Next Article in Journal
Duplication and Segregation of Centrosomes during Cell Division
Next Article in Special Issue
Pathophysiology and Therapy of Associated Features of Migraine
Previous Article in Journal
HHcy Induces Pyroptosis and Atherosclerosis via the Lipid Raft-Mediated NOX-ROS-NLRP3 Inflammasome Pathway in apoE−/− Mice
Previous Article in Special Issue
ATP-Sensitive Potassium Channels in Migraine: Translational Findings and Therapeutic Potential
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Could Experimental Inflammation Provide Better Understanding of Migraines?

by
Philip Victor Reducha
1,2,
Lars Edvinsson
1,3 and
Kristian Agmund Haanes
1,2,*
1
Department of Clinical Experimental Research, Glostrup Research Institute, Copenhagen University Hospital, Rigshospitalet Glostrup, 2600 Glostrup, Denmark
2
Department of Biology, Section of Cell Biology and Physiology, University of Copenhagen, 1017 Copenhagen, Denmark
3
Division of Experimental Vascular Research, Department of Clinical Sciences, Lund University Hospital, 221 00 Lund, Sweden
*
Author to whom correspondence should be addressed.
Cells 2022, 11(15), 2444; https://doi.org/10.3390/cells11152444
Submission received: 17 June 2022 / Revised: 29 July 2022 / Accepted: 4 August 2022 / Published: 6 August 2022
(This article belongs to the Special Issue Migraine Neuroscience: From Experimental Models to Target Therapy)

Abstract

:
Migraines constitute a common neurological and headache disorder affecting around 15% of the world’s population. In addition to other mechanisms, neurogenic neuroinflammation has been proposed to play a part in migraine chronification, which includes peripheral and central sensitization. There is therefore considerable evidence suggesting that inflammation in the intracranial meninges could be a key element in addition to calcitonin gene-related peptide (CGRP), leading to sensitization of trigeminal meningeal nociceptors in migraines. There are several studies that have utilized this approach, with a strong focus on using inflammatory animal models. Data from these studies show that the inflammatory process involves sensitization of trigeminovascular afferent nerve terminals. Further, by applying a wide range of different pharmacological interventions, insight has been gained on the pathways involved. Importantly, we discuss how animal models should be used with care and that it is important to evaluate outcomes in the light of migraine pathology.

1. Pathophysiology of Migraine

Migraine is a common neurological and headache disorder affecting around 15% of the world’s population (≈1 billion people, reported migraine in the previous year) and are three times more prevalent in women (around 18%) than in men (around 6%) [1,2]. It is the second cause of disability worldwide (behind only lower back pain) and the leading cause of neurological disability. Migraines share some similarities with other primary headache disorders in multiple aspects which can lead to misdiagnosis, but migraine also have several unique features. Various probable migraine triggers have been identified including genetic factors such as channelopathies that are familial or sporadic: maternal inheritance of mitochondrial encephalopathy, lactic acidosis, and stroke-like episodes; angiopathy inheritance, to name a few [3].
For a long time, the vascular theory of migraine was dominant. The theory was coined by Harold Wolff in the 1940s. He hypothesized that vasodilation of cranial vessels was the cause of migraines [4]. However, the vascular theory of migraine has since been challenged and by many replaced by neuronal theories which include the central and/or the peripheral nervous systems in migraine pathophysiology. Nevertheless, the vasculature has remained an important factor in migraine pathophysiology as the neurovascular theory of migraine is currently dominating. The neurovascular theory suggests involvement of the trigeminovascular system (TGVS) in the ictal phase of a migraine attack [5]. Neurogenic inflammation (NI) has also been proposed to play a part in migraine chronification. NI would consist of vasodilation, mast cell degranulation, plasma protein extravasation, platelet aggregation and the involvement of vasoactive peptides and inflammatory mediators such as calcitonin gene-related peptide (CGRP), substance P, and neurokinin A, as well as of prostaglandins [6]. However, besides CGRP, the involvement in migraine pathophysiology of the other neuropeptides, as well as mast cell degranulation and plasma extravasation is unclear and needs further research [7]. Peripheral and central sensitization have been proposed as likely causes of pain and involved in the process of chronification of migraine [8].
Sensitization is the process where the response to a stimulus is facilitated and the response itself increased in magnitude. The activation of endogenous inflammatory mediators (which are yet to be fully recognized) is believed to sensitize peripheral trigeminovascular neurons (or first-order neurons) through constant innervation of the dural meninges. Peripheral sensitization likely causes the throbbing nature of the headache, and CGRP has been observed to be involved in initiating and sustaining peripheral sensitization. The constant stimulation of the peripheral TGVS causes sensitization of second-order neurons (trigeminocervical complex, trigeminal nucleus caudalis (TNC)), third-order neurons (trigeminothalamic), as well as neurons in the brainstem and thalamus, causing central sensitization. Central sensitization may cause muscle tenderness, aversion to touch, as well as cephalic and extracephalic allodynia and could possibly be the reason that migraines turn chronic [9,10]. The neurovascular theory builds on the view that vasodilatation of cranial arteries excites the sensory perivascular nerves which cause either immediate or delayed headache pain [11]. The view suggests that vasodilatation activates cyclic nucleotides and potassium channels, resulting in a local rise in potassium ions which causes activation of the sensory nerve terminals in the dura mater [12].
The initiation of migraines is also unclear but likely occurs centrally, as fMRI scans have shown increased activation of the hypothalamus during the prodrome phase [13]. Cortical spreading depression (CSD) has been proposed to play a significant role in the initiation of a migraine attack. However, only 25% of migraineurs experience aura symptoms [14]. In migraines without aura (MoA), if a minor reduction in brain blood flow occurs it is not observed with current methods. Therefore, arguments have been made that CSD only causes aura in patients that have a specific genetic predisposition, which would explain the lack of experience of aura symptoms in MoA patients. Therefore, the role of CSD as a general migraine trigger has been questioned [15]. Regarding CSD and inflammation, there is no evidence of the depression itself being caused by inflammation. However, experimental CSD can trigger neurogenic meningeal inflammation and subsequently activate the trigeminovascular system [16,17]. However, this has not been confirmed in humans.
Although the origin and cause of migraine is not fully known, CGRP is an established key component in migraine attacks. For example, the plasma levels of CGRP are greatly increased during migraine attacks [18], and inhibition of CGRP release decreases both plasma levels of CGRP and the severity of migraine symptoms [18,19]. The role of CGRP in migraines is associated with CGRP being a strong vasodilator and involved in pain transmission from various parts of the head to the central nervous system. Therapies targeting CGRP release or its canonical receptor in the TGVS have been proven to ease symptoms and alleviate pain [5,20,21]. Binding of CGRP to its receptor (between RAMP1 and CLR) leads to conformational changes of the CGRP receptor, and activation of the Gαs subunit. The activated Gαs subunit binds to adenylate cyclase (transmembrane protein), which converts adenosine triphosphate (ATP) into the second messenger known as cyclic adenosine 3’,5’-cyclic monophosphate (cAMP). Intracellular cAMP levels are increased and causes an intracellular downstream signalling cascade. The increase in intracellular cAMP activates cAMP-dependent protein kinase A (PKA) signalling, which results in the activation of cAMP response element-binding protein (CREB), which finally affects gene transcription [22]. In the context of migraines, the increase in intracellular cAMP levels has been suggested to cause opening of hyperpolarization-activated cyclic nucleotide gated channels (HCN) or ATP-sensitive potassium (KATP) channels that could lead to hyperexcitability of the membrane and possibly be the reason for migraine pain [23].
CGRP has been observed in regions of the central nervous system (cortex, thalamus, hypothalamus, TNC etc.) [24,25]. Additionally, CGRP release has been linked to peripheral areas relevant to pain sensation during the ictal phase of migraine, such as the dura mater and trigeminal ganglion (TG) [18,26]. CGRP and the CGRP receptor are localized in proximity and in parallel to each other, as CGRP is found in great numbers in unmyelinated C-fibers, while CGRP receptors are localized in the Nodes of Ranvier of myelinated Aδ-fibers, both in the TG and dura mater [27]. The TGVS, where CGRP is found in high concentrations, consist of the trigeminal nerve (or fifth cranial nerve, CN V) which has three distinctive branches: the ophthalmic (V1), the maxillary (V2), and the mandibular (V3) branch. As reviewed by Walker and colleagues [28], V1 innervates and supplies strictly sensory information to intracranial dura matter and brain vessels as well as extracranially to the upper facial areas forehead and the scalp, and to orbit and the eyes, upper nasal cavities and skin above the eyes and upper parts of the nose. V2 innervates and conveys strictly sensory information to mid-facial regions, including inferior nasal cavities and sinuses, as well as the upper jaw. V3 innervates sensory information and has motor functions by innervating the muscles of lower facial areas, and the mandibles. In the context of migraines, it is believed that nociceptive information caused by the excitation of C-fibers and in turn of Aδ-fibers, passes along the trigeminal nerves (V1, and to lesser extend V2-V3) in a series of action potentials and merges in the TG. From the TG, the sensory information is sent further to second-order neurons in the trigeminocervical complex (TCC which contains: TNC, dorsal horn and upper cervical spinal horn C1-C2). The nociceptive information is passed on deeper into the central areas and processed by the nuclei of the brain stem, the thalamus and the hypothalamus. Visual, auditory, olfactory, somatosensory, and motor cortices receive the nociceptive signals and are believed to cause the characteristic migraine symptoms such as phonophobia, photophobia, headache pain or cognitive dysfunction [9].
CGRP is believed to play a big role during this communication of pain information [5,29]. It is postulated that during the ictal phase of migraines, neurons in the dura mater and other peripheral neurons receive increased stimulation causing the release of CGRP from C-fibers. CGRP binds to the CGRP receptor located in the Aδ-fibers, which further stimulates neurons, causing extended depolarization. Depolarization and CGRP release are continued along the trigeminovascular pathway, creating nociceptive signalling. The initial release of CGRP in the periphery is likely caused by the excitation of neural membranes from action potentials or through antidromic signalling, causing an opening of ion channels such as voltage gated sodium (Na+) ion channels. The opening of voltage-gated ion channels leads to an influx of calcium ions (Ca2+) into the cytosol provoking exocytosis of CGRP from vesicles and into the synaptic cleft; that CGRP can later be detected in blood samples from migraine patients [18].

