Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2515 2023-06-09 09:44:02 |
2 format correct Meta information modification 2515 2023-06-09 10:20:59 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Masood, T.; Lakatos, S.; Rosta, J. TRPA1 as a Factor in Migraine-Related Intracranial Hypersensitivity. Encyclopedia. Available online: https://encyclopedia.pub/entry/45380 (accessed on 27 July 2024).
Masood T, Lakatos S, Rosta J. TRPA1 as a Factor in Migraine-Related Intracranial Hypersensitivity. Encyclopedia. Available at: https://encyclopedia.pub/entry/45380. Accessed July 27, 2024.
Masood, Thannoon, Szandra Lakatos, Judit Rosta. "TRPA1 as a Factor in Migraine-Related Intracranial Hypersensitivity" Encyclopedia, https://encyclopedia.pub/entry/45380 (accessed July 27, 2024).
Masood, T., Lakatos, S., & Rosta, J. (2023, June 09). TRPA1 as a Factor in Migraine-Related Intracranial Hypersensitivity. In Encyclopedia. https://encyclopedia.pub/entry/45380
Masood, Thannoon, et al. "TRPA1 as a Factor in Migraine-Related Intracranial Hypersensitivity." Encyclopedia. Web. 09 June, 2023.
TRPA1 as a Factor in Migraine-Related Intracranial Hypersensitivity
Edit

The transient receptor potential ankyrin 1 (TRPA1) has gained more attention in migraine-related research. The involvement of the TRPA1 receptor in migraine headaches is proposed by the fact that TRPA1 may be a target of some migraine-triggering factors. Although it is doubtful that activation of TRPA1 alone is sufficient to induce pain, behavioral studies have demonstrated that TRPA1 is involved in injury- and inflammation-induced hypersensitivity. Here, the functional relevance of TRPA1 in headaches and its therapeutic potential was introduced, mainly focusing on its role in the development of hypersensitivity, referring to its altered expression in pathological conditions, and its functional interaction with other TRP channels.

TRPA1 migraine hypersensitivity

1. Migraine

Head pain and migraine-associated research is a marginal but notable field of pain investigation. As the most common chronic headache disorder, migraine affects more than 10% of the population [1]. Despite intensive migraine research, the pathomechanisms of migraine and other primary headaches are still unclear.
Migraine is a complex neurological disorder characterized by throbbing head pain, typically localized to one side of the head [2]. Migraine is often accompanied by diverse vegetative and sensory symptoms such as nausea, vomiting, and increased sensitivity to light and sound [3][4]. The recurring headache attacks, with intense pain sensations localized to the head and neck region, can last from hours to days. Chronic migraine is frequently associated with other neurological symptoms, e.g., visual disturbances such as scotoma, blurred vision, and disturbed contours of objects—taken together called a ‘visual aura’ [5]. The phenomenon of aura is also applied to other sensory disturbances, such as increased peripheral pain sensation, paresthesia, dysphasia, or olfactory hallucinations [6]. According to different studies, 10–12% of the population is affected, and two-thirds of migraine patients are women [7][8]. Due to migraine’s high prevalence and financial burden, pain research studies have made a huge effort for decades to reveal the underlying mechanisms. In spite of intensive research and a large number of clinical studies, the pathomechanism of migraine, however, has not yet been clearly revealed.
Experimental migraine studies mainly focus on developing intracranial pain as the most readily detectable symptom of migraine. Headaches categorized as primary, such as migraine, cluster- or tension-type headaches, are classified based on the absence of any specific underlying reason [9][10]. Specifically, pain sensitivity is restricted primarily to the blood vessels of the meninges, such as the dura mater encephali, pia mater, and the major cerebral vessels. The sensory innervation of these structures is raised from the trigeminal nerve. It is well-established that the activation of trigeminal nerves is closely related to head pain [11]. The trigeminovascular system involving primary sensory neurons in the trigeminal ganglion (TG), the trigeminal afferents, and their target structures such as meningeal and cerebral blood vessels [12], as well as the spinal trigeminal nuclei, remains a major focus of current concepts regarding the origins of head pain.

