You are on page 1of 12

Pathophysiology 18 (2011) 81–92

Optic neuritis: A mechanistic view


Erik V. Burton a , Benjamin M. Greenberg a , Elliot M. Frohman a,b,∗
a Department of Neurology, University of Texas Southwestern Medical Center at Dallas, United States
b Department of Ophthalmology, University of Texas Southwestern Medical Center at Dallas, United States
Received 28 February 2010; received in revised form 16 March 2010; accepted 8 April 2010

Abstract
Acute demyelinating optic neuritis is a condition of the optic nerves characterized by inflammation, demyelination, and neurodegeneration.
Optic neuritis is a relatively common demyelinating event, strongly associated with multiple sclerosis. A number of clinical, radiographic,
retinal imaging, and electrophysiologic techniques have provided significant insight into the pathologic and pathophysiologic mechanisms of
optic neuritis and its related disorder multiple sclerosis. The development of validated biomarkers within the anterior visual system has paved
the way for novel investigations aimed at characterizing the processes of axonal loss and neurodegeneration, neuroprotection, and perhaps
even neurorestoration strategies.
© 2010 Elsevier Ireland Ltd. All rights reserved.

Keywords: Multiple sclerosis; Visual evoked potentials; Optical coherence tomography pupil light reflex; Optic neuritis; Retinal nerve fiber layer; Macular
volume

1. Introduction ropathy (AION) [9] and Leber’s hereditary optic neuropathy


(LHON).
The optic nerve is a simple, well-defined white-matter Optic neuritis is a frequent “forme fruste” of MS. It is
tract emanating from single-ordered neurons within the reti- estimated that 40% of patients with MS experience ON as
nal macular zone. It is commonly affected in demyelinating their first clinical demyelinating event, while up to 80% of
disorders, such as acute demyelinating optic neuritis (ON), patients with MS experience ON at some point during the
and has well-defined clinical outcome measures. Optic neuri- course of their illness [10,11]. A number of studies includ-
tis is an inflammatory condition of the afferent visual system ing the optic neuritis treatment trial (ONTT) have looked at
that may be associated with a variety of autoimmune or other factors associated with the risk of developing MS in
infectious etiologies [1–6]. When associated with swelling the setting of isolated ON (Table 2) [12–14]. The high asso-
of the optic disc it is referenced as papillitis or anterior optic ciation of ON and MS, coupled with evidence from clinical
neuritis; when the disc appears normal it is referred to as trials demonstrating reduction in risk of progression and dis-
retrobullbar optic neuritis. Optic neuritis is strongly associ- ability among selected patients with early disease modifying
ated with multiple sclerosis (MS), though it may be idiopathic therapy, underscores the importance of early and accurate
or associated with numerous other conditions including neu- diagnosis and management.
romyelitis optica (NMO) [7,8]. There exists a number of As a common demyelinating event strongly associated
conditions which may result in ON (Table 1) and a number of with MS, assessment of the afferent visual system in ON
conditions from which ON must be differentiated clinically has provided significant insights into the pathophysiology of
and pathologically including anterior ischemic optic neu- MS in general. Essentially all of the cardinal pathologic fea-
tures of MS brain and spinal cord pathology are recapitulated
∗ Corresponding author at: Department of Neurology, University of Texas
within the anterior visual system including demyelination,
Southwestern Medical Center at Dallas, 5323 Harry Hines Blvd., Dallas, TX
axonal and neuronal degeneration, and astrogliosis [15–18].
75235, United States. Tel.: +1 214 645 0555; fax: +1 214 645 0556. Here we present a broad review of ON including the clinical
E-mail address: elliot.frohman@utsouthwestern.edu (E.M. Frohman). characteristics and anatomic considerations with particular

0928-4680/$ – see front matter © 2010 Elsevier Ireland Ltd. All rights reserved.
doi:10.1016/j.pathophys.2010.04.009
82 E.V. Burton et al. / Pathophysiology 18 (2011) 81–92

Table 1 ON may also present in chronic and subclinical (occult)


Causes of optic neuritis. forms.
Idiopathic The optic neuritis treatment trial assessed the benefits of
Multiple sclerosis corticosteroid treatment on visual recovery in ON, and the
Neuromyelitis optica (Devic’s)
Viral infections (measles, mumps, varicella)
relationship between ON and MS. Data on 457 patients,
Mycoplasma aged 18–46 years, with ON was collected. Patients typi-
Postviral, postvaccine cally experienced a decline in vision over a 7–10-day period
Mononucleosis with a highly variable degree of visual acuity (VA) loss [12].
Herpes zoster Some patients experienced little change in central vision and
Contiguous inflammation of the meninges, orbits, and sinuses
Granulomatous inflammation (sarcoidosis, Wegener’s,
retained a VA of 20/20 or better while no light perception
tuberculosis, syphilis) vision was present in only 3% of participants at presen-
Intraocular inflammation tation. A progression of vision loss beyond 2 weeks was
unusual in this study, and typical cases exhibited at least
some recovery within 30 days of symptom onset [22,23].
emphasis on the pathology and pathophysiology of ON and A number of additional historic features are strongly asso-
the clinical, electrophysiologic, neuroimaging, and retinal ciated with ON in addition to reduced VA (Table 2). In the
imaging correlates. ONTT, 92% of patients reported concomitant pain, partic-
ularly with eye movements [12]. Flashes of light may be
associated with ON and may be induced by eye movement or
2. Clinical characteristics of optic neuritis certain sounds, a phenomenon referred to as Moore’s light-
ning streaks [24–26].
Optic neuritis is very frequently the initial clinical man- Clinical findings in ON apart from reduced VA com-
ifestation of MS, and is one of the most common clinical monly include a relative afferent papillary defect (RAPD),
events during the course of that illness [10,11]. There is a gen- which is almost invariably present in the affected eye, unless
der predilection for women, who are three times as likely as the fellow optic nerve is affected pathologically such that
men to develop ON. The incidence of ON in persons at high light transmission is slowed (as with demyelination) rela-
risk for developing MS is approximately 3–5/100,000 per tively symmetrically. Central and cecocentral (a central field
year [19–21]. In lower-risk regions the incidence is approx- depression extending into the blind spot) visual field defects
imately one per 100,000 per year [19–21]. While the acute are common in patients with ON. Other variations include
form of ON is most commonly recognized by the clinician, arcuate defects, unilateral nasal and temporal hemianopsias,
quadrant defects, and diffuse field depression (48%) [27].
Table 2
Funduscopic examination was normal in approximately two-
Features of typical and non-typical optic neuritis. thirds of patients with classic ON (hence the adage, ‘the
Typical optic neuritis Non-typical optic neuritis
patient sees nothing and the physician sees nothing’) in the
ONTT, while in the remainder the optic disc appears swollen,
Unilateral (rarely bilateral) Bilateral and sequential
optic neuritis
as in anterior ON (papillitis) [12]. Pallor of the optic disc is not
Young Caucasian women (3F:1M)a Absence of light present acutely after ON, but rather takes weeks to months in
perception order to become manifest, and is explained by the process of
Acute to subacute onset (stabilization Progression of vision loss astrogliosis within the optic nerve head in conjunction with
by 2 weeks) beyond 2 weeks axonal loss [28].
Variable reduced acuity (better than Poor recovery
no light perceptiona )
A number of retinal and clinical features are suggestive of
Reduced color vision Absence of periocular non-typical ON (Table 2) [14]. Patients with atypical features
pain (such as hemorrhage) on presentation had a lower risk of
Recovery begins by day 30 (>90%) Severe papillitis developing MS, particularly when a baseline brain MRI scan
Periocular pain, particularly with eye Peripapillary hemorrhage was normal [13,14].
movement (>90%)a
Moore’s lightning streaks Retinal exudates
Visual recovery occurred rapidly following ON in most
Uhthoff’s phenomena Insidious onset patients in the ONTT and continued for up to a year in
RAPD (if unilateral) the affected eye [29,30]. The mean VA in the affected eye
Central visual field deficit at one year following onset of ON was 20/15, with 90%
No papillitis in 2/3a of patients having VA better than 20/40 [23]. Even fol-
Normal retina (except retinal venous
sheathing)a
lowing severe visual loss (no light perception or finger
Optic disc pallor (4–6 weeks from counting only) 49% of such patients regained VA 20/20
onset) or better. Though marked improvement in high contrast
MRI variable(at least one T2 lesiona ) VA abnormalities was typical, abnormalities on low-contrast
Non-typical features are associated with a low risk of developing MS. sensitivity, color perception, and visual field analyses per-
a Typical features associated with high risk for developing MS.
sisted in approximately 30% of affected eyes and in 14–19%
E.V. Burton et al. / Pathophysiology 18 (2011) 81–92 83