2. Inflammation in Migraine

During the years there have been much work, mainly in rodents, suggesting a significant role of NI in migraines. Overall, the current view is that in single migraine attacks there is limited evidence of inflammation. However, many migraine cases evolve over time to be more severe and even enter a chronic phase. It is therefore possible that inflammation might play a role in migraine chronification, which is the current working hypothesis (Figure 1). To understand such a notion would assist in providing better therapy.
In the 1980s, the theory “sterile neurogenic inflammation” of the dural meninges was presented [30]. The involvement of inflammation in migraines was further supported by preclinical studies that showed significant inhibition of plasma protein extravasation with classic anti-migraine drugs, ergots and sumatriptan [31,32,33,34]. Considerable evidence suggested that NI in the intracranial meninges could be a key element resulting in the sensitization of trigeminal meningeal nociceptors in migraine [35,36,37,38].
Major cytokines have been implicated in the TGVS pathway resulting in NI, including tumor necrosis factor (TNF)-α, IL-1β and IL-6 [39,40]. There are studies that observed inflammatory molecules, in particular TNF-α, in plasma, serum or urine samples. These data suggested presence of inflammation in patients when comparing the levels during attacks and in attack-free intervals [41,42]. In animals and particularly in the knock-in animal model of familial hemiplegic migraine type-1, it has been shown that mRNA expression of TNF-α was increased in the TG [43].
Elevated TNF-α serum levels in humans were found even outside of attacks, supporting a role of TNF-α in migraine [44] the. However, data is variable as some have shown that TNF-α is only elevated immediately after the attack and is lower in-between attacks [42]. For other cytokines, such as IL-1β and IL-6, these were reported to be increased during a migraine attack [45]. Further, an increase in IL-6 and TNF-α has been detected in the jugular venous blood during a migraine attack [46]. Although changes in TNF-α and other cytokines can be confirmed, it is not clear as to their significance in migraine pathology. In migraine patients, there is often an increase in inflammatory markers over the course of the disease.
As CGRP might be released during the 72 h of the migraine attack [18], it could potentially evoke continuous activation of C- and Aδ-fibers. This in turn would lead to an increase in expression of inflammatory cytokines, not only in the dura mater, but possibly also in neuronal cell bodies, which are localized in the TG and the TNC. Here the term “neurogenic neuroinflammation”, which is defined by inflammatory reactions in the nervous system in response to neuronal activity, can be applied [47]. Combined with central effects which play an important part in causing pericranial allodynia, hyperpathia and muscular pericranial hyperactivity, part of the pathology could also be linked to the peripheral effects of CGRP (or inflammatory sensitization). For example, migraine-proving agents generate periorbital allodynia, for example in animals [48,49]. This is likely due to the release of inflammatory mediators (bradykinin, prostaglandins, etc.) from nerve endings or cells of the immune system [50,51,52]. Worth noting is that peptidergic fibers were not observed along or in between the striated muscle bundles of the temporal or the occipital regions, only in the skin or accompanying the blood vessels [53]. Furthermore, a part of this sensitization could also occur at the nodes of Ranvier [27]. How the nodes of Ranvier could relate to inflammation remains to be explored. There are several studies that have utilized the case that migraines are linked to meningeal/dural activation of nociceptive fibers. Data from these models supports that sensitization of trigeminovascular afferent nerve terminals. This has been studied using several different methods and particularly electrophysiology has given insight into changes in meningeal excitability [54], and has provided insight in modulation of ion channel excitability [54,55] and responses to migraine therapeutics [55]. Data using an inflammatory model has given evidence that triptans can induce hyperpolarizing shifts in meningeal nociceptive fibers [55]. Further, some sex specific difference can also be observed using this model, even in electrophysiological properties [56], adding value of this model in relation to sex differences in migraine pathology in humans.

3. The Inflammatory Soup (IS) Model and Its Variants

The most common inflammatory migraine model is to surgically insert a cannula that reaches the dura of anesthetized animals and inject inflammatory mediators through the cannula. This model was first applied by Oshinsky and colleagues [57] and has been applied in various migraine research papers since. Certain variations of the model have been used, notably in the type of animals used, changes in surgical procedures or variations in injection volume, time points and frequency. Besides using cannulas, some groups use supradural catheters [58].
The application of chemical stimulants through this model and its variants has led to various physiological and behavioural changes in animals. These findings allow us to evaluate the relevance of this model as a tool for migraine research and act as reference points if one desires to use this model. Here we will present some of the main findings. First, we look at the effect of IS. IS applied to the dura causes periorbital and/or plantar allodynia [57,58], however most often the allodynia is observed after chronic administration. Additionally, other behavioural impairments have been observed. There are reports of increased face rubbing [59] and grooming [60] which are indicative of head pain and/or irritancy. Furthermore, increased rest and freezing behaviour, and decreased exploratory behaviour have been reported [61], which are behaviours also experienced by migraine patients. On the contrary, Liu and colleagues [62] observed increased distance travelled by IS rats in the Open Field Test (OFT), which was argued to be due to higher anxiety levels.
Some other notable findings are the changes at the cellular level, and we have summarized all the below findings in supplementary tables, notably the effects in TG (supplementary Table S1), brain/spinal cord (supplementary Table S2) and TNC (supplementary Table S3). The levels of c-Fos, a marker of neuronal activation and nociception, have been observed at higher expression levels in the TG and TNC [63,64,65]. Increases of other nociceptive and neuronal activation markers in the TGVS have been observed, such as increases in cAMP [64], PKA [64], ERK1/2 [64], CREB [64], NGF [66], PKC [66], tyrosine receptor kinase B (TrkB) [67] and EAAT3 [67]. Increases of inflammatory markers have also been observed in the TGVS, notably an increase in IL-1β [65,68], IL-18 [69], TNF-α [65,70], p38 [67], BDNF [67], Iκβ [71], TRIF [71], TLR4 [71], Myd88 [71] and NF-κβ [71], FKN [72], CX3CR1 [72] and PGE2 [73]. Also, an increased presence of microglia [59,67,69,70,71,74,75] and astrocytes [69,74,75] has been reported. As mentioned previously, releases of neuropeptides such as CGRP, substance P and PACAP-38 are notable components of NI and these peptides are also linked to migraine pathophysiology, especially CGRP. Chronic application of IS on the dura causes upregulation of CGRP genes [62,63,66,67,73,75,76,77,78,79,80,81,82,83] and Substance P [76,80,81,82] in the TGVS. PACAP-38 is increased in the plasma of migraineurs during attacks but reduced in chronic migraineurs [84,85]. Long-term IS infusion decreased PACAP-38 which the authors argue is due to PACAP being depleted after chronic administration [86].
Other notable pathological alterations caused by IS application to the dura include a decrease of 5-HT in the brain [65], an increase in the 5-HT7 receptor expression [64], a decrease in α7nACh receptor expression in the hippocampus [75], an increase in nitric oxide (NO) [79], increases in the warm receptor TRPV1 [65] and cold receptor TRPM8 [87] expression, decreases in SIRT1 and PGC-1α [80], mitochondrial dysfunction [80,88], an altered functional connectivity between various brain regions [68,89,90,91,92], an increase in EphB2/EphrinB2 receptors [81], increased synaptic plasticity in TNC [76,79,81] but decreased plasticity in the hippocampus [93], an increase in glutamate in the TNC [94], an increase in mGluR5 [95], a decrease in GABA [94], decreases in GABABR1 and GABABR2 [94], increases in mTOR and autophagy [95], an increase in ASIC3 receptor expression [66,82], the increase of nNOS in the brain of rhesus monkeys [83], the impairment of descending inhibitory pathways [96], an increase in nerve growth factor (NGF) [66], increases of the P2Y14 receptor [59] and the P2X4 receptor expression [67], an increase in BBB permeability [74,97], the increase of VEGF [97], the sensitization of trigeminovascular neurons [96,98,99,100,101], an increase in white matter in different brain regions [102] and changes in the gut microbiota [65]. The application of other chemical stimulants that are inflammation-relevant have also been investigated. A chronic dural application of IL-6, a pro-inflammatory cytokine increased in the blood and serum of migraineurs [103], has been reported to cause periorbital and plantar allodynia [104,105,106,107], upregulate ERK1 [107] in the TG and sensitize dural afferents [107].
Treatments that can reduce the pathological changes of the chemical stimulants may give indications of what could be potential migraine treatments and could be used as positive controls in future studies that will use this model. Treatments that managed to reduce periorbital and/or plantar allodynia or other pathologies induced by IS application to the dura include sumatriptan [65,97,108], ketoprofen (NSAID) [73], nimesulide (NSAID) [73], etoricoxib (NSAID) [73], baclofen (GABAB receptor agonist) [95], propranolol (β-blocker) [101], amitriptyline (tricyclic antidepressant) [93], minocycline (tetracycline-derived antibiotic) [70], rapamycin (mTOR inhibitor) [95], anti-IL18 [69], anti-NGF [66], flunarizine (calcium channel blocker) [62], H89 (PKA inhibitor) [94], chelerythrin (PKC blocker) [66], APETx2 (ASIC3 inhibitor) [82], MPEP (mGluR5 antagonist) [95], PNU-282987 (α7nACh receptor agonist) [75], TNP-ATP (P2X inhibitor) [67], ibudilast (PDE inhibitor) [70], electroacupuncture [64,100,109], EphB1-Fc (EphB receptor inhibitor; only 0.5 µg) [81], TAK-242 (TLR4 inhibitor) [71], naltrexone (TLR4 antagonist) [70], PP2 (NR2B-pTyr inhibitor) [76], genistein (protein tyrosine kinase inhibitor) [76], ANA-12 (TrkB antagonist) [67], SIRT1720 (SIRT1 activator) [80], xiongmatang extract [62], and wuzhuyu decoction [65]. Meanwhile, treatments using APETx2 [104], ANA-12 [104], anisomycin (protein synthesis inhibitor) [110], 4EGI-1 (translation initiation inhibitor) [110] or U0126 (MEK inhibitor) [107] alleviated IL-6 induced periorbital/plantar allodynia.

