1,767
Views
11
CrossRef citations to date
0
Altmetric
Editorial

What can photophobia tell us about dry eye?

, , &
Pages 321-324 | Received 17 May 2016, Accepted 08 Aug 2016, Published online: 23 Aug 2016

1. Photophobia and dry eye

Photophobia, or an abnormal experience of pain to light, is a frequent complaint in patients with sensations of ocular dryness. This editorial will review potential pathways of photophobia and discuss the possibility that its presence is suggestive of central sensitization as a component of dry eye (DE). Potential approaches to the treatment of central sensitization in DE and ways to address photophobia will also be discussed.

2. Frequency and morbidity of DE symptoms

DE is a prevalent condition, with approximately 15% of various populations over 50 years reporting symptoms of visual disturbances and/or sensations of dryness [Citation1]. Alongside these traditional complaints, patients may also report other ocular sensations using descriptors such as ‘aching’ (56%) and ‘hot burning’ (53%) to describe their eye pain [Citation2]. In a population of 236 veterans with mild or greater DE symptoms (dry eye questionnaire 5 score ≥ 6), we found that 75% reported pain sensitivity to light (photophobia); 39% of whom rated the sensitivity to be of moderate or greater severity (≥4 on scale of 0–10). In a similar manner, 64% reported pain sensitivity to wind; 32% of whom rated the sensitivity to be of moderate or greater severity. It is no surprise, then, that DE symptoms, including pain complaints, affect the quality of life of millions of Americans [Citation1]. A study utilizing the Impact of DE on Everyday Life questionnaire found that DE symptoms correlated with difficulties in physical and mental functioning [Citation3]. Utility assessment revealed that patients with severe DE symptoms have utility scores in the same range of conditions as class III/IV angina [Citation4].

3. What can we learn from the coexistence of photophobia in patients with other DE symptoms?

As above, we have found that the majority of patients with symptoms of ocular dryness report some degree of photophobia. However, photophobia is not unique to DE but can be found in a wide number of conditions including intraocular inflammation, iris abnormalities, retinal degenerations, migraines, blepharospasm, and other neurologic pathology. Interestingly, photophobia has also been reported in blind patients [Citation5]. Implied by such variation in causation is a multiplicity of diverse pathways converging on a common pain response.

4. Pathways of photophobia

The trigeminal nerve and its nuclei (the nucleus caudalis in particular [Citation6]) arise as obvious unifying culprits that can tie in photophobia and DE. The V1 distribution of the trigeminal nerve supplies nociceptive innervation to the ocular and orbital structures as well as the meninges. The primary afferents throughout V1 converge onto the trigeminal nucleus caudalis which then conveys sensory information into the central nervous system, projecting nociceptive sensation through the parabrachial and thalamic nuclei and onto the cortex and subcortical centers involved in the experience of pain [Citation7]. Light activates the trigeminal system as can be seen with increased firing and c-fos-positive neurons in the trigeminal nucleus caudalis in a rat model [Citation6].

While the thalamic nuclei mediate the transmission of the sensory component of the painful experience to noxious stimuli or to light in photophobia, pathways involving the parabrachial nuclei mediate the affective/motivational/behavioral components involved in the manifestations of photophobia and DE [Citation8Citation10]. This explains the multidimensional nature of the painful experience in patients with DE and/or photophobia, conditions not infrequently accompanied by high levels of emotional distress [Citation11], as well as the aversive behaviors of patients who protect their eyes with sun glasses, avoid intense light, avoid exposure to wind, and exhibit preferences for dark rooms and isolation.

More than one diverse circuit has been identified in humans and/or animal models that tie in light with the trigeminal–thalamic–cortical pathway [Citation7].

  1. As light is the primary stimulus for photophobia, the classical pathways of light perception are involved. These pathways include rods, cones (as primary light sensors), bipolar and amacrine cells, retinal ganglion cells and then the optic nerve, thalamic geniculate nucleus, and occipital cortex. Yet, some afferent fibers instead of continuing to the lateral geniculate nucleus (as they do for visual input) project to the olivary pretectal nucleus, which activates downstream the superior salivatory nucleus, and the pterygopalatine ganglion. This parasympathetic activation causes activation and sensitization of ocular trigeminal afferents via mechanisms that include uveal vasodilation and release of neuroinflammatory mediators (e.g. calcitonin gene-related peptide, vasoactive intestinal peptide, and nitric oxide). Trigeminal afferents from the eye (heavily expressed in ocular blood vessels) may then convey pain signals to nucleus caudalis, thalamus, and higher brain centers involved in pain perception, and via the parabrachial nucleus to areas driving the affective/aversive and behavioral manifestations of photophobia [Citation7].

