Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Diencephalic and brainstem mechanisms in migraine

Key Points

  • Migraine is a chronic and disabling brain disorder that typically manifests as attacks of one-sided, often throbbing head pain that is worsened by movement and is associated with nausea and sensitivity to light (photophobia) and sound (phonophobia).

  • Migraine is thought to involve activation of the trigeminovascular system, which includes the efferent projections to the pain-producing dura mater and large intracranial vessels, and the afferent projection to the central trigeminal nucleus caudalis (TNC). It is a pivotal relay centre that passes nociceptive information from the cranial vasculature to the brainstem and higher processing centres.

  • Several studies have demonstrated that structures within the brainstem, midbrain and forebrain — areas that are known to be involved in the modulation of the trigeminovascular system — are active during spontaneous migraine. This suggests that dysfunction within these regions may be responsible for a migrainous phenotype.

  • In the pons, the superior salivatory nucleus (SuS) has a reflex connection with the TNC and provides the cells of origin of the parasymapathetic outflow to the cranial vasculature. Activation contributes to the autonomic symptoms in primary headache and has now been shown to also activate and exacerbate neuronal firing in the TNC.

  • In the midbrain, the ventrolateral periaqueductal grey and rostral ventromedial medulla provide descending control of trigeminovascular nociceptive responses, controlled by 'on'- and 'off'-cell activation. Additionally, connections with hypothalamic, thalamic and limbic nuclei, which are involved in homeostatic processes, suggest that 'on'–'off' cell activation, or the dysregulation of these cells, may be involved in triggering migraine and may contribute to migrainous symptoms.

  • In the forebrain, the hypothalamus is involved in the control of the sleep–wake cycle, feeding, thirst, urination and arousal — behaviours that are altered during premonitory symptoms. Disruption to the regular functioning of these behaviours can also serve as triggers for migraine. Bidirectional connections with the SuS may provide a link between the potential site of origin of migraine triggers, premonitory symptoms and other migraine symptoms, and descending modulation of trigeminovascular nociceptive traffic.

  • The thalamus is a major centre for processing nociceptive information. Sensitization of third-order trigeminovascular nociceptive neurons in thalamic nuclei is likely to contribute to the cutaneous allodynia that is experienced by patients. Sensitized neurons of the posterior thalamus that receive projections from areas of the visual cortex respond to bright light with exacerbated firing. This may provide a neural explanation for the photophobia that is exhibited by patients who suffer from migraine.

  • We believe that the explanation that best accounts for the multifaceted migrainous symptomology is a dysfunction of the brainstem or diencenphalic nuclei that process nociceptive inputs of the craniovascular afferents. The bidirectional connections of the many brainstem and diencephalic nuclei mean that dysfunction can create a brain state that produces many simultaneous symptoms. The next best theory involves a sequential sensitization of trigeminovascular synapses up to the thalamus that requires an initial peripheral nociceptive event. This cannot explain many centrally mediated migraine triggers or the premonitory symptoms that precede any pain.

Abstract

Migraine is a common and complex brain disorder. Although it is clear that head pain is a key manifestation of the disorder for most patients, what drives the activation of neuronal pain pathways in susceptible patients is less obvious. There is growing evidence that migraine pathophysiology may, in part, include dysfunction of subcortical structures. These include diencephalic and brainstem nuclei that can modulate the perception of activation of the trigeminovascular system, which carries sensory information from the cranial vasculature to the brain. Dysfunction of these nuclei, and their connections to other key brain centres, may contribute to the cascade of events that results in other symptoms of migraine — such as light and sound sensitivity — thus providing a comprehensive explanation of the neurobiology of the disorder.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Neuronal pathways involved in migraine pathophysiology.
Figure 2: Descending modulation of trigeminovascular nociceptive transmission through midbrain nuclei.
Figure 3: Descending modulation of trigeminovascular nociceptive transmission by hypothalamic nuclei.
Figure 4: Trigeminovascular and visual inputs to the thalamus and visual cortex.

Similar content being viewed by others

References

  1. Goadsby, P. J., Lipton, R. B. & Ferrari, M. D. Migraine — current understanding and treatment. N. Engl. J. Med. 346, 257–270 (2002).

    Article  CAS  PubMed  Google Scholar 

  2. Lipton, R. B. et al. Migraine prevalence, disease burden, and the need for preventive therapy. Neurology 68, 343–349 (2007).

    Article  CAS  PubMed  Google Scholar 

  3. Lipton, R. B., Stewart, W. F., Diamond, S., Diamond, M. L. & Reed, M. Prevalence and burden of migraine in the United States: data from the American Migraine Study II. Headache 41, 646–657 (2001).

    Article  CAS  PubMed  Google Scholar 

  4. Stewart, W. F., Ricci, J. A., Chee, E. & Morganstein, D. Lost productive work time costs from health conditions in the United States: results from the American Productivity Audit. J. Occup. Environ. Med. 45, 1234–1246 (2003).

    Article  PubMed  Google Scholar 

  5. Andlin-Sobocki, P., Jonsson, B., Wittchen, H. U. & Olesen, J. Cost of disorders of the brain in Europe. Eur. J. Neurol. 12, 1–27 (2005).

    Article  PubMed  Google Scholar 

  6. Headache Classification Committee of the International Headache Society. The international classification of headache disorders. 2nd ed. Cephalalgia Suppl. 24, 9–160 (2004).

  7. Ferrari, M. D. Treatment of migraine attacks with sumatriptan. The Subcutaneous Sumatriptan International Study Group. N. Engl. J. Med. 325, 316–321 (1991).

    Article  Google Scholar 

  8. Humphrey, P. P. A. et al. Preclinical studies on the anti-migraine drug, sumatriptan. Eur. Neurol. 31, 282–290 (1991).

    Article  CAS  PubMed  Google Scholar 

  9. May, A. & Goadsby, P. J. The trigeminovascular system in humans: pathophysiologic implications for primary headache syndromes of the neural influences on the cerebral circulation. J. Cereb. Blood Flow Metab. 19, 115–127 (1999).

    Article  CAS  PubMed  Google Scholar 

  10. Goadsby, P. J. Pathophysiology of migraine. Neurologic Clinics 27, 335–360 (2009).

    Article  PubMed  Google Scholar 

  11. Goadsby, P. J., Charbit, A. R., Andreou, A. P., Akerman, S. & Holland, P. R. Neurobiology of migraine. Neuroscience 161, 327–341 (2009).

    Article  CAS  PubMed  Google Scholar 

  12. Dalkara, T., Nozari, A. & Moskowitz, M. A. Migraine aura pathophysiology: the role of blood vessels and microembolisation. Lancet Neurol. 9, 309–317 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Eikermann-Haerter, K. & Moskowitz, M. A. Animal models of migraine headache and aura. Curr. Opin. Neurol. 21, 294–300 (2008).

    Article  PubMed  Google Scholar 

  14. Eikermann-Haerter, K. & Ayata, C. Cortical spreading depression and migraine. Curr. Neurol. Neurosci. Rep. 10, 167–173 (2010).

    Article  PubMed  Google Scholar 

  15. Wolthausen, J., Sternberg, S., Gerloff, C. & May, A. Are cortical spreading depression and headache in migraine causally linked? Cephalalgia 29, 244–249 (2009).

    Article  CAS  PubMed  Google Scholar 

  16. Goadsby, P. J. Migraine, aura and cortical spreading depression: why are we still talking about it? Ann. Neurol. 49, 4–6 (2001).

    Article  CAS  PubMed  Google Scholar 

  17. Ayata, C. Cortical spreading depression triggers migraine attack: pro. Headache 50, 725–730 (2010).

    Article  PubMed  Google Scholar 

  18. Charles, A. Does cortical spreading depression initiate a migraine attack? Maybe not. Headache 50, 731–733 (2010).

    Article  PubMed  Google Scholar 

  19. McNaughton, F. L. & Feindel, W. in Physiological Aspects Of Clinical Neurology (ed. Rose, F. C.) 279–293 (Blackwell Scientific, Oxford, 1977).