4. Dural Activation by Complete Freund’s Adjuvant (CFA)

We have, in addition to experiments with IS, used additional preclinical models involving inflammatory stimulation of the peripheral parts of the trigeminal system to simulate the chronic nature of migraines without receptor activation. Firstly, we studied inflammatory pathways in cultured rat TG cells [111,112]. Secondly, we administered CFA (Complete Freund’s Adjuvant) into the temporomandibular joint (TMJ), which elicited activation of TG [113]. Thirdly, we developed a new animal model for trigeminal activation using chemical stimulation of the dura mater with CFA [114] to test whether application of CFA on a small part of the surface of the dura mater can cause long-term activation of the TG, and thus provide a model of migraine chronification [115]. Lastly, we evaluated whether activation of TNC and central sensitization occurs following CFA-induced activation of the dura mater [116].
Using cultures of isolated trigeminal neurons as a model for studies of neurons and glial cells, we found enhanced expression of CGRP in both neurons and satellite glia cells (SGCs) following inflammation. One of the most notable changes was in the mitogen-activated protein (MAP) kinase phosphorylation. The findings indicate that activation of a MAP kinase–dependent inflammatory signal pathway is involved in over-expression of CGRP in nociceptive neurons and could participate in generating pain hypersensitivity [117]. Looking further into in vivo inflammation, we showed that administration of CFA into the TMJ elicits activation of TG by increased expression of pERK1/2 (phosphorylated extracellular signal-regulated protein kinase), pp38 (phosphorylated p38 MAPK/ERK signaling pathway), CaMKII (Calmodulin-dependent protein kinase II), NF-κB and DREAM (Downstream regulatory element antagonist modulator) after 2 and 10 days. By applying CFA to induce local inflammation in the TMJ, this caused an upstream inflammation response in the TG where the TMJ sensory fibers have their cell bodies [113].
The inflammatory response does not only involve neurons, but importantly also SGCs, which together represent one anatomical and functional unit [113]. The idea of stimulating the nerves with inflammatory mediators was first shown by Takeda and colleagues, who injected CFA in the TMJ, and detected trigeminal activation. Importantly IL-1 receptor increased in the TG neurons, and this was combined with a potentiated excitability of Aδ-/C- fibers [118]. Furthermore, they later showed that the increased Aδ-fiber activity caused by inflammation could be blocked by an IL-1 receptor type 1 antagonist [119]. Others have shown that application of inflammatory substances onto the dura mater or chemical stimulation of the dural receptive fields causes hypersensitivity to mechanical and thermal stimulation together with direct activation of the TG [120].
The main application on the inflammatory investigation in rodents was the idea of applying the local inflammation of dura mater nociceptive fibers and the following activation in the TG. In the first papers we showed that the application of CFA [120,121] onto the dural surface indeed activated the TG, with the most remarkable changes seen in the expression of pERK1/2, IL-1β and CGRP positive nerve fibers in the TG [115,122]. This illustrates that the application of CFA onto the dura mater could be used as an animal model for long-term activation of the TGVS [115]. We further used this model to show that the application of CFA also induced activation (increased expression of c-Fos) of the central part of the TGVS: the TNC and C1-C2 regions of the spinal cord [116]. Interestingly, this inflammation could be blocked by the administration of a kynurenic acid analogue (SZR72), which is a precursor of an excitotoxin antagonist and anti-inflammatory substance [116,122]. Linking CFA to neurogenic neuroinflammation, masseteric injection of CFA caused spontaneous orofacial pain behaviors, neuronal activation in the TNC, and the release of interleukin-6 (IL-6). In this short-term study, pretreatment with a CGRP antagonist reduced pain behaviors. Nevertheless, IL-6 was unaffected by MK-8825 [123] but the study would probably have benefitted from a longer time frame than the 24 h used.

5. IS vs. CFA

Questions should also be asked about what type of inflammation model would best fit a migraine model. There are some notable differences between the IS and CFA. Usually, the IS consists of inflammatory mediators endogenous to what the body would release during an immune response like histamine, bradykinin, prostaglandins E2 and serotonin, and they are usually mixed into acidic PBS. Meanwhile, CFA is a water/oil emulsion consisting of dried and inactivated mycobacteria (most often Mycobacterium tuberculosis). IS induces a sterile inflammatory environment while CFA can go beyond and involve the adaptive immunity as well [124]. CFA is therefore generating local inflammatory markers, without direct receptor-induced activation which is opposite to IS.
One could argue that in the case of most migraines, it is more likely that a sterile inflammation is occurring caused by tissue damage and repair mechanisms initiated from prolonged neuronal activity and surplus of energy used, giving the upper hand to IS as being more relevant in migraines. However, CFA should not be dismissed, as CFA has been reliably used for multiple inflammation and pain models, and as mentioned earlier CFA creates similar cellular and sensitizing effects as IS when applied to the dura [115,116,125]. Further, one could question if some components of the IS could lead to indirect release of other signaling molecules such as ATP [126,127]. In addition to reactions to histamine and bradykinin, this has also been reported for other compounds known to induce itch [128,129] or pain [130]. Secondary effects from ATP could then affect mast cells [131], sensory neurons [132,133] or for example stimulate the release of CGRP [134], further complicating the interpretation of the data.
One could also consider whether the recipe in the IS should be revised, adding some inflammatory mediators that have been observed in the serum of migraine patients like IL-6, CGRP and TNF-α. Arguments could be made whether 5-HT should be removed from the IS mixture as 5-HT have potential anti-migraine effects, as triptans and ditans are agonists of 5-HT1B/D/F. receptors. However, as mentioned earlier, applying IS reduces 5-HT in the TGVS in one paper [65] and it is still uncertain whether 5-HT levels fluctuates in migraineurs [135]. Another aspect to take under consideration when using an inflammation model, is the possibility of surgical procedures creating inflammation on their own, and so using SHAM animals should be highly considered. In addition, anaesthesia, antibiotics and painkillers used for recovery could potentially have a modulatory effect on the induced inflammation.

6. Evaluating Outcomes in Animal Inflammation Models

The final goal of the animal research into migraines and the inflammation aspects thereof is to gain mechanistic insight into the pathology, we therefore address some important considerations when evaluating outcomes in inflammation studies. It is important when looking for translational aspects and identifying novel therapeutic targets that the model and outcomes represent true relatable outcomes. The main concept is that inflammatory stimulation of the dura will affect the nociceptive input from the meninges, with a headache most likely occurring in the animals due to the activated nociceptors from the meninges (Figure 2).
The inflammatory approach usually includes a craniotomy, which is an invasive procedure, might have relevance in interpreting the data. As an example, the paper by Laborc and colleagues highlights that the mere attachment of a rat in a stereotaxic frame can activate the trigeminal system [136]. This highlights the additions made by using both fresh controls and also shams in the uncovering of the true role of inflammation, and its separation from the invasive measures [137]. As we have seen above, some of the endpoints are similar as seen in human migraine attacks, but these studies will always be limited by the lack of direct communication with animals. Below, we discuss some limitations linked to outcomes in animal models.
Although photophobia is a known symptom of migraines [138], the general mechanism of migraine-induced photophobia is still unclear. The convergence of optical signals from retinal photoreceptors and nociceptive signals from the TGVS to the CNS is the most likely cause of photophobic experiences in migraineurs [139]. Investigating photophobia is generally a hard task due to the lack of well-established investigation tools and the naturally light-aversive behaviour of laboratory rodents as they are anxious about being seen by predators [140]. Nonetheless, a Light/Dark (LD) box test has previously been used in migraine research [141].
The LD box test does not exclusively capture photophobic behaviour, as it also can capture anxiety-like behaviour and measure locomotor activity [142]. A parameter like “time spend in light zone” is the most indicative of photophobic behaviour. In most studies to date, no difference was observed in similar LD box tests, in rodents where CFA has been administered [143,144]. However, CFA administration in those studies were not used in the context of migraine research and were targeted at inducing inflammation in the plantar region. Studies using other inflammatory mediators, such as NTG, have reported more light-aversive behaviours in NTG treated rodents compared to untreated rodents [145].
Other known behavioural migraine symptoms include anxiety and aversion of physical activity because of experienced pain [146]. The OFT has been reported to capture exploratory and anxiety-like behaviour, as well been able to measure locomotor activity [147]. One of the strongest arguments for including this outcome is the concept that physical activity is negatively influenced by headaches, which could be detected in the lack of exploratory behaviour. This is exemplified by part of the diagnostic criterion for migraines, where the headache is supposedly worsened by physical activity, or that the patient will avoid activity [148].
Although like the LD box test in this regard, the OFT differs by having a larger area for exploration and the lack of a high-intensity light. Anxious or depressed rats are expected to spend less time in the centre zone, while worse locomotor and exploratory performances are more associated with experiencing pain. Care should be taken with sequential testing as the animal would lose the novelty and exploratory features of OFTs, admittedly causing inconsistency in performances and/or defeating the purpose of the test. Either way, different strategies should be approached before initial tests to make sure to produce the most consistent results possible.
Previous migraine studies reported both periorbital and plantar allodynia when rodents were administered with migraine factors to the dura, such as IS [78,149]. The main question is whether the Von Frey threshold is capturing headache and not a cutaneous hypersensitivity, as both are seen in migraine patients. One important issue is that allodynia is not headache, and therefore therapeutic approaches might lead to treatments for allodynia and not for headache.

7. Concluding on CGRP and Inflammation—A Perspective

It is no surprise that inflammation is believed to play a potential part in migraine pathophysiology when CGRP is both a key player of migraine pathophysiology and NI. However, it is still unclear whether CGRP is an inducer of inflammation or a by-product of it. Subsequently, the question arises whether inflammation follows the activated areas in migraines (for example is caused by CGRP) or if inflammation causes the activation (and CGRP release). Nevertheless, one would postulate that anatomical separation between the hemispheres would occur, to give rise to the one-sidedness of migraine pain. CGRP even exerts anti-inflammatory properties in some tissue, questioning its role as a pro-inflammatory mediator and its safety as a systemic drug target, as blocking the anti-inflammatory properties could result in inflammation-related consequences elsewhere [150].
Neuroinflammation is typically viewed as a protective response to tissue damage. The main question in migraine is therefore also what type of tissue damage could lead to this initial tissue damage. One possibility could be a neuronal energy deficit. Indeed, this has mainly been put forward by Borkum [151], in light of data on a suggested energy deficit in migraines, particularly by a lower phosphocreatine-to-creatine ratio and an increased concentration of ADP [152,153]. In addition, migraines are associated with mitochondrial disease [154] and with subjects undergoing a fasting state [155]. One potential explanation could be a link to increased membrane potential as the cells struggle to maintain the resting voltage, which could potentially lead to CSD [156] or increased sensitivity in peripheral nerve fibers [27]. Following a potential energy deficit, an increase in inflammatory markers would be expected to protect the brain [157]. This neuroinflammation would control the severity and possible progression, but could also have detrimental effects [158]. The following neuroinflammation has in other systems been showed both to induce and aggravate neurodegeneration, but at the same time aids in the recovery of neurons [159]. This balance is clearly important when addressing inflammation as a potential clinical target.
Nonetheless, using CGRP-targeted drugs on inflammation models could inform us both of the role of CGRP in inflammation and whether inflammation plays an important role (if at all) in migraine pathophysiology. From what we could gather, the effects of anti-CGRP on IS and CFA have yet been investigated. However, triptans appear to mitigate the effects of both IS [65,101,105] and CFA [160,161], but whether it is due to their CGRP targeting or separate anti-inflammatory effects is not clear. Similarly, α-CGRP(8-37), a CGRP receptor antagonist, was able to reduce allodynia from intraplantar CFA [162]. Mixed data exist on the effect of gepants on CFA. The gepant BIBN4096BS (olcegepant) applied topically, intravenously but not spinally [163], was able to alleviate inflammatory pain from subcutaneously plantar administered CFA. However, intrathecal administration of BIBN4096BS did not reduce the increased nNOS that intraplantar CFA caused [164].
Interestingly, when applying CGRP unto the dura mater, it only causes hypersensitivity in female rodents, at least at doses up to 3.8 μg [106], which could be caused by differences in oestrogen signalling [165]. Therefore, if one was to test the efficacy of CGRP-targeting drugs in the inflammation models, one should be wary to also include female rodents if no effects of the drugs is found in the males. An increase in CGRP release at baseline values or stimulated dura mater compared to fresh rats could indicate a sensitized dura mater. Additionally, a significant reduction in CGRP release could indicate a desensitized dura mater from repeated stimulation. To our knowledge, none of the inflammatory models have tested this outcome. An obesity model showcased an increase of basal CGRP in the dura in rats fed a high fat/high sucrose diet. In a previous study, an increase in CGRP expression was observed in the TG couple hours post dural in vivo CFA administration [115]. It could therefore be interesting to test whether basal levels of CGRP in an inflammation model would rise.
There are some studies showing that corticosteroids are useful in managing resistant, severe, recurrent or prolonged migraine attacks (such as status migraenosus) in the emergency department [166]. The questions in our opinion remain as to whether these migraine cases such as status migraenosus should be considered an acute migraine or as a short chronic migraine (at least when it comes to the biology/pathology). Likewise, one could argue that if NSAIDs are effective treatment in some migraine patients, inflammation must therefore play an important pathological role. However it should be noted that these drugs have modes of actions that goes beyond their anti-inflammatory properties, such as their analgesic and antipyretic effects [167], which should be addressed in future studies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cells11152444/s1, Table S1: Summary of pathological changes in the TG, Table S2: Summary of pathological changes in the Brain/Spinal cord, Table S3: Summary of pathological changes in the TNC.