Table 3
Imaging and evoked potential changes reflecting pathophysiologic mechanisms in optic neuritis and their associated clinical changes.
Pathophysiology Imaging and evoked potential changes reflecting pathophysiology Clinical features associated with pathophysiology
Inflammation Gadolinium enhancement Reversible visual deficits
Increased optic nerve thickness Pain
Increased RNFL thickness Retinal venous sheathing
Decreased VEP amplitude and conduction block

Demyelination Decreased MTR Uhthoff’s phenomena


Prolonged VEP latency and conduction block Moore’s lightning streaks
Increased MD Potentially reversible visual deficits
Decreased FA

Neurodegeneration
Axonal loss Decreased optic nerve thickness Fixed visual deficit
Decreased RNFL thickness
Decreased MTR
Atrophy
Increased MD
Decreased FA
Decreased VEP amplitude
Neuronal loss Decreased macular volume Fixed visual deficit
Cortical atrophy
Decreased cortical MTR
Decreased VEP amplitude

Repair
Remyelination Increased MTR Visual recovery
Normalization of VEP latency
Adaptation fMRI changes Improved visual function
Adapted from Kolappan et al. [41].
RNFL: retinal nerve fiber layer; VEP: visual evoked potential; MTR: magnetization transfer ratio; MD: mean diffusivity; FA: fractional anisotropy; fMRI:
functional magnetic resonance imaging.

of contralateral eyes (suggesting occult optic neuropathy) lesion, however, raises the risk of developing clinically def-
[31–34]. inite MS to 72%. The risk of developing MS after 5 years
Additional diagnostic evaluation is generally unnecessary in those patients with normal initial MRI’s remained low but
in the setting of typical ON as other etiologies were unlikely in those with MRI lesions the risk remained significant. Of
in the ONTT [12]. A more aggressive assessment should be those individuals with lesions on an initial MRI, male sex,
considered when non-typical features of ON are present and optic disc swelling, poor vision, and no pain were associated
may include testing for lupus, Lyme disease, sarcoid, syphilis, with a lower risk for development of MS [39].
West Nile virus, human immunodeficiency virus, ehrlichio- Many of the clinical characteristics of ON have partic-
sis, Leber’s optic neuropathy, and neuromyelitis optica [22]. ularly strong correlations with underlying pathologic and
To determine risk of progression to MS, oligoclonal banding pathophysiologic changes (Table 3).
in the CSF was strongly predictive but not independent of
MRI and is of limited utility when the clinical course and
imaging features are typical for demyelinating disease [35]. 3. Anatomy of the afferent visual system
The risk of developing MS following an event of ON
can be estimated reliably, with MRI having been established The retina consists of a number of neuronal components
as the most important predictor. Approximately 50–70% including photoreceptors, bipolar cells, and ganglion cell
of patients with ON will have periventricular white-matter neurons and their axons, which form the retinal nerve fiber
abnormalities consistent with demyelination on an initial layer (RNFL), the most superficial layer of the retina. The
MRI scan [29,36–38]. In 115 patients with clinically iso- RNFL consists of unmyelinated axons that later become
lated ON, 70% of the patients had abnormal brain lesions myelinated after passing posteriorly through the lamina cri-
demonstrated by MRI and 27% had spinal cord lesions [37]. bosa to form the optic nerves. Optic nerve axons from the
According to the ONTT, the cumulative risk of developing nasal hemiretina represent greater than 50% of the optic nerve
MS following optic neuritis was 50% by 15 years. In patients axons and decussate in the chiasm to join the contralateral
with no white-matter lesions the risk of developing MS was axons of the temporal hemiretina to form the optic tract. Most
approximately 25%. The presence of even one white-matter optic tract axons terminate in the lateral geniculate nucleus
84 E.V. Burton et al. / Pathophysiology 18 (2011) 81–92

(LGN). Neurons of the LGN project to the primary visual and fat suppressed fast spin echo sequences (fsFSE), which
cortex via the optic radiations. This three-neuron pathway to reduce T2 hyperintensity of the optic nerve. Combinations
the primary visual cortex forms the retino-geniculo-calcarine of sequences with fast fluid attenuation inversion recovery
pathway for vision. (FLAIR) can reduce CSF signal, and three Tesla scanning
A minority of the optic nerve axons are organized into and surface coils improve sensitivity and spatial resolution
two other principal retinal projections. These include the [47,49,50]. STIR was shown in one study to detect lesions
retino-mesencephalic system, in which axons project via the in 84% of patients with symptomatic ON. Individuals with
brachium of the superior colliculus to innervate the nucleus longer lesions and/or optic canal involvement had slow or
of the optic tract (NOT) for mediation of the pupillary light poor visual recovery [51]. STIR and SPIR sequences are lim-
reflexes, and the retino-hypothalamic tract (innervating the ited by long acquisition times. Nevertheless, combinations
suprachiasmatic nucleus) involved in circadian rhythm regu- of sequences with FLAIR protocols have been shown to be
lation, neuroendocrine reflex arcs, and mood states. superior to simple STIR and SPIR sequences in detecting
optic neuritis [46]. Newer techniques such as DTI, MTI, and
MRS provide additional information on optic nerve structural
4. Imaging and electrophysiology of optic neuritis architecture [52–63].
Retinal imaging provides information about the patho-
Inflammation, demyelination, axonal transection, phyiology of ON and MS and correlates with certain
astrogliosis, and atrophy, along with limited repair and functional, radiographic, and physiologic changes [64–74].
plasticity, are prominent pathologic features of ON and its The retina is unique in that the RNFL is composed of unmyeli-
related condition MS. These pathologic features are the nated axons with an admixture of glia (approximate 80/20%
primary cause of neurologic deficits in ON and MS and composition). There are a number of techniques that are use-
correlate with specific clinical findings [17,40]. Techniques ful in the evaluation of the RNFL along with other retinal
such as visual evoked potentials (VEPs), T1 gadolinium structures, such as OCT, laser polarimetry (GDx), and Heidel-
enhanced and non-enhanced magnetic resonance imaging berg retinal tomography (HRT). With OCT, high-resolution
(MRI), magnetization transfer imaging (MTI), magnetic images of the internal retinal structures are generated by an
resonance spectroscopy (MRS), diffusion tensor imaging optical beam scanned across the retina [65].
(DTI), optical coherence tomography (OCT), and scanning
laser polarimetry (GDx) provide important information
regarding structure and function in ON [41] (Table 3). 5. Pathophysiology of inflammation in optic neuritis
Visual evoked potentials (VEPs) measure cortically gen-
erated potentials via repetitive visual stimulation. Visual The earliest pathologic findings in MS include inflamma-
evoked potentials provide information about electric trans- tion with lymphocytic infiltration of the leptomeninges and
mission of frequency-coded messages along ganglion cell Virchow–Robin spaces [16], as well as perivenular infiltrates
axons within the afferent visual system and are frequently of lymphocytes, plasma cells, and macrophages [75,76].
abnormal following ON [42,43]. Whole field VEP and Analogously, retinal venous sheathing or periphlebitis may
multifocal VEP (mfVEP) have been used clinically and be visualized in ON, and may be associated with increased
experimentally in the evaluation of ON and MS, and correlate risk of conversion to MS [77–79]. The presence of vascular
with a number of pathologic and clinical findings. Electro- inflammation, lesions and other pathologic changes early in
physiologic testing, such as VEP, is of limited clinical utility the retina, a structure free of myelin and oligodendrocytes,
but is often considered useful when there is paucity of data suggests that vascular changes may be a primary event in
to confirm a diagnosis of MS [44]. The mfVEP may be use- demyelinating diseases such as ON and MS.
ful in some selected patients when the distinction between Inflammation is not confined to the retina and optic nerves
optic nerve and retinal disease is in question, or evidence of but may involve any portion of the afferent visual system
subclinical optic nerve dysfunction is sought [45]. including the optic chiasm and optic tracts. Optochiasmatic
Optic nerve MRI has recently been utilized to study the arachnoiditis may be a prominent and early finding in MS,
natural history of ON in vivo. It can potentially differentiate and may be grossly visualized [80]. Involvement of the
among various pathologic components of ON, but ultimately retrochiasmal pathways including the optic tracts and radia-
the specificity is modest at best. MRI of the optic nerve is tions is seen in 13% of idiopathic inflammatory optic neuritis
challenging owing to the small size of this targeted struc- and are often associated with subclinical changes in visual
ture, errors associated with subject mobility, and confounding function [81,82]. In addition to inflammation and lesions
influence of surrounding anatomic structures including fat, within the optic nerves, chiasm, and tracts, the lateral genic-
bone, CSF, and vascular pulsations. A number of techniques ulate nucleus (LGN) and visual cortex are not infrequently
have been developed to overcome these limitations such as fat involved [83,84].
suppression, fast sequencing, 3D acquisition, surface coils, Inflammation culminates in early demyelination. Ini-
and high field MRI [46–50]. These include short-tau inversion tially the outer lamellae of myelin are stripped away by
recovery (STIR), selective partial inversion recovery (SPIR), macrophages in a successive fashion. The result is a ‘bare-
E.V. Burton et al. / Pathophysiology 18 (2011) 81–92 85

