these patients [36,46].
Furthermore, biochemical investigations have exhibited altered mitochondrial dynamics and biogenesis in trigeminal ganglion neurons following repeated inflammatory soup sensitization in rats [145]. Mitochondria exhibited fragmented structure, reduced DNA copy number, and alterations in mRNA and protein regulatory factors [145]. This disturbance was hypothesised to increase oxidative stress in tissue and potentially decrease the threshold for CSD events. Overall, these studies begin to pinpoint functional deficits in mitochondrial function, which may prove useful as therapeutic targets.

3.3. Supplementation of Animals

The use of Mg2+ supplementation in vivo has begun to unveil its role in headache pathology. Mg2+ has a vital function in NMDA receptors, imparting voltage sensitivity by blocking the receptor’s channel at resting membrane potentials and being removed during depolarization to allow Ca2+ influx and the activation of downstream signalling cascades. Trigeminal nerves are activated following direct stimulation of NMDA-receptors, making them important targets in migraine pain [146]. In mouse central neuron populations, the absence of Mg2+ increases the permeability of the receptor to cations [147], which may provide a mechanism for spontaneous CSD activity in rat hippocampal slices in low extracellular Mg2+ [49]. Mg2+ deficiency in headache patients, therefore, may be indicative of central sensitivity and the loss of NMDA-receptor blockade.
Unsurprisingly, these receptors have become an important target for migraine therapeutic development. MK-801 an NMDA-receptor antagonist, has demonstrated the ability to attenuate trigeminal nerve signalling in rats and cats [146,148]. Moreover, it has also demonstrated a role in abolishing CSD when induced in a rat brain [149]. Interestingly MgSO4 treatment was able to significantly reduce the number of CSDs induced when compared to untreated tissue [149]. MgSO4 treatment also increased latency times to anoxic depolarisation following cardiac arrest [149]. In rat brain slices, Mg2+ was able to improve recovery and maintain ATP levels during CSD in anoxic conditions, supporting its role in aiding respiration [150]. The ability of Mg2+ to block synaptic transmission and thereby reduce postsynaptic neuronal activation, lessens the metabolic burden required for homeostasis [150]. Moreover, the ability to perturbate spreading depression in anoxic conditions may be due to the role of Mg2+ in preserving energy mechanisms and ATP levels [149].

4. Conclusions

There is a wealth of clinical data characterising mitochondrial and metabolic dysfunction in migraine patients, similar to that of mitochondrial cytopathies [48]. Further in vivo investigation of these deficits demonstrates altered threshold for headache generating mechanisms including CSD. However, reproducibly pinpointing the exact deficits has been difficult to achieve both clinically and in vivo, which may contribute toward the mixed successes of nutraceutical trials. Understanding the metabolic deficiencies which increase susceptibility to CSD events may offer an approach to therapeutically recover mitochondrial energetic depletion, protect those susceptible to permanent tissue damage, and improve clinical phenotype.

Author Contributions

Conceptualization, O.G. and A.J.S.; methodology, O.G., S.P.M. and A.J.S.; software, O.G.; formal analysis, O.G. and A.J.S.; writing—original draft preparation, O.G., S.P.M., B.R.W., and A.J.S.; writing—review and editing, B.R.W., D.F. and G.G.L.; All authors have read and agreed to the published version of the manuscript.

Funding

O.G. is funded by a Brain Research UK PhD studentship, G.G.L. is supported by a Wellcome Trust Senior Fellowship (104612/Z/14/Z), A.J.S. is funded by a Sir Jules Thorn Award for Biomedical Science.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Figure was created using Biorender.com.

Conflicts of Interest

O.G.; consultancy work for Invex therapeutics (2020), S.P.M.; Invex therapeutics advisory board (2019); Heidelberg engineering speaker fees (2019), B.R.W.; consultancy work for Invex therapeutics (2020), A.J.S.; Novartis and Allergan Advisory board. Speaker fees Novartis. Invex therapeutics, company director with salary and stock options (2019, 2020).