    Google Scholar 

  20. Penfield, W. & F., M. Dural headache and innervation of the dura mater. Arch. Neurol. Psychiatry 44, 43–75 (1940).

    Article  Google Scholar 

  21. Ray, B. S. & Wolff, H. G. Experimental studies on headache. Pain sensitive structures of the head and their significance in headache. Arch. Surg. 41, 813–856 (1940).

    Article  Google Scholar 

  22. Goadsby, P. J. & Zagami, A. S. Stimulation of the superior sagittal sinus increases metabolic activity and blood flow in certain regions of the brainstem and upper cervical spinal cord of the cat. Brain 114, 1001–1011 (1991).

    Article  PubMed  Google Scholar 

  23. Kaube, H., Keay, K. A., Hoskin, K. L., Bandler, R. & Goadsby, P. J. Expression of c-Fos-like immunoreactivity in the caudal medulla and upper cervical spinal cord following stimulation of the superior sagittal sinus in the cat. Brain Res. 629, 95–102 (1993).

    Article  CAS  PubMed  Google Scholar 

  24. Hoskin, K. L., Zagami, A. S. & Goadsby, P. J. Stimulation of the middle meningeal artery leads to Fos expression in the trigeminocervical nucleus: a comparative study of monkey and cat. J. Anat. 194, 579–588 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. 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. 79, 964–982 (1998).

    Article  CAS  PubMed  Google Scholar 

  26. Knight, Y. E. et al. Patterns of fos expression in the rostral medulla and caudal pons evoked by noxious craniovascular stimulation and periaqueductal gray stimulation in the cat. Brain Res. 1045, 1–11 (2005).

    Article  CAS  PubMed  Google Scholar 

  27. Bartsch, T. & Goadsby, P. J. Stimulation of the greater occipital nerve induces increased central excitability of dural afferent input. Brain 125, 1496–1509 (2002).

    Article  PubMed  Google Scholar 

  28. Bartsch, T. & Goadsby, P. J. Increased responses in trigeminocervical nociceptive neurons to cervical input after stimulation of the dura mater. Brain 126, 1801–1813 (2003).

    Article  CAS  PubMed  Google Scholar 

  29. Bartsch, T. & Goadsby, P. J. Anatomy and physiology of pain referral in primary and cervicogenic headache disorders. Headache Curr. 2, 42–48 (2005).

    Article  Google Scholar 

  30. Liu, Y., Broman, J., Zhang, M. & Edvinsson, L. Brainstem and thalamic projections from a craniovascular sensory nervous centre in the rostral cervical spinal dorsal horn of rats. Cephalalgia 29, 935–948 (2009).

    Article  CAS  PubMed  Google Scholar 

  31. Hoskin, K. L., Bulmer, D. C. E., Lasalandra, M., Jonkman, A. & Goadsby, P. J. Fos expression in the midbrain periaqueductal grey after trigeminovascular stimulation. J. Anat. 198, 29–35 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Lambert, G. A., Hoskin, K. L. & Zagami, A. S. Cortico-NRM influences on trigeminal neuronal sensation. Cephalalgia 28, 640–652 (2008).

    Article  CAS  PubMed  Google Scholar 

  33. Edelmayer, R. M. et al. Medullary pain facilitating neurons mediate allodynia in headache-related pain. Ann. Neurol. 65, 184–193 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Burstein, R. & Jakubowski, M. Unitary hypothesis for multiple triggers of the pain and strain of migraine. J. Comp. Neurol. 493, 9–14 (2005).

    Article  PubMed  Google Scholar 

  35. Malick, A. & Burstein, R. Cells of origin of the trigeminohypothalamic tract in the rat. J. Comp. Neurol. 400, 125–144 (1998).

    Article  CAS  PubMed  Google Scholar 

  36. Malick, A., Strassman, R. M. & Burstein, R. Trigeminohypothalamic and reticulohypothalamic tract neurons in the upper cervical spinal cord and caudal medulla of the rat. J. Neurophysiol. 84, 2078–2112 (2000).

    Article  CAS  PubMed  Google Scholar 

  37. Burstein, R., Cliffer, K. D. & Giesler, G. J. Jr. Direct somatosensory projections from the spinal cord to the hypothalamus and telencephalon. J. Neurosci. 7, 4159–4164 (1987).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Burstein, R., Cliffer, K. D. & Giesler, G. J. Jr. Cells of origin of the spinohypothalamic tract in the rat. J. Comp. Neurol. 291, 329–344 (1990).

    Article  CAS  PubMed  Google Scholar 

  39. Benjamin, L. et al. Hypothalamic activation after stimulation of the superior sagittal sinus in the cat: a Fos study. Neurobiol. Dis. 16, 500–505 (2004).

    Article  CAS  PubMed  Google Scholar 

  40. Malick, A., Jakubowski, M., Elmquist, J. K., Saper, C. B. & Burstein, R. A neurohistochemical blueprint for pain-induced loss of appetite. Proc. Natl Acad. Sci. USA 98, 9930–9935 (2001). This study shows that sensitization of trigeminovascular neurons disrupts feeding in rodents, possibly through activation of hypothalamic neurons. It demonstrates a mechanism by which migraine in patients may alter food intake.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Matsushita, M., Ikeda, M. & Okado, N. The cells of origin of the trigeminothalamic, trigeminospinal and trigeminocerebellar projections in the cat. Neuroscience 7, 1439–1454 (1982).

    Article  CAS  PubMed  Google Scholar 

  42. Shigenaga, Y. et al. The cells of origin of cat trigeminothalamic projections: especially in the caudal medulla. Brain Res. 277, 201–222 (1983).

    Article  CAS  PubMed  Google Scholar 

  43. Williams, M. N., Zahm, D. S. & Jacquin, M. F. Differential foci and synaptic organization of the principal and spinal trigeminal projections to the thalamus in the rat. Eur. J. Neurosci. 6, 429–453 (1994).

    Article  CAS  PubMed  Google Scholar 

  44. Veinante, P., Jacquin, M. F. & Deschenes, M. Thalamic projections from the whisker-sensitive regions of the spinal trigeminal complex in the rat. J. Comp. Neurol. 420, 233–243 (2000).

    Article  CAS  PubMed  Google Scholar 

  45. Zagami, A. S. & Lambert, G. A. Stimulation of cranial vessels excites nociceptive neurones in several thalamic nuclei of the cat. Exp. Brain Res. 81, 552–566 (1990).

    Article  CAS  PubMed  Google Scholar 

  46. Zagami, A. S. & Lambert, G. A. Craniovascular application of capsaicin activates nociceptive thalamic neurones in the cat. Neurosci. Lett. 121, 187–190 (1991).

    Article  CAS  PubMed  Google Scholar 

  47. Burstein, R. et al. Thalamic sensitization transforms localized pain into widespread allodynia. Ann. Neurol. 68, 81–91 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Jasmin, L., Burkey, A. R., Card, J. P. & Basbaum, A. I. Transneuronal labeling of a nociceptive pathway, the spino-(trigemino-)parabrachio-amygdaloid, in the rat. J. Neurosci. 17, 3751–3765 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Kunkler, P. E. & Kraig, R. P. Hippocampal spreading depression bilaterally activates the caudal trigeminal nucleus in rodents. Hippocampus 13, 835–844 (2003).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Tracey, I. Imaging pain. Br. J. Anaesth. 101, 32–39 (2008).

    Article  CAS  PubMed  Google Scholar 

  51. Derbyshire, S. W. et al. Pain processing during three levels of noxious stimulation produces differential patterns of central activity. Pain 73, 431–445 (1997).