2. Transient Receptor Potential Ankyrin 1 (TRPA1)

In the last two decades, members of the TRP (transient receptor potential) superfamily have been identified to play a pivotal role in peripheral nociception and the sensation of pain [13][14][15][16]. TRPA1—formerly known as ankyrin-like with transmembrane domains protein 1 [ANKTM1], originally described by Story et al. [17]—is expressed by a small subpopulation of dorsal root ganglion neurons, as well as trigeminal neurons [18][19]. TRPA1 is activated by various noxious compounds such as allyl isothiocyanate (AITC), the pungent agent of mustard oil (MO), horseradish, and cinnamaldehyde [20][21]. Acrolein, an irritant compound found in tear gas and cigarette smoke, was the first substance to be identified as a ‘specific TRPA1 activator’, and, furthermore, electrophilic substances such as Δ9-tetrahydrocannabinol, icilin, and 2-aminoethoxydiphenyl borate (2APB) were also cited as TRPA1 activators [20][22].
The functional importance of the TRPA1 channel in peripheral pain sensation is suggested by the fact that TRPA1 expression is typically attributed to the small-sized neurons, commonly referred to as nociceptors. TRPA1-mediated pain sensation was reported in subjects following the administration of a synthetic TRPA1 agonist, suggesting that TRPA1 activation alone is sufficient to evoke pain in patients [23]. In addition, the importance of TRPA1 in human pain sensation is indicated by a genetic disorder in which TRPA1 is affected. Familial episodic pain syndrome, a genetic disorder accompanied by episodic body pain, has been shown to be associated with a TRPA1 deficiency [24].

3. TRPA1 in Migraine-Related Research

3.1. The Presence of TRPA1 in the Trigeminal System

The migraine-inducing effect of environmental irritants has been associated with the activity of certain pain-related receptors over time. In recent decades, the classical TRP channel TRPV1, the mechanosensor Piezo channels, or the purinergic receptor PY have been identified as potential targets of certain migraine-triggering substances. In 2011, when Kunkler et al. reported TRPA1-mediated vasodilation in meningeal vessels, TRPA1 came into the focus of migraine-related research [25]. As TRPA1 has gained more attention in relation to head pain, the ratio of the TRPA1-expressing neurons in the intracranial sensory system involving the TG has become the subject of investigation. By using an in situ hybridization technique, it was found that TRPA1 mRNA is expressed in 36.5% of all trigeminal primary sensory neurons [26]. Similar results are obtained by functional studies, as the application of the TRPA1 agonist AITC stimulates nearly 35% of rat trigeminal neurons [20]. TG is responsible for the sensory innervation of the entire head region, including the facial skin, nasal mucosa, intracranial structures, meningeal vasculature, etc. Focusing on head pain, the neurons responsible for innervating the pain-sensitive structures within the skull must be distinguished from the overall population. Applying whole-cell patch-clamp recording, approximately 40% of the dural projection neurons responded to the application of the TRPA1 agonist MO, representing a slightly higher proportion than the total trigeminal neuronal population [27]. This result supports the substantial role of TRPA1 in intracranial sensory processes, as more than one-third of primary sensory neurons—considered potential pain-sensitive neurons—are involved in TRPA1-mediated reactions. Interestingly, only a small percentage of trigeminal neurons projecting to the dura show TRPA1-immunoreactivity [28]. However, the TRPA1-immunoreactive neurons cluster around most of the dural projection neurons, suggesting potential cross-excitation between the neighboring neurons.