References

  1. James, S.L.; Abate, D.; Abate, K.H.; Abay, S.M.; Abbafati, C.; Abbasi, N.; Abbastabar, H.; Abd-Allah, F.; Abdela, J.; Abdelalim, A.; et al. Global, regional, and national incidence, prevalence, and years lived with disability for 354 diseases and injuries for 195 countries and territories, 1990-2017: A systematic analysis for the Global Burden of Disease Study 2017. Lancet 2018, 392, 1789–1858. [Google Scholar] [CrossRef] [Green Version]
  2. Stovner, L.J.; Nichols, E.; Steiner, T.J.; Abd-Allah, F.; Abdelalim, A.; Al-Raddadi, R.M.; Ansha, M.G.; Barac, A.; Bensenor, I.M.; Doan, L.P.; et al. Global, regional, and national burden of migraine and tension-type headache, 1990–2016: A systematic analysis for the Global Burden of Disease Study 2016. Lancet Neurol. 2018, 17, 954–976. [Google Scholar] [CrossRef] [Green Version]
  3. Steiner, T.J.; Stovner, L.J.; Birbeck, G.L. Migraine: The seventh disabler. J. Headache Pain 2013, 14, 1. [Google Scholar] [CrossRef] [Green Version]
  4. Matharu, N.A.B.; Surat, T.; Giorgio, L.; Mariam, T.; Marjan, J.; Manjit, M. Quality of life in primary headache disorders: A review. Cephalalgia 2015, 36, 67–91. [Google Scholar] [CrossRef]
  5. Lipton, R.B.; Bigal, M.E.; Kolodner, K.; Stewart, W.F.; Liberman, J.N.; Steiner, T.J. The family impact of migraine: Population-based studies in the USA and UK. Cephalalgia 2003, 23, 429–440. [Google Scholar] [CrossRef] [PubMed]
  6. Stewart, W.F.; Lipton, R.B.; Simon, D. Work-related disability: Results from the American migraine study. Cephalalgia 1996, 16, 231–238. [Google Scholar] [CrossRef]
  7. Foundation, W. Migraine’s Impact on Employment in Europe. What Can be Done to Improve Work Outcomes for People with Migraine? Lancaster University: Lancaster, UK, 2019; p. 54. Available online: https://www.lancaster.ac.uk/media/lancaster-university/content-assets/documents/lums/work-foundation/Migraines-impact-on-employment-in-Europe-FINAL-pub-vA-accessible.pdf (accessed on 15 April 2021).
  8. Kernick, D.; Stapley, S.; Hamilton, W. GPs’ classification of headache: Is primary headache underdiagnosed? Br. J. Gen. Pract. J. R. Coll. Gen. Pract. 2008, 58, 102–104. [Google Scholar] [CrossRef] [Green Version]
  9. Loder, E. Triptan therapy in migraine. N. Engl. J. Med. 2010, 363, 63–70. [Google Scholar] [CrossRef] [Green Version]
  10. Headache Classification Committee of the International Headache Society (IHS). The International Classification of Headache Disorders, 3rd edition. Cephalalgia 2018, 38, 1–211. [Google Scholar] [CrossRef]
  11. Wolff, H.G. Headache and other head pain. In Headache and Other Head Pain; Oxford University Press: Oxford, UK, 2007; p. 648. [Google Scholar]
  12. Goadsby, P.J.; Lipton, R.B.; Ferrari, M.D. Migraine—Current understanding and treatment. N. Engl. J. Med. 2002, 346, 257–270. [Google Scholar] [CrossRef] [Green Version]
  13. Moulton, E.A.; Burstein, R.; Tully, S.; Hargreaves, R.; Becerra, L.; Borsook, D. Interictal dysfunction of a brainstem descending modulatory center in migraine patients. PLoS ONE 2008, 3, e3799. [Google Scholar] [CrossRef] [PubMed]
  14. Eftekhari, S.; Warfvinge, K.; Blixt, F.W.; Edvinsson, L. Differentiation of nerve fibers storing CGRP and CGRP receptors in the peripheral trigeminovascular system. J. Pain 2013, 14, 1289–1303. [Google Scholar] [CrossRef] [Green Version]
  15. Goadsby, P.J.; Edvinsson, L.; Ekman, R. Vasoactive peptide release in the extracerebral circulation of humans during migraine headache. Ann. Neurol 1990, 28, 183–187. [Google Scholar] [CrossRef] [PubMed]
  16. Lassen, L.H.; Haderslev, P.A.; Jacobsen, V.B.; Iversen, H.K.; Sperling, B.; Olesen, J. CGRP may play a causative role in migraine. Cephalalgia 2002, 22, 54–61. [Google Scholar] [CrossRef] [PubMed]
  17. Hadjikhani, N.; Sanchez Del Rio, M.; Wu, O.; Schwartz, D.; Bakker, D.; Fischl, B.; Kwong, K.K.; Cutrer, F.M.; Rosen, B.R.; Tootell, R.B.; et al. Mechanisms of migraine aura revealed by functional MRI in human visual cortex. Proc. Natl. Acad. Sci. USA 2001, 98, 4687–4692. [Google Scholar] [CrossRef] [Green Version]
  18. Leao, A.A.P. Spreading depression of activity in the cerebral cortex. J. Neurophysiol. 1944, 7, 359–390. [Google Scholar] [CrossRef]
  19. Sugaya, E.; Takato, M.; Noda, Y. Neuronal and glial activity during spreading depression in cerebral cortex of cat. J. Neurophysiol. 1975, 38, 822–841. [Google Scholar] [CrossRef]
  20. Kraig, R.P.; Nicholson, C. Extracellular ionic variations during spreading depression. Neuroscience 1978, 3, 1045–1059. [Google Scholar] [CrossRef]
  21. Strong, A.J.; Fabricius, M.; Boutelle, M.G.; Hibbins, S.J.; Hopwood, S.E.; Jones, R.; Parkin, M.C.; Lauritzen, M. Spreading and synchronous depressions of cortical activity in acutely injured human brain. Stroke 2002, 33, 2738–2743. [Google Scholar] [CrossRef] [Green Version]
  22. Dohmen, C.; Sakowitz, O.W.; Fabricius, M.; Bosche, B.; Reithmeier, T.; Ernestus, R.I.; Brinker, G.; Dreier, J.P.; Woitzik, J.; Strong, A.J.; et al. Spreading depolarizations occur in human ischemic stroke with high incidence. Ann. Neurol. 2008, 63, 720–728. [Google Scholar] [CrossRef]
  23. Tozzi, A.; de Iure, A.; Di Filippo, M.; Costa, C.; Caproni, S.; Pisani, A.; Bonsi, P.; Picconi, B.; Cupini, L.M.; Materazzi, S.; et al. Critical role of calcitonin gene-related peptide receptors in cortical spreading depression. Proc. Natl. Acad. Sci. USA 2012, 109, 18985–18990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Moskowitz, M.A.; Nozaki, K.; Kraig, R.P. Neocortical spreading depression provokes the expression of c-fos protein-like immunoreactivity within trigeminal nucleus caudalis via trigeminovascular mechanisms. J. Neurosci. 1993, 13, 1167–1177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Ho, T.W.; Mannix, L.K.; Fan, X.; Assaid, C.; Furtek, C.; Jones, C.J.; Lines, C.R.; Rapoport, A.M. Randomized controlled trial of an oral CGRP receptor antagonist, MK-0974, in acute treatment of migraine. Neurology 2008, 70, 1304. [Google Scholar] [CrossRef] [PubMed]