    So, light, via the aforementioned pathways, may produce increased activation of the nucleus caudalis and generate subsequent perception of its input as painful. Interestingly, other craniofacial pain conditions also involve active parasympathetic pathways, namely trigeminal autonomic cephalalgias (e.g. cluster headaches). Functional imaging suggests that these syndromes activate a normal human trigeminal-parasympathetic reflex with clinical signs of cranial sympathetic dysfunction being secondary [Citation7].

    The contents of the orbit also receive sympathetic innervation via the short ciliary nerves to the blood vessels and the long ciliary nerves to the pupil. Sympathetic efferents have also been shown to reach the cornea [Citation12], and this may be relevant to pain states too. This is implied from the notion that stimulation of the superior cervical sympathetic ganglion can cause pain in humans [Citation13], while its blockade produces analgesia in patients with intractable facial pain who have failed trigeminal section [Citation13].

  2. Other circuits originate from light sensing but not image-forming pathways, and this may explain the fact that even blind patients may experience photophobia [Citation5]. These originate from intrinsically photosensitive retinal ganglion cells (IPRGCs), with melanopsin rather than rhodopsin as photopigment. Melanopsin responds maximally to light at a wavelength of 480 nm. In humans, only a minority (0.2–0.8%) of ganglion cells are intrinsically photosensitive [Citation14].

    These neurons respond to light (but in a non-image-forming manner) and via the optic nerve project also to the olivary pretectal nuclei, but also to thalamic nuclei bypassing the nucleus caudalis. Therefore, thalamic nuclei receive convergent nociceptive input from ocular primary afferents and from other trigeminal neurons via the nucleus caudalis, as well as direct input from the IPRGC projecting neurons. As a result of this convergence, they may get sensitized, and may convey pain-like signaling in response to light to higher brain centers involved in pain perception [Citation7].

  3. Finally, intrinsically photosensitive ganglion-like cells that contain melanopsin are also present in mammalian iris [Citation15]. So, light sensed by these afferents may activate trigeminal blink reflex even after transection of the optic nerve that contains the pathways from the image-forming and non-image-forming cells.

5. How does this pathway tie in with DE?

DE starts at the ocular surface with findings that include changes in tear composition, tear instability, hyperosmolarity, and inflammation. Corneal nociceptors are sensitive to such changes as their terminal endings are in close contact with the tear film [Citation16]. For example, ocular surface abnormalities such as hyperosmolarity and inflammation have been found to sensitize cold and polymodal afferent receptors, respectively [Citation17]. It is well known that prolonged and intense nociceptive input generates a state of increased excitability and synaptic potentiation in central nociceptive pathways [Citation18]. This phenomenon has complex electrophysiologic, molecular, and genetic determinants and clinically manifests as enhanced spontaneous pain, increased responsiveness to painful stimuli (hyperalgesia), perception of non-painful stimuli as painful (allodynia), distortion of the normal sensory experience, and a variety of several other unpleasant sensations. Central sensitization has been implicated as a significant factor in generating and maintaining other chronic craniofacial pain states, via increased brain excitability and decreased descending inhibitory to pain signaling [Citation19]. Thus, the finding of photophobia in patients with ‘typical’ DE symptoms may suggest that central sensitization underlies at least a portion of their symptoms.

Activation of glia as part of a neuroinflammatory response, triggered by an interplay of noxious injury, genetic, and molecular mechanisms, also significantly contributes to alteration of the processing of sensory signals by the central nervous system so that they are perceived as painful [Citation20]. This is also the case in the trigeminal sensory system [Citation21]. Because by central sensitization, any focal noxious process can evolve into widespread hypersensitivity to non-painful and painful stimuli [Citation22], input of light signaling may generate perception and behaviors suggestive of pain (i.e. photophobia or ‘allodynia to light’).

6. How does this affect DE treatment?

We have found that patients with photophobia have a more chronic disease course and report more severe symptoms than their counterparts with DE symptoms but without photophobia [Citation23]. In addition, in a cross-sectional study, patients with photophobia were more likely to report that their ocular pain was not completely relieved by artificial tears (hypromellose 0.4%) [Citation24]. Despite the limitations of our studies [Citation23,Citation24], taken together, these data suggest that patients with photophobia may have a chronic disease course that is more resistant to topical therapies. In these patients, a multimodal approach may be beneficial, including treating tear abnormalities and ocular surface damage with ocular surface protection and anti-inflammatory agents, and modulating somatosensory dysfunction. More research is needed, however, to understand which drugs, through which route, are best suited to modulate corneal somatosensory function.