3.2. Relevance of TRPA1 in Migraine-Like Pain

NO was the first experimentally confirmed migraine-provoking factor, and its experimental administration has since been used as a widely accepted migraine model [29]. Tear gas was also considered a headache trigger, as in 29% of seizure events, patients complained about post-seizure headaches [30]. In recent decades, several natural compounds have been identified as headache-provoking factors: camphor, tear gas, formalin, cigarette smoke, or alcoholic beverages—all these agents are well-known headache triggers [31]. TRPA1 was originally identified as a ‘nociceptive’ receptor activated by thermal and mechanical stimuli; however, more and more headache-inducing factors have also been revealed to be TRPA1 activators. Several migraine triggers, such as formalin, cigarette smoke, and even tear gas, have been shown to act through the TRPA1 channel [25][32][33][34]. Moreover, the classical trigger, nitric oxide (NO), has also been found to induce pain sensations through the activation of TRPA1 [35].
In addition, an endogenous product, 4-hydroxynonenal, which is associated with several chronic diseases, has been shown to act as a potential activator of TRPA1, supporting the relevance of TRPA1-activation in chronic pain disorders such as migraine [36].
All of the above studies confirmed that certain migraine-inducing agents are potential activators of TRPA1; however, they did not unambiguously establish a direct link between the headache and TRPA1 receptor activation. It has recently been shown that TRPA1 contributes to migraine-associated hypersensitivity, since the TRPA1 antagonist, ADM_12, prevented the nitroglycerine-induced hyperalgesia in both the acute and chronic migraine models [37].
It was recognized early that headache attacks are accompanied by intracranial vascular reactions [38][39]. Vascular events were originally thought to cause the sensation of pain during headache attacks. However, the direct relationship between head pain and vascular reactions was disproved by modern imaging studies two decades ago [40]. At the same time, meningeal vasodilation is a clear consequence of the activation of chemosensitive trigeminal afferents, so the monitoring of intracranial vascular reactions following trigeminal nerve stimulation serves as an approved model for migraine.
The phenomenon that inhalation of environmental irritants frequently triggers a migraine attack has led researchers to investigate the functional connection between nasal stimulation and headache-related meningeal events. The intranasal administration of chemical compounds and recording the evoked meningeal reactions provide a specific aspect of headache research. The putative role of TRPA1 in intracranial sensitization is supported by Kunkler et al. [25]. A nasal application of the TRPA1 activators MO and acrolein has been shown to increase the meningeal blood flow in a CGRP-dependent way. Subsequently, this group also demonstrated the sensitizing effect of TRPA1 activation on trigeminovascular responses, as chronic exposure to acrolein enhances TRPA1-mediated meningeal vasodilation and leads to facial allodynia [41]. In 2012, an experimental study showed that activation of TRPA1 via an injected dural cannula increased the pain sensation in the facial region [27]. Another behavioral study proved that umbellulone (UMB), the active compound of Umbellularia californica (‘headache tree’), causes trigeminal pain via TRPA1 [42]. It found that in 40% of rat trigeminal ganglion neurons, the administration of UMB could stimulate TRPA1, and intranasal and intravenous administration induced nociceptive behavior in a TRPA1-dependent manner

4. Possible Cooperation between TRPA1 and TRPV1

It is well-documented that other classical TRP channels, e.g., the transient receptor potential vanilloid 1 (TRPV1), have a pivotal role in inflammation-related hypersensitivity such as thermal hyperalgesia and allodynia [43][44][45]. Hypersensitivity can develop as a result of a functional modification of the TRPV1 by phosphorylation-induced sensitization [46][47][48]. Considering the importance of intracranial sensitization in the pathomechanism of migraine, the functional sensitization of TRPA1, TRPV1, and other TRP receptors involved in intracranial pain sensation at the receptor level should be discussed as a provocative factor in migraine attacks.
Based on its structure, TRPA1 belongs to the TRP superfamily. As a common feature, TRP subfamilies possess a tetrameric structure, each involving six transmembrane domains (S1 to S6) with a highly conservated ankyrin repeat and a ’TRP domain’ [16]. TRPA1 is activated by irritants and pungent agents (e.g., acrolein, AITC) through a reversible covalent modification of specific cysteine residues in the N-terminal [49]. Being an ion channel protein, a pore is formed between the S5 and S6 domains. Molecular modeling and sequence alignment revealed individual residues within the S5 and S6 segments, which are responsible for the opening and activation of the channel [43].
Despite the homology, TRP channels respond to a wide range of stimuli and possess diverse gating mechanisms, ion selectivity, and functions. However, the structural similarities may provide a basis for interaction between distinct TRP receptors when co-expressed [50]. The cooperation between TRPA1 and TRPV1 is suggested by a high-order co-expression of the two TRP receptors. It was earlier described that nearly all TRPA1-sensitive neurons also express TRPV1 [17]. Functional studies proved that the two channels can co-work in a synergistic manner [43][44]. Interestingly, the inflammatory mediator, bradykinin, has been found to activate both TRPA1 and TRPV1, and it seems that both receptors are required for the sensory neurons’ maximum response elicited by bradykinin [45]. The TRPA1 channel activity has been shown to be regulated and modulated by the expression of TPRV1 in trigeminal neurons [51].
Novel research has revealed possible cooperation between TRPA1 and TRPV1 by attaching distinct subunits to create structurally atypical heteromers. Fisher et al. created TRPA1:TRPV1 complexes in cultured sensory neurons. The heteromers that formed were found to be less sensitive to TRPV1 agonists; therefore, it was concluded that TRPV1 activity is inhibited by its combination with TRPA1 subunits [52]. TRPA1 function also appears to be negatively regulated by the complex formation, as TRPA1:TRPV1 heteromers failed to respond to TRPA1 agonists. All TRPs are intended to form tetrameric assemblies, but different TRP subfamilies are able to combine their subunits to form heterotetramer complexes. Although the combination of TRP subunits is not so obvious, the interaction is provided by special modules in the transmembrane domains [53]. The research group also identified a unique domain on the intracellular terminals of TRPV1 that is thought to be responsible for the formation of the TRPA1:TRPV1 complex. The ablation of the identified residues reduced the possibility of the TRPA1:TRPV1 interaction.
Tmem100, a conserved transmembrane protein, has been identified as a key regulator of TRPA1:TRPV1 cooperation [54]. This impressive study demonstrated that Tmem100 can physically interact with both TRPA1 and TRPV1 and reduce their association. Tmem100 appears to regulate TRPA1 activity in a TRPV1-dependent manner: in the absence of TRPV1, Tmem100 reduced the activity of the TRPA1. Weng et al. concluded that Tmem100 enhances TRPA1 activity by lowering the potential of TRPV1 to form complexes. They designed a peptide that blocks Tmem100 and consequently inhibits TRPA1 activity. The method could provide specific selective suppression of TRPV1- and TPRA1-mediated pain. The functional significance of the TRPA1:TRPV1 complex in the nociceptive pathway is under ongoing investigation and has not yet been clarified. Such physical interactions between two receptors can alter the receptor properties and excitability, so the cooperation of TRPA1 and TRPV1 should be carefully considered in chronic pain-related disease, such as in migraine, studies (Figure 1).
Figure 1. The significance of TRPA1 activation in the development of migraine-related hyperalgesia. The figure shows co-expression of the TRP channels TRPA1 and TRPV1 by primary sensory trigeminal neurons and illustrates a possible interaction between the two channels. Following activation of TRPA1 or TRPV1, heterotetrameric channels can be formed, which may display modified characteristics of both channels, leading to an increase in sensitivity and a decrease in the threshold of the affected neuron. The proposed interaction between TRPs may play a role in migraine-related hypersensitivity.
The TRPA1 channel is also known to form complexes with other ‘pain-related’ receptors, such as the N-methyl-d-aspartate receptor and the delta-opioid receptor [55]. These complexes may also contribute to the modification of the pain sensation and may be involved in the transduction of pain.