  26. Hewitt, D.J.; Aurora, S.K.; Dodick, D.W.; Goadsby, P.J.; Ge, Y.; Bachman, R.; Taraborelli, D.; Fan, X.; Assaid, C.; Lines, C.; et al. Randomized controlled trial of the CGRP receptor antagonist MK-3207 in the acute treatment of migraine. Cephalalgia 2011, 31, 712–722. [Google Scholar] [CrossRef]
  27. Olesen, J.; Diener, H.C.; Husstedt, I.W.; Goadsby, P.J.; Hall, D.; Meier, U.; Pollentier, S.; Lesko, L.M. Calcitonin gene-related peptide receptor antagonist BIBN 4096 BS for the acute treatment of migraine. N. Engl. J. Med. 2004, 350, 1104–1110. [Google Scholar] [CrossRef] [Green Version]
  28. Dodick, D.W.; Goadsby, P.J.; Silberstein, S.D.; Lipton, R.B.; Olesen, J.; Ashina, M.; Wilks, K.; Kudrow, D.; Kroll, R.; Kohrman, B.; et al. Safety and efficacy of ALD403, an antibody to calcitonin gene-related peptide, for the prevention of frequent episodic migraine: A randomised, double-blind, placebo-controlled, exploratory phase 2 trial. Lancet Neurol. 2014, 13, 1100–1107. [Google Scholar] [CrossRef]
  29. Dodick, D.W.; Goadsby, P.J.; Spierings, E.L.H.; Scherer, J.C.; Sweeney, S.P.; Grayzel, D.S. Safety and efficacy of LY2951742, a monoclonal antibody to calcitonin gene-related peptide, for the prevention of migraine: A phase 2, randomised, double-blind, placebo-controlled study. Lancet Neurol. 2014, 13, 885–892. [Google Scholar] [CrossRef]
  30. Ashina, H.; Iljazi, A.; Amin, F.M.; Ashina, M.; Lipton, R.B.; Schytz, H.W. Interrelations between migraine-like headache and persistent post-traumatic headache attributed to mild traumatic brain injury: A prospective diary study. J. Headache Pain 2020, 21, 134. [Google Scholar] [CrossRef]
  31. Mollan, S.P.; Hoffmann, J.; Sinclair, A.J. Advances in the understanding of headache in idiopathic intracranial hypertension. Curr. Opin. Neurol. 2019, 32, 92–98. [Google Scholar] [CrossRef] [Green Version]
  32. Yiangou, A.; Mitchell, J.L.; Fisher, C.; Edwards, J.; Vijay, V.; Alimajstorovic, Z.; Grech, O.; Lavery, G.G.; Mollan, S.P.; Sinclair, A.J. Erenumab for headaches in idiopathic intracranial hypertension: A prospective open-label evaluation. Headache J. Head Face Pain 2021, 61, 157–169. [Google Scholar] [CrossRef]
  33. Yiangou, A.; Mitchell, J.L.; Vijay, V.; Grech, O.; Bilton, E.; Lavery, G.G.; Fisher, C.; Edwards, J.; Mollan, S.P.; Sinclair, A.J. Calcitonin gene related peptide monoclonal antibody treats headache in patients with active idiopathic intracranial hypertension. J. Headache Pain 2020, 21, 116. [Google Scholar] [CrossRef] [PubMed]
  34. Ashina, H.; Iljazi, A.; Al-Khazali, H.M.; Christensen, C.E.; Amin, F.M.; Ashina, M.; Schytz, H.W. Hypersensitivity to Calcitonin Gene-Related Peptide in Post-Traumatic Headache. Ann. Neurol. 2020, 88, 1220–1228. [Google Scholar] [CrossRef]
  35. Ashina, H.; Moskowitz, M.A. Shared biological foundations of post-traumatic headache and migraine. Headache J. Head Face Pain 2021. [Google Scholar] [CrossRef] [PubMed]
  36. Uncini, A.; Lodi, R.; Di Muzio, A.; Silvestri, G.; Servidei, S.; Lugaresi, A.; Iotti, S.; Zaniol, P.; Barbiroli, B. Abnormal brain and muscle energy metabolism shown by 31P-MRS in familial hemiplegic migraine. J. Neurol. Sci. 1995, 129, 214–222. [Google Scholar] [CrossRef]
  37. Bigal, M.E.; Bordini, C.A.; Tepper, S.J.; Speciali, J.G. Intravenous magnesium sulphate in the acute treatment of migraine without aura and migraine with aura. A randomized, double-blind, placebo-controlled study. Cephalalgia Int. J. Headache 2002, 22, 345–353. [Google Scholar] [CrossRef]
  38. Corbo, J.; Esses, D.; Bijur, P.E.; Iannaccone, R.; Gallagher, E.J. Randomized clinical trial of intravenous magnesium sulfate as an adjunctive medication for emergency department treatment of migraine headache. Ann. Emerg. Med. 2001, 38, 621–627. [Google Scholar] [CrossRef]
  39. Kemp, G.J. Non-invasive methods for studying brain energy metabolism: What they show and what it means. Dev. Neurosci. 2000, 22, 418–428. [Google Scholar] [CrossRef]
  40. Reyngoudt, H.; Achten, E.; Paemeleire, K. Magnetic resonance spectroscopy in migraine: What have we learned so far? Cephalalgia 2012, 32, 845–859. [Google Scholar] [CrossRef] [Green Version]
  41. Barbiroli, B.; Montagna, P.; Cortelli, P.; Funicello, R.; Iotti, S.; Monari, L.; Pierangeli, G.; Zaniol, P.; Lugaresi, E. Abnormal brain and muscle energy metabolism shown by 31P-magnetic resonance spectroscopy in patients affected by migraine with aura. Neurology 1992, 42, 1209. [Google Scholar] [CrossRef] [PubMed]
  42. Reyngoudt, H.; Paemeleire, K.; Descamps, B.; De Deene, Y.; Achten, E. 31P-MRS demonstrates a reduction in high-energy phosphates in the occipital lobe of migraine without aura patients. Cephalalgia Int. J. Headache 2011, 31, 1243–1253. [Google Scholar] [CrossRef] [Green Version]
  43. Welch, K.M.A.; Levine, S.R.; Andrea, G.; Schultz, L.R.; Helpern, J.A. Preliminary observations on brain energy metabolism in migraine studied by in vivo phosphorus 31 NMR spectroscopy. Neurology 1989, 39, 538. [Google Scholar] [CrossRef]
  44. Sacquegna, T.; Lodi, R.; De Carolis, P.; Tinuper, P.; Cortelli, P.; Zaniol, P.; Funicello, R.; Montagna, P.; Barbiroli, B. Brain energy metabolism studied by 31P-MR spectroscopy in a case of migraine with prolonged aura. Acta Neurol. Scand. 1992, 86, 376–380. [Google Scholar] [CrossRef]