Outside the eye, the management of neuropathic pain depends on pain severity, underlying pathophysiology, and systemic comorbidities. Typically, first-line therapies for neuropathic pain are the alpha 2 delta ligand antiepileptics (gabapentin, pregabalin), followed by serotonin–norepinephrine reuptake inhibitors (duloxetine, venlafaxine) as second-line agents and tricyclic antidepressants (nortriptyline, amitriptyline) as third-line agents due to their side effects. Combination therapies (antiepileptics and antidepressants) are also used in cases where monotherapy provides only partial relief. In addition, depending on pain severity and failure of other treatments, some opioids (tramadol, buprenorphine) can be used in selected patients, together with the therapies above. Topical agents (diclofenac, lidocaine, and capsaicin) are also used even as first-line therapies or as parts of multimodal therapies or in specific conditions such as in the treatment of postherpetic neuralgia [Citation25Citation27].

More aggressive measures (e.g. nerve blocks, peripheral and/or central stimulation) are used in patients who have failed conservative therapy or if there are specific indications, such as for neuropathic pain localized in the area of innervation of a specific nerve and related to neuropathy of that nerve. Sympathetic blocks of the superior cervical ganglion, for example, have already been shown to attenuate ocular pain and photophobia [Citation28]. In addition, delivering all these therapies in a multidisciplinary approach (cognitive, behavioral, and physical therapy) is important.

Taking into consideration the affective and aversive manifestations of photophobia, that generate maladaptive behaviors and emotional distress, cognitive behavioral or other appropriate therapy aiming at behavioral modifications should be considered as an integral part of any comprehensive therapeutic approach. For example, in the context of avoidance maladaptive behaviors, addressing photophobia in DE may require some additional considerations. First, the use of sunglasses indoors should be strongly discouraged [Citation14]. Wearing dark glasses indoors causes dark-adaption and only makes light sensitivity worse. Instead, tinted lenses, such as FL-41 (which blocks light at 480 nm; the wavelength at which IPRGCs are maximally sensitive) have been shown to improve light sensitivity in a variety of clinical situations, including migraine and blepharospasm [Citation14].

With many agents available to treat neuropathic pain, research is needed to understand which of these approaches will be beneficial in addressing DE symptoms in those patients who continue to have persistent symptoms despite using currently approved therapies.

Declaration of interest

The authors have no relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript. This includes employment, consultancies, honoraria, stock ownership or options, expert testimony, grants or patents received or pending, or royalties.

Additional information

Funding

This paper was supported by the Department of Veterans Affairs, Veterans Health Administration, Office of Research and Development, Clinical Sciences Research EPID-006-15S (Dr. Galor), NIH Center Core Grant P30EY014801, Research to Prevent Blindness Unrestricted Grant, and NIH NIDCR R01 DE022903 (Dr. Levitt). In addition, we would like to note that the contents of this study do not represent the views of the Department of Veterans Affairs or the United States Government.