5. Possible Therapeutic Interventions

Supposing the relevant role of TRPA1 in peripheral nociception, the inhibition of TRPA1 appears to be a promising approach in pain therapy.
The first specific TRPA1 antagonists, such as HC–030031, AP–18, and A–967079, were developed and tested more than 10 years ago. Since then, additional TRPA1 antagonists have been developed, and some have also advanced to human trials [56][57][58][59]. However, differences in TRPA1 across species restrict translational research, and most clinical trials have been discontinued due to pharmacokinetic deficiencies; others have yet to be reported.
Desensitization is a well-known and widely studied phenomenon and provides a possible analgesic method in pain research: following initial activation, certain agonists, due to the accumulation of incoming Ca2+, may cause the inactivation of the receptor channel and render the channel unresponsive to any further stimulus. This stimulus-evoked desensitization is characteristic of all TRP channels. The therapeutic effect of the desensitization of TRPA1 is supported by a study that showed that safranal, an analgesic plant compound, provides its analgesic effect through the selective desensitization of TRPA1 [60]. Another analgesic peptide, crotalphine, has also been shown to provide its antinociceptive effect through the long-lasting desensitization of TRPA1 [61].
Extract of butterbur (Petasites hybridus), which is used in traditional medicine, is known to have anti-migraine effects. A clinical study confirmed the preventive effect of butterbur in the treatment of migraine [62]. Nearly two decades later, isopetasin, an active component of butterbur, was shown to have a desensitizing effect on TRPA1, which effect is suggested to be responsible for the anti-migraine action [63].