Fig. 1. Axial (A) and coronal (B) T1 weighted gadolinium enhanced MR images of acute left optic neuritis (arrows).

naked’ axon that may later be partially remyelinated or, barrier which permits the extravasation of gadolinium. Optic
instead, be invested by proliferating astrocytes (gliosis) nerve enhancement on MRI is associated with impaired visual
[75,76]. It is gliosis in proximity to demyelinated axons that acuity, desaturation of color perception, eye pain, an RAPD,
gives the optic disk its pale white or grey color. and delayed P100 latency and reduced amplitude on VEP test-
A number of clinical findings in the early stage of ON may ing [85,86]. Occasionally, inflammation of the optic chiasm
be associated with inflammation in addition to demyelina- may produce an expanded region of enhancement, whereas
tion, including reduced visual acuity and visual field defects, some patients may have focal leptomeningeal enhancement
along with corresponding changes in electrophysiology and along the chiasm [90–92]. Length of the enhancing portion
imaging metrics [85,86]. Evaluation of the afferent visual of the optic nerve also correlates with the severity of visual
system by these objective methods provides key insights into impairment [93]. Resolution of enhancement occurs over the
the pathophysiologic underpinnings of inflammation and its following month after symptom inception and is associated
impact upon the integrity of visual system function. with improvements in visual systemic metrics including VEP
Cortically generated potentials via visual stimulation in amplitude, and to lesser degree VEP latency. Collectively,
VEPs have been observed to be persistently delayed follow- these findings suggest that acute inflammation is associated
ing optic neuritis, even following visual recovery [42,43]. with blood brain/nerve barrier compromise, and results in
In MS and ON, the VEP P100 latency is characteristically reversible conduction block [85].
delayed, typically with preserved waveform amplitude. How-
ever, in early or severe ON, the P100 amplitude may be
reduced, temporally dispersed (due to axons firing at vary- 6. Pathophysiology of demyelination in optic neuritis
ing velocities secondary to variability in demyelination), or
associated with complete conduction block (severe attenua- Histopathologic features of demyelination within the optic
tion or absent amplitude). Acute reduction in amplitude, as nerves, chiasm, and optic tract parallel those of the brain.
well as complete conduction block, is likely related to active Demyelination occurs early in ON and in association with
inflammation and blood brain/nerve barrier breach, and is acute inflammation with macrophage ingestion of myelin.
associated with corresponding changes in visual function. Demyelination may include the entire cross-section of the
With resolution of the acute stage of ON the VEP ampli- nerve over several centimeters or may be more limited in
tude often normalizes, while the P100 latency often remains distribution. The entire nerve and optic chiasm may show
delayed [85]. Visual field defects identified in optic neuritis features of demyelination and chronically appear structurally
are highly variable and correspond anatomically to lesions smaller than normal (likely the result of myelin and axonal
within the afferent visual system, and have the potential to be loss). Remyelination occurs, albeit sparsely in the acute MS
correlated to abnormalities on mfVEP [87,88]. plaque [94].
Gadolinium enhancement (Fig. 1) is identified in approx- Once a demyelinated plaque forms in ON or MS the lesion
imately 90% of optic nerves during the acute phase of may become one of three types: a chronic inactive plaque,
optic neuritis [89]. Enhancement is a sensitive finding in a chronic active plaque, or a shadow plaque [15,18,94–96].
acute ON, is absent in chronic optic neuropathy (with some Chronic inactive plaques are characterized by complete or
important conspicuous exceptions such as with sarcoid and near-complete absence of myelin and are often found in
neoplasm’s), and is thought to be a vascular phenomenon the optic nerve and chiasm of patients with stable disease
resulting from breach in the integrity of the blood brain/nerve (particularly in those with disease duration of 20–30 years)
86 E.V. Burton et al. / Pathophysiology 18 (2011) 81–92

[15]. Chronic active plaques are associated with ongoing axonal loss, inflammation, and gliosis. Alternately, atrophy
demyelination and inflammation in conjunction with promi- and T1 hypointensities within the CNS strongly correlate
nent axonal transection [17]. When substantial remyelination with axonal and neuronal loss in MS. Newer MRI techniques
occurs, which does not begin until 1 month or more following such as DTI and MTI may improve our ability to sensitively
the onset of demyelination, it becomes known as a shadow and specifically detect changes in tissue architecture within
plaque [95,96]. the brain and optic nerve of MS patients [52–55].
Demyelinated axons have impaired ability to transmit MTI measures the rapid transfer of magnetization between
action potentials due to disruption of saltatory conduction two pools of protons, those bound in macromolecules such
via the nodes of Ranvier. This may result in complete or as myelin and axons, and free protons such as those in the
intermittent conduction block or slowed conduction. As men- extracellular space. Magnetic transfer ratio (MTR) is the mea-
tioned above, these changes have been confirmed by VEP surement of this effect and is an indirect measure of integrity
P100 latency analysis, which were delayed following ON and of macromolecular structure which is in greater concentra-
remained delayed following visual recovery [42,43]. Delayed tion in white versus grey matter [63]. MTR is significantly
latency is strongly associated with demyelination follow- reduced in affected optic nerves and is inversely correlated
ing optic neuritis, while changes in amplitude correlate with with VEP latency, consistent with demyelination [60,61].
inflammation in conjunction with axonal loss [85]. Demyeli- MTR trends upward (generally signifying improved tissue
nation can be highly restricted to a focal area of the visual structure and perhaps remyelination) after reaching a nadir at
system and as such may not be identified by standard VEP approximately 8 months following ON [113,114]. However,
techniques. As such, mfVEP has been established as a more some evidence suggests that MTR may also correlate with
sensitive detection strategy than conventional VEP in patients axonal and neuronal atrophy [61] limiting its specificity.
with MS [97]. DTI has shown dynamic changes in early and remote
Demyelinated axons are thermolabile such that small ON [57]. DTI describes the diffusion properties of water
physiologic changes in core body temperature are associated molecules in tissue along various planes. In geometrically
with changes in nerve conduction properties, most typi- ordered tissue (such as the optic nerve) water motion is
cally characterized by slowing or complete block. This likely facilitated along axons, whereas it is impeded perpendicular
accounts for the so-called Uhthoff’s phenomena (reversible to the main axis of the tract, thereby resulting in direc-
and typically stereotyped and recurrent clinical symptoms) tional selectivity diffusion characteristic. The metric of this
though a number of other physiologic changes have been directional selectivity is calculated as anisotropy, and mea-
implicated (e.g. exercise, infection, and psychologic stress) sured as fractional anisotropy (FA) (again increased along
[98–100]. In 1889 Uhthoff reported on the reversible blur- ordered pathways). The FA is reduced in the context of tis-
ring of vision in MS patients following exercise [101]. The sue damage, which disrupts the organized and directional
most widely accepted physiologic mechanism for this clinical movement of water along the direction of axons (so-called
finding is reversible conduction block in partially demyeli- axial diffusivity). The perpendicular impedence (so-called
nated axons with reduced safety threshold [102] for certain radial diffusivity) to diffusion is reduced with demyelination
metabolic and physiologic changes including small changes (because of less integrity and loss of tissue barriers), which
in temperature [103], blood pH, calcium, sodium, and potas- increases mean diffusivity (MD) (i.e. water can move more
sium [104,105]. A number of studies using VEPs have freely as a consequence of reduced tissue organization). In
confirmed that small elevations in body temperature result in addition to these important correlates of myelin disruption,
impaired conduction in demyelinated axons, and that changes evidence also suggests that there is a relationship between
in conduction are associated with new or worsening visual DTI metrics and axonal loss [62].
deficits, albeit temporarily [106–110]. Likewise active cool-
ing may improve conduction and result in reversal (often
rapidly) of visual deficits [111,112]. 8. Neurodegeneration in optic neuritis
In addition to impaired conduction, demyelinated axons
have abnormally increased (described as ephaptic electric Classically described as an inflammatory demyelinating
discharges) excitability resulting in a number of clinical phe- process with relative sparing of axons, Trapp et al. demon-
nomena in MS and ON, such as the perception phosphenes strated extensive axonal transection as a consistent feature
upon eye movements in dim lighting conditions (Moore’s in MS lesions [17]. We now recognize that axonal injury
lightning streaks) [24–26]. and atrophy represent prominent and early pathologic fea-
tures of MS and ON [18,115,116]. Atrophy is the result of
a combination of demyelination and axonal loss. White mat-
7. Neuroradiologic correlates of optic neuritis ter is composed of about 50% axons, 25% myelin, with glia
and other tissue elements comprising the remainder of this
Changes on conventional MRI studies including T2 hyper- tissue compartment. Given these proportions, axonal degen-
intensities are non-specific and may be associated with a eration would appear to prominently contribute to the process
number of pathologic processes including demyelination, of white-matter atrophy in MS [48].
E.V. Burton et al. / Pathophysiology 18 (2011) 81–92 87