Author Contributions

Conceptualization, P.V.R., L.E. and K.A.H.; methodology, P.V.R., L.E. and K.A.H.; validation, P.V.R., L.E. and K.A.H.; investigation, P.V.R., L.E. and K.A.H.; resources, L.E. and K.A.H.; writing—original draft preparation, P.V.R., L.E. and K.A.H.; writing—review and editing, P.V.R., L.E. and K.A.H.; visualization, P.V.R. and K.A.H.; supervision, L.E. and K.A.H.; project administration, K.A.H.; funding acquisition, L.E. and K.A.H. All authors have read and agreed to the published version of the manuscript.

Funding

P.V.R. and K.A.H. was supported by a Lundbeck foundation Fellowship (R345-2020-1977). The funders had no role in the design and content of the review.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Artwork was based on Servier Medical Art by Servier is licensed under a Creative Commons Attribution 3.0 Unported License (smart.servier.com, accessed on 16 June 2022) parts were created with BioRender.com (accessed on 16 June 2022).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bigal, M.E.; Lipton, R.B. The epidemiology, burden, and comorbidities of migraine. Neurol. Clin. 2009, 27, 321–334. [Google Scholar] [CrossRef] [PubMed]
  2. Ashina, M.; Katsarava, Z.; Do, T.P.; Buse, D.C.; Pozo-Rosich, P.; Özge, A.; Krymchantowski, A.V.; Lebedeva, E.R.; Ravishankar, K.; Yu, S.; et al. Migraine: Epidemiology and systems of care. Lancet 2021, 397, 1485–1495. [Google Scholar] [CrossRef]
  3. Bron, C.; Sutherland, H.G.; Griffiths, L.R. Exploring the Hereditary Nature of Migraine. Neuropsychiatr. Dis. Treat. 2021, 17, 1183–1194. [Google Scholar] [CrossRef] [PubMed]
  4. Evans, R.W. Wolff’s "Headache and Other Head Pain": Happy 70th Anniversary! Headache 2019, 59, 443. [Google Scholar] [CrossRef]
  5. Haanes, K.A.; Edvinsson, L. Pathophysiological Mechanisms in Migraine and the Identification of New Therapeutic Targets. CNS Drugs 2019, 33, 525–537. [Google Scholar] [CrossRef]
  6. Matsuda, M.; Huh, Y.; Ji, R.R. Roles of inflammation, neurogenic inflammation, and neuroinflammation in pain. J. Anesth. 2019, 33, 131–139. [Google Scholar] [CrossRef]
  7. Riesco, N.; Cernuda-Morollón, E.; Pascual, J. Neuropeptides as a Marker for Chronic Headache. Curr. Pain Headache Rep. 2017, 21, 18. [Google Scholar] [CrossRef]
  8. Torres-Ferrus, M.; Ursitti, F.; Alpuente, A.; Brunello, F.; Chiappino, D.; Vries, T.D.; di Marco, S.; Ferlisi, S.; Guerritore, L.; Gonzalez-Garcia, N.; et al. School of Advanced Studies of European Headache, From transformation to chronification of migraine: Pathophysiological and clinical aspects. J. Headache Pain 2020, 21, 42. [Google Scholar] [CrossRef]
  9. Dodick, D.W. A Phase-by-Phase Review of Migraine Pathophysiology. Headache 2018, 58, 4–16. [Google Scholar] [CrossRef] [Green Version]
  10. Bernstein, C.; Burstein, R. Sensitization of the trigeminovascular pathway: Perspective and implications to migraine pathophysiology. J. Clin. Neurol. 2012, 8, 89–99. [Google Scholar] [CrossRef] [Green Version]
  11. Ashina, M.; Hansen, J.M.; O, A.D.B.; Olesen, J. Human models of migraine—short-term pain for long-term gain. Nat. Rev. Neurol. 2017, 13, 713–724. [Google Scholar] [CrossRef]
  12. Al-Karagholi, M.A.; Ghanizada, H.; Nielsen, C.A.W.; Hougaard, A.; Ashina, M. Opening of ATP sensitive potassium channels causes migraine attacks with aura. Brain 2021, 144, 2322–2332. [Google Scholar] [CrossRef] [PubMed]
  13. Schulte, L.H.; May, A. The migraine generator revisited: Continuous scanning of the migraine cycle over 30 days and three spontaneous attacks. Brain 2016, 139, 1987–1993. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kikkeri, N.S.; Nagalli, S. Migraine with Aura. Treasure Island (FL); StatPearls Publishing LLC: Treasure Island, FL, USA, 2022.
  15. Borgdorff, P. Arguments against the role of cortical spreading depression in migraine. Neurol. Res. 2018, 40, 173–181. [Google Scholar] [CrossRef] [Green Version]
  16. Bolay, H.; Reuter, U.; Dunn, A.K.; Huang, Z.; Boas, D.A.; Moskowitz, M.A. Intrinsic brain activity triggers trigeminal meningeal afferents in a migraine model. Nat. Med. 2002, 8, 136–142. [Google Scholar] [CrossRef] [PubMed]
  17. Zhang, X.; Levy, D.; Kainz, V.; Noseda, R.; Jakubowski, M.; Burstein, R. Activation of central trigeminovascular neurons by cortical spreading depression. Ann. Neurol. 2011, 69, 855–865. [Google Scholar] [CrossRef] [Green Version]
  18. 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]
  19. Edvinsson, L.; Haanes, K.A.; Warfvinge, K.; Krause, D.N. CGRP as the target of new migraine therapies—successful translation from bench to clinic. Nat. Rev. Neurol. 2018, 14, 338–350. [Google Scholar] [CrossRef]
  20. Ramos, M.L.; Pascual, J. AMG 334 CGRP antibody for migraine: Time to celebrate? Lancet Neurol. 2016, 15, 347–349. [Google Scholar] [CrossRef]
  21. Negro, A.; Martelletti, P. Gepants for the treatment of migraine. Expert. Opin. Investig. Drugs 2019, 28, 555–567. [Google Scholar] [CrossRef]
  22. Russo, A.F. Calcitonin gene-related peptide (CGRP): A new target for migraine. Annu. Rev. Pharmacol. Toxicol. 2015, 55, 533–552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Haanes, K.A.; Edvinsson, L. Hyperpolarization through ATP-sensitive potassium channels; relevance to migraine pathology. Brain 2020, 143, e13. [Google Scholar] [CrossRef] [PubMed]
  24. Mulderry, P.K.; Ghatei, M.A.; Bishop, A.E.; Allen, Y.S.; Polak, J.M.; Bloom, S.R. Distribution and chromatographic characterisation of CGRP-like immunoreactivity in the brain and gut of the rat. Regul. Pept. 1985, 12, 133–143. [Google Scholar] [CrossRef]
  25. Warfvinge, K.; Edvinsson, L. Distribution of CGRP and CGRP receptor components in the rat brain. Cephalalgia 2019, 39, 342–353. [Google Scholar] [CrossRef] [PubMed]
  26. Edvinsson, J.C.; Reducha, P.V.; Sheykhzade, M.; Warfvinge, K.; Haanes, K.A.; Edvinsson, L. Neurokinins and their receptors in the rat trigeminal system: Differential localization and release with implications for migraine pain. Mol. Pain 2021, 17, 17448069211059400. [Google Scholar] [CrossRef] [PubMed]
  27. Edvinsson, J.C.A.; Warfvinge, K.; Krause, D.N.; Blixt, F.W.; Sheykhzade, M.; Edvinsson, L.; Haanes, K.A. C-fibers may modulate adjacent Adelta-fibers through axon-axon CGRP signaling at nodes of Ranvier in the trigeminal system. J. Headache Pain 2019, 20, 105. [Google Scholar] [CrossRef] [Green Version]
  28. Walker, H.K. Cranial Nerve V, Clinical Methods: The Trigeminal Nerve, the History, Physical, and Laboratory Examinations; Walker, H.K., Hall, W.D., Hurst, J.W., Eds.; Butterworth Publishers: Boston, MA, USA, 1990. [Google Scholar]
  29. Cottrell, G.S. CGRP Receptor Signalling Pathways. Handb. Exp. Pharmacol. 2019, 255, 37–64. [Google Scholar]
  30. Moskowitz, M.A. The neurobiology of vascular head pain. Ann. Neurol. 1984, 16, 157–168. [Google Scholar] [CrossRef]
  31. Markowitz, S.; Saito, K.; Moskowitz, M.A. Neurogenically mediated plasma extravasation in dura mater: Effect of ergot alkaloids. A possible mechanism of action in vascular headache. Cephalalgia 1988, 8, 83–91. [Google Scholar] [CrossRef]
  32. Shepherd, S.L.; Williamson, D.J.; Beer, M.S.; Hill, R.G.; Hargreaves, R.J. Differential effects of 5-HT1B/1D receptor agonists on neurogenic dural plasma extravasation and vasodilation in anaesthetized rats. Neuropharmacology 1997, 36, 525–533. [Google Scholar] [CrossRef]
  33. Buzzi, M.G.; Moskowitz, M.A. The antimigraine drug, sumatriptan (GR43175), selectively blocks neurogenic plasma extravasation from blood vessels in dura mater. Br. J. Pharmacol. 1990, 99, 202–206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Moskowitz, M.A. Neurogenic versus vascular mechanisms of sumatriptan and ergot alkaloids in migraine. Trends Pharmacol. Sci. 1992, 13, 307–311. [Google Scholar] [CrossRef]
  35. Theoharides, T.C.; Spanos, C.; Pang, X.; Alferes, L.; Ligris, K.; Letourneau, R.; Rozniecki, J.J.; Webster, E.; Chrousos, G.P. Stress-induced intracranial mast cell degranulation: A corticotropin-releasing hormone-mediated effect. Endocrinology 1995, 136, 5745–5750. [Google Scholar] [CrossRef] [PubMed]
  36. Theoharides, T.C.; Donelan, J.; Kandere-Grzybowska, K.; Konstantinidou, A. The role of mast cells in migraine pathophysiology. Brain Res. Rev. 2005, 49, 65–76. [Google Scholar] [CrossRef]
  37. Dalkara, T.; Zervas, N.T.; Moskowitz, M.A. From spreading depression to the trigeminovascular system. Neurol. Sci. 2006, 27, S86–S90. [Google Scholar] [CrossRef]
  38. Levy, D. Endogenous mechanisms underlying the activation and sensitization of meningeal nociceptors: The role of immuno-vascular interactions and cortical spreading depression. Curr. Pain Headache Rep. 2012, 16, 270–277. [Google Scholar] [CrossRef]