References

  1. Burch, R.C.; Buse, D.C.; Lipton, R.B. Migraine: Epidemiology, Burden, and Comorbidity. Neurol. Clin. 2019, 37, 631–649.
  2. Solomon, S.; Cappa, K.G.; Smith, C.R. Common migraine: Criteria for diagnosis. Headache 1988, 28, 124–129.
  3. Han, S.M.; Kim, K.M.; Cho, S.-J.; Yang, K.I.; Kim, D.; Yun, C.-H.; Chu, M.K. Prevalence and characteristics of cutaneous allodynia in probable migraine. Sci. Rep. 2021, 11, 2467.
  4. Ray, J.C.; Cheema, S.; Foster, E.; Gunasekera, L.; Mehta, D.; Corcoran, S.J.; Matharu, M.S.; Hutton, E.J. Autonomic symptoms in migraine: Results of a prospective longitudinal study. Front. Neurol. 2022, 13, 1036798.
  5. Luda, E.; Bo, E.; Sicuro, L.; Comitangelo, R.; Campana, M. Sustained visual aura: A totally new variation of migraine. Headache 1991, 31, 582–583.
  6. Peatfield, R.C.; Gawel, M.J.; Rose, F.C. Asymmetry of the aura and pain in migraine. J. Neurol. Neurosurg. Psychiatry 1981, 44, 846–848.
  7. Russell, M.B.; Rasmussen, B.K.; Thorvaldsen, P.; Olesen, J. Prevalence and sex-ratio of the subtypes of migraine. Int. J. Epidemiol. 1995, 24, 612–618.
  8. Rasmussen, B.K.; Olesen, J. Migraine with aura and migraine without aura: An epidemiological study. Cephalalgia Int. J. Headache 1992, 12, 221–228; discussion 186.
  9. Nappi, G.; Agnoli, A.; Manzoni, G.C.; Nattero, G.; Sicuteri, F. Classification and diagnostic criteria for primary headache disorders (Ad Hoc Committee IHS, 1988). Funct. Neurol. 1989, 4, 65–71.
  10. Lipton, R.B.; Goadsby, P.; Silberstein, S.D. Classification and epidemiology of headache. Clin. Cornerstone 1999, 1, 1–10.
  11. Feindel, W.; Penfield, W.; McNaughton, F. The tentorial nerves and Iocalization of intracranial pain in man. Neurology 1960, 10, 555–563.
  12. Moskowitz, M.A. The neurobiology of vascular head pain. Ann. Neurol. 1984, 16, 157–168.
  13. Caterina, M.J.; Rosen, T.A.; Tominaga, M.; Brake, A.J.; Julius, D. A capsaicin-receptor homologue with a high threshold for noxious heat. Nature 1999, 398, 436–441.
  14. Lee, H.; Iida, T.; Mizuno, A.; Suzuki, M.; Caterina, M.J. Altered thermal selection behavior in mice lacking transient receptor potential vanilloid 4. J. Neurosci. Off. J. Soc. Neurosci. 2005, 25, 1304–1310.
  15. Bautista, D.M.; Siemens, J.; Glazer, J.M.; Tsuruda, P.R.; Basbaum, A.I.; Stucky, C.L.; Jordt, S.-E.; Julius, D. The menthol receptor TRPM8 is the principal detector of environmental cold. Nature 2007, 448, 204–208.
  16. Montell, C. The TRP superfamily of cation channels. Sci. STKE Signal Transduct. Knowl. Environ. 2005, 2005, re3.
  17. Story, G.M.; Peier, A.M.; Reeve, A.J.; Eid, S.R.; Mosbacher, J.; Hricik, T.R.; Earley, T.J.; Hergarden, A.C.; Andersson, D.A.; Hwang, S.W.; et al. ANKTM1, a TRP-like Channel Expressed in Nociceptive Neurons, Is Activated by Cold Temperatures. Cell 2003, 112, 819–829.
  18. Karashima, Y.; Talavera, K.; Everaerts, W.; Janssens, A.; Kwan, K.Y.; Vennekens, R.; Nilius, B.; Voets, T. TRPA1 acts as a cold sensor in vitro and in vivo. Proc. Natl. Acad. Sci. USA 2009, 106, 1273–1278.
  19. Kobayashi, K.; Fukuoka, T.; Obata, K.; Yamanaka, H.; Dai, Y.; Tokunaga, A.; Noguchi, K. Distinct expression of TRPM8, TRPA1, and TRPV1 mRNAs in rat primary afferent neurons with adelta/c-fibers and colocalization with trk receptors. J. Comp. Neurol. 2005, 493, 596–606.
  20. Jordt, S.-E.; Bautista, D.M.; Chuang, H.-H.; McKemy, D.D.; Zygmunt, P.M.; Högestätt, E.D.; Meng, I.D.; Julius, D. Mustard oils and cannabinoids excite sensory nerve fibres through the TRP channel ANKTM1. Nature 2004, 427, 260–265.
  21. Bandell, M.; Story, G.M.; Hwang, S.W.; Viswanath, V.; Eid, S.R.; Petrus, M.J.; Earley, T.J.; Patapoutian, A. Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin. Neuron 2004, 41, 849–857.
  22. Babes, A.; Zorzon, D.; Reid, G. Two populations of cold-sensitive neurons in rat dorsal root ganglia and their modulation by nerve growth factor. Eur. J. Neurosci. 2004, 20, 2276–2282.
  23. Andersen, H.H.; Lo Vecchio, S.; Gazerani, P.; Arendt-Nielsen, L. Dose-response study of topical allyl isothiocyanate (mustard oil) as a human surrogate model of pain, hyperalgesia, and neurogenic inflammation. Pain 2017, 158, 1723–1732.
  24. Kremeyer, B.; Lopera, F.; Cox, J.J.; Momin, A.; Rugiero, F.; Marsh, S.; Woods, C.G.; Jones, N.G.; Paterson, K.J.; Fricker, F.R.; et al. A gain-of-function mutation in TRPA1 causes familial episodic pain syndrome. Neuron 2010, 66, 671–680.
  25. Kunkler, P.E.; Ballard, C.J.; Oxford, G.S.; Hurley, J.H. TRPA1 receptors mediate environmental irritant-induced meningeal vasodilatation. Pain 2011, 152, 38–44.
  26. Nagata, K.; Duggan, A.; Kumar, G.; García-Añoveros, J. Nociceptor and Hair Cell Transducer Properties of TRPA1, a Channel for Pain and Hearing. J. Neurosci. 2005, 25, 4052–4061.
  27. Edelmayer, R.M.; Le, L.N.; Yan, J.; Wei, X.; Nassini, R.; Materazzi, S.; Preti, D.; Appendino, G.; Geppetti, P.; Dodick, D.W.; et al. Activation of TRPA1 on dural afferents: A potential mechanism of headache pain. Pain 2012, 153, 1949–1958.
  28. Huang, D.; Li, S.; Dhaka, A.; Story, G.M.; Cao, Y.-Q. Expression of the transient receptor potential channels TRPV1, TRPA1 and TRPM8 in mouse trigeminal primary afferent neurons innervating the dura. Mol. Pain 2012, 8, 66.
  29. Iversen, H.K. Experimental headache in humans. Cephalalgia Int. J. Headache 1995, 15, 281–287.
  30. Anderson, P.J.; Lau, G.S.; Taylor, W.R.; Critchley, J.A. Acute effects of the potent lacrimator o-chlorobenzylidene malononitrile (CS) tear gas. Hum. Exp. Toxicol. 1996, 15, 461–465.
  31. Kelman, L. The triggers or precipitants of the acute migraine attack. Cephalalgia Int. J. Headache 2007, 27, 394–402.
  32. Rozen, T.D. Cluster headache as the result of secondhand cigarette smoke exposure during childhood. Headache 2010, 50, 130–132.
  33. Brône, B.; Peeters, P.J.; Marrannes, R.; Mercken, M.; Nuydens, R.; Meert, T.; Gijsen, H.J.M. Tear gasses CN, CR, and CS are potent activators of the human TRPA1 receptor. Toxicol. Appl. Pharmacol. 2008, 231, 150–156.
  34. McNamara, C.R.; Mandel-Brehm, J.; Bautista, D.M.; Siemens, J.; Deranian, K.L.; Zhao, M.; Hayward, N.J.; Chong, J.A.; Julius, D.; Moran, M.M.; et al. TRPA1 mediates formalin-induced pain. Proc. Natl. Acad. Sci. USA 2007, 104, 13525–13530.
  35. Miyamoto, T.; Dubin, A.E.; Petrus, M.J.; Patapoutian, A. TRPV1 and TRPA1 mediate peripheral nitric oxide-induced nociception in mice. PLoS ONE 2009, 4, e7596.
  36. Trevisani, M.; Siemens, J.; Materazzi, S.; Bautista, D.M.; Nassini, R.; Campi, B.; Imamachi, N.; Andrè, E.; Patacchini, R.; Cottrell, G.S.; et al. 4-Hydroxynonenal, an endogenous aldehyde, causes pain and neurogenic inflammation through activation of the irritant receptor TRPA1. Proc. Natl. Acad. Sci. USA 2007, 104, 13519–13524.