Thinning of the RNFL, which consists of axons devoid of ous pathologic changes including axonal loss and neuronal
myelin, occurs following ON but also in MS patients without degeneration. While MTR correlates well with demyelina-
a history of ON [71], and appears to correlate with cerebral tion in white matter, it was also shown to correlate well
atrophy and disease-related disability in MS [69]. Transec- with thinning of the RNFL, suggesting that MTR is reduced
tion of axons has been demonstrated in acute and chronic by axonal loss as well as by demyelination [61]. MTR has
MS lesions, leading to both Wallerian as well as a dying also been shown to be reduced in the visual cortex of indi-
back degeneration [18]. Axonal densities are reduced in the viduals having had optic neuritis (at least 6 months prior
optic nerves and tracts (30–45% reduction) of SPMS patients to imaging), further supporting trans-synaptic degeneration
[69]. as a significant etiology of neuronal and axonal atrophy
The lateral geniculate nucleus (LGN) and visual cortex are [129].
frequently affected subsequent to ON, secondary to reduced Trip et al. showed that DTI measures of MD were
transmission mechanisms in addition to trans-synaptic degen- increased and FA was decreased in optic nerves affected by
erative changes. In MS patients, there is significant neuronal ON and that these changes correlated with VEP amplitude
cell body atrophy noted in the LGN, particularly within the signifying axonal loss [62]. DTI parameters such as FA of
parvocellular layers, as well as similar changes in the pri- the optic nerve appear to correlate with functional, structural,
mary visual cortices [84]. The pathologic changes in the and physiologic tests of vision [57]. An additional study of
LGN and visual cortex may be a consequence of primary optic radiation specific damage showed that FA and perpen-
demyelination and inflammation, as a consequence of Wal- dicular diffusivity (MD) correlated well with OCT measures
lerian degeneration secondary to lesions of the connecting of retinal injury and visual disability [72]. These findings,
white-matter tracts or as a result of trans-synaptic degen- in addition to MTI changes, underscore the importance of
eration due to lesions in an anatomically linked remote trans-synaptic changes that may occur in MS and ON as a
region [84]. Evangelou et al. investigated the anterior visual source of disability. Mechanisms of trans-synaptic changes
pathways of eight postmortem cases of MS and found that are important for normal development as well as disability
in addition to significant atrophy and axonal loss in the and/or recovery in MS and ON and may represent a target for
optic nerve and tract, there was also a selective atrophy of neuroprotective drugs.
smaller neurons (parvocellular layer of LGN) consistent with
trans-synaptic degeneration. This suggests an increased sus-
ceptibility of smaller axons to MS related injury [117], as 9. Retinal imaging and optic neuritis
has also been suggested by studies of achromatic [118] and
chromatic [119] contrast sensitivity, VEP [120], temporal fre- The retina is unique in that the most superficial layer, the
quency discrimination [121] and pupillary light reflex metrics RNFL, contains axons and glia in the absence of myelin.
[122]. Axonal loss in the RNFL has consistently been demonstrated
Though VEP amplitude can be acutely reduced in early early in MS and following optic neuritis [64,68,73,130]. Reti-
ON, reductions are also observed in chronic ON and correlate nal axonal loss on red-free photography may be seen in eyes
with both a reduction in RNFL thickness (implying axonal with a history of acute optic neuritis (AON) associated with
loss) and macular volume (signifying loss of ganglion cell corresponding visual field changes. Visual confirmation of
neurons) [123]. nerve fiber layer loss during fundoscopy or retinal photog-
MRI measures of brain and spinal cord atrophy corre- raphy requires a reduced thickness of approximately 50% in
late well with disability in MS [124,125]. A study evaluating the RNFL (Fig. 2) [131].
optic nerve atrophy in patients with unilateral optic neuri- Advancements in the application of OCT has facilitated
tis demonstrated significant reduction in cross-sectional area objective analysis of neurodegeneration and axonal loss in
(11.2 mm2 ) of affected eyes compared to fellow (12.9 mm2 ) the anterior visual system [132]. With OCT, high-resolution
and control (12.8 mm2 ) eyes, with the degree of atrophy images of the internal retinal structures are generated by an
correlating to disease duration thereby suggesting ongoing optic beam scanned across the retina (Fig. 3). It is noninvasive
axonal loss following ON [126]. Additional studies have cor- and can sensitively and rapidly measure changes in the retina
related visual dysfunction with the degree of atrophy [127]. that are a consequence of MS and may provide an additional
Other studies have corroborated some of these findings [128]. means of monitoring therapeutic response to neuroprotective
Trip et al. found that optic nerve area was reduced by 30% agents [133,134].
in patients with incomplete recovery following unilateral ON Multiple groups have reported reduction in RNFL thick-
and that RNFL thickness and macular volume reduction cor- ness in eyes with a prior history of ON [64,73,130,135]. The
related significantly with optic atrophy [74]. Visual evoked RNFL thickness is 110–120 ␮m by the age of 15 years. Fol-
potential amplitude correlated with optic nerve atrophy, while lowing ON, about 75% of patients sustain a 10–40 ␮m loss
latency did not, suggesting that optic nerve atrophy on MRI of RNFL thickness that appears to be unaffected by corti-
is a function of axonal loss [123]. costeroid or disease modifying therapy [64]. Measurements
MTI and DTI provide insight into the natural history of RNFL correlate with high- and low-contrast visual acu-
of ON in vivo and can partially differentiate among vari- ity (more sensitively with the latter), visual field changes,
88 E.V. Burton et al. / Pathophysiology 18 (2011) 81–92

Fig. 2. Fundus appearance of eyes following left optic neuritis in a subject with multiple sclerosis. To the left side of the illustrations we see the normal retina
and optic disc of the unaffected right eye and the corresponding nerve fiber layer pattern corresponding to an average retinal nerve fiber layer (RNFL) thickness
of 106.4 ␮m. To the right side of the illustration we see the affected eye with optic disc pallor most prominent to the temporal segment and the corresponding
nerve fiber layer pattern of thinning corresponding to an average RNFL thickness of 60.9 ␮m with arrows indicating the thinning associated with the temporal
pallor. These changes had associated abnormalities in visual function including decreased high contrast visual acuity and low-contrast visual acuity.

brain atrophy, and subtype of MS [68]. Reductions in 10. Recovery and plasticity in optic neuritis
RNFL thickness and macular volume are significantly cor-
related with reductions in visual function [64,66,68,73] as Recovery has been demonstrated following ON in many
well as measures of disability (EDSS and MSSS) in MS clinical parameters including high contrast VA. Remyelina-
[69–70,136]. tion is a reparative process in MS whereby myelin is laid down
RNFL thinning in eyes not previously affected by an attack around demyelinated axons. When effective, axonal con-
of ON has been reported in RRMS, SPMS and PPMS. This duction and neurologic function is restored. Unfortunately
is least evident in the RRMS and most prominent in SPMS the process is incomplete and variable with the majority of
[66,137]. lesions failing to have functionally significant remyelination.