  39. Johnson KW, B.H. Neurogenic Inflammatory Mechanisms; Lipincott Williams & Wilkins: Philadelphia, PA, USA, 2006. [Google Scholar]
  40. Lassen, L.H.; Ashina, M.; Christiansen, I.; Ulrich, V.; Olesen, J. Nitric oxide synthase inhibition in migraine. Lancet 1997, 349, 401–402. [Google Scholar] [CrossRef]
  41. Covelli, V.; Munno, I.; Pellegrino, N.M.; Altamura, M.; Decandia, P.; Marcuccio, C.; di Venere, A.; Jirillo, E. Are TNF-alpha and IL-1 beta relevant in the pathogenesis of migraine without aura? Acta Neurol. Napoli 1991, 13, 205–211. [Google Scholar]
  42. Perini, F.; D’Andrea, G.; Galloni, E.; Pignatelli, F.; Billo, G.; Alba, S.; Bussone, G.; Toso, V. Plasma cytokine levels in migraineurs and controls. Headache 2005, 45, 926–931. [Google Scholar] [CrossRef]
  43. Franceschini, A.; Vilotti, S.; Ferrari, M.D.; van den Maagdenberg, A.M.; Nistri, A.; Fabbretti, E. TNFalpha levels and macrophages expression reflect an inflammatory potential of trigeminal ganglia in a mouse model of familial hemiplegic migraine. PLoS ONE 2013, 8, e52394. [Google Scholar] [CrossRef]
  44. Tanure, M.T.; Gomez, R.S.; Hurtado, R.C.; Teixeira, A.L.; Domingues, R.B. Increased serum levels of brain-derived neurotropic factor during migraine attacks: A pilot study. J. Headache Pain 2010, 11, 427–430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Yucel, M.; Kotan, D.; Ciftci, G.G.; Ciftci, I.H.; Cikriklar, H.I. Serum levels of endocan, claudin-5 and cytokines in migraine. Eur. Rev. Med. Pharmacol. Sci. 2016, 20, 930–936. [Google Scholar] [PubMed]
  46. Sarchielli, P.; Alberti, A.; Baldi, A.; Coppola, F.; Rossi, C.; Pierguidi, L.; Floridi, A.; Calabresi, P. Proinflammatory cytokines, adhesion molecules, and lymphocyte integrin expression in the internal jugular blood of migraine patients without aura assessed ictally. Headache 2006, 46, 200–207. [Google Scholar] [CrossRef] [PubMed]
  47. Edvinsson, L.; Haanes, K.A.; Warfvinge, K. Does inflammation have a role in migraine? Nat. Rev. Neurol. 2019, 15, 483–490. [Google Scholar] [CrossRef]
  48. de Logu, F.; Landini, L.; Janal, M.N.; Puma, S.L.; de Cesaris, F.; Geppetti, P.; Nassini, R. Migraine-provoking substances evoke periorbital allodynia in mice. J. Headache Pain 2019, 20, 18. [Google Scholar] [CrossRef] [Green Version]
  49. Nakamura-Craig, M.; Gill, B.K. Effect of neurokinin A, substance P and calcitonin gene related peptide in peripheral hyperalgesia in the rat paw. Neurosci. Lett. 1991, 124, 49–51. [Google Scholar] [CrossRef]
  50. Birrell, G.J.; McQueen, D.S.; Iggo, A.; Coleman, R.A.; Grubb, B.D. PGI2-induced activation and sensitization of articular mechanonociceptors. Neurosci. Lett. 1991, 124, 5–8. [Google Scholar] [CrossRef]
  51. Wang, H.; Ehnert, C.; Brenner, G.J.; Woolf, C.J. Bradykinin and peripheral sensitization. Biol. Chem. 2006, 387, 11–14. [Google Scholar] [CrossRef]
  52. Schaible, H.G.; Schmidt, R.F. Excitation and sensitization of fine articular afferents from cat’s knee joint by prostaglandin E2. J. Physiol. 1988, 403, 91–104. [Google Scholar] [CrossRef] [Green Version]
  53. Uddman, R.; Edvinsson, L.; Jansen, I.; Stiernholm, P.; Jensen, K.; Olesen, J.; Sundler, F. Peptide-containing nerve fibres in human extracranial tissue: A morphological basis for neuropeptide involvement in extracranial pain? Pain 1986, 27, 391–399. [Google Scholar] [CrossRef]
  54. Harriott, A.M.; Gold, M.S. Electrophysiological properties of dural afferents in the absence and presence of inflammatory mediators. J. Neurophysiol. 2009, 101, 3126–3134. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Harriott, A.M.; Scheff, N.N.; Gold, M.S. The complex actions of sumatriptan on rat dural afferents. Cephalalgia 2012, 32, 738–749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Scheff, N.N.; Gold, M.S. Sex differences in the inflammatory mediator-induced sensitization of dural afferents. J. Neurophysiol. 2011, 106, 1662–1668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Oshinsky, M.L.; Gomonchareonsiri, S. Episodic dural stimulation in awake rats: A model for recurrent headache. Headache 2007, 47, 1026–1036. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Edelmayer, R.M.; Vanderah, T.W.; Majuta, L.; Zhang, E.T.; Fioravanti, B.; de Felice, M.; Chichorro, J.G.; Ossipov, M.H.; King, T.; Lai, J.; et al. Medullary pain facilitating neurons mediate allodynia in headache-related pain. Ann. Neurol. 2009, 65, 184–193. [Google Scholar] [CrossRef] [Green Version]
  59. Zhu, P.; Dong, X.; Xu, H.; Wan, Q.; Guo, Q.; Wang, J.; Xiao, H.; Yan, L. Microglial P2Y14 receptor contributes to central sensitization following repeated inflammatory dural stimulation. Brain Res. Bull. 2021, 177, 119–128. [Google Scholar] [CrossRef]
  60. Hu, G.; Zhang, M.; Su, M.; Zhang, Q.; Wu, H.; Wang, X.; Dong, Z.; Yu, S. Wider range of allodynia in a rat model of repeated dural nociception compared with infraorbital nerve chronic constriction injury. Neurosci. Lett. 2018, 666, 120–126. [Google Scholar] [CrossRef]
  61. Melo-Carrillo, A.; Lopez-Avila, A. A chronic animal model of migraine, induced by repeated meningeal nociception, characterized by a behavioral and pharmacological approach. Cephalalgia 2013, 33, 1096–1105. [Google Scholar] [CrossRef]
  62. Liu, D.; Chang, L.; Wang, J.; Zhu, Q.; Jiang, N.; Azhar, M.; Ahmed, A.; Zeng, G. Effect of Xiongmatang Extract on Behavioral and TRPV1-CGRP/CGRP-R Pathway in Rats with Migraine. Front. Pharmacol. 2022, 13, 835187. [Google Scholar] [CrossRef]
  63. Fan, S.; Xiao, Z.; Zhu, F.; He, X.; Lu, Z. A new comorbidity model and the common pathological mechanisms of migraine and epilepsy. Am. J. Transl. Res. 2017, 9, 2286–2295. [Google Scholar]
  64. Liu, L.; Xu, X.B.; Qu, Z.Y.; Zhao, L.P.; Zhang, C.S.; Li, Z.J.; Lyu, T.L.; Wang, X.F.; Jing, X.H.; Li, B. Determining 5HT(7)R’s Involvement in Modifying the Antihyperalgesic Effects of Electroacupuncture on Rats With Recurrent Migraine. Front. Neurosci. 2021, 15, 668616. [Google Scholar] [CrossRef] [PubMed]
  65. Nan, N.; Gong, M.X.; Wang, Q.; Li, M.J.; Xu, R.; Ma, Z.; Wang, S.H.; Zhao, H.; Xu, Y.S. Wuzhuyu Decoction relieves hyperalgesia by regulating central and peripheral 5-HT in chronic migraine model rats. Phytomedicine 2022, 96, 153905. [Google Scholar] [CrossRef] [PubMed]
  66. Zhou, H.; Wang, X.; Wang, S.; Liu, C.; Fu, Q.; Qin, G.; Zhou, J.; Chen, L. Inhibition of Nerve Growth Factor Signaling Alleviates Repeated Dural Stimulation-induced Hyperalgesia in Rats. Neuroscience 2019, 398, 252–262. [Google Scholar] [CrossRef]
  67. Liu, C.; Zhang, Y.; Liu, Q.; Jiang, L.; Li, M.; Wang, S.; Long, T.; He, W.; Kong, X.; Qin, G.; et al. P2X4-receptor participates in EAAT3 regulation via BDNF-TrkB signaling in a model of trigeminal allodynia. Mol. Pain 2018, 14, 1744806918795930. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Jia, Z.; Tang, W.; Zhao, D.; Yu, S. Disrupted functional connectivity between the periaqueductal gray and other brain regions in a rat model of recurrent headache. Sci. Rep. 2017, 7, 3960. [Google Scholar] [CrossRef] [Green Version]
  69. Gong, Q.; Lin, Y.; Lu, Z.; Xiao, Z. Microglia-Astrocyte Cross Talk through IL-18/IL-18R Signaling Modulates Migraine-like Behavior in Experimental Models of Migraine. Neuroscience 2020, 451, 207–215. [Google Scholar] [CrossRef] [PubMed]
  70. Wieseler, J.; Ellis, A.; McFadden, A.; Stone, K.; Brown, K.; Cady, S.; Bastos, L.F.; Sprunger, D.; Rezvani, N.; Johnson, K.; et al. Supradural inflammatory soup in awake and freely moving rats induces facial allodynia that is blocked by putative immune modulators. Brain Res. 2017, 1664, 87–94. [Google Scholar] [CrossRef] [PubMed]
  71. Su, M.; Ran, Y.; He, Z.; Zhang, M.; Hu, G.; Tang, W.; Zhao, D.; Yu, S. Inhibition of toll-like receptor 4 alleviates hyperalgesia induced by acute dural inflammation in experimental migraine. Mol. Pain 2018, 14, 1744806918754612. [Google Scholar] [CrossRef] [Green Version]
  72. Zhou, Y.; Zhang, L.; Hao, Y.; Yang, L.; Fan, S.; Xiao, Z. FKN/CX3CR1 axis facilitates migraine-Like behaviour by activating thalamic-cortical network microglia in status epilepticus model rats. J. Headache Pain 2022, 23, 42. [Google Scholar] [CrossRef]
  73. Vellani, V.; Moschetti, G.; Franchi, S.; Giacomoni, C.; Sacerdote, P.; Amodeo, G. Effects of NSAIDs on the Release of Calcitonin Gene-Related Peptide and Prostaglandin E2 from Rat Trigeminal Ganglia. Mediat. Inflamm. 2017, 2017, 9547056. [Google Scholar] [CrossRef] [Green Version]
  74. Fried, N.T.; Maxwell, C.R.; Elliott, M.B.; Oshinsky, M.L. Region-specific disruption of the blood-brain barrier following repeated inflammatory dural stimulation in a rat model of chronic trigeminal allodynia. Cephalalgia 2018, 38, 674–689. [Google Scholar] [CrossRef] [PubMed]
  75. Liu, Q.; Liu, C.; Jiang, L.; Li, M.; Long, T.; He, W.; Qin, G.; Chen, L.; Zhou, J. α7 Nicotinic acetylcholine receptor-mediated anti-inflammatory effect in a chronic migraine rat model via the attenuation of glial cell activation. J. Pain Res. 2018, 11, 1129–1140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Wang, X.Y.; Zhou, H.R.; Wang, S.; Liu, C.Y.; Qin, G.C.; Fu, Q.Q.; Zhou, J.Y.; Chen, L.X. NR2B-Tyr phosphorylation regulates synaptic plasticity in central sensitization in a chronic migraine rat model. J. Headache Pain 2018, 19, 102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Wu, B.; Wang, S.; Qin, G.; Xie, J.; Tan, G.; Zhou, J.; Chen, L. Protein Kinase C γ Contributes to Central Sensitization in a Rat Model of Chronic Migraine. J. Mol. Neurosci. 2017, 63, 131–141. [Google Scholar] [CrossRef]
  78. Stucky, N.L.; Gregory, E.; Winter, M.K.; He, Y.Y.; Hamilton, E.S.; McCarson, K.E.; Berman, N.E. Sex differences in behavior and expression of CGRP-related genes in a rodent model of chronic migraine. Headache 2011, 51, 674–692. [Google Scholar] [CrossRef]
  79. Liang, X.; Wang, S.; Qin, G.; Xie, J.; Tan, G.; Zhou, J.; McBride, D.W.; Chen, L. Tyrosine Phosphorylation of NR2B Contributes to Chronic Migraines via Increased Expression of CGRP in Rats. Biomed. Res. Int. 2017, 2017, 7203458. [Google Scholar] [CrossRef]
  80. Liang, J.; Zhou, X.; Wang, J.; Fei, Z.Y.; Qin, G.C.; Zhang, D.K.; Zhou, J.Y.; Chen, L.X. Upregulation of silent information regulator 1 alleviates mitochondrial dysfunction in the trigeminal nucleus caudalis in a rat model of chronic migraine. Neuroreport 2021, 32, 144–156. [Google Scholar] [CrossRef]
  81. Wang, J.; Fei, Z.; Liang, J.; Zhou, X.; Qin, G.; Zhang, D.; Zhou, J.; Chen, L. EphrinB/EphB Signaling Contributes to the Synaptic Plasticity of Chronic Migraine Through NR2B Phosphorylation. Neuroscience 2020, 428, 178–191. [Google Scholar] [CrossRef]
  82. Wang, S.; Wu, B.X.; Liu, C.Y.; Qin, G.C.; Yan, W.H.; Zhou, J.Y.; Chen, L.X. Expression of ASIC3 in the Trigeminal Nucleus Caudalis Plays a Role in a Rat Model of Recurrent Migraine. J. Mol. Neurosci. 2018, 66, 44–52. [Google Scholar] [CrossRef]
  83. Chen, N.; Su, W.; Cui, S.H.; Guo, J.; Duan, J.C.; Li, H.X.; He, L. A novel large animal model of recurrent migraine established by repeated administration of inflammatory soup into the dura mater of the rhesus monkey. Neural. Regen. Res. 2019, 14, 100–106. [Google Scholar]
  84. Tuka, B.; Helyes, Z.; Markovics, A.; Bagoly, T.; Szolcsányi, J.; Szabó, N.; Tóth, E.; Kincses, Z.T.; Vécsei, L.; Tajti, J. Alterations in PACAP-38-like immunoreactivity in the plasma during ictal and interictal periods of migraine patients. Cephalalgia 2013, 33, 1085–1095. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Han, X.; Dong, Z.; Hou, L.; Wan, D.; Chen, M.; Tang, W.; Yu, S. Interictal plasma pituitary adenylate cyclase-activating polypeptide levels are decreased in migraineurs but remain unchanged in patients with tension-type headache. Clin. Chim. Acta 2015, 450, 151–154. [Google Scholar] [CrossRef] [PubMed]
  86. Han, X.; Ran, Y.; Su, M.; Liu, Y.; Tang, W.; Dong, Z.; Yu, S. Chronic changes in pituitary adenylate cyclase-activating polypeptide and related receptors in response to repeated chemical dural stimulation in rats. Mol. Pain 2017, 13, 1744806917720361. [Google Scholar] [CrossRef] [PubMed]
  87. Burgos-Vega, C.C.; Ahn, D.D.; Bischoff, C.; Wang, W.; Horne, D.; Wang, J.; Gavva, N.; Dussor, G. Meningeal transient receptor potential channel M8 activation causes cutaneous facial and hindpaw allodynia in a preclinical rodent model of headache. Cephalalgia 2016, 36, 185–193. [Google Scholar] [CrossRef] [PubMed]
  88. 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]
  89. Jia, Z.; Yu, S.; Tang, W.; Zhao, D. Altered functional connectivity of the insula in a rat model of recurrent headache. Mol. Pain 2020, 16, 1744806920922115. [Google Scholar] [CrossRef] [Green Version]
  90. Jia, Z.; Chen, X.; Tang, W.; Zhao, D.; Yu, S. Atypical functional connectivity between the anterior cingulate cortex and other brain regions in a rat model of recurrent headache. Mol. Pain 2019, 15, 1744806919842483. [Google Scholar] [CrossRef]
  91. Becerra, L.; Bishop, J.; Barmettler, G.; Kainz, V.; Burstein, R.; Borsook, D. Brain network alterations in the inflammatory soup animal model of migraine. Brain Res. 2017, 1660, 36–46. [Google Scholar] [CrossRef]
  92. Bishop, J.; Becerra, L.; Barmettler, G.; Chang, P.C.; Kainz, V.; Burstein, R.; Borsook, D. Modulation of brain networks by sumatriptan-naproxen in the inflammatory soup migraine model. Pain 2019, 160, 2161–2171. [Google Scholar] [CrossRef]
  93. Zhang, M.; Liu, Y.; Hu, G.; Kang, L.; Ran, Y.; Su, M.; Yu, S. Cognitive impairment in a classical rat model of chronic migraine may be due to alterations in hippocampal synaptic plasticity and N-methyl-D-aspartate receptor subunits. Mol. Pain 2020, 16, 1744806920959582. [Google Scholar] [CrossRef]
  94. Zeng, X.; Niu, Y.; Qin, G.; Zhang, D.; Zhou, J.; Chen, L. Deficiency in the function of inhibitory interneurons contributes to glutamate-associated central sensitization through GABABR2-SynCAM1 signaling in chronic migraine rats. FASEB J. 2020, 34, 14780–14798. [Google Scholar] [CrossRef] [PubMed]
  95. Niu, Y.; Zeng, X.; Qin, G.; Zhang, D.; Zhou, J.; Chen, L. Downregulation of metabotropic glutamate receptor 5 alleviates central sensitization by activating autophagy via inhibiting mTOR pathway in a rat model of chronic migraine. Neurosci. Lett. 2021, 743, 135552. [Google Scholar] [CrossRef] [PubMed]
  96. Boyer, N.; Dallel, R.; Artola, A.; Monconduit, L. General trigeminospinal central sensitization and impaired descending pain inhibitory controls contribute to migraine progression. Pain 2014, 155, 1196–1205. [Google Scholar] [CrossRef] [PubMed]
  97. Mi, X.; Ran, L.; Chen, L.; Qin, G. Recurrent Headache Increases Blood-Brain Barrier Permeability and VEGF Expression in Rats. Pain Physician. 2018, 21, E633–E642. [Google Scholar] [PubMed]
  98. Levy, D.; Zhang, X.C.; Jakubowski, M.; Burstein, R. Sensitization of meningeal nociceptors: Inhibition by naproxen. Eur. J. Neurosci. 2008, 27, 917–922. [Google Scholar] [CrossRef] [Green Version]
  99. Lambert, G.A.; Davis, J.B.; Appleby, J.M.; Chizh, B.A.; Hoskin, K.L.; Zagami, A.S. The effects of the TRPV1 receptor antagonist SB-705498 on trigeminovascular sensitisation and neurotransmission. Naunyn Schmiedeberg’s Arch. Pharmcol. 2009, 380, 311–325. [Google Scholar] [CrossRef]
  100. Qu, Z.; Liu, L.; Yang, Y.; Zhao, L.; Xu, X.; Li, Z.; Zhu, Y.; Jing, X.; Wang, X.; Zhang, C.S.; et al. Electro-acupuncture inhibits C-fiber-evoked WDR neuronal activity of the trigeminocervical complex: Neurophysiological hypothesis of a complementary therapy for acute migraine modeled rats. Brain Res. 2020, 1730, 146670. [Google Scholar] [CrossRef]
  101. Boyer, N.; Signoret-Genest, J.; Artola, A.; Dallel, R.; Monconduit, L. Propranolol treatment prevents chronic central sensitization induced by repeated dural stimulation. Pain 2017, 158, 2025–2034. [Google Scholar] [CrossRef]
  102. Jia, Z.; Tang, W.; Zhao, D.; Hu, G.; Li, R.; Yu, S. Volumetric abnormalities of the brain in a rat model of recurrent headache. Mol. Pain 2018, 14, 1744806918756466. [Google Scholar] [CrossRef]
  103. Togha, M.; Jahromi, S.R.; Ghorbani, Z.; Ghaemi, A.; Rafiee, P. Evaluation of Inflammatory State in Migraineurs: A Case-control Study. Iran. J. Allergy Asthma Immunol. 2020, 19, 83–90. [Google Scholar] [CrossRef]
  104. Burgos-Vega, C.C.; Quigley, L.D.; Avona, A.; Price, T.; Dussor, G. Dural stimulation in rats causes brain-derived neurotrophic factor-dependent priming to subthreshold stimuli including a migraine trigger. Pain 2016, 157, 2722–2730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Avona, A.; Price, T.J.; Dussor, G. Interleukin-6 induces spatially dependent whole-body hypersensitivity in rats: Implications for extracephalic hypersensitivity in migraine. J. Headache Pain 2021, 22, 70. [Google Scholar] [CrossRef] [PubMed]
  106. Avona, A.; Burgos-Vega, C.; Burton, M.D.; Akopian, A.N.; Price, T.J.; Dussor, G. Dural Calcitonin Gene-Related Peptide Produces Female-Specific Responses in Rodent Migraine Models. J. Neurosci. 2019, 39, 4323–4331. [Google Scholar] [CrossRef] [Green Version]