  37. Demartini, C.; Greco, R.; Magni, G.; Zanaboni, A.M.; Riboldi, B.; Francavilla, M.; Nativi, C.; Ceruti, S.; Tassorelli, C. Modulation of Glia Activation by TRPA1 Antagonism in Preclinical Models of Migraine. Int. J. Mol. Sci. 2022, 23, 14085.
  38. Tunis, M.M.; Wolff, H.G. The hemodynamic analysis of cranial artery pulse wave contours in vascular headache of the migraine type. Trans. Am. Neurol. Assoc. 1952, 56, 22–25.
  39. Kobari, M.; Meyer, J.S.; Ichijo, M.; Imai, A.; Oravez, W.T. Hyperperfusion of cerebral cortex, thalamus and basal ganglia during spontaneously occurring migraine headaches. Headache 1989, 29, 282–289.
  40. May, A.; Büchel, C.; Turner, R.; Goadsby, P.J. Magnetic resonance angiography in facial and other pain: Neurovascular mechanisms of trigeminal sensation. J. Cereb. Blood Flow Metab. Off. J. Int. Soc. Cereb. Blood Flow Metab. 2001, 21, 1171–1176.
  41. Kunkler, P.E.; Zhang, L.; Pellman, J.J.; Oxford, G.S.; Hurley, J.H. Sensitization of the Trigeminovascular System following Environmental Irritant Exposure. Cephalalgia Int. J. Headache 2015, 35, 1192–1201.
  42. Nassini, R.; Materazzi, S.; Vriens, J.; Prenen, J.; Benemei, S.; De Siena, G.; la Marca, G.; Andrè, E.; Preti, D.; Avonto, C.; et al. The “headache tree” via umbellulone and TRPA1 activates the trigeminovascular system. Brain J. Neurol. 2012, 135, 376–390.
  43. Caterina, M.J.; Leffler, A.; Malmberg, A.B.; Martin, W.J.; Trafton, J.; Petersen-Zeitz, K.R.; Koltzenburg, M.; Basbaum, A.I.; Julius, D. Impaired nociception and pain sensation in mice lacking the capsaicin receptor. Science 2000, 288, 306–313.
  44. Davis, J.B.; Gray, J.; Gunthorpe, M.J.; Hatcher, J.P.; Davey, P.T.; Overend, P.; Harries, M.H.; Latcham, J.; Clapham, C.; Atkinson, K.; et al. Vanilloid receptor-1 is essential for inflammatory thermal hyperalgesia. Nature 2000, 405, 183–187.
  45. Chan, C.L.H.; Facer, P.; Davis, J.B.; Smith, G.D.; Egerton, J.; Bountra, C.; Williams, N.S.; Anand, P. Sensory fibres expressing capsaicin receptor TRPV1 in patients with rectal hypersensitivity and faecal urgency. Lancet Lond. Engl. 2003, 361, 385–391.
  46. Ogawa, N.; Kurokawa, T.; Fujiwara, K.; Polat, O.K.; Badr, H.; Takahashi, N.; Mori, Y. Functional and Structural Divergence in Human TRPV1 Channel Subunits by Oxidative Cysteine Modification. J. Biol. Chem. 2016, 291, 4197–4210.
  47. Kumar, R.; Hazan, A.; Geron, M.; Steinberg, R.; Livni, L.; Matzner, H.; Priel, A. Activation of transient receptor potential vanilloid 1 by lipoxygenase metabolites depends on PKC phosphorylation. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2017, 31, 1238–1247.
  48. Robilotto, G.L.; Mohapatra, D.P.; Shepherd, A.J.; Mickle, A.D. Role of Src kinase in regulating protein kinase C mediated phosphorylation of TRPV1. Eur. J. Pain Lond. Engl. 2022, 26, 1967–1978.
  49. Zhang, L.; Kunkler, P.E.; Knopp, K.L.; Oxford, G.S.; Hurley, J.H. Role of intraganglionic transmission in the trigeminovascular pathway. Mol. Pain 2019, 15, 1744806919836570.
  50. Minke, B.; Cook, B. TRP Channel Proteins and Signal Transduction. Physiol. Rev. 2002, 82, 429–472.