    Article  CAS  PubMed  Google Scholar 

  52. Burstein, R. & Jakubowski, M. Neural substrate of depression during migraine. Neurol. Sci. 30, S27–S31 (2009).

    Article  PubMed  Google Scholar 

  53. Afridi, S. K. et al. A PET study exploring the laterality of brainstem activation in migraine using glyceryl trinitrate. Brain 128, 932–939 (2005).

    Article  CAS  PubMed  Google Scholar 

  54. Bahra, A., Matharu, M. S., Buchel, C., Frackowiak, R. S. & Goadsby, P. J. Brainstem activation specific to migraine headache. Lancet 357, 1016–1017 (2001).

    Article  CAS  PubMed  Google Scholar 

  55. May, A., Bahra, A., Buchel, C., Frackowiak, R. S. & Goadsby, P. J. Hypothalamic activation in cluster headache attacks. Lancet 352, 275–278 (1998).

    Article  CAS  PubMed  Google Scholar 

  56. Weiller, C. et al. Brain stem activation in spontaneous human migraine attacks. Nature Med. 1, 658–660 (1995). This was the first demonstration of brainstem activation during a spontaneous migraine attack, and led to the hypothesis of brainstem dysfunction in migraine.

    Article  CAS  PubMed  Google Scholar 

  57. Sprenger, T. & Goadsby, P. J. What has functional neuroimaging done for primary headache.and for the clinical neurologist? J. Clin. Neurosci. 17, 547–553 (2010).

    Article  PubMed  Google Scholar 

  58. Stephan, K. E. & Friston, K. J. Analyzing effective connectivity with fMRI. Wiley Interdiscip. Rev. Cogn. Sci. 1, 446–459 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  59. Sprenger, T. et al. Abnormal interictal large-scale brain network connectivity in episodic migraine. Headache 50, 71 (2010).

    Article  Google Scholar 

  60. Mainero, C., Boshyan, J. & Hadjikhani, N. Altered functional MRI resting-state connectivity in periaqueductal gray networks in migraine. Ann. Neurol. 11 Jul 2011 (doi:10.1002/ana.22537).

    Article  PubMed  PubMed Central  Google Scholar 

  61. Percheron, G. Thalamus. in The Human Nervous System (eds Paxinos, G. & May, J.) 592–675 (Elsevier, Amsterdam, 2003).

    Google Scholar 

  62. Denuelle, M., Fabre, N., Payoux, P., Chollet, F. & Geraud, G. Hypothalamic activation in spontaneous migraine attacks. Headache 47, 1418–1426 (2007). Together with reference 55, this was the first demonstration of hypothalamic activation during cluster headache and migraine, respectively. The studies put forward the argument that central changes may be responsible for the pain, vascular and autonomic symptoms that occur in cluster headache and migraine, and emphasize the importance of the brain in these headache disorders

    PubMed  Google Scholar 

  63. Afridi, S. K. et al. A positron emission tomographic study in spontaneous migraine. Arch. Neurol. 62, 1270–1275 (2005).

    Article  PubMed  Google Scholar 

  64. Fields, H. L. Is there is a facilitating component to central modulation of pain? J. Pain 1, 71–78 (1992).

    Google Scholar 

  65. Young, R. F. & Brechner, T. Electrical stimulation of the brain for relief of intractable pain due to cancer. Cancer 57, 1266–1272 (1986).

    Article  CAS  PubMed  Google Scholar 

  66. Mayer, D. J. Analgesia produced by electrical stimulation of the brain. Prog. Neuropsychopharmacol. Biol. Psychiatry 8, 557–564 (1984).

    Article  CAS  PubMed  Google Scholar 

  67. Sandkuhler, J. & Gebhart, G. F. Relative contributions of the nucleus raphe magnus and adjacent medullary reticular formation to the inhibition by stimulation in the periaqueductal gray of a spinal nociceptive reflex in the pentobarbital-anesthetized rat. Brain Res. 305, 77–87 (1984).

    Article  CAS  PubMed  Google Scholar 

  68. Fields, H. L., Basbaum, A. I., Clanton, C. H. & Anderson, S. D. Nucleus raphe magnus inhibition of spinal cord dorsal horn neurons. Brain Res. 126, 441–453 (1977).

    Article  CAS  PubMed  Google Scholar 

  69. Porreca, F., Ossipov, M. H. & Gebhart, G. F. Chronic pain and medullary descending facilitation. Trends Neurosci. 25, 319–325 (2002).

    Article  CAS  PubMed  Google Scholar 

  70. Bartsch, T., Knight, Y. E. & Goadsby, P. J. Activation of 5-HT(1B/1D) receptor in the periaqueductal gray inhibits nociception. Ann. Neurol. 56, 371–381 (2004). The authors showed that triptans that are targeted to the vlPAG can attenuate dural-evoked nociceptive trigeminovascular neurons — the first clear indication that triptans can act in areas other than the TNC to exert their therapeutic affects.

    Article  CAS  PubMed  Google Scholar 

  71. Spencer, S. E., Sawyer, W. B., Wada, H., Platt, K. B. & Loewy, A. D. CNS projections to the pterygopalatine parasympathetic preganglionic neurons in the rat: a retrograde transneuronal viral cell body labeling study. Brain Res. 534, 149–169 (1990).

    Article  CAS  PubMed  Google Scholar 

  72. Lai, T.-H., Fuh, J.-L. & Wang, S.-J. Cranial autonomic symptoms in migraine: characteristics and comparison with cluster headache. J. Neurol. Neurosurg. Psychiatry 80, 1116–1119 (2009).

    Article  PubMed  Google Scholar 

  73. Goadsby, P. J. Pathophysiology of cluster headache: a trigeminal autonomic cephalalgia. Lancet Neurol. 1, 251–257 (2002).

    Article  PubMed  Google Scholar 

  74. Akerman, S., Holland, P. R., Lasalandra, M. P. & Goadsby, P. J. Oxygen inhibits neuronal activation in the trigeminocervical complex after stimulation of trigeminal autonomic reflex, but not during direct dural activation of trigeminal afferents. Headache 49, 1131–1143 (2009). This is the first study to show a mechanism of action for the therapeutic effects of oxygen in cluster headache, and a lack of therapeutic effect in migraine. There is also a clear demonstration that SuS stimulation can cause activation in the medullary and cervical dorsal horn.

    Article  PubMed  Google Scholar 

  75. Okamoto, K., Tashiro, A., Chang, Z. & Bereiter, D. A. Bright light activates a trigeminal nociceptive pathway. Pain 149, 235–242 (2010).

    Article  PubMed  PubMed Central  Google Scholar 

  76. Goadsby, P. J. Sphenopalatine ganglion stimulation increases regional cerebral blood flow independent of glucose utilization in the cat. Brain Res. 506, 145–148 (1990).

    Article  CAS  PubMed  Google Scholar 

  77. Goadsby, P. J. Effect of stimulation of facial nerve on regional cerebral blood flow and glucose utilization in cats. Am. J. Physiol. 257, R517–R521 (1989).

    CAS  PubMed  Google Scholar 

  78. Goadsby, P. J. & MacDonald, G. J. Extracranial vasodilation mediated by vasoactive intestinal polypeptide (VIP). Brain Res. 329, 285–288 (1985).

    Article  CAS  PubMed  Google Scholar 

  79. Goadsby, P. J. & Edvinsson, L. Human in vivo evidence for trigeminovascular activation in cluster headache. Neuropeptide changes and effects of acute attacks therapies. Brain 117, 427–434 (1994).

    Article  PubMed  Google Scholar 

  80. Goadsby, P. J. & Edvinsson, L. Neuropeptide changes in a case of chronic paroxysmal hemicrania-evidence for trigemino-parasympathetic activation. Cephalalgia 16, 448–450 (1996).