References

  • The epidemiology of dry eye disease: report of the Epidemiology Subcommittee of the International Dry Eye WorkShop (2007). Ocul Surf. 2007;5(2):93–107.
  • Kalangara JP, Galor A, Levitt RC, et al. Characteristics of ocular pain complaints in patients with idiopathic dry eye symptoms. Eye Contact Lens. 2016. PMID: 26925537. [Epub ahead of print].
  • Pouyeh B, Viteri E, Feuer W, et al. Impact of ocular surface symptoms on quality of life in a United States veterans affairs population. Am J Ophthalmol. 2012;153(6):1061–1066e1063.
  • Schiffman RM, Walt JG, Jacobsen G, et al. Utility assessment among patients with dry eye disease. Ophthalmology. 2003;110(7):1412–1419.
  • Amini A, Digre K, Couldwell WT. Photophobia in a blind patient: an alternate visual pathway. Case report. J Neurosurg. 2006;105(5):765–768.
  • Okamoto K, Thompson R, Tashiro A, et al. Bright light produces Fos-positive neurons in caudal trigeminal brainstem. Neuroscience. 2009;160(4):858–864.
  • Digre KB, Brennan KC. Shedding light on photophobia. Journal of Neuro-Ophthalmology: the Official Journal of the North American Neuro-Ophthalmology Society. 2012;32(1):68–81.
  • Meng ID, Hu JW, Benetti AP, et al. Encoding of corneal input in two distinct regions of the spinal trigeminal nucleus in the rat: cutaneous receptive field properties, responses to thermal and chemical stimulation, modulation by diffuse noxious inhibitory controls, and projections to the parabrachial area. J Neurophysiol. 1997;77(1):43–56.
  • Aicher SA, Hegarty DM, Hermes SM. Corneal pain activates a trigemino-parabrachial pathway in rats. Brain Research. 2014;1550:18–26.
  • Aicher SA, Hermes SM, Hegarty DM. Corneal afferents differentially target thalamic- and parabrachial-projecting neurons in spinal trigeminal nucleus caudalis. Neuroscience. 2013;232:182–193.
  • Fernandez CA, Galor A, Arheart KL, et al. Dry eye syndrome, posttraumatic stress disorder, and depression in an older male veteran population. Invest Ophthalmol Vis Sci. 2013;54(5):3666–3672.
  • Marfurt CF. Sympathetic innervation of the rat cornea as demonstrated by the retrograde and anterograde transport of horseradish peroxidase-wheat germ agglutinin. J Comp Neurol. 1988;268(2):147–160.
  • Davis L, Pollock L. The role of the sympathetic nervous system in the production of pain in the head. Arch Neurol Psychiatry. 1932;27:282–293.
  • Katz BJ, Digre KB. Diagnosis, pathophysiology and treatment of photophobia. Surv Ophthalmol. 2016;61(4):466–477. doi:10.1016/j.survophthal.2016.02.001.
  • Xue T, Do MT, Riccio A, et al. Melanopsin signalling in mammalian iris and retina. Nature. 2011;479(7371):67–73.
  • Guthoff RF, Wienss H, Hahnel C, et al. Epithelial innervation of human cornea: a three-dimensional study using confocal laser scanning fluorescence microscopy. Cornea. 2005;24(5):608–613.
  • Parra A, Gonzalez-Gonzalez O, Gallar J, et al. Tear fluid hyperosmolality increases nerve impulse activity of cold thermoreceptor endings of the cornea. Pain. 2014;155(8):1481–1491.
  • Woolf CJ. Central sensitization: implications for the diagnosis and treatment of pain. Pain. 2011;152(3 Suppl):S2–15.
  • Noseda R, Burstein R. Migraine pathophysiology: anatomy of the trigeminovascular pathway and associated neurological symptoms, cortical spreading depression, sensitization, and modulation of pain. Pain. 2013;154(Suppl 1):S44–53.
  • Ren K, Dubner R. Neuron-glia crosstalk gets serious: role in pain hypersensitivity. Curr Opin Anaesthesiol. 2008;21(5):570–579.
  • Chiang CY, Wang J, Xie YF, et al. Astroglial glutamate-glutamine shuttle is involved in central sensitization of nociceptive neurons in rat medullary dorsal horn. J Neuroscience Offl J Soc Neurosci. 2007;27(34):9068–9076.
  • Burstein R, Jakubowski M, Garcia-Nicas E, et al. Thalamic sensitization transforms localized pain into widespread allodynia. Ann Neurol. 2010;68(1):81–91.
  • Galor A, Zlotcavitch L, Walter SD, et al. Dry eye symptom severity and persistence are associated with symptoms of neuropathic pain. Br J Ophthalmol. 2015;99(5):665–668.
  • Galor A, Batawi H, Felix ER, et al. Incomplete response to artificial tears is associated with features of neuropathic ocular pain. Br J Ophthalmol. 2016;100(6):745–749.
  • Kremer M, Salvat E, Muller A, et al. Antidepressants and gabapentinoids in neuropathic pain: mechanistic insights. Neuroscience. 2016. doi:10.1016/j.neuroscience.2016.06.057. PMID: 27401055 [Epub ahead of print].
  • Mehta S, McIntyre A, Janzen S, et al. Spinal cord injury rehabilitation evidence team. Systematic review of pharmacologic treatments of pain after spinal cord injury: an update. Arch Phys Med Rehabil. 2016;97(8):1381–1391. doi:10.1016/j.apmr.2015.12.023.
  • Jones RC 3rd, Lawson E, Backonja M. Managing neuropathic pain. Med Clin North Am. 2016;100(1):151–167.
  • Fine PG, Digre KB. A controlled trial of regional sympatholysis in the treatment of photo-oculodynia syndrome. Journal of Neuro-Ophthalmology: the Official Journal of the North American Neuro-Ophthalmology Society. 1995;15(2):90–94.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.