  45. Schulz, U.G.; Blamire, A.M.; Corkill, R.G.; Davies, P.; Styles, P.; Rothwell, P.M. Association between cortical metabolite levels and clinical manifestations of migrainous aura: An MR-spectroscopy study. Brain 2007, 130, 3102–3110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Montagna, P.; Cortelli, P.; Monari, L.; Pierangeli, G.; Parchi, P.; Lodi, R.; Iotti, S.; Frassineti, C.; Zaniol, P.; Lugaresi, E.; et al. 31P-magnetic resonance spectroscopy in migraine without aura. Neurology 1994, 44, 666–669. [Google Scholar] [CrossRef]
  47. Boska, M.D.; Welch, K.M.A.; Barker, P.B.; Nelson, J.A.; Schultz, L. Contrasts in cortical magnesium, phospholipid and energy metabolism between migraine syndromes. Neurology 2002, 58, 1227. [Google Scholar] [CrossRef] [PubMed]
  48. Barbiroli, B.; Montagna, P.; Martinelli, P.; Lodi, R.; Iotti, S.; Cortelli, P.; Funicello, R.; Zaniol, P. Defective brain energy metabolism shown by in vivo 31P MR spectroscopy in 28 patients with mitochondrial cytopathies. J. Cereb. Blood Flow Metab. 1993, 13, 469–474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Mody, I.; Lambert, J.D.; Heinemann, U. Low extracellular magnesium induces epileptiform activity and spreading depression in rat hippocampal slices. J. Neurophysiol. 1987, 57, 869–888. [Google Scholar] [CrossRef]
  50. Laurant, P.; Touyz, R.M.; Schiffrin, E.L. Effect of magnesium on vascular tone and reactivity in pressurized mesenteric resistance arteries from spontaneously hypertensive rats. Can. J. Physiol. Pharm. 1997, 75, 293–300. [Google Scholar] [CrossRef]
  51. Kuno, M.; Takahashi, T. Effects of calcium and magnesium on transmitter release at Ia synapses of rat spinal motoneurones in vitro. J. Physiol. 1986, 376, 543–553. [Google Scholar] [CrossRef] [Green Version]
  52. Lodi, R.; Lotti, S.; Cortelli, P.; Pierangeli, G.; Cevoli, S.; Clementi, V.; Soriani, S.; Montagna, P.; Barbiroli, B. Deficient energy metabolism is associated with low free magnesium in the brains of patients with migraine and cluster headache. Brain Res. Bull. 2001, 54, 437–441. [Google Scholar] [CrossRef]
  53. Ramadan, N.M.; Halvorson, H.; Vande-Linde, A.; Levine, S.R.; Helpern, J.A.; Welch, K.M. Low brain magnesium in migraine. Headache 1989, 29, 590–593. [Google Scholar] [CrossRef] [PubMed]
  54. Talebi, M.; Savadi-Oskouei, D.; Farhoudi, M.; Mohammadzade, S.; Ghaemmaghamihezaveh, S.; Hasani, A.; Hamdi, A. Relation between serum magnesium level and migraine attacks. Neurosciences 2011, 16, 320–323. [Google Scholar] [PubMed]
  55. Sarchielli, P.; Coata, G.; Firenze, C.; Morucci, P.; Abbritti, G.; Gallai, V. Serum and Salivary Magnesium Levels in Migraine and Tension-Type Headache. Results in a Group of Adult Patients. Cephalalgia 1992, 12, 21–27. [Google Scholar] [CrossRef]
  56. Mauskop, A.; Altura, B.T.; Cracco, R.Q.; Altura, B.M. Chronic daily headache—One disease or two? Diagnostic role of serum ionized magnesium. Cephalalgia 1994, 14, 24–28. [Google Scholar] [CrossRef]
  57. Peikert, A.; Wilimzig, C.; Köhne-Volland, R. Prophylaxis of Migraine with Oral Magnesium: Results From A Prospective, Multi-Center, Placebo-Controlled and Double-Blind Randomized Study. Cephalalgia 1996, 16, 257–263. [Google Scholar] [CrossRef]
  58. Köseoglu, E.; Talaslioglu, A.; Gönül, A.S.; Kula, M. The effects of magnesium prophylaxis in migraine without aura. Magnes Res. 2008, 21, 101–108. [Google Scholar] [PubMed]
  59. Pfaffenrath, V.; Wessely, P.; Meyer, C.; Isler, H.R.; Evers, S.; Grotemeyer, K.H.; Taneri, Z.; Soyka, D.; Göbel, H.; Fischer, M. Magnesium in the prophylaxis of migraine—A double-blind placebo-controlled study. Cephalalgia 1996, 16, 436–440. [Google Scholar] [CrossRef]
  60. Demirkaya, S.; Vural, O.; Dora, B.; Topçuoğlu, M.A. Efficacy of intravenous magnesium sulfate in the treatment of acute migraine attacks. Headache 2001, 41, 171–177. [Google Scholar] [CrossRef]
  61. Mauskop, A.; Altura, B.T.; Cracco, R.Q.; Altura, B.M. Intravenous magnesium sulphate relieves migraine attacks in patients with low serum ionized magnesium levels: A pilot study. Clin. Sci. 1995, 89, 633–636. [Google Scholar] [CrossRef] [Green Version]
  62. Frank, L.R.; Olson, C.M.; Shuler, K.B.; Gharib, S.F. Intravenous magnesium for acute benign headache in the emergency department: A randomized double-blind placebo-controlled trial. Can. J. Emerg. Med. 2004, 6, 327–332. [Google Scholar] [CrossRef] [Green Version]
  63. Kelman, L. The Triggers or Precipitants of the Acute Migraine Attack. Cephalalgia 2007, 27, 394–402. [Google Scholar] [CrossRef]
  64. Nadelson, C. Sport and exercise-induced migraines. Curr. Sports Med. Rep. 2006, 5, 29–33. [Google Scholar] [CrossRef]
  65. Rainero, I.; Limone, P.; Ferrero, M.; Valfrè, W.; Pelissetto, C.; Rubino, E.; Gentile, S.; Lo Giudice, R.; Pinessi, L. Insulin Sensitivity is Impaired in Patients with Migraine. Cephalalgia 2005, 25, 593–597. [Google Scholar] [CrossRef]
  66. Cavestro, C.; Rosatello, A.; Micca, G.; Ravotto, M.; Pia Marino, M.; Asteggiano, G.; Beghi, E. Insulin Metabolism is Altered in Migraineurs: A New Pathogenic Mechanism for Migraine? Headache J. Head Face Pain 2007, 47, 1436–1442. [Google Scholar] [CrossRef]
  67. Shaw, S.W.J.; Johnson, R.H.; Keogh, H.J. Metabolic changes during glucose tolerance tests in migraine attacks. J. Neurol. Sci. 1977, 33, 51–59. [Google Scholar] [CrossRef]
  68. Kim, J.H.; Kim, S.; Suh, S.I.; Koh, S.B.; Park, K.W.; Oh, K. Interictal metabolic changes in episodic migraine: A voxel-based FDG-PET study. Cephalalgia 2010, 30, 53–61. [Google Scholar] [CrossRef]
  69. Fumal, A.; Laureys, S.; Di Clemente, L.; Boly, M.; Bohotin, V.; Vandenheede, M.; Coppola, G.; Salmon, E.; Kupers, R.; Schoenen, J. Orbitofrontal cortex involvement in chronic analgesic-overuse headache evolving from episodic migraine. Brain 2006, 129, 543–550. [Google Scholar] [CrossRef]
  70. Magis, D.; D’Ostilio, K.; Thibaut, A.; De Pasqua, V.; Gerard, P.; Hustinx, R.; Laureys, S.; Schoenen, J. Cerebral metabolism before and after external trigeminal nerve stimulation in episodic migraine. Cephalalgia 2017, 37, 881–891. [Google Scholar] [CrossRef] [PubMed]
  71. Edmond, J.; Robbins, R.A.; Bergstrom, J.D.; Cole, R.A.; de Vellis, J. Capacity for substrate utilization in oxidative metabolism by neurons, astrocytes, and oligodendrocytes from developing brain in primary culture. J. Neurosci. Res. 1987, 18, 551–561. [Google Scholar] [CrossRef] [PubMed]
  72. Barbanti, P.; Fofi, L.; Aurilia, C.; Egeo, G.; Caprio, M. Ketogenic diet in migraine: Rationale, findings and perspectives. Neurol. Sci. 2017, 38, 111–115. [Google Scholar] [CrossRef] [PubMed]
  73. Sirven, J.; Whedon, B.; Caplan, D.; Liporace, J.; Glosser, D.; O’Dwyer, J.; Sperling, M.R. The Ketogenic Diet for Intractable Epilepsy in Adults: Preliminary Results. Epilepsia 1999, 40, 1721–1726. [Google Scholar] [CrossRef] [PubMed]
  74. Strahlman, R.S. Can Ketosis Help Migraine Sufferers? A Case Report. Headache J. Head Face Pain 2006, 46, 182. [Google Scholar] [CrossRef] [PubMed]
  75. Di Lorenzo, C.; Currà, A.; Sirianni, G.; Coppola, G.; Bracaglia, M.; Cardillo, A.; De Nardis, L.; Pierelli, F. Diet transiently improves migraine in two twin sisters: Possible role of ketogenesis? Funct. Neurol. 2013, 28, 305–308. [Google Scholar] [CrossRef] [PubMed]
  76. Di Lorenzo, C.; Coppola, G.; Sirianni, G.; Di Lorenzo, G.; Bracaglia, M.; Di Lenola, D.; Siracusano, A.; Rossi, P.; Pierelli, F. Migraine improvement during short lasting ketogenesis: A proof-of-concept study. Eur. J. Neurol. 2015, 22, 170–177. [Google Scholar] [CrossRef] [PubMed]
  77. Di Lorenzo, C.; Coppola, G.; Bracaglia, M.; Di Lenola, D.; Evangelista, M.; Sirianni, G.; Rossi, P.; Di Lorenzo, G.; Serrao, M.; Parisi, V.; et al. Cortical functional correlates of responsiveness to short-lasting preventive intervention with ketogenic diet in migraine: A multimodal evoked potentials study. J. Headache Pain 2016, 17, 58. [Google Scholar] [CrossRef] [Green Version]
  78. Maggioni, F.; Margoni, M.; Zanchin, G. Ketogenic diet in migraine treatment: A brief but ancient history. Cephalalgia 2011, 31, 1150–1151. [Google Scholar] [CrossRef]
  79. Kossoff, E.H.; Huffman, J.; Turner, Z.; Gladstein, J. Use of the modified Atkins diet for adolescents with chronic daily headache. Cephalalgia 2010, 30, 1014–1016. [Google Scholar] [CrossRef]
  80. Arngrim, N.; Schytz, H.W.; Britze, J.; Amin, F.M.; Vestergaard, M.B.; Hougaard, A.; Wolfram, F.; de Koning, P.J.H.; Olsen, K.S.; Secher, N.H.; et al. Migraine induced by hypoxia: An MRI spectroscopy and angiography study. Brain 2016, 139, 723–737. [Google Scholar] [CrossRef] [Green Version]
  81. Amery, W.K. Brain hypoxia: The turning-point in the genesis of the migraine attack? Cephalalgia 1982, 2, 83–109. [Google Scholar] [CrossRef]
  82. Grimaldi, D.; Tonon, C.; Cevoli, S.; Pierangeli, G.; Malucelli, E.; Rizzo, G.; Soriani, S.; Montagna, P.; Barbiroli, B.; Lodi, R.; et al. Clinical and neuroimaging evidence of interictal cerebellar dysfunction in FHM2. Cephalalgia 2010, 30, 552–559. [Google Scholar] [CrossRef]
  83. Sándor, P.S.; Dydak, U.; Schoenen, J.; Kollias, S.S.; Hess, K.; Boesiger, P.; Agosti, R.M. MR-spectroscopic imaging during visual stimulation in subgroups of migraine with aura. Cephalalgia 2005, 25, 507–518. [Google Scholar] [CrossRef] [PubMed]
  84. Watanabe, H.; Kuwabara, T.; Ohkubo, M.; Tsuji, S.; Yuasa, T. Elevation of cerebral lactate detected by localized 1H-magnetic resonance spectroscopy in migraine during the interictal period. Neurology 1996, 47, 1093–1095. [Google Scholar] [CrossRef] [PubMed]
  85. Reyngoudt, H.; Paemeleire, K.; Dierickx, A.; Descamps, B.; Vandemaele, P.; De Deene, Y.; Achten, E. Does visual cortex lactate increase following photic stimulation in migraine without aura patients? A functional 1H-MRS study. J. Headache Pain 2011, 12, 295–302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Santos, E.; Sanchez-Porras, R.; Dohmen, C.; Hertle, D.; Unterberg, A.W.; Sakowitz, O.W. Spreading depolarizations in a case of migraine-related stroke. Cephalalgia 2012, 32, 433–436. [Google Scholar] [CrossRef] [PubMed]
  87. Proia, P.; Amato, A.; Contrò, V.; Monaco, A.L.; Brusa, J.; Brighina, F.; Messina, G. Relevance of lactate level detection in migrane and fibromyalgia. Eur. J. Transl. Myol. 2019, 29, 8202. [Google Scholar] [CrossRef] [Green Version]
  88. Yavuz Altunkaynak, M.Ö. Devrimsel Harika Ertem, Betül Güveli, Filiz Uzun Okay, Zerrin Yıldırım, Belgin Mutluay, Ayten Ceyhan Dirican, Emine Altunkaynak, Ayhan Köksal, Sevim Baybaş. Serum lactic acid and pyruvic acid levels in patients with migraine and tension type headache. Dusunen Adam J. Psychiatry Neurol. Sci. 2013, 26, 276–280. [Google Scholar]
  89. Okada, H.; Araga, S.; Takeshima, T.; Nakashima, K. Plasma lactic acid and pyruvic acid levels in migraine and tension-type headache. Headache 1998, 38, 39–42. [Google Scholar] [CrossRef]
  90. Welch, K.M.; Levine, S.R.; D’Andrea, G.; Helpern, J.A. Brain pH in migraine: An in vivo phosphorus-31 magnetic resonance spectroscopy study. Cephalalgia 1988, 8, 273–277. [Google Scholar] [CrossRef]
  91. Abe, K.; Fujimura, H.; Nishikawa, Y.; Yorifuji, S.; Mezaki, T.; Hirono, N.; Nishitani, N.; Kameyama, M. Marked reduction in CSF lactate and pyruvate levels after CoQ therapy in a patient with mitochondrial myopathy, encephalopathy, lactic acidosis and stroke-like episodes (MELAS). Acta Neurol. Scand. 1991, 83, 356–359. [Google Scholar] [CrossRef]
  92. Barbiroli, B.; Iotti, S.; Lodi, R. Improved brain and muscle mitochondrial respiration with CoQ. An in vivo study by ^{31}P-MR spectroscopy in patients with mitochondrial cytopathies. Biofactors 1999, 9, 253–260. [Google Scholar] [CrossRef]