    Article  CAS  PubMed  Google Scholar 

  81. Hosoya, Y., Matsushita, M. & Sugiura, Y. A direct hypothalamic projection to the superior salivatory nucleus neurons in the rat. A study using anterograde autoradiographic and retrograde HRP methods. Brain Res. 266, 329–333 (1983).

    Article  CAS  PubMed  Google Scholar 

  82. Hosoya, Y., Sugiura, Y., Ito, R. & Kohno, K. Descending projections from the hypothalamic paraventricular nucleus to the A5 area, including the superior salivatory nucleus, in the rat. Exp. Brain Res. 82, 513–518 (1990).

    Article  CAS  PubMed  Google Scholar 

  83. Knight, Y. E. & Goadsby, P. J. The periaqueductal grey matter modulates trigeminovascular input: a role in migraine? Neuroscience 106, 793–800 (2001).

    Article  CAS  PubMed  Google Scholar 

  84. Knight, Y. E., Bartsch, T. & Goadsby, P. J. Trigeminal antinociception induced by bicuculline in the periaqueductal gray (PAG) is not affected by PAG P/Q-type calcium channel blockade in rat. Neurosci. Lett. 336, 113–116 (2003).

    Article  CAS  PubMed  Google Scholar 

  85. Knight, Y. E., Bartsch, T., Kaube, H. & Goadsby, P. J. P/Q-type calcium-channel blockade in the periaqueductal gray facilitates trigeminal nociception: a functional genetic link for migraine? J. Neurosci. 22, RC213 (2002). References 83–85 were the first studies to demonstrate that manipulation of P/Q-type calcium channels in the vlPAG have a descending modulatory effect on dural-evoked nociceptive trigeminovascular neurons — with implications for brainstem involvement in pathophysiology of migraine and familial hemiplegic migraine.

  86. Ophoff, R. A. et al. Familial hemiplegic migraine and episodic ataxia type-2 are caused by mutations in the Ca2+ channel gene CACNL1A4. Cell 87, 543–552 (1996).

    Article  CAS  PubMed  Google Scholar 

  87. Basbaum, A. I., Clanton, C. H. & Fields, H. L. Three bulbospinal pathways from the rostral medulla of the cat: an autoradiographic study of pain modulating systems. J. Comp. Neurol. 178, 209–224 (1978).

    Article  CAS  PubMed  Google Scholar 

  88. Holstege, G. & Kuypers, H. G. The anatomy of brain stem pathways to the spinal cord in cat. A labeled amino acid tracing study. Prog. Brain Res. 57, 145–175 (1982).

    Article  CAS  PubMed  Google Scholar 

  89. Fields, H. L., Vanegas, H., Hentall, I. D. & Zorman, G. Evidence that disinhibition of brain stem neurones contributes to morphine analgesia. Nature 306, 684–686 (1983).

    Article  CAS  PubMed  Google Scholar 

  90. Mason, P. Deconstructing endogenous pain modulations. J. Neurophysiol. 94, 1659–1663 (2005).

    Article  CAS  PubMed  Google Scholar 

  91. Fields, H. L., Bry, J., Hentall, I. & Zorman, G. The activity of neurons in the rostral medulla of the rat during withdrawal from noxious heat. J. Neurosci. 3, 2545–2552 (1983).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Gao, K. & Mason, P. Physiological and anatomic evidence for functional subclasses of serotonergic raphe magnus cells. J. Comp. Neurol. 439, 426–439 (2001).

    Article  CAS  PubMed  Google Scholar 

  93. Gao, K. & Mason, P. Serotonergic Raphe magnus cells that respond to noxious tail heat are not ON or OFF cells. J. Neurophysiol. 84, 1719–1725 (2000).

    Article  CAS  PubMed  Google Scholar 

  94. Mason, P. Physiological identification of pontomedullary serotonergic neurons in the rat. J. Neurophysiol. 77, 1087–1098 (1997).

    Article  CAS  PubMed  Google Scholar 

  95. Fields, H. L., Heinricher, M. M. & Mason, P. Neurotransmitters in nociceptive modulatory circuits. Annu. Rev. Neurosci. 14, 219–245 (1991).

    Article  CAS  PubMed  Google Scholar 

  96. Fields, H. L. & Heinricher, M. M. Anatomy and physiology of a nociceptive modulatory system. Philos. Trans. R. Soc. Lond. B 308, 361–374 (1985).

    Article  CAS  Google Scholar 

  97. Ellrich, J., Ulucan, C. & Schnell, C. Are 'neutral cells' in the rostral ventro-medial medulla subtypes of on- and off-cells? Neurosci. Res. 38, 419–423 (2000).

    Article  CAS  PubMed  Google Scholar 

  98. Mason, P. Ventromedial medulla: pain modulation and beyond. J. Comp. Neurol. 493, 2–8 (2005).

    Article  PubMed  Google Scholar 

  99. Leung, C. G. & Mason, P. Physiological properties of raphe magnus neurons during sleep and waking. J. Neurophysiol. 81, 584–595 (1999).

    Article  CAS  PubMed  Google Scholar 

  100. Foo, H. & Mason, P. Movement-related discharge of ventromedial medullary neurons. J. Neurophysiol. 93, 873–883 (2005).

    Article  CAS  PubMed  Google Scholar 

  101. Foo, H. & Mason, P. Discharge of raphe magnus ON and OFF cells is predictive of the motor facilitation evoked by repeated laser stimulation. J. Neurosci. 23, 1933–1940 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Foo, H. & Mason, P. Sensory suppression during feeding. Proc. Natl Acad. Sci. USA 102, 16865–16869 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Baez, M. A., Brink, T. S. & Mason, P. Roles for pain modulatory cells during micturition and continence. J. Neurosci. 25, 384–394 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Foo, H. & Mason, P. Brainstem modulation of pain during sleep and waking. Sleep Med. Rev. 7, 145–154 (2003).

    Article  CAS  PubMed  Google Scholar 

  105. Foo, H., Crabtree, K. & Mason, P. The modulatory effects of rostral ventromedial medulla on air-puff evoked microarousals in rats. Behav. Brain Res. 215, 156–159 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Vera-Portocarrero, L. P., Ossipov, M. H., King, T. & Porreca, F. Reversal of inflammatory and noninflammatory visceral pain by central or peripheral actions of sumatriptan. Gastroenterology 135, 1369–1378 (2008).

    Article  CAS  PubMed  Google Scholar 

  107. Holland, P. R., Akerman, S., Lasalandra, M. P. & Goadsby, P. J. Antinociceptive effects of orexin A in the vlPAG are blocked by 5-HT1B/1D receptor antagonism. Headache 48, S1–S67 (2008).

    Article  Google Scholar 

  108. Akerman, S., Holland, P. R., Lasalandra, M. P. & Goadsby, P. J. Activation of the CB1 receptor in the vlPAG inhibits nociceptive dural inputs in the TCC via neurons that are modulated by the 5-HT1B/1D receptor. Headache 48, S1–S67 (2008).

    Article  Google Scholar 

  109. Bigal, M. E. et al. Acute migraine medications and evolution from episodic to chronic migraine: a longitudinal population-based study. Headache 48, 1157–1168 (2008).

    Article  PubMed  Google Scholar 

  110. Hilairet, S., Bouaboula, M., Carriere, D., Le Fur, G. & Casellas, P. Hypersensitization of the Orexin 1 receptor by the CB1 receptor: evidence for cross-talk blocked by the specific CB1 antagonist, SR141716. J. Biol. Chem. 278, 23731–23737 (2003).

    Article  CAS  PubMed  Google Scholar 

  111. Ellis, J., Pediani, J. D., Canals, M., Milasta, S. & Milligan, G. Orexin-1 receptor-cannabinoid CB1 receptor heterodimerization results in both ligand-dependent and -independent coordinated alterations of receptor localization and function. J. Biol. Chem. 281, 38812–38824 (2006).