  51. Salas, M.M.; Hargreaves, K.M.; Akopian, A.N. TRPA1-mediated responses in trigeminal sensory neurons: Interaction between TRPA1 and TRPV1. Eur. J. Neurosci. 2009, 29, 1568–1578.
  52. Fischer, M.J.M.; Balasuriya, D.; Jeggle, P.; Goetze, T.A.; McNaughton, P.A.; Reeh, P.W.; Edwardson, J.M. Direct evidence for functional TRPV1/TRPA1 heteromers. Pflüg. Arch.-Eur. J. Physiol. 2014, 466, 2229–2241.
  53. Clapham, D.E. TRP channels as cellular sensors. Nature 2003, 426, 517–524.
  54. Weng, H.-J.; Patel, K.N.; Jeske, N.A.; Bierbower, S.M.; Zou, W.; Tiwari, V.; Zheng, Q.; Tang, Z.; Mo, G.C.H.; Wang, Y.; et al. Tmem100 Is a Regulator of TRPA1-TRPV1 Complex and Contributes to Persistent Pain. Neuron 2015, 85, 833–846.
  55. Cortés-Montero, E.; Rodríguez-Muñoz, M.; Ruiz-Cantero, M.D.C.; Cobos, E.J.; Sánchez-Blázquez, P.; Garzón-Niño, J. Calmodulin Supports TRPA1 Channel Association with Opioid Receptors and Glutamate NMDA Receptors in the Nervous Tissue. Int. J. Mol. Sci. 2020, 22, 229.
  56. Eid, S.R.; Crown, E.D.; Moore, E.L.; Liang, H.A.; Choong, K.-C.; Dima, S.; Henze, D.A.; Kane, S.A.; Urban, M.O. HC-030031, a TRPA1 selective antagonist, attenuates inflammatory- and neuropathy-induced mechanical hypersensitivity. Mol. Pain 2008, 4, 48.
  57. Hu, H.; Tian, J.; Zhu, Y.; Wang, C.; Xiao, R.; Herz, J.M.; Wood, J.D.; Zhu, M.X. Activation of TRPA1 channels by fenamate nonsteroidal anti-inflammatory drugs. Pflug. Arch. 2010, 459, 579–592.
  58. Chen, J.; Joshi, S.K.; DiDomenico, S.; Perner, R.J.; Mikusa, J.P.; Gauvin, D.M.; Segreti, J.A.; Han, P.; Zhang, X.-F.; Niforatos, W.; et al. Selective blockade of TRPA1 channel attenuates pathological pain without altering noxious cold sensation or body temperature regulation. Pain 2011, 152, 1165–1172.
  59. Mesch, S.; Walter, D.; Laux-Biehlmann, A.; Basting, D.; Flanagan, S.; Miyatake Ondozabal, H.; Bäurle, S.; Pearson, C.; Jenkins, J.; Elves, P.; et al. Discovery of BAY-390, a Selective CNS Penetrant Chemical Probe as Transient Receptor Potential Ankyrin 1 (TRPA1) Antagonist. J. Med. Chem. 2023, 66, 1583–1600.
  60. Li Puma, S.; Landini, L.; Macedo, S.J.; Seravalli, V.; Marone, I.M.; Coppi, E.; Patacchini, R.; Geppetti, P.; Materazzi, S.; Nassini, R.; et al. TRPA1 mediates the antinociceptive properties of the constituent of Crocus sativus L., safranal. J. Cell. Mol. Med. 2019, 23, 1976–1986.
  61. Bressan, E.; Touska, F.; Vetter, I.; Kistner, K.; Kichko, T.I.; Teixeira, N.B.; Picolo, G.; Cury, Y.; Lewis, R.J.; Fischer, M.J.M.; et al. Crotalphine desensitizes TRPA1 ion channels to alleviate inflammatory hyperalgesia. Pain 2016, 157, 2504–2516.
  62. Lipton, R.B.; Göbel, H.; Einhäupl, K.M.; Wilks, K.; Mauskop, A. Petasites hybridus root (butterbur) is an effective preventive treatment for migraine. Neurology 2004, 63, 2240–2244.
  63. Benemei, S.; De Logu, F.; Li Puma, S.; Marone, I.M.; Coppi, E.; Ugolini, F.; Liedtke, W.; Pollastro, F.; Appendino, G.; Geppetti, P.; et al. The anti-migraine component of butterbur extracts, isopetasin, desensitizes peptidergic nociceptors by acting on TRPA1 cation channel. Br. J. Pharmacol. 2017, 174, 2897–2911.
More
Information
Subjects: Biology
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , ,
View Times: 303
Revisions: 2 times (View History)
Update Date: 09 Jun 2023
1000/1000
Video Production Service