  107. Yan, J.; Melemedjian, O.K.; Price, T.J.; Dussor, G. Sensitization of dural afferents underlies migraine-related behavior following meningeal application of interleukin-6 (IL-6). Mol. Pain 2012, 8, 6. [Google Scholar] [CrossRef] [Green Version]
  108. Burgos-Vega, C.C.; Quigley, L.D.; Santos, G.T.D.; Yan, F.; Asiedu, M.; Jacobs, B.; Motina, M.; Safdar, N.; Yousuf, H.; Avona, A.; et al. Non-invasive dural stimulation in mice: A novel preclinical model of migraine. Cephalalgia 2019, 39, 123–134. [Google Scholar] [CrossRef]
  109. Qu, Z.; Liu, L.; Zhao, L.; Xu, X.; Li, Z.; Zhu, Y.; Zhang, C.; Jing, X.; Wang, X.; Li, B.; et al. Prophylactic Electroacupuncture on the Upper Cervical Segments Decreases Neuronal Discharges of the Trigeminocervical Complex in Migraine-Affected Rats: An in vivo Extracellular Electrophysiological Experiment. J. Pain Res. 2020, 13, 25–37. [Google Scholar] [CrossRef] [Green Version]
  110. Lackovic, J.; Price, T.J.; Dussor, G. De novo protein synthesis is necessary for priming in preclinical models of migraine. Cephalalgia 2021, 41, 237–246. [Google Scholar] [CrossRef] [PubMed]
  111. Kristiansen, K.A.; Edvinsson, L. Neurogenic inflammation: A study of rat trigeminal ganglion. J. Headache Pain 2010, 11, 485–495. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Kuris, A.; Xu, C.B.; Zhou, M.F.; Tajti, J.; Uddman, R.; Edvinsson, L. Enhanced expression of CGRP in rat trigeminal ganglion neurons during cell and organ culture. Brain Res. 2007, 1173, 6–13. [Google Scholar] [CrossRef] [PubMed]
  113. Csati, A.; Edvinsson, L.; Vecsei, L.; Toldi, J.; Fulop, F.; Tajti, J.; Warfvinge, K. Kynurenic acid modulates experimentally induced inflammation in the trigeminal ganglion. J. Headache Pain 2015, 16, 99. [Google Scholar] [CrossRef] [Green Version]
  114. Burstein, R.; Jakubowski, M. Analgesic triptan action in an animal model of intracranial pain: A race against the development of central sensitization. Ann. Neurol. 2004, 55, 27–36. [Google Scholar] [CrossRef] [PubMed]
  115. Lukacs, M.; Haanes, K.A.; Majlath, Z.; Tajti, J.; Vecsei, L.; Warfvinge, K.; Edvinsson, L. Dural administration of inflammatory soup or Complete Freund’s Adjuvant induces activation and inflammatory response in the rat trigeminal ganglion. J. Headache Pain 2015, 16, 564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Lukacs, M.; Warfvinge, K.; Tajti, J.; Fulop, F.; Toldi, J.; Vecsei, L.; Edvinsson, L. Topical dura mater application of CFA induces enhanced expression of c-fos and glutamate in rat trigeminal nucleus caudalis: Attenuated by KYNA derivate (SZR72). J. Headache Pain 2017, 18, 39. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Tajti, J.; Kuris, A.; Vecsei, L.; Xu, C.B.; Edvinsson, L. Organ culture of the trigeminal ganglion induces enhanced expression of calcitonin gene-related peptide via activation of extracellular signal-regulated protein kinase 1/2. Cephalalgia 2011, 31, 95–105. [Google Scholar] [CrossRef] [PubMed]
  118. Takeda, M.; Tanimoto, T.; Kadoi, J.; Nasu, M.; Takahashi, M.; Kitagawa, J.; Matsumoto, S. Enhanced excitability of nociceptive trigeminal ganglion neurons by satellite glial cytokine following peripheral inflammation. Pain 2007, 129, 155–166. [Google Scholar] [CrossRef] [PubMed]
  119. Takeda, M.; Takahashi, M.; Matsumoto, S. Contribution of activated interleukin receptors in trigeminal ganglion neurons to hyperalgesia via satellite glial interleukin-1beta paracrine mechanism. Brain Behav. Immun. 2008, 22, 1016–1023. [Google Scholar] [CrossRef]
  120. Burstein, R.; Yamamura, H.; Malick, A.; Strassman, A.M. Chemical stimulation of the intracranial dura induces enhanced responses to facial stimulation in brain stem trigeminal neurons. J. Neurophysiol. 1998, 79, 964–982. [Google Scholar] [CrossRef] [Green Version]
  121. Strassman, A.M.; Raymond, S.A.; Burstein, R. Sensitization of meningeal sensory neurons and the origin of headaches. Nature 1996, 384, 560–564. [Google Scholar] [CrossRef]
  122. Lukacs, M.; Warfvinge, K.; Kruse, L.S.; Tajti, J.; Fulop, F.; Toldi, J.; Vecsei, L.; Edvinsson, L. KYNA analogue SZR72 modifies CFA-induced dural inflammation- regarding expression of pERK1/2 and IL-1beta in the rat trigeminal ganglion. J. Headache Pain 2016, 17, 64. [Google Scholar] [CrossRef] [Green Version]
  123. Romero-Reyes, M.; Pardi, V.; Akerman, S. A potent and selective calcitonin gene-related peptide (CGRP) receptor antagonist, MK-8825, inhibits responses to nociceptive trigeminal activation: Role of CGRP in orofacial pain. Exp. Neurol. 2015, 271, 95–103. [Google Scholar] [CrossRef]
  124. Janeway, C.A., Jr. Approaching the asymptote? Evolution and revolution in immunology. Cold Spring Harb. Symp. Quant. Biol. 1989, 54, 1–13. [Google Scholar] [CrossRef] [PubMed]
  125. Spekker, E.; Laborc, K.F.; Bohár, Z.; Nagy-Grócz, G.; Fejes-Szabó, A.; Szűcs, M.; Vécsei, L.; Párdutz, Á. Effect of dural inflammatory soup application on activation and sensitization markers in the caudal trigeminal nucleus of the rat and the modulatory effects of sumatriptan and kynurenic acid. J. Headache Pain 2021, 22, 17. [Google Scholar] [CrossRef] [PubMed]
  126. Pinheiro, A.R.; Paramos-de-Carvalho, D.; Certal, M.; Costa, C.; Magalhães-Cardoso, M.T.; Ferreirinha, F.; Costa, M.A.; Correia-de-Sá, P. Bradykinin-induced Ca2+ signaling in human subcutaneous fibroblasts involves ATP release via hemichannels leading to P2Y12 receptors activation. Cell Commun. Signal. 2013, 11, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Pinheiro, A.R.; Paramos-de-Carvalho, D.; Certal, M.; Costa, M.A.; Costa, C.; Magalhaes-Cardoso, M.T.; Ferreirinha, F.; Sevigny, J.; Correia-de-Sa, P. Histamine induces ATP release from human subcutaneous fibroblasts, via pannexin-1 hemichannels, leading to Ca2+ mobilization and cell proliferation. J. Biol. Chem. 2013, 288, 27571–27583. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Kowal, J.M.; Haanes, K.A.; Christensen, N.M.; Novak, I. Bile acid effects are mediated by ATP release and purinergic signalling in exocrine pancreatic cells. Cell Commun. Signal 2015, 13, 28. [Google Scholar] [CrossRef] [Green Version]
  129. Beckel, J.M.; Gomez, N.M.; Lu, W.; Campagno, K.E.; Nabet, B.; Albalawi, F.; Lim, J.C.; Boesze-Battaglia, K.; Mitchell, C.H. Stimulation of TLR3 triggers release of lysosomal ATP in astrocytes and epithelial cells that requires TRPML1 channels. Sci. Rep. 2018, 8, 5726. [Google Scholar] [CrossRef]
  130. Matsuka, Y.; Neubert, J.K.; Maidment, N.T.; Spigelman, I. Concurrent release of ATP and substance P within guinea pig trigeminal ganglia in vivo. Brain Res. 2001, 915, 248–255. [Google Scholar] [CrossRef]
  131. Koroleva, K.; Gafurov, O.; Guselnikova, V.; Nurkhametova, D.; Giniatullina, R.; Sitdikova, G.; Mattila, O.S.; Lindsberg, P.J.; Malm, T.M.; Giniatullin, R. Meningeal Mast Cells Contribute to ATP-Induced Nociceptive Firing in Trigeminal Nerve Terminals: Direct and Indirect Purinergic Mechanisms Triggering Migraine Pain. Front. Cell. Neurosci. 2019, 13, 195. [Google Scholar] [CrossRef]
  132. Haanes, K.A.; Edvinsson, L. Expression and characterization of purinergic receptors in rat middle meningeal artery-potential role in migraine. PLoS ONE 2014, 9, e108782. [Google Scholar] [CrossRef]
  133. Yegutkin, G.G.; Guerrero-Toro, C.; Kilinc, E.; Koroleva, K.; Ishchenko, Y.; Abushik, P.; Giniatullina, R.; Fayuk, D.; Giniatullin, R. Nucleotide homeostasis and purinergic nociceptive signaling in rat meninges in migraine-like conditions. Purinergic Signal 2016, 12, 561–574. [Google Scholar] [CrossRef] [Green Version]
  134. Haanes, K.A.; Labastida-Ramirez, A.; Blixt, F.W.; Rubio-Beltran, E.; Dirven, C.M.; Danser, A.H.; Edvinsson, L.; MaassenVanDenBrink, A. Exploration of purinergic receptors as potential anti-migraine targets using established pre-clinical migraine models. Cephalalgia 2019, 39, 1421–1434. [Google Scholar] [CrossRef] [PubMed]
  135. Dussor, G. Serotonin, 5HT1 agonists, and migraine: New data, but old questions still not answered. Curr. Opin. Support Palliat Care 2014, 8, 137–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Laborc, K.F.; Spekker, E.; Bohar, Z.; Szucs, M.; Nagy-Grocz, G.; Fejes-Szabo, A.; Vecsei, L.; Pardutz, A. Trigeminal activation patterns evoked by chemical stimulation of the dura mater in rats. J. Headache Pain 2020, 21, 101. [Google Scholar] [CrossRef] [PubMed]
  137. Harriott, A.M.; Strother, L.C.; Vila-Pueyo, M.; Holland, P.R. Animal models of migraine and experimental techniques used to examine trigeminal sensory processing. J. Headache Pain 2019, 20, 91. [Google Scholar] [CrossRef] [PubMed]
  138. Vingen, J.V.; Pareja, J.A.; Støren, O.; White, L.R.; Stovner, L.J. Phonophobia in migraine. Cephalalgia 1998, 18, 243–249. [Google Scholar] [CrossRef]
  139. Noseda, R.; Copenhagen, D.; Burstein, R. Current understanding of photophobia, visual networks and headaches. Cephalalgia 2019, 39, 1623–1634. [Google Scholar] [CrossRef]
  140. Farajdokht, F.; Babri, S.; Karimi, P.; Mohaddes, G. Ghrelin attenuates hyperalgesia and light aversion-induced by nitroglycerin in male rats. Neurosci. Lett. 2016, 630, 30–37. [Google Scholar] [CrossRef]
  141. Vuralli, D.; Wattiez, A.S.; Russo, A.F.; Bolay, H. Behavioral and cognitive animal models in headache research. J. Headache Pain 2019, 20, 11. [Google Scholar] [CrossRef]
  142. Bourin, M.; Hascoët, M. The mouse light/dark box test. Eur. J. Pharmacol. 2003, 463, 55–65. [Google Scholar] [CrossRef]
  143. Parent, A.J.; Beaudet, N.; Beaudry, H.; Bergeron, J.; Bérubé, P.; Drolet, G.; Sarret, P.; Gendron, L. Increased anxiety-like behaviors in rats experiencing chronic inflammatory pain. Behav. Brain Res. 2012, 229, 160–167. [Google Scholar] [CrossRef] [Green Version]
  144. Hofmann, L.; Karl, F.; Sommer, C.; Üçeyler, N. Affective and cognitive behavior in the alpha-galactosidase A deficient mouse model of Fabry disease. PLoS ONE 2017, 12, e0180601. [Google Scholar] [CrossRef] [Green Version]
  145. Sufka, K.J.; Staszko, S.M.; Johnson, A.P.; Davis, M.E.; Davis, R.E.; Smitherman, T.A. Clinically relevant behavioral endpoints in a recurrent nitroglycerin migraine model in rats. J. Headache Pain 2016, 17, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Lantéri-Minet, M.; Radat, F.; Chautard, M.H.; Lucas, C. Anxiety and depression associated with migraine: Influence on migraine subjects’ disability and quality of life, and acute migraine management. Pain 2005, 118, 319–326. [Google Scholar] [CrossRef] [PubMed]
  147. Kraeuter, A.K.; Guest, P.C.; Sarnyai, Z. The Open Field Test for Measuring Locomotor Activity and Anxiety-Like Behavior. Methods Mol. Biol. 2019, 1916, 99–103. [Google Scholar] [PubMed]
  148. Ferrari, M.D. Migraine. Lancet 1998, 351, 1043–1051. [Google Scholar] [CrossRef]
  149. Moye, L.S.; Pradhan, A.A.A. Animal Model of Chronic Migraine-Associated Pain. Curr. Protoc. Neurosci. 2017, 80, 9.60.1–9.60.9. [Google Scholar] [CrossRef]
  150. Ray, J.C.; Allen, P.; Bacsi, A.; Bosco, J.J.; Chen, L.; Eller, M.; Kua, H.; Lim, L.L.; Matharu, M.S.; Monif, M.; et al. Inflammatory complications of CGRP monoclonal antibodies: A case series. J. Headache Pain 2021, 22, 121. [Google Scholar] [CrossRef]
  151. Borkum, J.M. Brain Energy Deficit as a Source of Oxidative Stress in Migraine: A Molecular Basis for Migraine Susceptibility. Neurochem. Res. 2021, 46, 1913–1932. [Google Scholar] [CrossRef]
  152. 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]
  153. Younis, S.; Hougaard, A.; Vestergaard, M.B.; Larsson, H.B.W.; Ashina, M. Migraine and magnetic resonance spectroscopy: A systematic review. Curr. Opin. Neurol. 2017, 30, 246–262. [Google Scholar] [CrossRef]
  154. Tiehuis, L.H.; Koene, S.; Saris, C.G.J.; Janssen, M.C.H. Mitochondrial migraine; a prevalence, impact and treatment efficacy cohort study. Mitochondrion 2020, 53, 128–132. [Google Scholar] [CrossRef] [PubMed]
  155. Dalkara, T.; Kilic, K. How does fasting trigger migraine? A hypothesis. Curr. Pain Headache Rep. 2013, 17, 368. [Google Scholar] [CrossRef]
  156. Stuart, S.; Griffiths, L.R. A possible role for mitochondrial dysfunction in migraine. Mol. Genet. Genom. 2012, 287, 837–844. [Google Scholar] [CrossRef] [PubMed]
  157. More, S.V.; Kumar, H.; Kim, I.S.; Song, S.Y.; Choi, D.K. Cellular and molecular mediators of neuroinflammation in the pathogenesis of Parkinson’s disease. Mediat. Inflamm 2013, 2013, 952375. [Google Scholar] [CrossRef] [PubMed]
  158. Kielian, T. Multifaceted roles of neuroinflammation: The need to consider both sides of the coin. J. Neurochem. 2016, 136, 5–9. [Google Scholar] [CrossRef] [Green Version]
  159. Lucas, S.M.; Rothwell, N.J.; Gibson, R.M. The role of inflammation in CNS injury and disease. Br. J. Pharmacol. 2006, 147, S232–S240. [Google Scholar] [CrossRef] [Green Version]
  160. Winters, B.L.; Jeong, H.J.; Vaughan, C.W. Inflammation induces developmentally regulated sumatriptan inhibition of spinal synaptic transmission. Br. J. Pharmacol. 2020, 177, 3730–3743. [Google Scholar] [CrossRef]
  161. Cornelison, L.E.; Woodman, S.E.; Durham, P.L. Inhibition of Trigeminal Nociception by Non-invasive Vagus Nerve Stimulation: Investigating the Role of GABAergic and Serotonergic Pathways in a Model of Episodic Migraine. Front Neurol. 2020, 11, 146. [Google Scholar] [CrossRef]
  162. Yarwood, R.E.; Imlach, W.L.; Lieu, T.; Veldhuis, N.A.; Jensen, D.D.; Herenbrink, C.K.; Aurelio, L.; Cai, Z.; Christie, M.J.; Poole, D.P.; et al. Endosomal signaling of the receptor for calcitonin gene-related peptide mediates pain transmission. Proc. Natl. Acad. Sci. USA 2017, 114, 12309–12314. [Google Scholar] [CrossRef] [Green Version]
  163. Hirsch, S.; Corradini, L.; Just, S.; Arndt, K.; Doods, H. The CGRP receptor antagonist BIBN4096BS peripherally alleviates inflammatory pain in rats. Pain 2013, 154, 700–707. [Google Scholar] [CrossRef]
  164. Wang, D.; Ruan, L.; Hong, Y.; Chabot, J.G.; Quirion, R. Involvement of PKA-dependent upregulation of nNOS-CGRP in adrenomedullin-initiated mechanistic pathway underlying CFA-induced response in rats. Exp. Neurol. 2013, 239, 111–119. [Google Scholar] [CrossRef] [PubMed]
  165. Warfvinge, K.; Krause, D.N.; Maddahi, A.; Edvinsson, J.C.A.; Edvinsson, L.; Haanes, K.A. Estrogen receptors alpha, beta and GPER in the CNS and trigeminal system—molecular and functional aspects. J. Headache Pain 2020, 21, 131. [Google Scholar] [CrossRef] [PubMed]
  166. Woldeamanuel, Y.W.; Rapoport, A.M.; Cowan, R.P. The place of corticosteroids in migraine attack management: A 65-year systematic review with pooled analysis and critical appraisal. Cephalalgia 2015, 35, 996–1024. [Google Scholar] [CrossRef] [PubMed]
  167. Gunaydin, C.; Bilge, S.S. Effects of Nonsteroidal Anti-Inflammatory Drugs at the Molecular Level. Eurasian J. Med. 2018, 50, 116–121. [Google Scholar] [CrossRef]
Figure 1. Chronification of migraine. Small CGRP-containing neurons (green), and larger CGRP receptor-containing neurons (pink) are surrounded by satellite glial cells. We postulate that activation of the satellite glial cells (indicated by a change in color from orange to red) sensitizes the glia/neuron subunit, thereby decreasing the migraine threshold and leading to increased migraine frequency.
Figure 1. Chronification of migraine. Small CGRP-containing neurons (green), and larger CGRP receptor-containing neurons (pink) are surrounded by satellite glial cells. We postulate that activation of the satellite glial cells (indicated by a change in color from orange to red) sensitizes the glia/neuron subunit, thereby decreasing the migraine threshold and leading to increased migraine frequency.
Cells 11 02444 g001
Figure 2. Sensitization in animal models. The chronification of migraines can to some extent be mimicked by the application of inflammatory substances such as Inflammatory Soup (IS) or Complete Freund’s Adjuvant (CFA) to the dura and meninges of rodents. Experimental data suggest that this activates satellite glial cells (indicated by a change in colour from orange to red) sensitizes (black arrow) the trigeminal ganglion (TG), and the trigeminal nucleus caudalis (TNC).
Figure 2. Sensitization in animal models. The chronification of migraines can to some extent be mimicked by the application of inflammatory substances such as Inflammatory Soup (IS) or Complete Freund’s Adjuvant (CFA) to the dura and meninges of rodents. Experimental data suggest that this activates satellite glial cells (indicated by a change in colour from orange to red) sensitizes (black arrow) the trigeminal ganglion (TG), and the trigeminal nucleus caudalis (TNC).
Cells 11 02444 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Reducha, P.V.; Edvinsson, L.; Haanes, K.A. Could Experimental Inflammation Provide Better Understanding of Migraines? Cells 2022, 11, 2444. https://doi.org/10.3390/cells11152444

AMA Style

Reducha PV, Edvinsson L, Haanes KA. Could Experimental Inflammation Provide Better Understanding of Migraines? Cells. 2022; 11(15):2444. https://doi.org/10.3390/cells11152444

Chicago/Turabian Style

Reducha, Philip Victor, Lars Edvinsson, and Kristian Agmund Haanes. 2022. "Could Experimental Inflammation Provide Better Understanding of Migraines?" Cells 11, no. 15: 2444. https://doi.org/10.3390/cells11152444

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