    Article  CAS  PubMed  Google Scholar 

  112. Finn, D. P. et al. Effects of direct periaqueductal grey administration of a cannabinoid receptor agonist on nociceptive and aversive responses in rats. Neuropharmacology 45, 594–604 (2003).

    Article  CAS  PubMed  Google Scholar 

  113. Martin, W. J., Tsou, K. & Walker, J. M. Cannabinoid receptor-mediated inhibition of the rat tail-flick reflex after microinjection into the rostral ventromedial medulla. Neurosci. Lett. 242, 33–36 (1998).

    Article  CAS  PubMed  Google Scholar 

  114. Meng, I. D., Manning, B. H., Martin, W. J. & Fields, H. L. An analgesia circuit activated by cannabinoids. Nature 395, 381–383 (1998).

    Article  CAS  PubMed  Google Scholar 

  115. Meng, I. D. & Johansen, J. P. Antinociception and modulation of rostral ventromedial medulla neuronal activity by local microinfusion of a cannabinoid receptor agonist. Neuroscience 124, 685–693 (2004).

    Article  CAS  PubMed  Google Scholar 

  116. Vaughan, C. W., Connor, M., Bagley, E. E. & Christie, M. J. Actions of cannabinoids on membrane properties and synaptic transmission in rat periaqueductal gray neurons in vitro. Mol. Pharmacol. 57, 288–295 (2000).

    CAS  PubMed  Google Scholar 

  117. Moreau, J. L. & Fields, H. L. Evidence for GABA involvement in midbrain control of medullary neurons that modulate nociceptive transmission. Brain Res. 397, 37–46 (1986).

    Article  CAS  PubMed  Google Scholar 

  118. Goadsby, P. J., Cittadini, E. & Cohen, A. S. Trigeminal autonomic cephalalgias: paroxysmal hemicrania, SUNCT/SUNA and hemicrania continua. Semin. Neurol. 30, 186–191 (2010).

    Article  PubMed  Google Scholar 

  119. Panda, S. & Hogenesch, J. B. It's all in the timing: many clocks, many outputs. J. Biol. Rhythms 19, 374–387 (2004).

    Article  CAS  PubMed  Google Scholar 

  120. Settle, M. The hypothalamus. Neonatal Netw. 19, 9–14 (2000).

    Article  CAS  PubMed  Google Scholar 

  121. Goder, R. et al. Polysomnographic findings in nights preceding a migraine attack. Cephalalgia 21, 31–37 (2001).

    Article  CAS  PubMed  Google Scholar 

  122. Dalkvist, J., Ekbom, K. & Waldenlind, E. Headache and mood: a time-series analysis of self-ratings. Cephalalgia 4, 45–52 (1984).

    Article  CAS  PubMed  Google Scholar 

  123. Giffin, N. J. et al. Premonitory symptoms in migraine: an electronic diary study. Neurology 60, 935–940 (2003).

    Article  CAS  PubMed  Google Scholar 

  124. Fleetwood-Walker, S. M., Hope, P. J. & Mitchell, R. Antinociceptive actions of descending dopaminergic tracts on cat and rat dorsal horn somatosensory neurones. J. Physiol. 399, 335–348 (1988).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  125. Holland, P. & Goadsby, P. J. The hypothalamic orexinergic system: pain and primary headaches. Headache 47, 951–962 (2007).

    Article  PubMed  Google Scholar 

  126. Bartsch, T., Levy, M. J., Knight, Y. E. & Goadsby, P. J. Differential modulation of nociceptive dural input to [hypocretin] orexin A and B receptor activation in the posterior hypothalamic area. Pain 109, 367–378 (2004).

    Article  CAS  PubMed  Google Scholar 

  127. Bartsch, T., Levy, M. J., Knight, Y. E. & Goadsby, P. J. Inhibition of nociceptive dural input in the trigeminal nucleus caudalis by somatostatin receptor blockade in the posterior hypothalamus. Pain 117, 30–39 (2005).

    Article  CAS  PubMed  Google Scholar 

  128. Matharu, M. S., Levy, M. J., Meeran, K. & Goadsby, P. J. Subcutaneous octreotide in cluster headache: randomized placebo-controlled double-blind crossover study. Ann. Neurol. 56, 488–494 (2004).

    Article  CAS  PubMed  Google Scholar 

  129. Levy, M. J., Matharu, M. S., Bhola, R., Meeran, K. & Goadsby, P. J. Octreotide is not effective in the acute treatment of migraine. Cephalalgia 25, 48–55 (2005).

    Article  CAS  PubMed  Google Scholar 

  130. Strassman, A. M., Raymond, S. A. & Burstein, R. Sensitization of meningeal sensory neurons and the origin of headaches. Nature 384, 560–564 (1996).

    Article  CAS  PubMed  Google Scholar 

  131. Dahlstrom, A. & Fuxe, K. Evidence for existence of monoamine-containing neurons in central nervous system. I. Demonstration of monoamines in cell bodies of brain stem neurons. Acta Physiol.Scand, Suppl. 232, 1–55 (1964).

    Google Scholar 

  132. Skagerberg, G., Bjorklund, A., Lindvall, O. & Schmidt, R. H. Origin and termination of the diencephalo-spinal dopamine system in the rat. Brain Res. Bull. 9, 237–244 (1982).

    Article  CAS  PubMed  Google Scholar 

  133. Holstege, J. C. et al. Distribution of dopamine immunoreactivity in the rat, cat and monkey spinal cord. J. Comp. Neurol. 376, 631–652 (1996).

    Article  CAS  PubMed  Google Scholar 

  134. Bergerot, A., Storer, R. J. & Goadsby, P. J. Dopamine inhibits trigeminovascular transmission in the rat. Ann. Neurol. 61, 251–262 (2007).

    Article  CAS  PubMed  Google Scholar 

  135. Charbit, A. R., Akerman, S., Holland, P. R. & Goadsby, P. J. Neurons of the dopaminergic/calcitonin gene-related peptide A11 cell group modulate neuronal firing in the trigeminocervical complex: an electrophysiological and immunohistochemical study. J. Neurosci. 29, 12532–12541 (2009). This was that first study to introduce the A11 dopaminergic nucleus into migraine pathophysiology. Stimulation of the A11 can inhibit dural-evoked nociceptive trigeminovascular neurons, whereas lesioning the nucleus facilitates the dural and facial nociceptive responses. It implicates the A11 in the hypothesis that diencephalic dysfunction could contribute to migraine pathophysiology.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Clemens, S., Rye, D. & Hochman, S. Restless legs syndrome: revisiting the dopamine hypothesis from the spinal cord perspective. Neurology 67, 125–130 (2006).

    Article  PubMed  Google Scholar 

  137. Ondo, W. G., He, Y., Rajasekaran, S. & Le, W. D. Clinical correlates of 6-hydroxydopamine injections into A11 dopaminergic neurons in rats: a possible model for restless legs syndrome. Mov. Disord. 15, 154–158 (2000).

    Article  CAS  PubMed  Google Scholar 

  138. Burstein, R., Yarnitsky, D., Goor-Aryeh, I., Ransil, B. J. & Bajwa, Z. H. An association between migraine and cutaneous allodynia. Ann. Neurol. 47, 614–624 (2000). In references 138 and 47, the authors show that nearly 80% of patients experience cutaneous allodynia during migraine attack. By studying patients during migraine attack and animal models of central sensitization, the authors show that allodynia probably occurs through sensitization of third-order trigeminovascular neurons at the level of the thalamus.

    Article  CAS  PubMed  Google Scholar 

  139. Shields, K. G. & Goadsby, P. J. Propranolol modulates trigeminovascular responses in thalamic ventroposteromedial nucleus: a role in migraine? Brain 128, 86–97 (2005).