  93. Sándor, P.S.; Di Clemente, L.; Coppola, G.; Saenger, U.; Fumal, A.; Magis, D.; Seidel, L.; Agosti, R.M.; Schoenen, J. Efficacy of coenzyme Q10 in migraine prophylaxis: A randomized controlled trial. Neurology 2005, 64, 713. [Google Scholar] [CrossRef] [PubMed]
  94. Shoeibi, A.; Olfati, N.; Soltani Sabi, M.; Salehi, M.; Mali, S.; Akbari Oryani, M. Effectiveness of coenzyme Q10 in prophylactic treatment of migraine headache: An open-label, add-on, controlled trial. Acta Neurol. Belg. 2017, 117, 103–109. [Google Scholar] [CrossRef]
  95. Rozen, T.D.; Oshinsky, M.L.; Gebeline, C.A.; Bradley, K.C.; Young, W.B.; Shechter, A.L.; Silberstein, S.D. Open Label Trial of Coenzyme Q10 as A Migraine Preventive. Cephalalgia 2002, 22, 137–141. [Google Scholar] [CrossRef] [PubMed]
  96. Dahri, M.; Tarighat-Esfanjani, A.; Asghari-Jafarabadi, M.; Hashemilar, M. Oral coenzyme Q10 supplementation in patients with migraine: Effects on clinical features and inflammatory markers. Nutr. Neurosci. 2019, 22, 607–615. [Google Scholar] [CrossRef] [PubMed]
  97. Hershey, A.D.; Powers, S.W.; Vockell, A.L.; Lecates, S.L.; Ellinor, P.L.; Segers, A.; Burdine, D.; Manning, P.; Kabbouche, M.A. Coenzyme Q10 deficiency and response to supplementation in pediatric and adolescent migraine. Headache 2007, 47, 73–80. [Google Scholar] [CrossRef]
  98. Arts, W.F.; Scholte, H.R.; Bogaard, J.M.; Kerrebijn, K.F.; Luyt-Houwen, I.E. NADH-CoQ reductase deficient myopathy: Successful treatment with riboflavin. Lancet 1983, 2, 581–582. [Google Scholar] [CrossRef]
  99. Boehnke, C.; Reuter, U.; Flach, U.; Schuh-Hofer, S.; Einhaupl, K.M.; Arnold, G. High-dose riboflavin treatment is efficacious in migraine prophylaxis: An open study in a tertiary care centre. Eur. J. Neurol. 2004, 11, 475–477. [Google Scholar] [CrossRef]
  100. Schoenen, J.; Jacquy, J.; Lenaerts, M. Effectiveness of high-dose riboflavin in migraine prophylaxis. A randomized controlled trial. Neurology 1998, 50, 466–470. [Google Scholar] [CrossRef] [PubMed]
  101. Thys-Jacobs, S. Alleviation of migraines with therapeutic vitamin D and calcium. Headache 1994, 34, 590–592. [Google Scholar] [CrossRef]
  102. Togha, M.; Razeghi Jahromi, S.; Ghorbani, Z.; Martami, F.; Seifishahpar, M. Serum Vitamin D Status in a Group of Migraine Patients Compared With Healthy Controls: A Case–Control Study. Headache J. Head Face Pain 2018, 58, 1530–1540. [Google Scholar] [CrossRef]
  103. Mottaghi, T.; Khorvash, F.; Askari, G.; Maracy, M.R.; Ghiasvand, R.; Maghsoudi, Z.; Iraj, B. The relationship between serum levels of vitamin D and migraine. J. Res. Med. Sci. Off. J. Isfahan Univ. Med. Sci. 2013, 18, S66–S70. [Google Scholar]
  104. Song, T.J.; Chu, M.K.; Sohn, J.H.; Ahn, H.Y.; Lee, S.H.; Cho, S.J. Effect of Vitamin D Deficiency on the Frequency of Headaches in Migraine. J. Clin. Neurol. 2018, 14, 366–373. [Google Scholar] [CrossRef] [PubMed]
  105. Hussein, M.; Fathy, W.; Abd Elkareem, R.M. The potential role of serum vitamin D level in migraine headache: A case-control study. J. Pain Res. 2019, 12, 2529–2536. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Mottaghi, T.; Askari, G.; Khorvash, F.; Maracy, M.R. Effect of Vitamin D supplementation on symptoms and C-reactive protein in migraine patients. J. Res. Med. Sci. Off. J. Isfahan Univ. Med. Sci. 2015, 20, 477–482. [Google Scholar] [CrossRef]
  107. Ghorbani, Z.; Rafiee, P.; Fotouhi, A.; Haghighi, S.; Rasekh Magham, R.; Ahmadi, Z.S.; Djalali, M.; Zareei, M.; Razeghi Jahromi, S.; Shahemi, S.; et al. The effects of vitamin D supplementation on interictal serum levels of calcitonin gene-related peptide (CGRP) in episodic migraine patients: Post hoc analysis of a randomized double-blind placebo-controlled trial. J. Headache Pain 2020, 21, 22. [Google Scholar] [CrossRef] [Green Version]
  108. Zandifar, A.; Masjedi, S.s.; Banihashemi, M.; Asgari, F.; Manouchehri, N.; Ebrahimi, H.; Haghdoost, F.; Saadatnia, M. Vitamin D Status in Migraine Patients: A Case-Control Study. Biomed. Res. Int. 2014, 2014, 514782. [Google Scholar] [CrossRef]
  109. Costa, B.; Comelli, F.; Bettoni, I.; Colleoni, M.; Giagnoni, G. The endogenous fatty acid amide, palmitoylethanolamide, has anti-allodynic and anti-hyperalgesic effects in a murine model of neuropathic pain: Involvement of CB1, TRPV1 and PPARγ receptors and neurotrophic factors. Pain 2008, 139, 541–550. [Google Scholar] [CrossRef]
  110. Mazzari, S.; Canella, R.; Petrelli, L.; Marcolongo, G.; Leon, A. N-(2-Hydroxyethyl)hexadecanamide is orally active in reducing edema formation and inflammatory hyperalgesia by down-modulating mast cell activation. Eur. J. Pharmacol. 1996, 300, 227–236. [Google Scholar] [CrossRef]
  111. Lo Verme, J.; Fu, J.; Astarita, G.; La Rana, G.; Russo, R.; Calignano, A.; Piomelli, D. The Nuclear Receptor Peroxisome Proliferator-Activated Receptor-α Mediates the Anti-Inflammatory Actions of Palmitoylethanolamide. Mol. Pharmacol. 2005, 67, 15. [Google Scholar] [CrossRef]
  112. Minnich, A.; Tian, N.; Byan, L.; Bilder, G. A potent PPARα agonist stimulates mitochondrial fatty acid β-oxidation in liver and skeletal muscle. Am. J. Physiol. Endocrinol. Metab. 2001, 280, E270–E279. [Google Scholar] [CrossRef] [Green Version]
  113. Annunziata, C.; Lama, A.; Pirozzi, C.; Cavaliere, G.; Trinchese, G.; Di Guida, F.; Nitrato Izzo, A.; Cimmino, F.; Paciello, O.; De Biase, D.; et al. Palmitoylethanolamide counteracts hepatic metabolic inflexibility modulating mitochondrial function and efficiency in diet-induced obese mice. FASEB J. 2020, 34, 350–364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Gómez-Boronat, M.; Isorna, E.; Conde-Sieira, M.; Delgado, M.J.; Soengas, J.L.; de Pedro, N. First evidence on the role of palmitoylethanolamide in energy homeostasis in fish. Horm. Behav. 2020, 117, 104609. [Google Scholar] [CrossRef]