Fig. 3. High-resolution OCT image of normal human macula. NFL, nerve fiber layer; IPL, inner plexiform layer; OPL, outer plexiform layer; RPE, retinal
pigment epithelium; ONL, outer nuclear layer; INL, inner nuclear layer; IS, inner segment; OS, outer segment; ELM, external limiting membrane; GCL,
ganglion cell layer.
E.V. Burton et al. / Pathophysiology 18 (2011) 81–92 89

Chronic inactive lesions, which are often located in the optic in ON, and correlates with clinical measures of visual func-
nerves and chiasm, are associated with complete or near- tion, evoked potentials, and retinal imaging. The newer MTI
complete absence of myelin and depopulation of oligoden- and DTI techniques have provided insight into the pro-
droglia from the center of the lesion. Chronic inactive lesions cesses of axonal loss and demyelination in ON and MS, and
may have some degree of preservation of demyelinated future application may enhance our ability to render prognos-
axons though there may be a prominent reduction in axonal tic predictions, correlate to certain clinical features such as
density. Early recovery of vision following ON has been cor- Uhthoff’s phenomena, and ultimately help monitor response
related with reduced conduction block and corresponding to novel treatments. Future imaging techniques including
improvement in VEP amplitude, however, latency typically higher magnetic field strength scanners, surface coils, and
remains prolonged. MTR, which correlates inversely with fast acquisition methods, will improve spatial resolution and
VEP latency, may improve after reaching a nadir 8 months reduce noise while MR spectroscopy may provide useful
following ON suggestive of remyelination [113,114]. information regarding cell specific changes in ON.
Functional MRI (fMRI) is a method of measuring brain VEP and mfVEP techniques have provided complimen-
activity that takes advantage of oxygenated and deoxy- tary information regarding the physiologic changes seen in
genated states of hemoglobin, having different MRI signals MRI, while retinal imaging has yielded direct evidence of
where functionally activated brain regions have increased axonal loss via RNFL thickness measures and neuronal loss
oxygen requirements. fMRI has been used extensively and via macular volume, both supporting early neurodegeneration
has been employed to demonstrate cortical plasticity in MS in ON and MS. Future studies correlating physiologic decay
[59]. Numerous studies have demonstrated reduced activa- in action potential propagation in ON eyes with structural
tion within the visual cortex in ON corresponding to reduced changes utilizing MTR, DTI, and OCT would provide further
neuronal input. Neural plasticity has been demonstrated in insight into ON and Uhthoff’ phenomena. Such character-
ON with increased activation in areas outside the visual ization would provide an objective means for quantifying
cortex in addition to correlation with VEP latency, sug- Uhthoff’s phenomena and response to various interventions
gesting an association with demyelination within the optic such as body surface cooling and potassium channel block-
nerve. Reorganization has been localized to the lateral occip- ade.
ital cortex and is associated with visual recovery [56,138]. Ultimately it would appear that ON is a good model
Further evidence for plasticity has been found in the LGN for understanding multiple sclerosis, as it is a very com-
[83]. mon clinical syndrome and the organization of the afferent
Proton MR spectroscopy (MRS) is another method of visual apparatus has been systematically characterized. Opti-
characterizing the pathophysiologic changes in the brain. cal coherence tomography is potentially a useful technique
MRS of the brain has been very helpful in providing cell spe- for monitoring the effects of neuroprotective therapies using
cific measures of neuronal damage. Additional MRS changes RNFL thickness for axonal loss and macular volume for neu-
provide evidence for repair involving non-neuronal brain ronal loss. Magnetization transfer ratio, DTI, and VEP are
cells [139]. sensitive measures of pathologic changes ON and in com-
bination with OCT may be effective for evaluation of repair
such as with remyelination.
11. Conclusions

Optic neuritis can now be sensitively characterized both References


structurally and functionally with the application of new
technologies including MRI of the afferent visual pathways, [1] Optic nerve decompression surgery for nonarteritic anterior ischemic
RNFL thickness and macular volume measurements by OCT, optic neuropathy (NAION) is not effective and may be harmful. The
and evoked potentials (conventional and mfVEP). Notwith- Ischemic Optic Neuropathy Decompression Trial Research Group,
JAMA 273 (1995) 625–632.
standing these important and powerful advancements, ON [2] A.C. Arnold, Evolving management of optic neuritis and multiple
remains a clinical diagnosis. Further, conventional brain MRI sclerosis, Am. J. Ophthalmol. 139 (2005) 1101–1108.
sequences continue to play a significant role in the evalua- [3] L.J. Balcer, S.L. Galetta, Optic neuritis, in: R.E. Rakel, E.T. Bope
tion of isolated ON in determining risk of MS [14,140]. VEP (Eds.), Conn’s Current Therapy, W.B. Saunders, Philadelphia, 2004,
has limited clinical utility (beyond the acumen of the experi- pp. 187–190.
[4] R. Foroozan, L.M. Buono, P.J. Savino, R.C. Sergott, Acute demyeli-
enced neurologist) but is useful in evaluating for occult optic nating optic neuritis, Curr. Opin. Ophthalmol. 13 (2002) 375–380.
neuropathy. Retinal imaging is non-specific but does pro- [5] E.M. Frohman, T.C. Frohman, D.S. Zee, R. McColl, S. Galetta, The
vide evidence of early axonal and neuronal degeneration, and neuro-ophthalmology of multiple sclerosis, Lancet Neurol. 4 (2005)
may reveal further evidence of occult optic neuropathy in the 111–121.
appropriate clinical settings. [6] N.J. Newman, V. Biousse, Hereditary optic neuropathies, Eye (Lond.)
18 (2004) 1144–1160.
The application of optic nerve MRI has been particularly [7] D.M. Wingerchuk, V.A. Lennon, S.J. Pittock, C.F. Lucchinetti, B.G.
useful in understanding the temporal profile of inflamma- Weinshenker, Revised diagnostic criteria for neuromyelitis optica,
tion, demyelination, axonal loss, gliosis, and neuronal loss Neurology 66 (2006) 1485–1489.
90 E.V. Burton et al. / Pathophysiology 18 (2011) 81–92

[8] D.M. Wingerchuk, W.F. Hogancamp, P.C. O’Brien, B.G. Wein- [29] R.W. Beck, P.A. Cleary, Optic neuritis treatment trial. One-year
shenker, The clinical course of neuromyelitis optica (Devic’s follow-up results, Arch. Ophthalmol. 111 (1993) 773–775.
syndrome), Neurology 53 (1999) 1107–1114. [30] R.W. Beck, P.A. Cleary, Recovery from severe visual loss in optic
[9] J.F. Rizzo 3rd, S. Lessell, Optic neuritis and ischemic optic neu- neuritis, Arch. Ophthalmol. 111 (1993) 300.
ropathy. Overlapping clinical profiles, Arch. Ophthalmol. 109 (1991) [31] R.W. Beck, R.L. Gal, M.T. Bhatti, et al., Visual function more than
1668–1672. 10 years after optic neuritis: experience of the optic neuritis treatment
[10] Y. Kuroiwa, H. Shibasaki, Clinical studies of multiple sclerosis in trial, Am. J. Ophthalmol. 137 (2004) 77–83.
Japan. I. A current appraisal of 83 cases, Neurology 23 (1973) [32] G.G. Celesia, D.I. Kaufman, M. Brigell, et al., Optic neuritis: a
609–617. prospective study, Neurology 40 (1990) 919–923.
[11] U. Leibowitz, M. Alter, Optic nerve involvement and diplopia as initial [33] J.L. Keltner, C.A. Johnson, J.O. Spurr, R.W. Beck, Visual field profile
manifestations of multiple sclerosis, Acta Neurol. Scand. 44 (1968) of optic neuritis. One-year follow-up in the Optic Neuritis Treatment
70–80. Trial, Arch. Ophthalmol. 112 (1994) 946–953.
[12] The clinical profile of optic neuritis. Experience of the optic neuritis [34] M.J. Menage, D. Papakostopoulos, J.C. Dean Hart, S. Papakostopou-
treatment trial. Optic Neuritis Study Group, Arch. Ophthalmol. 109 los, Y. Gogolitsyn, The Farnsworth-Munsell 100 hue test in the first
(1991) 1673–1678. episode of demyelinating optic neuritis, Br. J. Ophthalmol. 77 (1993)
[13] The 5-year risk of MS after optic neuritis. Experience of the optic 68–74.
neuritis treatment trial. Optic Neuritis Study Group, Neurology 49 [35] L.A. Rolak, R.W. Beck, D.W. Paty, W.W. Tourtellotte, J.N. Whitaker,
(1997) 1404–1413. R.A. Rudick, Cerebrospinal fluid in acute optic neuritis: experience
[14] R.W. Beck, J.D. Trobe, P.S. Moke, et al., High- and low-risk pro- of the optic neuritis treatment trial, Neurology 46 (1996) 368–372.
files for the development of multiple sclerosis within 10 years after [36] R.W. Beck, J. Arrington, F.R. Murtagh, P.A. Cleary, D.I. Kaufman,
optic neuritis: experience of the optic neuritis treatment trial, Arch. Brain magnetic resonance imaging in acute optic neuritis. Experience
Ophthalmol. 121 (2003) 944–949. of the Optic Neuritis Study Group, Arch. Neurol. 50 (1993) 841–846.
[15] W. Bruck, C. Stadelmann, Inflammation and degeneration in multiple [37] C.M. Dalton, P.A. Brex, K.A. Miszkiel, et al., Spinal cord MRI in
sclerosis, Neurol. Sci. 24 (Suppl. 5) (2003) S265–267. clinically isolated optic neuritis, J. Neurol. Neurosurg. Psychiatry 74
[16] J.H. Noseworthy, C. Lucchinetti, M. Rodriguez, B.G. Weinshenker, (2003) 1577–1580.
Multiple sclerosis, N. Engl. J. Med. 343 (2000) 938–952. [38] L. Jacobs, F.E. Munschauer, S.E. Kaba, Clinical and magnetic reso-
[17] B.D. Trapp, J. Peterson, R.M. Ransohoff, R. Rudick, S. Mork, L. Bo, nance imaging in optic neuritis, Neurology 41 (1991) 15–19.
Axonal transection in the lesions of multiple sclerosis, N. Engl. J. [39] Multiple sclerosis risk after optic neuritis: final optic neuritis treatment
Med. 338 (1998) 278–285. trial follow-up, Arch. Neurol. 65 (2008) 727–732.
[18] B.D. Trapp, R. Ransohoff, R. Rudick, Axonal pathology in multiple [40] P. O’Connor, Key issues in the diagnosis and treatment of multiple
sclerosis: relationship to neurologic disability, Curr. Opin. Neurol. 12 sclerosis. An overview, Neurology 59 (2002) S1–33.
(1999) 295–302. [41] M. Kolappan, A.P. Henderson, T.M. Jenkins, et al., Assessing struc-
[19] M. Wakakura, R. Minei-Higa, S. Oono, et al., Baseline features of ture and function of the afferent visual pathway in multiple sclerosis
idiopathic optic neuritis as determined by a multicenter treatment and associated optic neuritis, J. Neurol. 256 (2009) 305–319.
trial in Japan. Optic Neuritis Treatment Trial Multicenter Cooperative [42] A.M. Halliday, W.I. McDonald, J. Mushin, Delayed visual evoked
Research Group (ONMRG), Jpn. J. Ophthalmol. 43 (1999) 127–132. response in optic neuritis, Lancet 1 (1972) 982–985.
[20] M. Wakakura, K. Mashimo, S. Oono, et al., Multicenter clinical trial [43] A.M. Halliday, W.I. McDonald, J. Mushin, Visual evoked response in
for evaluating methylprednisolone pulse treatment of idiopathic optic diagnosis of multiple sclerosis, Br. Med. J. 4 (1973) 661–664.
neuritis in Japan. Optic Neuritis Treatment Trial Multicenter Coop- [44] G.S. Gronseth, E.I. Ashman, Practice parameter: the usefulness of
erative Research Group (ONMRG), Jpn. J. Ophthalmol. 43 (1999) evoked potentials in identifying clinically silent lesions in patients
133–138. with suspected multiple sclerosis (an evidence-based review): Report
[21] D.I. Kaufman, J.D. Trobe, E.R. Eggenberger, J.N. Whitaker, Practice of the Quality Standards Subcommittee of the American Academy of
parameter: the role of corticosteroids in the management of acute Neurology, Neurology 54 (2000) 1720–1725.
monosymptomatic optic neuritis. Report of the Quality Standards [45] D.C. Hood, J.G. Odel, B.J. Winn, The multifocal visual evoked poten-
Subcommittee of the American Academy of Neurology, Neurology tial, J. Neuroophthalmol. 23 (2003) 279–289.
54 (2000) 2039–2044. [46] A. Jackson, S. Sheppard, R.D. Laitt, A. Kassner, D. Moriarty, Optic
[22] R.W. Beck, The optic neuritis treatment trial. Implications for clinical neuritis: MR imaging with combined fat- and water-suppression tech-
practice. Optic Neuritis Study Group, Arch. Ophthalmol. 110 (1992) niques, Radiology 206 (1998) 57–63.
331–332. [47] S. Karim, R.A. Clark, V. Poukens, J.L. Demer, Demonstration of sys-
[23] R.W. Beck, P.A. Cleary, M.M. Anderson Jr., et al., A randomized, tematic variation in human intraorbital optic nerve size by quantitative
controlled trial of corticosteroids in the treatment of acute optic neu- magnetic resonance imaging and histology, Invest. Ophthalmol. Vis.
ritis. The Optic Neuritis Study Group, N. Engl. J. Med. 326 (1992) Sci. 45 (2004) 1047–1051.
581–588. [48] D.H. Miller, F. Barkhof, J.A. Frank, G.J. Parker, A.J. Thompson,
[24] F.A. Davis, D. Bergen, C. Schauf, I. McDonald, W. Deutsch, Move- Measurement of atrophy in multiple sclerosis: pathological basis,
ment phosphenes in optic neuritis: a new clinical sign, Neurology 26 methodological aspects and clinical relevance, Brain 125 (2002)
(1976) 1100–1104. 1676–1695.
[25] W.I. McDonald, D. Barnes, The ocular manifestations of multiple [49] S. Ueki, Y. Fujii, H. Matsuzawa, et al., Assessment of axonal degener-
sclerosis. 1. Abnormalities of the afferent visual system, J. Neurol. ation along the human visual pathway using diffusion trace analysis,
Neurosurg. Psychiatry 55 (1992) 747–752. Am. J. Ophthalmol. 142 (2006) 591–596.
[26] R.F. Moore, Subjective “Lightning Streaks”, Br. J. Ophthalmol. 19 [50] M. Weigel, W.A. Lagreze, A. Lazzaro, J. Hennig, T.A. Bley, Fast and
(1935) 545–547. quantitative high-resolution magnetic resonance imaging of the optic
[27] J.L. Keltner, C.A. Johnson, J.O. Spurr, R.W. Beck, Baseline visual nerve at 3.0 tesla, Invest. Radiol. 41 (2006) 83–86.
field profile of optic neuritis. The experience of the optic neuritis [51] D.H. Miller, M.R. Newton, J.C. van der Poel, et al., Magnetic res-
treatment trial. Optic Neuritis Study Group, Arch. Ophthalmol. 111 onance imaging of the optic nerve in optic neuritis, Neurology 38
(1993) 231–234. (1988) 175–179.
[28] L.I. Balcer, Clinical practice. Optic neuritis, N. Engl. J. Med. 354 [52] M. Filippi, Magnetization transfer MRI in multiple sclerosis and other
(2006) 1273–1280. central nervous system disorders, Eur. J. Neurol. 10 (2003) 3–10.
E.V. Burton et al. / Pathophysiology 18 (2011) 81–92 91

[53] M. Filippi, M. Bozzali, M. Rovaris, et al., Evidence for widespread that axonal loss is a substrate of MRI-detected atrophy, Neuroimage
axonal damage at the earliest clinical stage of multiple sclerosis, Brain 31 (2006) 286–293.
126 (2003) 433–437. [75] C.W. Adams, The onset and progression of the lesion in multiple
[54] M. Filippi, M.A. Rocca, G. Comi, The use of quantitative magnetic- sclerosis, J. Neurol. Sci. 25 (1975) 165–182.
resonance-based techniques to monitor the evolution of multiple [76] W. Bruck, P. Porada, S. Poser, et al., Monocyte/macrophage differ-
sclerosis, Lancet Neurol. 2 (2003) 337–346. entiation in early multiple sclerosis lesions, Ann. Neurol. 38 (1995)
[55] Y. Ge, R.I. Grossman, J.S. Babb, J. He, L.J. Mannon, Dirty-appearing 788–796.
white matter in multiple sclerosis: volumetric MR imaging and mag- [77] T. Engell, Neurological disease activity in multiple sclerosis patients
netization transfer ratio histogram analysis, Am. J. Neuroradiol. 24 with periphlebitis retinae, Acta Neurol. Scand. 73 (1986) 168–172.
(2003) 1935–1940. [78] T. Engell, B. Krogsaa, H. Lund-Andersen, Breakdown of the
[56] N. Levin, T. Orlov, S. Dotan, E. Zohary, Normal and abnormal fMRI blood-retinal barrier in multiple sclerosis measured by vitreous fluo-
activation patterns in the visual cortex after recovery from optic neu- rophotometry, Acta Ophthalmol. (Copenh.) 64 (1986) 583–587.
ritis, Neuroimage 33 (2006) 1161–1168. [79] S. Lightman, W.I. McDonald, A.C. Bird, et al., Retinal venous sheath-
[57] R.T. Naismith, J. Xu, N.T. Tutlam, et al., Disability in optic neuri- ing in optic neuritis. Its significance for the pathogenesis of multiple
tis correlates with diffusion tensor-derived directional diffusivities, sclerosis, Brain 110 (Pt 2) (1987) 405–414.
Neurology 72 (2009) 589–594. [80] R.A. Bell, D.M. Robertson, D.A. Rosen, A.W. Kerr, Optochias-
[58] M.A. Rocca, F. Agosta, B. Colombo, et al., fMRI changes matic arachnoiditis in multiple sclerosis, Arch. Ophthalmol. 93 (1975)
in relapsing-remitting multiple sclerosis patients complaining of 191–193.
fatigue after IFNbeta-1a injection, Hum. Brain Mapp. 28 (2007) [81] G.T. Plant, A.G. Kermode, G. Turano, et al., Symptomatic retrochi-
373–382. asmal lesions in multiple sclerosis: clinical features, visual evoked
[59] M.A. Rocca, E. Pagani, M. Absinta, et al., Altered functional and potentials, and magnetic resonance imaging, Neurology 42 (1992)
structural connectivities in patients with MS: a 3-T study, Neurology 68–76.
69 (2007) 2136–2145. [82] M.A. Rosenblatt, M.M. Behrens, P.H. Zweifach, et al., Magnetic res-
[60] J.W. Thorpe, G.J. Barker, S.J. Jones, et al., Magnetisation transfer onance imaging of optic tract involvement in multiple sclerosis, Am.
ratios and transverse magnetisation decay curves in optic neuritis: J. Ophthalmol. 104 (1987) 74–79.
correlation with clinical findings and electrophysiology, J. Neurol. [83] K. Korsholm, K.H. Madsen, J.L. Frederiksen, A. Skimminge, T.E.
Neurosurg. Psychiatry 59 (1995) 487–492. Lund, Recovery from optic neuritis: an ROI-based analysis of LGN
[61] S.A. Trip, P.G. Schlottmann, S.J. Jones, et al., Optic nerve magnetiza- and visual cortical areas, Brain 130 (2007) 1244–1253.
tion transfer imaging and measures of axonal loss and demyelination [84] J.H. Simon, R.P. Kinkel, L. Jacobs, L. Bub, N. Simonian, A. Wallerian
in optic neuritis, Mult. Scler. 13 (2007) 875–879. degeneration pattern in patients at risk for MS, Neurology 54 (2000)
[62] S.A. Trip, C. Wheeler-Kingshott, S.J. Jones, et al., Optic nerve diffu- 1155–1160.
sion tensor imaging in optic neuritis, Neuroimage 30 (2006) 498–505. [85] B.D. Youl, G. Turano, D.H. Miller, et al., The pathophysiology
[63] S.D. Wolff, R.S. Balaban, Magnetization transfer contrast (MTC) and of acute optic neuritis. An association of gadolinium leakage with
tissue water proton relaxation in vivo, Magn. Reson. Med. 10 (1989) clinical and electrophysiological deficits, Brain 114 (Pt 6) (1991)
135–144. 2437–2450.
[64] F. Costello, S. Coupland, W. Hodge, et al., Quantifying axonal loss [86] B.D. Youl, G. Turano, A.D. Towell, et al., Optic neuritis: swelling and
after optic neuritis with optical coherence tomography, Ann. Neurol. atrophy, Electroencephalogr. Clin. Neurophysiol. Suppl. 46 (1996)
59 (2006) 963–969. 173–179.
[65] E.M. Frohman, J.G. Fujimoto, T.C. Frohman, P.A. Calabresi, G. Cut- [87] A. Klistorner, H. Arvind, T. Nguyen, et al., Multifocal VEP and
ter, L.J. Balcer, Optical coherence tomography: a window into the OCT in optic neuritis: a topographical study of the structure-function
mechanisms of multiple sclerosis, Nat. Clin. Pract. Neurol. 4 (2008) relationship, Doc. Ophthalmol. 118 (2009) 129–137.
664–675. [88] A. Klistorner, C. Fraser, R. Garrick, S. Graham, H. Arvind, Corre-
[66] A.P. Henderson, S.A. Trip, P.G. Schlottmann, et al., An investigation lation between full-field and multifocal VEPs in optic neuritis, Doc.
of the retinal nerve fibre layer in progressive multiple sclerosis using Ophthalmol. 116 (2008) 19–27.
optical coherence tomography, Brain 131 (2008) 277–287. [89] J.F. Rizzo 3rd, C.M. Andreoli, J.D. Rabinov, Use of magnetic reso-
[67] R.C. Sergott, E. Frohman, R. Glanzman, A. Al-Sabbagh, The role nance imaging to differentiate optic neuritis and nonarteritic anterior
of optical coherence tomography in multiple sclerosis: expert panel ischemic optic neuropathy, Ophthalmology 109 (2002) 1679–1684.
consensus, J. Neurol. Sci. 263 (2007) 3–14. [90] P. Demaerel, W. Robberecht, I. Casteels, et al., Focal leptomeningeal
[68] J.B. Fisher, D.A. Jacobs, C.E. Markowitz, et al., Relation of visual MR enhancement along the chiasm as a presenting sign of multiple
function to retinal nerve fiber layer thickness in multiple sclerosis, sclerosis, J. Comput. Assist. Tomogr. 19 (1995) 297–298.
Ophthalmology 113 (2006) 324–332. [91] E. Kerty, N. Eide, P. Nakstal, R. Nyberg-Hansen, Chiasmal optic
[69] E. Gordon-Lipkin, B. Chodkowski, D.S. Reich, et al., Retinal nerve neuritis, Acta Ophthalmol. (Copenh.) 69 (1991) 135–139.
fiber layer is associated with brain atrophy in multiple sclerosis, Neu- [92] N.J. Newman, S. Lessell, J.M. Winterkorn, Optic chiasmal neuritis,
rology 69 (2007) 1603–1609. Neurology 41 (1991) 1203–1210.
[70] E. Grazioli, R. Zivadinov, B. Weinstock-Guttman, et al., Retinal nerve [93] M.J. Kupersmith, T. Alban, B. Zeiffer, D. Lefton, Contrast-enhanced
fiber layer thickness is associated with brain MRI outcomes in multi- MRI in acute optic neuritis: relationship to visual performance, Brain
ple sclerosis, J. Neurol. Sci. 268 (2008) 12–17. 125 (2002) 812–822.
[71] J.G. Kerrison, T. Flynn, W.R. Green, Retinal pathologic changes in [94] C.S. Raine, E. Wu, Multiple sclerosis: remyelination in acute lesions,
multiple sclerosis, Retina 14 (1994) 445–451. J. Neuropathol. Exp. Neurol. 52 (1993) 199–204.
[72] D.S. Reich, S.A. Smith, E.M. Gordon-Lipkin, et al., Damage to the [95] J.W. Prineas, R.O. Barnard, E.E. Kwon, L.R. Sharer, E.S. Cho, Multi-
optic radiation in multiple sclerosis is associated with retinal injury ple sclerosis: remyelination of nascent lesions, Ann. Neurol. 33 (1993)
and visual disability, Arch. Neurol. 66 (2009) 998–1006. 137–151.
[73] S.A. Trip, P.G. Schlottmann, S.J. Jones, et al., Retinal nerve fiber layer [96] J.W. Prineas, R.O. Barnard, T. Revesz, E.E. Kwon, L. Sharer, E.S.
axonal loss and visual dysfunction in optic neuritis, Ann. Neurol. 58 Cho, Multiple sclerosis. Pathology of recurrent lesions, Brain 116 (Pt
(2005) 383–391. 3) (1993) 681–693.
[74] S.A. Trip, P.G. Schlottmann, S.J. Jones, et al., Optic nerve atrophy and [97] L.K. Grover, D.C. Hood, Q. Ghadiali, et al., A comparison of multi-
retinal nerve fibre layer thinning following optic neuritis: evidence focal and conventional visual evoked potential techniques in patients
92 E.V. Burton et al. / Pathophysiology 18 (2011) 81–92

with optic neuritis/multiple sclerosis, Doc. Ophthalmol. 117 (2008) [121] G.T. Plant, R.F. Hess, Temporal frequency discrimination in optic
121–128. neuritis and multiple sclerosis, Brain 108 (Pt 3) (1985) 647–676.
[98] G.B. Scholl, H.S. Song, S.H. Wray, Uhthoff’s symptom in optic neu- [122] S.I. Moro, M.L. Rodriguez-Carmona, E.C. Frost, G.T. Plant, J.L. Bar-
ritis: relationship to magnetic resonance imaging and development of bur, Recovery of vision and pupil responses in optic neuritis and
multiple sclerosis, Ann. Neurol. 30 (1991) 180–184. multiple sclerosis, Ophthalmic Physiol. Opt. 27 (2007) 451–460.
[99] F.E. Lepore, Uhthoff’s symptom in disorders of the anterior visual [123] A. Klistorner, H. Arvind, T. Nguyen, et al., Axonal loss and myelin
pathways, Neurology 44 (1994) 1036–1038. in early ON loss in postacute optic neuritis, Ann. Neurol. 64 (2008)
[100] J.B. Selhorst, R.F. Saul, Uhthoff and his symptom, J. Neuroophthal- 325–331.
mol. 15 (1995) 63–69. [124] N.A. Losseff, L. Wang, H.M. Lai, et al., Progressive cerebral atrophy
[101] W. Uhthoff, Untersuchungen uber die bei der multiplen Herdsklerose in multiple sclerosis. A serial MRI study, Brain 119 (Pt 6) (1996)
vorkommenden Augenstorungen, Archiv für Psychiatrie und Ner- 2009–2019.
venkrankheiten 20 (1889). [125] N.A. Losseff, S.L. Webb, J.I. O’Riordan, et al., Spinal cord atrophy
[102] F.A. Davis, The multiple sclerosis “hot-bath test”: a putative model and disability in multiple sclerosis. A new reproducible and sensitive
in the intact frog, Mt. Sinai. J. Med. 41 (1974) 93–96. MRI method with potential to monitor disease progression, Brain 119
[103] M. Rasminsky, The effects of temperature on conduction in demyeli- (Pt 3) (1996) 701–708.
nated single nerve fibers, Arch. Neurol. 28 (1973) 287–292. [126] S.J. Hickman, P.A. Brex, C.M. Brierley, et al., Detection of optic
[104] F.A. Davis, Axonal conduction studies based on some considerations nerve atrophy following a single episode of unilateral optic neuri-
of temperature effects in multiple sclerosis, Electroencephalogr. Clin. tis by MRI using a fat-saturated short-echo fast FLAIR sequence,
Neurophysiol. 28 (1970) 281–286. Neuroradiology 43 (2001) 123–128.
[105] F.A. Davis, F.O. Becker, J.A. Michael, E. Sorensen, Effect of [127] S.J. Hickman, C.M. Brierley, P.A. Brex, et al., Continuing optic nerve
intravenous sodium bicarbonate, disodium edetate (Na2EDTA), and atrophy following optic neuritis: a serial MRI study, Mult. Scler. 8
hyperventilation on visual and oculomotor signs in multiple sclerosis, (2002) 339–342.
J. Neurol. Neurosurg. Psychiatry 33 (1970) 723–732. [128] M. Inglese, A. Ghezzi, S. Bianchi, et al., Irreversible disability and
[106] R.F. Saul, G. Hayat, J.B. Selhorst, Visual evoked potentials during tissue loss in multiple sclerosis: a conventional and magnetization
hyperthermia, J. Neuroophthalmol. 15 (1995) 70–78. transfer magnetic resonance imaging study of the optic nerves, Arch.
[107] H.E. Persson, C. Sachs, Visual evoked potentials elicited by pat- Neurol. 59 (2002) 250–255.
tern reversal during provoked visual impairment in multiple sclerosis, [129] B. Audoin, K.T. Fernando, J.K. Swanton, A.J. Thompson, G.T. Plant,
Brain 104 (1981) 369–382. D.H. Miller, Selective magnetization transfer ratio decrease in the
[108] H.E. Persson, C. Sachs, Provoked visual impairment in multiple scle- visual cortex following optic neuritis, Brain 129 (2006) 1031–1039.
rosis studied by visual evoked responses, Electroencephalogr. Clin. [130] V. Parisi, G. Manni, S.A. Gandolfi, M. Centofanti, G. Colacino, M.G.
Neurophysiol. 44 (1978) 664–668. Bucci, Visual function correlates with nerve fiber layer thickness in
[109] W.B. Matthews, D.J. Read, E. Pountney, Effect of raising body tem- eyes affected by ocular hypertension, Invest. Ophthalmol. Vis. Sci. 40
perature on visual and somatosensory evoked potentials in patients (1999) 1828–1833.
with multiple sclerosis, J. Neurol. Neurosurg. Psychiatry 42 (1979) [131] H.A. Quigley, E.M. Addicks, Quantitative studies of retinal nerve
250–255. fiber layer defects, Arch. Ophthalmol. 100 (1982) 807–814.
[110] A. Lossos, S. Dotan, O. Abramsky, Uhthoff’s symptom, Neurology [132] E. Frohman, F. Costello, R. Zivadinov, et al., Optical coherence
45 (1995) 599. tomography in multiple sclerosis, Lancet Neurol. 5 (2006) 853–863.
[111] S.R. Schwid, M.D. Petrie, R. Murray, et al., A randomized controlled [133] E.M. Frohman, F. Costello, O. Stuve, et al., Modeling axonal degener-
study of the acute and chronic effects of cooling therapy for MS, ation within the anterior visual system: implications for demonstrating
Neurology 60 (2003) 1955–1960. neuroprotection in multiple sclerosis, Arch. Neurol. 65 (2008) 26–35.
[112] B.J. Scherokman, J.B. Selhorst, E.A. Waybright, B. Jabbari, G.E. [134] R.C. Sergott, Historical perspective and future prospective for retinal
Bryan, C.G. Maitland, Improved optic nerve conduction with inges- nerve fiber loss in optic neuritis and multiple sclerosis, Int. Ophthal-
tion of ice water, Ann. Neurol. 17 (1985) 418–419. mol. Clin. 47 (2007) 15–24.
[113] S.J. Hickman, A.T. Toosy, S.J. Jones, et al., Serial magnetization [135] V. Parisi, M. Manni, M. Spadaro, et al., Correlation between mor-
transfer imaging in acute optic neuritis, Brain 127 (2004) 692–700. phological and functional retinal impairment in multiple sclerosis
[114] S.J. Hickman, A.T. Toosy, K.A. Miszkiel, et al., Visual recovery patients, Invest. Ophthalmol. Vis. Sci. 40 (1999) 2520–2527.
following acute optic neuritis—a clinical, electrophysiological and [136] J. Sepulcre, M. Murie-Fernandez, A. Salinas-Alaman, A. Garcia-
magnetic resonance imaging study, J. Neurol. 251 (2004) 996–1005. Layana, B. Bejarano, P. Villoslada, Diagnostic accuracy of retinal
[115] I.V. Allen, G. Glover, R. Anderson, Abnormalities in the macroscopi- abnormalities in predicting disease activity in MS, Neurology 68
cally normal white matter in cases of mild or spinal multiple sclerosis (2007) 1488–1494.
(MS), Acta Neuropathol. Suppl. 7 (1981) 176–178. [137] M. Pulicken, E. Gordon-Lipkin, L.J. Balcer, E. Frohman, G. Cutter,
[116] B.D. Trapp, L. Bo, S. Mork, A. Chang, Pathogenesis of tissue injury P.A. Calabresi, Optical coherence tomography and disease subtype in
in MS lesions, J. Neuroimmunol. 98 (1999) 49–56. multiple sclerosis, Neurology 69 (2007) 2085–2092.
[117] N. Evangelou, D. Konz, M.M. Esiri, S. Smith, J. Palace, P.M. [138] A.T. Toosy, S.J. Hickman, K.A. Miszkiel, et al., Adaptive cortical
Matthews, Size-selective neuronal changes in the anterior optic plasticity in higher visual areas after acute optic neuritis, Ann. Neurol.
pathways suggest a differential susceptibility to injury in multiple 57 (2005) 622–633.
sclerosis, Brain 124 (2001) 1813–1820. [139] S. De, D.M. Rabin, E. Salero, P.L. Lederman, S. Temple, J.H. Stern,
[118] G.T. Plant, R.F. Hess, Regional threshold contrast sensitivity within Human retinal pigment epithelium cell changes and expression of
the central visual field in optic neuritis, Brain 110 (Pt 2) (1987) alphaB-crystallin: a biomarker for retinal pigment epithelium cell
489–515. change in age-related macular degeneration, Arch. Ophthalmol. 125
[119] K.T. Mullen, G.T. Plant, Colour and luminance vision in human optic (2007) 641–645.
neuritis, Brain 109 (Pt 1) (1986) 1–13. [140] P.A. Brex, O. Ciccarelli, J.I. O’Riordan, M. Sailer, A.J. Thompson,
[120] G.T. Plant, Transient visually evoked potentials to sinusoidal grat- D.H. Miller, A longitudinal study of abnormalities on MRI and dis-
ings in optic neuritis, J. Neurol. Neurosurg. Psychiatry 46 (1983) ability from multiple sclerosis, N. Engl. J. Med. 346 (2002) 158–164.
1125–1133.

You might also like