    Article  PubMed  Google Scholar 

  140. Andreou, A. P., Shields, K. G. & Goadsby, P. J. GABA and valproate modulate trigeminovascular nociceptive transmission in the thalamus. Neurobiol. Dis. 37, 314–323 (2010).

    Article  CAS  PubMed  Google Scholar 

  141. Summ, O., Charbit, A. R., Andreou, A. P. & Goadsby, P. J. Modulation of nocioceptive transmission with calcitonin gene-related peptide receptor antagonists in the thalamus. Brain 133, 2540–2548 (2010).

    Article  PubMed  Google Scholar 

  142. Shields, K. G. & Goadsby, P. J. Serotonin receptors modulate trigeminovascular responses in ventroposteromedial nucleus of thalamus: a migraine target? Neurobiol. Dis. 23, 491–501 (2006).

    Article  CAS  PubMed  Google Scholar 

  143. Andreou, A. P. & Goadsby, P. J. Therapeutic potential of novel glutamate receptor antagonists in migraine. Expert Opin. Investig. Drugs 18, 789–803 (2009).

    Article  CAS  PubMed  Google Scholar 

  144. Noseda, R. et al. A neural mechanism for exacerbation of headache by light. Nature Neurosci. 13, 239–245 (2010). References 75 and 144 show a neural mechanism for the photophobia that is experienced during migraine. Reference 75 provides a direct link between light intensity and activation of the TNC. Reference 144 shows that dura-sensitive nociceptive trigeminovascular neuronal responses in the posterior thalamus are exacerbated by increasing light intensity, and that they project to the sensory and visual cortices.

    Article  CAS  PubMed  Google Scholar 

  145. Neugebauer, V. & Li, W. Processing of nociceptive mechanical and thermal information in central amygdala neurons with knee-joint input. J. Neurophysiol. 87, 103–112 (2002).

    Article  PubMed  Google Scholar 

  146. Hotopf, M., Mayou, R., Wadsworth, M. & Wessely, S. Temporal relationships between physical symptoms and psychiatric disorder. Results from a national birth cohort. Br. J. Psychiatry 173, 255–261 (1998).

    Article  CAS  PubMed  Google Scholar 

  147. Lazarov, N. E. et al. Amygdalotrigeminal projection in the rat: an anterograde tracing study. Ann. Anat. 193, 118–126 (2011).

    Article  PubMed  Google Scholar 

  148. Dehbandi, S., Speckmann, E. J., Pape, H. C. & Gorji, A. Cortical spreading depression modulates synaptic transmission of the rat lateral amygdala. Eur. J. Neurosci. 27, 2057–2065 (2008).

    Article  PubMed  Google Scholar 

  149. Sink, K. S., Walker, D. L., Yang, Y. & Davis, M. Calcitonin gene-related peptide in the bed nucleus of the stria terminalis produces an anxiety-like pattern of behavior and increases neural activation in anxiety-related structures. J. Neurosci. 31, 1802–1810 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  150. Calandre, E. P., Bembibre, J., Arnedo, M. L. & Becerra, D. Cognitive disturbances and regional cerebral blood flow abnormalities in migraine patients: their relationship with the clinical manifestations of the illness. Cephalalgia 22, 291–302 (2002).

    Article  CAS  PubMed  Google Scholar 

  151. Valfre, W., Rainero, I., Bergui, M. & Pinessi, L. Voxel-based morphometry reveals gray matter abnormalities in migraine. Headache 48, 109–117 (2008).

    Article  PubMed  Google Scholar 

  152. Kim, J. H. et al. Regional grey matter changes in patients with migraine: a voxel-based morphometry study. Cephalalgia 28, 598–604 (2008).

    Article  CAS  PubMed  Google Scholar 

  153. Tessitore, A. et al. Interictal cortical reorganization in episodic migraine without aura: an event-related fMRI study during parametric trigeminal nociceptive stimulation. Neurol. Sci. 32, S165–S167 (2011).

    Article  PubMed  Google Scholar 

  154. Peroutka, S. J. Neurogenic inflammation and migraine: implications for the therapeutics. Mol. Interv. 5, 304–311 (2005).

    Article  CAS  PubMed  Google Scholar 

  155. Moulton, E. A. et al. Interictal dysfunction of a brainstem descending modulatory center in migraine patients. PLoS ONE 3, e3799 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Stankewitz, A., Aderjan, D., Eippert, F. & May, A. Trigeminal nociceptive transmission in migraineurs predicts migraine attacks. J. Neurosci. 31, 1937–1943 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Stankewitz, A. & May, A. The phenomenon of changes in cortical excitability in migraine is not migraine-specific—a unifying thesis. Pain 145, 14–17 (2009).

    Article  PubMed  Google Scholar 

  158. Uddman, R. & Edvinsson, L. Neuropeptides in the cerebral circulation. Cerebrovasc. Brain Metab. Rev. 1, 230–252 (1989).

    CAS  PubMed  Google Scholar 

  159. Goadsby, P. J., Edvinsson, L. & Ekman, R. Release of vasoactive peptides in the extracerebral circulation of humans and the cat during activation of the trigeminovascular system. Ann. Neurol. 23, 193–196 (1988).

    Article  CAS  PubMed  Google Scholar 

  160. Goadsby, P. J. & Duckworth, J. W. Effect of stimulation of trigeminal ganglion on regional cerebral blood flow in cats. Am. J. Physiol. 253, R270–R274 (1987).

    CAS  PubMed  Google Scholar 

  161. Tran-Dinh, Y. R., Thurel, C., Cunin, G., Serrie, A. & Seylaz, J. Cerebral vasodilation after the thermocoagulation of the trigeminal ganglion in humans. Neurosurgery 31, 658–662 (1992).

    CAS  PubMed  Google Scholar 

  162. Goadsby, P. J., Knight, Y. E., Hoskin, K. L. & Butler, P. Stimulation of an intracranial trigeminally-innervated structure selectively increases cerebral blood flow. Brain Res. 751, 247–252 (1997).

    Article  CAS  PubMed  Google Scholar 

  163. Zagami, A. S., Goadsby, P. J. & Edvinsson, L. Stimulation of the superior sagittal sinus in the cat causes release of vasoactive peptides. Neuropeptides 16, 69–75 (1990).

    Article  CAS  PubMed  Google Scholar 

  164. Goadsby, P. J., Edvinsson, L. & Ekman, R. Vasoactive peptide release in the extracerebral circulation of humans during migraine headache. Ann. Neurol. 28, 183–187 (1990).

    Article  CAS  PubMed  Google Scholar 

  165. Gallai, V. et al. Vasoactive peptide levels in the plasma of young migraine patients with and without aura assessed both interictally and ictally. Cephalalgia 15, 384–390 (1995).

    Article  CAS  PubMed  Google Scholar 

  166. Fanciullacci, M., Alessandri, M., Figini, M., Geppetti, P. & Michelacci, S. Increase in plasma calcitonin gene-related peptide from the extracerebral circulation during nitroglycerin-induced cluster headache attack. Pain 60, 119–123 (1995).

    Article  CAS  PubMed  Google Scholar 

  167. Knight, Y. E., Edvinsson, L. & Goadsby, P. J. Blockade of calcitonin gene-related peptide release after superior sagittal sinus stimulation in cat: a comparison of avitriptan and CP122,288. Neuropeptides 33, 41–46 (1999).

    Article  CAS  PubMed  Google Scholar 

  168. Knight, Y. E., Edvinsson, L. & Goadsby, P. J. 4991W93 inhibits release of calcitonin gene-related peptide in the cat but only at doses with 5HT(1B/1D) receptor agonist activity? Neuropharmacology 40, 520–525 (2001).

    Article  CAS  PubMed  Google Scholar 

  169. Roon, K. I. et al. No acute antimigraine efficacy of CP-122288, a highly potent inhibitor of neurogenic inflammation: results of two randomized, double-blind, placebo-controlled clinical trials. Ann. Neurol. 47, 238–241 (2000).