  115. Hesselink, J.M.K. Chronic idiopathic axonal neuropathy and pain, treated with the endogenous lipid mediator palmitoylethanolamide: A case collection. Int. Med. Case Rep. J. 2013, 6, 49–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Chirchiglia, D.; Cione, E.; Caroleo, M.C.; Wang, M.; Di Mizio, G.; Faedda, N.; Giacolini, T.; Siviglia, S.; Guidetti, V.; Gallelli, L. Effects of Add-On Ultramicronized N-Palmitol Ethanol Amide in Patients Suffering of Migraine With Aura: A Pilot Study. Front. Neurol. 2018, 9, 674. [Google Scholar] [CrossRef] [PubMed]
  117. Papetti, L.; Sforza, G.; Tullo, G.; Alaimo di Loro, P.; Moavero, R.; Ursitti, F.; Ferilli, M.A.N.; Tarantino, S.; Vigevano, F.; Valeriani, M. Tolerability of Palmitoylethanolamide in a Pediatric Population Suffering from Migraine: A Pilot Study. Pain Res. Manag. 2020, 2020, 3938640. [Google Scholar] [CrossRef] [Green Version]
  118. Andreou, A.P.; Summ, O.; Charbit, A.R.; Romero-Reyes, M.; Goadsby, P.J. Animal models of headache: From bedside to bench and back to bedside. Expert Rev. Neurother. 2010, 10, 389–411. [Google Scholar] [CrossRef] [PubMed]
  119. Bergerot, A.; Holland, P.R.; Akerman, S.; Bartsch, T.; Ahn, A.H.; MaassenVanDenBrink, A.; Reuter, U.; Tassorelli, C.; Schoenen, J.; Mitsikostas, D.D.; et al. Animal models of migraine: Looking at the component parts of a complex disorder. Eur. J. Neurosci. 2006, 24, 1517–1534. [Google Scholar] [CrossRef]
  120. van den Maagdenberg, A.M.; Pietrobon, D.; Pizzorusso, T.; Kaja, S.; Broos, L.A.; Cesetti, T.; van de Ven, R.C.; Tottene, A.; van der Kaa, J.; Plomp, J.J.; et al. A Cacna1a knockin migraine mouse model with increased susceptibility to cortical spreading depression. Neuron 2004, 41, 701–710. [Google Scholar] [CrossRef]
  121. Leo, L.; Gherardini, L.; Barone, V.; De Fusco, M.; Pietrobon, D.; Pizzorusso, T.; Casari, G. Increased Susceptibility to Cortical Spreading Depression in the Mouse Model of Familial Hemiplegic Migraine Type 2. PLoS Genet. 2011, 7, e1002129. [Google Scholar] [CrossRef] [Green Version]
  122. Noseda, R.; Burstein, R. Migraine pathophysiology: Anatomy of the trigeminovascular pathway and associated neurological symptoms, CSD, sensitization and modulation of pain. Pain 2013, 13, 1–9. [Google Scholar] [CrossRef] [Green Version]
  123. Yao, G.; Huang, Q.; Wang, M.; Yang, C.L.; Liu, C.F.; Yu, T.M. Behavioral study of a rat model of migraine induced by CGRP. Neurosci. Lett. 2017, 651, 134–139. [Google Scholar] [CrossRef]
  124. Vergen, J.; Hecht, C.; Zholudeva, L.V.; Marquardt, M.M.; Hallworth, R.; Nichols, M.G. Metabolic imaging using two-photon excited NADH intensity and fluorescence lifetime imaging. Microsc. Microanal. 2012, 18, 761–770. [Google Scholar] [CrossRef] [Green Version]
  125. Kasischke, K.A.; Lambert, E.M.; Panepento, B.; Sun, A.; Gelbard, H.A.; Burgess, R.W.; Foster, T.H.; Nedergaard, M. Two-photon NADH imaging exposes boundaries of oxygen diffusion in cortical vascular supply regions. J. Cereb. Blood Flow Metab. 2011, 31, 68–81. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Galeffi, F.; Somjen, G.G.; Foster, K.A.; Turner, D.A. Simultaneous monitoring of tissue PO2 and NADH fluorescence during synaptic stimulation and spreading depression reveals a transient dissociation between oxygen utilization and mitochondrial redox state in rat hippocampal slices. J. Cereb. Blood Flow Metab. 2011, 31, 626–639. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Takano, T.; Tian, G.F.; Peng, W.; Lou, N.; Lovatt, D.; Hansen, A.J.; Kasischke, K.A.; Nedergaard, M. Cortical spreading depression causes and coincides with tissue hypoxia. Nat. Neurosci. 2007, 10, 754–762. [Google Scholar] [CrossRef]
  128. Carlson, A.P.; Carter, R.E.; Shuttleworth, C.W. Vascular, electrophysiological, and metabolic consequences of cortical spreading depression in a mouse model of simulated neurosurgical conditions. Neurol. Res. 2012, 34, 223–231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Lauritzen, M.; Jøsrgensen, M.B.; Diemer, N.H.; Gjedde, A.; Hansen, A.J. Persistent oligemia of rat cerebral cortex in the wake of spreading depression. Ann. Neurol. 1982, 12, 469–474. [Google Scholar] [CrossRef] [PubMed]
  130. Fabricius, M.; Lauritzen, M. Transient hyperemia succeeds oligemia in the wake of cortical spreading depression. Brain Res. 1993, 602, 350–353. [Google Scholar] [CrossRef]
  131. Chang, J.C.; Shook, L.L.; Biag, J.; Nguyen, E.N.; Toga, A.W.; Charles, A.C.; Brennan, K.C. Biphasic direct current shift, haemoglobin desaturation and neurovascular uncoupling in cortical spreading depression. Brain 2010, 133, 996–1012. [Google Scholar] [CrossRef]
  132. Sonn, J.; Mayevsky, A. Responses to Cortical Spreading Depression under Oxygen Deficiency. Open Neurol. J. 2012, 6, 6–17. [Google Scholar] [CrossRef] [Green Version]
  133. Khennouf, L.; Gesslein, B.; Lind, B.L.; van den Maagdenberg, A.M.J.M.; Lauritzen, M. Activity-dependent calcium, oxygen, and vascular responses in a mouse model of familial hemiplegic migraine type 1. Ann. Neurol. 2016, 80, 219–232. [Google Scholar] [CrossRef] [PubMed]
  134. Hoffmann, U.; Sukhotinsky, I.; Eikermann-Haerter, K.; Ayata, C. Glucose modulation of spreading depression susceptibility. J. Cereb. Blood Flow Metab. Off. J. Int. Soc. Cereb. Blood Flow Metab. 2013, 33, 191–195. [Google Scholar] [CrossRef] [PubMed]
  135. Bures, J.; Buresova, O. Activation of latent foci of spreading cortical depression in rats. J. Neurophysiol. 1960, 23, 225–236. [Google Scholar] [CrossRef]
  136. Costa-Cruz, R.R.G.; Guedes, R.C.A. Cortical spreading depression during streptozotocin-induced hyperglycaemia in nutritionally normal and early-malnourished rats. Neurosci. Lett. 2001, 303, 177–180. [Google Scholar] [CrossRef]
  137. Astrup, J.; Norberg, K. Potassium activity in cerebral cortex in rats during progressive severe hypoglycemia. Brain Res. 1976, 103, 418–423. [Google Scholar] [CrossRef]
  138. Sprenger, T.; Ruether, K.V.; Boecker, H.; Valet, M.; Berthele, A.; Pfaffenrath, V.; Wöller, A.; Tölle, T.R. Altered Metabolism in Frontal Brain Circuits in Cluster Headache. Cephalalgia 2007, 27, 1033–1042. [Google Scholar] [CrossRef] [PubMed]
  139. Feuerstein, D.; Backes, H.; Gramer, M.; Takagaki, M.; Gabel, P.; Kumagai, T.; Graf, R. Regulation of cerebral metabolism during cortical spreading depression. J. Cereb. Blood Flow Metab. 2016, 36, 1965–1977. [Google Scholar] [CrossRef]
  140. Csiba, L.; Paschen, W.; Mies, G. Regional changes in tissue pH and glucose content during cortical spreading depression in rat brain. Brain Res. 1985, 336, 167–170. [Google Scholar] [CrossRef]
  141. Scheller, D.; Kolb, J.; Tegtmeier, F. Lactate and pH change in close correlation in the extracellular space of the rat brain during cortical spreading depression. Neurosci. Lett. 1992, 135, 83–86. [Google Scholar] [CrossRef]
  142. Berthet, C.; Lei, H.; Thevenet, J.; Gruetter, R.; Magistretti, P.J.; Hirt, L. Neuroprotective role of lactate after cerebral ischemia. J. Cereb. Blood Flow Metab. Off. J. Int. Soc. Cereb. Blood Flow Metab. 2009, 29, 1780–1789. [Google Scholar] [CrossRef]
  143. Fried, N.T.; Moffat, C.; Seifert, E.L.; Oshinsky, M.L. Functional mitochondrial analysis in acute brain sections from adult rats reveals mitochondrial dysfunction in a rat model of migraine. Am. J. Physiol. Cell Physiol. 2014, 307, C1017–C1030. [Google Scholar] [CrossRef] [Green Version]
  144. Li, F.; Qiu, E.; Dong, Z.; Liu, R.; Wu, S.; Yu, S. Protection of flunarizine on cerebral mitochondria injury induced by cortical spreading depression under hypoxic conditions. J. Headache Pain 2011, 12, 47–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Dong, X.; Guan, X.; Chen, K.; Jin, S.; Wang, C.; Yan, L.; Shi, Z.; Zhang, X.; Chen, L.; Wan, Q. Abnormal mitochondrial dynamics and impaired mitochondrial biogenesis in trigeminal ganglion neurons in a rat model of migraine. Neurosci. Lett. 2017, 636, 127–133. [Google Scholar] [CrossRef] [PubMed]
  146. Bereiter, D.A.; Bereiter, D.F.; Hathaway, C.B. The NMDA receptor antagonist MK-801 reduces Fos-like immunoreactivity in central trigeminal neurons and blocks select endocrine and autonomic responses to corneal stimulation in the rat. Pain 1996, 64, 179–189. [Google Scholar] [CrossRef]
  147. Nowak, L.; Bregestovski, P.; Ascher, P.; Herbet, A.; Prochiantz, A. Magnesium gates glutamate-activated channels in mouse central neurones. Nature 1984, 307, 462–465. [Google Scholar] [CrossRef] [PubMed]
  148. Classey, J.D.; Knight, Y.E.; Goadsby, P.J. The NMDA receptor antagonist MK-801 reduces Fos-like immunoreactivity within the trigeminocervical complex following superior sagittal sinus stimulation in the cat. Brain Res. 2001, 907, 117–124. [Google Scholar] [CrossRef]
  149. van der Hel, W.S.; van den Bergh, W.M.; Nicolay, K.; Tulleken, K.A.; Dijkhuizen, R.M. Suppression of cortical spreading depressions after magnesium treatment in the rat. Neuroreport 1998, 9, 2179–2182. [Google Scholar] [CrossRef]
  150. Kass, I.S.; Cottrell, J.E.; Chambers, G. Magnesium and cobalt, not nimodipine, protect neurons against anoxic damage in the rat hippocampal slice. Anesthesiology 1988, 69, 710–715. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Perturbations in metabolic pathways as demonstrated in patient and animal studies. Neuroimaging studies in patients have exhibited disturbances in the ATP-PCR pathway which indicate a low free energy availability coupled with an increased metabolic rate. Glucose hypometabolism has been noted in migraine patients in addition to increased lactate, a product of anaerobic glycolysis. In animal studies however glycolysis is increased due to the rapid increase in ATP demand following headache mechanisms. This rapid incline also results in rapid NADH oxidation, which may lead to tissue hypoxia and anaerobic respiration.
Figure 1. Perturbations in metabolic pathways as demonstrated in patient and animal studies. Neuroimaging studies in patients have exhibited disturbances in the ATP-PCR pathway which indicate a low free energy availability coupled with an increased metabolic rate. Glucose hypometabolism has been noted in migraine patients in addition to increased lactate, a product of anaerobic glycolysis. In animal studies however glycolysis is increased due to the rapid increase in ATP demand following headache mechanisms. This rapid incline also results in rapid NADH oxidation, which may lead to tissue hypoxia and anaerobic respiration.
Life 11 00415 g001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Grech, O.; Mollan, S.P.; Wakerley, B.R.; Fulton, D.; Lavery, G.G.; Sinclair, A.J. The Role of Metabolism in Migraine Pathophysiology and Susceptibility. Life 2021, 11, 415. https://doi.org/10.3390/life11050415

AMA Style

Grech O, Mollan SP, Wakerley BR, Fulton D, Lavery GG, Sinclair AJ. The Role of Metabolism in Migraine Pathophysiology and Susceptibility. Life. 2021; 11(5):415. https://doi.org/10.3390/life11050415

Chicago/Turabian Style

Grech, Olivia, Susan P. Mollan, Benjamin R. Wakerley, Daniel Fulton, Gareth G. Lavery, and Alexandra J. Sinclair. 2021. "The Role of Metabolism in Migraine Pathophysiology and Susceptibility" Life 11, no. 5: 415. https://doi.org/10.3390/life11050415

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

Article Metrics

Back to TopTop