    Article  CAS  PubMed  Google Scholar 

  170. Earl, N. L., McDonald, S. A. & Lowry, M. T. Efficacy and tolerability of the neurogenic inflammation inhibitor, 4991W93 in the acute treatment of migraine. Cephalalgia 19, 357 (1999).

    Google Scholar 

  171. Doods, H. et al. Pharmacological profile of BIBN4096BS, the first selective small molecule CGRP antagonist. Br. J. Pharmacol. 129, 420–423 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Olesen, J. et al. Calcitonin gene-related peptide receptor antagonist BIBN 4096 BS for the acute treatment of migraine. N. Engl. J. Med. 350, 1104–1110 (2004). This proof of concept clinical study demonstrated that CGRP receptor antagonists were efficacious in the acute treatment of migraine without cardiovascular implication, and may represent a new class of compound for patients who suffer from migraine.

    Article  CAS  PubMed  Google Scholar 

  173. Ho, T. W. et al. Randomized controlled trial of an oral CGRP receptor antagonist, MK-0974, in acute treatment of migraine. Neurology 70, 1304–1312 (2008).

    Article  CAS  PubMed  Google Scholar 

  174. Connor, K. M. et al. Randomized, controlled trial of telcagepant for the acute treatment of migraine. Neurology 73, 970–977 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Hewitt, D. J. et al. Randomized controlled trial of the CGRP receptor antagonist, MK-3207, in the acute treatment of migraine. Cephalalgia 31, 712–722 (2011).

    Article  PubMed  Google Scholar 

  176. Diener, H.-C. et al. BI 44370 TA, an oral CGRP antagonist for the acute treatment of migraine attacks: results from a phase II study. Cephalalgia 31, 573–584 (2011).

    Article  PubMed  Google Scholar 

  177. Rasmussen, B. K. & Olesen, J. Migraine with aura and migraine without aura: an epidemiological study. Cephalalgia 12, 221–228 (1992).

    Article  CAS  PubMed  Google Scholar 

  178. Olesen, J. Cerebral and extracranial circulatory disturbances in migraine: pathophysiological implications. Cerebrovasc. Brain Metab. Rev. 3, 1–28 (1991).

    CAS  PubMed  Google Scholar 

  179. Cutrer, F. M. et al. Perfusion-weighted imaging defects during spontaneous migrainous aura. Ann. Neurol. 43, 25–31 (1998).

    Article  CAS  PubMed  Google Scholar 

  180. Hadjikhani, N. et al. Mechanisms of migraine aura revealed by functional MRI in human visual cortex. Proc. Natl Acad. Sci. USA 98, 4687–4692 (2001). The clearest demonstration so far of cortical signal changes that are consistent with CSD during migraine aura.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Sanchez del Rio, M. et al. Perfusion weighted imaging during migraine: spontaneous visual aura and headache. Cephalalgia 19, 701–707 (1999).

    Article  CAS  PubMed  Google Scholar 

  182. Lauritzen, M. Pathophysiology of the migraine aura. The spreading depression theory. Brain 117, 199–210 (1994).

    Article  PubMed  Google Scholar 

  183. Leao, A. A. P. Spreading depression of activity in cerebral cortex. J. Neurophysiol. 7, 359–390 (1944).

    Article  Google Scholar 

  184. Brennan, K. C. et al. Distinct vascular conduction with cortical spreading depression. J. Neurophysiol. 97, 4143–4151 (2007).

    Article  PubMed  Google Scholar 

  185. Bolay, H. et al. Intrinsic brain activity triggers trigeminal meningeal afferents in a migraine model. Nature Med. 8, 136–142 (2002).

    Article  CAS  PubMed  Google Scholar 

  186. Zhang, X. et al. Activation of meningeal nociceptors by cortical spreading depression: implications for migraine with aura. J. Neurosci. 30, 8807–8814 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Zhang, X. et al. Activation of central trigeminovascular neurons by cortical spreading depression. Ann. Neurol. 69, 855–865 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  188. Moskowitz, M. A., Bolay, H. & Dalkara, T. Deciphering migraine mechanisms: clues from familial hemiplegic migraine genotypes. Ann. Neurol. 55, 276–280 (2004).

    Article  CAS  PubMed  Google Scholar 

  189. Haerter, K., Ayata, C. & Moskowitz, M. A. Cortical spreading depression: a model for understanding migraine biology and future drug targets. Headache Curr. 2, 97–103 (2005).

    Article  Google Scholar 

  190. Ebersberger, A., Schaible, H.-G., Averbeck, B. & Richter, F. Is there a correlation between spreading depression, neurogenic inflammation, and nociception that might cause migraine headache? Ann. Neurol. 41, 7–13 (2001).

    Article  Google Scholar 

  191. Lambert, G. A., Michalicek, J., Storer, R. J. & Zagami, A. S. Effect of cortical spreading depression on activity of trigeminovascular sensory neurons. Cephalalgia 19, 631–638 (1999).

    Article  CAS  PubMed  Google Scholar 

  192. Piper, R. D., Lambert, G. A. & Duckworth, J. W. Cortical blood flow changes during spreading depression in cats. Am. J. Physiol. 261, H96–H102 (1991).

    CAS  PubMed  Google Scholar 

  193. Goadsby, P. J., Seylaz, J. & Mraovitch, S. in Migraine and other headaches: the vascular mechanisms (ed. Olesen, J.) 181–6 (Raven Press, New York, 1991)

    Google Scholar 

  194. Yokota, C. et al. Unique profile of spreading depression in a primate model. J. Cereb. Blood Flow Metab. 22, 835–842 (2002).

    Article  PubMed  Google Scholar 

  195. Mayevsky, A. et al. Cortical spreading depression recorded from the human brain using a multiparametric monitoring system. Brain Res. 740, 268–274 (1996).

    Article  CAS  PubMed  Google Scholar 

  196. Strong, A. J. et al. Spreading and synchronous depressions of cortical activity in acutely injured human brain. Stroke 33, 2738–2743 (2002).

    Article  PubMed  Google Scholar 

  197. Fabricius, M. et al. Cortical spreading depression and peri-infarct depolarization in acutely injured human cerebral cortex. Brain 129, 778–790 (2006).

    Article  PubMed  Google Scholar 

  198. Moncada, S., Palmer, R. M. & Higgs, E. A. Nitric oxide: physiology, pathophysiology, and pharmacology. Pharmacol. Rev. 43, 109–142 (1991).

    CAS  PubMed  Google Scholar 

  199. Lassen, L. H. et al. CGRP may play a causative role in migraine. Cephalalgia 22, 54–61 (2002).

    Article  CAS  PubMed  Google Scholar 

  200. Schytz, H. W. et al. PACAP38 induces migraine-like attacks in patients with migraine without aura. Brain 132, 16–25 (2009).

    Article  PubMed  Google Scholar 

  201. Wienecke, T., Olesen, J. & Ashina, M. Prostaglandin I2 (epoprostenol) triggers migraine-like attacks in migraineurs. Cephalalgia 30, 179–190 (2010).

    Article  CAS  PubMed  Google Scholar 

  202. Iversen, H. K., Olesen, J. & Tfelt-hansen, P. Intravenous nitroglycerin as an experimental-model of vascular headache — basic characteristics. Pain 38, 17–24 (1989).

    Article  CAS  PubMed  Google Scholar 

  203. Olesen, J., Iversen, H. K. & Thomsen, L. L. Nitric oxide supersensitivity: a possible molecular mechanism of migraine pain. Neuroreport 4, 1027–1030 (1993).

    Article  CAS  PubMed  Google Scholar 

  204. Olesen, J., Thomsen, L. L. & Iversen, H. Nitric oxide is a key molecule in migraine and other vascular headaches. Trends Pharmacol. Sci. 15, 149–153 (1994).

    Article  CAS  PubMed  Google Scholar 

  205. Olesen, J., Thomsen, L. L., Lassen, L. H. & Olesen, I. J. The nitric oxide hypothesis of migraine and other vascular headaches. Cephalalgia 15, 94–100 (1995).

    Article  CAS  PubMed  Google Scholar 

  206. Asghar, M. S. et al. Dilation by CGRP of middle meningeal artery and reversal by sumatriptan in normal volunteers. Neurology 75, 1520–1526 (2010).

    Article  CAS  PubMed  Google Scholar 

  207. Iversen, H. K. & Olesen, J. Headache induced by a nitric oxide donor (nitroglycerin) responds to sumatriptan. A human model for development of migraine drugs. Cephalalgia 16, 412–418 (1996).

    Article  CAS  PubMed  Google Scholar 

  208. Akerman, S., Williamson, D. J., Kaube, H. & Goadsby, P. J. The effect of anti-migraine compounds on nitric oxide-induced dilation of dural meningeal vessels. Eur. J. Pharmacol. 452, 223–228 (2002).

    Article  CAS  PubMed  Google Scholar 

  209. Humphrey, P. P., Feniuk, W. Mode of action of the anti-migraine drug sumatriptan. Trends Pharmacol. Sci. 12, 444–446 (1991).

    Article  CAS  PubMed  Google Scholar 

  210. Schoonman, G. G. et al. Migraine headache is not associated with cerebral or meningeal vasodilatation-a 3T magnetic resonance angiography study. Brain 131, 2192–2200 (2008). References 206 and 210 are two studies that show the complexity of the arguments over whether vascular changes are a necessary factor in the pain in migraine. Reference 210 was able to demonstrate migraine headache without meningeal or cerebral changes, whereas reference 206 shows migraine headache with substantial meningeal and cerebral vessel dilation.

    Article  CAS  PubMed  Google Scholar 

  211. Asghar, M. S. et al. Evidence for a vascular factor in migraine. Ann. Neurol. 69, 635–645 (2011).

    Article  PubMed  Google Scholar 

  212. Hoskin, K. L., Kaube, H. & Goadsby, P. J. Sumatriptan can inhibit trigeminal afferents by an exclusively neural mechanism. Brain 119, 1419–1428 (1996). This study showed that sumatriptan can have an action that is exclusively neural and not at the level of the vasculature, and was the first indication that drugs without actions at the vasculature may be therapeutic in migraine.

    Article  PubMed  Google Scholar 

  213. Rahmann, A. et al. Vasoactive intestinal peptide causes marked cephalic vasodilation, but does not induce migraine. Cephalalgia 28, 226–236 (2008).

    Article  CAS  PubMed  Google Scholar 

  214. Tassorelli, C. & Joseph, S. A. Systemic nitroglycerin induces fos immunoreactivity in brain-stem and forebrain structures of the rat. Brain Res. 682, 167–181 (1995).

    Article  CAS  PubMed  Google Scholar 

  215. Lambert, G. A., Donaldson, C., Boers, P. M. & Zagami, A. S. Activation of trigeminovascular neurons by glyceryl trinitrate. Brain Res. 887, 203–210 (2000).

    Article  CAS  PubMed  Google Scholar 

  216. Petersen, K. A., Nilsson, E., Olesen, J. & Edvinsson, L. Presence and function of the calcitonin gene-related peptide receptor on rat pial arteries investigated in vitro and in vivo. Cephalalgia 25, 424–432 (2005).

    Article  CAS  PubMed  Google Scholar 

  217. Storer, R. J., Akerman, S. & Goadsby, P. J. Calcitonin gene-related peptide (CGRP) modulates nociceptive trigeminovascular transmission in the cat. Br. J. Pharmacol. 142, 1171–1181 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Cumberbatch, M. J., Williamson, D. J., Mason, G. S., Hill, R. G. & Hargreaves, R. J. Dural vasodilation causes a sensitization of rat caudal trigeminal neurones in vivo that is blocked by a 5-HT1B/1D agonist. Br. J. Pharmacol. 126, 1478–1486 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  219. Paxinos, G. & Watson, C. The Rat Brain In Stereotaxic Coordinates (Elsevier Academic Press, London, 2005).

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Peter J. Goadsby.

Ethics declarations

Competing interests

P.J.G. is a paid member of the boards of the following companies: Allergan, Colucid, MAP Pharmaceuticals, Merck Sharpe and Dohme (MSD), Neuralieve, Neuraxon, ATI (Autonomic Technologies Incorporated), Boston Scientific, Coherex, Eli-Lilly, Medtronic, Linde gases and Bristol Myers Squibb (BMS). In addition, he is a paid consultant for Pfizer and Air Products and has provided paid expert testimony for MedicoLegal advice. The institution at which P.J.G. works receives grant funding from GlaxoSmithKline, MAP, MSD, Neuralieve, Boston Scientific and Amgen. P.J.G. has also received payment for lectures from MSD, Pfizer, Allergan and Mennarini, and has been paid to provide educational slides to the American Headache Society.

Glossary

Aura

Fully reversible neurological symptoms that typically occur before a migraine and that move from one part of a limb, or of the body, to another. These symptoms can include homonymous visual symptoms (that is, flickering lights, spots or lines, or complete loss of vision), sensory symptoms (that is, pins and needles, or numbness) and dysphasic speech.

Brainstem

The most posterior (stem-like) part of the brain, adjoining the cerebral hemispheres. The brainstem is structurally continuous with the spinal cord and comprises the pons, medulla oblongata and midbrain.

Cortical spreading depression

(CSD). A slowly propagating wave (2–6 mm min−1) of sustained strong neuronal depolarization that generates transient and intense spike activity, followed by neural suppression that can last for several minutes.

Trigeminocervical complex

(TCC). A group of spinal cord regions that include the trigeminal nucleus caudalis in the cervicomedullary junction and the superficial layers of the high cervical region dorsal horn at the C1 and C2 level of the spinal cord.

Quintothalamic tract

Also known as the trigeminothalamic tract. The tract that carries sensory information from the head and face to the thalamus through the trigeminal nucleus.

Positron emission tomography

(PET). A nuclear imaging technique that produces three-dimensional images of the brain by detecting photons that are emitted by a positron-emitting radionuclide tracer.

Midbrain

Also known as the mesencephalon. The region that is situated between the pons and the diencephalon, and that includes the periaqueductal grey (PAG).

Hyperaemia

Also known as hyperperfusion. An increase in blood flow.

Oligaemia

Also known as hypoperfusion. A decrease in blood flow.

Diencephalic structures

Structures of the diencephalon, which is the posterior part of the brain. The diencephalon is anterior to the brainstem and includes the hypothalamus, thalamus, metathalamus and epithalamus.

Trigeminal autonomic cephalalgias

(TACs). Primary headache disorders that are characterized by unilateral head pain occurring in association with ipsilateral cranial autonomic features. There are three major pathophysiological features: trigeminal distribution of pain, cranial autonomic features and an episodic pattern of attacks.

Central sensitization

An enhanced response of central neurons by activation of peripheral nociceptors.

Allodynia

The perception of pain from a stimulus that is normally considered innocuous or a stimulus that does not normally produce pain.

Referred pain

In the context of headache, referred pain is the area of the head where spontaneous pain is felt; in a literal sense it is pain in a location other than the site of stimulus. In the case of migraine, the stimulus is likely to be either central or at the very least intracranial, whereas the pain is perceived in the extracranial region.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Akerman, S., Holland, P. & Goadsby, P. Diencephalic and brainstem mechanisms in migraine. Nat Rev Neurosci 12, 570–584 (2011). https://doi.org/10.1038/nrn3057

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nrn3057

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing